Você está na página 1de 147

0123440151363738938165

5 3 938
15 4 813 938 38383
!"#$%$&"%' ()*+,(
-*++()*./($&%#+()*0)*+()*1)*()*"#2 )*
3"4#5"5( !6+()*7(+(
8!$(95+",+((()*&$
5*6+5$8+(($$+
:5#5$:
'; (
;+)*$<=+:(+$%$(>&((8?.95?8+$/+)*$
=+:(+$%$(>&((8?.95?1($-?' +
0 56%
"#+)*@"%&"5<AB?C"#+DEAA
8+(8+(($$++($ "&9$$(+$1)*()*"#F+F#+$*!#+:&"%5F?

42567289$:;<=%<3>>>
45 14242 1 4 2

!

"#!!"##$%#$&'()*'
+*#,-(./!(#,0(1')23)/.

012345745829

2 14  

012314567891
1 46714 82319 2 9 
9 714 732123149
!"#$%"&'#'&#$(&##)'&#"#$%#%
!"#$%"&'#'&#$*+",#-."#&.'#'&#$*+")#."""+"/"#&"'&#%0%+%"
)122.%'3.4%'3.3
5*&31678#39:;<"%=%#/+#>?7@AAB

C$)>+#*D"(+E+&"*7@AA
<&&+#*++/.3!$)"+$,#)'&#"#$%F">'+)+$..?$+.#%"
++#/"&>F?$++"%F#.?#%"%>,$+F$+'>"%>F"%?&+$%#?
F"%#"&?)$$$)>#%*?+$+.#%*$+$+G#?G#$)+#$+)+F##$%
$,)'&#+3
H+#%.#%I+F"%>3
0DJ!KL64M46NN@4@NNO4P
0DD!@KNP4@6LN
D"(+E+&"*IF';QH3R3JRSA@A6A6Q4P7@A6<"%
H$%1@@NK27N@L2KPKO4@Q&,"T1@@NK27N@L2KPKO4K
0%+%1GGG3"(+3.Q4F"#&1#%,$U"(+3.

Zusammenfassung
Versagensprognose von metallischen Werkstoe ist von groem Interesse in der Automobilindustrie, da sie einen wesentlichen Beitrag zur Verbesserung der Crashsicherheit von
Karosseriebauteilen liefert. In der vorliegenden Arbeit wird der Einuss des Spannungszustandes, insbesondere des Lode-Winkel-Parameters (oder der dritten deviatorischen Spannungsinvariante), auf die Schadigungsmodellierung diskutiert und mit Hilfe experimenteller
und numerischer Untersuchungen validiert. Die Implementierung dieser Modellierung f
uhrt
zu einer Erweiterung des Schadigungsmodells GISSMO (Generalized Incremental Stress State
ur 3D-Anwendungen um die Ber
uckdependant damage MOdel) von Neukamm et al. [14] f
sichtigung des Lode-Winkel-Parameters. Der Spannungszustand wird somit durch zwei Spannungszustand-Parameter eindeutig deniert: Die Triaxialitat und der Lode-Winkel-Parameter.
In Folge davon ergibt sich die Duktilitat (oder Bruchdehnung) als Funktion von Triaxialitat
und Lode-Winkel-Parameter.
Der Triaxialitat- und Lode-Winkel-Parameter-Raum wird mit einer Reihe von Versuchen
f
ur das Dualphasenstahl DP600 abgetastet. Versuche mit gekerbten Rundzugproben, glatten
Flachzugproben und Nakazima-Proben werden durchgef
uhrt, um das jeweilige Materialverhalten f
ur die Lode-Winkel-Parameter-Werte 1, 0 und -1 zu untersuchen. Zusatzlich werden
f
ur dazwischenliegende Spannungszustande Versuche mit Buttery-Proben und Flachzugproben durchgef
uhrt.
Die Bruchdehnung eines jeden Versuchs wird durch eine Kombination experimenteller und
numerischer Ergebnissen bestimmt. Parallel werden entsprechende Spannungszustandsparameter mit Hilfe von Gewichtungsfunktionen numerisch bestimmt. Diese in der Arbeit
vorgeschlagene Gewichtungsfunktionen hangen von der nichtlinearen Schadigungsakkumulationsformulierung im GISSMO-Schadigungsmodell.
Die vom Spannungszustand abhangige Bruchdehnungsformulierung wird in den kommerziellen FE-Code LS-DYNA implementiert. Eine auf einem Spannungszustand basierende
analytische Beschreibung der Bruchdehnungsache mit neun Parametern wird vorgeschlagen.
Auerdem wird eine mathematische Funktion der Bruchdehnung basierend auf der biharmonischen Spline-Methode vorgestellt.
Die Untersuchungen zeigen den Einuss des Lode-Winkel-Parameters auf die Duktilitat

ii
des untersuchten Materials DP600. Mit dem vorgestellten Ansatz zur Materialparameteran
passung wird eine gute Ubereinstimmung
zwischen experimentellen und numerischen KraftWeg-Kurven erreicht.

Abstract
Numerical fracture prediction of metals is of great interest in automotive industry, since it is
an eective way to improve crashworthiness of car body parts. In the present thesis, the eect
of stress state on damage modeling with the focus on the Lode angle parameter (or third
deviatoric stress invariant) is discussed and validated by experimental and numerical studies.
The numerical implementation is integrated to the damage model GISSMO (Generalized
Incremental Stress State dependant damage MOdel) as an extension, which was proposed by
Neukamm et al. [14]. The model is extended for 3D usage by utilization of the Lode angle
parameter. The stress state is dened with two stress state parameters, stress triaxiality and
Lode angle parameter uniquely. The material ductility (or fracture strain) is considered as a
function of the stress triaxiality and Lode angle parameter.
The stress triaxiality and Lode angle parameter space is covered with a series of tests
for the dual-phase steel DP600. Tests of axisymmetric notched round specimens, grooved
at specimens and Nakazima were conducted to study the material behavior for Lode angle
parameter equal to 1, 0 and -1, respectively. Additionally, for the intermediate stress states,
tests of buttery specimens and at specimens were carried out.
The representative fracture strain of each test is obtained by combining experimental and
numerical results. The corresponding stress state parameters are determined numerically by
using proposed weighting functions, which depend on the nonlinear damage accumulation
formulation in the GISSMO damage model.
The stress state dependent fracture strain formulation is implemented into the commercial
nite element code LS-DYNA. A nine-parameter stress state dependent analytical fracture
locus is proposed. In addition, a mathematical fracture strain function based on biharmonic
spline method is introduced.
The investigations on the material DP600 show the inuence of Lode angle parameter on
the fracture strain. With the introduced calibration approach a good correlation is obtained
between experimental and numerical force-displacement curves.

iv

Acknowledgments
There are many people who have contributed to the work presented in this thesis directly
and indirectly. Without their help and support it would not be possible to nish the current
work.
First of all, I would like to thank Professor Weichert for his guidance and support throughout the thesis. His critical comments kept me on the right track and increased my productivity.
I would like to express my appreciation to Professor Maier for the constructive remarks.
I would like to give my special thanks to the supervisor Dr. Sven David Wolkerling
for his friendliness guidance and support in last three years. I appreciated the constructive
discussions.
I have been very happy to work with Dr. Markus Feucht and Frieder Neukamm. I appreciated the interesting discussions and guidance very much. The discussions we had, improved
my understanding in the application of the ductile fracture models in crashworthiness simulations signicantly. I would also like to Paul Du Bois for his collaboration and for his fruitful
discussions.
I would like to thank to the head of the team Steen Hampel for giving me an opportunity
doing the research in his team.
I am thankful to Dr. Levent Aktay for the reviews during my PhD and continuous support
within three years.
I would like to thank to Dr. Gerhard Summ for the support concerning software and
very friendly discussions. I would like to express my graduates to team members Elzbieta
Skurski, Peter K
ummerlen, David Moncayo, Olivier Cousigne, Maxime Dagonet, Christian
Bou Farhat and Siu Ping Li for their friendship.
The meetings in IWM Freiburg improved my understanding in the subject damage modeling for crashworthiness simulations further. I would like to thank to Dr. Sun for the good
cooperation. I would like to thank to Dr. Florance Andreux for the cooperative work with numerical studies. Thanks are due to Clemense Fehrenbach and Dennis Holletzek for conducting
the experiments.
Finally, I would like to express my gratitude to my parents and my sister, who supported
me continuously in the last three years.

vi

Contents
1 Introduction

1.1

Background and motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2

Thesis objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.3

Outline of the thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2 A Brief Review of Damage Modeling Approaches

2.1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.2

Stress state characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.3

Damage modeling approaches . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.3.1

Micromechanical models . . . . . . . . . . . . . . . . . . . . . . . . . .

2.3.2

Conventional continuum mechanics models . . . . . . . . . . . . . . . . 13

2.3.3

Continuum damage mechanics . . . . . . . . . . . . . . . . . . . . . . . 14

3 Stress State Dependence of Damage Modeling


3.1

3.2

17

Plasticity model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.1.1

Experiments on stress triaxiality dependence . . . . . . . . . . . . . . . 19

3.1.2

Experiments on Lode angle parameter dependence . . . . . . . . . . . . 20

Material ductility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.2.1

Stress triaxiality dependence . . . . . . . . . . . . . . . . . . . . . . . . 21

3.2.2

Lode angle parameter dependence . . . . . . . . . . . . . . . . . . . . . 22

4 The GISSMO Damage Model

27

4.1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

4.2

Damage rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2.1

Damage variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

4.2.2

Experimental damage quantication

4.2.3

Damage accumulation rule . . . . . . . . . . . . . . . . . . . . . . . . . 30

. . . . . . . . . . . . . . . . . . . 30

4.3

Fracture strain denition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

4.4

Instability and coupling of damage with plasticity . . . . . . . . . . . . . . . . 33

viii

CONTENTS

5 Experimental Program
35
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
5.2 Un-notched and notched at specimens . . . . . . . . . . . . . . . . . . . . . . 37
5.3
5.4

5.2.1 Anisotropy of the material . . . . . . . . . . . . . . . . . . . . . . . . . 39


Grooved at specimens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
Axisymmetric notched round specimens . . . . . . . . . . . . . . . . . . . . . . 44

5.5
5.6

Nakazima tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
Buttery tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

6 Determination of Fracture Strain and Stress State


53
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
6.2 Discussion on mesh size eects . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
6.3 Determination of the stress-strain curve . . . . . . . . . . . . . . . . . . . . . . 56
6.4
6.5
6.6

Weighting function for stress state parameters . . . . . . . . . . . . . . . . . . 58


Damage exponent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
Numerical simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.6.1
6.6.2
6.6.3
6.6.4

6.7

Flat specimens . . . . . . . . . . . . . .
Grooved at specimens . . . . . . . . . .
Axisymmetric notched round specimens .
Nakazima tests . . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

62
66
69
73

6.6.5 Buttery tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75


Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

7 Determination of Fracture Locus


87
7.1 Simulation with analytical fracture strain denition . . . . . . . . . . . . . . . 87
7.1.1 Calibration of parameters . . . . . . . . . . . . . . . . . . . . . . . . . 87
7.2

7.3

7.1.2 Numerical simulations . . . . . . . . . . . . .


Simulation with mathematical fracture strain surface
7.2.1 Fracture strain surface determination . . . . .
7.2.2 Numerical simulations . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

90
97
97
98

Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

8 Conclusions and Future Research


105
8.1 Conclusions of the present thesis . . . . . . . . . . . . . . . . . . . . . . . . . . 105
8.2 Future research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
A Derivation of stress state dependent fracture function

109

Bibliography

122

List of Figures
1.1

2.1
2.2

Investigation of numrical modeling of fracture at dierent levels: Elementary


specimens, components and full car. . . . . . . . . . . . . . . . . . . . . . . . .

Representation of the stress vector in Haigh-Westergaard space (1 , 2 , 3 ) and


cylindrical coordinates (r, , z). . . . . . . . . . . . . . . . . . . . . . . . . . .

Geometrical representation of the principal stresses (1 , 2 , 3 ), deviatoric principal stresses (s1 , s2 , s3 ) on the octahedral plane. . . . . . . . . . . . . . . . . .

2.3

Illustration of two fracture mechanisms: Flat dimple fracture and shear dimple
fracture. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3.1

Representation of yield surfaces, Lode angle independent (von-Mises and DruckerPrager) and Lode angle dependent (Tresca and Mohr-Coulomb). . . . . . . . . 18

3.2

Comparison of equivalent stress-strain curves for (a) 2024-T351 aluminum extracted from tension, torsion and compression experiments [5] and (b) 5083
aluminum extracted from torsion and compression experiments [6]. . . . . . . . 20

3.3

Average stress triaxiality vs. fracture strain curves obtained from notched axisymmetric bars for steels and aluminum alloys. . . . . . . . . . . . . . . . . . 23

3.4

Comparison of average stress triaxiality-fracture strain curves for axisymmetric


tension = 1 and plane strain = 0 for steel grades and aluminum alloys. . . 24

3.5

Bounding curves of the fracture locus proposed by Xue and Wierzbicki [7] and
Bai and Wierzbicki [8]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

4.1

Integration of the GISSMO damage model to the process chain consisting of


metal forming and crashworthiness simulations. . . . . . . . . . . . . . . . . . 28

4.2

Representative volume element subjected to uniaxial tension. . . . . . . . . . . 29

4.3

Evolution comparison of linear damage, nonlinear damage and normalized void


volume fraction of GTN model with respect to plastic strain. . . . . . . . . . . 32

4.4

Proposed fracture locus and formulations for bound limit curves. . . . . . . . . 33

5.1

Schematic representation of AHSS compared to low strength and HSS steels. . 36

LIST OF FIGURES
5.2

Layout of machined specimens on the DP600 sheet plate for at, grooved at
and axisymmetric notched round specimens. . . . . . . . . . . . . . . . . . . . 36

5.3

Illustration of covered stress state by the investigated sets of specimens and


tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

5.4

The un-notched and notched at specimens. . . . . . . . . . . . . . . . . . . . 38

5.5

Normalized force-displacement curves obtained from the tests of un-notched


and notched at specimens. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

5.6

Fracture surface of the un-notched at specimen. . . . . . . . . . . . . . . . . 39

5.7

The post-mortem of un-notched and notched at specimens. . . . . . . . . . . 39

5.8

SEM micrograph of the microstructure of undeformed DP600 steel and fracture


surface of the un-notched at specimen. . . . . . . . . . . . . . . . . . . . . . . 40

5.9

Representation of the thickness and width of the specimen before and after
the deformation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

5.10 Evolutions of the r value and normalized force with respect to global normalized displacement for three duplicate tests of un-notched at specimen. . . . . 41
5.11 Critical dimensions for the grooved at specimens. . . . . . . . . . . . . . . . . 42
5.12 Evolution of the Lode angle parameter with respect to equivalent plastic strain
with constant tm =0.8mm for dierent ratios of w/tm and t/tm . . . . . . . . . 42
5.13 The geometry and dimensions of the grooved at specimens. . . . . . . . . . . 43
5.14 The normalized force-displacement curves obtained from tests of grooved at
specimens. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.15 The grooved at specimens after fracture. . . . . . . . . . . . . . . . . . . . . 44
5.16 The geometry and dimensions of the axisymmetric notched round specimens. . 45
5.17 Normalized force-displacement curves obtained from the tests of axisymmetric
notched round specimens. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.18 The post-mortem of axisymmetric notched round specimens. . . . . . . . . . . 45
5.19 The geometry and dimensions of the Nakazima test blank geometries. . . . . . 46
5.20 Punch force-displacement curves obtained from Nakazima tests. . . . . . . . . 47
5.21 The post-mortem of Nakazima tests blanks. . . . . . . . . . . . . . . . . . . . 47
5.22 Illustration of the buttery specimen. . . . . . . . . . . . . . . . . . . . . . . . 49
5.23 The geometry and dimensions of the buttery specimen. . . . . . . . . . . . . 49
5.24 Buttery test bench set up for 60 loading angle. . . . . . . . . . . . . . . . . 50
5.25 Representation of tested buttery specimen loading angles. . . . . . . . . . . . 50
5.26 Representation of the tracking points used for measurements of buttery tests. 50
5.27 Normalized force-displacement curves obtained from the buttery tests. . . . . 51

LIST OF FIGURES

xi

6.1

Comparison of the numerical simulations for the mesh size of 0.1mm, 0.05mm
and 0.025mm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

6.2

Comparison of the experimental engineering stress-strain curves with the numerical one calculated with the calibrated true stress-strain curve. . . . . . . . 58

6.3

Stepwise methodology applied by Tai [9]. . . . . . . . . . . . . . . . . . . . . . 60

6.4

Damage quantication with the methodology of density measurement [10] and


the GISSMO damage rule application. . . . . . . . . . . . . . . . . . . . . . . 61

6.5

Plastic strain, stress triaxiality and Lode angle parameter distribution at the
critical section of the un-notched at specimen for global displacement at the
onset of necking and at fracture. . . . . . . . . . . . . . . . . . . . . . . . . . . 63

6.6

Comparison of the experimental and numerical normalized force-displacement


curves. The plastic strain with respect to global normalized displacement at
the critical location is shown with the blue curve. . . . . . . . . . . . . . . . . 63

6.7

Evolution of two stress state parameters with respect to plastic strain and
weighted stress state parameters for the un-notched at specimen. . . . . . . . 64

6.8

Comparison of the experimental and numerical normalized force-displacement


curves for notched at specimens. . . . . . . . . . . . . . . . . . . . . . . . . . 64

6.9

Evolution of the stress triaxiality and Lode angle parameter with respect to
plastic strain at critical location and weighted stress state parameters for the
notched at specimens. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

6.10 Plastic strain, stress triaxiality and Lode angle parameter distributions at
fracture formation for the grooved at specimens with groove radii R=1mm
and R=4mm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
6.11 Comparison of the experimental and numerical normalized force-displacement
curves and evolution of plastic strain at the center of specimens for the grooved
at specimens. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.12 Evolution of the stress triaxiality with respect to plastic strain and weighted
stress triaxiality value for the grooved at specimens. . . . . . . . . . . . . . . 68
6.13 The weighted Lode angle parameter values and evolutions of the Lode angle
parameter with respect to plastic strain for the grooved at specimens. . . . . 69
6.14 Distribution of the plastic strain, stress triaxiality and Lode angle parameter
at fracture initiation deformation for the geometries with notch radii R=1mm
and 4mm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.15 Normalized force-displacement curves and evolution of the plastic strain at the
center of specimens for the axisymmetric notched round specimens. . . . . . . 71

xii

LIST OF FIGURES
6.16 Evolution of the stress triaxiality with respect to plastic strain and weighted
stress triaxiality values for the specimens with notch radii equal to 0.5mm,
1mm, 2mm and 4mm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
6.17 Histories of Lode angle parameter with respect to plastic strain and weighted
Lode angle parameter values for the axisymmetric specimens. . . . . . . . . . . 72
6.18 FE model (1/4) of Nakazima tests and a close view of the critical region. . . . 73
6.19 Plastic strain, stress triaxiality and Lode angle parameter distribution for the
Nakazima tests at fracture formation. . . . . . . . . . . . . . . . . . . . . . . . 75
6.20 Comparison of the force-displacement curves obtained from numerical simulations to the experimental ones and evolution of plastic strain at critical
location with respect to displacement for the Nakazima tests. . . . . . . . . . . 76
6.21 The weighted stress state values and evolution of the stress triaxiality and Lode
angle parameter with respect to the plastic strain for the Nakazima specimens
with blank width of 70mm and 90mm. . . . . . . . . . . . . . . . . . . . . . . 76
6.22 A view of nite element model discretization of the central region of the buttery specimen. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
6.23 Plastic strain distribution contours for the buttery specimens just before the
fracture initiation for all loading angles. . . . . . . . . . . . . . . . . . . . . . . 77
6.24 Stress triaxiality and plastic strain evolution on the surface and middle of
thickness of buttery specimens for the loading angles (a) 10 compression
and (b) 60 tension. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6.25 Normalized force-displacement curves and evolution of the plastic strain at
critical location for buttery test simulations. . . . . . . . . . . . . . . . . . . 80
6.26 Comparison of the deformation just before fracture initiation under loading
angle 0 shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
6.27 Evolution of the stress triaxiality with respect to plastic strain and weighted
stress triaxiality values for simulations of buttery tests. . . . . . . . . . . . . 81
6.28 Evolution of the Lode angle parameter with respect to plastic strain and
weighted Lode angle parameter values for simulations of buttery tests. . . . . 82
6.29 Representation of the histories and weighted stress state values obtained from
numerical simulations in the plane of stress triaxiality and Lode angle parameter for the buttery tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
6.30 Fracture strain with respect to weighted stress triaxiality and calibrated JC
equations for test sets of axisymmetric notched round specimens and grooved
at specimens. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
6.31 Fracture strain versus stress triaxiality for all tests. . . . . . . . . . . . . . . . 84

LIST OF FIGURES

xiii

6.32 Representation of the weighted stress state values in the plane of stress triaxiality and Lode angle parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . 84
6.33 Illustration of the possible additional test types in order to cover negative
stress triaxialities on the plane of stress triaxiality and Lode angle parameter.

