You are on page 1of 259

A MULTI-SCALE DAMAGE PERCOLATION

MODEL OF DUCTILE FRACTURE



by

CLIFF BUTCHER


Master of Science in Engineering
University of New Brunswick
Fredericton, New Brunswick, 2007

A Thesis Submitted in Partial Fulfillment of
the Requirements for the Degree of

DOCTOR OF PHILOSOPHY

IN MECHANICAL ENGINEERING

in the Graduate Academic Unit of Mechanical Engineering

Supervisor: Dr. Z. Chen, Dept. of Mechanical Engineering
Examining Board: Dr. A. Simoneau (chair), Dept. of Mechanical Engineering
Dr. R.J. Rogers, Dept. of Mechanical Engineering
Dr. A.G. Gerber, Dept. of Mechanical Engineering


University of New Brunswick
July, 2011
Cliff Butcher, 2011
Fredericton, New Brunswick, 2011
ii
Abstract

The traditional approach to modeling ductile fracture involves homogenizing the
microstructure of a material into a simple, equivalent geometry from which the relevant
constitutive laws can be derived. While attractive from a modeling perspective, critical
details of the microstructure are lost in this homogenization process such as the particle
size, shape, orientation, distribution and degree of clustering. Since void initiation and
evolution is a highly localized phenomenon originating within heterogeneous particle
clusters, these models fail to accurately predict fracture. These limitations can be
overcome using a promising new technique known as damage percolation modeling that
requires no idealizations or approximations of the microstructure. In this approach,
digital imaging techniques or x-ray microtomography can be used to obtain the particle
distribution in a material. Using this information, micromechanical models are applied
to characterize void and crack formation leading to failure at the individual particle
scale. The damage percolation model represents the future in material modeling as it
directly relates changes in the local microstructure to the overall material behaviour.

In this research, the first fully-coupled multi-scale damage percolation model has
been developed to predict fracture in advanced materials with heterogeneous particle
distributions. In the first phase of this work, a sophisticated damage percolation model is
developed using the latest micromechanical models to characterize void nucleation,
growth, and coalescence in three-dimensions for general loading conditions. A novel
strategy is proposed to determine the stress state within the second-phase particles and
inclusions to facilitate the development of a void nucleation model based solely upon
the particle properties.

The percolation model was implemented into a commercial finite-element code
using so-called percolation elements to capture the complex stress- and strain-
gradients that develop during deformation. A particle field generator is developed and
integrated into the finite-element code to create representative particle distributions
within the percolation elements to provide stochastic predictions of fracture that reflect
iii
the experimental variation. Finally, the percolation model was validated numerically and
experimentally for an automotive-grade aluminum alloy in a notched tensile test used
for material characterization. The complete multi-scale percolation model predicts
fracture as a direct consequence of the stress state, material properties and the local
conditions within the microstructure.




















iv
Acknowledgements

I wish to express my most sincere appreciation and deepest gratitude to my
supervisor, Dr. Zengtao Chen, for his unwavering support, patience, and commitment to
achieving the highest levels of academic excellence. I am very fortunate to have a
supervisor who has allowed me to pursue my research wherever it has led me,
sometimes astray, and for providing the encouragement and guidance to find my way
back. It is this freedom that has kept the research both interesting and challenging and I
am very proud of our work together.

I gratefully appreciate the friendship of other members of the research group who
have made it a pleasure to come to the lab everyday.

Finally, I would like to gratefully acknowledge the financial support of NSERC,
AUTO21 Network of Centres for Excellence and the University of New Brunswick.












v
Table of Contents

Abstract.............................................................................................................................ii
Acknowledgements..........................................................................................................iv
List of Figures..................................................................................................................xi
List of Tables.................................................................................................................xxi
Nomenclature...............................................................................................................xxii
Chapter 1 - Damage-induced Ductile Fracture.....................................................1
1.1 Introduction................................................................................................................1
.1 Deformation of ductile materials...........................................................................1
.2 Deformation of ductile materials containing void defects.....................................2
.3 Damage-based constitutive models........................................................................3
.4 Motivation for this research...................................................................................4
1.2 Role of material damage............................................................................................5
.1 Void nucleation......................................................................................................5
.2 Void growth...........................................................................................................9
.3 Void shape evolution...........................................................................................10
.4 Void coalescence.................................................................................................12
1.3 Damage-based yield criteria....................................................................................18
1.4 Void evolution and coalescence within clusters......................................................22
1.5 Damage percolation modeling.................................................................................24
.1 Role of the void and particle distribution.............................................................24
.2 Previous damage percolation models...................................................................24
1.6 Outline of Thesis: A multi-scale damage-percolation model of ductile
fracture....................................................................................................................28
Chapter 2 - Modeling Void Growth to Coalescence: Validation and
Calibration......................................................................................................................30
2.1 Establishing confidence in the analytical void evolution models............................30
2.2 Modeling void evolution within a unit cell .............................................................31
2.3 Analysis of an axisymmetric unit cell .....................................................................32
.1 Boundary conditions............................................................................................33
vi
.2 Stress state and microstructure evolution ............................................................34
.3 Identification of coalescence................................................................................36
2.4 Numerical solution procedure..................................................................................38
2.5 Unit cell simulation results......................................................................................39
.1 Penny-shaped void...............................................................................................40
.2 Oblate void...........................................................................................................41
.3 Spherical void......................................................................................................43
.4 Prolate void..........................................................................................................44
.5 Selection of a minimum void aspect ratio............................................................45
2.6 Theoretical models for void growth, shape and coalescence...................................47
.1 Yield criterion......................................................................................................47
.2 Void growth, shape evolution and coalescence...................................................48
2.7 Evaluation of the analytical void evolution models.................................................50
.1 Void coalescence..................................................................................................50
.2 Void growth.........................................................................................................52
.3 Void shape............................................................................................................54
2.8 Calibration of the void evolution models.................................................................57
.1 Void growth.........................................................................................................57
.2 Void shape evolution...........................................................................................60
2.9 Evaluation of the complete constitutive model........................................................63
2.10 Summary................................................................................................................65
Chapter 3 - Integration of a Particle-based Homogenization Theory into a
General Damage-based Constitutive Model ...............................................................67
3.1 Void nucleation in a damage-based constitutive model ..........................................67
3.2 Review of particle-based homogenization theories ................................................68
3.3 Selection of a homogenization theory .....................................................................72
3.4 Particle-based homogenization model ....................................................................73
.1 Effective moduli of a randomly-oriented composite...........................................73
.2 Average stress in the composite and its constituents...........................................75
.3 Average strain in the composite and its constituents...........................................75

vii
3.5 Procedure for integrating a particle-based homogenization theory
into an existing damage-based constitutive model .................................................76
3.6 Iterative solution for the effective secant moduli ....................................................82
3.7 Application of the particle-based homogenization scheme into a
Gurson-based constitutive model for ductile fracture .............................................83
.1 Yield criterion......................................................................................................84
.2 Void growth, shape evolution..............................................................................84
.3 Void coalescence..................................................................................................84
3.8 Evaluation of the full constitutive model ................................................................85
3.9 Results .....................................................................................................................86
.1 Average stress in a material containing voids and hard elastic
particles in a ductile matrix..................................................................................86
.2 Influence of the particle shape on void damage and ductile fracture...................87
3.10 Summary ...............................................................................................................91
Chapter 4 - A Multi-scale Percolation Model of Ductile Fracture............................93
4.1 Development of the percolation model....................................................................93
4.2 Principal assumptions .............................................................................................93
.1 Assumption of a homogeneous stress state No void and particle
interactions...........................................................................................................95
.2 Assumption of internal necking coalescence as the dominant coalescence
mode.....................................................................................................................95
.3 Assumption of size-effects being negligible above the micron-scale..................96
.4 Neglecting the influence of texture effects and grains.........................................96
.5 Terminology used in development of the percolation model...............................97
4.3 Macroscopic models ...............................................................................................97
.1 Relevant finite-element kinematics....................................................................100
.2 Percolation element types..................................................................................101
.3 Constitutive model for material softening.........................................................104
.4 Degradation of the elastic moduli......................................................................106
.5 Global coalescence and failure of the percolation element................................107
4.4 Microscopic models ..............................................................................................108
viii
.1 Object kinematics...............................................................................................109
.2 Void growth and shape evolution......................................................................114
.3 Void nucleation..................................................................................................118
.4 Neighbours element and object.......................................................................125
.5 Void coalescence................................................................................................127
4.5 Summary ...............................................................................................................136
Chapter 5 - Development and Integration of a Particle Field Generator into
the Percolation Model .................................................................................................138
5.1 Acquiring measured particle fields .......................................................................138
5.2 Procedure for generating a particle field ...............................................................139
.1 Particle field basics ..........................................................................................140
.2 Generation of random variables .......................................................................141
.3 Generation of the global percolation volume ...................................................143
.4 Object generation .............................................................................................144
.5 Generation of objects within clusters ...............................................................145
.6 Particle properties ............................................................................................149
.7 Object constraints .............................................................................................150
.8 Identification of the parent element..................................................................152
5.3 Numerical implementation ....................................................................................152
5.4 Application of the particle field generator to an Al-Mg alloy...............................154
.1 Material characterization .................................................................................155
.2 Characterization of the particles and void dimensions ...................................157
.3 Characterization of void and particle clusters .................................................159
.4 Parameters used in the particle field generation process for AA5182 ............161
5.5 Results ...................................................................................................................162
.1 Object dimensions .............................................................................................162
.2 Spatial distribution of the objects ......................................................................165
.3 Particle and void volume fractions ....................................................................167
5.6 Summary ...............................................................................................................169
Chapter 6 - Application of the Percolation Model to Predict Fracture in a Notched
Tensile Test Specimen of AA5182 Sheet....................................................................171
ix
6.1 Application of the percolation model.....................................................................171
6.2 Notch tensile test experiments ..............................................................................171
6.3 Material characterization .......................................................................................174
.1 Material properties ............................................................................................174
.2 Void nucleation by the constituent particles .....................................................175
.3 Material properties of the constituent particles .................................................176
6.4 Finite-element model ............................................................................................177
.1 Mesh and boundary conditions .........................................................................177
.2 Identification of the fracture parameters ...........................................................179
6.5 Results and discussion ..........................................................................................179
.1 Ligament strain and axial strain at failure ..........................................................179
.2 Porosity and nucleation ......................................................................................181
.3 Nucleation stress and strain ................................................................................183
.4 Void and crack size ............................................................................................187
.5 Shape of the voids and cracks ............................................................................189
.6 Comparison of the percolation model with a traditional
damage-based model ..........................................................................................190
6.6 Summary ...............................................................................................................193
Chapter 7 - Void Growth and Evolution in Void Clusters ..................................... 194
7.1 Void evolution within clusters...............................................................................194
7.2 Non-local void evolution ......................................................................................195
7.3 Void interactions ...................................................................................................196
7.4 Measuring void interactions ..................................................................................198
.1 Interactions in the coalescence model ...............................................................198
.2 Nonlocal weighting function .............................................................................200
.3 Non-local void growth and shape evolution .....................................................202
.4 Nonlocal void size effect ...................................................................................203
7.5 Application of the non-local model to a linear void cluster...................................204
7.6 Results and discussion ..........................................................................................206
.1 Two-void linear cluster with initial spacing ratio .............................................206
.2 Three-void linear cluster ...................................................................................207
x
.3 Three-void linear cluster aligned with the loading direction ............................211
7.7 Future work and current limitations of the model .................................................212
7.8 Summary ...............................................................................................................212
Chapter 8 Conclusions..............................................................................................214
Chapter 9 Recommendations for Future Work.....................................................216
References ....................................................................................................................219
Appendices ...................................................................................................................230
Appendix A Eshelby Tensors ................................................................................230
Appendix B Important Geometrical Relationships of the Unit Cell ......................232
Appendix C Ellipsoids ...........................................................................................234
C.1. General equation of an ellipsoid ..................................................................234
C.2. Radius of ellipsoid along a viewing direction .............................................234
C.3. Line projection of an ellipsoid .....................................................................235
C.4. Arbitrary stretching of an ellipsoid .............................................................236

xi
List of Figures

Figure 1.1: Ductile damage evolution, (a) Initial state; (b) void nucleation;
(c) void growth; (d) void coalescence............................................................ 3
Figure 1.2: Particle debonding (left) and cracking (right) in aluminum
6061 reinforced with Al
2
O
3
particles..............................................................5
Figure 1.3: Idealized spheroidal void shapes...................................................................11
Figure 1.4: Comparison of the semi-empirical equations for void growth and
aspect ratio evolution with the results of Pardoen and Hutchinson
(2000) at a stress triaxiality of unity ............................................................12
Figure 1.5: (a) Triangular array of cylindrical holes in AA-5052 sheet and
(b) coalescence due to shear localization under tensile loading ..................13
Figure 1.6: (a) Rectangular array of cylindrical holes in AA-5052 sheet and
(b) coalescence due to localized necking under tensile loading...................13
Figure 1.7: Scanning electron microscope fractographs of Weldox 420
steel illustrating two different rupture mechanisms .....................................14
Figure 1.8: (a) Dual population of large and small voids at beginning of
deformation; (b) Activation and growth of small voids in a
region of high stress surrounding the large voids; (c) Fracture
as the voids link up throughout the material ................................................15
Figure 1.9: Cubic unit cell containing an axisymmetric elliptical void
subjected to uniaxial loading.........................................................................18
Figure 1.10: Void-matrix aggregate containing a random distribution
of arbitrary voids and its representation as a spherical and
cylindrical unit cell......................................................................................19
Figure 1.11: Gurson (1977) yield function showing the reduction in
the macroscopic equivalent stress with increasing hydrostatic
stress and porosity ......................................................................................20
Figure 1.12: Experimentally observed flow localization paths of Geltmacher
et al. (1996) in test specimens with a minimum hole spacing of
two hole diameters and loaded in uniaxial tension (left) and
xii
equal-biaxial tension (right) .......................................................................23
Figure 1. 13: Thomson et al. s (1999) 3D unit cell model used to model the
three void cluster types, (a) linear, (b) planar, and (c) quasi-spherical
. . . . . . 23
Figure 1.14: Tessellated AA5182 particle fields of 2000 x 2000 pixels
corresponding to a size of 0.75 mm x 0.75 mm .........................................25
Figure 1.15: Predicted damage prior to profuse coalescence within
a particle field of AA5182 sheet under plane strain loading
at 23.9% major strain .................................................................................26
Figure 1.16: Comparison of the predicted average diameter of the broken particles
with the experiment data of Hadianfard et al. (2009) using the 2-D
percolation model of Butcher and Chen (2009) .........................................26
Figure 1.17: Predicted damage within an AA5182 aluminum particle field during
a stretch flange forming operation .............................................................27
Figure 1.18: Multi-scale percolation model applied to a finite-element simulation
of a notched tensile test ..............................................................................29
Figure 2.1: Procedure for approximating a hexagonal distribution of unit cells into
an axisymmetric cell model that can be reduced to a one-quarter, 2-D
geometry for finite-element modeling ........................................................31
Figure 2.2: Typical cell geometry and finite-element mesh for an initially spherical
void in a one-quarter axisymmetric unit cell (left) and a one-eight cubic
unit cell (right) .............................................................................................32
Figure 2.3: General geometry and boundary conditions used to model the one-
quarter axisymmetric unit cell ......................................................................33
Figure 2.4: Typical deformation history of a voided unit cell ........................................37
Figure 2.5: Typical relationship of the macroscopic radial and axial strains for a
voided unit cell .............................................................................................38
Figure 2.6: Typical macroscopic stress and strain curve for a voided unit cell ..............38
Figure 2.7: Porosity history of a penny-shaped void subjected to various triaxial
loadings ........................................................................................................41
Figure 2.8: Evolution of the void aspect ratio of an initially penny-shaped void
xiii
subjected to various triaxial loadings ...........................................................41
Figure 2.9: Porosity history of an initially oblate void subjected to various triaxial
loadings ........................................................................................................42
Figure 2.10: Evolution of the void aspect ratio of an initially oblate void subjected
to various triaxial loadings .........................................................................42
Figure 2.11: Porosity history of an initially spherical void subjected to various
triaxial loadings ..........................................................................................43
Figure 2.12: Evolution of the void aspect ratio of an initially spherical void
subjected to various triaxial loadings .........................................................44
Figure 2.13: Porosity history of an initially prolate void subjected to various
triaxial loadings ..........................................................................................45
Figure 2.14: Evolution of the void aspect ratio of an initially prolate void
subjected to various triaxial loadings .........................................................45
Figure 2.15: Comparison of void growth for different penny-shaped voids
for the extreme cases of the stress triaxiality .............................................46
Figure 2.16: Comparison of void shape evolution for different penny-shaped
voids for the extreme cases of the stress triaxiality ...................................47
Figure 2.17: Comparison of the numerical and predicted coalescence strains
using the coalescence model of Pardoen and Hutchinson (2000) in
Eq. (1.13) ....................................................................................................50
Figure 2.18: Comparison of the numerical and predicted coalescence strains
using the coalescence model of Benzerga (2002) .....................................51
Figure 2.19: Comparison of the analytical void growth model with the unit cell
data for an initially penny-shaped void ......................................................52
Figure 2.20: Comparison of the analytical void growth model with the unit cell
data for an initially oblate void ..................................................................53
Figure 2.21: Comparison of the analytical void growth model with the unit cell
data for an initially spherical void ..............................................................53
Figure 2.22: Comparison of the analytical void growth model with the unit cell
data for an initially prolate void .................................................................54
Figure 2.23: Comparison of the analytical void shape evolution model with the
xiv
unit cell data for an initially penny-shaped void ........................................55
Figure 2.24: Comparison of the analytical void shape evolution model with the
unit cell data for an initially oblate void ....................................................55
Figure 2.25: Comparison of the analytical void shape evolution model with the
unit cell data for an initially spherical void ................................................56
Figure 2.26: Comparison of the analytical void shape evolution model with the
unit cell data for an initially prolate void ...................................................56
Figure 2.27: Calibrated q
2
parameter for an initially penny-shaped void subjected
to various stress triaxialities .......................................................................57
Figure 2.28: Calibrated q
2
parameter for an initially spherical void subjected to
various stress triaxialities ...........................................................................58
Figure 2.29: Calibrated q
2
parameter for an initially prolate void subjected to
various stress triaxialities ...........................................................................58
Figure 2.30: Comparison of the calibrated void growth model with the unit cell
data for an initially penny-shaped void ......................................................59
Figure 2.31: Comparison of the calibrated void growth model with the unit cell
data for an initially spherical void ..............................................................59
Figure 2.32: Comparison of the calibrated void growth model with the unit cell
data for an initially prolate void .................................................................60
Figure 2.33: Comparison of the void aspect correlation function with the model
of Ragab (2004a) and the unit cell data .....................................................61
Figure 2.34: Comparison of the void shape correlation with the unit cell data
for an initially penny-shaped void .............................................................62
Figure 2.35: Comparison of the void shape correlation with the unit cell data
for an initially spherical void .....................................................................63
Figure 2.36: Comparison of the void shape correlation with the unit cell data
for an initially prolate void .........................................................................63
Figure 2.37: Comparison of the constitutive model and the unit cell data for
the maximum principal stress for an initially penny-shaped
void in uniaxial tension ..............................................................................64
Figure 2.38: Comparison of the constitutive model and the unit cell data for
xv
the porosity evolution of an initially penny-shaped void in uniaxial
tension ........................................................................................................64
Figure 2.39: Comparison of the constitutive model and the unit cell data
for the porosity evolution of an initially oblate void in uniaxial
tension ........................................................................................................65
Figure 3.1: Predictions of the tensile stress-strain curves in the longitudinal
direction for the composite reinforced with ellipsoidal inclusions
with a volume fraction of 25%. (a) n = 0.05; (b) n = 0.40 ..........................71
Figure 3.2: (a) Evolution of the plastic strain in the matrix as a function of
the applied strain. (b) Evolution of the von Mises equivalent
stress in the ellipsoids as a function of the applied strain ............................72
Figure 3.3: Equivalent stress response of the composite matrix with the applied
strain .............................................................................................................78
Figure 3.4: Schematic of the integration process of a particle-based
homogenization theory into a damage-based constitutive model ................80
Figure 3.5: Equivalent stress within the constituents of the voided bulk material
when subjected to uniaxial tension ..............................................................86
Figure 3.6: Normalized principal stress within the spherical elastic particles
and inclusions for three loading conditions .................................................87
Figure 3.7: Influence of the particle shape on damage development in the
coupled constitutive model compared to the porosity determined
using the standard model ..............................................................................88
Figure 3.8: Influence of the particle shape on the evolution of the void shape
in the coupled constitutive model relative to the void shape
predicted using the standard model ..............................................................89
Figure 3.9: Influence of the particle shape on the stress triaxiality within the
matrix material compared to the standard model where the triaxiality
is constant at a value of 1/3 ..........................................................................90
Figure 3.10: Variation of the plastic limit-load required for coalescence in
the coupled and standard (no particles) damage-based
constitutive models ....................................................................................91
xvi
Figure 4.1: Flow-chart of the macroscopic modeling process ........................................99
Figure 4.2: Finite-element discretization (Bonet and Wood, 1997) ..............................100
Figure 4.3: Local node numbering and coordinate systems for a 4-node (left)
and 8-node (right) isoparametric element ..................................................103
Figure 4.4: Flow-chart of the percolation modeling process ........................................109
Figure 4.5: Percolation element in the dimensionless element coordinate
system (left) and in the global coordinate system (right) where the
voids and particles have translated and rotated with the deformation
of the element .............................................................................................110
Figure 4.6: Digitally compressed montage of 500 x 500 m
2
area of an
undeformed AA6061 alloy (left) and a montage after being subjected
to a strain of 98% in torsion (right) ............................................................111
Figure 4.7: Comparison of the analytical rotation model of Kailasam and Ponte
Casteneda (1998) with the void rotation obtained from unit cell
simulations .................................................................................................112
Figure 4.8: Procedure for modeling an arbitrary ellipsoidal void as an
axisymmetric void as viewed by the principal loading direction ...............116
Figure 4.9: Stretching of a void in three arbitrary directions resulting in a new
void size and orientation ............................................................................117
Figure 4.9: Void nucleation by the cracking of an oblate ellipsoidal inclusion ...........119
Figure 4.10: SEM observation from Moulin et al. (2009) showing the intermetallic
particles in as- cast AA5182 (left) and the typical morphology of the
large Fe-rich particles (right) ...................................................................120
Figure 4.11: Variation of the nucleation stress with particle size and toughness .........123
Figure 4.12: Procedure for modeling void nucleation ..................................................125
Figure 4.13: Void coalescence geometry showing the in-plane tensile stress values
transverse to the inter-void ligament vector, w ........................................129
Figure 4.14: Void coalescence geometry showing the identification of the relevant
dimensions ...............................................................................................132
Figure 4.15: Equivalent void geometry for evaluation of the plastic limit-load
coalescence model ....................................................................................133
xvii
Figure 4.16: Two-dimensional schematic of coalescence between two voids
showing their bounding boxes and the creation of the new crack ...........134
Figure 4.17: Schematic of coalescence between a large and small void and the
resulting crack geometry ..........................................................................135
Figure 5.1: Procedure for generating a particle field within a block of
percolation elements ..................................................................................140
Figure 5.2: Schematic of the rejection sampling algorithm used to determine a
random variable, x, from a target probability density function ..................142
Figure 5.3: Orientation of an object relative to the rolling direction by a solid
angle,
r
......................................................................................................145
Figure 5.4: Schematic showing the possible measures used in calculating the
inter-object spacing ....................................................................................147
Figure 5.5: Schematic showing the creation of a stringer-type cluster .........................149
Figure 5.6: Schematic of optimization method of Lin and Han (2002) for the
minimum distance between two general ellipsoids ....................................151
Figure 5.7: Algorithm for the particle field generator ...................................................153
Figure 5.8: In-plane images of the particle distribution for AA5182 sheet for a
thickness of 1.0 mm (left) and 1.6 mm (right) ...........................................155
Figure 5.9: 3D view of intermetallic phases and voids in 1 mm thick AA5182 sheet
obtained by Maire et al. (2006): (a) as cast state; (b) the hot rolled state;
(c) cold rolled state; and (d) microstructure at the end of a tensile test .....156
Figure 5.10: Distributions of the semi-axes of the Fe-rich particles. Adapted from
Orlov (2006) .............................................................................................158
Figure 5.11: Distributions of the semi-axes of the Mg
2
Si particles. Adapted from
Orlov (2006) .............................................................................................158
Figure 5.12: Distributions of the semi-axes of the voids. Adapted from
Orlov (2006) .............................................................................................159
Figure 5.13: Distributions for the inter-object spacing as measured using the
center-to-center distance ..........................................................................160
Figure 5.14: Estimation of the absolute minimum distance between objects from
the center-to- center distance ...................................................................161
xviii
Figure 5.15: Generated particle field of AA5182 with a volume of 200 m x
200 m x 200 m .....................................................................................163
Figure 5.16: Distributions of the semi-axes of the Fe-rich particles .............................164
Figure 5.17: Distributions of the semi-axes of the Mg
2
Si particles ..............................164
Figure 5.18: Distributions of the semi-axes of the voids ..............................................165
Figure 5.19: Distributions of the minimum center-to-center distance between
objects ......................................................................................................166
Figure 5.20a: 95% Confidence intervals for the predicted volume fractions of the
voids and Mg
2
Si particles in AA5182 sheet ..........................................168
Figure 5.20b: 95% Confidence intervals for the predicted volume fractions of the
Fe-rich particles in AA5182 sheet .........................................................169
Figure 6.1: Geometry of a typical notched tensile sheet specimen ...............................172
Figure 6.2: Crack initiation at the notch root in a 1.5 mm thick AA5182 tensile
specimen with a notch ratio of 0.125 and radius of 1 mm .........................173
Figure 6.3: Comparison of the average experimental and predicted load-curve
for an AA5182 notched tensile specimen ..................................................173
Figure 6.4: Experimental flow stress relation for AA5182 sheet and its Voce law
representation .............................................................................................174
Figure 6.5: Variation of the material hardening exponent in AA5182 .........................175
Figure 6.6: (a) Boundary conditions in the one-eighth finite-element of a notched
tensile specimen with a notch ratio of 0.25 ................................................178
Figure 6.7: Porosity distribution and failure location in a notch tensile specimen
with a notch ratio of 0.25 ...........................................................................178
Figure 6.8: Comparison of the experimental and predicted 95% confidence
intervals for the axial strain at failure ........................................................180
Figure 6.9: Comparison of the experimental and predicted 95% confidence
intervals for the ligament strain at failure ..................................................180
Figure 6.10: Evolution of the equivalent plastic strain at the center of the notch
root ...........................................................................................................181
Figure 6.11: Comparison of the predicted porosity in the center of the notch root
with the experiment data of Orlov (2006) for a standard tensile
xix
specimen ...................................................................................................182
Figure 6.11: Comparison of the total number of voids and cracks predicted by the
percolation model with the experimental results of Orlov (2006) ...........182
Figure 6.12: Comparison of the average maximum principal stress in the particles
at void nucleation for each particle field considered ...............................184
Figure 6.13: Predicted average plastic strain of the particles upon nucleation via
cracking ....................................................................................................185
Figure 6.14: Variation of the nucleation strain with the particle size and stress state
using the phenomenological model of Butcher and Chen (2009a) ..........186
Figure 6.15: Predicted average equivalent radius of the voids in the various
particle fields ............................................................................................187
Figure 6.16: Predicted average equivalent radius of the cracks in the various
particle fields ............................................................................................188
Figure 6.17: SEM micrograph of the fracture surface at the notch root for an
AA5182 tensile specimen with a notch ratio of 0.125 .............................189
Figure 6.18: Evolution of the global aspect ratio (voids and cracks) in the
various particle fields ...............................................................................190
Figure 6.19: Predicted porosity in the notch root using the percolation model
and the Gurson-based model of Butcher and Chen (2011) who
considered both stress- and strain-controlled nucleation models .............191
Figure 6.20: Comparison of the predicted porosity due to nucleated voids of the
percolation model with the Gurson-based model used by Butcher
and Chen (2011) .......................................................................................192
Figure 7.1: Schematic of a non-local averaging procedure showing the
characteristic length ...................................................................................195
Figure 7.2: Fracture of a 45 array of 1 mm diameter holes in Al-Mg sheet ................197
Figure 7.3: Evolution of the plastic constraint factor and stress state for an
initially spherical void ................................................................................199
Figure 7.4: Evolution of the void aspect ratio for an initially spherical void ...............200
Figure 7.5: Evolution of the plastic limit load metric for an initially spherical void ....201
Figure 7.6: Small voids rapidly grow next to primary voids where the local
xx
stress triaxiality is high ..............................................................................203
Figure 7.7: Typical geometry of a solid percolation element containing a
linear cluster of voids .................................................................................205
Figure 7.8: Comparison of the predicted porosity for a two void linear cluster
with non-local interactions with the porosity when the voids remain
isolated (no interactions) ............................................................................206
Figure 7.9: Comparison of the predicted aspect ratio for a two void linear cluster
with non-local interactions with the aspect ratio when the voids
remain isolated (no interactions) ................................................................207
Figure 7.10: Comparison of the predicted porosity for a three void linear cluster
with non-local interactions with the porosity when the voids remain
isolated (no interactions) ..........................................................................208
Figure 7.11: Comparison of the predicted aspect ratio for a two void linear
cluster with non-local interactions with the aspect ratio when
the voids remain isolated (no interactions) ..............................................209
Figure 7.12: Predicted porosity for a three-void linear cluster relative to the
porosity trend for the isolated case (no interactions) ...............................210
Figure 7.13: Predicted aspect ratio for a three-void linear cluster relative to the
aspect ratio trend obtained for the isolated case (no interactions) ...........210









xxi

List of Tables

Table 3.1. Material parameters of the constituents in an AA5083-O alloy ....................85
Table 5.1. Chemical composition of the AA5182-O alloy ...........................................155
Table 5.2. Object densities (number per mm
3
)

in AA5182 sheet .................................157
Table 5.3. Volume fractions of the constituents in AA5182 sheet ...............................157
Table 6.1: Material parameters of AA5182 required by the percolation model ...........177





















xxii

Nomenclature


Standard notation is used throughout. Boldface symbols denote tensors with the
orders indicated by the context. All tensor components are defined in respect to the
global Cartesian system with the base vectors e
i
(i = 1, 2, 3) and the summation
convention is used for repeated indices, unless otherwise specified. A superscript T
denotes the transpose and a superposed dot indicates the time derivative. Let a, b, be
vectors, A, B, are second-order tensors and A, B, denote fourth-order tensors. The
second and fourth-order identity tensors are denoted as I and I, respectively.

Macroscopic quantities are denoted by upper-case letters and symbols such as the
macroscopic stress and strain tensors,
ij
, and E
ij
. Microscopic quantities are expressed
using the lower-case version of their macroscopic counterparts so that the stress and
strain tensors are,
ij
o and
ij
c . Quantities in the plastic regime will be denoted with a
superscript p unless stated otherwise. Volume averaged and effective/equivalent values
are denoted using an overbar symbol such as o for the matrix flow stress.


1
Chapter 1 Damage-induced ductile fracture

1.1 Introduction

To properly design engineering structures and manufacturing processes, the
engineer must be able to accurately predict the onset of material failure. The failure
mechanism is dependent upon many factors such as the environment, loading condition
and material composition. Depending on the application, material failure can be
classified as the onset of elastic instabilities such as buckling, permanent (plastic)
deformation, or the formation of a localized neck and fracture. From the perspective of
sheet metal forming, failure is defined by the formation of a localized neck, crack,
wrinkle or complete fracture since all of these effects result in an unacceptable
component. To successfully characterize the material behavior in a metal forming
operation requires knowledge of a yield criterion to identify the onset of plastic or
permanent deformation, a hardening rule and a failure criterion.

.1 Deformation of ductile materials

The von Mises yield criterion (1913) is commonly used to describe yielding of
ductile metals and is in good agreement with experimental observations (Norton, 2000).
The yielding mechanism in ductile materials is due to the relative sliding of the
materials atoms within their lattice structure. Sliding is caused by shear stress and thus
distorts the lattice where the energy due to this distortion is related to the applied shear
stress. The von Mises criterion predicts yielding when the distortional component of the
strain energy reaches the energy required to initiate yielding in a tensile test. The von
Mises effective stress is expressed as


2 2 2
eq 1 2 3 1 2 2 3 1 3
E = E +E +E E E E E E E (1.1)

where
1 3
E are the principal stresses in descending order. The material has yielded when
eq y
o E = , where
y
o is the yield strength determined from a tensile test. For most
engineering materials, the onset of yielding is accompanied by an increase in strength
2
due to the pile up of dislocations within the lattice. This strain-hardening behavior is
often described using a power law relationship between the yield stress and plastic strain
in the material. The von Mises yield criterion assumes the material to be isotropic, free
of internal defects and independent of the hydrostatic stress (negative pressure). The
hydrostatic component of the stress state does not appear in the yield criterion since it
deforms the material uniformly and introduces no distortional energy.

.2 Deformation of ductile materials containing microstructural imperfections

Unfortunately, the von Mises constitutive model cannot predict the fracture of
ductile metals since the material is assumed to be free of the internal defects that cause
failure. This limitation becomes acute in sheet metal forming where the formability of
the material is limited by the presence of initial micro-voids (porosity) and second-phase
particles and inclusions. A typical ductile material contains an initial porosity on the
order of 0.1 1% (Pardoen and Hutchinson, 2000). During plastic deformation, the pre-
existing voids grow while additional voids are nucleated at second-phase particles or
inclusions. Void growth and nucleation continue during deformation before linking up
to form micro-cracks (coalescence). The presence of micro-voids and cracks promotes
local softening of the material leading to localized necking. Strong stress- and strain-
gradients develop in the necked region, accelerating void growth and nucleation
resulting in widespread coalescence and failure of the material.

To model this fracture mechanism, the evolution of the voids is typically separated
into three idealized stages: void nucleation, growth and coalescence (Figure 1.1). While
easily categorized into independent regimes, the reality is that these mechanisms are
tightly interwoven and related to many additional factors such as the second-phase
particle/void distribution, void geometry, stress state, strain rate, material hardening and
temperature (Horstemeyer et al., 2003). Furthermore, the highly localized and dynamic
nature of ductile fracture makes observing void initiation and evolution extremely
difficult. Due to the complexity of the fracture mechanism, damage-based constitutive
models must employ many approximations and simplifications to reach a tractable
analytical or numerical solution.
3

Figure 1.1: Ductile damage evolution, (a) Initial state; (b) void nucleation; (c) void growth;
(d) void coalescence (Chen, 2004).

.3 Damage-based constitutive models

Originally developed in the 1970s, the damage-based constitutive model of Gurson
(1977) and its well-known extensions (Tvergaard, 1981; Tvergaard and Needleman,
1984) have been widely used to predict formability and void damage in porous ductile
materials, albeit with mixed results. This type of constitutive modeling is based upon the
assumption that a porous material can be decomposed into a periodic array of
representative volume elements (RVE) or unit cells that contain a single void. From this
unit cell geometry, Gurson (1977) developed a pressure-sensitive yield criterion that
accounts for material softening due to the presence of voids within the material. The
voids nucleate, grow and coalesce according to adopted sub-models.

In the subsequent years, Gursons modeling approach has led to increasingly rigorous
treatments and models that account for the void shape, anisotropy, size effects and
multiple-void populations (Golaganu et al., 1993, 1994; Pardoen and Hutchinson, 2000;
Benzerga, 2002; Wen et al., 2003; Fabregue and Pardoen, 2008). Despite these
advances, the models remain highly approximate and neglect critical factors such as the
particle/void distribution and local microstructure. To compensate for the numerous
approximations in these models, many calibration parameters have been introduced that
must be identified using both numerical and experimental techniques. The inherent
4
difficulty in identifying these parameters and their lack of transferability to different
stress states limits their industrial application (Zhang et al., 2000).

The fundamental limitation in Gurson-based models is that the voids are
homogenized throughout the material while ductile fracture is a highly localized process
originating within particle and void clusters. An equivalent void is used to describe the
porosity within the material. While advantageous from a modeling perspective,
describing void damage in an average sense inadvertently overlooks the very
mechanism that causes fracture: localization within heterogeneous void clusters. To
motivate the practical application of damage-based material models, a bottom-up
approach to fracture is required where damage evolution at the individual particle scale
is directly linked to the bulk behaviour of the material without homogenizing the
microstructure or introducing ad-hoc calibration parameters. Only by analyzing the
particle field of a material can ductile fracture be properly represented.

.4 Motivation for this research

This research seeks to address damage-induced ductile fracture within a consistent
framework where void growth, nucleation and coalescence are modeled within
measured particle fields to predict fracture. Unlike traditional damage models, the
localized nature of void nucleation, growth and coalescence is captured without relying
on assumptions regarding the particle/void size, distribution and orientation. In this
approach, influence of the particle distribution and localized nature of ductile fracture
can be directly linked to the bulk behavior of the material.

The following review of the literature pertinent to the current research will focus on
the three stages of damage-induced ductile fracture: void nucleation, growth and
coalescence. An emphasis is placed upon the various micromechanical modeling
techniques used to describe damage evolution within the constitutive models. The
influence of microstructural heterogeneity and the limitations of the traditional approach
to modeling ductile fracture are discussed, followed by a summary of the damage
percolation modeling technique that will be used in this work.
5
1.2 Role of material damage

.1 Void nucleation

Void nucleation occurs as second phase particles or inclusions crack or debond from
the matrix material during plastic deformation as shown in Figure 1.2. Void nucleation
is difficult to model since the particle may crack, or debond (partially or completely)
depending on the particle size, shape, distribution, stress state, temperature and strength
of the particle and interface (Gurland and Plateau, 1963; Gurland, 1972; Cox and Low,
1974; Tanaka et al., 1970; Argon et al., 1975a, b, c; Goods and Brown, 1979; Fisher and
Gurland, 1981a, b; Beremin, 1981; Horstemeyer and Gokhale, 1999; Horstemeyer et al.,
2000, 2003).


Figure 1.2: Particle debonding (left) and cracking (right) in aluminum 6061 reinforced
with Al
2
O
3
particles. The loading direction is horizontal (Kanetake et al., 1995).

The size and shape of the second-phase particles/inclusions play a major role in the
resulting nucleation mechanism. Large particles (radius greater than 1 m) are prone to
cracking at lower strains than smaller sub-micron particles which exhibit debonding
(Thomason, 1990). Nucleation occurs earlier for large particles because they contain
more internal surface defects, have a larger interface and are irregularly shaped. An
elongated particle aligned in the principal loading direction is more likely to crack while
debonding is more common if loaded transversely. Small particles tend to be spherical
or equiaxed and favour interface fracture due to the pile-up of dislocations at the
particle-matrix interface. Equiaxed particles are less sensitive to the loading direction
6
and nucleation is a competition between the cracking and debonding (Lassance et al.,
2006). Lower temperatures promote both cracking and debonding due to increased work
hardening rates (Horstemeyer et al., 2003).

To further complicate matters the nucleation mechanism can be considered to be a
stress- or strain-controlled process. The debonding of small particles is related to
dislocation kinetics and can thus be considered strain-dependent (McClintock, 1968a).
Particle debonding can also be described as stress-controlled since the separation of the
interface is dependant upon the hydrostatic stress (Needleman 1987; Shabrov and
Needleman, 2002). Alternatively, the nucleation of large particles due to cracking can be
described using a stress-controlled model since this is a brittle-type fracture (Thomason,
1990). Overall, nucleation is a complex and material specific phenomenon which is
difficult to model and observe experimentally. Consequently, void nucleation models
typically follow a continuum-based approach where the objective is to model the
average nucleation response of an alloy. The shortcomings of these nucleation models
are corrected (ideally) through the introduction of material specific calibration
parameters.

.1 Modeling void nucleation

Gurland and Plateau (1963) proposed one of the first nucleation models and adopted
an energy approach for particle cracking. It was assumed that a particle fractures when
the strain energy in the particle exceeds the surface energy of the newly formed crack
surface. Tanaka et al. (1970) performed a micromechanical analysis of a spherical
inclusion subjected to uniaxial tension and found that the energy criterion of Gurland
and Plateau (1963) is indeed a necessary condition for nucleation. Argon et al.
(1975a,b,c) proved that the energy requirement is a necessary but not sufficient
condition; the strength of the particle or interface must also be exceeded. The stress-
based model of Argon et al. (1975a, b, c) assumed that the critical interface stress,
N
o ,
could be approximated as the sum of the flow and hydrostatic stress components


N hyd
o o = + E (1.2)
7
where o is the matrix flow stress and
hyd
E is the macroscopic hydrostatic stress defined
as

1 2 3
hyd
3
E +E +E
E = (1.3)

Brown and Stobbs (1976) also developed a stress-based nucleation rule by relating
the stress at the particle-matrix interface with dislocation density


N
2
v d
b p
o
t
= (1.4)

where is the shear modulus,
d
p is the dislocation density and b
v
is the Burgers vector.
The dislocation density can be related to the critical strain required to nucleate a particle
based on particle size, r, and interface stress. The resulting nucleation criterion is


2 2
N
c 2
8
5.1
v
r
G b
t o
c = (1.5)

The Brown and Stobbs (1976) model incorporated the effect of void size and
observed that large particles tend to nucleate at smaller strains. Gurson (1977) suggested
that nucleation can be expressed in terms of equivalent plastic strain based on Gurlands
(1972) experimental results for nucleation in spheroidized steel. Following Gursons
result, Chu and Needleman (1980) theorized that the stress or strain required to nucleate
a void follows a normal distribution. The strain-controlled nucleation model is expressed
as

2
p
p N n
nucleation
N N
1
exp
2 2
f
f
s s
c c
c
t
(
| |
( =
|
(
\ .

(1.6)

where
nucleation
f is the porosity contributed by void nucleation and an over-dot symbol
indicates the time derivative;
p
c is the equivalent plastic strain,
n
f is the volume fraction
of void nucleating particles and
N
c and
N
s are the mean and standard deviation of the
nucleation strain. For stress-controlled nucleation
8

2
hyd N
n
nucleation hyd
N y N y
1
exp ( )
2 2
f
f
s s
o o
o
o o t
(
| | +E
(
= +E
|
|
(
\ .

(1.7)

where
N
o is the critical nucleation stress and
y
o is the initial yield stress.

Beremin (1981) also proposed a stress-based nucleation criterion by employing the
homogenization method of Bervellier and Zaoui (1979) to estimate the stress within the
inclusions. The current stress within the particle is related to the stress in the matrix
material as

N 1 y
( ) k o o o = E + (1.8)

where k is a parameter related to the composition of the particles and the loading
direction. The Beremin (1981) model can be coupled into the normal distribution model
in Eq. (1.7) by replacing the term
hyd
o + E with the stress obtained in Eq. (1.8).