85

7.1

Illustration of the generated analytical fracture surface. . . . . . . . . . . . . . 89

7.2

Comparison of the normalized force-displacement curves from numerical simulations and experiments for the notched specimens. . . . . . . . . . . . . . . . 90

7.3

Contour plot of the damage indicator D at fracture initiation for the at


specimens. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

7.4

Comparison of the normalized force-displacement response of grooved at specimens between the experiments and numerical simulations. . . . . . . . . . . . 92

7.5

Contour plot of the damage indicator D at fracture initiation for the grooved
at specimens. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

7.6

Comparison of the normalized force-displacement responses of axisymmetric


notched round specimens between the experiments and numerical simulations.

93

7.7

Contour plot of the damage indicator D at fracture initiation for the axisymmetric notched round specimens. . . . . . . . . . . . . . . . . . . . . . . . . . 93

7.8

A comparison of the force-displacement curves obtained from experiments and


numerical simulations of Nakazima tests. . . . . . . . . . . . . . . . . . . . . . 94

7.9

Contour plot of the damage indicator D at fracture initiation for the Nakazima
tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

7.10 Comparison of the post-mortem state of Nakazima tests between experiments


and numerical simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.11 Comparison of the experimental and numerical force-displacement responses
for the buttery tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
7.12 Damage indicator distribution at the gauge section of buttery tests. . . . . . 96
7.13 Comparison of the fractured specimens in numerical simulations and experiments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
7.14 Illustration of the generated mathematical fracture strain surface. . . . . . . . 99
7.15 Comparison of the normalized force-displacement curves from numerical simulations and experiments for the notched specimens. . . . . . . . . . . . . . . . 100
7.16 Comparison of the normalized force-displacement response of grooved at specimens between the experiments and numerical simulations. . . . . . . . . . . . 100
7.17 Comparison of the normalized force-displacement curves obtained from numerical simulations and experiments for the axisymmetric notched round specimens.101

xiv

LIST OF FIGURES
7.18 The force-displacement responses obtained from experiments and numerical
simulations with MFS and AFS for the Nakazima tests. . . . . . . . . . . . . . 101
7.19 The force-displacement response obtained from experiments and numerical
simulations with MFS and AFS for the buttery tests. . . . . . . . . . . . . . 102
7.20 Geometry, FE-model and boundary conditions of the torsion specimens. . . . . 103
7.21 Comparison of the max curves from experiments and numerical simulations
for the torsion specimens. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
A.1 Proposed fracture strain locus and bound curve formulations. . . . . . . . . . . 110

List of Tables
3.1

Johnson-Cook model fracture strain parameters for discussed steel grades and
aluminum alloys. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

5.1

The chemical composition of the dual phase steel DP600. . . . . . . . . . . . . 35

5.2

Calculated fracture strain from measured thicknesses for the Nakazima blanks
with width of 70mm and 90mm. . . . . . . . . . . . . . . . . . . . . . . . . . . 48

6.1

FE-models of 1/8 specimens for three dierent mesh size and corresponding
total number of elements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

6.2

Comparison of the numerical results of FE models with dierent discretization


at high deformation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

6.3

Calibrated parameters of Ludwik power law. . . . . . . . . . . . . . . . . . . . 58

6.4

Number of elements of 1/8 at-specimen FE models. . . . . . . . . . . . . . . 62

6.5

Fracture strain, weighted stress triaxiality and Lode angle parameter for at
un-notched and notched specimens. . . . . . . . . . . . . . . . . . . . . . . . . 65

6.6

Number of elements of the grooved at specimens ( 1/8 FE models). . . . . . . 66

6.7

Fracture strain, weighted stress triaxiality and Lode angle parameter for the
grooved at specimens. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

6.8

Number of elements of axisymmetric notched round specimens ( 1/8 FE models). 69

6.9

Fracture strain, weighted stress triaxiality and Lode angle parameter for axisymmetric notched round specimens. . . . . . . . . . . . . . . . . . . . . . . . 72

6.10 Comparison of the numerically determined fracture strains to the experimentally determined ones for the Nakazima tests with blank width of 70mm and
90mm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
6.11 Fracture initiation locations in the main plane and thickness directions of
buttery specimens for dierent loading angles. . . . . . . . . . . . . . . . . . 79
6.12 List of fracture strain and weighted stress state parameters for buttery tests.
7.1

80

List of fracture strain and weighted stress state parameters obtained from
numerical simulations of all tests. . . . . . . . . . . . . . . . . . . . . . . . . . 88

xvi

LIST OF TABLES
7.2

A list of calibrated parameters of the analytical fracture strain surface. . . . . 89

Nomenclature
Symbol

Meaning

a, b, c

Ludwik power law parameters.

g, h

weighting functions.

fading exponent of GISSMO damage model.

damage exponent.

vector of damage parameteres of fracture strain.

strain ratio in thickness and longitudinal direction.

A0

reference area perpendicular to vector


n.

A0

reference area in representative volume element.

(n)

Aef f

eective area perpendicular to vector


n.

Aef f

eective area in representative volume element.

+
0
C1,2
, C1,2
, C1,2

six fracure locus parameters proposed by Bai [8].

+
0
D1,2,3
, D1,2,3
, D1,2,3

nine parameters of proposed fracture locus.

damage.

Dc

critical damage.

D(n)

damage in the plane perpendicular to vector


n.

E0 , Eef f

Youngs modulus of undamaged and damaged material.

Fr

forming intensity in the GISSMO damage model.

force.

I1,2,3

rst, second and third stress invariants.

J1,2,3

rst, second and third deviatoric stress invariants.

L0 , L

initial gauge length and gauge displacement.

(n)

xviii

Nomenclature

Symbol

Meaning

moment.

relative ratio of principal stress deviators.

plastic strain tensor.

p,l

stress triaxiality dependent plastic strain at the begin of instability.

weighted stress triaxiality.

stress triaxiality.

alternative Lode angle denition.

Lode angle.

second denition of Lode angle parameter.

weighted Lode angle parameter.

Lode angle parameter.

engineering strain.

0 +

f , f , f

three bounding curves of the fracture stain surface.

f,m

the arithmetic mean of fracture strain.

stress state dependent fracture strain function.

equivalent plastic strain at fracture.

pi , pf

lower and upper bound (plastic strain) of the weigthing integral.

nk

equivalent plastic strain at the onset of necking.

px,y,z

maximum, intermediate, minimum principal plastic strains.

equivalent plastic strain.

true strain.

angle of rotation.

eective stress.

engineering stress.

eq

equivalent stress.

mean stress.

nk

equivalent stress at the onset of necking.

true stress.

Nomenclature

xix

Symbol

Meaning

, ij

stress tensor.

uts

ultimate tensile strength.

Cauchy stress.

max

maximum shear stress.

AFS

Analytical Fracture Strain surface.

AHSS

Advanced High Strength Steels.

CDM

Continuum Damage Mechanics.

DP600

Dual Phase steel 600.

GISSMO

Generalized Incremental Stress State dependant damage MOdel.

GTN

Gurson-Tvergaard-Needleman.

JC

Johnson-Cook.

MFS

Mathematical Fracture Strain surface.

RVE

Representative Volume Element.

SEM

Scanning Electron Microscopy.

xx

Nomenclature

Chapter 1
Introduction
1.1

Background and motivation

In recent years, the importance of numerical simulations in automotive industry increased signicantly, since strong competition leads to shorter development cycles. Numerical simulation
is a very ecient way to shorten the development cycle by reducing very costly experiments
in many aspects, such as crashworthiness, fatigue and noise vibration harshness analyses.
Recently, great eort has been spent in improving crashworthiness of the car body structures in order to comply with strict crash regulations. Structures can be optimized eectively
with crashworthiness simulations in means of passive safety, which is one the key subjects
in automotive industry. Understanding and modeling of fracture phenomena are vital for
accurate predictions in numerical simulations.
In academic world, fracture of materials is also of great interest and investigated deeply
in the range of nanoscale ( 1010 107 meters) to macroscale (> 103 meters) by many
researchers. On the other hand, in engineering community, due to the limitations of modern computer power and size of studied structures, the fracture phenomena is studied in
macroscale. Usually a systematic approach is followed by increasing the complexity gradually as depicted in Fig. 1.1. The characterization of material properties and validation of
numerical models are carried out with basic specimens. As next, the complexity is increased
and the prediction of the model is veried with component tests. Ultimate goal is the precise
prediction of deformation and fracture in the full car crashworthiness simulations. However,
numerical modeling of fracture may be a very challenging task and usually depends on many
inuence factors:
The proper plasticity model is crucial for damage modeling. Therefore many aspects
such as anisotropy, hardening type, hydrostatic pressure and Lode dependence of the
plasticity model should be considered. Which complexity is required to model the fracture of the investigated material and structure?

1 Introduction

Figure 1.1: Investigation of numrical modeling of fracture at dierent levels: Elementary


specimens (blue box), components (red box) and full car.
Commonly, loss of load carrying capacity is related to hydrostatic pressure dependence.
Should Lode dependence also be considered?
Damage accumulation rule is another important point, which determines the fracture
strain under complex loading. How should the hydrostatic pressure and Lode dependence be utilized in the damage accumulation rule?
The coupling of plasticity model and damage is still debatable. Does the coupling of
damage and plasticity model lead to better results and under which conditions should
be used?
The mentioned points determine the complexity of the numerical simulations used in crashworthiness simulations. The complexity of the modeling approach inuences directly the
calibration method and applicability to practical problems.

1.2

Thesis objectives

The objective of this thesis is the investigation of the stress state dependent damage modeling
with the focus on the Lode angle parameter (third deviatoric stress invariant). Two main
scientic contributions of the thesis are summarized as follows:
The formulation and application of the stress state dependent fracture strain denition.
The denition of the required tests and the corresponding calibration approach for the
proposed model.

1.2 Thesis objectives

The stress state dependent fracture strain formulation is integrated into the damage model
Generalized Incremental Stress State dependant damage MOdel (GISSMO), which was proposed by Neukamm et al. [14] to bridge the gap between metal forming and crashworthiness
simulations. Since the classical methods such as Forming Limit Diagram usually provide unsatisfactory results in metal forming simulations, the model is formulated as an incremental
continuum damage model which is based on the eective stress concept proposed by Lemaitre
[11]. Special emphasis is given to pre-damage and failure in the model. The model can be
used modular simultaneously with the anisotropic constitutive models in metal forming simulations and isotropic constitutive models in crashworthiness simulations also considering the
pre-damage.
The loading environment is room temperature and quasi-static loading. The material is
from the class of advanced high strength steel (AHSS) types Dual Phase steel 600 (DP600)
and the material is assumed to be isotropic.
The presented work includes numerical simulations and experimental study. A hybrid
approach, which combines the experimental and numerical results, is used in order to calibrate the damage model. Specimens are loaded until fracture in order to obtain the fracture
behavior under dierent stress states by varying specimen geometry and loading conditions.
The output of the experiments is force-displacement responses. The stress state and fracture
strain are determined with numerical simulations. In the current work, the material is assumed to fail at macroscopic crack formation (at orders of 0.1-0.2 mm). Thus in the current
research local stress and strain variables are used rather than far eld variables.
In order to realize the objective, the work is examined in subparts as follows:
Overview the literature on the subject ductile damage with the focus of stress state
parameters, stress triaxiality and Lode angle parameter.
Develop a fracture strain function, which depends on both stress state parameters.
Develop a nonlinear damage accumulation rule incorporating stress state.
Find appropriate specimen types, geometries and loading conditions covering wide
range of stress state for the validation.
Propose a calibration approach for the extended damage model.
Validate the proposed model and calibration approach with the numerical simulations
based on elementary specimen tests.

1.3

1 Introduction

Outline of the thesis

The present thesis consists of six chapters, an appendix and a bibliography of the references.
Each chapter handles a specic part of the research. The contents of the chapters are described
below.
Chapter 2 explains briey the required basics of ductile fracture such as stress state
parameters, stress triaxiality and Lode angle parameter which are the main focuses of the
current thesis. A brief review of damage modeling approaches is given under three groups,
micromechanical damage models, conventional continuum models with damage formulation
and continuum damage mechanics.
Chapter 3 gives a summary of literature about the stress state dependent damage plasticity and material ductility. The inuence of stress triaxiality and Lode angle parameter on
the damage plasticity and material ductility is the focus of the chapter. Experimental results
from literature for steel grades and aluminum alloys are presented and numerical approaches
incorporating stress state parameters are discussed.
Chapter 4 explains the damage model GISSMO briey. Stress triaxiality and Lode angle
parameter dependent nine parameter empirical fracture strain function is introduced. The
nonlinear damage rule used in modeling damage is discussed.
Chapter 5 is devoted to the experimental program, which consists of tests of notched and
un-notched at specimens, grooved at specimens, axisymmetric notched round specimens,
Nakazima and buttery specimens. The geometry of specimens and loading conditions are
introduced briey and results of tests are discussed.
Chapter 6 describes the numerical simulations without consideration of damage formulation. The focus of the chapter is the determination of the fracture strain and corresponding
stress state for each test mentioned in Chapter 5. The mesh size inuence is discussed with
numerical simulation results. Weighting functions for the stress state histories are introduced,
which are used in determination of the stress state parameters. Some methods and results
are briey reviewed in the context of damage exponent determination.
Chapter 7 introduces a methodology to generate the fracture locus with the results obtained in Chapter 6. Two dierent approaches are introduced as analytical fracture strain
denition and mathematical fracture strain denition. The rst one is based on the fracture
strain denition derived in Chapter 4, while the latter is a continuous smooth surface generated mathematically. The prediction of the fracture loci are investigated with numerical
simulations and the results are discussed.
Chapter 8 summarizes the contribution of the current thesis and suggests future research
topics.

Chapter 2
A Brief Review of Damage Modeling
Approaches
Abstract: In this chapter, the basics of ductile fracture is explained. The stress vector
is represented in Haigh-Westergaard space and in cylindrical coordinates. The representation is also given with three stress invariants. The stress state parameters, stress
triaxiality and Lode angle parameters are explained in detail. A brief overview of damage modeling approaches is given under three groups; micromechanical models, conventional continuum mechanics models with damage formulation and continuum damage
models. The advantages and disadvantages of the modeling approaches are discussed.

2.1

Introduction

In recent years, importance of prediction of damage is increased signicantly in many industrial applications. Numerical modeling of damage is of great interest, since the numerical
simulation tools are one of the most important links in product development cycles. In recent
four decades damage phenomenon is investigated in various aspects and many models have
been proposed to model damaging of materials.
The term ductile fracture is used ambiguously in literature mainly with two meanings.
In micromechanical investigations, it is mostly used as the fracture type resulting from nucleation, growth and coalescence of voids within the material. On the other hand, for large-scaled
applications it is referred to large deformation that materials exhibit before failure occurs.
In this thesis the term ductility is related to ductile fracture in a more specic sense and
it is used as equivalent plastic strain to fracture formation at the critical location, which can
be also called fracture strain by neglecting the elastic part for large deformations.
The investigations in the eld of strength of materials, which can be related in a way to
damage modeling of materials has started before almost 500 years with Leonardo da Vinci.
In his note Testing the Strength of Iron Wires of Various Lengths he investigated the limit

2 A Brief Review of Damage Modeling Approaches

loads for dierent iron wires experimentally [12]. Galileo carried out the rst tensile test
and concluded that strength of a bar is dependent on its cross-section and not its length
in his famous book Two New Sciences ([12]). The subject damage modeling has been
investigated by many researchers; however it is still an open subject.

2.2

Stress state characterization

The stress tensor ij of a material point has six components and can be represented in sixdimensional space. However, it is easier to deal with three principal components 1 , 2 , 3
and represent the stress state in Haigh-Westergaard space (Fig. 2.1) in which the coordinates
are three principal stresses. In Haigh-Westergaard space, a specic stress state is represented
2

z
P
N

plane

|NK| =

2
3 eq

K
|ON| =
O

3m

r
1

Figure 2.1: Representation of the stress vector in Haigh-Westergaard space (1 , 2 , 3 ) and


cylindrical coordinates (r, , z).
with a specic point. The depiction shows mainly the form for a combination of principal
stress components and does not consider the orientation of the principal stress components.

The stress vector OP can be also represented in cylindrical coordinate system (r, , z), as
shown in Fig. 2.1. The z axis is called hydrostatic axis, where all the principal stresses are
equal. The plane, which is passing through the origin O and is perpendicular to z axis, is

called the plane. In cylindrical coordinate system the stress vector OP can be decomposed

into two components as hydrostatic part ON and deviatoric part N P . The vector component

ON is perpendicular to octahedral plane and the vector N P is in the octahedral plane. The

magnitude of the vector ON is linearly related to hydrostatic stress

|ON | = 3m ,
(2.1)

2.2 Stress state characterization

where m is the mean stress and dened through


1
m = (1 + 2 + 3 ).
3

(2.2)

The magnitude of the vector N P is linearly related to equivalent stress

2
|N P | =
eq ,
3

(2.3)

where eq is the von Mises equivalent stress and dened through


1
eq =
(1 2 )2 + (1 3 )2 + (2 3 )2 .
2

(2.4)

The third coordinate dierentiates the stress state between tension, shear and compression.
In literature the Lode angle inuence is usually used as Lode dependence and for the Lode
angle there are dierent denitions. In [1315] the Lode angle is dened as the angle to the
positive direction of the 1 = 3 axis, i.e. a in Fig. 2.2. On the other hand, in [1619]
the Lode angle is represented as the positive angle to the positive direction of 1 axis, in
Fig. 2.2. In the present thesis, the latter is used. The Lode angle can be dened through the
|NL| =

2
2 = 1

|NM| =

2 = 3

|MP| =

2 = 3

3
2 s1

3
2 s2
3
2 s3

N 1 = 2
M
N

3 = 1

P
1 = 3

a
L

1
1 = 2

3 = 2
1 = 3

Figure 2.2: Geometrical representation of the principal stresses (1 , 2 , 3 ), deviatoric principal stresses (s1 , s2 , s3 ) on the octahedral plane.
relative ratio of principal stress deviators through
[
)]
(
2
1 1
1

= cot

,
3 2

(2.5)

2 A Brief Review of Damage Modeling Approaches

where the relative ratio of principal stress deviators is dened as


=

s2 s3
2 3
=
.
s1 s3
1 3

(2.6)

The alternative Lode angle denition a can be also expressed as a relation of relative ratio
of stress deviators

[
a = tan

]
1
(2 1) .
3

(2.7)

The stress state denition in cylindrical coordinate system can be also expressed with
stress invariants uniquely. Above discussed three cylindrical dimensions m , eq and are
described through
1
m = I1 ,
(2.8)
3
eq =

1
= arccos
3

3J2 ,

(
)
3 3J3
,
2J2 3/2

(2.9)

(2.10)

where I1 is the rst stress invariant, J2 and J3 are second and third deviatoric stress invariants.
It should be noted that with three stress invariants the stress state at a material point can
be dened uniquely.
It is convenient to work with the dimensionless pressure , which is dened as the ratio
of hydrostatic pressure to equivalent stress
=

m
.
eq

(2.11)

The parameter , often referred to stress triaxiality, has been used extensively in the ductile
damage investigations ([2025]). The Lode angle can be related to the normalized third
deviatoric invariant through

27 J3
3 3 J3
=
(2.12)
=
= cos(3).
2 eq 3
2 J2 3/2
A detailed derivation is presented in [17]. Hereinafter the normalized third deviatoric invariant
will be called Lode angle parameter, as it is a function of . The range of the Lode angle
parameter is 0 1, since the range of the Lode angle is 0 /3. It can be
showed that = 1 corresponds to axisymmetric tension, = 0 corresponds to generalized
shear condition (plane strain). The lower limit value = 1 corresponds to axisymmetric
compression or equi-biaxial tension [7]. Two dimensionless stress state parameters and
dene the direction of the stress vector in Haigh-Westergaard space.

2.3 Damage modeling approaches

In some investigations [26, 27] Lode angle parameter is dened alternatively as


=1

6
2
= 1 arccos .

(2.13)

It should be noted the two Lode angle parameter denitions and have the same values for
the limiting bounds axisymmetric deviatoric tension ( = 0), plane strain or generalized shear
( = /6) and axisymmetric deviatoric compression ( = /3), whereas for the midrange the
denitions dier slightly.
In order to dene the magnitude of the stress vector, additionally, equivalent stress eq
is required. It should be noted that stress triaxiality is a relation of stress invariants I1
and J2 , whereas the Lode angle parameter is formulated with stress invariants J2 and J3 .
Under proportional loading, which can be also described by constant stress vector direction
in Haigh-Westergaard space, the stress triaxiality and the Lode angle parameter remain
constant during loading.
Wierzbicki and Xue ([7]) showed that for plane stress state (3 = 0) there is a unique
relation between the stress triaxiality and Lode angle parameter as
(
)
27
1
2
=
.
(2.14)
2
3
It should be noted that in industrial applications the plane stress formulation is widely used,
which requires only one of two stress parameters (, ) to determine the stress state at a
material point. The focus of the thesis is 3-dimensional formulation and two stress parameters
and determines the stress state. The former is used to describe pressure dependence and
the letter describes Lode dependence.

2.3

Damage modeling approaches

Since early 1970s, there are many attempts to model the phenomenon of ductile fracture.
Mainly two dierent approaches were used for damage modeling in engineering materials. The
rst one is the microscopic approach, in which the evolution of microscopic variables such as
voids and cracks are examined. The second approach is the macroscopic approach in which
material is represented by its global response. In both groups, various numerical models are
suggested. These models can be discussed more specically in three groups: Micromechanical
models (I) and two groups of macroscopic models, uncoupled continuum models (II) and
continuum damage models (CDM) (III).

2.3.1

Micromechanical models

The microstructure of engineering alloys is usually very complex and contains inclusions and
second phase particles. From micromechanical point of view, ductile fracture is dened as

10

2 A Brief Review of Damage Modeling Approaches

material separation which is the result of void nucleation, evolution of existing micro voids
and cracks, followed by progressive void coalescence. Voids are rst initiated at material
defects such as inclusions. It should be also noted that in many engineering alloys voids
usually preexist. Mainly, void coalescence is the result of the failure of ligament between the
voids (a) perpendicular to loading direction or (b) in the localized shear direction [28]. In
the rst case the inuence of secondary voids is not signicant, coalescence of voids occurs
with necking of ligaments between primary voids. This type of failure occurs at high stress
triaxialities and the tensile mode causes at dimpled fracture morphology (see [29]-(a)). In
the second case, secondary voids initiate on the slip-band and cause ow localization which
is followed by void coalescence [29] as demonstrated in Fig. 2.3-(b).
(a)

(b)

20m

20m

Figure 2.3: Illustration of two fracture mechanisms: (a) Flat dimple fracture where fracture
occurs at high stress triaxialities as a result of void ligament necking followed by failure
and (b) shear dimple fracture where fracture occurs with deformation localization between
primary voids at low triaxialities [30].
In micromechanical models, material is handled as inhomogeneous cells with consideration of voids and micro cracks. Macroscopic material response is determined by the global
response of void containing cell. In these models the holes surrounding material matrix obeys
conventional continuum mechanics. The micromechanical modeling can be divided into three
subgroups as void nucleation modeling, void growth modeling and void coalescence modeling.
The void nucleation modeling studies are much limited compare to void growth and
void coalescence models. The reason of the limited studies is the diculty of the subject,
which requires the modeling of inclusions, interfaces between inclusions and material matrix.
The inclusions are assumed as elastic and brittle particles and void nucleation occurs by
fracture or debonding of inclusions. Beremin [31] calculated the maximum principal stress
within the inclusions using the results of Berveiller and Zaoui [32]. A most widely used model
was proposed by Needleman [33] by using cohesive model between inclusions and material
matrix, where the interfacial strengths are relatively weak. The inclusion fracture, which is

2.3 Damage modeling approaches

11

also related to void nucleation, was studied by Steglich and Brocks [34] and Shabrov et al.
[35].
The void nucleation stage is followed by void growth. A cylindrical void and a spherical
void containing material was studied by McClintock [20] and Rice and Tracy [21], respectively.
In these investigations the void surrounding material matrix was assumed as perfectly plastic
and analytical expressions dening void growth are derived. In both studies the eect of
stress triaxiality and plastic strain on the void growth is outlined. Rice and Tracy found that
the void growth is exponentially dependent on the stress triaxiality. In fact the most known
and widely used micromechanical model is Gurson-Tvergaard-Needleman (GTN), which is an
extended version of the model proposed by Gurson [36]. Gurson derived a pressure dependent
yield function from an isolated spherical void in a continuum. The void containing solid is
considered as pressure dependent, dilatant elasto-plastic continuum. The global response of
the solid is correlated to void volume fraction. The yield function of Gurson model is dened
as

eq 2
=
+ 2f cosh
M 2

3m
2M

[
]
1 + f 2 = 0,

(2.15)

where f is the void volume fraction, M is the yield stress of the undamaged matrix material,
eq and m are the macroscopic equivalent stress and the mean stress, respectively. The
matrix yield stress M is a function of matrix plastic strain P . With the assumption of
associative ow rule, the plastic ow function is the same as the yield function. The yield
criterion dened by Gurson implies that the macroscopic plastic strain rate can be found by
the normality rule

p =
,
(2.16)

where is the plastic multiplier, p and are plastic strain tensor and stress tensor, respec represents the time derivative. The rate of work equivalence implies
tively and ()
: p = (1 f )M P .

(2.17)

As above mentioned, the void coalescence occurs through internal ligament necking or
through deformation localization on the shear bands between voids. The rst mechanism was
investigated by Thomason [37]. Benzerga [38] introduced a micromechanical model for void
coalescence with consideration of internal ligament necking. The model is an enhanced version
of the model proposed by Thomason [37]. The second mechanism was studied by Brown and
Embury [39]. They concluded that the coalescence occurs when the spaces between voids
reaches to the order of void height.
The yield function proposed by Gurson predicts the complete loss of load carrying capacity
at f = 1, which means the material is only made of voids and it is unrealistically larger
than experimental results. Tvergaard investigated micromechanical behavior for materials

12

2 A Brief Review of Damage Modeling Approaches

containing periodic distributions of voids and introduced additional two parameters q1 and
q2 in order to improve the yield function proposed by Gurson (Eq. 2.15) [40, 41]. Tvergaard
and Needleman [42] introduced the function f , in order to improve the load carrying capacity
loss (void coalescence) limit,
{
f,
f (f ) =
fc +

1/q1 fc
(f
ff fc

for f fc
,
fc ) for fc f ff

(2.18)

where fc and ff are critical void volume fraction and void volume fraction at coalescence,
respectively. At small deformations the total void volume fraction evolution is described as
the sum of evolution of existing voids (f)growth and nucleation of new voids (f)nucleation
f = (f)nucleation + (f)growth .

(2.19)

Chu and Needleman [43] introduced the strain controlled void nucleation evolution term in a
statistical manner. A normal distribution with respect to matrix plastic strain is proposed
fnucleation = AN M ,
where the normal distribution function AN is dened as
(
{
})
fN
1 P N
exp
AN =
.
2
SN
SN 2

(2.20)

(2.21)

Here, fN is the total void volume fraction that can be nucleated, N and SN are the mean
value and standard deviation of the distribution of the plastic strain.
Although the GTN model is formulated to describe all three stages in metal failure, void
nucleation, growth and coalescence, it possesses some inherent drawbacks:
1. The model is inapplicable to model the localization and fracture for low stress triaxiality, shear dominated deformations, since it does not predict void growth and damage
under shear loading, when fN < ff . However, in engineering applications such as in
crashworthiness simulations, the stress state on the fractured components is complex
and fracture at low stress triaxialities is also observed.
2. The model does not include the Lode dependence.
3. The geometrical shape and shape of voids are considered in a simplied way. It is
assumed that the voids conserve their spherical shape also after deformation. Under
loading with high stress triaxialities the assumption may lead to correct results. On
the other hand at low stress triaxialities the shape of voids tends to be elliptical after
deformation.

2.3 Damage modeling approaches

13

4. The calibration of micromechanical parameters is very cumbersome and done usually


in a stochastic way with reverse engineering, since experimentally determination of
parameters is very complicated
5. Only void volume fraction is used as a measure of damage. Defects such as volumless
cracks are not considered, which cause also loss of load carrying capacity.
Because of above pointed out drawbacks, the GTN model was extended by many researchers.
Xue [44] and Nahshon and Hutchinson [45] modied the model in a phenomenological way
to account for shear deformation and for the Lode dependence. Gologanu and coworkers
[46] enhanced the model with consideration of axisymmetric elliptical voids. The enhanced
model of Gologanu is further modied by Pardoen and Hutchinson [28] with consideration
of void spacing eects. A comprehensive review of GTN model and its extension was given
by Lassance et al. [47].

2.3.2

Conventional continuum mechanics models

The deformation is a complex phenomenon at micro scale. Many ductile materials at macroscale
can be assumed as a continuum. In conventional continuum mechanics models with damage
formulation, macroscopic response is a combination of material matrix, defects like inclusions, voids and micro cracks and globally it is assumed to obey continuum mechanics laws.
The damage indicator is calculated separately and does not inuence the material plasticity. Because of the relative simplicity, in industrial applications the models are widely used.
The most widely used model is proposed by Johnson and Cook [24], which incorporates
phenomenologically the strain rate and temperature factors in the stress-strain function,
[
( )] [
(
)q ]
[
]
p
T T0
N
eq = A + Bf
1 + C ln
1
,
(2.22)
0
Tm T0
where A, B, C and N and q are ve model parameters, T , T0 , Tm are temperature, room
temperature and melting temperature, respectively and 0 is the reference strain rate. Under
proportional loadings the fracture strain f is dened as function of stress triaxiality , plastic
strain rate and temperature
[
]
(
)
T T0
D3
f () = D1 + D2 e
(2.23)
[1 + D4 ln(p /0 )] 1 + D5
Tm T0
where D1 , D2 , D3 , D4 , D5 are ve material constants. In this group of damage models, Lode
dependence was introduced rstly by Wilkins et al. [48] in a cumulative damage formulation. J2 plasticity is used as constitutive plasticity model. The proposed damage formulation
incorporates pressure and Lode dependence as
f (m , Aw ) = Dc (1 aw m ) (2 Aw )w ,

(2.24)

14

2 A Brief Review of Damage Modeling Approaches

where Dc , aw , , w are the model parameters to be calibrated and Aw is the stress eccentricity
and denotes the Lode dependence as
{
}
s2 s2
Aw = max
,
.
(2.25)
s1 s3
Recently Bai and Wierzbicki [8] introduced a stress triaxiality and Lode dependent plasticity
model and cumulative damage model.

2.3.3

Continuum damage mechanics

The Continuum Damage Mechanics (CDM) was proposed rstly by Kachanov [49] in the
framework of creep damage. It is further extended by Rabotnov [50] and Lemaitre [11].
The voids, micro cracks and their interactions are depicted in a phenomenological way. The
model is based on the eective stress concept, which is related to the load carrying area after
removing damaged part such as voids and micro cracks. Eective stress is dened as
=

,
1D

(2.26)

where is the Cauchy stress. A macroscopic response of the material is used in constitutive
and damage models. Bonora [51] dened the dissipation potential Fr as the sum of plastic
potential Fp and damage potential Fd , which are functions of associated variables
Fdp = Fp (, Rs , D) + Fd (Y , p , D),

(2.27)

where Rs is the isotropic hardening stress, Y is the damage energy release rate. The elastic
strain increment is dened as
eij =

1 + ij
kk

ij ,
E 1D E1D

(2.28)

where E and are Youngs modulus and Poissons ratio, respectively. The plastic strain
increment is dened by
pij =

Fp
3 s ij 1
=
.
ij
2 1 D eq

(2.29)

= Fd .
D
Y

(2.30)

Damage evolution is dened through

The weakening is also introduced into conventional plasticity models. Brvik et al. [52] introduced weakening factor to Johnson-Cook (JC) model
(
( )] [
)q ]
[
[
]
p
T T0
N
1
eq = (1 D) A + Bf
,
(2.31)
1 + C ln
0
Tm T0

2.3 Damage modeling approaches

15

where is the model constant in the rst part. The rest of the equation is same as in Eq. 2.22.
In a similar way Xue and Wierzbicki [53] introduced a weakening factor to von Mises plasticity
as
(
)
eq = 1 Dx M ,
(2.32)
where x is the weakening factor and M is the yield stress of the the material matrix.

16

2 A Brief Review of Damage Modeling Approaches

Chapter 3
Stress State Dependence of Damage
Modeling
Abstract: In this chapter, the stress state dependence of damage plasticity and material
ductility is investigated through numerical and experimental studies from the literature.
The experimental results of round bar tests under tension with dierent notch radii
for steel grades and aluminum alloys are discussed to point out the eect of stress
triaxiality (or hydrostatic pressure). The inuence of the Lode angle parameter (or
Lode dependence) is investigated by comparing the experimental results obtained from
tests with axisymmetric tension ( = 1) and plane strain ( = 0) stress state. The
literature overview concludes that stress state dependence of plastic ow and material
ductility varies depending on the material group. A literature overview of stress state
dependent plasticity and material ductility models is given with discussions.