The strain-controlled criterion in Eq. (1.6) has been widely used in the literature
because it is easily implemented into finite-element codes and is a monotonic increasing
quantity that is independent of the stress state. The stress-controlled nucleation model
has not received much attention as it can cause non-normality in the flow rule and
promote early flow localization (Saje et al., 1982). Recent works by Butcher et al.
(2006, 2009) and Butcher and Chen (2011) have shown that the stress-controlled
nucleation model in Eq. (1.7) can give good predictions of nucleation in aluminum-
magnesium and advanced high strength steels. The advantage of using a stress-based
nucleation model is that it is not sensitive to the load-path like the strain-controlled
model and can be transferable to general stress states.

Needleman (1987) numerically investigated the debonding of spherical particles
within a characteristic unit cell and observed that only a fraction of the hydrostatic stress
was present at the particle-matrix interface. A reduction factor, c, was introduced into
the model of Argon (1975a, b, c) in Eq. (1.2) as


N hyd
c o o o = + (1.9)
9

where 0.35 c ~ for spherical particles and 0.44 c ~ for cubic particles (Needleman, 1987;
Shabrov and Needleman, 2002). Thomson (2001) extended the work of Needleman (1987)
to investigate debonding within particle clusters and observed nucleation to be very
sensitive to the cluster arrangement and density. This study highlighted the localized
nature of the nucleation mechanism and the importance of accounting for the particle
distribution on void nucleation, growth and coalescence. It is important to mention that the
commonly used nucleation models of Chu and Needleman (1980) in Eq. (1.6) and (1.7) do
not account for the particle size, shape, or the stress state within the particle. None of the
nucleation models described above account for the particle distribution.

.2 Void growth

Void growth is the most understood stage of ductile fracture. Void growth is a
continuum plastic deformation process and therefore easier to model than void
nucleation or coalescence (Zhang and Skallerud, 2010). While void nucleation and
coalescence often occur suddenly, void growth is a relatively stable phase of
deformation and easier to observe experimentally and model analytically.

.1 Modeling void growth

Voids are typically assumed to be initially cylindrical, spherical or spheroidal and
change shape during loading. McClintock (1968b) first analyzed a cylindrical void in an
infinite work-hardening matrix and expressed void growth as a function of remote
strain. Rice and Tracey (1969) analyzed a spherical void in an infinite, perfectly plastic
material and demonstrated the exponential dependence of void growth upon the tensile
hydrostatic stress (negative pressure). The intensity of the hydrostatic stress is defined
by the stress triaxiality, T, which is the ratio of the hydrostatic to effective stress. Stress
triaxialities of 0, 1/3 and 2/3 correspond to pure shear, uniaxial tension and equal-biaxial
tension, respectively. Huang (1991) reworked the numerics in the Rice and Tracey
(1969) model to better reflect growth at lower stress triaxiality. The resulting void
growth rate is expressed as
10
( )
1
p
4
3
1.28 exp
2
f
T T
f
c
| |
=
|
\ .
for
1
1
3
T s s (1.10a)


p
3
1.28exp
2
f
T
f
c
| |
=
|
\ .
for 1 T > (1.10b)

The void growth model of Rice and Tracey (1969) has been continually improved to
yield results in good agreement with finite-element simulations (LeRoy et al., 1981;
Budiansky et al., 1982; Huang, 1991; Thomason, 1993). More recently, Liu et al. (2003)
extended the Rice-Tracey model using strain-gradient plasticity theory to account for the
void size effect as small voids tend to grow slower than large voids. The growth of sub-
micron sized voids is impeded by the toughening of the material at the void surfaces due
to the pile up of dislocations. It is important to note that all of these void growth models
are only valid for isolated voids as interactions with nearby voids are not included. The
void growth rate is significantly larger within void clusters (Thomson et al., 1999;
Bandstra and Koss, 2008) where severe stress- and strain-gradients accelerate both void
growth and coalescence.

Overall, the analytical solutions for void growth can provide decent agreement with
the numerical results (Ragab, 2004a) but are limited by their ability to account for void
shape effects. To provide a void growth rule more suited for engineering applications,
Ragab (2004a) compiled an exhaustive summary of the analytical and finite-element
results for void growth and shape evolution in the literature to develop a set of semi-
empirical equations for a wide range of materials, void shapes and stress states. Void
growth in the Ragab (2004a) model is derived from a reduced form of the growth law in
the Gurson (1977) model and will be discussed at length in Chapter 2.

.3 Void shape evolution

Naturally, the void can be expected to change its shape during deformation as the
voids are not internally pressurized and must deform according to the plastic flow of the
surrounding matrix material. The void growth rate and shape evolution are intrinsically
linked as the void shape induces anisotropy, altering the stress state and the growth rate
11
in a non-linear fashion. To simplify the modeling procedure, it is often assumed that the
voids can be approximated as axisymmetric ellipsoids (spheroids) as shown in Figure
1.3. In uniaxial tension, void shape evolution is significant as initially penny-shaped
voids can rapidly elongate in the principal loading direction into prolate voids.
Alternatively, at high stress triaxialities as found ahead of a crack tip, the voids will
grow laterally and become flat and oblate, regardless of the principal loading direction.
Penny-shaped voids are the exception to this trend and only appreciably grow in the
opening direction even at high stress triaxiality. The mechanics for modeling void shape
evolution are complex and no analytical model has yet been developed that can account
for the void shape evolution in different stress states. As such, the adoption of heuristic
parameters in these models is currently unavoidable (Keralavarma and Benzerga, 2010).

(b) oblate spheroid: W < 1 (c) sphere: W = 1
(d) prolate spheroid: W > 1
(a) penny-shaped spheroid: W 0

Figure 1.3: Idealized spheroidal void shapes

Along with his study of semi-empirical rules for void growth, Ragab (2004a)
consolidated the void shape evolution studies in the literature to develop a set of semi-
empirical evolution laws using a total strain formulation. The semi-empirical evolution
laws are functions of the nucleation strain,
N
c , equivalent plastic strain, initial void
aspect ratio,
i
W , hardening exponent, n, and stress triaxiality in the general form


i
( , , , , )
p
N
W g W T n c c = (1.11)

The void shape evolution laws of Ragab (2004a) in Eq. (1.11) will be described in detail
in Chapter 2. Void growth and shape evolution in the Ragab (2004a) equations assume
an isolated void and do not account for interactions with neighboring voids. The
12
advantages of these semi-empirical expressions for void growth and shape evolution are
that they are efficient to compute and ideally suited for implementation into a
percolation-type model that may contain thousands of isolated voids of different shapes.
A rate-based void evolution rule has also been widely used in the GLD (Gologanu,
Leblond and Devaux, 1993, 1994) constitutive model with good accuracy and this rule
could be used instead of the Ragab (2004a) model. The Ragab (2004a) model has been
shown to give very good agreement with the numerical results of Pardoen and
Hutchinson (2000) who employed the sophisticated GLD model (Figure 1.4).

Figure 1.4: Comparison of the semi-empirical equations for void growth and aspect ratio
evolution with the results of Pardoen and Hutchinson (2000) at a stress triaxiality of unity
(Ragab, 2004b). The voids aspect ratio in this figure is denoted by
1
.

.4 Void coalescence

Void coalescence is a sudden and rapidly occurring phenomenon where the micro-
voids in a material rapidly link up to form micro-cracks that propagate throughout the
material causing sudden failure. Void coalescence depends on many factors such as the
initial porosity, stress triaxiality, void size, shape, spacing and material hardening. There
are three idealized mechanisms of void coalescence: (i) primary void impingement, and
failure of the inter-void ligament due to (ii) shearing, or (iii) necking. It is generally
agreed that void coalescence is a combination of ligament necking and shearing
13
(Barsoum and Faleskog, 2007a, b) at low to moderate stress triaxialities with necking
coalescence occurring exclusively at high triaxialities. Shear coalescence (Tvergaard,
1982; Faleskog and Shih, 1997; Barsoum and Faleskog, 2007) has been attributed to the
formation of shear bands on at the micro-scale and is favored at low stress triaxiality,
low strain bi-axiality and low hardening. Coalescence by primary void impingement is
theoretically possible but instability and failure of the ligament would occur before the
voids would actually come into contact. Void coalescence due to inter-ligament shearing
and necking are shown in Figures 1.5 and 1.6, respectively.


(a) (b)
Figure 1.5: (a) Triangular array of cylindrical holes in AA-5052 sheet and (b) coalescence
due to shear localization under tensile loading (Weck, 2006). The loading direction is
vertical and the holes have an initial diameter of 10 m.


(a) (b)
Figure 1.6: (a) Rectangular array of cylindrical holes in AA-5052 sheet and (b) coalescence
due to localized necking under tensile loading (Weck, 2006). The loading direction is
vertical and the holes have an initial diameter of 10 m.

14
The fracture surface caused by ligament necking (Figure 1.7a) exhibits the classic
dimpled structure associated with ductile fracture while shearing exhibits a smoother
surface (Figure 1.7b) as the shear stress rotates and smears out the voids. Inspection of
the fracture surfaces reveals the presence of many void sizes indicating that coalescence
involves secondary populations of smaller voids (Tvergaard, 1982; Sun, 1995, Brocks et
al., 1996; Faleskog and Shih, 1997; Perrin and Leblond, 2000; Fabregue and Pardoen,
2009). These smaller voids nucleate between larger voids and link up to cause rapid
failure by compromising the integrity of the ligament and hastening its collapse as seen
in Figure 1.8. The investigation of the role of secondary voids on ductile fracture has
become an active area of research since coalescence within void clusters can no longer
be ignored (Fabregue and Pardoen, 2008). Aluminum-magnesium alloys are particularly
sensitive to this mechanism since fracture can be triggered by the coalescence between
two or three primary void clusters (Worswick et al., 2001; Chen, 2004).


Figure 1.7: Scanning electron microscope fractographs of Weldox 420 steel illustrating two
different rupture mechanisms (Barsoum and Faleskog, 2007b).


15

(a) (b) (c)
Figure 1.8: (a) Dual population of large and small voids at beginning of deformation; (b)
Activation and growth of small voids in a region of high stress surrounding the large
voids; (c) Fracture as the voids link up throughout the material.

.1 Modeling void coalescence

Coalescence models have typically been proposed based upon geometrical
considerations where coalescence occurs once a material specific value has been
reached. The critical geometric parameter could be the void radius, spacing and/or
porosity (McClintock, 1968b; Rice and Tracey, 1969; Brown and Embury, 1973; LeRoy
et al., 1981; Tvergaard and Needleman, 1984; Rousselier, 1987). The majority of work
on void coalescence has been performed by assuming the existence of a critical porosity.
Tvergaard and Needleman (1984) theorized void coalescence as the termination of
stable void growth and nucleation where the voids suddenly link-up throughout the
material resulting in failure. A critical porosity at coalescence, f
c
, and at failure, f
f
, were
theorized as intrinsic material parameters and used to identify the transition to
coalescence and subsequent loss of load bearing capacity.

Subsequent investigations by Becker (1987) and Koplik and Needlman (1988)
observed that the critical and failure porosities were dependent upon the material, stress
state and work hardening. Today, it is acknowledged that the critical porosity cannot be
considered a material constant because it is related to the initial porosity, void shape,
spacing, clustering, stress state, and strain (Zhang and Niemi, 1994a, b; Pardoen and
Delannay, 1998; Zhang et al., 2000; Pardoen and Hutchinson, 2000; Bandstra and Koss,
2008). While attractive for its simplicity, the assumption of a critical porosity is
unrealistic because it contains no physical foundation.
16
To account for the physical mechanism of void coalescence, Thomason (1985a, b,
1990) theorized coalescence as the competition between a stable, homogeneous
deformation mode and an unstable, localized deformation mode. The reasoning of
Thomason (1990) was that the influence of the voids on the material response is
negligible in the early stages of deformation and the material deforms homogenously.
The pre-existing voids will grow during this stage and additional voids will be nucleated
from inclusions and second-phase particles. Eventually, there is sufficient porosity to
influence the material response, reducing the energy required to transition to a localized
deformation mode within the void ligaments. The onset of coalescence is identified
when the energy required for both homogeneous and localized deformation modes
becomes equal. At this point, the plastic limit-load of the ligament has been reached and
deformation becomes localized in the ligament resulting in necking failure. Unlike the
previous phenomenological coalescence models, the plastic limit-load criterion was
rigorously derived from the void geometry and stress states associated with coalescence.

The pioneering work of Thomason (1990) has greatly contributed to the
understanding and modeling of internal necking coalescence and has received
widespread acceptance (Pardoen and Hutchinson, 2000; Zhang, 2000; Benzerga, 2002;
Zhang and Skallerud, 2010; Schyvaerts et al., 2010). The plastic limit-load of the
ligament has been reached when the following constraint is satisfied


( )
2
1
1
( , )
f
uc
C W
k _
_
o
E
> (1.12)

where 1
uc
k = for an axisymmetric unit cell and / 4
uc
k t = for a cubic cell. The plastic
constraint factor, C
f
, is a function of the void aspect ratio, W, and the void spacing ratio,
. Several variants of C
f
, have been proposed in the literature to account for different
void geometries. The constraint factor of Pardoen and Hutchinson (2000) is widely used
and is expressed as

2
1

f PH
C
W
_ |
o
_ _

| |
| |
= + |
|
|
\ .
\ .
(1.13)

17
where
2
0.1 0.22 4.8 n n o = + + and 1.24 | = .

A limitation of the Thomason (1990) and Pardoen and Hutchinson (2000) models is
that infinite ductilities are predicted for flat voids as 0 W . To remove this limitation,
Benzerga (2002) compiled the unit cell results of Golaganu (1997) who employed richer
velocity fields than Thomason (1990). A heuristic extension was introduced into the
model to obtain a plastic constraint factor of the form


2
1
2 1 2
1 1.3
0.1
0.1 0.02
f B
C
W
_
_ _ _


=
| |
+
|
+ +
\ .
(1.14)

The ligament size ratio or void spacing ratio is defined as the ratio of the lateral void
radius, R
x
, to the lateral void spacing, L
x
,

for a periodic arrangement of 3-D unit cells as
shown in Figure 1.9. The ligament size ratio evolves with the microstructure and is
related to the unit cell geometry as

1
3
x
x
R
L W
f
_

= =
| |
|
\ .
(1.15)

where is the aspect ratio of the unit cell and is a shape parameter specific to the
assumed cell geometry with 2/ 3 = for an axisymmetric unit cell and / 6 t = for a
cubic cell. Assuming the matrix material of the unit cell is incompressible, the evolution
of the unit cell aspect ratio, /
y x
L L = , can be expressed in rate form as


1
3
2
E = (1.16)

While void coalescence due to internal necking of the inter-void ligament can be
characterized using the plastic limit-load criterion, no such model exists for ligament
shearing or combined necking and shearing. Shear coalescence has often been observed
experimentally but few criteria have been developed aside from the work of McClintock
(1968b) and Richelsen and Tvergaard (1994). The importance of shear coalescence has
been overlooked since the numerical investigations of void growth and coalescence have
18
traditionally focused on a range of triaxiality of one to three where coalescence occurs
due to necking failure (Barsoum and Faleskog, 2007a, b).

2L
y
R
y
R
x
R
z
2L
z
2L
x

1

Figure 1.9: Cubic unit cell containing an axisymmetric elliptical void (
x z
R R = ) subjected
to uniaxial loading.

The development of void evolution and coalescence models in shear has gained
traction in recent years and is now an active area research area. Xue (2007) has modified
the McClintock (1968b) coalescence model to introduce a shear damage parameter for
coalescence. Alternatively, Butcher and Chen (2008) coupled the McClintock (1968b)
shear model with the plastic limit-load model of Benzerga (2002) to account for
combined shear and necking coalescence. Most recently, Schyvaerts et al. (2011) have
modeled internal necking coalescence in combined tension and shear, and Neilson and
Tvergaard (2011) have considered shearing between both primary and secondary voids
using a 2-D cell model.

1.3 Damage-based yield criteria

The void growth, nucleation and coalescence models discussed previously must be
coupled with a yield criterion to capture material softening and the evolution of the
stress state with progressive damage. Gurson (1977) was the first to propose a damage-
based yield criterion and flow rules for a porous ductile material based on the upper
bound theory of plasticity. The upper bound formulation was used to determine the
19
maximum macroscopic stresses required to sustain plastic flow. To simplify the
analysis, the material is assumed to have a periodic distribution of voids with each void
located at the center of a cylindrical or spherical unit cell. In this manner, the randomly
distributed voids in the representative volume are replaced with an equivalent void. In
reality, the void distribution is random with voids and particles of many sizes as shown
in Figure 1.10a. Gurson derived his model for both spherical (Figure 1.10b) and
cylindrical unit cells (Figure 1.10c). These voids resemble the shape of voids found
experimentally and serve to further simplify the analysis. Gurson defined porosity (void
volume fraction) in the unit cells as

Spherical:
3
3
a
f
b
= Cylindrical:
2
2
a
f
b
= (1.17a, b)

where a is the radius of the void and b the radius of the unit cell.
2
1
3

a
b
R

a
b
R
2
3
1


(a) (b) (c)

Figure 1.10: Void-matrix aggregate containing a random distribution of arbitrary voids
and its representation as a spherical and cylindrical unit cell.

Another advantage of Gursons model is its implicit accounting for material isotropy
as the voids are assumed to remain spherical or cylindrical during deformation. The
spherical model is geometrically isotropic while the cylindrical void experiences
transverse isotropy. Gursons yield criteria for spherical and cylindrical voids are

Spherical void:
2
eq hyd 2
2
3
2 cosh( ) 1 0
2
f f
o o
E E
u = + = (1.18a)
20
Cylindrical void:
2
eq hyd 2
2
3 3
2 cosh 1 0
2
f f
o o
| | E E
u = + =
|
|
\ .
(1.18b)

where o is the equivalent tensile flow stress in the matrix material, neglecting
variations in local stress, and
eq
E and
hyd
E are the effective macroscopic von Mises and
hydrostatic stress, respectively. Gursons formulation reduces to the von Mises yield
criterion for a damage-free material by setting 0 f = in Eq. (1.18). The relationship
between Gursons damage model and von Mises criterion with hydrostatic stress is
presented in Figure 1.11. Damage-based materials are sensitive to hydrostatic stress
because a tensile hydrostatic stress will expand the voids, softening the material and
resulting in earlier necking and failure. Conversely, a compressive hydrostatic stress will
increase ductility by closing the voids.


Figure 1.11: Gurson (1977) yield function showing the reduction in macroscopic equivalent
stress with increasing hydrostatic stress and porosity.

Void shape evolution is not considered in the Gurson model and the voids are
assumed to remain spherical or cylindrical during deformation. The spherical void
model is commonly employed since it better represents the ellipsoidal nature of voids in
a real material compared to the cylindrical geometry. Tvergaard (1981) introduced three
calibration parameters,
1
q ,
2
q and
3
q , into the Gurson (1977) yield criterion to obtain
21
better agreement with the numerical simulations of a voided material. For most ductile
materials,
1
1.25 1.5 q ~ ,
2
1 q ~ and
2
3 1
q q = are suitable choices. The Gurson-
Tvergaard (GT) yield function can be written as


2
eq hyd 2 2
1 2 1 2
3
2 cosh 1 0
2
fq q q f
o o
E E | |
u = + =
|
\ .
(1.19)

The GT model has been widely adopted in the literature and provides the basic
framework for damage-based modeling and many extensions and improvements have
been proposed. Notable improvements to the Gurson model to include the addition of
anisotropy (Liao et al., 1997; Chien et al., 2001; Wang et al., 2004), non-local damage
development (Leblond et al., 1994; Tvergaard and Needleman, 1995; Reusch et al.,
2003; Drabek, 2005) and most importantly, the incorporation of void shape (Gologanu
et al., 1993, 1994), size effects (Wen et al., 2005) and material anisotropy (Benzerga et
al, 2004). The latest advanced models of ductile fracture (Pardoen and Hutchinson,
2000; Benzerga, 2002; Lassance et al., 2006; Fabregue and Pardoen, 2008; Scheyvaerts
et al., 2011) are based upon the Golaganu-Leblond-Devaux (GLD) model which is a
Gurson-type yield criterion that explicitly accounts for void shape effects. The GLD
model is based upon a periodic distribution of spheroidal voids embedded within a
spheroidal unit cell and is expressed as


2 2 h h
( )
2( 1)( ) cosh ( 1) ( ) 0
zz xx
C
g g f g g f
q k
o o
E E + E E (
u = + + + + + =
(

(1.20)

h 2 2
2 (1 2 )
xx zz
o o E = E + E

2
2
2
4 3 1 -3
=
2 2
3 1
e gG
g C
F H F
e
_
k q
_
= = =


2 2
2 1 2
2 2 2 3
2 2
1 1
sin
2 2
e e
e
e e
o


= +

where f is the porosity, o is the flow stress,
xx yy
E = E and
zz
E are the axisymmetric
principal stresses; _ is a constant,
2
e is the eccentricity of the ellipsoidal matrix and
F , G and H are some mean values that can be assumed to be constants. The
complete model is not described here for brevity but the increasing complexity
22
associated with including void shape effects compared to the original Gurson model in
Eq. (1.19) is evident. The GLD model represents a marked improvement over the
Gurson model by accounting for the inter-relationships between void growth, shape,
triaxiality and material softening. However, with increasing accuracy comes increasing
complexity and the GLD model is difficult to implement into commercial finite-element
codes and its formulation is not easily amenable to arbitrary loading conditions. Similar
to its predecessors, the GLD model assumes the existence of a periodic arrangement of
voided unit cells and thus cannot reflect interactions within void clusters or the role of
the void distribution on fracture.

1.4 Void evolution and coalescence within clusters

The initiation, evolution and coalescence of voids within a material is ultimately
governed by the stress state and the void and particle distribution. It is well known that
voids and particles that are in close proximity to one another will interact and promote
rapid growth and coalescence. While void evolution will be accelerated within these
clusters, it is the coalescence mechanism that is the most sensitive to the distribution.
Coalescence is a direct consequence of the void distribution because coalescence cannot
occur unless there is a neighbour to coalesce with. Successful crack propagation relies
upon a favorable network of nearby voids to link-up with throughout the material. The
presence of heterogeneous particle and void clusters provides the perfect conditions for
a local crack to form and propagate to adjacent clusters. From this perspective, the onset
of coalescence within a single cluster can set the stage for plastic instability and fracture.

Numerical and experimental studies of void clusters typically involve introducing a
series of holes into a sample to create a desired void distribution and stretching it to
failure (Goto and Koss, 1996; Geltmacher et al., 1998; Bandstra and Koss, 1998, 2008;
Toi and Kang, 2005; Weck, 2006; Jones et al., 2007; Hu et al., 2008). The influence of
the void distribution and the stress state is clearly shown in Figure 1.12 from
Geltmacher et al. (1996) where the fracture path is different in biaxial tension because it
activates coalescence paths in two directions. In uniaxial tension, coalescence occurs
only in ligaments that are transverse to the loading direction. These studies are
23
informative and can provide insight into the behaviour within void clusters. However,
the experimental tests are limited to approximating voids as cylindrical holes that are
significantly larger than the actual voids.

Figure 1.12: Experimentally observed flow localization paths of Geltmacher et al. (1996) in
test specimens with a minimum hole spacing of two hole diameters and loaded in uniaxial
tension (left) and equal-biaxial tension (right).


Figure 1. 13: Thomsons 3D unit cell model used to model the three void cluster types, (a)
linear, (b) planar, and (c) quasi-spherical (Thomson et al. , 1999).

Another avenue to characterize void evolution within clusters is by creating unit cells
that contain multiple voids as shown in Figure 1.13. (Thomson, 1999; Horstemeyer et
al., 2000; Zhang and Chen, 2007). These simulations are valuable for determining the
influence of the cluster geometry on void growth and coalescence but are not
representative of the true microstructure as it assumes a periodic arrangement of void
24
clusters. As Figure 1.12 has demonstrated, coalescence in a random distribution of voids
does not resemble that of a unit cell. These types of simulations are only useful to
visualize deformation within clusters or to obtain benchmarks for the development
and/or calibration of void evolution and coalescence models.

1.5 Damage percolation modeling

.1 Role of the void distribution

The over-arching theme of this literature review has been to emphasize that the
assumption of a periodic void distribution is employed in all facets of this research area
to model ductile fracture. Significant efforts have gone into refining these models and
much progress has been made. However, there is a natural limit to the accuracy of these
models because they cannot properly account for the physical mechanism of the failure
mechanism: void initiation and evolution within heterogeneous clusters. A cynical
perspective is that these models are no better than a phenomenological continuum
damage model that has the same number of adjustable parameters. The continuum
damage model is also not constrained by ensuring that the parameters are realistic, just
that they work. Certainly a void-based constitutive model is more realistic but it is
debatable that is more useful. It is difficult to defend the advantages of a void-based
constitutive model compared to a continuum damage model except from the academic
argument of scientific purity. To obtain physically realistic fracture predictions one must
first have a physically realistic model.

.2 Previous damage percolation models

The limitation of homogenous damage models can be overcome using a technique
known as damage percolation modeling developed by Worswick et al. (2001) and
extended by Chen (2004) and Butcher and Chen (2009). In the damage percolation
model, digital imaging techniques are used to obtain the actual particle distribution in a
material (Figure 1.14). The image tessellation algorithm approximates the
void/particle/inclusion as an ellipsoid and extracts the particle size, shape, location, and
25
cluster and nearest neighbour lists. Using this information, micromechanical models are
applied to characterize void and crack formation leading to failure at the individual
particle scale (Figure 1.15 and 1.16). In this approach, all three stages of ductile fracture
are captured with damage originating within heterogeneous particle clusters.
Furthermore, ductile fracture in the percolation model naturally accounts for the
presence of multiple void sizes, shapes and clusters which cannot be included using a
traditional damage model. The damage percolation model is a powerful technique in
material modeling as it directly relates changes in the local microstructure to the overall
material behaviour.


Figure 1.14: Tessellated AA5182 particle field corresponding to a size of 0.75 mm x 0.75
mm. The image was obtained from a large scale plane-view image (Chen, 2004). The
rolling direction is horizontal.






26

Figure 1.15: Predicted damage prior to profuse coalescence within a particle field of
AA5182 sheet under plane strain loading at 23.9% major strain (Butcher and Chen, 2009).
The grey ellipses are second phase particles, the red ellipses are nucleated voids and the
black ellipses represent voids or cracks formed by coalescence. The rolling and loading
directions are horizontal.



Figure 1.16: Comparison of the predicted average diameter of the broken particles with
the experiment data of Hadianfard et al. (2009) using the 2-D percolation model of Butcher
and Chen (2009). The experimental range for the broken particles is presented as a shaded
band and three particle fields were evaluated in the model for each load ratio.

The previous percolation models of Worswick et al. (2001) and Chen et al. (2004)
showed great promise in predicting damage-induced ductile fracture but the framework
27
requires further development and validation before it can make deterministic predictions
of fracture. To date, the percolation models have been restricted to two-dimensional
particle fields and do not account for void orientations and neglect the stress state which
is critical for modeling void nucleation, growth and coalescence. Furthermore, the stress
state evolves considerably during deformation as a consequence of both (i) material
hardening and (ii) softening due to the presence of voids and geometric considerations.
Since the stress state is not considered, all aspects of these percolation models are strain-
controlled. This assumption significantly impedes the ability of the model to
quantitatively predict fracture under general loading conditions because strain-based
measures cannot uniquely define the stress state.

To apply the strain-based percolation model to a metal forming operation, Chen
(2004) extracted the nodal displacements from a finite-element simulation and mapped
them onto a large particle field for use in the percolation model. In this manner, the
strain-gradients throughout the material could be recreated within the particle field with
the percolation model acting as post-processor for the finite-element model. This
approach demonstrated the vast potential of the percolation model as localized damage
within heterogeneous clusters was reflected in a metal forming operation (Figure 1.17).
The approach of Chen (2004) represents a loose-coupling between the micro- and
macro-mechanical behaviour of the material and serves as the motivation for this work
as we seek to develop a fully-coupled and predictive damage percolation model.


Figure 1.17: Predicted damage within an AA5182 aluminum particle field during a stretch
flange forming operation. The stretch flange finite-element model is axisymmetric and a
two-dimensional particle field was used (Chen et al., 2003).

28
The percolation model of Chen et al. (2004) was extended by Orlov (2006) to include
a three-dimensional particle field obtained using x-ray micro-tomography. The
experimental tomography data was used to control damage evolution in the percolation
model to obtain predictions consistent with the experiment. The percolation model of
Orlov (2006) should be considered a proof-of-concept rather than a predictive model
because simple micromechanical models were employed and its performance is mainly
a function of using the experiment data to recreate the damage distribution. This is an
effective calibration that offsets its limitations. A damage percolation model that
employs a rigorous treatment of each stage of void evolution is required to predict the
experimental data of Orlov (2006) from first principles before the model can be
considered predictive.

1.6 Outline of Thesis: A multi-scale damage-percolation model of ductile
fracture

The objective of this work is to develop a multi-scale modeling strategy for
heterogeneous media where the evolution of the microstructure can be related to the
behaviour of the bulk material. The role of the heterogeneous particle distribution will
be accommodated by developing a sophisticated damage percolation model. Previous
percolation models have shown great promise but relied upon numerous assumptions
and modeling techniques so that they are qualitative in nature and not predictive. The
present percolation model will be designed with an emphasis on developing physically-
motivated evolution rules and sound modeling practices to obtain a deterministic and
predictive model. The stress state will be captured by integrating the percolation model
into a commercial finite-element code as shown in Figure 1.18. This research will
generate a cutting-edge research and design tool which will enable manufacturing
engineers to both simulate and optimize any type of sheet metal forming operation from
the micro-to-macro scale by modeling the actual microstructure of the material.


29

Finite-element model of
a notched tensile test
Finite-elements are mapped onto a
measured particle field
Micromechanical models are applied
to each particle to characterize local
void initiation and evolution leading
to fracture
Finite-element model of
a notched tensile test
Finite-elements are mapped onto a
measured particle field
Micromechanical models are applied
to each particle to characterize local
void initiation and evolution leading
to fracture
1
1

1
1
1

1

Figure 1.18: Multi-scale percolation model applied to a finite-element simulation of a
tensile test.

The development of the multi-scale percolation model and its application is
organized into the following chapters. Chapter 2 deals with the validation and
calibration of the analytical models for void evolution using finite-element simulations.
Chapter 3 is concerned with implementing a homogenization theory from particle-
reinforced plasticity into an existing damage-based framework to estimate the stress
within the inclusions. Chapter 4 provides the theoretical framework for the percolation
model. Chapter 5 will focus on the generation of representative particle distributions for
use in the percolation model. The complete percolation model is applied to a notched
tensile test in Chapter 6 and the predicted fracture strains and porosity trends are
validated with the available experiment data. Finally, Chapter 7 will propose a novel
non-local modeling strategy to account for void evolution within clusters.






30
Chapter 2 Modeling Void Growth to Coalescence:
Validation and Calibration

2.1 Establishing confidence in the analytical void evolution models

The physical foundation of a damage percolation model rests upon the accuracy of
the analytical sub-models that describe void nucleation, evolution, coalescence and
material softening. While numerous micromechanical models have been proposed in the
literature that can capture damage evolution, at least qualitatively, the veracity of these
models must be first evaluated over the large number of void shapes, sizes and stress
states that will be found within a real microstructure. Finite-element simulations of
axisymmetric unit cells containing different void geometries are required to provide the
numerical benchmarks for these evolution models. An emphasis will be placed upon the
evolution of penny-shaped voids as they are nucleated by particle cracking and have
received scant attention in the literature. It is imperative that the analytical evolution
models agree with the finite-element simulations for a voided material to establish
confidence in the predictions of the damage percolation model.

The organization of the chapter is as follows:

- Finite-element modeling and analysis of voided unit cells
- Discussion of finite-element simulations for a range of void aspect ratios and
initial porosities
- Evaluation of current micromechanical models to predict damage evolution
- Development of unit cell correlations for damage evolution
- Refinement of the yield criterion and evaluation of the complete constitutive
model to analytically predict void evolution and coalescence




31
2.2 Modeling void evolution in a unit cell

The voids within a bulk material are assumed to be distributed throughout the
material in such a manner that the material may be considered to be composed of
identical unit cells, each containing a single void at its center. Typically, the unit cells
are assumed to be cubic or axisymmetric. The axisymmetric unit cell is obtained by
assuming the material is composed of interlocking hexagonal unit cells. These
hexagonal unit cells can be approximated as cylindrical and then reduced to a one-
quarter, 2-D geometry for finite-element modeling as shown in Figure 2.1. Similarly, the
geometry of a 3-D cubic unit cell can be directly reduced to a one-eighth model due to
symmetry.


Figure 2.1: Procedure for approximating a hexagonal distribution of unit cells into an
axisymmetric cell model that can be reduced to a one-quarter, 2-D geometry for finite-
element modeling.

The predicted trends for damage evolution are somewhat dependant upon the type of
unit cell geometry (Kuna and Sun, 1995). The evolution of porosity is lowest in a 3-D
model and highest in the axisymmetric model due to the 3-D material containing more
material to constrain the growth of the void and delay coalescence. As seen in Figure
2.2, the cubic unit cell requires a significantly larger number of elements than the
axisymmetric model and requires 8-node solid brick elements that are significantly more
computationally expensive than the 4-node quadrilaterals used in the axisymmetric
model. As a result, the run-times for cubic cells tend to be prohibitive unless a specific
loading condition or geometry mandates their use. In the literature, most unit cell studies
have focused on the use of axisymmetric geometries as they are efficient and provide a
32
good estimate for damage evolution and fracture strains (Koplik and Needleman, 1988;
Kuna and Sun, 1995; Brocks et al., 1995; Pardoen and Hutchinson, 2000). It should be
emphasized that axisymmetric models provide a lower, and therefore, more conservative
estimate for the fracture strain (coalescence) than their 3-D counterparts. Axisymmetric
unit cells are considered in the present work to generate the void evolution trends
required for validation of the micromechanical models.


Figure 2.2: Typical cell geometry and finite-element mesh for an initially spherical void in
a one-quarter axisymmetric unit cell (left) and a one-eight cubic unit cell (right).

2.3. Analysis of an axisymmetric unit cell

The axisymmetric unit cell is analyzed using a cylindrical coordinate system denoted
by (
r
e ,
u
e ,
z
e ) along the r-radial, -ortho-radial and z-axial axis, respectively. As shown
in Figure 2.3, the geometry of the cylindrical unit cell is defined by a height of 2L
z
and
radius, L
r
and contains a spheroidal void with radii, R
z
and R
r
. The void shape is a
prolate spheroid when R R
z r
> , an oblate spheroid when R R
z r
< , or a sphere when
R R
z r
= . The geometry of the unit cell and void are defined using non-dimensional
parameters. The void is defined by its volume fraction or porosity, f, and its aspect ratio,
W, while the cell geometry is defined by its aspect ratio, . These parameters can also be
combined to calculate the void spacing ratio (ligament size ratio), , that plays a critical
role in void coalescence. These parameters all evolve during deformation of the cell and
their initial values are defined as
33

2
0 0
0
0 0
2
3
z r
z r
R R
f
L L
| |
=
|
\ .

0
0
0
z
r
R
W
R
=
0
0
0
z
r
L
L
=
1/3
0 0
0 0
0 0
3
2
r
r
f
R
L W

_
| |
= =
|
\ .
(2.1-2.4)

Figure 2.3: General geometry and boundary conditions used to model the one-quarter
axisymmetric unit cell.

.1 Boundary conditions

To enforce the assumption of a periodic array of unit cells, the faces of the unit cell
must remain straight and move as rigid planes during the deformation process so that the
cell remains cylindrical. The faces of the cell at L
r
r = and L
z
z = must have normal
displacements and retain their mutual orientations. This condition is easily accomplished
for one-quarter 2-D unit cell by designating the upper-right corner node as a master node
denoted as A in Figure 2.3 to control the displacements of the radial and axial planes
by using multipoint constraints. These constraints enable the lengths of the unit cell to
be defined at an arbitrary state as


A
0 r r r
L L u = +
A
0 z z z
L L u = + (2.5)

The formal displacement boundary conditions of the one-quarter unit-cell geometry
are stated as

34
0
r
u = along the axis 0 r = , 0
z
z L s s
0
z
u = on the bottom 0
r
r L s s , 0 z = (2.6)
A
r r
u u = on the lateral surface
r
r L = , 0
z
z L s s
A
z z
u u = on the top 0
r
r L s s ,
z
z L =

The boundaries of the cell are free of shear tractions and the surfaces of the void are
free of all tractions. Additionally, the matrix material is assumed to be a pure matrix
containing no secondary voids or particles. A void nucleation rule governing the
formation of secondary voids is not considered as the unit cell results will be dependant
upon that specific nucleation model and its assumed parameters. The above conditions
are applicable to all of the void geometries and stress states considered in this chapter.

.2 Stress state and microstructure evolution

The microscopic stress and strain tensors describe the stress state within the unit cell
and are denoted by
ij
o , and
ij
c . The macroscopic stress and strain tensor are applied at
the cell boundaries and are denoted as
ij
E and
ij
E , respectively. Since the cell surfaces
are free of shear, the applied stress and strain directions coincide with the principal
directions. The macroscopic principal and equivalent strains are given as


A
r
0
ln 1
r
r
u
E
L
| |
= +
|
\ .

A
z
0
ln 1
z
z
u
E
L
| |
= +
|
\ .

eq
2
3
z r
E E E = (2.7-2.9)

The remote true macroscopic principal stresses are calculated at any instant as the
average reaction force at the cell faces per current area through


r
0
1
z
r Lr
L
r
z
S dz
L
=
E =
}

z
2
0
2
r
z Lr
L
z
r
rS dr
L
=
E =
}
(2.10-11)

where S is the stress vector. The corresponding macroscopic hydrostatic stress,
hyd
E ,
equivalent stress,
eq
E , and stress triaxiality, T, are calculated as


1 2 3
hyd
3
E +E +E
E =
eq z r
E = E E
hyd
eq
T
E
=
E
(2.12-2.14)

35
The analysis can be simplified by defining a loading parameter, o , that can be
determined for a specified stress triaxiality as


3 1
3 2
r
z
T
T
o
E
= =
E +
(2.15)

where 0 o = for uniaxial tension ( 1/ 3 T = ) or -0.5 for pure shear ( 0 T = ).

The plastic strain increment,
p
c , and flow stress, o , within the unit cell can be
determined from work equivalence as

(1 )
p p
ij ij
f E c o =E (2.16)

by neglecting the elastic strains and assuming the matrix is incompressible so
that
p
ij ij
E E ~ . The microscopic plastic strain and flow stress of the unit cell are related by
the stress-strain relation of the matrix. In the present study, the ductile matrix is assumed
to be an isotropic, rate-independent material that hardens according to the power-law
relation

y y
1
n
p
E o
c
o o
| |
= +
|
\ .
(2.17)

where
y
o is the initial yield stress, E is the Youngs modulus and n is the hardening
exponent.

The void volume fraction of the unit cell can be computed from the updated
coordinates of the element nodes that define the surface of the void or from an
approximate analytical formula proposed by Koplik and Needleman (1988) using the
cell volume, V, as

e
0
0
1 (1 )
V V
f f
V V
A
= (2.18-20)

e
0 hyd
3(1 )
(1 )
V v
f
V E
A
= E
A 2 A
0 0
2
0 0
0
( ) ( )
r
r z z
z
r
L u L u V
V L L
+ +
=

36
It was confirmed that the approximate analytical method gave very good agreement
with numerical integration of the void surface as observed in previous studies (Koplik
and Needleman, 1988; Pardoen and Hutchinson, 2000; Siad et al., 2008) and will be
adopted in the present work as well to reduce the amount of post-processing. It should
be noted that this approximation in Eq. (2.18) is not valid for a porous matrix material
(secondary voids). The void aspect ratio, cell aspect ratio and the void spacing ratios can
be readily determined from the geometry using Eqs. (2.2 - 2.4).

.3 Identification of void coalescence

The termination of homogeneous deformation within the unit cell is marked by the
onset of void coalescence when deformation becomes unstable and localized within the
inter-void ligament while the material outside the ligament unloads elastically as shown
in Figure 2.4. The transition to void coalescence is identified by the transverse strain rate
approaching zero as the unit cell deforms in a type of uniaxial stretching mode,
regardless of the stress triaxiality. In this regime, the void expands rapidly in the
transverse direction as the ligament length approaches zero and the load-bearing
capacity of the material is lost. Physically, this represents the collapse of the ligament as
the void is beginning to link-up with a neighbouring void from an adjacent unit cell to
form a larger void. However, in the finite-element simulation, the ligament will shrink
and become highly deformed as the length approaches zero, ultimately requiring a re-
meshing scheme to model the post-coalescence regime.

In the present study, the finite-element simulations are terminated after the onset of
void coalescence as the results of the model in the post-coalescence regime are not of
prime importance. For the practical range of stress triaxiality found in sheet metal
forming operations ( 1 T s ), the collapse of the ligament occurs almost instantaneously
and thus the coalescence strain can be used as an excellent approximation to the fracture
strain. The procedure for identifying the onset of coalescence using the radial strain rates
is demonstrated in Figure 2.5 and the associated loss of load-bearing capacity with
coalescence is illustrated in Figure 2.6.

37

Figure 2.4: Typical deformation history of a voided unit cell. Initially, the deformation
mode is homogeneous and characterized by stable void growth until the onset of
coalescence when the deformation mode becomes unstable and localized within the
ligament. Note the abrupt change in the transverse growth of the void during coalescence
compared to the vertical displacement of the cell. These results were obtained for an
initially spherical void with an initial porosity of 0.1% and subjected to a stress triaxiality
of unity.


Figure 2.5: Typical relationship of the macroscopic radial and axial strains for a voided
unit cell. The onset of void coalescence is identified when the radial strain rate is zero.

38


Figure 2.6: Typical macroscopic stress and strain curve for a voided unit cell. Note that the
loss of load-bearing capacity of the material is approximately linear in the post-coalescence
regime.


2.4 Numerical solution procedure

To obtain general trends for damage evolution and coalescence within a unit cell, the
stress triaxiality must be kept constant throughout the cell during deformation. Since the
compressibility of the cell varies during deformation, the stress triaxiality will steadily
increase and fluctuate unless a control scheme is used to adjust the displacements of the
cell faces. This problem becomes acute near the onset of void coalescence where
deformation becomes unstable within the inter-void ligament. Fortunately, using
ABAQUS finite-element software, the modified Riks-algorithm (Riks, 1979) is ideally
suited to incrementing the load on the unit cell to maintain equilibrium. The Riks
method is an iterative scheme that allows equilibrium solutions to be found for
nonlinear, unstable problems such as buckling or an abrupt change in the deformation
mode.