3.1

Plasticity model

The fracture formation is an ultimate process, which follows large plastic deformation at a
material point with high gradients of plastic stress and strain. Therefore the used plasticity
model is very important in ductile fracture modeling.
In the past three centuries various yield criteria are proposed. Trescas well-known maximum shear stress criterion is actually introduced by Coulomb and it is one of the two most
commonly used criteria today. The second most common used yield criterion is the von Mises
criterion. In fact, it is rstly suggested by Maxwell in 1856. It should be noted that von Mises
yield surface only depends on the second stress invariant J2 ; while Tresca yield surface has
dependence on second deviatoric stress invariant J2 and the third deviatoric stress invariant
J3 , which can be also concluded as Lode dependence.
The above described two criteria have no hydrostatic pressure dependence, which means
under hydrostatic loading material never yields. However, some material can yield at su-

18

3 Stress State Dependence of Damage Modeling

ciently high hydrostatic tension. Mohr-Coulomb [54] model and Drucker-Prager [55] criteria
are the examples of the rst proposed models incorporating hydrostatic pressure inuence.
The former is the generalized version of the Tresca criterion, while the latter is the modied
version of von Mises criterion (Fig. 3.1). Hydrostatic pressure generally determines the size
(a) von Mises and Drucker-Prager

(b) Tresca and Mohr-Coulomb

Mohr-Coulomb

Drucker-Prager

Tresca

von Mises
O

Figure 3.1: Representation of yield surfaces (a) Lode angle independent: von Mises (hydrostatic pressure independent) and Drucker-Prager(hydrostatic pressure dependent) (b) Lode
angle dependent: Tresca (hydrostatic pressure independent) and Mohr-Coulomb (hydrostatic
pressure dependent).
of yield surface, whereas the Lode angle controls the shape of the yield surface. In the 20th
century various plasticity models incorporating hydrostatic pressure and Lode dependence
are proposed basically for the usage of rock and soil mechanics [5658]. On the other hand,
the inuence of hydrostatic pressure and Lode angle on the metal yielding has neglected.
The yield criteria used also as failure criteria by some researchers. For perfect plastic
materials, yielding also implies failure, since the stress level is constant between initial yielding
and fracture points. In this thesis yielding is used as begin of plastic ow, while failure is
used as the ultimate loss of material strength resulting from damage accumulation.
Even though most materials can be treated as isotropic approximately, strictly speaking,
all materials are anisotropic to some extend; that is, the material properties are not the
same in every direction. Hill proposed an orthotropic yield surface, which has three mutually
orthogonal planes and is an extension of von Mises isotropic plasticity [59]. Recently Barlat
et al. [60] presented two 3D yield functions with 13 and 18 parameters, respectively. There
are an abundance of various anisotropic models for plane stress (e.g. [6062]), but the subject
anisotropy is out of the scope of the thesis.
There are various experimental investigations, which showed that material can exhibit
complex plastic hardening behavior under reversal loading. In order to model the hardening

3.1 Plasticity model

19

behavior, kinematic hardening models [63,64] and combined kinematic-isotropic [65,66] hardening models were proposed. Since neglecting kinematic hardening does not induce signicant
errors under monotonic loadings, the subject will not be discussed further.
In some investigations, the dierence between plastic ow under uniaxial tension and
uniaxial compression is explained as the hydrostatic pressure dierence. In fact the deviatoric
stress state for these two loading cases is not same. The Lode angle parameter diers for
uniaxial tension as = 1 and uniaxial compression as = 1. In order to investigate the
inuence of hydrostatic pressure, the deviatoric stress state should be kept constant. Tensile
tests of axisymmetric round bars or plane strain specimens with dierent notches can be used
for that purpose.

3.1.1

Experiments on stress triaxiality dependence

The experimental investigations for the hydrostatic pressure (or in this thesis used form stress
triaxiality) inuence on the yield function can be tracked back to Bridgmans investigations.
Bridgman exposed axisymmetric bars to hydrostatic pressure in a pressure chamber [67],
which was later also used in pressure eect investigations by many researchers [6870]. Superimposed hydrostatic pressure does not change the deviatoric state parameter which is
a relationship between J2 and J3 (2.2). Bridgman tested the axisymmetric specimens under
conning pressures up to 3000Mpa, which are rarely seen in industrial structures. He concluded that hydrostatic pressure has no inuence on the plastic ow. Brownrigg et al. [71] and
Spitzig and Richmond [72] investigated the hydrostatic pressure eect on material yielding,
respectively for 1045 steel and aluminum grade 1100. Brownrigg et al. also concluded a weak
hydrostatic pressure dependence on the material yielding.
Axisymmetric specimens with dierent notch radii have been also used in the investigations of hydrostatic pressure inuence. Wilson studied 2024-T351 aluminum notched
round bars with dierent notches under tension and obtained good numerical results with
hydrostatic pressure dependent Prager-Drucker model [73]. On the other hand, for the same
aluminum alloy 2024-T351, Bao and Wierzbicki [25, 74] observed almost perfect correlation
between experiments and numerical simulations with pressure independent von Mises plasticity (J2 -plasticity). Xue also concluded that hydrostatic pressure inuence of the yielding
is debatable for the same aluminum alloy, 2024-T351 [14]. Experiments of 1045 steel round
bars with dierent notches were carried out by Bai et. al and good results are obtained for
all tests with J2 -plasticity , which indicates no pressure dependence of plasticity model [75].
Brvik et al. [76] and later Teng and Wierzbicki [77] performed numerical simulations of
smooth and notched round bars designed by Brvik et al. [78] for Weldox 460 E steel and
obtained good accordance between numerical and experimental load-displacement responses
with a pressure insensitive plasticity model. Barsoum conducted smooth and notched round

20

3 Stress State Dependence of Damage Modeling

bar tests for the medium-strength steel Weldox 460 and high-strength steel Weldox 960 and
it was observed that the pressure dependence on the initial yield strength is not signicant
[79]. Copolla et al. also conrmed for 6 dierent steel grades that plastic ow does not depend
on the hydrostatic pressure [80].

3.1.2

Experiments on Lode angle parameter dependence

The inuence of the Lode angle parameter can be investigated by keeping the stress triaxiality value constant. However usually the stress state variables are changing on the loading
path, which makes the separation of the inuence of two stress state variables a challenging
task. The inuence can be extracted from axisymmetric and plane strain specimens with
same stress triaxiality values. For both mentioned specimen types the Lode angle parameter remains constant on loading path as = 1 and = 0, respectively. A literature review
shows that experimental investigations on the inuence of Lode angle parameter on material
yielding are scarce. Bai et al. proposed a pressure and Lode dependent yield function [75].
It has been showed that the plasticity model parameters obtained from 2024-T351 axisymmetric specimens require up to 20% deviatoric state (Lode angle dependence) correction,
while the required correction for hydrostatic pressure is relatively small. Seidt [5] investigated the same aluminum alloy 2024-T351 widely. He compared the equivalent stress-strain
curves, which were extracted from axisymmetric tension, axisymmetric compression and torsion tests. Especially for the torsion loading case a signicant dierence (20%) was found
(Fig. 3.2-a), which also particularly conrms the results of Bai et al. . The curves extracted
from torsion and axisymmetric tests for the 5083 aluminum are presented in Fig. 3.2-b. It
(b) 5083 aluminum
700

600

600

Equivalent stress (Mpa)

Equivalent stress (Mpa)

(a) 2024-T351 aluminum


700

500
400
300
Tension
Torsion
Compression

200
100

500
400
300
200

Tension
Torsion

100

0
0

0.1

0.2

0.3

Equivalent plastic strain

0.4

0.4

0.8

1.2

Equivalent plastic strain

Figure 3.2: Comparison of equivalent stress-strain curves for (a) 2024-T351 aluminum extracted from tension, torsion and compression experiments [5] and (b) 5083 aluminum extracted from torsion and compression experiments [6].

3.2 Material ductility

21

can be argued that for the presented cases, the parameter stress triaxiality is also dierent
and can cause the mentioned dierence. The authors showed that the stress triaxiality has
no inuence on the yielding by comparing axisymmetric tests with dierent notches.
Recently the Lode dependence on the plasticity properties was also conrmed for cast
aluminum alloy [81, 82]. On other hand the investigations of Bai [8] showed that for the
steel 1045 the Lode dependence on material plasticity is not signicant. It was shown that
with J2 -plasticity it is possible to obtain good results for both axisymmetric and plane-strain
specimens.
As it is presented above, the Lode (or Lode angle parameter) and pressure (or stress
triaxiality) dependence of plastic ow is material dependent. Especially a correction of Lode
dependence is necessary for the discussed aluminum alloys. For the discussed steel grades,
pressure and Lode insensitive J2 -plasticity can be used without introducing signicant errors. By considering the literature survey and the investigated material DP600, numerical
simulations are run with J2 -plasticity in the current thesis.

3.2

Material ductility

As already stated, in this thesis the term ductility refers equivalent plastic strain at fracture.
Recently the stress triaxiality and Lode angle parameter dependence of material ductility has
been investigated by several researchers [8385]. In these investigations the inuence of the
two quantities on the material ductility is assumed to be separable as
f = g1 ()g2 (),

(3.1)

where g1 and g2 are the functions dependent on the stress triaxiality and Lode angle parameter, respectively. Xue dened the fracture strain as a product of reference fracture strain and
two functions g1 and g2 [83]. Bai [19] dened the fracture strain as a surface over the stress
triaxiality and Lode parameter. In this thesis the stress triaxiality and Lode angle parameter
eects on ductility are assumed to be separable.

3.2.1

Stress triaxiality dependence

The design of constant pressure experiments posses challenges. However, it is possible to


elucidate the existence and the general trend of fracture strain on the pressure with specic
type of experimental setups. The axisymmetric specimens with imposed hydrostatic pressure
in a high pressure vessel were tested by various researchers starting with the pioneering work
by Bridgman [67]. In 1970 to 1990 extensive experiments were conducted by many researchers
[70, 8690]. Most of these works reviewed by Lewandowski and Lowhaphandu [68]. In these

22

3 Stress State Dependence of Damage Modeling

works it has been generally observed that imposing larger hydrostatic pressures cause also
an increase in the fracture strain.
As in the plasticity studies, the pressure eect on material ductility is also investigated by
testing notched specimens under atmospheric pressure. Similar to adding conning pressure
on a round bar, the lateral tension is added by creating a circumferential notch around the
circular specimens. Brvik et al. [52,78] conducted axisymmetric notched tensile experiments
for Weldox 420E Steel and found out that fracture strain increases in an exponential manner
as the stress triaxiality increases. Barsoum et al. [29, 91] approved the exponential relation
for the Weldox steel 420 and 920. The exponential relationship between fracture strain and
triaxiality is also found for DH36 steel [8,15], 1045 steel [8] and for various steel grades [9294].
Various aluminum alloys were also investigated with axisymmetric notched specimens and
the similar tendency of monotonic increase in hydrostatic pressure has been observed for
2024-T351 [19, 95], 2021 [93] and 5083 [6] aluminum alloys.
For both methods, axisymmetric specimens under dierent superimposed hydrostatic
pressures and axisymmetric notched specimens under atmospheric pressure, the Lode angle parameter remains constant ( = 1) during loading. Similar trends for fracture strain
have been also observed for the plane strain tensile specimens with dierent notches [75, 96],
where the Lode angle parameter remains as = 0.
Rice and Tracy [21] proposed a stress triaxiality fracture strain, which has been derived
by analyzing void growth under hydrostatic loading
C2

f () = C1 e

(3.2)

where C1 and C2 are material constants. As discussed in previous chapter, JC fracture strain
equation (Eq. 2.23) is a similar formulation with an additional parameter when the strain
rate and temperature terms are neglected
(
)
f () = D1 + D2 eD3 .

(3.3)

In fact most of the above discussed experimental results can be described by the fracture strain
denition of Rice and Tracy or Johnson-Cook. In Fig. 3.3, the fracture strain is shown as a
function of the average stress triaxiality through loading for the above discussed experiments.
Calibrated JC fracture strain-average stress triaxiality curves are also depicted. The list of
the calibrated JC parameters is given in Table 3.1.

3.2.2

Lode angle parameter dependence

The determination of the Lode angle parameter inuence on the fracture strain is not trivial,
since the stress triaxiality value is usually not constant through loading. However, comparison
of tted stress triaxiality vs. fracture strain curves for axisymmetric tensile specimens ( = 1)

3.2 Material ductility

23

(a) Steel grades

(b) Aluminum alloys

1
Fitted curves
Experimental data

1.5
Weldox-420
DH36

Experimental data

0.8
Fracture strain

Fracture strain

Fitted curves

FE370
0.5

6061-T6
0.6
2021
0.4
2024-T351
0.2

5083-H116

25MnCr6
0

0
0.5

0.8

1.1

1.4

1.7

0.3

0.5

Average stress triaxiality

0.7

0.9

1.1

1.3

Average stress triaxiality

Figure 3.3: Average stress triaxiality vs. fracture strain for a) steels: Weldox-420 [29], DH36
[75], FE370 [97] and 25MnCr6 [80] b) aluminum alloys 2024-T351 [95], 5083-H116 [6], 6061T6 [94] and 2021 [93].
Table 3.1: Johnson-Cook model fracture strain parameters for discussed steel grades and
aluminum alloys.
Material

D1

D2

D3

-0.066

3.861

1.499

DH36 [75]

6.365

2.404

FE370 [97]

2.276

1.234

25MnCr6 [80]

3.839

2.097

2024-T351 [95]

6.859

2.451

5083-H116 [6]

0.640

1.448

6061-T6 [94]

2.604

2.966

2021 [93]

1.280

1.950

Weldox-420 [29]

and plane strain specimens ( = 0) can be compared to investigate the Lode dependence on
the material ductility.
Bai and Wierzbicki [75] examined DH36 and 1045 steel for axisymmetric specimens with
dierent notch radii and grooved at specimens with dierent groove radii. It has been
observed that Lode angle parameter inuence on fracture strain is material dependent. 1045
steel showed Lode dependence on the material ductility, whereas the inuence is scarce for
the DH36 steel (Fig. 3.4-a). The same situation has been also conrmed for aluminum alloys.

24

3 Stress State Dependence of Damage Modeling

Strong Lode dependence of material ductility is observed for aluminum alloy 2024-T351 [5].
On the other hand the inuence for the aluminum alloy is less signicant for 5083-H116 [6].
The inuence of the Lode angle parameter on material ductility varies through dierent stress
(a) Steel grades

(b) Aluminum alloys


0.8

Fracture strain

1.5

= 1 - Fitted curves
= 0 - Fitted curves
= 1 - Experimental data
= 0 - Experimental data

0.6
Fracture strain

= 1 - Fitted curves
= 0 - Fitted curves
= 1 - Experimental data
= 0 - Experimental data

5083-H116

0.4

1
DH36
1045

2024-T351

0.2

0.5

0
0.2

0.4

0.6

0.8

1.2

0.2

Avarage stress triaxiality

0.4

0.6

0.8

1.2

Avarage stress triaxiality

Figure 3.4: Comparison of average stress triaxiality-fracture strain curves for axisymmetric
tension = 1 and plane strain = 0 for steel grades and aluminum alloys. The tted curves
(solid curves for = 1 and dashed curves for = 1) and experimental average values (circles
for = 1 and squares for = 0) are plotted for a) steels: DH36 [75] and 1045 [75] b) for
aluminum alloys 5083-H116 [6] and 2024-T351 [5].
triaxiality ranges. Generally, the eect becomes weak for high stress triaxialities.
Fracture locus maps based on bounding curves has been proposed by several researchers.

0
The three bounding curves, +
f , f and f are the stress triaxiality dependent curves with
constant Lode angle parameter values, which correspond to axisymmetric deviatoric tension
= 1, plane strain or generalized shear = 0 and deviatoric compression = 1, respectively
(Fig. 3.5). Xue and Wierzbicki [7] extended the Wilkins model and Rice-Tracys criterion and

proposed a symmetric (+
f = f ) fracture locus depending on Lode angle parameter and stress
triaxiality

[
] 2
0
0
f (, ) = +
f f f ,

(3.4)

0
where +
f and f are stress triaxiality dependent Rice-Tracy criteria. The asymmetry of

the fracture has been investigated by various researchers. Using experimental data on three

geological materials, Chen and Hahn [98] showed that +


f > f . Bardet [99] studied the

experiments carried out by Nakai et al. [100] and also proposed +


f > f .
Theoretical analysis based on unit cell by Zhang et. al [101], Gao and Kim [102], showed
that for constant stress triaxiality values, fracture strain for axisymmetric biaxial tension is

3.2 Material ductility

25
Fracture strain

+
f
0f

Stress triaxiality
-1

0
Lode angle parameter ( or )

Figure 3.5: Bounding curves of the fracture locus proposed by Xue and Wierzbicki [7], f (,
) and Bai and Wierzbicki [8], f (, ).
+
higher than the axisymmetric uniaxial tension (
f > f ). Wan et al. [103] investigated the
+
growth of void in the unit cell and concluded also the same (
f > f ).
+
Bai and Wierzbicki [8] proposed an asymmetric fracture strain surface
f = f , which has
a parabolic behavior in the Lode angle parameter direction and Rice-Tracy type exponential

equation in the stress triaxiality direction.


[
]
1 +
1

0
0
f (, ) =
(f + f ) f 2 + (+

f ) + f
2
2 f
[
]
)
1( +
+

0
0
=
C exp(C2 ) + C1 exp(C2 ) C1 exp(C2 ) 2
2 1
)
1( +
()
+
C1 exp(C2+ ) C1 exp(C2 ) + C10 exp(C20 ),
2

(3.5)

where C1+ , C2+ , C10 , C20 , C1 and C2 are six fracture locus parameters. It should be noted that
Bai and Wierzbicki [8] used the second denition of Lode angle parameter .

26

3 Stress State Dependence of Damage Modeling

Chapter 4
The GISSMO Damage Model
Abstract: The phenomenological damage model GISSMO was proposed by Neukamm
et al. [14] in order to bridge the gap between metal forming and crashworthiness
simulations. The extension and the used features in the current thesis are discussed
in detail. The term damage is explained and experimental damage quantication
methods are discussed briey. The derivation of the nonlinear damage accumulation rule
is presented in detail. A stress triaxiality and Lode angle parameter dependent material
ductility (fracture strain) function is proposed. The other features of the damage model
such as material instability modeling and coupling of the damage with plasticity model
are discussed briey.

4.1

Introduction

The mechanical properties of material in crashworthiness simulations are assumed to be as


in material delivery state. The car body metal sheet parts are deformed in a metal forming
process before they are used as car components. Metal forming causes local deformations
on the metal sheet, which can change the material properties locally. Neglecting the predeformation can lead to wrong results in crashworthiness simulations.
In metal forming simulations, the classical models such Forming Limit Diagrams usually
predict relatively bad results. Therefore the damage model GISSMO is formulated as an
incremental continuum damage model, which is based on eective stress concept (see Eq. 2.26)
proposed by Lemaitre [11] in post-critical range of deformation.
In crashworthiness simulations, usually isotropic models such as J2 -plasticity are used.
On the other hand in metal forming more sophisticated anisotropic yield loci are considered.
Thus it is necessary to use dierent constitutive laws in the process chain. The damage model
GISSMO is used both in metal forming and crashworthiness simulations simultaneously with
constitutive plasticity models.
The integration of the damage model GISSMO into the process chain consisting of metal

28

4 The GISSMO Damage Model

forming and crashworthiness numerical simulations is illustrated in Fig. 4.1. The model is

b) Mapping

a) Metal forming simulation


p , t
Material model

, p , q

c) Crashworthiness simulation
0p , t0

Material model

, p , q

D
D, q

GISSMO

0
D 0 , q

GISSMO

Figure 4.1: Integration of the damage model GISSMO to the process chain consisting of metal
forming and crashworthiness simulations [2].

modularly integrated into material models in both stages. In the metal forming simulation
stage, the distribution of the plastic strain p , thickness t, damage D and internal variables
q distribution are calculated. In the intermediate step, the mentioned variables are mapped
from ne mesh to relatively coarse mesh, which is commonly used in crash simulations. In
the third stage the mapped variables are used as initial values for the crash simulation. The
damage model can be used as damage indicator uncoupled with a constitutive model as in
continuum material models with damage formulation discussed in subsection 2.3.2 or can
be coupled to the constitutive plastic models as continuum damage mechanics discussed in
subsection 2.3.3.
In the current thesis, the pre-deformation is not considered, since all the specimens are
machined from the sheet metal plate. Therefore the concerned part is only the crashworthiness
numerical simulations. As already discussed in Chapter 3, the plastic ow properties dier
for dierent type of metals. For example, in general the aluminum alloys have Lode angle
dependent plastic ow, whereas the plastic ow is not stress state dependent for the steels.
Furthermore some steels and aluminum alloys have signicant anisotropic plastic ow. Thus,
dierent complexity of constitutive model is required depending on the investigated material.
The damage model GISSMO can be used simultaneously with dierent constitutive models.
Therefore, in this thesis formulated stress triaxiality and Lode angle parameter dependent
fracture surface is integrated to the damage model GISSMO in order to use the denition
with dierent type of constitutive models.