The Riks-option within ABAQUS is employed to determine the displacements on the
unit cell faces in such a way to enforce the loading parameter, o , in Eq. (2.15) to be a
39
constant prescribed value based upon the specified stress triaxiality. The axial stress,
z
E ,
is set to unity and the corresponding radial stress is
r
o E = . Consequently, the axial
direction is the principal loading direction for all simulations in the present study since
1 o < for the range of stress triaxialities considered (1/ 3 3 T s s ). Convergence studies
were performed to test for mesh dependence and it was found to be minimal if the
element shape within the ligament is rectangular and the mesh is sufficiently refined
near the void surface as seen in Figures 2.2 and 2.4. The accuracy of the unit cell results
was verified through comparison with previous cell studies in the literature by Pardoen
and Hutchinson (2000), Lassance et al. (2008) and Scheyvaerts et al. (2010).

2.5 Unit cell simulation results

A large number of finite-element models have been developed with an emphasis
placed on small void aspect ratios (W <1) that best resemble the conditions observed in
particle cracking and debonding. Typical void shape evolution models are only valid for
an aspect ratio greater than 1/6 while voids nucleated by particle cracking have an aspect
ratio close to zero. The deformation of the void at small aspect ratios is essential for use
in the percolation model as real particles will nucleate voids via this mechanism. To best
represent an isolated void within a real material the initial porosity of the voids is taken
to be 0.01% and 0.1%. Finally, the simulations are stopped shortly after the onset of
coalescence as mesh distortion effects become significant, especially for initially flat
voids. Fortunately, the load drop, the void growth rate and shape evolution all become
linear within the post-coalescence regime and their behaviour can be reliably
extrapolated to estimate their value at fracture (Scheyvaerts et al., 2010).

The aspect ratios considered in the analysis are: 0.001, 0.01, 0.05, 1/6, 1 and 6 for
porosities of 0.1% and 0.01% and a hardening exponent of 0.10. To enable the results to
be generalized to any material, the parameters in the flow stress relation for the material
in Eq. (2.17) are
y
/ 500 E o = , which is applicable to a wide range of engineering
materials (Pardoen and Hutchinson, 2000). The stress triaxialities range from 1/3 to 3 to
40
capture void evolution in stress states ranging from uniaxial tension to the severe
conditions found ahead of a crack tip.

.1 Penny-shaped voids: W
o
= 1/100

The evolution of the porosity and the aspect ratio of an initially penny-shaped void
for a wide range of stress triaxialities are presented in Figures 2.7 and 2.8, respectively.
What is most remarkable about these figures is that the voids grow exceedingly fast and
in an approximately linear fashion in each stress state. In uniaxial tension (T = 1/3), the
void enlarges from 0.1% to nearly 9% while the aspect ratio changes by a factor of over
100 and is roughly spherical at fracture. The linear nature of the penny-shaped voids
evolution is attributed to it opening in the load direction and experiencing negligible
lateral growth. As most nucleated voids are initially penny-shaped, the proper modeling
of this geometry is of prime importance for the percolation model.


Figure 2.7: Porosity history of a penny-shaped void subjected to various triaxial loadings.

41

Figure 2.8: Evolution of the void aspect ratio of an initially penny-shaped void subjected to
various triaxial loadings.

.2 Oblate voids: W
o
= 1/6

The growth and shape evolution of an initially oblate void is presented in Figures 2.9
and 2.10, respectively. The expected non-linear response of the voids begins to emerge
where void growth is slow at low triaxiality and rapid at high triaxiality. The high
growth rate at high large stress triaxiality occurs because the void has sufficient height
to be expanded by the lateral stress. This is in contrast with the initially penny-shaped
void that is roughly independent of the hydrostatic stress.

42

Figure 2.9: Porosity history of an initially oblate void subjected to various triaxial
loadings.


Figure 2.10: Evolution of the void aspect ratio of an initially oblate void subjected to
various triaxial loadings.

Conversely, in uniaxial tension, the void becomes prolate and the void growth rate is
very slow compared to the penny-shaped case. Prolate voids do not grow fast at low
triaxiality because plastic flow becomes easier as the voids become more prolate and
43
needle-like as the voids laterally contract. Eventually, the void essentially closes itself
off and the matrix can deform almost as if the void were not present.

.3 Spherical voids: W
o
= 1

The well-known void evolution trends for spherical voids are presented in Figure
2.11 and 2.12. The void growth trends for spherical voids are similar to that of the oblate
void with the difference in the growth rates between stress states more exaggerated. The
aspect ratio shows a new development compared to the previous cases in that the void
becomes oblate at high stress triaxiality. The height of the void is sufficiently large for
the hydrostatic stress at high triaxiality to cause the void to grow laterally and cause the
aspect ratio to decrease despite the principal loading direction being vertical. This is a
counterintuitive but well documented phenomenon for voids at high stress triaxiality and
highlights the difficulties in developing a shape evolution model because the void grows
laterally when one would expect it to grow vertically.


Figure 2.11: Porosity history of an initially spherical void subjected to various triaxial
loadings.

44

Figure 2.12: Evolution of the void aspect ratio of an initially spherical void subjected to
various triaxial loadings.

.4 Prolate voids: W
o
= 6

Finally, the void growth and shape evolution for initially prolate voids are presented
in Figures 2.13 and 2.14. The trends are again similar to the spherical case with the
influence of the stress state more dramatic. What is most apparent is that the growth of
the voids is completely negligible in uniaxial tension as the void becomes extremely
prolate but stays relatively the same size. Void growth is also very small at a triaxiality
of 2/3 which corresponds to equal-biaxial tension. Since the void is initially prolate and
taller than it is wide, it is the most susceptible to lateral growth at high triaxiality and the
void shape becomes nearly spherical at fracture. From a modeling perspective, prolate
voids are not of paramount importance since their growth rates are so small in the
practical regime found in metal forming (T < 1).

45

Figure 2.13: Porosity history of an initially prolate void subjected to various triaxial
loadings.


Figure 2.14: Evolution of the void aspect ratio of an initially prolate void subjected to
various triaxial loadings.

.5 Selection of a minimum void aspect ratio

There are only so many finite-element simulations and unit cell geometries one can
consider and the argument can always be made for why different stress states or
46
geometries should have been included. Overall, a maximum initial aspect ratio of 6
seems reasonable for the modeling of prolate voids in the percolation model. Especially,
since most voids will be nucleated by particle cracking and debonding, they are best
described as penny-shaped. If penny-shaped cracks are the most important void
geometry then what is the minimum shape that can be considered? In the limit, the
penny-shaped void has no initial height, the aspect ratio is zero, and it cannot be
modeled using regular finite-element techniques. Mesh distortion issues also arise when
modeling increasingly small voids so there is a practical limit to the geometry one can
consider.


Figure 2.15: Comparison of void growth for different penny-shaped voids for the extreme
cases for the stress triaxiality.

Fortunately, the void growth rate and shape evolution for extremely flat voids are
essentially the same as shown in Figures 2.15 and 2.16, which is caused by deformation
primarily occurring in the opening direction of the void. The void opens in the loading
direction at a high enough rate that the solutions for aspect ratios of 1/1000 and 1/100
are very similar. This is an extremely fortuitous result as any void that is nucleated in
the percolation model by particle cracking can be assigned an arbitrary aspect ratio of
47
0.01, removing a parameter that otherwise would have had to been predetermined. A
similar result has also been reported by Lassance et al. (2007).

Figure 2.16: Comparison of void shape evolution for different penny-shaped voids for the
extreme cases for the stress triaxiality.

2.6. Theoretical models for void growth, shape and coalescence

With the benchmark trends for void growth and shape evolution established, the
validity of the analytical evolution models can be evaluated. It is important to state that
these models are all expected to perform at least reasonably well in some stress states, or
they would be without merit in the first place. However, for the purposes of percolation
modeling where the actual evolution of the voids in the material is to be described, the
performance of the models in a variety of conditions is required. This validation process
is required to establish confidence in the predictions of the percolation model.

.1 Yield criterion

The yield criterion of Gurson and Tvergaard (GT) described in Eq. (1.19) is adopted
to describe material softening of the material and will be used to integrate the stress state
in the analytical model. The extension of Ragab (2004a) is adopted to account for the
influence of the void shape in the GT model. In this extension, q
1
and q
2
parameters in
48
the yield function are no longer material constants but related to the stress state, void
shape and material hardening.

.2 Void growth, shape evolution and coalescence

The void growth and shape evolution model considered was proposed by Ragab
(2004a) as a set of semi-empirical equations obtained by a meta-analysis of the
definitive analytical and numerical studies of void growth and shape evolution in the
literature (McClintock, 1968; Rice and Tracey, 1969; Budiansky et al., 1982; Huang,
1991; Lee and Mear, 1992; Yee and Mear, 1996; Sovik and Thaulow, 1997; Pardoen
and Hutchinson, 2000). In his analysis, Ragab (2004) individually calibrated the q
i

parameters in the GT yield criterion as functions of the void shape, porosity, stress state
and hardening exponent. The calibrated (q
1
, q
2
) parameters not only enabled improved
modeling of an isolated void for use in the percolation model but also improved the
accuracy of the yield criterion by accounting for void shape effects while preserving the
relatively simple Gurson framework. This is very advantageous since other damage-
based models such as the GLD model and its variants (Gologanu et al., 1997; Benzerga,
2002) that account for void shape effects are difficult to implement compared to the GT
model. An approximate method for estimating the influence of the void shape on the
yield criterion is preferable to implementing these rigorous models.

.1 Void growth

The associated flow rule of the GT model yields the void growth relation


hyd
1 2 2
p
growth 1
1 hyd hyd
1 2 2
3
3 (1 ) sinh
2
3
3 sinh
2
f f q q q
f E
fq q q
o
o o
E | |

|
\ .
=
E E E | | | |
+
| |
\ . \ .
(2.21)

that was simplified by Ragab (2004a) by assuming small porosities and axisymmetric
deformation to obtain

49

growth hyd p
1 2 2 1
3 3
sinh
2 2
f
q q q E
f o
E | |
=
|
\ .
(2.22)

that was then calibrated to the void growth results available in the literature.

The semi-empirical equations of Ragab (2004a) for the variation of the q
i
parameters
in the GT model are valid for a large range of stress triaxiality from1/ 3 8/ 3 T s s , and
for hardening exponents up to 0.40. The q
i
relations are


2 3
1
q A BT CT DT = + + + (2.23)


2
2.28 3.55 3.84 A n n = +
2
0.92 1.32 0.32 B n n = +

2
0.53 2.31 2.35 C n n = +
2 3
0.10 0.27 0.70 1.78 D n n n = + +


2
q W
q
= (2.24)

( 1) 0.206ln( ) 0.266 0.02 W T n q < =
2 3 4
( 1) 3.484 11.614 13.72 6.54 1.06 0.2 W T T T T n q > = + + +

.2 Void shape evolution

For a constant strain-path, the evolution law of Ragab (2004a) for the void aspect
ratio is

( ) ( ) (
p
0 hyd 0 eq
ln / 1.1 2 / 1 ) W W n f E e o = + E + (2.25a)

where is defined for initially oblate (
0
1/ 6 1 W s s ) and prolate voids (
0
1 6 W s s ) as

( )
p
oblate 0 eq
ln( ) / 0.109 1.224 W E e = +
( )
p
prolate eq 0
0.535 0.0235 ln( ) E W e = + (2.25b, c)


.3 Void coalescence

A robust void coalescence criterion has been developed by Thomason (1990) to
predict the onset of internal necking coalescence between two adjacent voids as a
function of the microstructure geometry and stress state. The extension of Pardoen and
Hutchinson (2000) in Eq. (1.13) has been widely used in the literature as it accounts for
material hardening and has proven to give good coalescence predictions. However,
50
infinite ductilities are predicted by this variant of the plastic limit load for penny-shaped
cracks. Since penny-shaped voids will be used extensively in the percolation model, it is
of interest to also evaluate the model of Benzerga (2002) in Eq. (1.14).

2.7 Evaluation of the analytical void evolution models

.1 Void coalescence

It is of interest to first evaluate the performance of the coalescence models since
coalescence is predicted as a function of the void geometry. The coalescence model will
only be as accurate as the sub-models that describe the void geometry. The models of
Pardoen and Hutchinson (2000) and Benzerga (2002) were evaluated using the
geometry of the voids from the unit cell simulations to obtain their predictions under
ideal conditions. The predictions of each model with the numerical coalescence strains
are presented in Figures 2.17 and 2.18, respectively.

Figure 2.17: Comparison of the numerical and predicted coalescence strains using the
coalescence model of Pardoen and Hutchinson (2000) in Eq. (1.13).

Each coalescence model gives excellent agreement with the numerical coalescence
strain to a surprising level of accuracy. This is not entirely unexpected since these
models have been previously compared to unit cell simulations in the literature but not
51
for the number of different penny-shaped void geometries and stress states in the present
work. There is no clear advantage to using either coalescence model based upon the
present results. Even though the Pardoen and Hutchinson (2000) model accounts for
hardening but not flat voids, and the Benzerga model (2002) assumes a perfectly plastic
material, both models give very good predictions. The presence of flat voids does not
translate to infinite ductilities in practice since these voids rapidly deform to shapes that
are within the domain of the model. The strong agreement of both models attests to the
strong underpinnings of the Thomason (1990) model that they are based upon.


Figure 2.18: Comparison of the numerical and predicted coalescence strains using the
coalescence model of Benzerga (2002) in Eq. (1.14)

The important conclusion from this validation is that the coalescence models can
accurately describe void coalescence by internal necking. If the other analytical models
can predict void growth and shape evolution to a similar degree of accuracy, the
coalescence predictions will be sound and agree with that of the unit cell.




52
.2 Void growth

The analytical void growth rule was evaluated using the stress and strain data from
the unit cell simulations. The predictions of the model for penny-shaped cracks are
dismal as shown in Figure 2.19. In defense of the model, it was not designed for penny-
shaped voids and its limit for flat voids is an aspect ratio of 1/6. However, there is no
available void growth model for penny-shaped voids so one is left with few options.
This result highlights the importance of the validation process. Simply adopting this
model and extrapolating it to the penny-shape regime would lead to exceptionally poor
predictions for void evolution in the percolation model and these errors would go
undetected.


Figure 2.19: Comparison of the analytical void growth model with the unit cell data for an
initially penny-shaped void

The predictions of the analytical model for initial void shapes of 1/6, 1 and 6 are
presented in Figures 2.20-2.22, respectively. This is the regime that the model was
designed for and agreement is quite reasonable at low to moderate strains for a large
range of stress states and void shapes. This is not a trivial achievement. The semi-
empirical equations for void growth could certainly be used in a traditional damage-
based constitutive model where the porosity is representative of an equivalent void. In
53
this application, the model is meritorious as it offers a simple and closed-form
approximation that is valid for a wide range of conditions. However, for the percolation
model where individual voids are considered, the accuracy of the model leaves
something to be desired and will be calibrated in the proceeding section.


Figure 2.20: Comparison of the analytical void growth model with the unit cell data for an
initially oblate void


Figure 2.21: Comparison of the analytical void growth model with the unit cell data for an
initially spherical void.
54

Figure 2.22: Comparison of the analytical void growth model with the unit cell data for an
initially prolate void.

.3 Void shape

The predictions of the semi-empirical model of Ragab (2004a) for the void aspect
ratio for the range of initial shapes considered are presented in Figures 2.23-2.26. Unlike
the void growth model, the predictions for the initial penny-shaped void in Figure 2.23
are in very good agreement with the unit cell data. This is a surprising result since the
model was not calibrated for this regime. Overall, the model gives very good predictions
for the void aspect ratio for oblate, spherical and prolate voids in most stress states. The
shape evolution model is quite reasonable and suitable for use in a typical damage-based
constitutive model. However, a higher level of accuracy is desired for use in the
percolation model, especially at low stress triaxiality where void shape evolution is
considerable.

55

Figure 2.23: Comparison of the analytical void shape evolution model with the unit cell
data for an initially penny-shaped void.



Figure 2.24: Comparison of the analytical void shape evolution model with the unit cell
data for an initially oblate void.


56

Figure 2.25: Comparison of the analytical void shape evolution model with the unit cell
data for an initially spherical void.



Figure 2.26: Comparison of the analytical void shape evolution model with the unit cell
data for an initially prolate void.




57
2.8. Calibration of the void evolution models

.1 Void growth

The void growth model of Ragab (2004a) used a simplified version of the growth
rule obtained for the Gurson model. In the present work, we will use the original form of
the growth rule in Eq. (2.21) because unlike Ragab (2004a), the unit cell data has been
obtained first-hand and not taken from the literature. The q
2
parameter cannot be solved
for explicitly in Eq. (2.21) and a Newton-Raphson method is used to determine the q
2

value obtained at each time-step in the unit cell simulation. The calibrated q
2
value for
each void geometry and stress state were determined to create a library of highly
accurate void growth correlations. The calibrated q
2
values for several of the void shapes
considered are shown in Figures 2.27-2.29. The calibrated q
2
value converges to a value
of unity for a spherical void which is the value derived by Gurson (1977). The calibrated
void growth model is compared with the unit cell data in Figures 2.30-2.32 and the
agreement is excellent.


Figure 2.27: Calibrated q
2
parameter for an initially penny-shaped void subjected to
various stress triaxialities.

58

Figure 2.28: Calibrated q
2
parameter for an initially spherical void subjected to various
stress triaxialities.


Figure 2.29: Calibrated q
2
parameter for an initially prolate void subjected to various
stress triaxialities.


59

Figure 2.30: Comparison of the calibrated void growth model with the unit cell data for an
initially penny-shaped void.



Figure 2.31: Comparison of the calibrated void growth model with the unit cell data for an
initially spherical void.

60

Figure 2.32: Comparison of the calibrated void growth model with the unit cell data for an
initially prolate void.

.2 Void shape evolution

Although the semi-empirical equations of Ragab (2004a) give pretty good results for
the void aspect ratio, a calibration is required because void growth is tightly coupled to
the void shape through
2
q W
q
= . As a result of this exponential dependence, the
previous void growth calibration is in vain unless the aspect ratio is also highly accurate
because errors in the aspect ratio will be magnified in the growth rule. Unlike the growth
rule, the semi-empirical equation for the void shape is a pure correlation and its form is
not amenable to calibration. To ensure that the void aspect ratio is modeled to a high
degree of accuracy, a second-order polynomial was observed to be suitable to very large
strains. If higher strains are required, the curve is linearly extrapolated with the slope
selected to give good agreement to high strains. The high-strain regime is not of prime
concern in the percolation model as voids will tend to coalesce with their neighbours at
low to moderate strains. The correlation function for the aspect ratio is a function of the
equivalent strain and initial aspect ratio as

61

( )
2
2 eq 1 eq o eq
1 eq eq


l
l l l
W a E a E W E E
W b E E W E E
= + + s
= + >
(2.26)

where
l
E is the limit strain and a
1
, a
2
, and b
1
, are the calibration coefficients.

Care must be taken when calibrating the void aspect ratio because the trend for the
aspect ratio departs from that of an isolated void prior to coalescence as shown in Figure
2.33. Prior to coalescence, the void begins to interact with its neighbour, raising the
local stress triaxiality and flattening the void. These interactions must be avoided during
the calibration process as they are geometry dependent and the voids in the percolation
model are assumed to be isolated. The limit strain in the correlation is selected to ensure
interaction effects are avoided. The correlation functions for the different void
geometries are compared with the unit cell data in Figures 2.34-2.36, which shows that
they accurately predict the aspect ratio to high strains while avoiding interaction effects.


Figure 2.33: Comparison of the void aspect correlation function with the model of Ragab
(2004a) and the unit cell data. The limit strain is selected to void including the spurious
interaction effects prior to coalescence.

62

Figure 2.34: Comparison of the void shape correlation with the unit cell data for an
initially penny-shaped void.



Figure 2.35: Comparison of the void shape correlation with the unit cell data for an
initially spherical void.

63

Figure 2.36: Comparison of the void shape correlation with the unit cell data for an
initially prolate void.

2.9 Evaluation of the complete constitutive model

The final validation of this chapter is to evaluate the complete constitutive model
including the calibrated void evolution rules by integrating the stress state using the GT
yield surface. The stress state in the percolation model will be determined by this model
and these values will be used to evaluate the void evolution models. An advantage of the
calibration procedure used in the void growth model is that the q
2
parameter is intrinsic
to the GT yield surface and its calibration should also improve the yield surface to
capture material softening and void shape effects. The comparisons of the full
constitutive model with the unit cell data for an initially penny-shaped void are shown in
Figures 2.37 and 2.38. The full constitutive model gives very good agreement in
predicting the stress state, the porosity evolution and even coalescence. This is a
significant achievement since the integration of the stress state along with the void
evolution and coalescence models must all be working in concert to successfully
reproduce the cell results. The porosity distribution for an oblate void is shown in Figure
2.39 and the trends are also in good agreement.

64

Figure 2.37: Comparison of the constitutive model and the unit cell data for the maximum
principal stress for an initially penny-shaped void in uniaxial tension.


Figure 2.38: Comparison of the constitutive model and the unit cell data for the porosity
evolution of an initially penny-shaped void in uniaxial tension.

65
0.000
0.001
0.002
0.003
0.004
0.005
0.006
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
Equivalent macroscopic strain
P
o
r
o
s
i
t
y
Unit cell
Model
W
0
= 1/6 f
0
= 0.1% n = 0.10 T = 1/3

Figure 2.39: Comparison of the constitutive model and the unit cell data for the porosity
evolution of an initially oblate void in uniaxial tension.

2.10 Summary

A rigorous validation and calibration programme was undertaken to evaluate the
analytical models used to describe void evolution and coalescence. Finite-element
simulations of voided unit cells were performed to obtain benchmarks for the analytical
models for a wide range of stress states. Emphasis was placed upon penny-shaped voids
because they will be used extensively in the percolation model and the present evolution
rules in the literature do not apply to voids of this shape. The principal contributions of
this chapter are:

- The analytical models for void evolution were calibrated using the unit cell data
by developing a vast library of correlation parameters

- The semi-empirical void growth model of Ragab (2004a) was improved by
calibrating the model using first-hand unit cell data and extended to the penny-
shaped void regime

- The growth and shape evolution of isolated voids within the percolation model
66
will be representative of the actual case by virtue of the calibrated models

- Void coalescence can be accurately predicted using the plastic limit-load model
of Pardon and Hutchinson (2000) or Benzerga (2002)

- The GT yield criterion was improved by calibrating the q
2
parameter in the void
growth model. Although designed for spherical voids, this calibration enables the
model to give good results for penny-shaped to prolate voids.
67
CHAPTER 3
Integration of a Particle-based Homogenization Theory
into a General Damage-based Constitutive Model

3.1 Void nucleation in a damage-based constitutive model

Despite great strides in developing physically-motivated models for void growth,
shape evolution and coalescence, a suitable treatment for void nucleation remains an
open question. Accurate modeling of void nucleation is difficult within a Gurson-based
framework due to the intrinsic assumption that the material does not contain any second-
phase particles. Consequently, the nucleation models employed in these constitutive
models are overly simplistic as the particle shape, composition, stress state and load-
sharing are neglected, lumped into a single calibration parameter (Beremin, 1981) or
indirectly accounted for in a phenomenological manner (Chu and Needlman, 1980).
Furthermore, the growth and coalescence of existing voids is affected by the presence of
the particles since they alter the stress state within the material due to load-sharing with
the matrix. As such, there is a clear need for an efficient, albeit approximate method to
account for the role of the particles and inclusions on ductile fracture.

It is important to note that the lack of progress in developing physically sound
nucleation models is not for lack of effort but a result of the inherently complex nature
of the nucleation process. Void nucleation is very difficult to capture experimentally
since it is a relatively random and instantaneous event that cannot be captured in-situ
without the aid of high resolution x-ray tomography. Additionally, the local stress state
near a particle of interest is typically unknown as well as the particle composition and
mechanical properties. The nucleation mechanism can also occur by debonding or
particle cracking and is influenced by the particle size, distribution, strain rate and
temperature. From an engineering perspective, one can clearly see the attraction in
adopting a phenomenological nucleation model whose parameters can be adjusted to
give good agreement with the experiment data. Nevertheless, there is ample opportunity
to improve the physical foundation of the current nucleation models, especially in
68
regards to percolation modeling. For a damage percolation model to accurately predict
fracture, it cannot rely upon ad-hoc nucleation models since it considers each individual
particle within the microstructure.

The objective of this chapter is to develop a procedure to integrate a secant-based
homogenization technique for particle-reinforced plasticity into an existing damage-
based material model to determine the stress state within the second-phase
particles/inclusions. This model has been developed to predict ductile fracture of
industrial alloys during sheet metal forming operations where the loading is proportional
and the particle content is small. The proceeding sections of this chapter will introduce
the particle-based homogenization model and its integration into a general damage-
based constitutive model. The integration scheme is then applied to an enhanced GT
material model and the performance of the model is evaluated in several stress states
common to sheet metal forming. The ability to determine the stress state within the
particles will be used to model nucleation in Chapters 4 and 6. This chapter is organized
into the following sections:

- Brief discussion of homogenization theories for particle-reinforced plasticity

- Discussion of the particle-based homogenization theory used in the present work

- Development of the integration procedure for implementing the homogenization
scheme into a damage-based framework

- Implementation of the model into an enhanced GT material model

- Evaluation and application of the loosely-coupled model for a model aluminum
alloy

3.2 Review of particle-based homogenization theories

The development of particle-based homogenization theories to predict the bulk
behavior of a material from its constituents encompasses a large branch of materials
mechanics. As such, only a brief review of the development of homogenization
69
techniques is presented here with a focus on methods amenable for implementation into
a damage-based framework. Excellent reviews on the efforts to develop particle-based
homogenization techniques in the plastic-regime can be found in Ponte-Casteneda and
Suquet (1998) and Chaboche et al. (2005).

The framework for particle-based homogenization problems was pioneered by
Eshelby (1957) when he obtained closed-form solutions for the stress field within
ellipsoidal inclusions embedded in a matrix material by theorizing a stress-free
transformation strain (eigen-strain) between the matrix and inclusions. Eshelbys work
provided the analytical techniques required to estimate the average stress within the
inclusions and matrix material using the now famous, S tensor. Hill (1965) later
developed a rigorous solution for composite materials using the incremental theory of
anisotropic elasticity. Incremental approaches to the homogenization process are based
upon the tangent stiffness tensors of the constituents. It was soon recognized that this
approach led to overestimations in the flow stress of the material due to the anisotropic
nature of the tangent stiffness tensor in the plastic regime.

To obtain more realistic predictions for the composite flow stress, Berveiller and
Zaoui (1979) modified Hills incremental solution to obtain a total-strain formulation for
proportional loading using the secant moduli of the matrix material. This model was
further improved by Weng (1984, 1990) and Tandon and Weng (1986) using the mean
field methods of Mori and Tanaka (1973) to account for inclusion interactions. Although
the experimental material behaviour was better described using these models, errors
arose when the inclusions remained elastic, resulting in an overly stiff response and an
overestimation of the composite flow stress. The reason for this error was attributed to
the fact that the plastic strain in the matrix was determined from a reference equivalent
stress using the volumetric average of the matrix stress tensor. This equivalent stress can
be significantly lower than the phase-average of the equivalent stress due to the severe
stress-gradients that develop in the plastic regime, resulting in the composite stress to be
overestimated (Pierard et al., 2007).

70
This limitation led to the development of a modified secant method that defines the
stress field in the non-linear matrix phase using the second-order moment of the
volumetric stress tensor by Suquet (1995). The modified secant approach coincides with
the variational approach of Ponte Castenada (1996) and gives very good agreement with
the finite-element solution for a ductile matrix embedded with spherical elastic
inclusions (Segurado et al., 2002). Despite these improvements, a fundamental
limitation of secant-based methods is that they cannot be applied for non-proportional
loadings, resulting in a renewed interest in recent years to improve the incremental
formulations. Significant improvements have been obtained in the incremental models
by using only the isotropic component of the anisotropic tangent stiffness tensors
(Gonzalez and Llorca, 2000; Doghri and Ouaar, 2003).

Today, it is generally agreed that both secant and tangent-based homogenization
models can provide acceptable approximations to the material behavior of composite
materials in the elastic and plastic regimes (Pierard et al., 2007). As each type of model
has its advantages, the desired application of the model is perhaps the most important
factor in its selection. For materials with dilute concentrations of inclusions (less than
10%), Mueller and Mortensen (2006) found no discernable difference in the predictions
for the bulk and shear moduli of the composite when they evaluated five different
homogenization schemes in the plastic regime (secant, self-consistent, generalized self-
consistent, differential effective-medium, identical hard spheres approximation). This
result suggests that a simple homogenization model can be suitable for a large range of
engineering alloys.

A recent work by Pierard et al. (2007) provides additional insight into the importance
of selecting the appropriate homogenization scheme. In this work, a finite-element study
was conducted for a material containing aligned ellipsoidal inclusions to obtain an
exact solution to the homogenization problem. A classical secant (Mori and Tanaka,
1973), modified secant (Suquet, 1997) and a modified incremental model (Doghri and
Ouaar, 2003) were then evaluated to measure their predictive abilities. It was observed
that the modified secant method gave the best agreement with the numerical results at
71
the onset of plastic deformation while the classical method gave the best prediction for
the composite hardening rate in the fully plastic regime. The incremental model of
Doghri and Ouaar (2003) provided very good results in all of the cases considered. A
comparison of the results with the exact solution obtained using finite-element
simulations is shown in Figure 3.1.


Figure 3.1: Predictions of the tensile stress-strain curves in the longitudinal direction for
the composite reinforced with ellipsoidal inclusions with a volume fraction of 25%. (a) n =
0.05; (b) n = 0.40. The error bars represent the standard deviation in the numerical
simulations (Pierard et al., 2007).

What is most interesting from the work of Pierard et al. (2007) in regards to
percolation modeling is the stress predicted within the inclusions. This is the most
important feature of the homogenization model as its main purpose is to be incorporated
into a void nucleation model. It was revealed that while the incremental and modified
secant methods gave the best estimates for the composite stress, the stress within the
particles was best predicted using the secant method shown in Figure 3.2. This is an
important result and it emphasizes the fact that that the ability of a model to predict the
overall response of a material from its constituents does not mean that the predicted
stress within the constituents is accurate as well. It is also important to note that all of
the models underestimated the plastic strain in the matrix as the strong strain-gradients
that develop between inclusions are not captured. These results serve to highlight the
72
approximate nature of the homogenization models and one can see that in many cases,
there may not be a distinct advantage in selecting one model over the other. The specific
application and degree of accuracy in the metric of interest should be used in selecting
an appropriate model.


Figure 3.2: (a) Evolution of the plastic strain in the matrix as a function of the applied
strain. (b) Evolution of the von Mises equivalent stress in the ellipsoids as a function of the
applied strain. The composite was loaded in the longitudinal direction and the matrix
hardening coefficient was 0.40 (Pierard et al., 2007).

3.3 Selection of a homogenization theory

The materials of interest in the present work are industrial alloys such as automotive-
grade aluminum-magnesium alloys or high strength steels where the volume fraction of
second phase particles is on the order of a few percent. As such, the reinforcement of
these particles on the stress-strain curve is not of prime concern since they are generally
not considered at all in Gurson-based material models. The influence of the particles in
Gurson-based models is implicitly captured by using the experimental flow stress
relation and assuming it describes the behaviour of the virgin matrix. The objective of
this work is to augment existing damage-based constitutive models by improving their
ability to predict void nucleation through knowledge of the stress within the inclusions,
not to predict the macroscopic response of the material from homogenization theory.
The initiation and evolution of voids is the principal concern. If the material of interest
contains a large volume fraction of reinforcing particles, a sophisticated homogenization
73
scheme may be required to predict the overall response of the material to capture the
load loss as particles nucleate voids.

For the purpose of modeling void nucleation, the secant method of Tandon and Weng
(1986) will be used to estimate the stress within the inclusions. The advantages of
adopting a secant-based method are that they are computationally efficient, easily
implemented into numerical codes and can provide reasonable estimations for the stress
in the matrix and inclusions. The issue of computational efficiency is of paramount
importance in a damage percolation model because the homogenization procedure must
be repeated for each of the large number of particles within the microstructure at each
loading step. The adoption of a more sophisticated homogenization model is left for
future work.

3.4 Particle-based homogenization model

The particle-based homogenization scheme considered in this work was developed
by Tandon and Weng (1986) for a dual-phase composite subjected to a prescribed
traction. The composite is composed of three-dimensional, randomly-oriented elastic
particles embedded within a ductile matrix. The particles are spheroidal and
characterized by their elastic properties, volume fraction and aspect ratio. The
homogenization theory is valid for inclusion shapes ranging from flat discs, to
ellipsoids, to elongated fibres.

The homogenization model of Tandon and Weng (1986) is a secant-based approach
which utilizes Berveiller and Zaouis (1979) modification to the solution of Hill (1965)
for proportional loading and incorporates the mean-field method of Mori and Tanaka
(1973) to account for particle interactions. Additional details of the derivation of this
model can be found in Tandon and Weng (1986). The following section will present the
details of this theory relevant for integration with a damage-based constitutive model.



74
.1 Effective moduli of a randomly-oriented composite

The ductile matrix material is treated as phase 0 and the embedded elastic particles
are defined as phase 1 with a volume fraction, f
1
, and aspect ratio, W
1
. The respective
isotropic Poissons ratio, and bulk, shear and elastic moduli of the r-th phase are denoted
by v
r
,
r
,
r
and E
r
with a superscript, s, used to denote a secant quantity such as the
secant shear modulus of the matrix,
0
s
. The stress and strain tensors for each phase are
denoted by
( ) r
ij
o and
( ) r
ij
c which can be decomposed into their respective deviatoric and
hydrostatic components as
( ) ( ) ( )
hyd
r r r
ij
ij ij
o o o o ' = + with
( ) ( )
hyd
/ 3
r r
kk
o o = and
( ) ( ) ( )
hyd
r r r
ij
ij ij
c c o c = + with
( ) ( )
hyd
/ 3
r r
kk
c c = where ij o is the Kronecker delta. Any property or
quantity associated with the composite is denoted using an overbar symbol such as
ij
o for the composite stress tensor.
In a composite material the constituents are generally in a triaxial state of stress
that should be characterized using the equivalent measures for the stress, strain and
plastic strain as


( ) ( ) ( )
eq
(3/ 2)
r r r
ij ij
o o o ' ' =

( ) ( ) ( )
eq
(2/ 3)
r r r
ij ij
c c c ' ' =

p p p
eq
(2/ 3)
ij ij
c c c =
(3.1-3)

The components of the stress and strain in the composite are related through the
effective secant shear and bulk moduli as


' '
2
s
ij ij
o c =
hyd hyd
3
s
o k c = (3.4a, b)

To determine the stress and strain in the composite and its constituents during plastic
deformation, the effective secant elastic moduli of the matrix and composite must first
be determined as functions of the matrix plastic strain,
p(0)
eq c . The secant elastic modulus
and Poissons ratio of the matrix are expressed as


0 0
0 p(0)
eq
0 0 0
(0)
0 eq
3E 1
E
E (1 2 ) 1
E
s
s
s
v

c

o
= =
+
+

0
0 0
0
1 1
2 2
s
s
E
v v
E
| |
=
|
\ .
(3.5, 3.6)

75
The secant bulk and shear moduli are obtained using the standard isotropic
relations
0 0 0
E / 2(1 )
s s s
v = + and
0 0 0
E / 3(1 2 )
s s s
v k = . The matrix material is assumed to be
plastically incompressible and thus the secant bulk modulus remains constant at
0 0
s
k k = .
The effective secant moduli of the composite material can be determined as


0
1 2 1
(1 / )
s
s s
f p p
k
k =
+

0
1 2 1
(1 / )
s
s
s s
f q q

=
+
(3.7, 3.8)

where
is
p and
is
q are functions of the particle shape, volume fraction, elastic moduli of
the constituents and the fourth-order Eshelby (1957) S tensor in Appendix A. The
expressions for
is
p and
is
q are rather lengthy and are not presented here for brevity but
can be found in Tandon and Weng (1986).

.2 Average stress in the composite and its constituents

The respective hydrostatic and deviatoric stresses in each constituent are determined
from the composite stress tensor using the stress concentration factors,
( ) r
a and
( ) r
b as


( )
( )
hyd hyd
r
r
a o o =
( )
( )
r
r
ij ij
b o o ' ' =
( )
( )
r
r
eq eq
b o o = (3.9a-c)

which are defined for the matrix material and particles as


(0)
1
1/
s
a p =
( )
(1) (0)
1 1
1 (1 ) / a f a f = (3.10a-b)

(0)
1
1/
s
b q =
( )
(1) (0)
1 1
1 (1 ) / b f b f = (3.10c-d)

The stress in the composite and its constituents must be in equilibrium and thus


(0) (1)
1 1
(1 )
ij ij ij
f f o o o = + (3.11)

The onset of yielding of the matrix occurs when
(0) (0)
eq y / b o o > with
(0)
b determined
using the elastic moduli of the matrix in Eq. (3.5).



76
.3 Average strain in the composite and its constituents

Due to the presence of the elastic inclusions, the composite is not plastically
incompressible and the composite plastic strain must be determined from the unloading
process as

p
'
hyd
1 1 1 1
2 2 3 3
ij ij ij
s s
c o o o
k k
| | | |
= +
| |
\ . \ .
(3.12)

The matrix is assumed to be isotropic and obey J
2
plasticity (von Mises material) from
which the plastic strain components can be readily determined by integrating the J
2
flow
rule for proportional loading to yield


p(0)
eq p(0)
'
3
2
ij ij
c
c o
o
= (3.13)

The particles are assumed to remain elastic during deformation and the strain in the
particles can be expressed as


(1) ' (1)
hyd (1)
1 1
2 3
ij
ij ij
b a o o
c o
k
= + (3.14)

The solution for the secant moduli of the composite and subsequent stress and strain
in the constituents is nonlinear and an iterative solution is required in the plastic regime.
The required material parameters to determine the behaviour of the composite and stress
in the constituents are E
0
, v
0
, E
1
, v
1
, f
1
, and W
1
. A simple fixed-point algorithm can be
easily implemented to determine the required value for
p(0)
eq
c and is described in Tandon
and Weng (1986). However, this solution method must be modified in the present work
because the homogenization method is to be coupled with a damage-based constitutive
model to account for the influence of voids on the subsequent stress and strain in the
constituents.



77
3.5 Procedure for integrating a particle-based homogenization theory
into an existing damage-based constitutive model

To integrate a particle-based homogenization theory into a damage-based
constitutive model, we assume that the bulk material can be idealized as a three-phase
composite composed of a matrix material with embedded particles and voids. It is
assumed that this idealized three-phase composite can be decomposed into its
constituents by applying two successive homogenization schemes: (i) a Gurson-based
constitutive model for the voids embedded in a ductile composite matrix which is
composed of the particles and matrix material and (ii) separation of the composite
matrix into its constituents to determine the stress within the matrix and particles using a
secant-based homogenization scheme.

Let us consider a bulk material which contains both voids and hard elastic
particles/inclusions within a ductile matrix. To mitigate the influence of the voids, the
experimental flow stress relation for the bulk material can be obtained from a torsion or
compression-type test or from a tensile test if the initial porosity is negligible (Pardoen,
2006). This flow stress relation is essentially that of a two-phase composite composed of
the matrix and particles.

Now, the bulk material is subjected to a deformation process such as a sheet metal
forming operation where the pre-existing voids will grow and additional voids will be
nucleated from particle cracking and/or debonding from the matrix. The presence of the
voids results in material softening which further promotes void evolution resulting in
ductile fracture as the voids coalesce and link-up throughout the material. The influence
of the voids on the response of the bulk material can be described using a damage-based
constitutive model such as the Gurson (1977) model. Gurson-based models are the result
of a homogenization procedure for a material composed of voids embedded within a
virgin matrix. Therefore, a Gurson-based model can be applied to the bulk material to
account for void damage by modeling the bulk material as a material which contains
voids embedded within a so-called composite matrix. The composite matrix is
composed of the virgin matrix and particles whose composite behavior is described
78
using the experimental flow-stress relation for a two-phase composite as described
above.

The typical procedure for the integration of the Gurson-based material model is now
applied to the bulk material. The material has a void volume fraction or porosity, f, and
is subjected to a monotonic proportional loading with a macroscopic strain,
ij
E . The
stress state is then integrated using the damage-based yield surface to determine the
macroscopic stress,
ij
E . The equivalent plastic strain and flow stress within the
composite matrix are,
p
eq
E and
p
eq
( ) E o o = . Void evolution is very sensitive to the stress
state which is characterized by the stress triaxiality ratio defined as eq
hyd
/ T = E E .

Due to the presence of the voids, the bulk material is softer and thus the composite
matrix must work-harden to a greater extent to reach the applied strain of
ij
E than if no
voids were present. Since the voids do not contribute to load-sharing, the entire stress
must be borne by the matrix and particles. Therefore, from the perspective of the
constituents, it is equivalent to subjecting the composite matrix to a larger applied strain
denoted
ij
c , that results in the same equivalent stress and plastic strain as when softening
was considered. This effect is illustrated in Figure 3.3.


79
Figure 3.3: Equivalent stress response of the composite matrix with the applied strain.

The situation for the composite matrix now resembles that of a particle-based
homogenization problem for a prescribed traction. The equivalent stress state within the
composite matrix is known and the stress state in the constituents must be determined
that is in equilibrium with this prescribed stress. The secant-based homogenization
method of Tandon and Weng (1986) can now be applied to determine the effective
secant moduli of the composite matrix which satisfy the stress state defined by the
damage-based stress integration.

The general integration procedure is presented in Figure 3.4 and can be summarized
in the following steps:

(a) A bulk material containing voids and particles within a ductile matrix is
subjected to a macroscopic loading

(b) The voids within the material are homogenized into an equivalent void
embedded in a composite matrix

(c) A damage-based constitutive model is applied to determine the stress in the
composite matrix by accounting for the influence of the voids

(d) With the voids removed from the composite matrix, a particle-based
homogenization theory for a prescribed traction is applied to the composite
matrix to determine the stress within the matrix and particles

(e) The particles can now be tested for nucleation using an appropriate model for the
material

(f) The evolution of the voids must obey the plasticity of the matrix and the
traditional voided unit cell is obtained by combining the virgin matrix and the
isolated void. The standard models for void evolution are evaluated using the
stress state and plastic strain of the matrix.
80

Homogenize the random voids into a
single void and homogenize the particles to
an average size and shape
Bulk material:
Composed of random voids and particles
embedded within a ductile matrix
1
1


Figure 3.4: Schematic of the integration process of a particle-based homogenization theory into a damage-based constitutive
model.
81
To create a stress state in the composite matrix equivalent to that of the voided bulk
material, the applied strain,
ij
c , must be determined to produce a stress,
ij
o , subject to the
constraints that
eq
o o = ,
p p
eq eq
E c = and
hyd eq
/ T o o = . It is important to mention that the
equivalent stress,
eq
E , identified in the damage-based integration is not equal to the flow
stress within the composite matrix due to the presence of the voids. The procedure to
determine the applied strain,
ij
c , and stress,
ij
o , within the composite matrix using a
secant-based homogenization scheme for monotonic, proportional loading is described
in the following section.

First, the hydrostatic stress of the composite matrix can be computed directly from
the requirement of an equivalent triaxiality as


hyd hyd eq
( / ) o o = E E (3.15)

From the definition of the effective secant moduli in Eq. (3.4) and the equivalent stress
and strain in Eq. (3.1) and Eq. (3.2), the equivalent strain in the composite matrix is


eq
/ 3
s
c o = (3.16)

Only the deviatoric stress and strain of the composite matrix remain unknown and only
one is required to be pre-determined as they are related through the secant shear
modulus. The deviatoric strain in the composite is assumed to be proportional to the
deviatoric strain in the damage-based model or three-phase composite as


' '
ij ij
E c = (3.17)

and thus
eq eq
/ E c = , from which the stress and applied strain in the composite matrix
can be derived as

'
hyd
eq
2
3
ij ij ij
E
E
o
o o o
| |
= +
|
|
\ .

hyd '
p
eq
1 1
3 3
ij ij ij s s
E
E
o
o
c o
k
| |
= +
|
|
\ .
(3.18, 3.19)

82
which satisfies the requirements that
hyd eq
/ T o o = and eq o o = . The stress within the
composite matrix can be evaluated immediately following the stress integration of the
damage-based constitutive model while
ij
c requires knowledge of the effective secant
moduli which have yet to be determined.

3.6 Iterative solution for the effective secant moduli

An iterative procedure is required to determine the effective secant moduli of the
composite matrix from which the stress and strain within the constituents can be
determined. The iterative solution is developed through the final constraint that the
secant elastic moduli must result in a plastic strain in the composite matrix of
p p
eq eq
E c = .
The plastic strain of the composite matrix must account for the elastic heterogeneity of
the particles and from substituting Eq. (3.12) into Eq. (3.3) and utilizing Eq. (3.1), the
required secant shear modulus of the composite matrix is


( )
1
2
* p 2 2
eq hyd
(3 ) 2 1/ 1/
s s
E o o k k o

| |
= +
|
\ .
(3.20)

All of the parameters in Eq. (3.20) are constant during the iteration loop except for the
secant bulk modulus,
s
k , which varies with the plastic strain of the matrix.

A fixed-point algorithm is used to obtain the solution to the non-linear equations for
the secant moduli by iterating upon the secant elastic modulus of the matrix,
0
E
s
. A trial
value for
0
E
s
is assumed from which the subsequent trial values for the secant shear, bulk
and Poissons ratios of the matrix can be determined from Eq. (3.5) and (3.6) and used
to evaluate the expressions for
is
p and
is
q as well as
s
k in Eq.(3.7) and
* s
in
Eq.(3.20). A new estimate for
0
s
can be determined by setting
* s s
= in Eq. (3.8)
and evaluating Eq. (3.5) to obtain a new estimate for
0
E
s
. If this new value of
0
E
s
is equal
to the trial value, the solution has been obtained, otherwise the algorithm is repeated
using the current
0
E
s
as the new trial value. Upon convergence, the stress concentration
83
factors to determine the stress and strain tensors of the matrix and particles can be
evaluated using Eq. (3.9), Eq. (3.13) and Eq. (3.14). This algorithm generally achieves
convergence within three iterations to a tolerance of 0.001 and is straightforward in its
implementation.

3.7 Application of the particle-based homogenization scheme into a
Gurson-based constitutive model for ductile fracture

The integration of a homogenization scheme into a Gurson-based damage-
framework improves all aspects of the damage model and couples the various damage
mechanisms that are often considered independently. The advantages of the proposed
fully-coupled damage-based constitutive model are

- Modeling of void nucleation is improved through knowledge of the stress state
within the particles as a function of their shape, content and composition

- As particles nucleate voids, the stress state within the matrix becomes more severe
due to increased material softening (higher porosity) as fewer particles are
available for load-sharing.

- The stress state within the matrix and particles becomes progressively more severe
as the particle content decreases due to nucleation. This promotes additional
nucleation, void growth and material softening

- The plastic strain within the matrix material is higher than in the composite matrix
and thus promotes void evolution and coalescence

- The model reverts to its original formulation if the material does not contain any
second-phase particles or inclusions.

- The flow stress relation of the matrix material does not have to be predetermined.

84
- An absolute minimum number of new parameters have been introduced into the
damage-based framework. The elastic properties of the constituents as well as the
average particle shape and content are typically known from standard material
characterization techniques used in damage-based modeling.

- The particle-based homogenization scheme is computationally efficient and easily
implemented into any existing damage-based constitutive model

The proposed integration procedure will now be implemented into an enhanced
version of the well-known Gurson-Tvergaard (Gurson, 1977; Tvergaard, 1981)
constitutive model. A limitation of using a secant-based homogenization scheme is the
intrinsic assumption of monotonic proportional loading.

.1 Yield criterion

The Gurson-Tvergaard (GT) yield criterion from Eq. (1.19). is adopted in the present
work along with the semi-empirical extension by Ragab (2004a) for void shape
evolution through the use of the q
1
and q
2
parameters. The reader is directed to Chapters
1 and 2 for additional information on their development.

.2 Void growth and shape evolution

Void growth and shape evolution is modeled using the semi-empirical equations of
Ragab (2004a) in Eq. (2.22) and (2.25). The original form of these models, not the
calibrated versions from Chapter 2, is used in this application because only a general
study is being performed.

.3 Void coalescence

The plastic limit-load coalescence model of Pardoen and Hutchinson (2000) in Eq.
(1.13) is employed to model coalescence. Details of this coalescence model are
presented in Chapters 1 and 2.


85
3.8 Evaluation of the full constitutive model

The coupled GT constitutive model is applied to a model material that is subjected to
proportional plane-stress loading to represent a sheet metal forming operation. The
composite material is an aluminum-magnesium alloy, AA5083-O (Tohgo and
Mochizuki, 2002) and the material parameters are presented in Table 3.1. The void-
related material parameters for this alloy were not reported in Tohgo and Mochizuki
(2002) and typical damage values were arbitrarily assumed in order to provide general
qualitative trends for performance of the model. The voids are assumed to be pre-
existing within the bulk material (no nucleation). A nucleation rule is not considered at
present since the objective is to demonstrate the ability of the model to predict the stress
within the inclusions and matrix. For a specific material, the nucleation rule can easily
be added into the model to account for the dominant nucleation mode such as cracking
or debonding. A treatment for nucleation will be discussed in Chapter 4 in development
of the percolation model.

Table 3.1. Material parameters of the constituents in an AA5083-O alloy (Tohgo and
Mochizuki, 2002).
Constituent Material Parameter
Composite

y
150 MPa o = 0.17 n = v 0.29 =
Matrix
0
69 GPa E =
0
v 0.30 =
Particles
1
10% f =
1
90 GPa E =
1
v 0.17 =
Voids
o
0.1% f = 1/ 6
o
W =
o
10 =

The composite is subjected to proportional loading where the major and minor true
strain increments are related using the load ratio,
2 1
/ E E = , with ratios of -0.5, 0 and 1
corresponding to uniaxial tension ( 1/ 3 T = ), plane strain ( 1/ 3 T = ) and equal-biaxial
tension ( 2/ 3 T = ). The power-law hardening rule in Eq. (2.17) is used to describe the
flow stress in the material. The equations of the constitutive model have been
implemented into a C++ code and evaluated using an applied major strain increment of
86
0.001. The stress is integrated in the GT damage model using the forward-Eulerian
integration scheme of Worswick and Pelletier (1998) described in Chapter 4.

3.9 Results

.1 Average stress in a material containing voids and hard elastic particles in a
ductile matrix

The coupled damage-based constitutive model naturally separates the bulk material
containing the particles, voids and matrix into its constituents as shown in Figure 3.5.
The influence of the voids is embedded within the response of the composite matrix
whose stress is known from the GT stress integration. The addition of the particle-based
homogenization model into the GT framework enables the application of a stress-based
nucleation rule to predict particle cracking or debonding.
0.0
1.0
2.0
3.0
4.0
5.0
6.0
7.0
0 0.5 1 1.5 2 2.5 3
Equivalent strain E
eq
E
q
u
i
v
a
l
e
n
t

s
t
r
e
s
s

i
n

t
h
e

c
o
n
s
t
i
t
u
e
n
t
s
:

e
q
/

y
Uniaxial tension
Particles / Inclusions
Composite matrix
Matrix

Figure 3.5: Equivalent stress within the constituents of the voided bulk material when
subjected to uniaxial tension. The inclusions are spherical and the voids are initially oblate.
The stress has been normalized by the yield stress of the composite matrix.

When modeling nucleation, knowledge of the maximum principal stress is of prime
importance as the hard elastic particles may experience a brittle-type fracture or debond
in the principal loading direction. The maximum principal stress for any particle shape,
87
composition, and stress state is automatically determined within the coupled model as
shown in Figure 3.6 for several loading conditions.
0.0
1.0
2.0
3.0
4.0
5.0
6.0
7.0
0 0.5 1 1.5 2 2.5 3
Equivalent strain E
eq
S
t
r
e
s
s

i
n

t
h
e

p
a
r
t
i
c
l
e
s
:

1
/

y
Coalescence
Plane strain
Uniaxial tension
Equal-biaxial tension

Figure 3.6: Normalized principal stress within the spherical elastic particles/inclusions for
three loading conditions. The bulk material ruptures due to void coalescence when loaded
in plane strain and equal-biaxial tension.

.2 Influence of the particle shape on void damage and ductile fracture

The secondary benefit of the coupled damage model is that the reinforcing influence
of the second-phase particles and inclusions can be removed from the bulk material to
determine the stress state within the matrix material for improved modeling of void
evolution. It is important to note that the traditional damage-based model would give
similar predictions for void evolution if the flow stress relation of the matrix was pre-
determined from an experimental or numerical procedure. The coupled-GT model not
only eliminates this step but also provides the stress state in the matrix and the
particles/inclusions as a function of their shape and composition. To elucidate the
performance of the coupled-GT model, the standard GT model described in the previous
section is applied without accounting for the influence of the particles. The
improvement of the coupled model is then measured using the results of the standard
model as the performance baseline. The aspect ratio of the reinforcing particles is varied
88
from 1/6 to 6 to assess the sensitivity of the damage response to a change in the particle
geometry.

The inclusion shape has a dramatic influence upon void growth within the bulk
material as seen in Figure 3.7 where the porosity can reach levels of 50% or higher for
the ellipsoidal particles than if the particles were neglected. The spherical particles will
also lead to increased void growth but it is mild in comparison. As noted by Tandon and
Weng (1986), both the oblate and prolate particles are similar in their ability to reinforce
the material with flat particles or discs providing the best reinforcement shape. It is
important to observe the multiplying effect of the particles on void growth as a 10%
volume fraction of particles may yield a significantly larger increase in the predicted
void damage. The opposite trend is observed in the evolution of the void shape in Figure
3.8 where shape evolution is the highest when the material contains spherical particles.

0
10
20
30
40
50
60
70
0 0.5 1 1.5 2 2.5
Equivalent strain E
eq
%

D
i
f
f
e
r
e
n
c
e
:

P
o
r
o
s
i
t
y
W
1
= 1
W
1
= 1/6
W
1
= 6
Uniaxial tension

Figure 3.7: Influence of the particle shape on damage development in the coupled
constitutive model compared to the porosity determined using the standard model.
89
0
5
10
15
20
25
30
35
40
45
0 0.5 1 1.5 2 2.5
Equivalent strain E
eq
%

D
i
f
f
e
r
e
n
c
e
:

V
o
i
d

a
s
p
e
c
t

r
a
t
i
o
Uniaxial tension
W
1
= 1/6
W
1
= 6
W
1
= 1

Figure 3.8: Influence of the particle shape on the evolution of the void shape in the coupled
constitutive model relative to the void shape predicted using the standard model.

It may seem counter-intuitive that increased void growth did not lead to increased
shape change but this effect is a result of the particle interactions with the matrix stress.
From inspection of the stress triaxiality in Figure 3.9, the matrix triaxiality is the highest
for flat particles and the lowest for the spherical particles. The better the reinforcing
properties of the particles (higher load-sharing with the matrix), the more the stress state
within the matrix is altered. In fact, the stress triaxiality decreases within the elastic
particles from spheres to discs and therefore the triaxiality must increase in the matrix to
maintain equilibrium. Since the voids are not internally pressurized, they will deform
according to the plasticity, and stress within the matrix and their evolution is very
sensitive to the hydrostatic stress (triaxiality). It is this sensitivity to the triaxiality that
results in the significant void growth in the coupled model when the material contains
ellipsoidal particles.
90
0
5
10
15
20
25
30
35
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2 2.25
Equivalent strain E
eq
%

D
i
f
f
e
r
e
n
c
e
:

T
r
i
a
x
i
a
l
i
t
y

i
n

t
h
e

m
a
t
r
i
x
_
W
1
= 1
W
1
= 1/6
W
1
= 6
Uniaxial tension

Figure 3.9: Influence of the particle shape on the stress triaxiality within the matrix
material compared to the standard model where the triaxiality is constant at a value of 1/3.

The stress triaxiality is also responsible for the decreased void shape evolution for
the ellipsoidal inclusions. The initial void shape in this study was assumed to be an
oblate ellipsoid with an aspect ratio of 1/6 to resemble the void shape due to particle-
cracking or partial debonding. The growth of flat and oblate voids occurs predominantly
in the loading direction as the void tends to open-up vertically rather than expanding
laterally. Consequently, the largest amount of void shape change will occur in uniaxial
tensile loading with a triaxiality of 1/3. Increasing the triaxiality leads to increased
lateral void growth and a more spherical void and thus a smaller aspect ratio. Since the
spherical particles are least effective at increasing the triaxiality in the matrix, the voids
will experience the most elongation, and the least growth compared to a composite
reinforced with ellipsoidal particles.

Due to the coupled nature of the model, the resulting inter-relationships may not be
immediately apparent but the net effect of the interactions can be observed in the
coalescence model. The plastic limit-load criterion in Eq. (1.13) is a powerful
coalescence model since it relates the geometry of the microstructure on the right-hand
side (RHS) of the equation to that of the stress state in the material on the left-hand side
91
(LHS). As shown in Figure 3.9 for plane strain loading, the net effect of the particles
and voids on the stress state in the matrix shifts the stress state upwards to a more severe
one. This stress state results in the evolution of the microstructure through the void
volume fraction, aspect ratio and spacing ratio. As the microstructure evolves, the
energy required to achieve a localized deformation mode decreases which shifts the
plastic constraint to the left in Figure 3.10. The intersection of the geometric plastic
constraint with the stress state signifies the onset of coalescence and fracture. The
highly-coupled nature of the model is revealed in Figure 3.10 as void coalescence and
fracture does not occur in the standard damage-based model while the coupled model
predicts fracture at a much more modest strain.

0.0
1.0
2.0
3.0
4.0
5.0
0 0.25 0.5 0.75 1 1.25 1.5 1.75
Equivalent strain E
eq
P
l
a
s
t
i
c

c
o
n
s
t
r
a
i
n
t

f
a
c
t
o
r
LHS of Eq.(1.13): Standard model
LHS of Eq.(1.13): Coupled model
RHS of Eq.(1.13): Standard model
RHS of Eq.(1.13): Coupled model
Coalescence
Plane strain: W
1
= 1/6
Microstructure evolution
Stress state

Figure 3.10: Variation of the plastic limit-load required for coalescence in the coupled and
standard (no particles) damage-based constitutive models.

3.10 Summary

An approximate integration scheme has been presented to implement a secant-based
homogenization theory for particle-reinforced plasticity into an existing damage-based
constitutive model for ductile fracture. The resulting model can account for the
influence of the second-phase particles on void growth, shape evolution, coalescence
92
and material softening. The stress state within particles can also be determined as a
function of the particle content, shape and elastic properties. Additionally, the stress
state within the matrix material can be estimated with no prior knowledge of its
hardening profile. The resulting model will be implemented into a damage percolation
model in Chapter 4 to determine the stress state within individual inclusions to predict
void nucleation. The present model is best suited for application to sheet metal forming
of damage-sensitive industrial alloys where the loading is proportional and the volume
fraction of the second-phase particles is low.

93
Chapter 4 A Multi-scale Damage Percolation
Model of Ductile Fracture

4.1 Development of the percolation model

In this chapter, the void growth and evolution models developed in the previous
chapters are amalgamated into a complete framework to model damage evolution in a
material with a heterogeneous particle distribution. The void evolution models are
adapted to a general three-dimensional particle field and stress state by removing the
periodic assumption used in their development in Chapter 2. The percolation model is
then integrated into a commercial finite-element code, LS-DYNA, to create so-called
percolation elements by mapping the particle distributions to the elements. This
enables the percolation model to be directly coupled to the macro-scale as damage
evolution at the micron-scale will control the stress state in the elements, and ultimately,
that of the entire structure via the finite-element method. This represents a significant
contribution in the modeling of ductile fracture as the particle distribution can be
accounted for in finite-element simulations of metal forming operations.

The development of a sophisticated constitutive model of this scale requires careful
consideration and justification of each step in the modeling process to ensure that model
has a strong physical foundation. The development of the percolation model will be
separated into two main categories: macroscopic models and microscopic models. The
macroscopic models are used to transition from the micro-scale to the macro-scale and
vice versa.


4.2 Principal assumptions

The principal advantage of the damage percolation model is that the microstructure
no longer is assumed to be periodic and a measured particle/void distribution can be
used to model ductile fracture. A concerted effort has been made in this work to improve
the physical foundation of the percolation model and significant progress achieved in
reducing the number of assumptions inherent in the modeling of ductile fracture.
94
Nevertheless, extending the micromechanical models to a general three-dimensional
case requires numerous approximations and assumptions. The principal assumptions
used in the development of the percolation model are:

- The stress state within the element is uniform and homogeneous. Local stress
and strain heterogeneity within void and particle clusters is neglected

- The particles, voids and cracks do not interact. Voids/cracks grow as isolated
voids.

- Void evolution models obtained using axisymmetric unit cells can provide a
reasonable approximation to the growth of voids in a real material if the shear
stress is small compared to the tensile stress.

- The particles, voids and cracks remain ellipsoidal during deformation.

- Nucleation via debonding is not explicitly modeled. An emphasis is placed upon
void nucleation via particle cracking.

- Void coalescence occurs by necking failure of the inter-void ligament. Ligament
shearing is not considered.

- Size effects are negligible at the length scales considered and continuum-based
models for void initiation and evolution are valid at the micron-scale and higher.
Deformation of voids and particles that are smaller than 1 m would require
considering dislocation dynamics.

- The grains and texture-related effects are not accounted for in the model. It is
assumed that the bulk properties of the particle field are isotropic although
deformation of the voids at the local-scale is dependent upon the loading
direction.

95
Each of the above assumptions will introduce a degree of uncertainty into the model
with the major limitations being the assumption of a uniform stress distribution, no
shear coalescence, negligible size effects and neglecting anisotropy and the grains in the
material. The rationale for these assumptions is discussed in the proceeding sections.

.1 Assumption of a homogeneous stress state No void and particle interactions

Although the stress state is assumed to be homogeneous within the percolation
element, the stress state is not homogeneous because severe local strain-gradients
develop within particle and void clusters. As a result, the voids and particles will interact
and not evolve as isolated objects. This assumption is unlikely to be overcome without
explicit modeling each of void and particle using finite-element techniques; however, it
can be improved by using a sufficient number of percolation elements. In the limit, each
percolation element would contain a single void/particle and the isolated assumption
becomes realistic. A scheme for modeling the interactions of voids within clusters is
proposed in Chapter 7.

.2 Assumption of internal necking coalescence as the dominant coalescence mode

The omission of a shear-based coalescence model is an unfortunate limitation of the
present work and is attributed to the lack of robust model akin to the plastic limit-load
criterion for internal necking coalescence. Butcher and Chen (2009) proposed a shear-
extension to the plastic limit-load for combined tension and shear based upon the work
of Xue (2007). However, this model is better suited for a general constitutive model and
not for individual voids. A recent work by Schyvearts et al. (2011) has found that the
plastic limit-load model can give good predictions for necking coalescence in combined
tension and shear by accounting for the void orientation with shear. The percolation
model is also expected to perform well in this situation as it accounts for the orientation
of the voids and cracks as well as their rotation when they evolve and change shape in
shear loading.



96
.3 Assumption of size-effects being negligible above the micron-scale

All multi-scale models are only valid for a range of length scales. The percolation
model has been designed to model void initiation and evolution from the micron-scale to
the macro-scale. The typical size of a second-phase particle in many alloys is on the
order of several microns or larger in diameter. It is assumed that continuum-scale
models can be used to describe the deformation of the particles and voids at this scale.
Size effects are unavoidable and it is well known that sub-micron sized voids tend to
grow slower than their larger counterparts (Liu et al., 2003; Tvergaard and Niordson,
2004; Wen et al., 2005). At the sub-micron scale, dislocation dynamics must be
considered using a strain-gradient plasticity model (Fleck and Hutchinson, 1997) that
includes a length scale parameter for the size effects. Fortunately, a numerical study by
Wen et al. (2005) has found that the void size effect is limited in uniaxial tension and for
a small volume fraction of voids. Since most metal forming operations occur in the low
triaxiality regime (T < 1) and have failure porosities on the order of several percent at
most, the size effect can be neglected as a first-order approximation of the material
behavior. An extension of the percolation model that accounts for dislocation dynamics
on void nucleation and evolution would improve the physical foundation of the model
and enable it to incorporate events from the nano-scale to the macro-scale.

.4 Neglecting the influence of texture effects and grains

The percolation model does not account for influence of material texture or for the
presence of grains in the material. This assumption implies that it is the presence of the
micro-voids in the material that are responsible for the promotion of localization and
fracture. For example, in a uniaxial tensile test, geometric softening causes the material
to form a localized neck. The stress state in the center of the necked region becomes
severe and spurs void nucleation and growth, softening the material in the neck and
causing additional necking. Ultimately, the sample fails as the voids coalesce throughout
the necked region. This type of fracture mechanism can be well described using a
damage-based approach such as the percolation model.

97
This type of fracture process is not always the case since anisotropy and the grain
structure of the material can be responsible for the initial localization by the formation
of shear bands. These shear bands provide the high local plastic strains required to drive
void nucleation and growth, softening the material in the shear band and promoting
additional localization until fracture occurs. In this situation, the isotropic percolation
model would overestimate the strain and porosity at fracture since higher levels of
porosity would be required to trigger the localization process without the presence of the
shear bands.

Overall, the percolation model provides an excellent tool for the modeling of ductile
fracture and can describe a large portion of the fracture process. A future extension of
the percolation model should account for material anisotropy in the yield criterion and
void evolution models. The influence of the grains could be accounted for by coupling
the percolation model with a crystal plasticity model. The coupled crystal plasticity
percolation model would provide insight into the complete fracture process by
accounting for the entire microstructure.

.5 Terminology used in development of the percolation model

The particle field is assumed to be composed of particles, voids and cracks and
referred to as objects if the modeling treatment is the same for each constituent.
Particles are modeled as a single class but their composition and type may be variable to
accommodate different particle types in the material. Cracks are subjected to the same
modeling treatments as voids but are defined as a crack to discriminate between the
primary and nucleated voids with the voids that form due to coalescence. A crack is
formed by the coalescence between a void and a crack, two voids or two cracks.

4.3 Macroscopic models

Quantities at the element scale are defined as macroscopic because the stress state
within the element is obtained by homogenizing the voids within the microstructure
using the GT yield criterion. In this definition, the element is analogous to the unit cell
98
described in Chapter 2 except the cell contains a particle field instead of a single void.
The purpose of the macroscopic modeling process is to accept the nodal displacements
from the finite-element code and determine the stress and strain according to the
adopted constitutive model for the element. The failure criterion for the element is then
evaluated and the element is deleted or the stress returned to the finite-element solver.
This is the typical procedure for any finite-element program with the only notable
difference being that the constitutive model is rather complex. An overview of the
macroscopic modeling process is presented in Figure 4.1.

The elements provide the link to the relevant length scales for engineering (mm and
higher) because the stress state within the elements controls the deformation of the
global structure. Therefore, a brief review of the relevant kinematics of the finite-
elements is required as they play a major role in the percolation model.

99
Call Subroutine for the Constitutive Model
Update Stress Tensor
and Plastic Strain
Determine the Macroscopic Stress Tensor
and Plastic Strain in the Element
Test for
Element Failure
Delete Element
Yes
No
Call User Defined Subroutine for the Percolation Elements
Obtain Nodal Displacements of the Elements from the FE Code
Finite-Element Simulation at Time t
Call the Percolation Subroutine and Update Global Porosity
Compute the Strain Increment of the
Element from the Nodal Displacements
Divide the Nodal Displacements and
Strain Increment into Sub-increments
Reached Number of
Sub-increments?
No

Figure 4.1: Flow-chart of the macroscopic modeling process


100
.1 Relevant finite-element kinematics

Finite-element discretization is achieved in the initial (reference) configuration by
using either four or eight node isoparametric elements to interpolate the position of the
element nodes, X
a
, during deformation as

1
( , , )
n
a a
a
N q ,
=
=

X X

(4.1)

where

( , , )
a
N q , are the standard shape functions defined in dimensionless element
coordinates and n is the number of nodes (Figure 4.2). During deformation the current
position of the nodes, ( , , )
a
x y z x

as well as the nodal velocities, v
a
, can be expressed in
terms of the shape functions as


1
( )
n
a a
a
N t
=
=

x x


1
n
a a
a
N
=
=

v v (4.2, 4.3)


Figure 4.2: Finite-element discretization (Bonet and Wood, 1997)

By enforcing the displacement of the nodes by an arbitrary increment, u
a
, the
displacement can be interpolated as

1
n
a a
a
N
=
=

u u

(4.4)

101
The deformation gradient tensor, F, maps the element from its initial configuration to its
deformed configuration and can be interpolated within the element using the relations


1
n
a
a
N
F
=
c
=
c

x
x


1
N N
J

c c
=
c c x

1
n
a
a
a
N
J
=
c
=
c

(4.5-4.7)

where J is the Jacobian matrix. The deformation gradient can be used to determine the
strain measure of interest such as the Cauchy-Green strain. The strain matrix for specific
elements will be discussed in a subsequent section.

The velocity gradient tensor, L, is related to the rate of change of the deformation
gradient as

1
L FF

=

(4.8)

The deformation gradient tensor can be decomposed into a stretch and rotation tensor.
The rotational rate of deformation can be measured using the antisymmetric spin tensor,
, as

1
( )
2
T
L L O =

(4.9)
.2 Percolation element types

The types of percolation elements considered at present work are eight-node
isoparametric brick elements and four-node isoparametric plane elements. The type of
element integration such as reduced-point integration is controlled by the finite-element
code. The geometry and local node numbering of the elements are shown in Figure 4.3.


.1 Eight-node isoparametric brick element

The geometry and local node numbering of the eight-node brick element is shown in
Figure 4.3. The center of the element is computed from the element nodal points as


1
1
n
c a
a
n
=
=

x x

(4.10)

102
where n is the number of element nodes. The half lengths of the element sides (a, b, c)
are

( ) ( ) ( ) ( ) 2 1 3 4 6 5 7 8
1
8
a x x x x x x x x + + + (

=

(4.11a)

( ) ( ) ( ) ( ) 4 1 3 2 7 6 8 5
1
8
b y y y y y y y y + + + (

=

(4.11b)

( ) ( ) ( ) ( ) 5 1 6 2 7 3 8 4
1
8
c z z z z z z z z + + + (

=

(4.11c)

Any point of interest within the element can be transformed to dimensionless element
coordinates using the relations


2
c
x x
a


=


2
c
y y
b
q

=


2
c
z z
c
,

= (4.12a-c)


The shape functions for an isoparametric eight-node brick element at a node, a, are


1
(1 )(1 )(1 )
8
a a a a
N q q , , = + + +

(4.13)

with derivatives

1
a a
a
a
N
N
| | c
=
|
c +
\ .



(4.14)

The strain in the element can be interpolated using the shape functions. For an eight-
node brick element, the macroscopic strain tensor is defined as


1 1
0 0
0 0
0 0
B
2
0
2
2
0
0
a
a
x
y
a
a
n n
z
ij a a a
a a
a a xy
a
yz
a a
xz
a a
N
x
N
E
y
E
N
u
E z
E v
N N
E
w
y x
E
N N
E
z y
N N
z x
= =
c (
(
c
(
c
(
(
c
(
c
(

(
c
= = =
( ` `
c c
(
)
( c c
(
c c
(
)
( c c
(
c c
(
(
c c

u

(4.15)
103
1 2
4 3

x
y
2a
2b
1
2 3
4
5
6 7
8

y
x
z
2a
2b
2c

Figure 4.3: Local node numbering and coordinate systems for a 4-node (left) and 8-node
(right) isoparametric element.

.2 Four-node isoparametric plane element

The relevant equations for a four-node isoparametric element can be readily
determined from the equations for the eight-node element above by replacing the factor
of 1/8 with 1/4 and setting any z-related quantities to zero. The strain tensor for the
plane elements is slightly different and is written as


1 1
0
0
0
0 B
0 0
2
a
x
a
n n
a y
ij a a z
a a a z
z
xy
a a
N
x
E
N
u E
y
E E
v E
E
N N
y x
c
= =
c (
(
c
(
c

(

(
c
= = + = +
` ` `
(
)

(
)

( )
c c
(
c c (


u

(4.16)

where the strain increment in the z-direction (thickness direction) is not computed from
the shape functions. If the element is a plane strain element then 0.
z
E = If the element
is a plane stress element, then
z
E will be iteratively determined during the stress
integration routine to obtain the value of
z
E that results in 0.
z
E = Once the proper strain
in the thickness direction has been determined, the component of the deformation
gradient in this direction can be obtained. Assuming no shearing in the thickness
104
directions, for a plane strain element, 0.
z
F = For plane stress, the determinant of the
Jacobian can be used to solve for F
z
using the cumulative strains as


1
x y z
z
x y xy yx
E E E
F
F F F F
+ + +
=


(4.17)

.3 Constitutive model to account for material softening

The percolation material model was written as a user-defined subroutine for LS-
DYNA to integrate the stress state, analyze the microstructure for void evolution and
return the stress tensor and plastic strain to the finite-element program. The extended
Gurson-based yield criterion described in Chapter 2 with the calibrated q
2
parameter is
used to account for material

softening by using the global porosity,
d
f , and the average
q
2

value of the voids and cracks as


2
eq hyd
2 2
, ,
1 hyd eq 2 hyd eq 1
3
2 ( , ) cosh ( , ) 1 0
2
d d
f q n q q n f
o o
E E | | | |
u = + E E E E =
| |
\ . \ .

(4.18)
where

1 1

v c
n n
v c
d i i
i i
f f f
= =
= +


2 2 2
1 1
1 1
v c
n n
v v c c
i i j j
i j
v c
q f q f q
f f
= =
= +


(4.19, 4.20)


with the subscripts v and c denote quantities for the voids and cracks, respectively, and
an overbar symbol denotes a global quantity. The q
1
parameter does not require an
averaging procedure since it is a function of the stress triaxiality and hardening exponent
and these quantities are assumed to be homogeneous in the element. It is worth noting
that summation convention for tensor manipulation is temporarily turned off so far for
consistency with the literature.

.1 Implementation of the yield criterion

The numerical integration of damage-based constitutive models is made difficult
because their dependence upon the hydrostatic stress creates a yield surface with normal
vectors that can vary significantly over the time step, mandating very small time steps.
105
The integration scheme of Worswick and Pelletier (1998) has been employed in the
present work since it is designed for explicit time-integration and uses a forward-
Eulerian scheme. A brief synopsis of the integration procedure is provided here and the
reader is referred to Worswick and Pelletier (1998) for additional details.

At the beginning of an integration cycle or time step, the stress state lies on or inside
the yield surface. Using the macroscopic strain increment determined from the nodal
displacements of the element,
ij
E A , the macroscopic stress increment is assumed to be
elastic and a trial or predictor stress,
T
ij
E , is calculated as


T 0 e
ij ij ijkl kl
C E E = E + A (4.21)

where
0
ij
E is the stress at the beginning of the increment,
kl
E A is the strain increment
and
e
ijkl
C is the elastic stiffness tensor. The yield criterion in Eq. (4.18) is then evaluated
at the trial stress and if
T
( ) 0
ij
u E > , the increment is elastic-plastic and the stress must be
corrected so that the condition for plastic flow is met and
ij
( ) 0 u E = . Otherwise,
if
T
ij
( ) 0 u E < , the stress state is elastic and the trial stress is correct.

Since the normal to the Gurson (1977) yield surface changes rapidly due to its
exponential dependence upon the hydrostatic stress, the strain increment is divided into
sub-increments so that the yield surface normal ,
ij
o , can be accurately estimated at the
trial stress as

T
hyd T T 1
hyd
3
( ) sinh
2 2
d
ij ij ij d d ij
q
f

o o k o
o o
| | E
= E E +
|
|
\ .

(4.22)

where
ij
o is the Kronecker delta and
d
and
d
k are the degraded shear and bulk moduli
of the material.

106
A scalar variable known as the plastic multiplier, d , is used along with the yield
surface normal to return the stress to the yield surface and thus satisfy
ij
( ) 0 u E = . The
corrected stress can be expressed as

T
ij ij ij
d o E =E (4.23)

The plastic multiplier cannot be explicitly determined due to the nonlinear nature of the
yield surface and it is solved using an iterative Newton-Raphson method as


( ) ( ) ( )
1
/
( )
i i i
d
d d d
d d

+
| | u
=
|
\ .
(4.24)

until the value of d is found that satisfies the yield criterion, ( ) 0
ij
u E = , to a specified
tolerance. This approximation is very efficient with typically only one or two iterations
required to achieve convergence to a tolerance of 0.0001. Once the trial stress has been
corrected, the equivalent microscopic plastic strain increment of the matrix
material,
p
eq
c A , can be determined from work equivalence as


(1 )
p
ij ij p
eq
f
c
c
o
E
A =

(4.25)

The void nucleation, growth, shape evolution, rotation and coalescence sub-models are
then evaluated for the current strain increment and the global porosity and related
parameters updated for the next sub-increment.

.4 Degradation of the elastic moduli

The elastic constants will degrade with the evolution of the porosity in the material
and contribute to additional material softening. From a practical perspective, the
porosity in most metals is too small to be a significant factor, but should be accounted
for in the model for completeness. Since this is a secondary effect, the relations of Mori
and Tanaka (1973) for spherical voids can provide a good approximation for the
materials of interest and are well suited for random and clustered distributions of voids
(Kachanov et al., 1994; Cramer and Sevostianov, 2009). The degraded bulk and shear
107
moduli of a voided material are expressed as a function of their initial values and the
porosity as

4 (1 )
4 3
d
d
d
f
f
k
k
k

=
+

(1 )
6 12
1 ( )
9 8
d
d
d
f
f

k
k

=
+
+
+
(4.26, 4.27)

The elastic modulus and Poisson ratio can be computed from the bulk and shear moduli
using the standard isotropic relations. The elastic constants of the average particle in the
material are computed as a weighted average of the composition of each particle as


1 1
1

pt p
n n
p p p
ij j p
i j
f
f

= =
=



1 1
1

pt p
n n
p p p
ij j p
i j
f
f
k k
= =
=


1 1
pt p
n n
p
p ij
i j
f f
= =
=

(4.28-4.30)

where f
p
is the total particle volume fraction and the subscripts i and j correspond to the
particle type and particle number.

.5 Global coalescence and failure of the element

The onset of profuse void coalescence signaling failure of the particle field is easily
identified because the voids will rapidly link-up throughout the field to form a single
crack that encompasses the entire field. Failure of the particle field is identified by
homogenizing all of the individual voids into an equivalent void and evaluating the
coalescence model in Eq. (1.13) at each time-step. Since the porosity increases so
rapidly at the onset of profuse coalescence, the choice of global failure criterion was not
an important factor. Assuming a global failure porosity of 2% as suggested by Chen
(2004) for AA5182 sheet led to similar results identified by using Eq. (1.13) and by
visual means. Ultimately, the coalescence criterion of Pardoen and Hutchinson (2000) in
Eq. (1.13) was adopted since it does not introduce any additional parameters into the
model.

An additional advantage of this modeling treatment is that fracture is computed using
the same procedure used in standard damage-based constitutive models. In this manner,
the percolation model could be viewed as a sophisticated void evolution sub-model that
is used in a standard Gurson-type model of ductile fracture. To compute the plastic
108
limit-load for the global equivalent void, the void is assumed to be located at the center
of the element and the element is treated as the unit cell. The global aspect ratio is
computed as the weighted average of the voids and cracks in the material as


1 1
1 1
v c
n n
v v c c
i i j j
i j
v c
W f W f W
f f
= =
= +

(4.31)

The ligament spacing ratio of the global void is computed from the unit cell geometry as

1
3
d e
uc
f
W

_
q
| |
=
|
\ .
(4.32)

where / 6
uc
q t = for a cubic unit cell and
e
is the aspect ratio of the element with
respect to the principal loading direction. The plastic limit-load criterion in Eq. (1.13)
can now be evaluated using W , _ , the principal macroscopic stress,
1
, and the material
flow stress, o .


4.4 Microscopic models

The micromechanical modeling procedure used in the percolation routine is
presented in Figure 4.4. The term microscopic defines a quantity that is measured or
defined within the element such as the equivalent plastic strain. At the macro-scale, the
stress and damage are uniform and homogeneous and an equivalent void is used to
account for material softening. Within the element at the micro-scale, the void and
particle distributions are not homogeneous and vary within the element volume. The
proceeding sections will extend the existing void growth, shape evolution and
coalescence models to a general particle field subjected to an arbitrary loading.

109
Update Centroids of All Objects in the Element
using the Nodal Displacement Increment
Evaluate the Deformation gradient, Spin Tensor and Plastic Spin Tensor
for All Objects and Update the Object Orientation
Determine the Stress State within the Particles
and Evaluate the Nucleation Model
Nucleation?
Create Penny-Shaped Void
Yes
Test for Coalescence for ALL Void
and Crack Neighbours
Search for Neighbour Elements, Particles,
Voids and Cracks using Characteristic
Lengths
Evaluate Growth and Shape Evolution of the Voids and Cracks
Coalescence
or Overlap?
Merge Objects to
Create a New Crack
Update the Total Porosity in the Material
and Other Volume Averaged Statistics
Call the Percolation Subroutine and Update Global Porosity
Return Averaged Properties to the Constitutive Model to
Account for Material Softening and Failure of the Element
Yes

Figure 4.4: Flow-chart of the percolation modeling process

.1 Object kinematics

.1 Object translation

The objects (particles, voids and cracks) will translate and rotate within the element
during deformation. The location of the object is fixed within the dimensionless element
coordinate system but its position will change in the global system as the element
deforms as shown in Figure 4.5. The centroids of each object are updated at each time
step using the nodal displacements of the objects parent element and its shape functions
110
evaluated at the object centroid. The position of the object center in the global system is
calculated using Eq. (4.2).
x
y

u
1
v
1 v
2
u
2
u
3
u
4
v
4 v
3

Element coordinate system


Global coordinate system

Figure 4.5: Percolation element in the dimensionless element coordinate system (left) and
in the global coordinate system (right) where the voids and particles have translated and
rotated with the deformation of the element.

.2 Object rotation

The rigid body rotation of the objects is computed from the material spin tensor of
the element in Eq. (4.9). However, this rotation is not sufficient to describe the complete
motion of the objects within the material as they may rotate due to plastic deformation
within the element. In uniaxial tension, the material spin is zero but the objects will still
rotate to align themselves with the principal loading direction. This effect should be
considered because of its implications on the fracture mechanism because the objects
will orient themselves to create favourable conditions for nucleation, growth and
coalescence. Voids are more likely to coalesce when they are aligned transversely to the
principal loading direction and particles will crack transversely to the loading direction.
Particle and void rotation is especially important in torsion as shown in Figure 4.6 where
the voids and particles will rotate to align themselves at 45 to the loading direction.

111

Figure 4.6: Digitally compressed montage of 500 x 500 m
2
area of an undeformed AA6061
alloy (left) and a montage after being subjected to a strain of 98% in torsion (right)
(Agrawal et al., 2002).

The influence of plasticity-induced rotation has largely been neglected in the
modeling of damage-induced ductile fracture. Typically, the voids have been assumed to
remain stationary or to rotate with the material spin (Benzerga, 2002; Pardoen, 2007).
More recent models by Keralavarma and Benzerga (2010) and Schyvaerts et al. (2011)
have employed the nonlinear homogenization solution of Kailasam and Ponte Casteneda
(1998) in its reduced form for voids as used by Aravas and Ponte Casteneda (2004).
These solutions are based upon a rigorous homogenization scheme to obtain expressions
for the plastic spin rate in a composite material. This model has been shown to give very
good agreement with the void rotation as validated by Schyvaerts et al. (2011) using
finite-element simulations (Figure 4.7).

The model of Kailasam and Ponte Casteneda (1998) will also be adopted to account
for particle and void/crack rotations within the percolation element, albeit in a slightly
reduced version. The solution of Kailasam and Casteneda (1998) can account for any
number of different particle types in addition to the voids as well as their distributions.
However, the computation of the spin rates is not straightforward and requires many
intermediate calculations of fourth-order tensors and matrix operations. With thousands
of particles and voids within the percolation element and different particle types to be
112
evaluated, this can be a computationally expensive process. Therefore, the rotation rate
of each object type will be evaluated by neglecting interactions with the other object
types, considerably reducing the expression for the rotation rate.

-1.6
-1.2
-0.8
-0.4
0.0
0.4
0.8
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Shear strain
V
o
i
d

o
r
i
e
n
t
a
t
i
o
n

a
n
g
l
e


(
r
a
d
)

-
Unit cell
Rotation model with
unit cell data
W
0
= 1
W
0
= 1/3
W
0
= 3

Figure 4.7: Comparison of the analytical rotation model of Kailasam and Ponte Casteneda
(1998) with the void rotation obtained from unit cell simulations (Schyvaerts et al., 2010).


By also neglecting the distributional effects of the ellipsoidal objects within the
material, the average deformation rate, D
o
, and the average spin,
o
, of the local
representative ellipsoidal object can be expressed as


:
o o p
D E = A :
o o p
E O = OC (4.33, 4.34)

where O is the material spin rate of the element computed in Eq. (4.9) and A
o
and C
o

are fourth-order concentration tensors for the objects and are defined as


o M 1 1
[ (1 ) ( [ ] )]
o o
f

= A I S I L L (4.35)


( )
1
1
o M M
(1 ) (1 )
o o o o
f f

(
= +
(

C L L L S H (4.36)

where
o
f

is the object volume fraction; L
M
and L
o
are the fourth-order viscosity
(elastic) tensors of the matrix and object that are defined as a function of their shear and
113
bulk moduli as L = (3, 2). The fourth order tensors, S and , are Eshelby tensors
(1957) whose expressions are given in Appendix A. For the rotation of the voids and
cracks, the expressions in Eq. (4.35) and Eq. (4.36) are computed by setting
o
L = 0. In
the limit that the object volume fraction approaches zero, the solutions reduce to those
of Eshelby (1957).