4.2 Damage rule

4.2
4.2.1

29

Damage rule
Damage variables

The word damage is used as the relative measure of change in physical properties of partial
damaged material compared to the undamaged material. These physical properties include
elastic modulus, remaining ductility, hardness and local mass density.
From micromechanical point of view, ductile fracture is a process, which follows nucleation, growth and coalescence of micro voids and cracks. In micromechanical models damage
is usually dened as the void fraction in a Representative Volume Element (RVE). At a
specic void fraction, which is material dependent, the material is assumed to fail.
The damage variable was rstly introduced for creep by Kachanov [49], and later by
Rabotnov [50] in the framework of continuum damage mechanics. From the physical point
of view, damage can be dened as the reduction of the nominal section area in the RVE
which is the result of micro cracks and micro voids. Murakami and Ohno [104] were the rst

introducing the plane with normal vector


n , cutting RVE as represented in Fig. 4.2.
(a) Damaged configuration

(b) Effective undamaged configuration


F

Removing voids
and cracks

Ae

A0

F
Ad

Figure 4.2: Representative volume element subjected to uniaxial tension. (a) Volume with
damage and (b) eective load carrying material volume.

The damage D(n) on the plane with the normal vector


n is dened through
(n)

D
(n)

(n)

=1

Aef f
(n)

(4.1)

A0

(n)

where A0 is the reference area of cross-section and Aef f is eective resisting area after subtracting the micro cracks and micro voids from reference area. It should be noted that in this

case the damage denition depends on the direction of normal vector


n . Therefore according
to denition the damage should be dened as a tensor. Chaboche [105] introduced a fourthorder damage tensor and second order damage tensors were introduced by Kachanov [106],
Krajcinovic [107], Chow and Wang [108], Murakami [109], Hammi et al. [110] and Br
unig

30

4 The GISSMO Damage Model

[111]. However in these models, the identication of parameters becomes a very complicated
and challenging task with limited published examples in the literature [112,113], which makes
the models for industrial applications not practical.
By assuming the damage as isotropic, Eq. 4.1 can be written as
D =1

Aef f
,
A0

(4.2)

where D is isotropic scalar damage, Aef f /A0 is the ratio independent of chosen orientation.
In many industrial applications the damage indicator has been modeled as a scalar value and
good results have been obtained. [114116]. Therefore the damage is represented by a scalar
quantity in the present thesis.

4.2.2

Experimental damage quantification

The experimental determination of the damage possesses some diculties and usually it is
modeled as an internal variable. Lemaitre and Chaboche [117] proposed Youngs modulus
reduction as a measure of damage,
D =1

Eef f
,
E0

(4.3)

where E0 and Eef f are Youngs modulus of undamaged and damaged materials, respectively.
Alves et al. [118] and Bonora et al. [119] used the proposed method for dierent stress
triaxialities. It should be noted that, with this method an average value through the critical
cross-section of the specimen is obtained. However for materials with high ductility, plastic
strain and damage localization occur especially at the center of critical cross-sections of
axisymmetric round specimens. Thus the mentioned model can lead to slightly conservative
results for the materials with high ductility. X-ray micro-tomography was used by Weck et
al. [120] and Tasan [10] to observe void (damage volume fraction) evolution. Other damage
quantication methods, micro indentation method based on hardness measurements was
examined by Tasan et al. [121] and brittle-fracture methodology was proposed by Hoefnagels
[122].

4.2.3

Damage accumulation rule

In recent years various damage accumulation rules have been proposed. These criteria were
basically empirical and based on observation and experience. The cumulative strain damage models assume that fracture occurs due to the plastic deformation history. For a given
material, damage occurs, when the damage indicator integral reaches the critical value Dc

h(eld variables)dp = Dc ,
(4.4)

4.2 Damage rule

31

where h is a weighting function, which is usually a function of stress invariants or stress


components ij , plastic strain p , plastic strain rate p , temperature T and current damage
D. Normalization both sides of the integral lead to

g(eld variables)dp = 1.
(4.5)
The simplest and most commonly used damage potential is the linear function which is
dened through
p
D= .
(4.6)
f
For a given fracture strain, the damage increment is dened as
dD =

1
dp ,
f

(4.7)

which was also used by many researchers [24, 123125]. In this case the damage increment
is linear with respect to the plastic strain increment. However, some researchers showed
that damage increment and plastic strain increment do not have necessarily linear relation.
Bonora [51], Bai et al. [126] and Xue [126] proposed nonlinear damage accumulation rules.
Recently Weck et al. [127, 128] investigated metallic sheets containing laser drilled holes.
It has been concluded that the void growth rate is nonlinear with respect to local strain.
Tasan [10] measured the void growth at dierent plastic strain levels with Scanning Electron
Microscopy (SEM) and concluded a nonlinear void growth with respect to plastic strain. In
micromechanical models such as proposed by Gurson [36] and Barsoum and Falskog [91], the
void volume fraction, which can be also considered as damage, increases nonlinearly with
respect to increasing plastic strain.
In the GISSMO damage model, the damage potential under proportional loading is dened
as a power function
( )n
p
D=
,
(4.8)
f
where n is the damage exponent, which denes the nonlinearity level of the damage accumulation. Corresponding damage increment can be dened as
dD =

n (n1)
p
dp = g(f , p , n)dp ,
f n

(4.9)

where g is a weighting function. The fracture strain f in the formulation is stress state
dependent. More precisely f is a function of stress state parameters, the stress triaxiality
and Lode angle parameter . A similar damage evolution equation was used by Xue [129]. It
should be noted that the damage increment depends on the current p . In micromechanical
models such as Gurson model, void volume fraction evolution is calculated from the current

32

4 The GISSMO Damage Model

volume fracture. By substituting Eq. 4.8 into Eq. 4.9, one gets the damage evolution increment
used in the GISSMO damage model
n1
n
D n dp = g(f , D, n)dp ,
f (, )

dD =

(4.10)

where in the weighting function g instead of current p , current D is used. From this point
of view, the phenomenological damage evolution rule proposed in GISSMO damage model is
motivated by micromechanical models. Under proportional loading conditions both formulations lead to identical results, whereas the damage prediction of the formulations diers
under non-proportional loading through loading path. Under proportional loading the damage increment is integrated as

D=

g(f , D, n)dp = 1.

(4.11)

The initial and end conditions for the integral in Eq. 4.11 correspond to D = 0 at p = 0 and
D = 1 at p = f . Depending on the chosen damage exponent n, the functional g represents
an innite number of solutions for the integral. Linear damage accumulation of JC n = 1,
nonlinear damage accumulation with n = 2 and evolution of normalized void volume fraction
in GTN with respect to plastic strain are compared in Fig. 4.3. Under complex loading
1
Johnson-Cook n=1
GISSMO n=2
Gurson

D or f /ff

0.8
0.6
0.4
0.2
0

0.2

0.4

0.6

0.8

p /f

Figure 4.3: Evolution comparison of linear damage (n = 1), nonlinear damage (n = 2) and
normalized void volume fraction of GTN model with respect to plastic strain.
histories (non-proportional loading) the damage is accumulated incrementally and material
fails when integral reaches unity

D=

g(f , D, n)dp = 1.
0

(4.12)

4.3 Fracture strain denition

4.3

33

Fracture strain definition

A similar but extended fracture locus proposed by Bai and Wierzbicki (Eq. 3.5) is introduced
in dependence of and . Three bound curves (Fig. 4.4) axisymmetric tension = 1, plane
strain or generalized shear = 0 and axisymmetric deviatoric compression = 1 are dened
as reduced form of JC equation (Eq. 3.3). The dependence of the Lode angle parameter is

+
+
+
+
f = D1 + D2 exp(D3 )

Fracture strain
0f = D10 + D20 exp(D30 )

f = D1 + D2 exp(D3 )

-1

Figure 4.4: Proposed fracture locus and formulations for bound limit curves.
2

described with a second order polynomial a + b + c functional. The functional coecients


a, b and c are found with respect to three limiting curves and the 9 parameter asymmetric
fracture locus f is introduced.
[

]
)
)
1(
1(
+
0

+
+
0
0
f (, ) =
D1 + D1 D1 +
D2 exp(D3 ) + D2 exp(D3 ) D2 exp(D3 ) 2
2
2
[
]
) 1( +
)
1( +

+
D1 D1 +
D2 exp(D3 ) D2 exp(D3 )
2
2
+ D10 + D20 exp(D30 ),
(4.13)
where D1 , D2 , D3 , D10 , D20 , D30 , D1+ , D2+ and D3+ are nine parameters of the fracture locus.
Many special cases can be recovered from the proposed fracture locus. By omitting D1 , D10 ,
D1+ terms the function becomes the Eq. 3.5 proposed by Bai and Wierzbicki. By requiring

the symmetry conditions (+


f = f ) additional to omitted terms, the function becomes the
Eq. 3.4 proposed by Xue and Wierzbicki. By neglecting the Lode angle parameter inuence
( = 0), JC equation (Eq. 3.3) is obtained.

34

4.4

4 The GISSMO Damage Model

Instability and coupling of damage with plasticity

The modeling of local material stability is of great interest in metal forming simulations.
Although in literature some methods [130] are proposed to model the instability of material,
it is dicult to acquire the required stress and strain rates in experiments [3]. In damage
model GISSMO the instability variable Fr is determined using an evolution equation for the
forming intensity in a cumulative way
dFr =

n1
n
Fr n dp ,
p,l ()

(4.14)

where p,l is the stress triaxiality dependent equivalent plastic strain at the begin of instability and n is the exponent, which is also used in damage accumulation. In crashworthiness
simulations, usage of limited mesh size requires regularization of dissipated energy to the
crack formation. The regularization issue is treated with coupling the damage with stress
tensor and dening element size dependent fracture strain. Damage D and stress tensor are
coupled through
{
(
(
)m ) for D Dc ;
=
(4.15)
c
1 DD
for D > Dc ,
1Dc
where Dc is the critical damage, which is discussed in detail in [131] and m is the fading
exponent, which controls the rate of stress reduction on the element.

Chapter 5
Experimental Program
Abstract: The present chapter discusses the experimental program presented in the
current thesis. Flat un-notched and notched specimens with two notch radii, grooved
at specimens with four groove radii, axisymmetric notched round specimens with four
notch radii, buttery tests with ve dierent loading angles and Nakazima tests with two
dierent blank geometries are investigated. Geometry of specimens, loading conditions
and experimental force-displacement results are presented with discussions. In-plane
anisotropy in two orthogonal directions is briey discussed based on experimental results
of un-notched specimens.

5.1

Introduction

Recently, Advanced High Strength Steels (AHSS) are used widely in automotive industry
since they show good strength, formability and weldability behavior. Also their cost is lower
than equivalent heat-treated alloys since the desired characteristics are achieved directly from
hot rolling. In Fig. 5.1, AHSS are compared with the low strength and HSS steels. In this
thesis, DP600 from AHSS family is investigated. DP-steels are low-carbon steels and contain
large amount of manganese and silicon and small amount of microalloying. The chemical
composition of the investigated material DP600 is given in Table 5.1.
Table 5.1: The chemical composition of the dual phase steel DP600.

Weight %

Si

Mn

Al

0.14

0.50

2.0

0.07

0.015

0.015

0.005

The experimental program consists of sets of un-notched and notched at specimens,


axisymmetric notched round specimens, at grooved specimens, Nakazima tests and Buttery
tests. All specimens were machined from the DP600 sheet metal plates with thickness of

36

5 Experimental Program
70

Total Elongation (%)

60
50

IF

40

MILD IFHS
BH

30

CM

20

TRI
P

DP

- CP

HSLA

10

MS

0
0

200

500

800

1100

1400

1700

Tensile Strength (Mpa)

Figure 5.1: Comparison of AHSS (transformation-induced plasticity (TRIP) steels, dual phase
(DP) steels and martensitic (MS) steels) with low strength and HSS steels [132].
2mm, which were provided by the company Alcan. The tests of at specimens, axisymmetric
notched round specimens, at grooved specimens were conducted in Fraunhofer Institute
for Mechanics of Materials IWM, the Nakazima tests in Department of Ferrous Metallurgy
IEHK-RWTH and buttery tests in Institute of General Mechanics IAM-RWTH.
The experiments were conducted at room temperature and type of loading is quasi-static.
All specimens are machined from the rolling direction in order to avoid anisotropy eects
(Fig. 5.2).
(a) Flat and axisymmetric round specimens

(b) Grooved flat specimens

Flat specimens
Axisymmetric notched round specimens
Rolling direction

Figure 5.2: Layout of machined specimens on the DP600 sheet plate (a) at and axisymmetric
notched round specimens (b) grooved at specimens.
As explained in previous chapters, material ductility depends on the stress state, which can

5.2 Un-notched and notched at specimens

37

be experimentally acquired with dierent specimen types and loading conditions. Usually the
parameters of the numerical damage models can not be determined from experiments directly,
which necessitates a hybrid methodology combining numerical simulations and experiments.
Each specimen or loading condition represents a specic stress state, which is dened with
two stress state parameters, stress triaxiality and Lode angle parameter in this thesis.
The covered stress states by the investigated sets of tests are depicted in Fig. 5.3.
Lode angle
parameter
1

tter
fly

Flat specimens

Bu

0.5

tes

ts

Plane stress

Axisymmetric notched
round specimens

0
-1

-0.5

0.5

Grooved flat
specimens
1

Stress triaxiality
1.5

Plane strain
-0.5

-1

Nakazima
tests

Figure 5.3: Illustration of covered stress state by the investigated sets of specimens and tests.

5.2

Un-notched and notched flat specimens

Un-notched and two notched specimens with Radius (R)=2mm and 4mm are machined from
the rolling direction of the sheet metal. The specimens were loaded with 0.01mm/s machine
speed to complete separation. For each geometry three duplicate tests were conducted. The
un-notched specimens are used to extract the material properties under uniaxial stress state,
whereas the two notched specimens are used to cover higher stress triaxiality values. The
geometry and dimensions used in this set of study are shown in Fig. 5.4.
The displacement was measured directly on the specimens by a gauge and the gauge
length L0 for the un-notched and notched at specimens is 10mm and 30mm, respectively.
The experimental results for the un-notched and notched specimens are presented in
Fig. 5.5. The results are presented as normalized force and displacement curves; the recorded
force F is normalized with the undeformed critical cross-section perpendicular to loading
direction A0 and the measured gauge displacement L is normalized with initial gauge
length L0 .

38

5 Experimental Program
Notched R=2mm

20

L0 =30

R2
80

L0 =30

60

L0 =10

Lc =14

18

2
2

R4
80

4.1

4.1

4.1

Notched R=4mm
9

Un-notched

18

R5

10
10

10

20

20

Figure 5.4: The un-notched and notched at specimens. From left to right: un-notched, notch
radii R=2mm and 4mm.
(b) Notched: R=2mm and R=4mm
0.8

0.6

0.6
F/A0 (Gpa)

F/A0 (Gpa)

(a) Un-notched
0.8

0.4

0.2

Noched R=2mm
Noched R=4mm

0.4

0.2

0
0

0.1

0.2

0.3
L/L0

0.4

0.5

0.01

0.02

0.03
0.04
L/L0

0.05

0.06

Figure 5.5: Normalized force-displacement curves obtained from the tests of (a) un-notched
at specimens and (b) notched at specimen (R=4mm and 2mm) tests.
It is observed that normalized force-displacement curves for the un-notched and notched
specimens do not vary to maximum force. The notched specimens with R=4mm and R=2mm
also do not show scatter in the normalized force-displacement levels after maximum points to
complete separation. However, the normalized displacement at fracture varies for the notched
specimens (downward arrows in Fig. 5.5-(b)). The observed maximum dierence is 5%, which
is observed for tests of the notched specimen with R=4mm.
Ductile materials usually form a localized, inhomogeneous high strain location before
fracture [133]. The fracture occurs in the localized zone, since the order of plastic strain
is signicantly higher than the rest of the specimen. Depending on the specimen geometry
and material properties, a second shear-band deformation can be also formed, which leads

5.2 Un-notched and notched at specimens

39

the second localized zone [134, 135]. The un-notched at specimens show a signicant diuse
necking zone but not a second localized shear band localization (Fig. 5.6). The specimens after

Figure 5.6: Fracture surface of the un-notched at specimen.


complete separation are shown in Fig. 5.7. The SEM micrograph of microstructure is shown
Un-notched

Notched R=2mm

Notched R=4mm

Figure 5.7: The post-mortem of un-notched at specimens and notched at specimens with
notch radii R=2mm and 4mm.
in Fig. 5.8-(a). The white and dark zones are ferritic and martensitic zones, respectively. The
fracture surface shows a typical ductile fracture with high stress triaxiality, where dimples
are observed Fig. 5.8-(b). The average size of the voids in the fracture surface is 5 m.