The ellipsoidal object is defined as having semi-axes (a, b, c) where a >b >c are
corresponding vectors, n
1
, n
2
, n
3
. The microstructural spin, , in the global frame is
defined as

( )
2 2
3
3 2 2
, 1
1
: ( 1)
2
i j
i j o o
i j j i i j
i j
i j
i j
w w
w w
D w
w w
e
=
=
=
+
(
= O + + =

n n n n n n (4.37)

where w
i
are the aspect ratios of the object and are


1
c
w
a
=

2
c
w
b
=

3
1 w =
(4.38a-c)


Finally, the plastic spin of the object is expressed as


p
w O = O

(4.39)

and the rotation rates of the semi-axes of the ellipsoid are


p
i i
= O n n

(4.40)

The plastic strain will be zero if the material spin is zero and the object is aligned in
the loading direction. Otherwise, the object will rotate until it is aligned with the
principal loading direction. In the case that two of the aspect ratios are equal such as
1 2
w w = ,
12
p
O in Eq. (4.37) becomes indeterminate and is set to zero (Aravas and Ponte
Casteneda, 2004). If the object is spherical, the material is locally isotropic the plastic
spin vanishes since C
o
= 0.





114
.3 Implications of the rotation model on the percolation model

The adoption of the above rotation model significantly improves the physical
foundation of the percolation model and has implications on void nucleation, growth
and coalescence as these are all related to the direction of the maximum principal stress.
During plastic deformation, the predictions of the rotation model are improved by
computing the elastic tensor of the matrix, L
M
, using the secant shear modulus,
s
,
determined in Chapter 3 as part of the homogenization process to determine the stress
within the particles. In this manner, another level of integration is achieved in the
percolation model.

The rotation model can be evaluated for the bulk values of each object type and
applied uniformly to rotate each object within the particle field. However, this method
neglects the individual object dimensions and any strain-gradients that exist within the
element. This has been accounted for in the implementation of the model by evaluating
the material deformation gradient, spin tensor and strain increment at the centroid of
each object in the element using the equations in section 4.3.1. The equations for the
local plastic spin can then be determined using the above procedure by assuming that the
local strain rate is a close estimate to the local plastic strain rate. This implementation
provides an immense contribution to the model when shear loading is present as the
shear-induced rotational distribution is captured by the model. Objects located in the
center of the element will experience less rotation than the objects near the surface of
the element where the shear traction is applied. Overall, the treatment of particle and
void rotation in the percolation model is physically sound and superior to the existing
implementations of this model.

.2 Void growth and shape evolution

Modeling of the growth and shape evolution of the voids and cracks in the
percolation model requires certain assumptions since the evolution models are designed
for axisymmetric ellipsoids (spheroids) and not for general 3-D conditions. The main
challenge in adapting these models to the general case is their implicit dependence upon
115
the loading direction. By assuming a periodic distribution of axisymmetric voids, the
void aspect ratio can be defined as a state variable with a definitive initial value that can
evolve during deformation. However, in the general case, the loading direction is not
constrained to a specific direction and the aspect ratio is not an independent variable, but
a function of the loading direction. This point is best illustrated if we consider a penny-
shaped void that has just nucleated from a cracked particle. If the principal loading
direction happens to be aligned with the opening direction of the void, it will appear as a
penny-shaped void as viewed from the loading direction and there is no issue. If the
loading direction happens to be transverse to the void opening direction, the penny-
shaped void appears as an extremely prolate or needle-shaped void that will experience
negligible growth and shape evolution. Certainly the void growth and shape evolution
rules for a penny-shaped void do not apply in this case.

With this in mind, all void/crack growth and shape evolution will be governed by the
principal loading direction and how the geometry of the void appears from this
direction. The development of general three-dimensional void evolution models that
explicitly accounts for the viewing direction would be a welcome contribution to the
percolation model.

.1 Procedure for modeling void growth and shape evolution

Eliminating the void aspect ratio as a state variable requires the determination of the
void dimensions and shape relative to the principal loading direction. Consider an
arbitrary void or crack with the semi-axes (a, b, c) corresponding to the vectors n
1
, n
2
,
n
3
. The principal loading direction is defined by the vector u, as seen in Figure 4.8. The
distance from the center of the void to its surface along the direction u, is denoted R
1
. A
line-projection of the void is then taken with u as the viewing direction to obtain an
ellipse with semi-axes R
2
and R
3
that coincide with the p and n directions. The
reconstructed geometry of the void as viewed from the principal loading direction is an
ellipsoid with semi-axes, R
1
, R
2
, R
3
and the equivalent aspect ratio is defined as

2 3
u
R
W
R R
=
(4.41)
116

This aspect ratio is used to evaluate the void growth and shape evolution models but the
porosity of the actual void is used in these models and not the porosity of the equivalent
void.

n
(b) void projection (a) ellipsoidal void
p
u
(c) model ellipsoidal void
=
1 1
axi 2 3
R R
W =
R R R
E
1
1
R
3
R
2
R
3
R
2
R
1
R
, E
1
u
a
b
c
n
2
n
3
n
1

Figure 4.8: Procedure for modeling an arbitrary ellipsoidal void as an axisymmetric void
as viewed by the principal loading direction.

The void growth and shape evolution rates can now be determined from knowledge of
this effective aspect ratio. The void growth rate, / f f , is determined using the calibrated
void growth rule of Ragab (2004) using the unit cell correlations obtained in Chapter 2.
Similarly, the evolution rate of the effective void aspect ratio,

/ W W , is also determined
using the unit cell correlations in Chapter 2. These evolution rates were obtained for the
equivalent void and are not truly representative of the actual void since these growth
rates are not aligned with the semi-axes of the actual void. A clever application of the
unit cell geometric relationships from Appendix B enable the calculation of the radial
growth rates of the equivalent axisymmetric void as a function of the principal strain
rates as

1
1 2 3
1
1
2
3
R f W
E E E
R f W
| |
= + + + +
|
\ .
(4.42)

117

2 3
1 2 3
2 3
1
3
R R f W
E E E
R R f W
| |
= = + + +
|
\ .
(4.43)

The evolution rates of the equivalent void semi-axes are taken as the growth rates of the
actual void but in the directions u, p, n. The problem can now be treated as the
stretching of an ellipsoid along three arbitrary directions as shown in Figure 4.9. The
stretch values of the void in the u, p, n directions are


1
1
1
1
R
S
R
= +
2
2
2
1
R
S
R
= +
3 2
S S = (4.44a-c)

The solution for the arbitrary stretching of an ellipsoid requires an eigenvalue
solution where the eigenvalues are related to the new void semi-axes with their
orientation defined by the eigenvectors. The solutions for the line projection of an
ellipsoid, the distance from an ellipsoid center to its surface and for the arbitrary
stretching of an ellipsoid are presented in Appendix C.

(a) stretching of the void in the u,p,n directions
2
2
2
1
R
S =
R
+
1
1
1
1
R
S =
R
+
3
3
3
1
R
S =
R
+
(b) new void size and orientation
n
1
n
3
n
2
c
a
b

Figure 4.9: Stretching of a void in three arbitrary directions resulting in a new void size
and orientation

The novelty of this modeling procedure is that the evolution of the void orientation is
naturally accounted for in the model, and this is not captured using the standard growth
and shape models in the literature. The void will naturally grow and rotate itself to be
aligned with the preferential loading direction. If the directions of the semi-axes and
118
principal loading directions are aligned, the void will not rotate. Additionally, the
growth and shape evolution rates are calculated using the library of unit cell correlations
and therefore provide very good estimates for the evolution rates. Coupling the proposed
evolution model with the plasticity-induced rotation model will provide a strong
physical foundation for void evolution in the percolation model.

.3 Void nucleation

The dominant void nucleation mechanisms in ductile materials are particle cracking
and interface fracture (debonding). Particle cracking occurs primarily in strong, rigid
particles in a brittle-type fracture in the direction transverse to the maximum principal
stress. Void nucleation due to interface separation is strongly related to the particle
geometry and the interface strength between the inclusion and the surrounding matrix.
Similar to particle cracking, debonding occurs preferentially in the principal loading
direction unless the stress triaxiality is high.

From a modeling perspective, particle cracking and debonding will occur
simultaneously in a material and thus two nucleation models and their parameters must
be determined for each type of inclusion present. The identification of the appropriate
material parameters for even a single nucleation model is a complicated process as
nucleation is exceptionally difficult to measure experimentally. Consequently, most
treatments of nucleation rely on a phenomenological model such as a normal-
distribution (Chu and Needleman, 1980) with either assumed or calibrated parameters
obtained from a combination of finite-element simulations and experiment data (Butcher
and Chen, 2011).

In the percolation model, an approximate treatment for nucleation is required to
reduce the number of required nucleation parameters while retaining a physically sound
description of the nucleation mechanism. A proper accounting of nucleation requires
both a nucleation model and a treatment for determining the initial dimensions and
orientation of the nucleated void. For practical considerations, an emphasis will be
placed upon developing an accurate representation of particle cracking as this is the
119
dominant nucleation mode in many advanced automotive alloys. Additionally, the
nucleated void by particle cracking gives rise to a penny-shaped void whose growth and
evolution is relatively well understood from finite-element simulations in Chapter 2. The
implementation of a sophisticated particle debonding criterion into the percolation
framework is left for future work.

.1 Modeling void nucleation using penny-shaped voids

In particle cracking, the void is assumed to nucleate transversely to the maximum
principal direction as illustrated in Figure 4.9. This cracking behaviour is similar for
both oblate and prolate ellipsoidal inclusions with prolate inclusions particularly prone
to cracking. At the instant of nucleation, the particle is in two pieces that have yet to
appreciably separate and the void is said to be penny-shaped. From unit cell simulations
of penny-shaped voids, it has been shown that the growth of the void and response of
the material converges for aspect ratios lower than 1/100. This is an extremely beneficial
result that allows us to assume an aspect ratio for the nucleated void with the confidence
that is a physically sound approximation.

Figure 4.9: Void nucleation by the cracking of an oblate ellipsoidal inclusion. Nucleation
occurs in the same manner for a prolate ellipsoid.





120
.2 A nucleation model for particle cracking

A void nucleation model for particle cracking is fundamentally a model of brittle
fracture and must account for the stress within the particle, particle size, shape and
composition. It is well known that hard large particles are predisposed to cracking since
they are irregularly shaped and experience higher stresses via load sharing with the
matrix compared to their smaller neighbours. From a statistical perspective, a large
particle has a higher probability of containing internal defects than a smaller particle.
This effect can be clearly seen in the SEM micrographs in Figure 4.10 that show the
presence of surface cracks on the particles in an as cast AA5182 alloy. These cracks are
a result of the solidification process and act as preferential sites for particle break-up
during the hot and cold rolling processes. In a metal forming operation, these particles
are smaller but still contain cracks that can nucleate voids. Since brittle fracture is
associated with negligible plastic deformation, the particles can be assumed to remain
elastic during the deformation process, enabling the use of linear elastic fracture
mechanics to relate the stress state with the distribution of internal cracks within the
inclusions.


Figure 4.10: SEM observation from Moulin et al. (2009) showing the intermetallic particles
in as- cast AA5182 (left) and the typical morphology of the large Fe-rich particles (right)

To model the break-up of the large intermetallic particles in AA5182 during the
rolling process, Moulin et al. (2009) performed an extensive finite-element and
experimental study of irregularly shaped inclusions. It was observed that the stress
distribution within these particles is not homogeneous and that only certain regions of
121
the particle will achieve the stress required for particle break-up. These high stress
regions were termed the active volume of the inclusion, V
a
. Since the particles are
brittle and contain cracks, the fracture mechanism is one of mode I failure and is a
function of the critical material toughness, K
1C
, and the distribution of the cracks.
Assuming a homogeneous distribution of cracks with an effective length, a
eff
, the crack
will propagate according to Griffiths criterion for mode I fracture when the principal
stress in the particle satisfies the inequality


1
P
1
C
eff
K
a
o
t
>

(4.46)

This equation originates from the energy requirement that the stored elastic energy
within the inclusion must satisfy the energy released by the creation of the two fracture
surfaces. By assuming a homogenous distribution of cracks, the inclusion is expected to
contain a crack that is perpendicular to the principal loading direction that can propagate
to cause fracture. The critical length of this crack, a
c
, that must be present within the
active volume of the inclusion can be estimated as


2
1
1
C
c
th
K
a
o
o
t o
| |
~ ~
|
\ .

(4.47)

where is a geometry parameter to account for various effects such as crack blunting
and interactions and
th
o is a threshold stress for the maximum principal stress within the
active volume. The critical volume that contains a crack transverse to the loading
direction is defined as
3 3
c c
V a = and particle fracture will occur when the critical volume
is equal to the active volume (Moulin et al., 2009)


6 3
1

C
a c
th
K
V V
o
t o
| | | |
> >
| |
\ . \ .

(4.48)

Stress state and nucleation

The criterion of Moulin et al. (2009) in Eq. (4.48) is ideally suited for the modeling
of void nucleation since it is a volume-based metric that accounts for the stress within
122
the inclusion, the material toughness and has a strong physical foundation based upon
the energy required for crack propagation via the Griffith criterion. However, the model
cannot be readily implemented into the percolation model because the threshold stress
and active volume of the particles are functions of the particle size, shape, composition
and loading condition. In Moulin et al. (2009), they were determined by using finite-
element techniques.

Fortunately, since the inclusions in the percolation model are assumed to be
ellipsoidal, the stress and strain distributions within the inclusion are uniform as proven
by Eshelby (1957). This favourable result indicates that the activated volume in Eq.
(4.48) must be equal to the volume of the inclusion and that the threshold stress must be
equal to the principal stress in the inclusion. The homogenization scheme of Chapter 3
for particle-reinforced plasticity can now be used to determine the stress within the
particle as a function of the particle size, shape and composition. The fracture criterion
can now be expressed in terms of a nucleation stress


p
1 N
o o >

(4.49)

*
1
*
1 1
1/6
1
1
C
N C C
p
K
K K
V
o o o
t
= = ~


where
*
1C
K is the effective critical toughness of the particle material and is the sole
parameter in the nucleation model. The K
1c
value can be estimated if the composition of
the particle is known or else it can be identified by calibrating the effective
*
1C
K with the
experimental nucleation data. A realistic range for the value of K
1c
for brittle materials is
1 10 MPam
1/2
. The geometry factor, o , does not have to be determined and can be
lumped into the
*
1C
K

value.

123
0 100 200 300 400 500 600
0
500
1000
1500
2000
2500
N
u
c
l
e
a
t
i
o
n

s
t
r
e
s
s


o
N


(
M
P
a
)
Particle volume (m
3
)


K
1C
= 2 MPa-m
1/2
K
1C
= 4 MPa-m
1/2
K
1C
= 6 MPa-m
1/2

Figure 4.11: Variation of the nucleation stress with particle size and toughness

The variation of the particle nucleation stress with the volume is presented in Figure
4.11. The criterion captures the particle size-effect where small particles nucleate at high
strains while large particles nucleate at low stresses and are roughly independent of the
particle size. The nucleation model also predicts that brittle phases are more likely to
crack than more ductile phases. It should be emphasized that the nucleation stress is
deeply coupled with the particle size, shape, composition, fracture toughness and the
applied stress state as

1
( , , , , , )
p p p p p
N N c ij
f W E v K o o = E

(4.50)

and that successful prediction of nucleation relies upon the interrelationships between
these variables.

The principal contribution of this nucleation treatment is that it is physically realistic
and does not contain any calibration parameters. The percolation model is now
completely deterministic where fracture is a natural consequence of void nucleation and
evolution and no adjustable parameters are employed. The particle distribution, stress
state and material properties are solely responsible for the fracture process. All
124
parameters are intrinsic material properties that can be quantified or estimated such as
the K
1c
value.

.3 Determination of the initial dimensions for a nucleated void

A three-dimensional ellipsoidal inclusion with semi-axes (a, b, c) and orientation
vectors (n
1
, n
2
, n
3
) is embedded within a ductile matrix as shown in Figure 4.12. The
matrix and inclusion are subjected to an external loading that gives rise to a maximum
principal stress,
1
, along the direction, u. The loading is severe enough to induce the
particle to fracture or debond from the matrix to form a void. It is assumed that the
particle cracks through its center in the direction transverse to the principal loading
direction. The dimensions and orientation of the void in this plane are obtained by
sectioning the inclusion to form an ellipse on the p-n plane where the vectors p and n
correspond to the semi-axes of the ellipse denoted a
s
and b
s
. The void height, c
W
, is
determined from the assumed aspect ratio for the void at nucleation, W
N
. For penny-
shaped voids, any value for W
N
< 0.01 can accurately represent the void. The height of
the void can then be calculated as


axi W N N s s
c W R W a b = = (4.51)

Finally, the void semi-axes, (c
W
, a
s
, b
s
), and their respective directions, (u, p, n), are
ordered so that the semi-axes are now defined as (a, b, c) with directions (n
1
, n
2
, n
3
) and
a b c. The inclusion that nucleated the void is then removed from further analysis as
it is assumed that the particle can only crack once, has a negligible ability to reinforce
the material (true for particle cracking) and that any pieces of the broken particle do not
significantly impact the subsequent growth of the void (Lassance et al., 2007).

125

(a) ellipsoidal inclusion in a ductile matrix (b) cross-sectional area of the nucleated void

(c) penny-shaped void geometry (d) ordered void geometry and orientation

Figure 4.12: Procedure for modeling void nucleation

When the principal stresses are comparable as in equal-biaxial tension the particle may
crack transverse to either direction. In this case, the cracking direction is selected as the
direction that has the largest particle aspect ratio because prolate particles are more
prone to cracking than oblate particles.

.4 Neighbours element and object

Since fracture is a local phenomenon originating within specific regions of the
material the objects must have knowledge of their surroundings. This is accomplished
by creating neighbour lists on the macro- and the micro-scale through the use of
characteristic length parameters. If the distance from the centroid of an object to the
centroid of a neighbour object is less than the characteristic length, it is appended to a
axi W N N s s
c W R W a b = =
126
neighbour list of that object. The size of the characteristic length controls the sphere of
influence of the object. In the percolation model, up to four distinct characteristic
lengths can be input to determine element neighbours, particle neighbours and
void/crack neighbours.

At the global level, a characteristic length can be defined for the elements, L
e
, to
locate the neighbouring percolation elements and identify the objects within them for
potential interactions. This characteristic length is not a significant parameter except for
crack propagation prior to final fracture as it enables the crack front to progress through
neighbouring elements. For the most part, this parameter is used to reduce the number of
interaction searches for each time step by limiting the size of the global neighbourhood.
The simplest choice of selecting element neighbours is by using common nodes so that
only the surrounding elements are considered. If more elements are required, the center-
to-center distance between elements is used to identify the neighbours.

Within the global neighbourhood defined by the element length, up to three
additional characteristic lengths can be adopted for:

- Particle-particle neighbours: L
pp

- Particle-void/crack neighbours: L
pv

- Void-void, void-crack, or crack-crack neighbours: L
vc


It is important to have different characteristic lengths to be able to easily test for
different interactions. For example, in modeling void nucleation within particle clusters,
the particle-particle characteristic length will play a central role which is not likely to be
the same as for void-crack interactions. The particle-void characteristic length could be
used in modeling nucleation since the proximity of a particle to a crack will increase its
propensity to nucleate a void. In the simplest case where no interactions are considered
(isolated voids and particles) a single length can be used to define the local
neighbourhood. In this case, any reasonable choice for this value such as five average
diameters will be sufficient to identify the neighbouring particles, voids and cracks for
nucleation and coalescence.
127
In the event of void coalescence, the new void inherits the neighbours of its parent
particle and the new void is added to the lists of its neighbours. Similarly, when two
voids or cracks coalesce, the new crack will inherit the neighbour lists both parents and
is added to these neighbours lists as well. In this manner, coalescence will naturally
sweep through the particle field, inheriting neighbours and coalescing with them as
cracks propagate through the elements. The advantage of this method is that it is
computationally efficient, physically sound, and relegates the characteristic length to a
third-tier parameter. Of course, if the computational resources are available, every object
within every element can be compiled into a neighbour list and no characteristic lengths
are required.

.5 Void coalescence

As with void growth and shape evolution, departing from a periodic microstructure
containing axisymmetric voids in favour of a general distribution of 3-D ellipsoids
introduces challenges for void coalescence modeling. The plastic limit-load coalescence
model has been widely used in the literature as well as validated in Chapter 2. The
model is robust and can give very strong predictions of coalescence if one happens to
have two identical voids horizontally aligned transverse to the principal loading
direction. Certainly, this is not the general case and some modifications are required to
utilize this criterion in the percolation model. Butcher and Chen (2009) have appended
the plastic limit-load model in a prior 2-D version percolation model. Scheyvaerts et al.
(2010) considered the possibility of coalescence on angles in a periodic microstructure
in plane strain tension with shear. The modeling of void coalescence in a general
particle field will be separated into four categories that require consideration:

- Determination of the stress state when the voids are located in different
elements

- Identification of the coalescence plane and appropriate stress transverse to the
arbitrarily oriented ligament

128
- Determination of the spacing ratio and equivalent void geometry for two
arbitrary ellipsoidal voids

- Merging procedure to create the coalesced crack (void)

.1 Coalescence between elements

The stress state within each percolation element is assumed to be uniform since the
presence of the voids and particles are not explicitly modeled in the finite-element
program. Therefore, if the coalescence criterion is to be evaluated for two voids located
in different parent elements, the average stress tensor of the elements is computed as


e1 e2
2
ij ij
c
ij
E + E
E = (4.52)

This is a rather simplistic approach to estimating the stress tensor in the region between
the two elements; however, it is reasonable if the element resolution is such that the
variation in the stress state between elements is small. Future work could take advantage
of the integration points within the element to better estimate the stress in a specific
region of the element.

.2 Identification of the maximum stress transverse to the ligament

The plastic limit-load coalescence model assumes that the neighbouring voids are
aligned so that the center-to-center vector is transverse to the maximum principal stress.
Obviously this is not the general case and this condition must be relaxed to be applied
for a general void distribution.

First, consider two arbitrary ellipsoidal voids subjected to an arbitrary loading where
the principal loading direction is not transverse to the inter-void ligament as defined by
the center-to-center vector, w, in Figure 4.13. In a two-dimensional model, a simple
stress transformation can be used to obtain the stress transverse to the ligament and the
evaluation of the criterion can proceed. In the three-dimensional case, a plane exists that
is transverse to the ligament and the stress will vary within this plane. The maximum
129
tensile stress within this plane must be determined to evaluate the plastic limit-load
criterion. This stress will be referred to as the maximum in-plane tensile stress and is
denoted,
1
ip
E . The second in-plane stress that is transverse to both
1
ip
E and the ligament
is denoted,
2
ip
E .


Figure 4.13: Void coalescence geometry showing the in-plane tensile stress values
transverse to the inter-void ligament vector, w.

The in-plane stresses can be determined by selecting an arbitrary coordinate system
using the vectors u, v, and w. Within this plane exists an angle, , that will rotate u and v
about w to be aligned with the in-plane stress vectors. The solution for these vectors has
been framed as an optimization problem to enable its extension in future work to
minimize the plastic limit-load by accounting for the variation in the dimensions of the
voids as a function of . For an assumed vector, v = [v
1,
v
2,
v
3
], and center-to-center
vector, w = [w
1,
w
2,
w
3
], the trial in-plane stress is defined using a vector, s = [s
1,
s
2,
s
3
],
as a function of by


1
( )
ip T c
ij
u E = E s s (4.53)

130
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) ( )
2 2
1 1 1 1 1 2 3 2 1 3 2 3
2 2
2 1 2 3 1 2 2 2 2 3 1 3
2 2
3 1 3 2 1 2 3 1 2 3 3 3
(1 ) cos (1 cos ) sin (1 cos ) sin
(1 cos ) sin (1 ) cos (1 cos ) sin
(1 cos ) sin (1 cos ) sin (1 ) cos
s v v w v v v w v v v w
s v v v w v v w v v v w
s v v v w v v v w v v w
u u u u u
u u u u u
u u u u u
= + + + +
= + + + +
= + + +


An iterative Newton-Raphson search quickly converges to the solution by iterating with
respect to . The value of in the n-th +1 iteration is


1
1 2 2
1
( ) /
( ) /
ip
n n ip
d d
d d
u u
u u
u u
+
E
=
E
(4.54)

Once has been determined, the vectors u and v can be rotated to the optimal directions
and the in-plane stress values can be determined using the standard stress
transformations

1
ip T c
ij
E = E u u

2
ip T c
ij
E = E v v (4.55)

In many cases, void coalescence will be triggered by the maximum in-plane tensile
stress but coalescence can also occur in the second in-plane stress direction depending
on the void alignment and stress state. It is cautioned that a preoccupation with the
maximum tensile stress governing coalescence can lead to erroneous predictions in
equal-biaxial stretching when
2 1
ip ip
E ~ E . In the interest of being conservative,
coalescence will also be evaluated in the second in-plane stress direction whenever
2
0
ip
E > . Knowledge of the tensile stress along the vector, w, is not important even in
triaxial loading since the void spacing ratio will be zero as the voids are aligned in this
direction and the plastic limit-load to coalescence will be infinite. This type of
coalescence is rare and is known as necklace coalescence and cannot be predicted by
the plastic limit-load model.

.3 Effective geometry for the evaluation of coalescence

A procedure must be developed to effectively homogenize the two voids into an
equivalent geometry of two identical voids to be amenable to the evaluation of the
coalescence model. The plastic limit-load criterion requires knowledge of the void
aspect ratio, W, spacing ratio, , and the maximum tensile stress transverse to the
131
ligament. The maximum in-plane tensile stress has been determined in the optimization
procedure above. The void spacing ratio is readily determined from the center-to-center
distance,
2 1 2 c c c c
d = x x , between the void centroids, x
ci
, as


1 2
2
w w
c
c c
R R
d
_
+
= (4.56)

where R
w1
and R
w2
are the distances from the voids, measured from their center to their
surface along the center-to-center vector, w, as seen in Figure 4.14. The center-to-center
distance is used for modeling coalescence so that the spacing ratio reduces to its original
definition in the plastic limit-load formulation when the voids are aligned and have the
same orientation. It is important to state that
c
in Eq. (4.56) is not the absolute
minimum spacing ratio between the two voids. The minimum spacing ratio is not
employed because it is computationally expensive and requires an iterative solution for
every void-neighbour pair in every element at each time step. Overall, the center-to-
center spacing ratio is a good approximation to the minimum spacing ratio provided the
voids are not very close together and have similar orientations.

The principal approximations and uncertainties introduced into the coalescence
model are related to the determination of the equivalent aspect ratio. The voids have
arbitrary sizes and orientations and are far from the idealized geometry of two identical
spheroidal voids. The aspect ratio of the equivalent axisymmetric void is defined as


u
eq
axi
R
W
R
=

(4.57)

where R
u
is the radius along the in-plane stress direction and R
axi
is the equivalent
axisymmetric radius. The height of the equivalent spheroidal void, R
u
, can be estimated
by averaging the distance from the void centers to their surfaces along the in-plane stress
direction as

1 1
2
u u
u
R R
R
+
=

(4.58)

132
To determine the equivalent axisymmetric radius, each void is first sectioned
transverse to the in-plane stress direction to obtain an ellipse with semi-axes (a
si
, b
si
) as
shown in Figure 4.14. For clarity, R
w1
and R
w2
are shown to be aligned with the section
of the ellipse in Figure 4.14 but the semi-axes of the sectioned ellipsoid are generally not
equal to the R
w
distance due to the void orientation.

w
ip
1
, E u
x
y
z
Void 1
ip
2
, E v
c
1
E
Void 2
c
2
E
b
s1
a
s1
a
s2
b
s2
R
u 1
R
u 2
R
w2
R
w1

Figure 4.14: Void coalescence geometry showing the identification of the relevant
dimensions

Since the locations of the semi-axes in the section planes may be different for the two
voids, these dimensions cannot be directly averaged. Instead an equivalent axisymmetric
radius is defined for each void as


1 1 1 axi s s
R a b

=
2 2 2 axi s s
R a b

= (4.59)

and the axisymmetric radius of the equivalent void and its aspect ratio are computed as


1 2
2
axi axi
axi
R R
R

+
=


1 1
1 2
u u
eq
axi axi
R R
W
R R

+
=
+
(4.60, 4.61)

The coalescence criterion in Eq.(1.13) can now be evaluated using W
eq
,
c
,
1
ip
E , and the
matrix flow stress, o . The geometry of the voids used in the coalescence model is
shown in Figure 4.15.
133


Figure 4.15: Equivalent void geometry for evaluation of the plastic limit-load coalescence
model

.4 Creation of a new crack: Merging operation

When the coalescence criterion is satisfied for two voids/cracks, a new crack must be
created by merging the two voids together. The merging operation plays a significant
role in the fracture process since the size of the resulting crack will influence
coalescence with other neighbouring voids. The simplest merging operation is to use the
maximum void dimensions of the two voids to make a large crack that encompasses
both voids as in Worswick et al. (2001). The disadvantage of this model is that the crack
size is uncharacteristically large and coalescence quickly spreads throughout the particle
field.

The bounding box method presented in Figure 4.16 is used to define the new crack
dimensions. The crack is assumed to be oriented with its semi-axes along the u, v, w
vectors identified in the previous section for the in-plane stresses. The dimensions and
points that will define the bounding box for the new crack must be identified in the three
planes defined by u, v, w. Each void bounding box will have four points that lie above
and four points that lie below the plane and the distances from these points to the plane
134
are calculated as D
a1i
, D
b1i
for void 1 and D
a2i
, D
b2i
for void 2. The semi-axis of the crack
in this plane is then computed as


1 2 1 2
max( , ) max( , )
1...4
2
a i a i b i b i
D D D D
R i
+
= = (4.62)


Figure 4.16: Two-dimensional schematic of coalescence between two voids showing their
bounding boxes and the creation of the new crack.


This process is repeated for each plane to obtain the three semi-axes of the crack. The
center of the crack can be determined by constructing the bounding box for the crack by
using one void centroid as the reference point. The centroid is the midpoint of the
bounding box coordinates.

The maximum distances above and below the planes are employed to ensure that the
crack will progressively enlarge as it coalesces with other voids and cracks. The average
void dimensions could be used but only for void-void coalescence and not for void-
crack or crack-crack coalescence as the amplification effect of the crack size will be
135
lost. If the voids that compose the crack belong to different parent elements, the parent
element of the crack is identified as the element with the minimum distance from the
element center to the crack centroid.

Overall, this merging procedure is about as good as one can achieve without
resorting to an optimization algorithm to determine the minimum ellipsoid that envelops
both voids. This merge procedure works very well for most void geometries except for
coalescence between a large and small void where the crack size is overestimated as
shown in Figure 4.17. Fortunately, this does not become significant until large cracks
have first formed which typically does not occur until prior to fracture when coalescence
is sweeping throughout the element and failure is imminent. Nevertheless, future work
could extend the merge operation to account for the relative sizes of the voids in
determining the size of the new crack.


Figure 4.17: Schematic of coalescence between a large and small void and the resulting
crack geometry.


.5 Mandatory coalescence: Object intersections

It is not uncommon for voids and cracks to overlap as they evolve with deformation
or coalesce with their neighbours. This type of void/crack impingement is not realistic
since the material between the voids will fail before the voids come into contact
(Thomason, 1990). As a result, when a void or crack overlaps with another neighbouring
136
void, the coalescence process is enforced using the merging process described above.
The minimum distance algorithm of Lin and Han (2002) described in Chapter 5 is
evaluated prior to testing for coalescence at each time step. There are no restrictions
placed upon particles coming into contact with voids, cracks or other particles for the
sole reason of reducing the number of intersection tests. Each percolation element may
contain thousands of particles and their overlapping is of secondary importance
compared to the voids. A worthwhile extension would be to treat the particles as rigid
bodies that cannot intersect to model the creation of deformation-induced particle
clusters as the particles pile up with their neighbours. The proximity of a particle next to
an existing crack would also promote premature void nucleation.

4.5 Summary

Damage percolation modeling provides a unique tool to study and understand ductile
fracture in heterogeneous materials and the model has made significant in-roads into
developing a physically sound framework to model fracture from first principles. The
significant features of the present percolation model are:

- The model accepts a three-dimensional particle distribution of arbitrary
ellipsoidal particles and voids.

- The particle fields are mapped to finite-elements to capture the development of
the complex stress- and strain-gradients that develop in the microstructure.

- The initiation and evolution of damage at the micro-scale controls the bulk
material behavior of the element and ultimately the structure in the finite-
element model

- The Gurson-based yield criterion has been calibrated through the development of
a library of correlations to model void evolution.

- The stress state within the particles is determined based upon a homogenization
scheme and is a function of the particle shape, size and composition.
137

- A sophisticated void nucleation model for particle cracking is used to predict
nucleation as a function of the particle stress (from homogenization theory),
composition, fracture toughness, size and shape.

- Physically sound treatments for void growth, shape and coalescence have been
developed that place no restriction on the void shape or orientation

- The void growth, shape evolution and coalescence models have been calibrated
and validated from an extensive study of voided unit cells.

- An advanced rotation law based upon homogenization theory has been adopted
to account for the plastic spin of the particles, voids and cracks within the
material

- No material calibration parameters are required by the model or have been
introduced. All parameters are directly related to the material such as the particle
composition, flow stress relation of the matrix, etc. The only parameter that
could be calibrated with experiment would be the fracture toughness of the
particles if it cannot be predetermined. Otherwise fracture is a natural
consequence of microstructure evolution.

The theoretical development of the percolation model concludes with this chapter.
The subsequent chapters will develop a particle field generator to populate the
percolation elements with statistically representative particle fields. Finally, the
complete finite-element percolation model will be evaluated in Chapter 6 to predict
damage initiation and failure in an aluminum-magnesium alloy.



138
Chapter 5 Development and Integration of a Particle Field
Generator into the Percolation Model

5.1 Acquiring measured particle fields

There are two distinct challenges in the development of a multi-scale damage
percolation model: (i) the development and application of the micromechanical models
used to predict damage initiation and evolution and (ii) obtaining the experimental
particle distributions for the model to use. While the development of the
micromechanical models and implementation of the percolation model is a significant
endeavor, obtaining the experimental particle distributions is equally as challenging, and
arguably more tedious.

Throughout the development phase of the percolation model, an emphasis has been
placed on generalizing and automating the model to expand its use from an academic
exercise to a predictive model that can be applied to a practical metal forming operation.
Despite these efforts, obtaining the required experimental particle distributions remains
a difficult process, especially for 3-D fields, and will often not be feasible due to cost
considerations or access to the required equipment. For most materials, the basic
statistics that describe the particle distribution are known or can be reliably estimated
such as the average particle size, shape, spacing and volume fraction. In light of this
assumption, it is of great interest to develop a particle field generator that can create
representative particle fields using the experiment data (if available) or from assumed
statistical distributions. The incorporation of a particle field generator into the
percolation model will enable fracture to be treated in a stochastic manner where the
variation in the microstructure will lead to a range of predicted fracture strains.
139
The organization of this chapter will discuss the basic procedure for generating the
particle field within the percolation model followed by sections discussing the various
techniques and modeling treatments of the algorithm. Finally, the particle field generator
will be used to generate representative particle fields of an aluminum-magnesium alloy
and validated using the experimental results in the literature.

5.2 Procedure for generating a particle field

The procedure for generating a particle field begins with the finite-element model.
First, the algorithm identifies the elements that have been set as percolation-type
elements from the input file generated using the finite-element software, LS-DYNA.
The node information of these elements is used to identify the corners of the volume to
determine a single element block that is termed the global percolation volume. The
volume is then populated with objects based upon their packing densities and arranged
into clusters. The element volume is then decomposed into its constituent elements and
the individual objects are assigned to their parent element as shown in Figure 5.1. These
steps are performed automatically at the start of a finite-element simulation as part of the
pre-processing routine of the percolation model. The subsequent sections will discuss
the decisions and modeling treatments required in each stage of the particle generation
process, starting with the sampling techniques used to define the objects.
140
1
2
3
4
5
6 7
8

y
x
z
2a
2b
2c
(a) Percolation elements identified from
the global finite element mesh
(b) Percolation elements combined to create a global
element to define volume for particle field generation
(c) Particles and voids are generated
within the volume
(d) Particles and voids are assigned to parent elements

Figure 5.1: Procedure for generating a particle field within a block of percolation elements

.1 Particle field basics
The particle field may contain any number of particle types in addition to the voids.
No restrictions are made on the properties of the particles and voids except that they
must be ellipsoidal. Since the voids and particles are modeled as ellipsoids and
141
generated using the same procedure, they will simply be referred to as objects in the
development of the particle field generator. The generator requires the creation of
random numbers from a uniform distribution that are denoted as r(a, b), where a and b,
correspond to the bounds for the distribution. The particle field generator has been
constructed to populate either 8-node brick or 4-node plane isoparametric elements. The
procedures in this chapter are presented only for the brick element to provide a general
3-D solution. The procedures for the 2-D element are readily determined by constraining
the solution to one plane.

.2 Generation of random variables

The particle field generator algorithm recreates the particle and void distributions
through knowledge of the probability density functions (PDF) for the relevant variables
such as the object semi-axes (radii), orientation angles and inter-object spacings. If the
approximate distribution is known, the variables can be generated using one of the
available density functions:

- constant value (no distribution)
- uniform distribution
- normal (Gaussian) distribution
- log-normal distribution
- Weibull distribution (two or three parameter variants)
Bi-modal variants of these distributions have also been implemented. However, if the
experimental distribution is known, the random variable can be generated from this
distribution using a rejection-sampling technique. This method provides greater
confidence in the resulting particle field since the variables of interest will have the
same statistical distributions as experimentally observed.




142
Rejection-sampling

A rejection-sampling algorithm has been adopted to obtain a random variable, x,
from its experimental PDF. First, the curve data that defines the density function is pre-
processed to determine maximum probability in the curve, P
max
, as well as the left and
right bounds for the variable, defined as a and b. The operation of the rejection-sampling
algorithm is as follows:

- Randomly generate a trial value for the variable, x
T
, that lies within [a, b].
- Evaluate the probability at the trial value, P(x
T
), by interpolating the density
curve data
- Generate a random number, u, that lies within the range [0, P
max
]
- Accept x
T
if u P(x
T
). Otherwise, reject x
T
and repeat the process with a new
trial x
T
.
A graphical representation of the rejection-sampling algorithm is presented in Figure
5.2.

x
Target pdf
a b
P(x)
P
max
x
T
P(x
T
)
u = r
u
(0,P
max
)
Accept x
T
if u P(x
T
)

Figure 5.2: Schematic of the rejection-sampling algorithm used to determine a random
variable, x, from a target probability density function (PDF)



143
.3 Generation of the global percolation volume

The nodal coordinates of all of the percolation elements are known at run-time based
upon the list of percolation element ids input into the finite-element input card deck.
The corner nodes that define the global element block are sorted from the nodal data and
the positions of these nodes are denoted as
G
x . The half-lengths of the global element
block, , ,
G G G
a b c , are computed using the procedure in Chapter 4. If the finite-element
simulation uses 2-D plane stress elements, the global element block will be defined as a
2-D plane element and its side lengths and properties evaluated accordingly.

The center of the global percolation volume is computed by averaging the centroids
of the individual elements that comprise the volume as


1
1
e
n
G a
c c
e a
x
n
=
=

x

(5.1)

where
e
n is the number of percolation elements

The outward facing normal vectors, n
i
, for each face of the global element must be
determined to create the constraints that define the volume during the particle generation
process. The equations for each plane that describes a face of the elements is calculated
as

0
p
d = n p -

(5.2)

where p is a point on the plane and d
p
is calculated by evaluating the equation using
known point on the plane such as a nodal point.

.4. Object generation

.1 Number of objects

The number of each type of object to be generated within the volume is determined
from the measured object densities,
i
, that are a measure of the number of objects per
144
unit volume. This metric is chosen because it is easily measured experimentally
compared to using volume fractions such as the porosity for the voids. This metric will
lead to slight variations in the global volume fractions of the objects since the sizes of
the objects will vary according to their distributions. Once the volume of the element
block, V
G
, has been computed, the number of objects of the i-th object type, N
i
, to be
created is

( ) /
i G
N objects vol V =

(5.3)

The use of object densities also simplifies the implementation of the algorithm because
the generation process can be terminated upon reaching the desired number of objects.
Alternatively, if a target volume fraction of an object is used to control the generation
process, an iterative procedure would be required to continually remove and create
objects to meet this target.

.2 Generation of a point inside the element block

A point, p = (x, y, z), within the element block can be generated by selecting three
random numbers, r
1-3
, to yield


1
(2 (0,1) 1)
G
G c
x =a r x +

(2 (0,1) 1)
G
G c
y =b r y +


3
(2 (0,1) 1)
G
G c
z =c r z + (5.4)


.3 Generation of object semi-axes

All objects are general three-dimensional ellipsoids with semi-axes R
1
, R
2
, R
3
. Since
the density functions for these radii must be obtained experimentally, they correspond to
a specific direction. For sheet metals, these are the rolling, transverse and short
transverse or through-thickness directions. These axes could also correspond to the
directions of anisotropy in a general material. The directions for the rolling, transverse
and thickness directions are denoted as r, t
r
, and t
t
, respectively. After the radii have
been generated, they are sorted and reclassified into the traditional form for ellipsoids
with a, b, c, as the semi-axes, where a > b > c.


145
.4 Generation of object orientation

Unlike previous percolation models (Worswick et al., 2001; Chen, 2004; Orlov,
2006), no restrictions are placed upon the orientation of the objects within the field. In
sheet materials, the particles and voids tend to naturally align themselves with the
rolling direction. To account for the orientation of the objects, a preferential direction
can be specified along with the density function for the solid angle,
r
, from which the
objects semi-major axis deviates. An example of this effect is shown in Figure 5.3.


Figure 5.3: Orientation of an object relative to the rolling direction by a solid angle,
r
.

.5 Generation of objects within clusters

.1 Object generation sequence

The sequence for generating the objects of a specific type can be random or follow a
specific order, such as first generating all of the particles of type 1 followed by all of the
particles of type 2 and then the voids. The object generation sequence plays a role in
determining the degree of clustering in the material by controlling the population of
available neighbours. If the object type is randomly determined, the degree of clustering
of the objects will be higher as the pool of potential neighbours is kept to the current
number of objects. Early in the generation process, the number of available neighbours
is small and the same neighbours could be selected repeatedly, leading to excessive
146
clustering. Conversely, if an entire population of particles is generated before a
population that requires neighbours such as voids, the voids will not be as near each
other as they have selected their neighbouring particle from a larger pool of potential
neighbours.

.2 Distance between objects

Accurate knowledge of the density distributions for the distance between object types
is critical to recreate the spatial distribution of the objects within the material.
Experimentally, the distance between objects is commonly measured using the inter-
particle dilatational spacing (IPDS) (Worswick et al., 2001; Chen, 2004; Orlov, 2006).
This value is determined by calculating the distances between a particle or void and all
of its surrounding neighbours. The minimum distance is called the IPDS and the process
is repeated for all of the particles within the microstructure to develop the density
function. Naturally, an IPDS distribution exists for each object and neighbour
combination in the material such as IPDS distribution for voids and particles of type 1,
or the distribution between particles of type 1 and type 2 and so forth. In the present
work, the IPDS will be simply referred to as the inter-object spacing, denoted D
s
, to
provide a general term that can refer to voids or particles.

There is some uncertainty surrounding the calculation of the inter-object spacing.
Typically, the spacing is taken as the minimum of all of the center-to-center distances
between an object and all of its neighbours. Using this definition, D
s
= min(D
1
,D
2
) in
Figure 5.4. However, by inspection of Figure 5.4, it can be seen that the minimum
spacing, D
min
, between two objects can be very different compared to the center-to-
center distance. The center-to-center spacing becomes a better estimate of the minimum
spacing when the objects are far apart and have similar orientations. While more
difficult to compute, the absolute minimum distance should provide a better
representation for the object spacing than the minimum center-to-center distance
because it accounts for the size of the objects. For this reason, the minimum inter-object
spacing, D
s
, is used in the generation of the particle fields.

147
D
1
Object
Neighbour 1
Neighbour 2
D
2
D
mi n

Figure 5.4: Schematic showing the possible measures used in calculating the inter-object
spacing.

.3 Clusters

The voids and second-phase particles/inclusions in sheet materials are preferentially
aligned in the rolling direction. Consequently, the rolling process creates aligned
clusters or stringers where neighbouring voids and particles are linearly aligned in the
rolling direction. In this work, we will consider two types of clusters: general and
stringer. A general cluster is formed by simply placing one object (void or particle) next
to an existing object at a distance specified from the density function for the inter-object
distances between the two object types. Stringer-type clusters are similar to the general
type with a constraint on the location that an object can be placed next to a neighbour as
shown in Figure 5.5. An object will be placed within a general cluster when a random
number from zero to one is greater than the probability of clustering, P
c
, such that


(0,1) P
c c
r > (5.5)

If the object has been selected to be within a cluster, then the cluster will be of the
stringer-type if

(0,1) P
s s
r > (5.6)

where P
s
, is the probability the object will be in a stringer-type cluster.


148
Object is not in a cluster: (0,1) P
c c
r <

If the trial object does not lie within a cluster, the centroid of the object, x, is
randomly selected within the percolation volume using Eq. (5.4).

Stringer-type cluster: (0,1) P
c c
r >

and (0,1) P
s s
r >



The following process is used to create a stringer-type cluster:

- An existing object is randomly selected to be the neighbour of the trial object
being generated and is located at
N
x
- The center-to-center vector between the two objects is denoted as
( ) , ,
c c c
l m n c =
and the rolling direction vector is defined as
( ) , ,
r r r
l m n r =

- The probability density function for the stringer solid angle,
s
, between c and r
is known for the two object types

- The probability density function is known for the inter-object distance, D
s
,
between the two object types in the stringer

- The center-to-center vector is determined that satisfies the stringer angle
requirement

- The minimum distance is computed between the ellipsoid and its neighbour. An
iterative process is performed until the minimum distance along this vector is D
s
.
This process ensures that the two objects do not overlap and identifies the
centroid, x, of the object.

- If no suitable neighbour is found after a set number of attempts, a new trial
object is created with a different orientation and dimensions. The neighbour
search is then resumed.

149

Figure 5.5: Schematic showing the creation of a stringer-type cluster

General cluster: (0,1) P
c c
r >

and (0,1) P
s s
r <



If the new object is located within a general cluster, the process is the same as for the
stringer cluster. The only difference between the models is in computing the center-to-
center vector between the two objects. In this case, the position of the trial object, x,
relative to its neighbour is


( ) 1 2 3
( 1,1), ( 1,1), ( 1,1)
N s
r r r
D

+
c
x = x
c

(5.7)

where c is composed of three random numbers generated from a uniform distribution.

.6 Particle properties

The properties of the particles can also be assigned if their probability density
functions are known. This is an important feature of the generator as the composition of
the particles certainly varies in a real material and will have a significant effect on void
nucleation and ultimately fracture. The distributions for the elastic modulus, E,
Poissons ratio, v, and the yield strength of the particles,
y
, can be varied in the current
version of the generator.


150
.7 Object constraints

An object will be rejected from the particle field if any of the following constraints
are violated:

.1 The centroid of the object lies inside the percolation volume

The centroid of an object, x, lies within the global element block when the following
constraint is true for each face:

0
i i
d < n x - (5.9)

where n
i
, are the outward facing normal vectors for the planes that define the element
faces and d
i
is a constant of the plane.



.2 The object does not touch or overlap with any other object

No object is allowed to touch or overlap with another object. This constraint is
enforced by evaluating the minimum distance between an object and every other object
in the particle field. The objects are touching if the minimum distance is zero or the
minimum distance algorithm does not converge.

.3 Minimum distance between objects

The minimum distance between two arbitrary ellipsoids is a non-linear optimization
problem that requires an iterative solution. In this work, we adopt the geometric-based
algorithm of Lin and Han (2002) as it is simple, efficient and has very good convergence
properties. Only a brief review of the algorithm is presented.

Consider two ellipsoids denoted E
1
and E
2
as shown in Figure 5.6. At the k-th
iteration, having two points on the respective ellipsoids denoted x and y, a ball is
constructed within each ellipsoid that is tangent to the current point on the ellipsoid. The
centers of the ellipsoids provide a good estimate for the initial values of x and y. The
line segment
1 2
c c between the two ball centers is then tested to see if it lies within both
ellipsoids. If this is true, the ellipsoids are overlapping and the distance is reported as
151
zero. Otherwise, a new point x
k+1
is computed as the intersection of the line segment
1 2
c c with the boundary of the ellipsoid E
1
. The same process is used to determine the
new point y on the ellipsoid E
2
. The algorithm will converge so that the distance
1 2
( , )
k k
d E E = y x will be the minimum distance between the ellipsoids.

y
k
c
1
c
2
x
k
x
k+1
y
k+1
E
1
E
2

Figure 5.6: Schematic of optimization method of Lin and Han (2002) for the minimum
distance between two general ellipsoids

.4 The object does not touch or intersect with any of the facing planes that define the
percolation volume

The minimum distance between an object and a plane must be determined to test if
an object has intersected the boundary of the element block. This problem could be
simplified to the case of testing for an intersection between a plane and ellipsoid but the
minimum distance is preferred as it provides a more general solution. If the generated
particle field were to be meshed in a finite-element model for homogenization studies, a
minimum distance between the ellipsoids and the cell walls would have to be enforced
(Pierard et al., 2007). The minimum distance between a plane and an ellipsoid can be
solved by adapting the method of Lin and Han (2002). In this case, the equation of the
second ellipsoid is replaced with the equation of the plane and the algorithm converges
rapidly.

.5 Neighbour constraints

The type of permissible neighbour can be specified for each object. This information
is used in modeling the clusters/nearest neighbours. The type of neighbour that an object
may have is important for determining the spacing between the objects for accurate
152
representation of object clusters. The algorithm allows for a probability distribution for
the spacing between each type of neighbour in the material. This is important as the
distribution for the spacing between voids and different particle types can be used in the
model if they are known.

.8 Identification of the parent element

Each object can belong to only a single parent element. An object may be intersected
by the boundary between two percolation elements, but it belongs to the element in
which its centroid resides. An object with its centroid at x = (x, y, z) within the global
volume is located within its parent element when the following three constraints are
satisfied:

1 1
2
c
x x
a

s s

1 1
2
c
y y
b

s s

1 1
2
c
z z
c

s s (5.8a-c)



where x
c
, y
c
, z
c
denote the center of the parent element that has side lengths of 2a, 2b,
2c. These constraints are evaluated for each element within the element block until Eqs.
(5.8a-c) are satisfied for the current object.

5.3 Numerical implementation

The particle field generator algorithm was written in C++ and has been directly
ported into the percolation code. A Matlab code was developed for visualization of the
particle fields. The particle generator was ported into the finite-element percolation code
so that a new particle field could be automatically generated at run-time. This feature
provides the foundation for developing a heuristic version of the model where a large
number of simulations can be performed to capture the statistical variation in the
material behaviour. The algorithm for the particle field generation process is shown in
Figure 5.7.
153
Determine number of each object to be generated from object densities,
i
Generate the object
Assign object semi-axes (a ,b ,c ) and orientation, , from respective distributions
Select neighbour type:
specified or randomly
Randomly select a solid angle, ,
for the center-to-center vector
Select solid angle, , for the center-to-
center vector from its distribution
Select object type to be generated:
Randomly or specified from a sequence
Determine centroid of the object
- location must satisify and the minimum distance D
s
Evaluate constraints
- no intersections with other objects
- no intersections with volume boundaries
If the object is a particle, assign properties, E, v and
y
, from respective distributions
Assemble percolation elements into a global element block
Is the object
in a cluster?
r (0,1) P
c
Is the cluster
a stringer?
r (0,1) P
s
Determine object centroid
- randomly position object
Constraints
satisfied?
Number of
objects reached?
Deconstruct the element block and assign objects to parent elements
and send the object data to the percolation model
Yes
Yes
Yes
No
No
No
No
Obtain minimum spacing, D
s
,
between the object and neighbour
r (0,1) P
s
Randomly select
neighbour
Yes


Figure 5.7: Algorithm for the particle field generator
154
5.4. Application of the particle field generator to an Al-Mg alloy

The performance of the particle field generator rests upon knowledge of the
probability density distributions for the governing variables. There are two commonly
used methods to measure a particle distribution: x-ray microtomography (3-D) or from
high resolution optical images (2-D). An image tessellation algorithm is then applied to
extract the particle and void information such as the centroid, semi-axes, orientation and
nearest neighbour information.

X-ray microtomography is the obvious choice for obtaining a three-dimensional
particle distribution if the particles in the material of interest are on the order of one
micron or larger. X-ray tomography studies of ductile fracture have long suffered from
issues of resolution but recent advances have improved the resolution to about 0.7 m at
the state-of-the-art synchrotron in Grenoble, France (Maire et al., 2006; Orlov, 2006).
However, access to synchrotrons capable of achieving this degree of resolution is
limited. As a result, obtaining 2-D images from metallographic samples is the standard
method of measuring a particle distribution as it is inexpensive, accessible, and has very
good resolution of 0.139 m per pixel (Chen, 2004). A three-dimensional distribution
can be obtained by successive sectioning of the sample and assembling the 2-D images
but this is an extremely time-consuming process and may not be appropriate for
materials with fine particles (Mangan et al., 1999). A typical 2-D image obtained using
the sectioning and polishing method is presented in Figure 5.8.

Fortunately, a high-resolution x-ray microtomography study of the material of
interest, AA5182 alloy, was performed by Orlov (2006). These results will be employed
in the present work to generate the 3-D particle distributions for the percolation model.
The resolution of the pixels within the voxel images was 0.7 m which can capture most
particles in the AA5182 sheet. Particles that are too small to be detected at this
resolution are not expected to have a significant influence on the fracture strain
predictions of the model because it is the large particles that will primarily nucleate
voids. An experimental study of this alloy by Hadianfard et al. (2008) found that the
average diameter of a broken particle (void nucleation) ranged from about 4.5 m to
155
7.25 m. From this study, we can conclude that the resolution used by Orlov (2006) of
0.7 m was sufficient to capture the particles that contribute to the fracture mechanism.


Figure 5.8: In-plane images of the particle distribution for AA5182 sheet for a thickness of
1.0 mm (left) and 1.6 mm (right). The rolling direction is horizontal and the transverse
direction is vertical (Chen, 2004).

.1 Material characterization

As-cast 5xxx series aluminum-magnesium alloys contain coarse intermetallic
particles that may be larger than 100 m in size. These particles then break up during
the rolling process into particle fragments of 1-10 m in size (Maire et al., 2006) as
illustrated in Figure 5.9. There are two main types of intermetallic particles in AA5182
sheet: iron-rich and Mg
2
Si particles. The chemical composition of the alloy is presented
in Table 5.1. The particles are oriented and elongated along the rolling direction and
exhibit a strong degree of clustering. The initial voids in the material are located within
particle clusters and transverse to the rolling direction, suggesting they were formed by
particle cracking during the rolling process.

Table 5.1. Chemical composition of the AA5182-O alloy (Chen, 2004).
Element Si Fe Mg Mn Cu
Wt% 0.08 0.21 4.6 0.33 0.04


156

Figure 5.9: 3D view of intermetallic phases and voids in 1 mm thick AA5182 sheet obtained
by Maire et al. (2006): (a) as cast state; (b) the hot rolled state; (c) cold rolled state; and (d)
microstructure at the end of a tensile test. Light grey spots show the iron-rich particles,
dark grey spots are Mg
2
Si-particles and black spots are pores.

The tessellated particle fields were extracted from the voxel images by Orlov (2006)
using a matrix erosion technique that approximated the voids and particles as ellipsoids
that are aligned along the rolling, transverse or through-thickness directions of the
material. The majority of the second-phase objects in AA5182 are Fe-rich (83%), with
about 10% Mg
2
Si particles located near Fe-rich particles with the remaining 7% being
157
voids. The object densities for the material are presented in Table 5.2 with the volume
fractions in Table 5.3. The proximity of the voids to the particles was measured by the
number of particle-void interfaces (PVI).

Table 5.2. Object densities (number per mm
3
)

in AA5182 sheet (Orlov, 2006)
Fe-rich Mg
2
Si Voids Fe-rich PVI Mg
2
Si-PVI
659127 77045 55889 28103 33439

Overall, the percentages of the voids that shared an interface with the Fe-rich and Mg
2
Si
particles were reported as 46% and 54% respectively. This result indicates that the
probability of a void nucleating during the cold rolling process is approximately equal
for both particle types. This is a fortuitous result as no special measures must be taken in
the particle generation process to position voids adjacent to a preferential particle type.

Table 5.3. Volume fractions of the constituents in AA5182 sheet (Orlov, 2006). The sample
volume was 70 m x 70 m x 70 m.
Fe-rich Mg
2
Si Voids Matrix
0.00483 0.000485 0.000529 0.994156

.2 Characterization of the particles and void dimensions

The probability density functions for the ellipsoid semi-axes of the Fe-rich and
Mg
2
Si particles are presented in Figures 5.10-12, respectively. These distributions were
adapted from the work of Orlov (2006). As a consequence of the hot rolling process, the
particles should be elongated and flattened in the rolling and transverse directions and
compacted in the thickness direction. These trends were observed in the distributions as
the semi-axes in the rolling directions are been shifted to the right compared to the other
directions. The semi-axes are smallest in the thickness direction, giving credence to the
assumption that they have been compacted during rolling. The distributions for the voids
semi-axes in Figure 12 are largest in the transverse direction due to their formation by
158
particle cracking, which occurs transverse to the principal loading direction (rolling
direction). Similar to the particles, the voids also show evidence of being compressed
during the rolling process since the semi-axis is smallest in the thickness direction.
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
0 1 2 3 4 5
Semi-axes of Fe -rich particles (m)
P
r
o
b
a
b
i
l
i
t
y

D
e
n
s
i
t
y

_
Rolling direction
Transverse direction
Thickness direction

Figure 5.10: Distributions of the semi-axes of the Fe-rich particles. Adapted from Orlov
(2006).
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
0 1 2 3 4 5
Semi-axes of the Mg
2
Si particles (m)
P
r
o
b
a
b
i
l
i
t
y

D
e
n
s
i
t
y


Rolling direction
Transverse direction
Thickness direction

Figure 5.11: Distributions of the semi-axes of the Mg
2
Si particles. Adapted from Orlov
(2006).
159
0
0.2
0.4
0.6
0.8
1
0 1 2 3 4 5
Semi-axes of the voids (m)
P
r
o
b
a
b
i
l
i
t
y

D
e
n
s
i
t
y


1
Rolling direction
Transverse direction
Thickness direction

Figure 5.12: Distributions of the semi-axes of the voids. Adapted from Orlov (2006).

.3 Characterization of void and particle clusters

The probability density functions for the center-to-center distances between objects
in AA5182 sheet is presented in Figure 5.13. The distribution for the Fe-rich particles
was obtained by only considering other Fe-rich particle neighbours. The Mg
2
Si
distribution includes both Fe-rich and Mg
2
Si neighbours and the void distribution was
measured for all three object types. It can be seen that the degree of clustering for the
Fe-rich and Mg
2
Si particles is normally distributed while the voids have a strong peak at
a distance of about 2-4 m due to their nucleating in particle clusters during the rolling
process. This peak can be described as a first-order cluster where the clustering then
becomes more distributed and similar to the trends for the particles.

The manner in which the distance distributions were obtained has clear implications
on generating the random particle fields and the generation sequence of the objects and
neighbour constraints. The following assumptions are made:

- Fe-rich particles will be generated first and can only have Fe-rich neighbours

160
- The Mg
2
Si particles will be generated second and can have Mg
2
Si or Fe-rich
neighbours

- The voids will be generated last and can be located next to a void or either
particle type
0.00
0.01
0.02
0.03
0.04
0.05
0 10 20 30 40 50
Center-to-center distance between objects (m)
P
r
o
b
a
b
i
l
i
t
y

D
e
n
s
i
t
y
_
Fe-rich
Mg2Si
Voids

Figure 5.13: Distributions for the inter-object spacing as measured using the center-to-
center distance

A second issue regarding the measurement of the inter-object spacings is that the
center-to-center distance was used instead of the absolute minimum distance between
the objects. As discussed previously and shown in Figure 5.4, these values can be
significantly different when modeling an object cluster. Since the absolute minimum
distance, D
s
, is used in the present work, the required minimum distance will be
estimated from the center-to-center distance, D
c2c
, as


min 2 c c o n
D D R R =

(5.10)

where R
o
and R
n
are the distances from the object/neighbour center to its edge along the
center-to-center vector as shown in Figure 5.14.

161

Figure 5.14: Estimation of the absolute minimum distance between objects from the
center-to- center distance

.4 Parameters used in the particle field generation process for AA5182

To generate particle fields that are in accordance with the experimental results of
Orlov (2006), certain considerations must be made so that the predicted distributions are
consistent with the measurement techniques used to obtain. For example, the objects
must be aligned in the material directions since orientation was not considered in the
image tessellation process. The following criteria and parameters were used in
generating the particle fields.

- Volume of the particle field is 200 m x 200 m x 200 m
- All objects are aligned with the rolling, transverse and thickness directions. No
orientation angle.
- Rejection-sampling will be used to generate the object semi-axes
- All clusters are of the stringer type with a randomly selected solid angle of -20
to 20 relative to the rolling direction
- Fe-rich particles are generated first followed by the Mg
2
Si particles and then the
voids
- The Fe-rich particles are randomly distributed within the volume
162
- The probability of an Mg
2
Si particle being in a cluster is 80%. The spacing
between an Mg
2
Si particle and its neighbouring particle obeys a normal
distribution with a mean of 0.7 m and a standard deviation of 3 m. These
values were suggested by Orlov (2006).
- All voids are assumed to be within a cluster due to the assumption that the voids
were created by the rolling process
- The spacing between a void and its neighbouring particle is selected from a
uniform distribution from 1 m to 4.5 m. These values were determined
parameterically to give agreement with the experimental spacing distribution

5.5 Results
A typical particle field with a volume of 200 m x 200 m x 200 m is shown in
Figure 5.15 where the clustering of the voids and particles is evident along with their
preferential orientation along the rolling direction. The number of voids, Fe-rich and
Mg
2
Si particles in this volume are 447, 616, 5273, respectively. The number of objects
to be created is important for the rejection-sampling algorithm to recreate the
distributions for the object dimensions/properties. The resulting distribution will
converge to the target distribution provided a sufficient number of random samples are
taken.

.1 Object dimensions

The generated distributions for the semi-axes of the Fe-rich particles, Mg
2
Si particles
and voids are presented in Figures 5.16-18. The distributions of the Fe-rich partcle
dimensions are in excellent agreement with the distributions shown in Figure 5.10 since
5273 samples were taken from the target distribution in the sampling algorithm. The
distributions for the Mg
2
Si particles in Figure 17 exhibit more scatter since only 616
particles were generated. However, the general trends are in good agreement with the
experimental distribution in Figure 5.11 and is an acceptable representation. Finally, for
the voids in Figure 18, the scatter is again larger compared to the Mg
2
Si particles since
only 447 voids are created. Nevertheless, the general trends for the void distribution can
163
be seen when compared to the target distribution in Figure 5.12, especially in the
transverse direction. Overall, the rejection-sampling technique employed in the present
work can successfully recreate the target distribution provided a sufficient number of
samples are taken. For the model alloy, decent results can be obtained with 400-600
objects and excellent agreement at 5000 samples.



Figure 5.15: Generated particle field of AA5182 with a volume of 200 m x 200 m x 200
m.

164
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
0
200
400
600
800
1000
1200
Semi-axes of the Fe-rich particles (m)
F
r
e
q
u
e
n
c
y


Rolling direction
Transverse direction
Thickness direction

Figure 5.16: Distributions of the semi-axes of the Fe-rich particles
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
0
20
40
60
80
100
120
Semi-axes of Mg
2
Si particles (m)
F
r
e
q
u
e
n
c
y


Rolling direction
Transverse direction
Thickness direction

Figure 5.17: Distributions of the semi-axes of the Mg
2
Si particles
165
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
0
10
20
30
40
50
60
70
Semi-axes of the voids (m)
F
r
e
q
u
e
n
c
y


Rolling direction
Transverse direction
Thickness direction

Figure 5.18: Distributions of the semi-axes of the voids

.2 Spatial distribution of the objects

The ability of the particle field generator to recreate the spatial distribution of the
objects is a true test of the performance of the algorithm since it is not a straightforward
process as generating the object dimensions. The spatial distribution of the object is not
explicitly controlled and is a function of the clustering parameters (probability of
clustering, cluster type, allowable neighbours in the cluster), the inter-object spacing
distribution functions and even the sequence that the different object populations are
generated. To obtain object spacing values in accordance with the experimental results
of Orlov (2006), the dimensions of the particle field must match the experimental size of
200 m x 200 m x 200 m.

The probability density distributions for the spatial distributions are shown in Figure
5.19 and are in very good agreement with the experimental distributions in Figure 5.13.
A parametric study was performed to determine the inter-object spacings for the voids
and their neighbouring particles since the voids exhibit a bi-modal distribution with one
strong peak at 2-4 m and another at about 15 m, which is the average value for all
object types. A uniform distribution with a range of 1 m to 4.5 m enabled the
166
distribution of the voids to exhibit the characteristic bi-modal distribution. The Fe-rich
particles were generated randomly within the material and requires no assumed
distribution for the particle spacing. A preliminary study for the distribution of the
spacings of the Fe-particles was performed but was insensitive to the assumed values
due to the large number of particles, 5273, in the volume so that the resulting
distribution for the spacings was random. For the Mg
2
Si particles, a probability of
clustering of 80% with the spacing distribution following a normal distribution with a
mean and standard deviation of 0.7 m and 3 m, respectively. These values were
suggested by Orlov (2006). Finally, the stringer angle was varied as well but exhibited
no significant influence on the resulting object-spacings. So it was assumed to vary from
-20 to 20 relative to the rolling direction as this is a typical range observed in sheet
metals.
0 5 10 15 20 25 30 35 40 45 50
0
0.01
0.02
0.03
0.04
0.05
0.06
Center-to-center distance between objects D
c2c
(m)
P
r
o
b
a
b
i
l
i
t
y

D
e
n
s
i
t
y


Fe-rich and Mg
2
Si Particles
Voids
Particles and voids

Figure 5.19: Distributions of the minimum center-to-center distance between objects

Overall, the spatial distributions of the particles and voids in an AA5182 alloy can be
recreated using the proposed algorithm, albiet to the accuracy of experiment data
considered. It is important to mention that the clustering criterion used by Orlov (2006)
is only a first-order approximation to the actual clustering in the material and could be
improved in future work. The use of the minimum center-to-center object spacing
167
provides only a rough estimate of the true clustering behaviour as it neglects the
dimensions of the objects and therefore, their true spacings. The absolute minimum
distance between the objects should be used in conjunction with the center-to-center
distance to quantify the object spacing using a dimensionless spacing ratio.
Additionally, the average spacings surrounding the objects should also be reported to
obtain a radial spacing distribution. Therefore, although the particle field generator in
the present work provides good agreement with the experimental distributions, the
spatial reconstructions are approximate and may fail to account for certain effects.

.3 Particle and void volume fractions

The particle and void volume fractions are commonly the parameters of interest in
damage-based constitutive models such as the Gurson (1977) model. While often
assumed to be constant, it is not uncommon to see significant variation in the
experimental volume fraction measurements. It is therefore of interest to generate
particle fields of different sizes to observe the predicted variation of the particle and
volume fractions. A total of 20 different AA5182 particle fields were generated within
cubic elements with side lengths of 70 m, 100 m, 150 m and 200 m. The resulting
confidence intervals for the voids and Mg
2
Si particles are presented in Figure 5.20a and
the values for the Fe-rich particles in Figure 5.20b. The general trend is that the
variation in the volume fractions of the constituent objects decreases with increasing
element size. As the element size increases, a larger number of objects are generated and
the object dimensions consolidate about a mean value, giving rise to a consistent
prediction of the volume fraction. However, despite the decreased variation, the
statistical significance of this effect is weak since most of the confidence intervals
overlap.

For the purpose of predicting the object volume fractions, the size of the volume is
not a significant factor. This is an interesting result since the size distributions of the
objects is determined using rejection-sampling and therefore related to the element
volume through the object density. This result is encouraging since the volume fraction
of the voids in the element of 70 m
3
(19 voids) is virtually the same as in the element
168
of 200 m
3
(447 voids). Additionally, the volume fractions for the voids, Fe-rich and
Mg
2
Si particles reported by Orlov (2006) for a volume of 70 m
3
were 0.0529%,
0.483% and 0.0485%. The experimental values for the voids and Mg
2
Si particles fall
within the predicted range for that element size while the predictions for the Fe-rich
particles are above the experimental value at about 0.525%, which is still reasonable.
The generally good agreement with the experimental volume fractions is encouraging
for the performance of the particle field generator. It is also important to note that the
experimental object densities were taken as constant values and future work should
determine a tolerance for these values to better capture the variation in the volume
fractions.
0.04
0.044
0.048
0.052
0.056
0.06
0 0.05 0.1 0.15 0.2 0.25
Element side length (mm)
V
o
l
u
m
e

f
r
a
c
t
i
o
n

(
%
)
Voids
Mg2Si particles

Figure 5.20a: 95% Confidence intervals for the predicted volume fractions of the voids and
Mg
2
Si particles in AA5182 sheet. 20 random particle fields were generated for each
element volume.
169
0.44
0.46
0.48
0.5
0.52
0.54
0.56
0.58
0.6
0 0.05 0.1 0.15 0.2 0.25
Element side length (mm)
V
o
l
u
m
e

f
r
a
c
t
i
o
n

(
%
)
Fe-rich particles

Figure 5.20b: 95% Confidence intervals for the predicted volume fractions of the Fe-rich
particles in AA5182 sheet. 20 random particle fields were generated for each element
volume.

5.6 Summary

A particle field generation algorithm was developed and implemented into the
percolation model to populate the elements with second-phase particles and initial voids.
The particle field generator was designed for a general material and uses the probability
density functions (assumed or experimental) to generate a representative particle field
within the element. The particle field generator can accept any number of different
particle types and accounts for the shape, orientation angle, cluster type, nearest
neighbour information and particle properties.

The proposed algorithm has been implemented into the percolation model as a pre-
processor to automatically create the particle distributions in the percolation elements at
the start of a simulation. By generating different representative particle fields of an
alloy, a series of finite-element simulations can be performed to capture the variation in
the fracture strains and predicted material properties of a material of interest. The
particle field generator algorithm was validated for an aluminum-magnesium alloy using
the microtomography results of Orlov (2006). The void and particle size distributions,
170
global volume fractions and inter-particle and void spacings were successfully generated
by the model and in good agreement with the experimental observations.
171
Chapter 6 Application of the Percolation Model to Predict
Fracture in a Notched Tensile Test Specimen of AA5182 Sheet


6.1 Application of the percolation model

The performance of the complete percolation model will be evaluated by applying
the model to a notched tensile test specimen of AA5182 sheet. All of the fundamentals
established in the previous chapters are used in this application. The calibrated void
evolution models in Chapter 2 are used to model the individual voids in the particle
field. The homogenization scheme of Chapter 3 is used to obtain the stress state in the
particles as a function of their composition, size and shape to predict nucleation using
the particle cracking-based model of Chapter 4. The particle field generator of Chapter 5
is used to generate five representative particle fields of AA5182 to capture the
experimentally observed variation in the fracture strains. The percolation elements
developed in Chapter 4 are placed at the notch root of the sample to capture the
initiation of a macro-crack as observed experimentally. Finally, the predicted fracture
strains, porosity and nucleation trends are compared and validated with the experiment
data and as well as the data available in the literature.

6.2 Notch tensile test experiments

Notched tensile specimens of 1.5 mm thick AA5182 sheet with a notch radius, R, of
3 mm gage length, L, of 12.5 mm, and total sample length of 80 mm were tested to
failure in an Instron 1332 testing machine with a crosshead velocity of 0.03 mm/s. The
notch ligament length is varied to achieve various stress states and is characterized using
the notch ratio defined as

2R
w
= (6.1)

where w is the sample width. The notch ratio considered in the present work is 0.25
which gives rise to a ligament length of 18 mm. A total of six specimens for each notch
ratio were tested to failure. All of the samples were fabricated with the loading direction
172
aligned with the rolling direction of the sheet. The typical specimen geometry is
presented in Figure 6.1.


Figure 6.1: Geometry of a typical notched tensile sheet specimen. The loading and rolling
directions are horizontal. The specimen is symmetric about the x, y and z axes.

Notched tensile specimens are used to alter the stress state in the ligament and
enforce localization to occur within the ligament. Two experimental fracture parameters
are used in the calibration of the nucleation models: (i) the ligament strain,
lig
f
c , and (ii)
the axial strain
a
f
c . The ligament strain is adopted because it is representative of
deformation where the specimen fractures. The axial strain at failure is used to provide a
metric that is independent of the fractured region and representative of bulk
deformation. The axial and ligament strains at fracture are defined as


a
ln
f
f
o
L
L
c =
| |
|
\ .

lig
lig
ln
lig
f
f
o
c =
| |
|
\ .
(6.2, 6.3)

where the initial ligament length is 2
o
lig w R = . The axial strain at failure is
recorded at either the appearance of a macro-crack at the notch root and not final failure
since the objective of the finite-element models is to predict the formation of a macro-
crack and not the subsequent tearing process of the ligament. Finite-element modeling
of this process would require additional modeling considerations such as the use of
173
cohesive elements. The implementation of the percolation model in LS-DYNA causes
element deletion upon failure and is not suitable for modeling material separation.

Tensile specimens with notch ratios less than 1/3 exhibit visible cracking at the
notch root prior to fracture as shown in Figure 6.2. The onset of cracking at the notch
root is readily determined from the experimental load-elongation curve as the
appearance of the crack corresponds to an abrupt load drop as shown in Figure 6.3.



Figure 6.2: Crack initiation at the notch root in a 1.5 mm thick AA5182 tensile specimen
with a notch ratio of 0.125 and radius of 1 mm.

0.0
2.0
4.0
6.0
8.0
10.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Elongation (mm)
L
o
a
d

(
k
N
)
Experiment
Finite-Element
Notch ratio = 0.25
Progressive
ligament failure
Crack forms at
the notch root

Figure 6.3: Comparison of the average experimental and predicted load-curve for an
AA5182 notched tensile specimen.
174
6.3 Material characterization

.1 Material properties

Aside from the particle distribution, the only parameters required by the percolation
model are the material properties of the matrix material and the constituent particles.
There are no calibration parameters in the model so care must be taken in determining
the material properties. The flow stress relation for 1.5 mm thick AA5182-O sheet was
determined in-situ using three standard tensile test specimens as shown in Figure 6.4.

0
50
100
150
200
250
300
350
400
0 0.1 0.2 0.3 0.4 0.5 0.6
Equivalent plastic strain
E
q
u
i
v
a
l
e
n
t

s
t
r
e
s
s

(
M
P
a
)
Experiment
Voce fit

Figure 6.4: Experimental flow stress relation for AA5182 sheet and its Voce law
representation.

The average flow stress response was expressed as a Voce (1948) hardening law:


( ) ( )
p
s s y
exp
|
o o o o o c
(
=
(

(6.4)

with
y
122.7 MPa, o =
s
398.1 MPa, o = 7.631 o = and 0.905 | = . AA5182 alloys exhibit
different hardening regimes as shown in Figure 6.5 where it initially resembles a high-
strain hardening material with a peak hardening exponent of about 0.33 before the flow
stress saturates and the matrix resembles that of a perfectly plastic material, n 0. The
175
strain hardening parameter of AA5182 is computed at each stage during deformation
using the relation

p
p
d
n
d
c o
o c
= (6.5)

Figure 6.5: Variation of the material hardening exponent in AA5182.

The modulus of elasticity was estimated from the tensile tests to be 65.33 GPa and
the Poissons ratio is assumed to be the standard 0.33 for aluminum alloys.

.2 Void nucleation by the constituent particles

The porosity and number of nucleated voids in AA5182 were measured using x-ray
microtomography by Orlov (2006) in a tensile sample at five strain levels. It was
observed that all of the nucleated voids were attributed to the Fe-rich particles and that
the Mg
2
Si particles did not appreciably nucleate voids. A similar result has been
reported by Agarwal et al. (2002) for an AA5083 alloy. From these results we can safely
assume that void nucleation by the Mg
2
Si particles is of minor importance. Therefore,
all of these particles are assumed to remain intact in the percolation model. The Fe-rich
particles are hard and brittle. It is assumed that they nucleate via particle cracking using
the nucleation model described in Chapter 4. It should be noted that even if a percentage
of the Mg
2
Si particles did nucleate voids, the volume fraction of this phase is only one-
176
tenth that of the Fe-rich particles and their contribution would be marginal by
comparison.

.3 Material properties of the constituent particles

The properties of the constituent intermetallic particles are not often known and must
be estimated from their composition. Fortunately, the composition of the Mg
2
Si particles
is well documented since this intermetallic has gained interest as a possible structural
material due to its low density of 1.99 g/cm3, high melting point, and compressive
strength of 1640 MPa (Xiong et al., 2007). The elastic modulus and Poissons ratio of
these particles are taken as 133.5 GPa and 0.161, respectively (Tani and Kido, 2008).
The fracture toughness of the Mg
2
Si particles is not required since they are not
considered for nucleation but have a fracture toughness ranging from 0.6 to 1.2 MPa-
m
1/2
.

The composition of the Fe-rich particles depends upon the casting process such as
casting method, cooling time and temperature. The general composition of the Fe-rich
particles are Al
3
(Fe, Mn), Al
6
(Fe, Mn), Al
8
Mg
5
and others such as Al
2
Fe
5
(Li and
Arnberg, 2004; Rathod and Katsuna, 2004). Li and Arnberg (2004) observed that the
majority of the Fe-rich particles in DC-cast AA5182 were A1
3
Fe and Al
m
Fe where m
was approximately 4 in a DC-cast AA5182 alloy. The Fe-rich particles are perhaps best
treated as a single class for the purposes of modeling nucleation and estimating their
composition. Experimental data for the Fe-rich particles is scarce. Yamagiwa et al.
(2003) estimated that the Fe-rich intermetallics have the elastic modulus of steel, 200
GPa, but with half the density. The Poisson ratio can be taken as about 0.30 as suggested
by Somekawa et al. (2004).

The fracture toughness of the Fe-rich particles is required to predict particle cracking
and is the controlling parameter in the percolation model. Since the percolation model
only contains this sole damage-related parameter, it can be determined by calibration
with experiment. In the interest of evaluating the true predictive ability of the model, it
should be first evaluated using only the material data. The fracture toughness for two
177
types of Fe-rich intermetallics in AA5182, Al3Fe and Al
2
Fe
5
, have been reported as 2.15
and 2.30 MPa-m
1/2
(Rathod and Katsuna, 2004). Since the fracture toughness values for
these two different particles are so similar, we can assume that these values are
representative of the other Fe-rich particles. This is a reasonable assumption in light of
the absence of material data and the validity of this assumption will be evaluated by
comparing the predictions of percolation model with experiment data. The material
parameters are presented in Table 6.1.

Table 6.1: Material parameters of AA5182 required by the percolation model.
Constituent E (GPa) v
y
(MPa)
K
1c
(MPa-
m
1/2
)
Matrix 65.33 0.33 122.7 -
Fe-rich particles 200 0.30 - 2.15-2.30
Mg
2
Si particles 113.5 0.161 - -

6.4 Finite-element model

.1 Mesh and boundary conditions

Due to symmetry, only one-eighth of the specimen geometry is represented in the
finite-element model. A velocity boundary condition is applied to the free-end of the
specimen. A sinusoidal velocity profile is used to limit inertial effects with a peak
velocity corresponding to a strain rate of 100 s
-1
. The boundary conditions and typical
mesh containing 24000 are presented in Figure 6.6. The mesh sensitivity for this notch
ratio and number of elements was negligible (Butcher and Chen, 2011).

For computational efficiency, only percolation elements are placed at the notch root
(Figure 6.6b) to capture the initiation of the macro-crack while the remaining elements
obey J
2
plasticity using the hardening rule in Eq. (6.1). At present, the percolation model
is computationally expensive and the size of the global particle field in this specific
model is limited to a size of 4000 particles split into four elements. The placement of
only several percolation elements at the notch root is acceptable for this specific notch
geometry because deformation is highly localized at the notch root. A previous study by
178
Butcher and Chen (2011) using a Gurson-based constitutive model has shown that the
only appreciable damage occurs within this region (Figure 6.7).

(a) (b)
Figure 6.6: (a) Boundary conditions in the one-eighth finite-element of a notched tensile
specimen with a notch ratio of 0.25. The z-symmetry boundary condition is not showed for
clarity. (b) Finite-element mesh showing the location of the percolation elements. The
remaining elements obey J
2
plasticity.


Figure 6.7: Porosity distribution and failure location in a notch tensile specimen with a
notch ratio of 0.25. The porosity distribution was obtained using the GT material model
with strain-controlled nucleation (Butcher and Chen, 2011).
179
.2 Identification of the fracture parameters

The finite-element simulation is terminated at the onset of element deletion
signifying the initiation of a macro-crack at the notch root (Butcher and Chen, 2011).
The resulting ligament and axial strains are then obtained from the finite-element model
and compared with the experiment values. The data in each of the percolation elements
is combined into a single data set and used for post-processing. The simulation is
repeated for five particle fields, P1-P5, generated using the distributions and procedures
in Chapter 5.

6.5 Results and discussion

.1 Ligament and axial strain at failure

The predictions of the percolation model for the axial and ligament strain with the
experimental values are shown in Figures 6.8 and 6.9, respectively. The fracture strains
are presented as 95% confidence intervals due to the stochastic nature of the percolation
model predictions and because of experiment data exhibits variation as well. It is
remarkable that the predictions of the percolation model are in excellent agreement for
the axial strain at failure and are reasonable for the ligament strain. Although only five
particle fields were considered, the variation in the axial strain measurements in the
percolation model is comparable to the experimental variation. It is important to state
that this agreement is not a result of a calibration and is a testament to the physical
foundation and predictive ability of the model.

Since the confidence intervals overlap, there is no significant difference between the
numerical and experimental predictions but the evidence is not overwhelming for the
ligament strain. Overall, the axial strain is the more reliable measure in this case because
it is directly measured by the extensometer and is located away from the fracture
surface. The ligament strain is difficult to measure with a micrometer because it is
distorted by the fracture process. The use of a digital-image-correlation system to map
the strain field within the ligament during the experiment and stop the test at the
formation of a crack at the notch root would be preferable in future studies.
180
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0.14
0.16
A
x
i
a
l

s
t
r
a
i
n

a
t

f
a
i
l
u
r
e

1
Experiment Percolation model

Figure 6.8: Comparison of the experimental and predicted 95% confidence intervals for
the axial strain at failure

-0.08
-0.07
-0.06
-0.05
-0.04
-0.03
-0.02
-0.01
0.00
T
r
u
e

l
i
g
a
m
e
n
t

s
t
r
a
i
n
Experiment Percolation model

Figure 6.9: Comparison of the experimental and predicted 95% confidence intervals for
the ligament strain at failure.




181
.2 Porosity and void nucleation

The stress state at the notch root is essentially that of pure uniaxial tension and the
stress triaxiality is approximately 1/3 throughout the deformation process. The evolution
of the plastic strain in the notch root with the sample elongation is shown in Figure 6.10.
Since the stress state is nearly identical to uniaxial tension, the predicted porosity and
nucleation trends can be compared with the x-ray tomography results of Orlov (2006)
from a standard uniaxial tension test without a notch. The predicted porosity
distributions of the five particle fields considered are shown in Figure 6.11. All of the
particle fields are in excellent agreement with the experimental porosity data at a plastic
strain up to 0.10 and show generally good agreement at the higher strain levels.

0.0
0.1
0.2
0.3
0.4
0.5
0.6
0 0.5 1 1.5 2
Elongation (mm)
E
q
u
i
v
a
l
e
n
t

p
l
a
s
t
i
c

s
t
r
a
i
n

Figure 6.10: Evolution of the equivalent plastic strain at the center of the notch root

All of the particle fields considered, P1-P5, exhibit the same behaviour where
deformation is relatively homogenous until the commencement of void nucleation at
higher strains. The start of nucleation is shortly followed by localized coalescence that
quickly becomes unstable and sweeps throughout the particle field causing failure. The
fracture porosities are in good agreement with the metallographic observations of Smerd
et al. (2005) who reported failure porosities on the order of 0.3%.

182
0.0%
0.1%
0.2%
0.3%
0.4%
0.5%
0 0.1 0.2 0.3 0.4 0.5 0.6
Equivalent plastic strain
P
o
r
o
s
i
t
y

Exp. - Surface and Center
Exp. - Surface
Exp. - Center
P1
P2
P3
P4
P5

Figure 6.11: Comparison of the predicted porosity in the center of the notch root with the
experiment data of Orlov (2006) for a standard tensile specimen.