5.2.1

Anisotropy of the material

The previous investigations for DP600 steel sheets with thickness (t) of 1.5mm showed that
in-plane anisotropy is not signicant for the investigated material [136]. The comparison of the
plastic strains in the thickness and width direction can be used to determine the anisotropy

40

5 Experimental Program

(a)

(b)

Figure 5.8: (a) SEM micrograph of the microstructure of undeformed DP600 steel (ferrite
phase is white and martensite phase is dark) and (b) fracture surface of the un-notched at
specimen.
between compared two directions. For this purpose typically used r value is dened through
r=

ln(w/w0 )
,
ln(t/t0 )

(5.1)

where t0 is the thickness before deformation and t is the deformed thickness of the at unnotched specimen. w0 and w are the width of specimen before and after the deformation,
respectively Fig. 5.9. However, due to the relative small dimension of the sheet metal thickness

on

cti

t0

ire
gd
n
i
ad
Lo

w
w0

Figure 5.9: Representation of the thickness and width of the specimen before and after the
deformation.
and sensitivity of r value to measurement errors, the thickness terms in the Eq. 5.1 are
substituted according to the postulate of constant volume in plasticity theory,
( )
ln ww0
(5.2)
r = ( L0 w 0 ) .
ln Lw
The r value for three un-notched at specimens is recorded for technical strains at 5%,
10%, 15%, 20% and after complete separation. The evolution of the r value with respect to

5.3 Grooved at specimens

41

normalized global displacement is demonstrated with normalized force-displacement curves in


Fig. 5.10. On the onset of necking the value r increases slightly and average value of r=0.92 is

F/A0

0.8

0.8

0.6

0.6
Test 1 F/A0
Test 2 F/A0
Test 3 F/A0
Test 1 r value
Test 2 r value
Test 3 r value

0.4

0.2

0.4

0.2

0
0

0.1

0.2

r value

0.3

0.4

0
0.5

L/L0

Figure 5.10: Evolutions of the r value and normalized force with respect to global normalized
displacement for three duplicate tests of un-notched at specimen.
recorded at fracture points, which indicates that the anisotropy between longitudinal direction
and thickness direction is not signicant.

5.3

Grooved flat specimens

The double side grooved at specimens were tested to investigate the material plasticity
and ductility under plane strain conditions, which corresponds to the Lode angle parameter
= 0, where the principal strain component in width direction is negligibly small compared
to the other two principal strain components. By introducing dierent groove radii, a range
of stress triaxiality at constant Lode angle parameter = 0 can be investigated (Fig. 5.3).
The specimens were cut from the rolling direction of the plates (Fig. 5.2-(b)) and loaded with
machine speed 0.01mm/s until fracture.
In order to ensure the plain strain conditions, the specimens must be dimensioned properly, since the dimensions thickness t, width w and minimum thickness at the center of groove
tm (see Fig. 5.11) have signicant inuence on the stress state. Bai and Wierzbicki [8] assigned
the ratios w/tm = 32.25 and t/tm = 3.125, Benzerga [38] applied the ratios w/tm = 16.6 and
t/tm = 3.
The geometry of the grooved at specimens was optimized with numerical simulations.
Numerical simulations for the specimen with groove radius R=2mm were carried out for
dierent ratios of w/tm and t/tm . Minimum thickness at the groove center was kept constant

42

5 Experimental Program

w
tm

Figure 5.11: Critical dimensions for the grooved at specimens.


tm =0.8mm and other two dimensions w and t were varied. The material point at the center of
the specimen was investigated. Inuence of dierent ratios of w/tm and t/tm on the evolution
of Lode angle parameter with respect to equivalent plastic strain is demonstrated in Fig. 5.12.
It is observed that, as expected, the w/tm ratio aects the evolution signicantly and for ratio

Lode angle parameter

1
w/tm
w/tm
w/tm
w/tm
w/tm
w/tm

0.8

0.6

= 6.25 and t/tm


= 6.25 and t/tm
= 12.5 and t/tm
= 12.5 and t/tm
= 25.0 and t/tm
= 25.0 and t/tm

= 2.5
= 5.0
= 2.5
= 5.0
= 2.5
= 5.0

0.4

0.2

0.2

0.4

0.6

0.8

Plastic strain

Figure 5.12: Evolution of the Lode angle parameter with respect to equivalent plastic strain
with constant tm =0.8mm for dierent ratios of w/tm and t/tm .
values greater than 12.5, plane strain conditions are achieved. By considering the machining
abilities and the thickness of metal sheet plates, the ratios of w/tm = 12.5 and w/tm = 2.5
were assigned to the specimens in this line of study. In order to investigate a range of stress
triaxiality under plane strain conditions, tests of specimens with grove radii equal to 0.5mm,
1mm, 2mm and 4mm were conducted Fig. 5.13.
The displacement was recorded by a clip gauge with a length of 20mm in all tests of
grooved at specimens. The normalized force-displacement curves for the grooved at specimens with groove radii R=0.5mm, 1mm, 2mm and 4mm are plotted together in Fig. 5.14
with two duplicate tests for each specimen geometry. The scattering of experimental results
concerning material plasticity is negligibly small. All duplicate curves are almost identical in terms of normalized force-displacement level. The fracture points are indicated with
downward arrows in Fig. 5.14. The maximum scattering of material ductility, which can be

5.3 Grooved at specimens

43

34

R0.5

30.5

10

R1

R2

0.8

0.8

R0.5

R1

R4
0.8

0.8

R2

R4

110

.1

22

9.5

15

19

30.5

9.5

L0 =20

Lc =25

R12

Figure 5.13: The geometry and dimensions of the grooved at specimens with groove radii
equal to 0.5mm, 1mm, 2mm and 4mm.

(b) R=2mm and R=4mm


1

0.8

0.8
F/A0 (Gpa)

F/A0 (Gpa)

(a) R=0.5mm and R=1mm


1

0.6
0.4

R=0.5mm Test 1
R=0.5mm Test 2
R=1.0mm Test 1
R=1.0mm Test 2

0.2
0

0.005

0.01
L/L0

0.6
0.4

R=2.0mm Test 1
R=2.0mm Test 2
R=4.0mm Test 1
R=4.0mm Test 2

0.2

0.015

0.02

0.005

0.01

0.015 0.02
L/L0

0.025

0.03

Figure 5.14: The normalized force-displacement curves obtained from tests of grooved at
specimens with groove radii (a) R=0.5mm and R=1mm and (b) R=2mm and R=4mm.

described as the end of force-displacement curves, was observed for the specimen geometry
with groove radius R=1mm (%8 in terms of global normalized displacements between two
duplicate tests).
The post-mortem of grooved at specimens are shown in Fig. 5.15.

44

5 Experimental Program
R=0.5mm

R=1mm

R=2mm

R=4mm

Figure 5.15: The grooved at specimens with groove radii R=0.5mm, 1mm, 2mm and 4mm
after fracture.

5.4

Axisymmetric notched round specimens

The tests of axisymmetric notched round specimens were carried out to characterize material
properties under axial symmetric stress state, which corresponds to Lode angle parameter
=1 (Fig. 5.3). By the variation of notch radius, it is possible to investigate the inuence of
the stress triaxiality for a specic range. The minimum value of the stress triaxiality that
can be obtained with this group of specimens is = 1/3, which is the stress state of the unnotched round specimen before the onset of necking. Decreasing the notch radius increases
the stress triaxiality value.
In this study, experiments of axisymmetric notched round specimens with notch radii
R = 0.5mm, 1mm, 2mm and 4mm were conducted. The geometry and dimensions of the
specimens are illustrated in Fig. 5.16.
The loading speed was set to 0.01mm/s in order to assure quasi-static loading conditions.
The displacement is measured with a gauge with length of 10mm. For each notch geometry two duplicate tests with the same loading conditions were conducted. The normalized
force-displacement curves are plotted in Fig. 5.17. The scatter of force-displacement levels is
negligible. The maximum dierence in material ductility is observed for the specimen with
smallest notch radius R=0.5mm as 13% in normalized displacement at fracture.
The post-mortem specimens are shown in Fig. 5.18.

5.4 Axisymmetric notched round specimens

45

R=0.5mm

R=1mm

R=2mm

R=4mm

Figure 5.16: The geometry and dimensions of the axisymmetric notched round specimens
with notch radii R=0.5mm, 1mm, 2mm and 4mm.
(b) R=2mm and R=4mm

0.8

0.8

0.6

0.6

0.4

F/A0 (Gpa)

F/A0 (Gpa)

(a) R=0.5mm and R=1mm

R=0.5mm Test 1
R=0.5mm Test 2
R=1.0mm Test 1
R=1.0mm Test 2

0.2

0.4

R=2.0mm Test 1
R=2.0mm Test 2
R=4.0mm Test 1
R=4.0mm Test 2

0.2

0
0

0.01

0.02

0.03

0.04

0.02

L/L0

0.04

0.06

L/L0

Figure 5.17: Normalized force-displacement curves obtained from the tests of axisymmetric
notched round specimens for notch radii (a) R=0.5mm and 1mm (b) R=2mm and 4mm.
R=0.5mm

R=1mm

R=2mm

R=4mm

Figure 5.18: The post-mortem of axisymmetric notched round specimens with notch radii
R=0.5mm, 1mm, 2mm and 4mm.

46

5.5

5 Experimental Program

Nakazima tests

Nakazima tests are commonly used to assess deformation under biaxial conditions in sheet
metals. The load is applied by a hemisphere punch on the metal blanks to fracture. A lubricant
between punch and blank (sheet metal) is used in order to overcome possible friction eects.
By changing the geometry of metal blank, a range of stress state between uniaxial tension
(=1/3, =1) and equi-biaxial tension (=2/3, =-1) can be covered [137].
For the current thesis, Nakazima tests with two dierent blank geometries were carried
out. The dimensions of the two blank geometries, with width of 90mm and 70mm are demonstrated in Fig. 5.19. In the blank with initial width of 90mm the stress state is equi-biaxial,
whereas the blank with width of 70mm has a slightly dierent stress state.
(a) 90mm

(b) 70mm
Punch
Punch

Blank
holder

8.75

12.5
12.5

Punch

8.75

R15
R15

Blank

70

87.5

90

107.5

Die

115
thickness =2

115

Figure 5.19: The geometry and dimensions of the Nakazima test blank geometries with initial
width of (a) 70mm and (b) 90mm.
For each blank geometry three duplicate tests were run. The punch force-displacement
curves are shown in Fig. 5.20. Since in Nakazima tests force does not ow through a section
on the blank, the force-displacement curves are not normalized as for the other specimen sets.
The duplicate tests show negligible scatter in force-displacement levels. The displacement of
punch at fracture diers between duplicate tests for the blank geometry with width of 70mm
and 90mm 5% and 4%, respectively.
The post-mortem of Nakazima blanks are illustrated in Fig. 5.21. It should be noted
that the fracture initiates in the middle and propagates through plane of symmetry. Under
equi-biaxial tension, the equivalent plastic strain after fracture can be calculated with the
assumptions of incompressibility and constant strain distribution through the thickness of
the specimens. The plastic strain in thickness direction pz is calculated through
pz

( )
t
= ln
,
t0

(5.3)

5.5 Nakazima tests

47
(a) Width=70mm

(b) Width=90mm

120

120
100

Test 1
Test 2
Test 3

80

Force (kN)

Force (kN)

100

60
40
20
0

Test 1
Test 2
Test 3

80
60
40
20

10
15
20
Displacement (mm)

25

10
15
20
Displacement (mm)

25

Figure 5.20: Punch force-displacement curves obtained from Nakazima tests with blank width
of (a) 90mm and (b) 70mm.
Width=70mm

Width=90mm

Fracture

Fracture

Figure 5.21: The post-mortem of Nakazima test blanks with width of 70mm and 90mm.

where t is the thickness after fracture and t0 is the thickness at undeformed state. Assumptions
of constant volume and similarity of other two in-plane principal plastic components (px =
px = pz /2) lead to
[
]
2
p =
(px )2 + (py )2 + (pz )2 = pz .
3

(5.4)

After fracture, the specimens were cut through the plane of symmetry, where the fracture
initiated and propagated. The thickness t is measured in the middle of the blank where the
fracture assumed to initiate and initial thickness t0 is the maximum measured thickness on
the cut specimens. The calculated fracture strains f and arithmetic mean of fracture strains
f,m for two blank geometries are presented in Table 5.2.

48

5 Experimental Program

Table 5.2: Calculated fracture strain from measured thicknesses for the Nakazima blanks with
width of 70mm and 90mm.
Geometry and Test

5.6

f = ln

( )

70mm-Test1

0.516

70mm-Test2

0.703

70mm-Test3

0.493

90mm-Test1

0.787

90mm-Test2

0.970

90mm-Test3

0.872

t
t0

f,m

0.570

0.876

Butterfly tests

The last of experiments are buttery tests, which cover a wide range of the stress triaxiality
and Lode angle parameter with a single specimen geometry (Fig. 5.3). The unique design
of the buttery specimen leads to deformation concentration at the center of the specimen
under dierent loading directions, which determines the fracture initiation location in the
specimen.
Mohr and Treitler [138] used buttery shaped specimens in order to investigate fracture
strain in the range 0-0.6 of stress triaxiality. Bai and Wierzbicki [8] calibrated the stress
triaxiality and Lode angle parameter dependent fracture locus for high strength steel
A710 with a single buttery specimen geometry loaded under dierent angles.
The buttery specimen investigated in this thesis, was optimized with numerical simulations before running the tests by considering also the maximum sheet metal thickness and
machining abilities. The geometry features two curvatures in order to force the strain localization to occur in the middle of specimen, where the thickness jump is maximum between
the gauge section and the shoulders (Fig. 5.22).
The detailed buttery specimen geometry and dimensions are shown in Fig. 5.23.
The buttery specimens assessed in the current study were machined and tested in Institute of General Mechanics (IAM) of RWTH Aachen University. The experiments were
carried out in a custom-made testing machine similar to that used by Mohr and Treitler
[138]. The test bench with subparts is illustrated in Fig. 5.24. The moving parts (2b, 3b, 4b
and 5 in Fig. 5.24) are connected to two guides (1), which move vertically. The loading angle
is adjusted by rotating the two inner discs (3a and 3b). A homogeneous displacement eld
on the shoulders of buttery specimen is obtained by xing the specimen to the inner discs

5.6 Buttery tests

49

Shoulder

Shoulder

Figure 5.22: Illustration of the buttery specimen.


50
17

12.75

8.1

12.65

5
R1.2

R4.

10.25

R18

.5

32

R6.

37.5

3.5

25

.93
R52

1
2

Figure 5.23: The geometry and dimensions of the buttery specimen.


with two grips (4a and 4b). The force is applied vertically by the actuator (5). In order to
eliminate dynamic eects, the tests were run at a very low speed, v=0.01mm/s. The test sets,
at un-notched and notched, at grooved, axisymmetric notched round and Nakazima cover
the range for high stress triaxiality values. Therefore, the loading angles in buttery tests are
mainly chosen to represent the stress states with low stress triaxialities. The specimens were
tested under ve loading angles, 10 compression , 0 shear, 10 tension, 20 tension and 60
tension (Fig. 5.25). The approximate range of stress triaxiality and Lode angle parameter
covered by buttery tests is depicted in Fig. 5.3
For each loading angle three duplicate tests were conducted. The displacement measurement was carried out optically by a camera. The measurement is carried out by tracking
two chosen material points on the specimen. The tracking material points were chosen on
the shoulders of the buttery specimens. The initial distance between track points were set
as 6.5mm for the loading angles 10 compression , 0 shear, 10 tension, 20 tension and as
5.0mm for the 60 tension (Fig. 5.26). The force signal measured by the testing machine,
which was synchronized with the optical displacement measurements.

50

5 Experimental Program

(a)
1

2a
Specimen
(b)
3a
4a
4b
60

3b
2b
5

Force

Figure 5.24: Buttery test bench set up for 60 loading angle. (a) Universal test machine
components: Inner discs (3a, 3b), grips (4a, 4b), connecting parts (2a, 2b), guides (1) and
force applying actuator (5). (b) Closer look to the specimen xed by grips after fracture.
10 Compression

0 Shear

10 Tension

20 Tension

60 Tension

Figure 5.25: Representation of tested buttery specimen loading angles, 10 compression, 0


shear, 10 tension, 20 tension and 60 tension.
(b)

45

6.

(a)

60

Figure 5.26: Representation of the tracking points used for measurements of buttery tests
with loading angles, (a) 10 compression, 0 shear, 10 tension, 20 tension and (b) 60
tension.

5.6 Buttery tests

51

The measured force is normalized with respect to cross-section area parallel to shoulders
at the center of specimens. The normalized force-displacement curves from the tests with ve
dierent loading angles are shown in Fig. 5.27. The scatter of normalized force-displacement
(b) 0 shear

0.7

0.7

0.6

0.6
F/A0 (Gpa)

F/A0 (Gpa)

(a) 10 compression

0.5
0.4
0.3
Test 1
Test 2
Test 3

0.2
0.1
0

0.1

0.2
0.3
L/L0

0.5
0.4
0.3
Test 1
Test 2
Test 3

0.2
0.1
0

0.4

0.7

0.7

0.6

0.6

0.5
0.4
0.3
Test 1
Test 2
Test 3

0.2
0.1
0

0.05

0.1

0.15
L/L0

0.1

0.15
0.2
L/L0

0.25

0.3

(d) 20 and 60 tension

F/A0 (Gpa)

F/A0 (Gpa)

(c) 10 tension

0.05

0.2

0.5

60 tension

0.4
0.3

20 tension

0.2

Test 1
Test 2
Test 3

0.1
0.25

0.05

0.1
L/L0

0.15

0.2

Figure 5.27: Normalized force-displacement curves obtained from the buttery tests with
loading angles (a) 10 compression (b) 0 shear (c) 10 (d) 20 and 60 tension.
levels for the duplicate tests is scarce, whereas the dierence in global displacement at fracture
is signicant. The minimum dierence at the normalized displacement at complete fracture
is 8% for the loading angle 20 compression, whereas maximum dierence was observed for
the tests with loading angles 10 compression and 60 tension as 19%. In fact the relative
high scatter may be the result of machining errors on the surface, since the gauge section has
a complex geometry. The fracture initiation occurs on the surface of the gauge section and
small machining irregularities on the surface may lead to change on the fracture initiation
times.

52

5 Experimental Program

Chapter 6
Determination of Fracture Strain and
Stress State
Abstract: In this chapter, a hybrid method combining the numerical simulations with
experimental results is discussed. The experimental global force-displacement responses
discussed in the previous chapter are used as reference for the numerical simulations. Numerical simulations are carried out to determine the numerical global force-displacement
responses and the components of stress and strain tensors. Material ductility (fracture
strain) and two stress state parameters, stress triaxiality and Lode angle parameter
are calculated from strain and stress tensors, respectively. A weighting function based
on the damage accumulation rule is proposed, since the stress state parameter values are not constant through the loading path. Numerical simulations are carried out
with isotropic J2 -plasticity. The damage accumulation is assumed to be nonlinear with
respect to plastic strain (n=2). The eect of mesh size on the physical quantities is
discussed briey.