50000
60000
70000
80000
90000
100000
110000
120000
130000
0 0.1 0.2 0.3 0.4 0.5 0.6
Equivalent plastic strain
T
o
t
a
l

n
u
m
b
e
r

o
f

v
o
i
d
s

a
n
d

c
r
a
c
k
s

p
e
r

m
m



3


1
Exp. - Surface and Center
Exp. - Surface
Exp. - Center
P1
P2
P3
P4
P5

Figure 6.11: Comparison of the total number of voids and cracks predicted by the
percolation model with the experimental results of Orlov (2006).

183
The number of voids in the particle field is representative of void nucleation and the
experimental and numerical predicted values are presented in Figure 6.11. The
predictions of all of the particle fields are in very good agreement with the results of
Orlov (2006) with little variation. The convergence of the nucleation predictions
demonstrates that only several particle fields are required to obtain the general results.
The porosity and fracture strains in the previous figures are expected to exhibit the most
variation because they are related to coalescence and that is strongly dependent which is
local microstructure. This agreement with the experimental nucleation trends is very
encouraging for the physical foundation of both the percolation and the nucleation
model. The void nucleation model relies upon estimation of the stress state within the
particle, its elastic properties, shape, size and fracture toughness as well as the intrinsic
physics related to the assumption of particle cracking. It is no small feat to have all of
these factors to align in a convincing fashion without the use of heuristics and
calibrations.

.3 Nucleation stress and strain

Traditional void nucleation models rely upon strain-based measurements to describe
nucleation using the plastic strain of the material. This is a convenient method because
the particles do not have to be explicitly modeled or the stress estimated within them
using a homogenization method. The bulk deformation is used to quantify the initiation
of voids and this can be used to recreate the expected trend in the simulation. The
problem with this methodology is that the nucleation model cannot be applied to any
other situation other than the one it was determined in. Strain measurements are path-
dependent while stress measurements are not. The plastic strain obtained in a tensile test
is not the same as the plastic strain in tube bending and hydroforming. It is of great
interest to adopt a stress-based nucleation model so that it can be transferred to different
stress states and used to predict fracture in metal forming operations.

184
0
100
200
300
400
500
600
700
800
900
1000
1100
1200
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
True axial strain
A
v
e
r
a
g
e

p
a
r
t
i
c
l
e

n
u
c
l
e
a
t
i
o
n

s
t
r
e
s
s

(
M
P
a
)
P1
P2
P3
P4
P5

Figure 6.12: Comparison of the average maximum principal stress in the particles at void
nucleation for each particle field considered.

The average particle principal stress at nucleation is shown in Figure 6.12. Large
particles crack at low stresses early in the deformation process while the smaller
particles do not nucleate until the later in the deformation process. The average
nucleation stress is approximately constant because the early nucleation of the large
particles leaves the population of roughly the same size. The predicted nucleation stress
for the Fe-rich particles in AA5182 is about 950 MPa and is a reasonable value for a
hard and brittle particle. It is expected that the Fe-rich particles would have a lower
nucleation stress than the yield stress of the Mg2Si particles of 1640 MPa who do not
appreciably nucleate voids. The predicted nucleation stress of the Fe-rich particles could
be validated using a micro-indentation method. The fracture toughness can also be
determined using this procedure to verify if the value of 2.30 MPa-m
1/2
is truly
representative of the particles.

185
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0 10 20 30 40 50 60 70 80 90 100 110 120
Average volume of a broken particle (m
3
)
E
q
u
i
v
a
l
e
n
t

p
l
a
s
t
i
c

s
t
r
a
i
n

a
t

n
u
c
l
e
a
t
i
o
n
1
P1
P2
P3
P4
P5

Figure 6.13: Predicted average plastic strain of the particles upon nucleation via cracking.

A strain-based nucleation criterion can also be developed by comparing the average
volume of a broken particle to the global plastic strain at which it nucleated as shown in
Figure 6.13. All of the particle fields are in good agreement with each other and display
the well-known behaviour that small voids will nucleate only at high strains while larger
particles display a negligible size effect and have a near constant nucleation stress or
strain. Orlov (2006) observed that no particles that had a volume smaller than 17.8 m
3

nucleated a void and this is in accordance with the predictions of the model

It is interesting to mention that the strain-based nucleation trend in Figure 6.13 is
quite similar to the phenomenological nucleation model of Butcher and Chen (2009a)
that was used in a previous 2-D percolation model in Figure 6.14. The advantage of the
present model is that these trends in Figure 6.13 were obtained from first-principles, do
not contain calibration parameters, and use the stress in the particles instead of the
matrix strain. Nevertheless, it is encouraging that the phenomenological model of
Butcher and Chen (2009a) can provide similar trends and could certainly be calibrated
using the percolation model results.

186

Figure 6.14: Variation of the nucleation strain with the particle size and stress state using
the phenomenological model of Butcher and Chen (2009a). This nucleation model was used
to predict nucleation in a previous version of the percolation model that was restricted to
2-D voids.

The particle-cracking void nucleation model has given predictions that are in
excellent agreement with the experimental data, especially considering that no
calibration parameters were employed. The agreement could no doubt be improved by
introducing calibration parameters such as a critical particle volume or stress. The
physics of the nucleation process are well represented with large particles nucleating at
low strains compared to smaller particles. It should be emphasized that this nucleation
rule does not account for particle clusters and this will play a role on the nucleation
process. A non-local averaging scheme could be employed to account for the influence
of neighbour particles when determining the stress in the particle. This process will
introduce a length-scale that will require calibration and was not considered at present to
avoid introducing parameters with unknown values. Future work will address the issue
of particle nucleation within clusters.




187
.4 Void and crack size

Figure 6.15 shows the variation of the average equivalent radius of the voids during
deformation. The average size of the voids steadily increases as the primary voids grow
evolve with the matrix. The average void size then dips as the nucleation begins and
creates penny-shaped voids of almost negligible size, increasing the number of voids
and reducing the average. The average void size quickly recovers from this dip and
grows at a faster rate since penny-shaped voids grow at a high rate in the opening
direction. The trends are similar for each of the particle fields considered and the
average void radius at fracture is approximately 1.3-1.7 m.

0.6
0.8
1.0
1.2
1.4
1.6
1.8
2.0
2.2
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
True axial strain
A
v
e
r
a
g
e

e
q
u
i
v
a
l
e
n
t

v
o
i
d

r
a
d
i
u
s

(

m
)


!
P1
P2
P3
P4
P5

Figure 6.15: Predicted average equivalent radius of the voids in the various particle fields

The average equivalent crack radius exhibits significant variation between particle
fields as shown in Figure 6.16. This is not unexpected since coalescence is primarily a
function of geometry and each particle field will contain different clusters that will
coalesce in different ways and directions. What can be gleaned from this figure is that
the general trend is the same for each particle field in that there are few coalescence
events before fracture. The crack size early in the deformation process is similar to that
of the voids, which is a result of coalescence between two voids that are in close
188
proximity and belong to the same particle cluster. Crack propagation begins and ends
abruptly as seen by the sharp vertical jumps in the radius in a small applied strain
increment. The average crack radius ranges from about 4 m to 8 m for the field
considered.

0.0
1.0
2.0
3.0
4.0
5.0
6.0
7.0
8.0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16
True axial strain
A
v
e
r
a
g
e

e
q
u
i
v
a
l
e
n
t

c
r
a
c
k

r
a
d
i
u
s

(

m
)
P1
P2
P3
P4
P5

Figure 6.16: Predicted average equivalent radius of the cracks in the various particle fields


The predicted ranges for the equivalent void and crack radii are in qualitative
agreement with the sizes of the voids obtained from an SEM micrograph of the fracture
surface at the notch root in Figure 6.17. The SEM micrograph was obtained for a similar
notch ratio of 0.125 that exhibits the same fracture behaviour. The dimpled fracture
surface corresponds to the size of the voids and the large dimples correspond to cracks
as defined in the percolation model. The presence of a large primary void near the top of
the image and the relatively uniform size of the smaller remaining voids is consistent
with a large void instigating the unstable coalescence process.

189

Figure 6.17: SEM micrograph of the fracture surface at the notch root for an AA5182
tensile specimen with a notch ratio of 0.125.

.5 Shape of the voids and cracks

The variation of the average aspect ratio of the voids and cracks is presented in
Figure 6.18. The initial voids are approximately spherical and elongate in the loading
direction, increasing the aspect ratio. The onset of void nucleation quickly reduces the
average aspect ratio since the nucleated voids are penny-shaped and have an aspect ratio
of 0.01. However, once nucleated, these penny-shaped voids exhibit remarkable growth
and shape evolution and the average aspect ratio of the voids recovers and slightly
surpasses its pre-nucleation value at fracture.

Overall, the average void aspect ratio is roughly spherical for the bulk of the
deformation process. This result has several implications for constitutive modeling. The
first being that the standard Gurson (1977) yield criterion could be used to account for
material softening of the bulk particle field. This is beneficial because the Gurson (1977)
model is widely available and simpler to implement than other more advanced models
that account for void shape. It should be emphasized that this does not mean that void
shape can be neglected in the constitutive model since the void growth predictions for a
constant spherical void would not at all be in accordance with penny-shaped voids. The
assumption of a spherical void is only valid for evaluating the yield criterion to account
190
for material softening and not the evolution of the individual or equivalent void. If only
a single void is considered as in the traditional damage-based constitutive models, the
void shape must be considered.

0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
1.8
2.0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
True axial strain
A
v
e
r
a
g
e

a
s
p
e
c
t

r
a
t
i
o

o
f

t
h
e

v
o
i
d
s

a
n
d

c
r
a
c
k
s
P1
P2
P3
P4
P5

Figure 6.18: Evolution of the global aspect ratio (voids and cracks) in the various particle
fields. The aspect ratio is determined relative to the principal loading direction.

.6 Comparison of the percolation model with a traditional damage-based model

It is of great interest to compare the predictions of the percolation model with that of
a typical damage-based constitutive model. Butcher and Chen (2011) employed the
classical Gurson-Tvergaard (GT) model for spherical voids to describe void growth and
material softening in AA5182 notched tensile specimens. Void nucleation was
represented using the stress- and strain-based continuum models of Chu and Needleman
(1980) in Eq. (1.6) and Eq. (1.7). The unit cell geometry used in the coalescence model
was determined based upon the average inter-particle spacings from 2-D images of the
material. The initial porosity and volume fraction of second-phase particles of Orlov
(2006) were employed to ensure that all of the parameters were based upon
experimental measurements. The mean nucleation stress and strain values were then
identified by calibrating the model to give fracture strains in accordance with the
191
experiment data. In short, considerable effort went into ensuring that the modeling
procedure was sound and that the parameters were physically realistic.

0.0%
0.1%
0.2%
0.3%
0.4%
0.5%
0.6%
0.7%
0.8%
0.9%
1.0%
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8
Elongation (mm)
P
o
r
o
s
i
t
y
Butcher and Chen (2011):
Butcher and Chen (2011):
P1
P3

N
= 0.375

N
= 466 MPa

Figure 6.19: Predicted porosity in the notch root using the percolation model and the
Gurson-based model of Butcher and Chen (2011) who considered both stress- and strain-
based nucleation models.

The comparison of the porosity in the notch root as predicted by Butcher and Chen
(2011) and the percolation model is presented in Figure 6.19. Only two of the particle
fields are presented for clarity. The predictions of the percolation model are in pretty
good agreement with that of the strain-controlled nucleation model. Porosity is low for
the majority of deformation prior to the onset of rapid nucleation causing fracture. The
porosity at failure is similar for both the continuum and percolation models. The ability
of the Gurson-based model to predict fracture at this realistic porosity is because the
void spacing was accounted for in defining the representative unit cell in the
coalescence model and that the assumption of a global spherical void is reasonable
(Figure 6.18). As established in Chapter 2, if the void geometry and spacing ratio are
well predicted, the coalescence model will give good predictions for failure.

192
From these results, the performance of the traditional model with proper material
characterization appears to be quite reasonable. Although the nucleation model was
calibrated to provide sufficient porosity to cause coalescence at the experimental
fracture strain, the resulting trend for the porosity and its value at coalescence were not
calibrated, so their agreement with the percolation model is encouraging. However, if
we dig deeper into the Gurson-based model, its predictions for nucleation expose the
flaws in the model as shown in Figure 6.20.

0.0%
0.1%
0.2%
0.3%
0.4%
0.5%
0.6%
0.7%
0.0 0.5 1.0 1.5 2.0
Elongation (mm)
V
o
i
d

n
u
c
l
e
a
t
i
o
n
1
Butcher and Chen (2011):
Butcher and Chen (2011):
P1
P3

N
= 0.375

N
= 466 MPa

Figure 6.20: Comparison of the predicted porosity due to nucleated voids of the
percolation model with the Gurson-based model used by Butcher and Chen (2011). The
constitutive model of Butcher and Chen (2011) considered continuum-based stress and
strain-controlled nucleation models for this alloy.

The porosity created by nucleated voids is extremely large compared to the
percolation model whose trend is too low to be clearly shown in the figure. The
contribution of the porosity due to nucleated voids in percolation model is negligible
because the voids are initially penny-shaped. The calibration of the nucleation parameter
results in the model predicting a massive amount of void nucleation in order to trigger
fracture at the desired experimental strain. As a result, the porosity due to nucleation is
0.4-0.5% which means that 80 to 100% of the second phase- particles have become
193
voids because the total particle fraction in this material is only about 0.5%. In reality, the
experimental percentage of particles that nucleate voids is a pallid 6.7% by comparison.
The calibration process of the nucleation model has given the model a decent prediction
of the porosity history and fracture strains but at the cost of the validity of the nucleation
predictions.

The purpose of this comparison was to emphasize the limitations of the traditional
damage-based constitutive models and the difficulties in calibrating them to offset their
limitations while giving physically realistic predictions. Recent extensions to these
models have improved their physical foundations, especially for void growth and
coalescence, but the nucleation treatments remain inadequate. The homogenization
techniques developed in Chapter 3 along with the nucleation model in Chapter 4 can be
employed in future studies to improve the physical foundation of these model.

6.6 Summary

The complete damage percolation model was used to predict fracture and damage
evolution in a notched tensile sheet specimen of AA5182 sheet. Representative particle
distributions were created and mapped to the percolation elements located at the notch
root where fracture initiates in the sample. The fracture strain, porosity, and nucleation
predictions of the model are in very good agreement with the experimental trends of
Orlov (2006).

No calibration or adjustable parameters were employed in the model and its good
predictions of the experiment data attest to the strong physical foundation of the model.
Fracture is a sole consequence of the stress state, material composition and the particle
distribution. The present percolation model represents a significant contribution to the
field of damage-induced ductile fracture by enabling a designer to optimize a material
microstructure by simulating its performance in a metal forming operation. Unlike
traditional damage-based models, the physical foundation of the model is sound and the
inter-relationships between the competing fracture mechanisms are well represented.

194
Chapter 7 Void Growth and Evolution in Void Clusters


7.1. Void evolution within clusters

All damage-based constitutive models, including the percolation model, employ void
growth and shape evolution models that assume the voids remain isolated and do not
interact with their neighbours. Unfortunately, this is not the case in a real material. Void
interactions within clusters directly lead to incipient coalescence and ultimately,
fracture. Consequently, there is great interest in developing a modelling treatment that
can account for the influence of interactions on the growth and evolution of voids within
a cluster.

Perhaps the most straightforward approach to modeling void clusters is to perform
finite-element simulations of unit cells that contain multiple voids to investigate the
increased growth rates and local coalescence (Thomson et al., 1998; Zhang and Chen,
2007). Evolution rules could then be developed to account for void interactions. This
method is useful to provide a benchmark for evaluating analytical models for void
clusters or for a specific and well-defined material with a clear cluster geometry that
could be reduced to a representative volume element for analysis. Unfortunately, this
methodology is impractical for a general treatment that is applicable to different
materials due to the myriad of void sizes, shapes, spacings and cluster arrangements that
occur in a real microstructure. There is a practical limit on how many numerical
simulations one can perform.

Due to this limitation, it is reasonable to assume that void evolution models will
always be developed for isolated voids. With this in mind, a method is required to apply
these models to account for void interactions. The subsequent sections in this chapter
will develop a novel treatment for the modeling of void interactions that accounts for
important factors such as the relative void spacing, void size, shape, stress state and
ligament orientation. The resulting model will be evaluated for a simple linear cluster of
oblate voids using a single percolation element subjected to uniaxial tensile loading.


195
7.2 Non-local void evolution

A non-local treatment for void evolution provides a promising approach to apply the
void evolution rates for isolated voids to void clusters. Non-local void evolution theories
were initially developed to remove the mesh dependence in finite-element simulations
using damage-based constitutive models (Pijaudier-Cabot and Bazant, 1987; Leblond et
al., 1994; Tvergaard and Needleman, 1997; Drabeck et al., 2005; Niordson, 2008). In
the non-local averaging procedure, a characteristic length-scale and a corresponding
weight function is used to measure the influence of the neighbouring voids on the
growth rate of the void of interest. In this method, the deformation of a single element
will no longer dominate the simulation and cause erroneous predictions for the porosity
and fracture strain. The general expressions for the nonlocal procedure for the void
growth rate are shown in Eq. (7.1) and (7.2) and illustrated in Figure 7.1. The growth
rate at a position, x, is compared with the growth rate of a neighbour, i, at position, y,
using a weighting function, w, that is a function of the distance, characteristic length, L,
and contains the calibration parameters, p
i
. The averaging procedure is then performed
repeatedly for all of the neighbouring voids.


( )
1
( )
N
nl local i
i
f f w
=
=

x x y
( )
calibration
( ) , ,
i
w g L p = x y x y (7.1, 7.2)


L

Figure 7.1: Schematic of a non-local averaging procedure showing the characteristic
length.

196
While non-local methods are viable methods for removing the mesh dependence,
their treatment for void interactions is somewhat ad-hoc and lacks a strong physical
foundation. Issues abound in regards to selecting a proper length-scale and weighting
function. A typical weighting function could be a normal or exponential distribution that
uses the distance between element centroids and the length-scale can be chosen as the
average void size, spacing, or arbitrarily assumed or calibrated (Tvergaard and
Needleman, 1997). Furthermore, the adoption of distance-based weight functions is
questionable as they are overly simplistic and are unable to account for the relevant
factors that influence void interactions such as the void spacing, shape, size, stress state
and cluster orientation. A distance-based metric is simply unable to describe these
effects. Consequently, the choice of the length-scale and parameters in the weighting
function must be selected to correct for these deficiencies, leading to non-transferable
parameters. As such, there is a clear need to develop a nonlocal treatment that explicitly
accounts for the void geometry and stress state by improving the physical foundation of
the weighting function.

7.3 Void interactions

To develop a proper weighting function to model void interactions one must first
establish a metric to quantify the interactions. The simplest treatment is to adopt a
distance-based metric so voids that are closer together will have more interactions.
However, no process in void evolution is that straightforward to model and the relative
spacing, stress state and geometry of the neighbouring voids all must be considered. The
issue is how one accounts for all of these factors without introducing so many
material/calibration parameters into the model as to make it impractical. As it happens,
the solution to all of these challenges can be found by studying the point when the void
interactions become critical: the onset of void coalescence.

In void coalescence, two neighbouring voids will rapidly grow laterally towards one
another as deformation localizes within the ligament prior to its complete collapse. An
197
experimental display of void interactions is presented in Figure 7.2 for an array of 1 mm
diameter holes drilled in 1 mm thick AA5182 sheet.


Figure 7.2: Fracture of a 45 array of 1 mm diameter holes in Al-Mg sheet. The loading
direction is vertical and coalescence occurs between the large transverse voids by ligament
failure.

As expected, the two voids that are transverse to the loading direction (vertical) have
grown to be larger and slightly more oblate than the vertical voids. The interactions
between these voids drives the void growth rate and results in a more oblate shaped void
as they laterally grow towards one another and eventually coalesce. The vertical voids
are smaller as they are aligned in the principal loading direction where coalescence is
rarely seen (Benzerga, 2002). From this example, the influence of the void spacing and
orientation relative to the principal loading direction on void evolution is clearly
demonstrated.

Great strides have been made in the modeling of the coalescence process using the
plastic-limit-load criterion (Thomason, 1990; Pardoen and Hutchinson, 2000) so that
coalescence can now be accurately predicted as a function of the void shape, spacing,
stress state and material hardening. If we view the coalescence process from the
perspective of void interactions, we can define the onset of void coalescence as the point
when interactions become unstable as the voids rapidly grow towards one another
regardless of the principal loading direction. In this manner, it becomes clear that the
coalescence model provides an intrinsic measure of the void interactions and can be
198
used as the metric in the non-local weight function. To develop our weighting function,
first we will turn our attention to the modeling of the void coalescence process.

7.4 Measuring void interactions

.1 Interactions in the coalescence model

The plastic limit-load coalescence model of Thomason (1990) was developed to
identify the onset of internal necking coalescence when deformation of a unit cell is no
longer homogeneous but localized within the inter-void ligament. At this stage, all
deformation occurs within the ligament, driving rapid void expansion as the ligament
necks down and collapses as the void links up with its neighbour to form a larger void or
crack. Since the ligament must support the load during deformation, Thomason (1990)
was able to develop a constraint factor, C
f
, to relate the macroscopic load applied on the
cell to the stress state that develops within the ligament as a function of the
microstructure geometry. As shown in Figure 7.3 and discussed in Chapter 2, expressing
the data from a unit cell simulation in the terms of the plastic limit-load model can
accurately characterize the coalescence process. When the voids are small and/or far
apart, the constraint factor is very large and the voids grow as isolated voids. As the
voids grow and evolve during deformation, the constraint factor decreases as the stress
state becomes steadily more severe due to material softening. When curves for the
constraint factor and the tensile stress state intersect, coalescence occurs and fracture is
imminent.
199
0
2
4
6
8
10
12
14
0.0 0.2 0.4 0.6 0.8 1.0
Equivalent macroscopic strain
T = 1 f
0
= 0.1% W
0
= 1 n = 0.10
2
1 1.24
( )
f
C n
W
_
o
_ _
| |
| |
= + |
|
|
\ .
\ .
( )
2
1
1 _ o
E
Coalescence

Figure 7.3: Evolution of the plastic constraint factor and stress state for an initially
spherical void.

As seen in the unit cell simulations discussed in Chapter 2, the void initially grows as
an isolated void before beginning to interact with its neighbour as the ligament necks
down. The onset of void interactions is best visualized by their influence on the void
aspect ratio as the aspect ratio evolution rate will decrease to zero prior to coalescence.
At this point, the void growth rate will rapidly increase as the void grows laterally
toward its neighbour (reducing the aspect ratio). The lateral void growth rate
overshadows the elongation of the void in the principal loading direction. The
interactions dominate the void evolution process with coalescence and ligament failure
occurring shortly thereafter as seen in Figure 7.4. By extrapolating the void aspect ratio
(assuming no void interactions), the extent of the interactions can be visualized.

200
0.5
1.0
1.5
2.0
2.5
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Equivalent macroscopic strain
V
o
i
d

a
s
p
e
c
t

r
a
t
i
o



W
Coalescence
T = 1 f
0
= 0.1% W
0
= 1 n = 0.10
Void interactions
Extrapolated
dW/W = 0
Isolated void

Figure 7.4: Evolution of the void aspect ratio for an initially spherical void.

.2 Nonlocal weighting function

From inspection of Figures 7.3 and 7.4, the plastic limit-load model can be used as
the metric in the weighting function for measuring void interactions instead of the inter-
element or inter-void distance. The coalescence model can be normalized so that a value
of zero corresponds to the onset of coalescence and a value of unity indicating that
coalescence is impossible. The resulting metric is defined as
L
P A and its application for
an initially spherical void is shown Figure 7.5


( )
2
lig
1
1
1
1
( , )
L
f
P
C W _ _ o
| |
E
A =
|
\ .
(7.3)

As clearly seen in Eq. (7.3), the interactions between two voids are related to the stress
state, void spacing, void shape and material hardening without any fitting or calibration
parameters. Additionally, the characteristic length parameter has been removed from the
model as the void spacing ratio, _ , accounts for the distance between voids. This is a
significant contribution as Eq. (7.3) clearly has a stronger physical foundation than using
a simple distance-based measure while removing a parameter from the non-local model.

201
0.0
0.2
0.4
0.6
0.8
1.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2
Equivalent macroscopic strain

P
L
Coalescence
T = 1 f
0
= 0.1% W
0
= 1 n = 0.10
( )
2
1
1
1
1
L
f
P
C _ o
| | E
A =
|
\ .
Isolated void
dW/W = 0:
Void interactions
are significant

Figure 7.5: Evolution of the plastic limit load metric for an initially spherical void.

With a metric for the interactions established, a weighting function can be proposed.
Since the appropriate form for the weighting function remains an open question, any
type can be assumed. In this case, an exponential weighting function seems to be a
reasonable choice as void growth and shape evolution are exponential in nature and void
interactions will likely rapidly increase as the voids become closer together. As such, the
following form for the weighting function, e , is proposed


( )
( )
2
1
exp
L
P
|
e | = A (7.4)

where
1
and
2
are calibration parameters. The value for
1
can be readily determined
since 0 e ~ when 1
L
P A ~ so that the voids will grow as isolated voids when they do not
interact. Therefore, setting 0.0001 e = when 1
L
P A = in Eq. (4) yields
1
9.21 | = . The
value of
2
is left as a calibration parameter and could be determined from finite-element
simulations of a unit cells or clustered voids. For example, in Figure 7.5, at a triaxiality
of unity for an initially spherical void, interactions appear to be negligible until
0.1
L
P A ~ . Values of
2
could be selected so that the growth rate and aspect ratio match
the finite-element behaviour.
202
.3 Non-local void growth and shape evolution

Since voids will generally grow faster in the presence of a neighbour, we assume that
the voids cannot grow any slower than the isolated void case. The non-local growth rate
is expressed in traditional formulations as


( )
nl
1
N
i L i
i
f f P f e
=
= + A

(7.5)

where N is the number of neighbours and
i
f is the isolated void growth rate of the i-th
neighbour.

While the void aspect ratio is normally not considered in non-local treatments, a
proper model must account for the decrease in the aspect ratio evolution rate associated
with interactions prior to coalescence. Since the aspect ratio, W, is defined in the
percolation model relative to the principal loading direction, the shape evolution rate
may increase or decrease depending on the alignment of the interacting voids. The
alignment of the voids is defined by the center-to-center vector (ligament direction)
between the two voids. If the inter-void ligament is transverse to the principal loading
direction as in typical unit cell simulations, the aspect ratio will decrease with
interactions and increase otherwise. As such, we introduce a sign function, g, based
upon the angle, , between the principal stress direction,
1

E , and the direction of the


maximum tensile stress transverse to the ligament,
lig
1

E . It is arbitrarily assumed that the


sign function will change from positive to negative when the angle is greater than 45.
The resulting nonlocal aspect ratio evolution rate is calculated as


( )
nl
1
( )
N
i i L i
i
W W g P W e
=
= + A

(7.6-7.8)

lig
1 1
lig
1 1

cos

E E
=
E E

1 / 4
( )
1 / 4
g
t

t
>
=

<





203
.4 Nonlocal void size effect

In addition to accounting for the void shape, spacing and stress state, the relative size
of the void compared to its neighbours must also be considered in the development of a
realistic treatment for void interactions. Let us consider two voids of the same shape
with one void much larger than the other. The void evolution rate of the small void
cannot be assumed to be equal to that of the large void in measuring the interactions as
its contribution to the large void will be exceedingly small. The reason for this is that the
large void will disturb the stress field surrounding it, increasing the local stress
triaxiality and spurring growth of the nearby smaller void. Conversely, the disruption of
the stress field surrounding the small void will be small as well, contributing little in
comparison to the disruption caused by its larger neighbour. This can be thought of as a
one-way interaction as it is the larger primary void that is driving the accelerated
growth of the smaller neighbour. It is this exact mechanism that drives the nucleation
and growth of a secondary population of voids that results in a dimpled fracture surface
as smaller voids or dimples can be seen to surround a larger, primary void, as shown in
Figure 7.6.


Figure 7.6: Small voids rapidly grow next to primary voids where the local stress
triaxiality is high. These voids influence their smaller neighbours in turn, weakening the
ligament and resulting in a dimpled fracture surface from the multiple size-scales of voids.
The loading direction is vertical.

The desired void size effect can be introduced into the weighting function using a
simple size function, h, of the relative porosity (volume fraction) of the void of interest
and its neighbour as
204

0 (no size effect)
( ) 1/ 3 (equivalent radius)
1 (unit cell)
s
Neighbour
s
f
h size
f
q
q

| |
= =

|
\ .

(7.9)


where
s
is a parameter for the void size and could be determined from finite-element
simulations of two different sized voids embedded within a unit cell or infinite material.
If the voids are the same size as in a periodic distribution of unit cells, then 0
s
q = . As
the void size is a measure of the relative void volumes, a value of 1/ 3
s
q =

is a
reasonable first choice as it represents the ratio of the equivalent radii of the two voids.
This modification to the model is advantageous as large voids will contribute to the
evolution of their smaller neighbours while the smaller neighbours will have a smaller
contribution to their larger neighbour. While phenomenological in nature, this
modification is expected to give a reasonable description of the interactions between
voids of different sizes. The final expressions for the non-local void growth and shape
evolution rate can now be expressed as


( )
nl
1
s
N
i
i L i
i
f
f f P f
f
q
e
=
| |
= + A
|
\ .

(7.10)

( )
nl
1
( )
s
N
i
i Li i
i
f
W W g P W
f
q
e
=
| |
= + A
|
\ .

(7.11)


7.5 Application of the non-local model to a linear void cluster

To demonstrate the performance of the proposed non-local model for void
interactions, a single 8-node solid brick percolation element will be considered that
contains a linear cluster of initially oblate voids subjected to uniaxial tension. First, a
linear cluster of two voids will be analyzed followed by a cluster containing three voids
where interactions will become severe for the central void in the cluster. The typical
percolation element geometry is shown in Figure 7.7 and the voids have been arranged
so that the element is essentially composed of either two or three unit cells with each
containing an isolated void.
205

R
2
R
1
R
3
R
2
R
1
R
3
d/2 d d/2
2L
1
2L
3
2L
2

1

Figure 7.7: Typical geometry of a solid percolation element containing a linear cluster of
voids.

The total initial porosity of the cell is maintained at 0.1% with an initial aspect ratio
of the voids of 1/6. The cell lengths are varied such that the void spacing ratio between
the voids within the cluster is 0.44814 as this value is commonly used in unit cell
studies. The hardening law of the matrix is the same as used in Chapter 2 for the unit
cells. The parameter in the non-local weighting function,
2
, is somewhat arbitrarily
assumed to be
2
= 1.62. This value was selected as it predicts coalescence of the two-
void linear cluster at the same fracture strain of 1.20 that was observed in unit cell
simulations for an initial spacing ratio of = 0.44814 for voids with the same initial
aspect ratio. The parameter for the void size effect is taken as 1/ 3
s
q =

but is a
secondary parameter since the voids are all initially the same size. Additionally, the void
growth and shape evolution rates obtained from the unit cell simulations of Chapter 2
for voids with an initial porosity of 0.1% are used to model the evolution of the voids in
the isolated case. These evolution rates are used to provide a benchmark to measure the
effect of void interactions.



206
7.6 Results and discussion

.1 Two-void linear cluster with initial spacing ratio

The two-void cluster closely represents two unit cells joined together to create the
single percolation element. As shown in Figure 7.8, when non-local interactions are
considered, the porosity trend deviates immediately at the onset of deformation due to
the initial close spacing of the voids. Overall, this effect is minimal on the global
porosity at first until an equivalent strain of about 0.40 where the isolated void growth
trend tapers off and growth slows down as the void becomes prolate. Due to the
presence of the interactions, the shape evolution is reduced as shown in Figure 7.9.
Since oblate voids grow faster than prolate voids, the slower evolution of the aspect
ratio when interactions are considered translates to higher void growth rates and the
strong divergence between the isolated and non-local cases.

0.000
0.002
0.004
0.006
0.008
0.010
0.012
0.0 0.4 0.8 1.2 1.6 2.0
Equivalent macroscopic strain
P
o
r
o
s
i
t
y



f
Global - Isolated void
Global - Interactions
Coalescence
W
0
= 1/6 T = 1/3 n = 0.10
Global void
Voids with interactions
Isolated void

Figure 7.8: Comparison of the predicted porosity for a two void linear cluster with non-
local interactions with the porosity when the voids remain isolated (no interactions).
.
Perhaps what is most interesting in Figure 7.9 is that when interactions are
considered, the slowing down of the void shape evolution rate prior to coalescence is
well represented and in qualitative agreement with the behaviour observed in unit cells
(Figure 7.4). This is a significant contribution as the flattening of the void is naturally
207
captured and is a consequence of microstructure evolution as the voids near the onset of
coalescence. If the standard non-local approach was adopted, only the void growth rate
would be treated to the non-local averaging process and the influence of interactions on
the void shape would be neglected. In this case, the void aspect ratio would have
continued to increase prior to coalescence as in the isolated case and no flattening would
occur prior to coalescence. Furthermore, the porosity in this treatment would be lower
since interactions would not slow the evolution of the aspect ratio. The void would
become prolate at a lower strain, resulting in a lower void growth rate.

0
2
4
6
8
0.0 0.4 0.8 1.2 1.6 2.0
Equivalent macroscopic strain
V
o
i
d

a
s
p
e
c
t

r
a
t
i
o



W
Coalescence
W
0
= 1/6 f
0
= 0.1% n = 0.10
Isolated void
Voids with interactions
Global aspect ratio

Figure 7.9: Comparison of the predicted aspect ratio for a two void linear cluster with non-
local interactions with the aspect ratio when the voids remain isolated (no interactions).

.2 Three-void linear cluster

To better capture the influence of non-local interactions, a central void is placed
between the existing two voids to create a three-void cluster while retaining the same
spacing ratio as before. The central void will strongly interact with its neighbour on
either side and is expected to grow faster and remain more oblate than its neighbours.
Additionally, a small size differential will develop as the central void grows faster due
its more interactions and influence the evolution of its neighbours in turn. The
208
neighbouring voids only experience significant interactions with the center void since
the spacing between the flanking voids is large enough to preclude any meaningful
interaction. This is one of the benefits of the model because no characteristic length is
required. Each void can be tested against all other voids since the plastic limit-load
model contains an intrinsic measure of the inter-void distance in the void spacing ratio.
If the voids are far apart, the spacing ratio is low and the plastic limit-load will be high,
driving the weighting function to zero and removing the influence of the neighbouring
void.


0.000
0.002
0.004
0.006
0.008
0.010
0.012
0.0 0.2 0.4 0.6 0.8 1.0
Equivalent macroscopic strain
P
o
r
o
s
i
t
y



f
Global - Isolated void Global
Isolated void Center void
Neighbour voids Coalescence
W
0
= 1/6 T = 1/3 n = 0.10
Global void
Individual voids

Figure 7.10: Comparison of the predicted porosity for a three void linear cluster with non-
local interactions with the porosity when the voids remain isolated (no interactions).

As shown in Figure 7.10, the evolution of the individual voids and global porosity is
much higher for the thee-void cluster with interactions compared to the isolated void
case. Similar to the two-void case with interactions, the voids evolve almost the same as
initially isolated voids until a strain of about 0.30 before interactions have a meaningful
influence on the total porosity in the material. The addition of the central void also has a
marked effect on the coalescence strain, reducing it from 1.20 to about 0.65. This type
of non-linear drop in the ductility seems reasonable as a three-void cluster would create
209
a much more severe local stress state within the cluster region than in the two-void
cluster, driving rapid void evolution and ligament collapse.

The influence of the non-local interactions on the void aspect ratio is quite significant
as shown in Figure 7.11. The shape of the central void evolves significantly slower
compared to its neighbours as it has a void on either side to coalesce with. The aspect
ratio evolution rate of the neighbouring voids slows down prior to coalescence but they
continue to elongate in the loading direction to become prolate ellipsoids. It is the
central void with its oblate void growth that ultimately triggers coalescence by reducing
the void spacing.

0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
1.8
2.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2
Equivalent macroscopic strain
V
o
i
d

a
s
p
e
c
t

r
a
t
i
o



W
Coalescence
W
0
= 1/6 f
0
= 0.1% n = 0.10
Isolated void
Center void (most interactions)
Neighbour voids
Global aspect ratio

Figure 7.11: Comparison of the predicted aspect ratio for a two void linear cluster with
non-local interactions with the aspect ratio when the voids remain isolated (no
interactions).

To clearly observe the departure of the interacting voids from the isolated case, the
porosity and aspect ratio are compared relative to the isolated case in Figures 7.12 and
7.13, respectively. The voids evolve as isolated voids in the early stages of deformation
when they are still small, relatively far apart, and the stress state in the material is still
developing. Only as the voids begin to encroach upon one another do interactions begin
to occur and the trends depart from the isolated case. Naturally, as coalescence becomes
210
inevitable, the interactions between the voids are strong enough to cause the voids to
grow laterally even though the loading direction is vertical.

1.0
1.1
1.2
1.3
1.4
1.5
1.6
1.7
1.8
0.0 0.4 0.8 1.2
Equivalent macroscopic strain
P
o
r
o
s
i
t
y

r
e
l
a
t
i
v
e

t
o

i
s
o
l
a
t
e
d

c
a
s
e
Coalescence
W
0
= 1/6 T = 1/3 n = 0.10
Center void (most interactions)
Neighbour voids
Global

Figure 7.12: Predicted porosity for a three-void linear cluster relative to the porosity trend
for the isolated case (no interactions)

0.0
0.2
0.4
0.6
0.8
1.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2
Equivalent macroscopic strain
A
s
p
e
c
t

r
a
t
i
o

r
e
l
a
t
i
v
e

t
o

i
s
o
l
a
t
e
d

c
a
s
e
Coalescence
W
0
= 1/6 f
0
= 0.1% T = 1/3 n = 0.10
Center void (most interactions)
Neighbour voids
Global

Figure 7.13: Predicted aspect ratio for a three-void linear cluster relative to the aspect
ratio trend obtained for the isolated case (no interactions)

211
What is interesting about this non-local approach is that the physical behaviour of the
voids is well described, albeit in an approximate manner. For example, in a real void
cluster, the local stress triaxiality would increase markedly as the material within the
cluster is highly constrained. The higher stress triaxiality fuels void expansion, softening
the material and further increasing the severity of the stress state. The increased stress
triaxiality within the cluster causes voids to grow more oblately (except for penny-
shaped voids), resulting in higher growth rates as oblate voids grow faster than prolate
voids. While the non-local model is unable to determine this local triaxiality within the
clusters to directly model these effects, it does capture these effects by adjusting the
isolated void evolution rates to mimic this behaviour. The physics of the coalescence
model govern the strength of the interactions that give rise to this non-linear evolution of
damage.

.3 Three-void linear cluster aligned with the loading direction

The orientation of the inter-void ligament is critical for void coalescence. Voids that
are aligned with the direction of the principal loading direction rarely coalesce
(Benzerga, 2002). This is also reflected within the coalescence model. The three-void
cluster was rotated 90 to be perfectly aligned in the vertical direction and collinear with
the loading direction. Since the stress state considered is uniaxial tension, the stress
component transverse to the ligaments is zero and no interactions will occur since
1
L
P A = . The voids within the cluster simply evolve as isolated voids. The model
automatically accounts for directional effects, unlike traditional non-local approaches
that use the distance between elements or voids as a metric.

In the traditional approach, the predicted trends for the cluster would be identical
regardless of the loading direction which is not realistic. In the present non-local model,
if the vertically aligned cluster was exposed to a more severe stress state such as biaxial
tension, the voids would begin to interact as a tensile load would be present transverse to
the ligament. The stress state can be envisaged as activating different coalescence paths,
leading to new opportunities for voids to interact with their neighbours. This is an
important attribute of the proposed model as variations within the stress state and void
212
orientations is naturally accounted for, which is encouraging for identifying unique and
transferable parameters for the weight function.

7.7 Future work and current limitations of the model

The present chapter is prelminary in nature and its primary objective has been to
propose an improved framework for non-local modeling of void evolution. A great deal
of work remains to fully develop the model and subsequent extensions of the work are
both welcome and encouraged. The principal issues and limitations in this work are:

- The appropriate weight function for non-local analysis is unknown. This is an
outstanding issue in this area and not specific to the present work.

- The same weighting function and parameters were used to average the void
growth rate and aspect ratio evolution rate. It is possible that this may not be the
case and requires further analysis. Finite-element simulations of two identically
shaped voids within a cluster could be performed to evaluate the validity of this
assumption.

- The void size modification is phenomenological in nature and should be
validated. The present form is straightforward and is expected to be easily
calibrated using finite-element simulations of a cluster containing different sized
voids.

- The coalescence model does not explicitly account for void shearing. The
development of an accurate coalescence model for void shearing remains an
active area of research.

7.8 Summary

A novel non-local treatment to account for the interactions within void clusters has
been proposed by using the coalescence model to define the severity of interactions
between neighbouring voids. The contributions of this work are considerable as the
213
physical foundation of the weighting function has been improved while simplifying the
non-local averaging procedure and removing material parameters. The principal
contributions of this work are:

- The stress state, void spacing, shape, cluster alignment and material hardening
are combined into a single metric for quantifying the severity of the interactions
using the plastic limit-load coalescence model

- The non-local averaging procedure is extended to consider the void shape as well
as the growth rate to capture a decrease in the shape evolution rate prior to
coalescence that also translates into higher void growth rates in this regime

- The influence of the relative sizes between a void and its neighbours is
considered when evaluating their effect on the growth rate

- The characteristic length, L, is no longer a material parameter and not required in
the model

- Numerical simulations of void clusters can be performed to create benchmarks
for calibration of the model and elucidate the appropriate weight function and its
parameters.

- The model can be applied to percolation models and applied to void clusters or
used in a typical finite-element simulation using a damage-based constitutive
model with a single void in each element.







214
8. Conclusions

A fully-coupled damage percolation finite-element model has been developed to
predict ductile fracture using only the material properties and knowledge of its particle
distribution. The initiation of void damage at the individual particle-scale is captured
along with the stress- and strain-gradients that develop in the microstructure through the
use of percolation elements. Representative particle distributions can be generated at
run-time of a finite-element simulation to populate the percolation elements and obtain a
range for the predicted formability of a material in a forming operation. For the first
time, a sophisticated damage percolation model has been developed that is predictive
and does not require calibration or adjustment parameters. The fracture process is a
natural consequence of the particle distribution, stress state and the material properties.
The role of a heterogeneous particle field can be successfully described using the
present model without homogenizing away the particle and void clusters that ultimately
govern the fracture mechanism.

The complete percolation model has been used to predict the formation of a macro-
crack in a notched tensile test specimen of AA5182 sheet. The experimental variation of
the fracture strains was well represented in the percolation model by considering five
representative particle fields. The predicted porosity and nucleation trends were
validated with the x-ray microtomography results of Orlov (2006) from a standard
tensile test specimen. The agreement in the predicted and experimental porosity and
nucleation trends was very encouraging and attests to the strong foundation of the
percolation sub-models.

Using the percolation model, the assumptions used in traditional damage-based
constitutive models can be challenged and validated to ensure that when these models
are used, the parameters will be representative of the material. In AA5182, the
percolation model has predicted that the global void aspect ratio of the particle field is
roughly spherical, indicating that the simple Gurson (1977) yield criterion could be used
to describe material softening in this material.
215
The principal contribution of this work has been the development of an extensive
framework to describe all stages of ductile fracture; void nucleation, growth, shape
evolution, rotation and coalescence. Each stage has received a rigorous and thoughtful
treatment to obtain physically realistic models that do not rely upon calibration
parameters. In removing the assumption of a particle distribution and demanding
realistic modeling treatments, many new modeling treatments had to be developed that
could also be used to improve traditional damage-based models. Notably, the
development of a novel strategy to couple a homogenization theory from particle-
reinforced plasticity into an existing damage framework to estimate the stress state
within the particles. This development has led to a physically accurate nucleation rule
that is related to the stress state, particle size, shape, elastic properties and fracture
toughness all quantities that can be experimentally determined.

Additionally, an innovative treatment for modeling void interactions within clusters
has been proposed using non-local theory. Unlike past non-local models that rely upon a
characteristic length-scale, the proposed model measures the strength of void
interactions using the coalescence criterion for internal necking. In this manner, the
stress state, void geometry, spacing and ligament orientation are combined into a single
metric to quantify interactions, providing the non-local model with a sound physical
foundation.

The future of the multi-scale percolation model is bright and this thesis has only
presented the basic capabilities of the model. This thesis completes what is essentially
the first phase of this research development of the percolation framework. An
emphasis has been placed upon building a solid foundation for the model to establish
confidence in its abilities to describe the various stages of void evolution. The next
phase of this research will apply the model to industrial sheet metal forming operations
and related experiments for material characterization. Unlike traditional damage-based
models, the percolation framework naturally unifies the complex inter-relationships
between void nucleation, growth, shape evolution, rotation and material softening in
regards to the spatial distributions of the voids and inclusions.
216
9. Recommendations for Future Work

While significant progress has been made in developing a robust percolation model
with few assumptions or calibration parameters, there remains ample opportunity to
extend the model to account for its current limitations.

9.1 Void nucleation

.1 Debonding

Void nucleation by particle debonding is currently not considered in the percolation
model. Modeling of the debonding process is not a trivial task and introduces many
uncertainties such as estimating the interface strength, size and shape effects and
representing the separation process.

.2 Incremental homogenization theory to estimate the particle stress

The homogenization theory used to estimate the particle stress in Chapter 3 uses a
secant-based approach and is therefore limited to proportional loading conditions. This
is suitable for many sheet metal forming operations but cannot be used when the strain-
path changes such as tube-bending and hydroforming. Implementing an incremental
homogenization scheme a manner similar to the method in Chapter 3 would remove this
limitation and be valid for general loading conditions.

9.2 Void growth and shape evolution

A substantial validation and calibration of the void growth and shape evolution
models was performed in Chapter 2 but was limited to axisymmetric stress states. This
is valid for many different loading conditions but cannot describe void evolution when
the shear stress is significant. As such, the present percolation model is not valid for
pure torsion since the triaxiality is zero and no growth will be predicted. Additionally,
the use of axisymmetric unit cells introduces complexity when attempting to generalize
the void geometry to 3-D loading as discussed in Chapter 4. Ideally, the void evolution
217
models will be calibrated using 3-D unit cells in stress states ranging from pure shear to
axisymmetric loading at high triaxialities.

9.3 Void coalescence

The present void coalescence model could be improved using the finite-element
simulation data for general 3-D voids subjected to a range of shear and tensile stress
states. The assumption of necking coalescence is valid at moderate to high stress
triaxiality but breaks down when the shear stress is significant such as in a stamping
operation. Many practical metal forming operations involve significant shear
components and the lack of a shear-based coalescence model is arguably the largest
limitation in the present work.

9.4 Material softening

From a practical perspective, the Gurson-Tvergaard yield criterion used in the
present work is sufficient to describe material softening because the porosity is
essentially negligible, even at failure, with a value of about 0.5% for AA5182. Frankly,
a damage-based yield criterion is not essential if the model is being used in a typical
metal forming operation. A welcome extension would be to replace the current isotropic
Gurson-Tvergaard constitutive model with anisotropic version since sheet metals can
exhibit significant anisotropy due to the rolling process. Recent works by Keralavarma
and Benzerga (2010) have shown that anisotropy can significantly influence all aspects
of void evolution.

9.5. Particle fields

The particle field generator should be extended to account for more descriptive
statistics than were used to recreate the particle fields of Orlov (2006). The radial
distribution function of the void and particle spacing would be a significant
improvement as using only the center-to-center distance does not account for the object
size and orientation when modeling clusters and identifying neighbours.

218
It is important to mention that the percolation model also accepts measured 2-D or 3-
D particle distributions instead of using the particle field generator. Ideally, these direct
particle fields would be used in the model if one had access to a synchrotron with
sufficient resolution to quantify the second-phase particles and voids.

9.6 Void clusters

The proposed non-local method for modeling void interactions within clusters is
promising but requires additional studies before it is used in the percolation model in a
practical application. Finite-element simulations of unit cells containing void clusters
can be used to identify the weight function and calibrate its parameters. Experimental
tests using sufficiently small holes to represent voids could also be employed to validate
the model.

9.7 Particle clusters

Void nucleation within particle clusters was not modeled in the present work
although the model has the capability. The length-scale parameter used to identify the
particle neighbours can be used to average the properties and geometry of the
surrounding neighbours. These averaged values can then be used in the particle-based
homogenization scheme to estimate the particle stress. In this manner, nearby particles
would form a cluster and the stress in these particles would be different than if the
particles were isolated. If the length-scale is large enough to encompass all of the
particles in the material, then the homogenization scheme regains its original
formulation as the average particle composition and shape are used to estimate the
stress. This method was not employed in the present work because the choice of length-
scale parameter would impact the fracture predictions and the percolation model would
then require calibration. Nucleation within particle clusters will be considered in a
forthcoming work.
219
References


Agrawal, H., Gokhale, A.M., Graham, S., Horstemeyer, M.F., Bamman, D.J. (2002a).
Rotations of brittle particles during plastic deformation of ductile alloys. Materials
Science and Engineering A, 328, 310-316.

Agarwal,H., Gokhale, A.M., Graham, S., Horsemeyer, M.F. (2002b). Anisotropy of
Intermetallic Particle Cracking Damage Evolution in an Al-Mg-Si Base Wrought
Aluminum Alloy under Uniaxial Compression. Metallurgical and Materials
Transactions A, 33, 3443-3448.

Aravas, N., Ponte Castaneda, P. (2004). Numerical methods for porous metals with
deformation-induced anisotropy. Computational Methods Applied Mechanical
Engineering 193, 3767-3805.

Argon, A.S., Im, J., Needleman, A. (1975a). Distribution of plastic strain and negative
pressure in necked steel and copper bars. Metallurgical Transactions 6A, 815-824.

Argon, A.S., Im, J., Safoglu, R. (1975b). Cavity formation from inclusions in ductile
fracture. Metallurgical Transactions 6A, 825-837.

Argon, A.S., Im., J., Safoglu, R. (1975c). Separation of inclusions in spheroidized 1045
steel, Cu-0.6% Cr alloy and maraging steel in plastic straining. Metallurgical
Transactions 6A, 839-851.

Bandstra J. P. , Goto D. M. , Koss D. A. (1998), Ductile failure as a result of a void-
sheet instability: experiment and computational modeling, Material Science and
Engineering A. 249, 46-54.

Bandstra, J., Koss, D. (2008). On the influence of void clusters on void growth and
coalescence during ductile fracture. Acta Materialia 56, 4429-4439.
Barsoum, I. and Faleskog, J. (2007a). Rupture mechanisms in combined tension and
shear experiments. International Journal of Solids and Structures 44, 1768-1786.

Barsoum, I. and Faleskog, J. (2007b). Rupture mechanism in combined tension and
shear Micromechanics. International Journal of Fracture 44, 5481-5498.

Becker, R. (1987). The effect of porosity distribution on ductile fracture. Journal of
Mechanics and Physics of Solids 35, 577-599.

Benzerga, A.A. (2002). Micromechanics of coalescence in ductile fracture. Journal of
Mechanics and Physics of Solids 50, 1331-1362.

Benzerga, A.A., Besson, J., Pineau, A. (2004). Anisotropic ductile fracture Part II:
Theory. Acta Materialia 52, 4639-4650.
220
Beremin, F.M. (1981). Cavity formation from inclusions in ductile fracture of A50S
steel. Metallurgical Transactions, 12A, 723-731.

Berveiller, M., Zaoui, A. (1979). An extension of the self-consistent scheme to
plastically-flowing polycrystals. Journal of the Mechanics and Physics of Solids 26,
325-344.

Brocks, W., Sun, D.Z., Honig, A. (1996). Verification of micromechanical models for
ductile fracture by cell model calculations. Computational Materials Science 7, 235-
241.

Brown L. M. , Embury J. D. (1973). The initiation and growth of voids at second phase
particles, In: Proc. 3rd Int. Conf. Strength. Met. Alloys, Institute of Metals, London,
164-169.

Brown, L.M., Stobbs, W.M. (1976). Work-hardening of copper-silica-5. Equilibrium
plastic relaxation by secondary dislocations. Philosophical Magazine 34, 351-372.

Budiansky, B., Hutchinson, J.W., Slutsky, S. (1982), Void growth and collapse in
viscous solids. Mechanics of solids The Rodney Hill 60th anniversary volume,
Pergamon press, pp.13-45.

Butcher, C., Chen, Z.T., Worswick, M.J. (2006). A lower bound damage-based finite
element simulation of stretch flange forming of Al-Mg alloys. International Journal of
Fracture 142, 289-298.

Butcher, C., Chen, Z.T., Bardelcik, A., Worswick, M. (2009). Damage-based finite-
element modeling of tube hydroforming. International Journal of Fracture, 155, 55-65.

Butcher, C., Chen, Z.T. (2009). A void coalescence model for combined tension and
shear. Modeling and Simulation in Materials Science and Engineering 17, 1-15.

Butcher, C., Chen, Z.T. (2009). Damage percolation modeling of void nucleation within
heterogeneous particle distributions. Modeling and Simulation in Materials Science and
Engineering 17, 075003.

Butcher, C., Chen, Z.T. (2009). A continuum void nucleation model for an Al-Mg alloy
sheet based on measured particle distribution. Acta Mechanica Solida Sinica, 22 (2009)
391-398.

Butcher, C., Chen, Z.T. (2011). Characterizing void nucleation in a damage-based
constitutive model using notched tensile sheet specimens. Theoretical and Applied
Fracture Mechanics, 55, 140-147.

221
Chaboche, J.L., Kanoute, P., Roos, A., 2005. On the capabilities of mean-field
approaches for the description of plasticity in metal-matrix composites. International
Journal of Plasticity 21, 14091434

Chein, W., Pan, J., Tang, S. (2001). Modified anisotropic Gurson yield criterion for
porous ductile sheet metals. Journal of Engineering Materials and Technology 132,
409-416.

Chen, Z.T. (2004). The role of heterogeneous particle distribution in the prediction of
ductile fracture. Ph.D. Thesis, University of Waterloo, Canada.

Chu, C., Needleman, A. (1980). Void nucleation effects in biaxially stretched sheets.
Journal of Engineering Materials and Technology 102, 249 -256.

Cox, T.B., Low, J.R.Jr. (1974). An investigation of the plastic fracture of AISI 4340 and
18 Nickel-200 grade maraging steels, Metallurgical Transactions 5 1457-1470.

Cramer, M., Sevostianov, I. (2009). Effect of pore distribution on elastic stiffness and
fracture toughness of porous materials. International Journal of Fracture 160, 189-196.

Drabek, T., Bohm, H.J. (2005). Damage models for studying ductile matrix failure in
composites. Computational Materials Science 32, 329-336.

Doghri, I., Ouaar, A. (2003). Homogenization of two-phase elasto-plastic composite
materials and structures: study of cyclic plasticity and numerical algorithms.
International Journal of Solids and Structures 40, 1681-1712.

Eshelby, J.D. (1957). The determination of the elastic field of an ellipsoidal inclusion,
and related problems. Proceedings of the Royal Society of London A 241, 376-396.

Fabregue, D., Pardoen, T. (2008). A constitutive model for elastoplastic solids
containing primary and secondary voids. Journal of the Mechanics and Physics of Solids
56, 719-741.

Faleskog, J., Shih, C. (1997). Micromechanics of coalescence I. Synergistic effects of
elasticity, plastic yielding and multi-size-scale voids. Journal of the Mechanics and
Physics of Solids 45, 21-50.

Fisher, J.R., Gurland, J. (1981a), Void nucleation in spheroidized carbon steels, Part 1:
Experimental, Metal Science May, 185-192.

Fisher, J.R., Gurland, J. (1981b), Void nucleation in spheroidized carbon steels, Part 2:
Model, Metal Science May, 193-202.

222
Fleck, N.A., Hutchinson, J.W. (1997). Strain gradient plasticity. In: Hutchinson, JU.W.,
Wu, T.Y. (Eds.), Advances in Applied Mechanics, vol. 33. Academic Press, New York,
pp. 295-361.

Geltmacher, A. B. ; Koss, D. A. ; Matic, P. ; Stout, M. G. (1996), A modeling study
of the effect of stress state on void linking during ductile fracture, Acta Materialia. ,
44, 2201-2210.

Gologanu, M., Leblond, J., Devaux, J. (1993). Approximate models for ductile metals
containing non-spherical voids case of axisymmetric prolate ellipsoidal cavities.
Journal of the Mechanics and Physics of Solids 41, 1723-1754.

Gologanu, M., Leblond, J., Devaux, J. (1994). Approximate models for ductile metals
containing non-spherical voids case of axisymmetric oblate ellipsoidal cavities.
Journal of Engineering Materials and Technology 116, 290-297.

Gonzalez, C., Segurado, J., Llorca, J. (2004). Numerical simulation of elasto-plastic
deformation of composites: evolution of stress microfields and implications for
homogenization models. Journal of the Mechanics and Physics of Solids, 52, 1573-
1593.

Goods, S., Brown, L. (1979). The nucleation of cavities by plastic deformation, Acta
Metallurgica 27, 1-15.

Goto D. M. , Koss D. A. (1996), An experimental model of the growth of neighbouring
voids during ductile fracture, Scripta Materialia, 35, 459-463.

Gurland, J., Plateau, J. (1963), The mechanism of ductile rupture of metals containing
inclusions, Transactions of the ASME 56, 442-454.

Gurland, J. (1972). Observations on the fracture of cementite particles in a spheroidized
1.05% steel deformed at room temperature, Acta Metallurgica 20, 735-741.

Gurson, A.L. (1977). Continuum theory of ductile rupture by void nucleation and
growth Part I. Yield criteria and flow rules for porous ductile media. Journal of
Engineering Materials and Technology 99, 2-15.

Hill, R. (1965). Continuum micro-mechanics of elastoplastic polycrystals. Journal of the
Mechanics and Physics of Solids 13, 89-101.

Hadianfard, M.J., Smerd, R., Winkler, S., Worswick, M. (2008). Effects of strain rate on
mechanical properties and failure mechanism of structural Al-Mg alloys, Materials
Science and Engineering A 492 (2008) 283-292.

Horstemeyer, M.F., Gokhale, A.M. (1999). A void-crack nucleation model for ductile
metals. International Journal of Solids and Structures 36, 5029-5055.
223
Horstemeyer, M.F., Lathrop, J., Gokhale, A.M., Dighe, M. (2000). Modeling stress state
dependent damage evolution in a cast Al-Si-Mg aluminum alloy. Theoretical and
Applied Fracture Mechanics 33, 31-47.

Horstemeyer, M.F., Ramaswamy, S. and Negrete, M. (2003). Using a micromechanical
finite element parametric study to motivate a phenomenological macroscale model for
void/crack nucleation in aluminum with a hard second phase, Mechanics of Materials
35, 675-687.

Hu, X., Wilkinson, D., Jian, M., Mishra, R. (2008). Modeling strain localization using a
plane stress two-particle model and the influence of grain level matrix inhomogeneity.
Journal of Engineering Materials and Technology 130, 1-10.

Huang, Y., Hutchinson, J., Tvergaard, V. (1991). Cavitation instabilities in elastic
plastic solids Journal of the Mechanics and Physics of Solids 39, 223241.

Huang, Y. (1991). Accurate dilatation rates for spherical voids in triaxial stress fields.
Journal of Applied Mechanics 58, 1084-6.

Jones, M.K., Horstemeyer, M.F., Belvin, A.D. (2007). A multiscale analysis of void
coalescence in nickel. Journal of Engineering Materials and Technology 129, 94-104.

Kachanov, M., Tsukrov, I., Shafir, B. (1994). Effective moduli of solids with cavities of
various shapes. Applied Mechanics Review 47, 151-174..

Kailasam, K., Ponte Castaneda, P. (1998). A general constitutive theory for linear and
nonlinear particulate media with microstructure evolution. Journal of the Mechanics and
Physics of Solids 46, 427-465.

Kanetake, N., Nomura, M., Choh, T. (1995). Continuous observation of microstructural
degradation during tensile loading of particle reinforced aluminum composites. Material
Science and Technology 11, 1246-1252.

Kayikcioglu, T., Gangal, A., Ozer, M. (2000). Reconstructing ellipsoids from three
projection contours. Pattern Recognition Letters 21, 959-968.

Keralavarma, S. M. , Benzerga, A. A. , (2010). A constitutive model for plastically
anisotropic solids with non-spherical voids. Journal of Mechanics and Physics of
Solids 58, 874-90

Koplik J. , Needleman A. (1988). Void growth and coalescence in porous plastic
solids. International Journal of Solids and Structures 24, 835-853.

Kuna, M. , Sun, D. Z. (1996). Three-dimensional cell model analyses of void growth
in ductile materials. International Journal of Fracture 81, 235-258.

224
Lassance, D., Scheyvaerts, F., Pardoen, T. (2006). Growth and coalescence of penny-
shaped voids in metallic alloys. Engineering Fracture Mechanics 73, 1009-1034.

Leblond, J., Perrin, G. and Devaux, J. (1994). Bifurcation effects in ductile metals with
damage delocalization. Journal of Applied Mechanics 61, 236-242.

Lee, B.J., Mear, M.E. (1992). Axisymmetric deformation of power-law solids
containing a dilute concentration of spheroidal voids. Journal of Mechanics and Physics
of Solids 40, 1805-1836.

LeRoy, G., Embury, J.D., Edward, G., Ashby, M.F. (1981), A model of ductile fracture
based on the nucleation and growth of voids. Acta Metallurgica 29, 1509-1522.

Li, Y.J., Arnberg, L. (2004). Solidification structures and phase selection of iron-bearing
eutectic particles in a DC-cast AA5182 alloy. Acta Materialia 52, 2673-2681.

Liao, K., Pan, J. and Tang, S. (1997). Approximate yield criteria for anisotropic porous
ductile sheet metals. Mechanics of Materials 26, 213-226.

Liu, B., Qiu, X., Huang, Y., Hwang, K., Li, M., Liu, C. (2003). The size effect on void
growth in ductile materials. Journal of Mechanics and Physics of Solids 51, 1171-1187.

Maire, E., Grenier, J.C., Daniel, D., Baldacci, A., Klocker, H., Bigot, A. (2006).
Quantitative 3D characterization of intermetallic phases in an Al-Mg industrial alloy by
X-ray microtomography. Scripta Materialia 55, 123-126.

Mangan, M.A., Spanos, G. and Kral, M.V. (1999) Correlation Between the
Crystallography and Morphology of Proeutectoid Cementite Precipitates. Acta
Materialia 47, 4263-74.

McClintock, F. (1968a). On the mechanics of fracture from inclusions, Ductility, ASM,
Metals Park, Ohio, 255-277

McClintock, F. (1968b). A criterion for ductile fracture by the growth of holes. Journal
of Applied Mechanics 35, 363-371.

Mori, T., Tanaka, K. (1973). Average stress in matrix and average elastic energy of
materials with misfitting inclusions. Acta Metallurgica 21, 571-574.

Needleman, A. (1987). Continuum model for void nucleation by inclusion debonding.
Journal of Applied Mechanics 54, 525-531.

Niordson, C. (2008). Void growth to coalescence in a non-local material. European
Journal of Mechanics and Solids 27, 222-233.

Norton, R. (2000). Machine design: an integrated approach. Prentice Hall, New Jersey.
225
Orlov, O. (2006). A three-dimensional damage percolation model. Ph.D. Thesis,
University of Waterloo, Waterloo, Ontario, Canada, 2006

Pardoen, T., Delannay, F. (1998). The coalescence of voids in prestrained notched round
copper bars. Fatigue and Fracture of Engineering Materials and Structures 21, 1459-
1472.

Pardoen, T., Hutchinson, J. (2000). An extended model for void growth and
coalescence. Journal of the Mechanics and Physics of Solids 48, 2567-2512.

Pardoen, T. (2006). Numerical simulation of low stress triaxiality ductile fracture.
Computers and Structures 84, 1641-1650.

Perrin G. , Leblond J. (2000). Accelerated void growth in porous ductile solids
containing two populations of cavities. International Journal of Plasticity 16, 91-120.

Pierard, O., Gonzalez, C., Segurado, J., Llorca, J., Doghri, I. (2007). Micromechanics of
elasto-plastic materials reinforced with ellipsoidal inclusions. International Journal of
Solids and Structures 44, 6945-6962.

Pijaudier-Cabot, G., Bazant, Z.P. (1987). Nonlocal damage theory. Journal of
Engineering Mechanics 113, 1512-1533.

Ponte-Castaneda, P., Suquet, P., (1998). Nonlinear composites. Advances in Applied
Mechanics 34, 171301.

Ragab, A.R. (2004a). Application of an extended void growth model with strain
hardening and void shape evolution to ductile fracture under axisymmetric tension.
Engineering Fracture Mechanics 71, 1515-1534.

Rathod, M.J., Kutsuna, M. (2004). Joining of Aluminum Alloy 5052 and
Low-Carbon Steel by Laser Roll Welding. Welding Research, January, 16-26.

Reusch, F., Svendsen, B., Klingbeil, D. (2003). Local and non-local Gurson-based
ductile damage and failure modeling at large deformation. European Journal of
Mechanics and Solids 22, 779-792.

Rice, J., Tracey, D. (1969). On the ductile enlargement of voids in triaxial stress fields.
Journal of the Mechanics and Physics of Solids 17, 201-217.

Richelsen, A.B., Tvergaard, V. (1994). Dilatant plasticity on upper bound estimates for
porous ductile solid. Acta Metallurgica 42, 25612577.

Riks, E. (1979). An incremental approach to the solution of snapping and buckling
problems. International Journal of Solids and Structures 15, 529-551.

226
Rousselier G. (1987). Ductile fracture models and their potential in local approach of
fracture. Nuclear Engineering and Design 105, 97-111.

Saje, M., Pan, J., Needleman, A. (1982). Void nucleation effects on shear localization in
porous plastic solids. International Journal of Fracture 19, 163-182.

Scheyvaerts, F., Pardoen, T., Onck, P.R. (2010). A new model for void coalescence by
internal necking International Journal of Damage Mechanics 19, 95-126

Scheyvaerts, F., Onck, P.R., Tekoglu, C., Pardoen, T. (2011). The growth and
coalescence of ellipsoidal voids in plane strain under combined shear and tension.
Journal of Mechanics and Physics of Solids 59, 373-397.

Segurado, J., Llorca, J. (2002). A numerical approximation to the elastic properties of
sphere-reinforced composites. Journal of the Mechanics and Physics of Solids 50, 2107-
2121.

Shabrov, M., Needleman, A. (2002). An analysis of inclusion morphology effects on
void nucleation. Modeling and Simulation in Materials Science and Engineering 10, 63-
183.

Siad, L., Ouali, M.O., Benabbes, A. (2008). Comparison of explicit and implicit finite
element simulations of void growth and coalescence in porous ductile materials.
Materials and Design 29, 319-329.

Somekawa, H., Tanaka, T., Sasaki, H., Kita, K., Inoue, A., Higashi, K. (2004). Diffusion
bonding in ultra-fine-grained Al-Fe alloy indicating high-strain-rate superplasticity. Acta
Materialia 52, 1051-1059.

Sovik, O., Thaulow, C. (1997). Growth of spheroidal voids in elastic-plastic solids.
Fatigue and Fracture in Engineering Materials and Structures 20, 1731-44.

Sun, Y. (1995). Constitutive equations for ductile materials containing large and small
voids. Mechanics of Materials 19, 119-127.

Suquet, P., (1995). Overall properties of non-linear composites: a modified secant
moduli theory and its link with Ponte Castenedas nonlinear variational procedure.
Comptes rendus de Academi des Sciences, Paris, Serie IIB 320, 563-571.

Suquet, P., 1997. Effective properties of nonlinear composites. In: Continuum
Micromechanics. CISM Course and Lecture Notes. pp. 197264

Tanaka, K., Mori, T., Nakamura, T. (1970), Cavity formation at the interface of a
spherical inclusion in a plastically deformed matrix, Philosophical Magazine 21, 267-
279.

227
Tandon, G.P., Weng, G.J. (1986). Average stress in the matrix and effective moduli of
randomly oriented composites. Composites Science and Technology 27, 111-132

Tani, J., Kido, H. (2008). Lattice dynamics of Mg
2
Si and Mg
2
Ge compounds from first-
principles calculations. Computational Materials Science 42, 531-536.

Thomason, P. (1985a). Three-dimensional models for the plastic limit-loads at incipient
failure of the intervoid matrix in ductile porous solids. Acta Metallurgica 33, 1079-
1085.

Thomason, P. (1985b). A three-dimensional model for ductile fracture by the growth
and coalescence of micro-voids. Acta Metallurgica 33, 1087-1095.

Thomason, P. (1990). Ductile fracture of metals. Pergamon Press Inc., Oxford.

Thomason, P.F. (1993). Ductile fracture by the growth and coalescence of microvoids of
non-uniform size and spacing. Acta Metallurgica 41, 2127-2134.

Thomason, P. (1998). A view on ductile-fracture modeling. Fatigue and Fracture of
Engineering Materials and Structures 21, 1105-1122.

Thomson C. I. A. , Worswick M. J. , Pilkey A. K. , Lloyd D. J. and Burger G. (1999).
Modeling void nucleation and growth within periodic clusters of particles, Journal of
the Mechanics and Physics of Solids 47, 1-26.

Tohgo, K., Mochizuki, K. (2002). Damage mechanics approach to material-dependency
of fracture toughness in aluminum alloys. International Journal of Damage Mechanics
11, 151-170.

Toi, Y., Kang, S.S. (2005). Mesoscopic natural element analysis of elastic moduli, yield
stress and fracture of solids containing a number of voids. International Journal of
Plasticity 21, 2277-2296.

Tvergaard, V. (1981). Influence of voids on shear band instabilities under plane strain
conditions. International Journal of Fracture 17, 389-407.

Tvergaard, V. (1982). Ductile fracture by cavity nucleation between larger voids.
Journal of the Mechanics and Physics of Solids 30, 265-286.

Tvergaard, V., Needleman, A. (1984). Analysis of the cup-cone fracture in a round
tensile test bar. Acta Metallurgica 32, 157-169.

Tvergaard, V., Needleman, A. (1995). Effects of non-local damage in porous plastic
solids. International Journal of Solids and Structures 8, 1063-1077.

228
Tvergaard, V., Needleman, A., 1997. Nonlocal effects on localization in a void-sheet.
International Journal of Solids and Structures 34, 22212238.


Tvergaard, V., Niordson, C. (2004). Nonlocal plasticity effects on interaction of
different size voids. International Journal of Plasticity 20, 107-120.

Voce, E. (1948). The relationship between stress and strain for homogeneous
deformation. Institute of Metals 74, 537562.

Wang, D., Pan, J., Liu, S. (2004). An anisotropic yield criterion for porous ductile sheet
metals with planar anisotropy. International Journal of Damage Mechanics 13, 7-33.

Weck, A. (2006). The Role of Coalescence on Ductile Fracture. Graduate Seminar,
McMaster University, Hamilton ON, Canada

Wen, J., Huang, Y., Hwang, K., Liu, C., Li, M. (2005). The modified Gurson model
accounting for the void size effect. International Journal of Plasticity 21, 381-395.

Weng, G.J. (1984). Some elastic properties of reinforced solids, with special reference
to isotropic ones containing spherical inclusions. International Journal of Engineering
Science 22, 845-856.

Weng, G.J. (1990). The overall elastoplastic stress strain relations of dual-phase metals.
Journal of the Mechanics and Physics of Solids 38, 419-441.

Worswick M. J. , Pick R. J. (1990). Void growth and constitutive softening in a
periodically voided solid. Journal of the Mechanics and Physics of Solids 38, 601-
625.

Worswick, M.J., Pelletier, P. (1998). Numerical simulation of ductile fracture during
high strain rate deformation. European Physical Journal Applied Physics 4, 257-267.

Worswick M. J. , Chen Z. T. , Pilkey A. K. , Lloyd D. , Court S. (2001). Damage
characterization and damage percolation modeling in aluminum alloy sheet. Acta
Materialia 49, 2791-2803.

Xiong, W., Qin, X., Wang, L. (2007). Densification Behaviour of Nanocrystalline
Mg2Si Compact in Hot Pressing. Journal of Material Science and Technology 23, 595-
598.

Xue, L. (2007). Constitutive modeling of void shearing effect in ductile fracture of
porous materials. Engineering Fracture Mechanics 75, 3343-3366.

Yamagiwa, K., Watanabe, Y., Fukui, Y., Kapranos, P. (2003). Novel Recycling System
of Aluminum and Iron Wastes in-situ Al-Al3Fe Functionally Graded Material
Manufactured by a Centrifugal Method. Materials Transactions 44, 2461-2467.

229
Yee, K.C., Mear, M.E. (1996). Effect of void shape on the macroscopic response of non-
linear porous solids. International Journal of Plasticity 12, 45-68.

Zhang, Y., Chen, Z. (2007). On the effect of stress triaxiality on void coalescence.
International Journal of Fracture 143, 105-112.

Zhang, Z. Niemi, E. (1994a). Studies on the ductility predictions by different local
failure criteria. Engineering Fracture Mechanics 48, 529-540.

Zhang, Z., Niemi, E. (1994b). Analyzing ductile fracture using dual dilational
constitutive equations. Fatigue and Fracture of Engineering Materials and Structures
17, 695-707.

Zhang, Z., Thaulow, C., Odegard, J. (2000). A complete Gurson model approach for
ductile fracture. Engineering Fracture Mechanics 67, 155-168.

Zhang, Z.L., Skallerud, B. (2010). Void coalescence with and without prestrain history,
International Journal of Damage Mechanics 19, 153-174.



230
Appendix A Eshelby Tensors


The components of Eshelbys fourth-order S and H tensors (Eshelby, 1957) for an
ellipsoidal inclusion embedded within an isotropic matrix with a Poissons ratio, v , can
be determined using the following procedure.

Select a coordinate system such that the semi-axes of the ellipsoid are a b c > > . The
non-zero components of the S tensor are now defined as


2
1111 11 1
S a QI RI = +
2
2222 22 2
S b QI RI = +
2
3333 33 3
S c QI RI = +


2
1122 12 1
S b QI RI =
2
1133 13 1
S c QI RI =
2
2211 12 2
S a QI RI =


2
2233 23 2
S c QI RI =
2
3311 13 3
S a QI RI =
2
3322 23 3
S b QI RI = (A1)

( )
2 2
1212 12 1 2
( )
2 2
Q R
S a b I I I = + + + ( )
2 2
2323 12 2 3
( )
2 2
Q R
S b c I I I = + + +

( )
2 2
3131 13 1 3
( )
2 2
Q R
S a c I I I = + + +
3
8 (1 )
Q
v t
=


1 2
8 (1 )
v
R
v t



and the non-zero components of the rotation tensor, H are


1 2
1212
8
I I
t

H =
2 3
2323
8
I I
t

H =
3 1
3131
8
I I
t

H = (A2)

The I parameters are computed from the expressions


( )
( )
1
2 2 2 2
4 abc F E
I
a c a c
t
=


( )
2 2
3
2 2 2 2
4 abc b a c
I E
ac b c a c
t
| |

=
|
|

\ .

2 1 3
4 I I I t =


2 1
12
2 2
I I
I
a b


3 1
13
2 2
I I
I
a c


3 2
23
2 2
I I
I
b c

(A3)

2
12 13
11
4 /
3
a I I
I
t
=
2
12 23
22
4 /
3
b I I
I
t
=
2
13 23
33
4 /
3
c I I
I
t
=

231
where F and E are elliptic integrals of the first and second kinds with amplitude and
modulus


( )
2 2
0
,
1 sin
dw
F F k
k w
u
u = =

}

( ) ( )
1
2 2
2
0
, 1 sin E E k k w dw
u
u = =
}
(A4, A5)

( ) ( )
2 2 2 2 2
/ k a b a c =
( )
1 2 2
sin 1 / c a u

=

The relations for the I parameters must be computed using a slightly different
procedure if the inclusion is a sphere or a spheroid. For an oblate spheroid ( a b c = > ),
the I coefficients that have changed from Eq. (A3) are

( )
2 2
1
1 2
3/ 2
2
2 2
2
cos 1
a c c c c
I I
a a c
a c
t


| |
= =
`
|
\ .
)

3 1
4 2 I I t =
13
12
2
4
I
I
a
t
= (A6)

while for prolate inclusions ( a b c > = )

( )
2 2
1
2 3
3/ 2
2
2 2
2
1 cosh
ac a a a
I I
c c c
a c
t


| |
= =
`
|
\ .
)

1 2
4 2 I I t =
12
23
2
4
I
I
b
t
= (A7)

Finally, for spherical inclusions ( a b c = = ), the non-zero components of the S tensor are


1111 2222 3333
7 5
15(1 )
v
S S S
v

= = =


1212 2323 3131
4 5
15(1 )
v
S S S
v

= = =

(A8)

1122 2233 3311 1133 2211 3322
5 1
15(1 )
v
S S S S S S
v

= = = = = =



and all of the terms in the H tensor are set to zero.
232
Appendix B Important Geometrical Relationships
of the Unit Cell


A number of important and useful relations can be obtained to describe the
microstructure of a voided unit cell and its evolution. First, consider a general voided
unit cell subjected to an arbitrary deformation process. The unit cell has the half-lengths
L
1
, L
2
, L
3
that are coincident with the principal loading directions. The unit cell contains
an arbitrarily oriented ellipsoidal void with semi-axes denoted as R
1
, R
2
and R
3
. The
porosity (void volume fraction) of the cell is defined as


void 1 2 3 1 2
3
2 1 2 3
cell 1 2 3 1 2
f V R R R W W
V L L L
_ _ _ _ _

= = = = (B1)

where
i
W are the void aspect ratios;
i
are the cell aspect ratios;
i
_ are the void spacing
(or ligament size) ratios and is a shape parameter specific to the assumed unit cell
with 2/ 3 = for an axisymmetric unit cell and / 6 t = for a cubic cell.

The void aspect ratios, spacing ratios, and cell aspect ratios are defined as


1
1
2
R
W
R
= ,
3
2
2
R
W
R
= ,
i
i
i
R
L
_ = ,
1
1
2
L
L
= ,
3
2
2
L
L
= (B2-B6)

Taking the derivative of Equations (1-6) with respect to time, it is straightforward to
obtain the following relations as a function of the macroscopic principal strain rates as


1 2 3
1 2 3
1 2 3
f R R R
f R R R
c c c = + + (B7)

1 1 2
1 1 2
W R R
W R R
= ,
2 3 2
2 3 2
W R R
W R R
= ,
i i
i
i i
R
R
_
c
_
= ,
1
1 2
1

c c

= ,
2
3 2
2

c c

= (B8-11)

From Equations (B7-11), the growth rates of the void semi-axes can be expressed as

233

1 1 2
1 2 3
1 1 2
1
2
3
R f W W
R f W W
c c c
| |
= + + + +
|
\ .
(B12)


2 1 2
1 2 3
2 1 2
1
3
R f W W
R f W W
c c c
| |
= + + +
|
\ .
(B13)


3 1 2
1 2 3
3 1 2
1
2
3
R f W W
R f W W
c c c
| |
= + + + +
|
\ .
(B14)

Alternatively, using only the void spacing ratios, the evolution rate of the porosity can
be expressed as

1 2 3
1 2 3
f
f
_ _ _
_ _ _
= + + (B15)

The above equations are valid for any type of unit cell or ellipsoidal void geometry
and reduce accordingly depending on the type of geometry considered. For example, an
axisymmetric unit cell with the void semi-axes coincident with the principal loading
directions:
2 2
/ 0 W W = ,
2 2
/ 0 = and
2 2 3 3
/ / _ _ _ _ = .
234
Appendix C - Ellipsoids

1. General equation of an ellipsoid

The expanded Cartesian form for a general ellipsoid is


2 2 2
0 Ax By Cz Dxy Exz Fyz Gx Hy Jz K + + + + + + + + + =

(C1)


where the coefficients G, H, J, K, are related to the ellipsoid centroid as


2
o o o
G Ax Dy Ez = 2
o o o
H By Dx Fz = 2
o o o
J Cz Ex Fy =


2 2 2
1
o o o o o o o o o
K Ax By Cz Dx y Ex z Fy z = + + + + +


The coefficients that define the ellipsoid semi-axes and directions are consolidated
into a matrix, P, as


/ 2 / 2
/ 2 / 2
/ 2 / 2
A D E
P D B F
E F C
| |
|
=
|
|
\ .
(C2)


to give the matrix form of an ellipsoid located at the origin


T
1 1
1 P = x x (C3)


The semi-axes and their directions can be determined from an eigenvalue solution of the
P matrix.

2. Radius of ellipsoid along a viewing direction

Consider a three-dimensional ellipsoid with semi-axes (a, b, c) that are oriented
along the respective directions (e
1
, e
2
, e
3
) with e
i
= [l
i
, m
i
, n
i
]
T
.
The viewing direction, u
= [u
1
, u
2
, u
3
]
T
,

can be transformed to a vector, v, the ellipsoidal frame as


R = v u

1 1 1
2 2 2
3 3 3
l m n
R l m n
l m n
| |
|
=
|
|
\ .
(C4)

235


The viewing vector within the ellipsoidal frame can be converted to a pair of spherical
angles as


1
2 1
tan ( / ) v v u

=

1
3
cos ( / ) v |

= v
(C5)

The distance from the center of the ellipsoid to its surface can then be easily determined
from the parametric formulation of the ellipse as


1
2 2 2
2
cos sin sin sin cos
R
a b c
u | u | |

(
| | | | | |
= + +
(
| | |
\ . \ . \ .
(

(C6)



3. Line projection of an ellipsoid

The line projection of a general three-dimensional ellipsoid can be determined using
the method of Kayikcioglu et al. (2000). Consider an arbitrary ellipsoid located at the
origin of a reference space
T
1
[ ] = x y z x that is being viewed along the direction defined
by
T
1 2 3
[ ] u u u = u . The equation of the ellipsoid in matrix form is


T
1 1
1 P = x x (C7)


The projection of the ellipsoid in a plane parallel to the viewing direction is an ellipse
with semi-axes (a
p
, b
p
) that is oriented in the space,
T
2
[ ] = u p n x , as illustrated in Figure
X. The projection space is related to the reference space by the transformation matrix,
Q, as

1 2
Q = x x (C8)


cos sin sin cos cos
sin cos sin cos
cos 0 sin
Q
u | u u |
u u u |
| |
| |
|
=
|
|
\ .
(C9)

where = atan2(u
2
, u
1
) and
1
3
cos ( / ) = u |

u .
236
projected ellipse
ellipsoid
view direction
n
x
y
z
p
u
a
p
b
p
u

Figure C1: Geometry of arbitrary ellipsoid and its line projection

The equation of the projected ellipse within the p-n plane is


T
2 2
1 S = x x (C10)

where S is the coefficient matrix of the ellipse and is obtained by


T
S =Q PQ (C11)

The orientation and length of the ellipse semi-axes are obtained from an eigenvalue
solution of the S matrix. The semi-axes are related to the eigenvalues,
i
, by
2
1/
p
a =
and
1
1/
p
b = . The orientation of the semi-axes in the reference frame can then be
readily determined using Eq. (8).

4. Arbitrary stretching of an ellipsoid

Consider a three-dimensional ellipsoid with semi-axes (a, b, c) that are oriented
along the respective directions (e
1
, e
2
, e
3
) with e
i
= [l
i
, m
i
, n
i
]
T
.
The ellipsoid is to be
stretched by the quantities (s
1
, s
2
, s
3
) in the respective directions (s
1
, s
2
, s
3
) with s
i
= [l
si
,
m
si
, n
si
]
T
. To determine the shape and orientation of the ellipsoid after stretching, the
coefficient matrix of the new ellipsoid must be computed. These coefficients can be
237
obtained by performing successive transformations of the points on the surface of a unit
sphere. First, the sphere is stretched and rotated to obtain the ellipsoid in its desired
orientation using the tensors, S
abc
and R
abc
. The points on the ellipsoid are then rotated to
the stretching system using, R
s
, stretched by, S
s
, and rotated back to the initial system.
The resulting transformation between the points on the surface of the ellipsoid, x
E
,

and
the unit sphere, x
S
, in the reference space of sphere x
S
=[x, y, z]
T
is expressed as


E S
T = x x

(C12-17)

T T
s s abc abc
T R S R S =


1 1 1
T
2 2 2
3 3 3
abc
l m n
R l m n
l m n
| |
|
=
|
|
\ .

0 0
0 0
0 0
abc
a
S b
c
| |
|
=
|
|
\ .

1 1 1
2 2 2
3 3 3
s s s
s s s s
s s s
l m n
R l m n
l m n
| |
|
=
|
|
\ .

1
2
3
0 0
0 0
0 0
s
s
S s
s
| |
|
=
|
|
\ .


The equation of a unit sphere in matrix form is expressed as


T
1
S S
= x x

(C18)


and by substituting in the transformation defined by Eq. (12), the equation of the new
ellipsoid can be obtained as


T
1
E E
P = x x

T
1 1
P T T

( =

(C19, C20)


The new orientation and semi-axes of the ellipsoid are obtained from an eigenvalue
solution of the P matrix. The semi-axes are related to the eigenvalues,
i
, as
3
1/ a = ,
2
1/ b =

and
1
1/ c = for the ordered eigenvalues:
1 2 3
> > .

You might also like