6.1

Introduction

Experimental investigations give information about the physical nature of the material. However, in reality, the experimental measurements are usually not sucient to calibrate the
complex numerical models. Ductile fracture is an occurrence at local material points where
the fracture initiates. In conventional experimental methods, usually the needed physical information can not be acquired, since it is not possible to perform direct measurements at
local fracture initiation points. Therefore, numerical simulations are required to reveal the
physical information in conjunction with experiments.
A hybrid methodology is proposed, which combines the numerical simulations and experimental results. The experimental global force-displacement responses discussed in chapter 5
are used as reference for the numerical simulations. Numerical simulations are carried out to

54

6 Determination of Fracture Strain and Stress State

determine the numerical global force-displacement responses and the components of stress
and strain tensors. Material ductility (fracture strain) in dependence of two stress state parameters, stress triaxiality and Lode angle parameter , which is the focus of the thesis,
are calculated from strain and stress tensors obtained in numerical simulations.
The determination of the fracture initiation point is vital for the determination of fracture
strain and corresponding stress state. For some specimen types and loading cases fracture
initiation location can be observed optically. However the observation can be a challenging
task, if the fracture initiates inside the specimen or propagates very fast. In the present study,
the fracture initiation locations are determined numerically according to maximum plastic
strain and stress triaxiality locations in the specimens. With a few exceptions, the location of
the maximum plastic strain and maximum stress triaxiality in the specimens are coincident
for the investigated tests.
As shown in chapter 3, plastic ow can be stress state dependent especially for some
aluminum alloys. On the other hand, the stress state dependency of plastic ow for steels is
scarce. Also by consideration of applicability to crashworthiness simulations, the numerical
simulations are run with isotropic J2 -plasticity1 .
In CDM and micromechanical models (section 2.3) the stress-strain curve is specied for
the matrix material and global response is determined as the response of RVE which consists
of matrix material and damage (voids and cracks). On the other hand, the macroscopic
stress-strain curve measured in experiments is the combined response of material hardening
and damaging process, which makes the separation of plastic hardening of matrix material
and damage experimentally impossible. This implies that the material deforms according to
classical plasticity rule, without coupling the plastic ow and damage [119, 139]. Therefore,
in this study classical J2 -plasticity is used and damage is calculated simultaneously but it
does not aect the plastic ow.
In current crashworthiness simulations the mesh size of FE-models varies in the range of
1mm and 10mm. Especially for the coarse mesh size the post-critical response and fracture
strain is strongly mesh size dependent. In fact with course mesh it is not possible to model the
post-critical response and fracture strain accurately, since the mesh size is often larger than
physical localization region. In order to regularize the dissipated energy through deformation
to fracture, the damage is coupled with plasticity formulation in the GISSMO damage model.
The degree of coupling of plasticity with damage as well as fracture strain is set as a function
of mesh size numerically [1, 4]. Investigation of mesh size dependency of plasticity model is
not in the scope of the current thesis, since ne mesh size of 0.05mm-0.1mm is used in the
investigations.
1

J2 -plasticity refers to *MAT24 (*MAT PIECEWISE LINEAR PLASTICITY) in the context of finite
element code LS-DYNA.

6.2 Discussion on mesh size eects

6.2

55

Discussion on mesh size effects

In the current thesis it is intended to investigate the physical quantities fracture strain f ,
stress triaxiality and Lode angle parameter without introducing numerical errors caused
by mesh size. Therefore, the numerical simulations are carried out with ne mesh size. The
geometrical dimensions of critical locations dier for dierent sets of specimens. For example,
the critical gauge section has a length of 14mm for the un-notched at specimen, whereas
the critical section for the grooved at specimen with the groove radii equal to 0.5mm has
length of around 2mm. The deformation is localized in the investigated tests. In order to
complete the numerical simulations in reasonable times, in the noncritical regions (with low
deformation) of the specimens a coarse mesh is used.
The convergence analysis of mesh size was carried out for the un-notched at specimen
geometry, since it exhibits the maximum fracture strain in the experimental program within
all tests. FE models with mesh size of 0.1mm, 0.05mm and 0.025mm are investigated. Due
to symmetry conditions, 1/8 of the full specimen was modeled with applied symmetry conditions. The total number of elements of the FE-models are given in Table 6.2.

Table 6.1: FE-models of 1/8 specimens for three dierent mesh size and corresponding total
number of elements.
Mesh size

0.1mm

0.05mm

0.025mm

Number of elements

29280

234240

710720

In the convergence study, the global force-displacement responses and evolution of the
stress state parameters, stress triaxiality and Lode angle parameter were compared.
In Fig. 6.1-(a) the normalized force-displacement curves are compared in the range of the
normalized displacement 0-0.45, which corresponds to maximum local plastic strain of around
150%. In Fig. 6.1-(b) at the critical location (center of the specimen) the evolution of the
stress triaxiality and Lode angle parameter with respect to plastic strain is shown.
The maximum normalized displacement at fracture for three duplicate tests is around 0.43.
It is observed that the global force-displacement responses of three models are identical before
the onset of highly localized deformation. Also evolution of the stress triaxiality and Lode
angle parameters shows a similar trend. The assessed quantities are compared for dierent
mesh sizes at plastic strain of 1.5 in Table 6.2 by choosing the numerical model with mesh
size 0.05mm as reference. It is concluded that even for localized high deformation the results
are converged at the mesh size of 0.05mm.

56

6 Determination of Fracture Strain and Stress State


(a)

(b)
1

0.8

0.8

or

F/A0 (Gpa)

0.6

0.4

0.6
0.4

0.1mm
0.05mm
0.025mm

0.2

0.1mm
0.05mm
0.025mm

0.2

0.1mm
0.05mm
0.025mm

0
0

0.1

0.2

0.3

0.4

0.3

0.6

0.9

1.2

1.5

L/L0

Figure 6.1: Comparison of the numerical simulations for the mesh size of 0.1mm, 0.05mm and
0.025mm. (a) Normalized force-displacement curves and (b) the stress triaxiality and Lode
angle parameter with respect to plastic strain at the center of the specimen.
Table 6.2: Comparison of the numerical results of FE models with dierent discretization at
high deformation. Model with mesh size of 0.05mm is chosen as the reference and relative
dierence of other two models.
F
A0

6.3

(Gpa) at

L
=0.45
L0

at p =1.5

at p =1.5

0.1mm

+3%

-3%

+8%

0.05mm

0.384 (Ref.)

0.835 (Ref.)

0.411 (Ref.)

0.025mm

-1%

+1%

-3%

Determination of the stress-strain curve

As rst step in the numerical simulations the equivalent stress-strain curve is obtained. Commonly, un-notched at or axisymmetric specimens are used in the determination process. In
the present thesis, for this purpose un-notched at specimens are used, since the manufacturing of axisymmetric round specimens from the sheet metal with thickness of 2mm is prone
to geometrical deviations. The middle one of three experimental curves (Fig. 5.5-(a) is taken
as the reference curve for the task of stress-strain curve determination.
The stress-strain curve is determined in three steps. In the rst step, the experimental
measured quantities force F and displacement L are recorded. The engineering stress e
and engineering strain e are determined through
e =

F
,
A0

(6.1)

6.3 Determination of the stress-strain curve


and
e =

57

L
,
L0

(6.2)

where the L0 is the initial gauge length.


In the second step, engineering stress-strain curve is transformed to true stress-strain
(t -t ) curve with the equations
t = e (1 + e ) ,

(6.3)

t = ln(e + 1).

(6.4)

and

In the J2 -plasticity the scalar physical quantities equivalent stress eq and equivalent plastic
strain p dene the plastic ow. Before onset of necking, the stress state through the gauge
section is homogeneous and stress tensor has one component (uniaxial tension), which is
equal to equivalent stress. Also the strain distribution is homogeneous in the gauge section
and the equivalent plastic strain is equal to the component in loading direction in the plastic
strain tensor. However beyond the onset of necking, deformation localization occurs, which
leads to an inhomogeneous stress state in the gauge section. Therefore calculation of true
stress-strain (t - t ) is valid before onset of necking, which occurs at 16% of true strain
for the current material. The necking process is initiated when Consideres condition [140]
is met. Consideres condition is fullled at the maximum engineering stress level, which is
called ultimate tensile strength uts . When the elastic strain part is ignored, the condition is
dened in terms of equivalent stress eq and equivalent plastic strain p as
eq
= eq ,
p

(6.5)

where the corresponding equivalent stress and equivalent plastic strain values are called
necking equivalent stress nk and necking equivalent plastic strain nk (0.16 for the DP600 ),
respectively.
In the third step the stress-strain curve is extrapolated for the strains beyond uniform
elongation. In recent years many extrapolation methods have been proposed [141,142]. However most of them are geometry and material dependent, which necessitates an iterative
approach. In the current study, beyond onset of necking the stress-strain is extrapolated with
the power law of Ludwik [143] where stress is dened as a function of equivalent plastic strain
eq = a + bp c ,

(6.6)

where a, b and c are three Ludwik parameters. In order to obtain a continuous dierentiable
curve the stress-strain curve is extrapolated by consideration of following initial conditions

deq
eq |p =nk = nk and
= nk .
(6.7)
dp p =nk

58

6 Determination of Fracture Strain and Stress State

Table 6.3: Calibrated parameters of Ludwik power law.


a
0.42

0.61

0.47

The two power law parameters a and c are determined by the two initial conditions, which
leaves the parameter b as the only adjustable parameter to control the behavior of the curve.
The parameter b is changed iteratively until a good correlation is obtained between the experimental engineering stress-strain curve and numerical one. The three calibrated parameters
are shown in Table 6.3. The comparison of the engineering stress-strain curves of experiments
and numerical simulation with the calibrated stress-strain curve is demonstrated in Fig. 6.2.
It should be noted that experimental curve lies in the middle is used for the calibration.

Engineering or true stress (Gpa)

1
0.9
0.8
0.7
0.6
necking

0.5
0.4
0.3

e e Experiments
e e Experiments
t t Simulation

0.2
0.1
0
0

0.1

0.2

0.3

0.4

0.5

Engineering or true strain

Figure 6.2: Comparison of the experimental engineering stress-strain curves with the numerical one calculated with the calibrated true stress-strain curve (blue).

6.4

Weighting function for stress state parameters

Each specimen is assumed to represent a single point in the space of stress triaxiality, Lode
angle parameter and fracture strain. However, usually the two stress state parameter values
do not remain constant on the loading path to fracture, which make the calibration procedure
complicated. A typical example can be seen in Fig. 6.1, where both stress triaxiality and Lode
angle parameter change signicantly after the onset of necking. Thus weighting functions are
used in order to obtain discrete values for the two stress state parameters, stress triaxiality

6.5 Damage exponent

59

and Lode angle parameter. The weighting function should be based on damage increment
formulation in order to be consistent. In the current thesis, the weighting function is based
on the damage increment formulation of GISSMO damage model (Eq. 4.10) and dened as
g=

n1
n
D n .
f (, )

(6.8)

The weighted stress triaxiality w and weighted Lode angle parameter w are dened as

1 pf
w =
g (p ) dp ,
(6.9)
D pi
and

1
w =
D

pf

g (p ) dp ,

(6.10)

pi

where pi and pf are the lower and upper bounds for the integral. In the current thesis,
the damage is assumed to accumulate as the material starts to deform plastically. Then,
the lower bound of the integral is equal to zero (pi = 0) and the upper bound is equal to
fracture strain (pf = f ) where damage D is equal to unity. It should be noted that the
damage exponent n eects the integrals signicantly. For linear damage accumulation n = 1
the whole deformation (plastic strain) range has the same signicance. On the other hand,
for high damage exponents the plastic strain range at high deformation has more signicance.
The damage exponent inuence is removed, if the stress state parameter values are constant
during loading. These kinds of specimens/tests are of great importance, since the calibration
errors, which can be caused from damage exponent are eliminated.

6.5

Damage exponent

As in previous section stated, the damage exponent has signicant inuence on the determination of stress state parameters, which corresponds to specic fracture strain. The
nonlinearity of the damage accumulation with respect to plastic strain has been shown by
many researchers [51,128]. However, determination of the damage exponent n is a signicant
challenge. One method is to carry out stepwise experiments, which was applied by Tai [9]
as depicted in Fig. 6.3. Two tests with dierent notch radii correspond to stress triaxiality
values 1 and 2 are run to fracture as the reference tests. The third test with the stress
triaxiality 1 is run until a specic plastic strain level and stopped. The deformed specimen
is machined as it has the notch radius of second specimen geometry in order to set the stress
triaxiality value to 2 and then loaded to fracture. The damage exponent is determined by
integrating the damage in stepwise experiments with the help of reference experiments. It
should be noted that the Lode angle parameter inuence on the fracture strain is eliminated,
since for the axisymmetric notched round specimens it has a constant value ( = 1) during

60

6 Determination of Fracture Strain and Stress State

eq
No. 1
1

2
eq

No. 2

2
eq

(a)

(b)

Figure 6.3: Stepwise methodology applied by Tai [9].

loading. On the other hand, for the materials with high ductility, the stress triaxiality value
usually changes during the loading and makes the described methodology dicult to apply
practically.
In fact step wise experiments were carried out rstly by Bridgman [67] in high pressure
vessels. He introduced the stress triaxiality dierence by adjusting the hydrostatic pressure.
Another method is the Nakazima tests followed by machining at specimens. However, in
this method not only stress triaxiality but also the Lode angle parameter is changed from
= 1 to = 1. The method is valid for plane stress state conditions.
By using Bridgmans stepwise experiments for ductile steel and assuming constant stress
triaxiality through the loading, Xue [14] calculated the damage exponent as n = 2.21 for a
similar damage accumulation formulation.
Recently, Tasan [10] investigated dierent damage quantication methods under two
main groups, material property-based methods and morphology-based methods for DP600,
which is also the investigated material in the current thesis. He concluded that the material property-based methodologies have higher accuracy than morphology-based methods. In
Fig. 6.4 damage calculated from a material property-based methodology (density measurement) is depicted. The damage rule in GISSMO damage model is applied to the start and
end points of the measured damage data for the damage exponents of n = 2 and n = 3.
By considering slightly better correlation with damage exponent n = 2 and determination of
Xue, in this thesis, as an assumption nonlinear damage accumulation with damage exponent
n = 2 is considered.

6.6 Numerical simulations

61

Damage (%)

5
4
3
n=3

n=2
2
1
0
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Equivalent strain

Figure 6.4: Damage quantication with the methodology of density measurement [10] and
GISSMO damage rule application for the damage exponents of n = 2 and n = 3.

6.6

Numerical simulations

The subject of the current section is the determination of the representative fracture strain,
stress triaxiality and Lode angle parameter values for each test. The numerical simulations
of the conducted tests discussed in the chapter 5 were carried out. The used plasticity model
is J2 -plasticity and all numerical simulations are run with the stress-strain curve obtained in
section 6.3. In this chapter damage is not considered in the numerical code.
For each test the fracture strain and two stress state parameters, stress triaxiality and
Lode angle parameter are determined at the location of fracture initiation. The fracture
initiation location can be observed optically. However, as already mentioned, usually it is a
very challenging task since fracture can be initiated inside the specimen or propagates very
fast during loading. In this case the two questions have to be answered; where and at which
global displacement the fracture initiates? The fracture initiation location for specimens can
be investigated experimentally by interrupting the tests at dierent displacement levels and
slicing the specimen followed by examination with a microscope. This approach has been
applied by some researchers for the investigation of axisymmetric round bars and it has
been found that crack initiates at the center of the specimen where stress triaxiality and
equivalent plastic strain are the largest [22]. The experimental force-displacement responses
can also give information about global displacement at fracture initiation. Usually there is
a signicant drop as the crack initiates [74]. In this section fracture initiation locations and
displacement at fracture are determined by consideration of following assumptions:
The fracture process is very fast and the global displacement dierence between fracture
initiation and signicant amount of material failure is relatively small.

62

6 Determination of Fracture Strain and Stress State


The fracture initiation is controlled by the stress triaxiality and plastic strain during
loading. Therefore the fracture initiates at the location with highest stress triaxiality
and plastic strain values.
Material damaging occurs as material starts to deform plastically.
The damage accumulation is nonlinear with respect to plastic strain and the damage
exponent (n) is equal to 2.

6.6.1

Flat specimens

The simulations are based on un-notched at specimens and notched at specimens with
radii of 2mm and 4mm. Due of the symmetry conditions, only a 1/8 of the specimens are
modeled. The number of elements of the FE models is listed in Table 6.4 The numerical
Table 6.4: Number of elements of 1/8 at-specimen FE models.
un-notched

notched R=4mm

notched R=2mm

Mesh size

0.05mm

0.05mm

0.05mm

Number of elements

234240

194880

246160

simulations are run with fully integrated 8-node brick elements (LS-DYNA ETYPE 2 2 ).
In Fig. 6.5 the plastic strain, stress triaxiality and Lode angle parameter distribution at
the critical section of the un-notched at specimen is shown for the global displacement at
the onset of necking and at fracture initiation. For all three at specimen geometries, the
material point at the center of the specimen has the highest value of the stress triaxiality
and plastic strain at fracture. Therefore the fracture is assumed to initiate at the center of
specimens.
For the un-notched at specimen, the comparison of the normalized force-displacement
curves obtained from numerical simulation and experiments is depicted in Fig. 6.6. The
evolution of the plastic strain at the center of the specimen with respect to normalized
global displacement is shown with the blue curve. The x marks in Fig. 6.6 indicate the
deformation state (onset of necking and fracture initiation) in Fig. 6.5. The experimental
curve in the middle is the reference curve for the numerical simulations. The fracture strain
is determined as 1.41 for the un-notched specimen. It is observed that beyond the onset of
necking, the evolution of the plastic strain is highly nonlinear with respect to normalized
global displacement.
2

ETYPE 2 refers to full integration for solid elements in LS-DYNA.

6.6 Numerical simulations

63

Plastic strain

(a)

1.50

0.8

1.0

1.35

0.7

0.9
0.8

1.20

0.6

1.05

0.5

0.90

0.4

0.75

0.3

0.60

0.2

0.7
0.6
0.5
0.4

0.1

0.45

0.3

0.30

(b)

Lode angle parameter

Stress triaxiality

0.2

0.15

0.1

Figure 6.5: Plastic strain, stress triaxiality and Lode angle parameter distribution at the
critical section of the un-notched at specimen for global displacement (a) at the onset of
necking and (b) at fracture.
1.6

0.8

1.4

0.7
(a)

1.2

Fracture

0.5

(b)

0.4

0.8

F/A0 - Experiment-1,2,3
F/A0 - Simulation
p - Simulation

0.3

0.6

0.2

0.4

0.1

0.2

0.1

0.2

0.3

0.4

0.5

Plastic strain

F/A0 (Gpa)

0.6

L/L0

Figure 6.6: Comparison of the experimental and numerical normalized force-displacement


curves (L0 =10mm). The plastic strain with respect to global normalized displacement at the
critical location is shown with the blue curve.
The evolution of the stress state variables, stress triaxiality and Lode angle parameter
with respect to plastic strain is illustrated in Fig. 6.7. It can be seen that, both stress state
parameters indicate exactly the stress state of uniaxial tension state before the onset of
necking. The weighted stress state parameters w and w are calculated with the assumption
of nonlinear damage accumulation (n=2) and fracture strain of 1.41.
The comparison of the experimental and numerical normalized force-displacement curves

64

6 Determination of Fracture Strain and Stress State

Stress triaxiality or
Lode angle parameter

0.8

0.6

0.4

0.2

0.3

0.6

0.9

1.2

1.5

Plastic strain

Figure 6.7: Evolution of two stress state parameters with respect to plastic strain and weighted
stress state parameters (squares) for the un-notched at specimen. x marks denote the
fracture strain.
for the notched specimens with notch radii R=2mm and 4mm are presented in Fig. 6.8.
It is observed that after the maximum force there is dierence between the experimental
0.7

1.4

0.6

1.2
F/A0 - R=4mm-Exp.1,2,3
F/A0 - R=4mm-Sim.
p - R=4mm-Sim.

0.4

1
0.8

F/A0 - R=2mm-Exp.1,2,3
F/A0 - R=2mm-Sim.
p - R=2mm-Sim.

0.3

0.6

0.2

0.4

0.1

0.2

0.01

0.02

0.03

0.04

0.05

0.06

Plastic strain

F/A0 (Gpa)

0.5

L/L0

Figure 6.8: Comparison of the experimental and numerical normalized force-displacement


curves (L0 =30mm) for notched at specimens R=4mm and 2mm. The plastic strains with
respect to global normalized displacement at the critical locations are shown with the blue
curves.
and numerical normalized force-displacement curves. Global displacements at fracture are

6.6 Numerical simulations

65

depicted with downward arrows for the experiments. The values for the numerical simulations
are depicted with x marks, which are calculated as the mean values of three duplicate tests
for both notch geometries. The fracture strains at the center of specimens are determined as
0.81 and 0.78 for the notch radii R=4mm and R=2mm, respectively. The evolution of the
stress state parameters with respect to plastic strain and weighted values at fracture strain
are demonstrated in Fig. 6.9.
1

Stress triaxiality or
Lode angle parameter

0.8

0.6

0.4
R=4mm
R=4mm

w R=4mm
w R=4mm

R=2mm
R=2mm

w R=2mm
w R=2mm

0.2

0
0

0.2

0.4

0.6

0.8

Plastic strain

Figure 6.9: Evolution of the stress triaxiality and Lode angle parameter with respect to plastic
strain at critical location and weighted stress state parameters (squares) for the notched at
specimens.
The fracture strain, weighted stress triaxiality and Lode angle parameter values for all
three specimens are represented in Table 6.5
Table 6.5: Fracture strain, weighted stress triaxiality and Lode angle parameter for at unnotched and notched specimens.
Un-notched

Notched R=4mm

Notched R=2mm

Fracture strain f

1.41

0.81

0.78

Weighted stress triaxiality w

0.61

0.63

0.67

Weighted Lode angle parameter w

0.73

0.59

0.36

66

6 Determination of Fracture Strain and Stress State

6.6.2

Grooved flat specimens

In this section the grooved at specimens with groove radii R=0.5mm, 1mm, 2mm and 4mm
are investigated. By consideration of symmetry conditions 1/8 of the specimens are modeled
with applied symmetry boundary conditions. Special attention is given for the mesh size in
the groove section and mesh size of 0.05mm is applied. Number of elements of FE models
used in the investigation are listed in Table 6.6. In the numerical simulations fully integrated
Table 6.6: Number of elements of the grooved at specimens ( 1/8 FE models).

Mesh size
Number of elements

R=0.5mm

R=1mm

R=2mm

R=4mm

0.05mm

0.05mm

0.05mm

0.05mm

40300

52600

65200

52800

8-node brick elements were used.


The fracture formation location is determined from numerical simulations by investigation
of the stress triaxiality and plastic strain distributions in the specimens. At fracture initiation,
the contour plot of plastic strain, stress triaxiality and Lode angle parameter in the groove
section are shown in Fig. 6.10 for two geometries with groove radii R=1mm and 4mm. It is
observed that plastic strains (Fig. 6.10-(a),(b)) and stress triaxiality (Fig. 6.10-(c),(d)) values
are the highest at the center of the specimens. Therefore, it is likely that fracture initiates
at the center. It should be emphasized that the stress state is plane strain (=0) at a wide
region on the plane which passes through the center perpendicular to loading direction and
it changes to uniaxial tension state at the edges. The size of the plane strain and transition
region is determined mainly with the thickness and width at the center of groove as explained
in section 5.3.
The comparison of the experimental normalized force-displacement curves with the numerical ones are shown in Fig. 6.11. In order to determine the fracture strain at the center
of the specimen, the evolutions of the plastic strain with respect to global normalized displacement are shown with blue curves. The fracture is assumed to propagate very fast after
the initiation and the global displacement dierence at fracture initiation and complete separation of specimen is assumed to be negligible. The x marks denote the fracture points,
which are determined as the average of end points of the curves of two duplicate tests for each
groove radius geometry. For all specimen geometries a good correlation is achieved between
experiments and numerical simulations in terms of force-displacement responses.
The histories of stress triaxiality with respect to plastic strain at the center of specimens
are shown in Fig. 6.12.
Except the geometry with notch radius R=0.5mm, the stress triaxiality shows a slightly

6.6 Numerical simulations

(a) Plastic strain R=1mm

67

0.7

(b) Plastic strain R=4mm


1.05

0.6

0.90

0.5

0.75

0.4

0.60

0.3

0.45

0.2

0.30

0.1

0.15

(c) Stress triaxiality R=1mm

(d) Stress triaxiality R=4mm


0.875

1.05

(e) Lode angle parameter R=1mm

0.90

0.750

0.75

0.625

0.60

0.500

0.45

0.375

0.30

0.250

0.15

0.125

1.00
0.75

(f) Lode angle parameter R=4mm

1.00
0.75

0.50

0.50

0.25

0.25

0.00

0.00

-0.25

-0.25

-0.50

-0.50

-0.75

-0.75

-1.00

-1.00

Figure 6.10: Plastic strain, stress triaxiality and Lode angle parameter distributions at fracture formation for the grooved at specimens with groove radii R=1mm ((a),(c),(e)) and
R=4mm ((b),(d),(f)).

increasing trend through deformation. The maximum dierence of stress triaxiality between
the beginning and end of deformation is around 23% for the specimen with notch radius
R=4mm, which is relatively low.
The evolution of the Lode angle parameter with respect to plastic strain is illustrated in
Fig. 6.13 for all groove radii.
It is observed that the Lode angle parameter converges very quickly to the plane stress
state ( = 0) and remains constant through the loading to fracture. The weighted stress
state values w and w are calculated by considering the damage exponent n=2 and shown
in Fig. 6.12 and Fig. 6.13.
The detailed information of fracture strain and weighted stress state values w and w is
given in Table 6.7.

68

6 Determination of Fracture Strain and Stress State

(a) Groove radius R=0.5mm

(b) Groove radius R=1mm

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

0
0

0
0.005 0.01 0.015 0.02 0.025 0.03
L/L0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

0
0

(d) Groove radius R=4mm


1

0.6

0.4

0.4

0.2

0.2
0
0.005 0.01 0.015 0.02 0.025 0.03

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

0.6

F/A0 (Gpa)

0.8
p

F/A0 (Gpa)

0.8

L/L0

0
0.005 0.01 0.015 0.02 0.025 0.03

F/A0 - Exp.-1,2,3
F/A0 - Sim.
p - Sim.

(c) Groove radius R=2mm

F/A0 (Gpa)

F/A0 (Gpa)

0
0

0
0.005 0.01 0.015 0.02 0.025 0.03

L/L0

L/L0

Figure 6.11: Comparison of the experimental and numerical normalized force-displacement


curves (L0 =20mm) and evolution of plastic strain at the center of specimens for the grooved
at specimens with groove radii (a) R=0.5mm, (b) R=1mm, (c) R=2mm and (d) R=4mm.
1.2

Stress triaxiality

1
0.8
0.6
0.4
0.2

R=0.5mm

w R=0.5mm

R=1mm

w R=1mm

R=2mm

w R=2mm

R=4mm

w R=4mm

0
0

0.1

0.2

0.3

0.4
0.5
0.6
Plastic strain

0.7

0.8

0.9

Figure 6.12: Evolution of the stress triaxiality with respect to plastic strain and weighted
stress triaxiality values for the grooved at specimens with groove radii equal to 0.5mm,
1mm, 2mm and 4mm. x marks denote the fracture strain.

6.6 Numerical simulations

69

Lode angle parameter

0.8

0.6

0.4

R=0.5mm

w R=0.5mm

R=1mm

w R=1mm

R=2mm

w R=2mm

R=4mm

w R=4mm

0.2

0
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Plastic strain

Figure 6.13: The weighted Lode angle parameter values and evolutions of the Lode angle
parameter with respect to plastic strain for the grooved at specimens with groove radii
equal to 0.5mm, 1mm, 2mm and 4mm.x marks denote the fracture strain.
Table 6.7: Fracture strain, weighted stress triaxiality and Lode angle parameter for the
grooved at specimens.
R=0.5mm

R=1mm

R=2mm

R=4mm

Fracture strain f

0.59

0.70

0.84

0.98

Weighted stress triaxiality w

1.03

0.88

0.80

0.76

Weighted Lode angle parameter w

0.01

0.01

0.01

0.01

6.6.3

Axisymmetric notched round specimens

Numerical simulations based on the set of axisymmetric notched round specimen tests were
investigated. The mesh size is chosen as 0.04mm at the critical location, which corresponds
to notch of the specimens. Due to symmetry only 1/8 of the specimens are modeled by
applying symmetry boundary conditions. The number of elements are listed in Table 6.8.
The numerical simulations were run with fully integrated 8-node brick elements.
Table 6.8: Number of elements of axisymmetric notched round specimens ( 1/8 FE models).

Mesh size
Number of elements

R=0.5mm

R=1mm

R=2mm

R=4mm

0.04mm

0.04mm

0.04mm

0.04mm

64929

73992

52872

51168

For the notch geometries R=1mm and R=4mm the distribution of plastic strain and two

70

6 Determination of Fracture Strain and Stress State

stress state parameters are shown in Fig. 6.14. It should be noted that Lode angle parameter
(a) Plastic strain R=1mm

0.8

(b) Plastic strain R=4mm

1.05

0.6

0.90

0.5

0.75

0.4

0.60

0.3

0.45

0.2

0.30

0.1

0.15

(c) Stress triaxiality R=1mm

(d) Stress triaxiality R=4mm


0.9

0.9

0.8

0.8

0.7

0.7

0.6

0.6

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

(e) Lode angle parameter R=1mm

1.20

0.7

1.00

(f) Lode angle parameter R=4mm

1.00

0.75

0.75

0.50

0.50

0.25

0.25

0.00

0.00

-0.25

-0.25

-0.50

-0.50

-0.75

-0.75

-1.00

-1.00

Figure 6.14: Distribution of the plastic strain, stress triaxiality and Lode angle parameter at
fracture initiation deformation for the geometries with notch radii R=1mm and 4mm.
is constant ( = 1) through the midplane for the specimens, which disables the Lode angle
parameter to be an inuence factor within the investigated notch geometries. For all notch
geometries, the regions with the highest plastic strain and stress triaxiality are coincident and
they correspond to the center of specimens (See. Fig. 6.14-(a),(c) for R=1mm and Fig. 6.14(b),(d) for R=4mm). Therefore it is reasonable to assume the center of the specimens as
fracture initiation location.
The force-displacement curves and history of plastic strain at the center of specimens
are shown in Fig. 6.15. In terms of maximum normalized force there is dierence of around
7% and 6% for the notch radii R=0.5mm and R=1mm, respectively. This may be caused
from machining uncertainties, since the dimensions of the specimens are small and machining smaller radii can lead to geometrical deviations. The other possible cause may be pressure

6.6 Numerical simulations

71

1.2

0.6

0.9

0.4

0.6

0.2

0.3

0.01 0.02 0.03 0.04 0.05 0.06

L/L0

0.8

1.2

0.6

0.9

0.4

0.6

0.2

0.3

0.8

F/A0 (Gpa)

(b) Notch radius R=1mm

F/A0 (Gpa)

(a) Notch radius R=0.5mm

F/A0 - Exp.-1,2
F/A0 - Sim.
p - Sim.

0.6

0.9

0.4

0.6

0.2

0.3

0.01 0.02 0.03 0.04 0.05 0.06


L/L0

L/L0

0.8

1.2

0.6

0.9

0.4

0.6

0.2

0.3

1.2
F/A0 (Gpa)

0.8

(d) Notch radius R=4mm

F/A0 (Gpa)

(c) Notch radius R=2mm

0.01 0.02 0.03 0.04 0.05 0.06

0.01 0.02 0.03 0.04 0.05 0.06


L/L0

Figure 6.15: Normalized force-displacement curves (L0 =10mm) and evolution of the plastic
strain at the center of specimens for the axisymmetric notched round specimens with notch
radii (a) R=0.5mm, (b) R=1mm, (c) R=2mm and (d) R=4mm.

dependence of plastic ow especially at high pressure values. It should be noted that the nonlinearity of the evolution of the plastic strain with respect to global displacement increases
as the notch radius increases, which indicates signicant deformation localization. The normalized displacement at fracture initiation (x marks) is determined as the mean of two
duplicate tests (downward arrows) for each notch geometry.
The evolutions of the stress triaxiality and Lode angle parameter with respect to plastic
strain are given in Fig. 6.16 and Fig. 6.17, respectively. It is observed that through the
loading, the stress triaxiality value changes signicantly especially for the notch radii R=2mm
and 4mm, in which the localization is severe compare to other notch geometries. As expected
the Lode angle parameter has a constant value = 1 for all specimen geometries through
loading Fig. 6.17. In both diagrams the determined fracture strains (x marks) and weighted
stress state values (squares) with damage exponent n = 2 are illustrated.
The fracture strains and weighted stress state parameters are listed in Table 6.9

72

6 Determination of Fracture Strain and Stress State


1.2

Stress triaxiality

1
0.8
0.6
0.4
0.2

R=0.5mm

w R=0.5mm

R=1mm

w R=1mm

R=2mm

w R=2mm

R=4mm

w R=4mm

0
0

0.2

0.4

0.6
Plastic strain

0.8

1.2

Figure 6.16: Evolution of the stress triaxiality with respect to plastic strain and weighted
stress triaxiality valuesfor the axisymmetric specimens with notch radii equal to 0.5mm,
1mm, 2mm and 4mm. x marks denote the fracture strain for the corresponding test.

Lode angle parameter

1
0.8
0.6
0.4
0.2

R=0.5mm

w R=0.5mm

R=1mm

w R=1mm

R=2mm

w R=2mm

R=4mm

w R=4mm

0
0

0.2

0.4

0.6
0.8
Plastic strain

1.2

Figure 6.17: Histories of Lode angle parameter with respect to plastic strain and weighted
Lode angle parameter values for the axisymmetric specimens with notch radii R=0.5mm,
1mm, 2mm and 4mm. x marks illustrate the fracture strain for the corresponding test.
Table 6.9: Fracture strain, weighted stress triaxiality and Lode angle parameter for axisymmetric notched round specimens.
R=0.5mm

R=1mm

R=2mm

R=4mm

Fracture strain f

0.67

0.88

1.08

1.19

weighted stress triaxiality w

0.95

0.82

0.74

0.70

weighted Lode angle parameter w

1.00

1.00

1.00

1.00

6.6 Numerical simulations

6.6.4

73

Nakazima tests

Numerical simulations of Nakazima tests were carried out for two blank geometries with
width of 70mm and 90mm (section 5.5). The fracture initiation occurs at the center of blanks
for the both geometries. The mesh size is chosen as 0.2mm at the center of specimens by
considering the relatively big dimensions of the blanks. As it can be seen in Fig. 5.19, the
blank geometries have two symmetry planes. Thus, 1/4 of the test components are modeled
and on the symmetry planes corresponding displacement and rotations are constrained. Main
factor in the consideration is to complete the numerical simulations within a reasonable
amount of time. The FE models of blanks with width of 70mm and 90mm have 144348 and
125726 8 node brick elements, respectively. The components, die, punch and punch holder
are modeled as rigid bodies, since the deformation is assumed to be negligible on these parts.
The FE model of the Nakazima test with blank width of 90mm is shown in Fig. 6.18. The
(a)

(b)

Die

Blank holder

Punch

Figure 6.18: (a) FE model (1/4) of Nakazima test with blank width of 90mm and (b) a close
view of the critical region.
numerical simulations are run with fully integrated solid elements. The friction coecient is
set to be 0.01 between punch and blank, since in experiments friction eects are minimized.
As in the experiments, in the numerical simulations the displacement and force responses are
measured on the punch.
It can be argued that, mesh size of 0.2mm is not consistent with the numerical simulations
of other tests. The mesh size inuence was investigated by developing models with mesh size of
0.1mm. The global force-displacement curve, plastic strain and two stress state parameters
at critical locations were compared. It was observed that the dierence in plastic strain
evolutions and stress state parameters obtained from the models with mesh size 0.1mm and
0.2mm is negligibly small and global-force-displacement curves are identical for both models.
It is concluded that the numerical results converge at the mesh size of 0.2mm for the Nakazima

74

6 Determination of Fracture Strain and Stress State

tests. In fact the deformation is not very localized for the Nakazima tests and fracture strains
are low compare to other tests, which also explains the convergence for relatively coarse mesh
size.
In the experiments, for both blank geometries fracture was observed through the symmetry plane at the center of the blanks (Fig. 5.21). The deformation at the center region is the
largest for both tests. However it is not highly concentrated. For both tests, the contour plot
of plastic strain at fracture initiation is shown in Fig. 6.19-(a),(b). The stress triaxiality value
is around 2/3 over a wide region on the upper side of blanks for both tests Fig. 6.19-(c),(d).
By considering the experimental and numerical results, the fracture is assumed to initiate at
the center of the upper surface of blanks and the fracture strain and stress state parameters
are determined according to these locations.
The force-displacement curves obtained by numerical simulations are compared to experimental curves in Fig. 6.20. For both tests there is a dierence in the force-displacement
responses between numerical simulations and experiments. The dierence in force level at
fracture initiation point is 12% and 6%, respectively. The plastic strain evolution at fracture initiation location with respect to global displacement is also shown in Fig. 6.20. The
fracture points are determined from the average of peaks of three duplicate tests and shown
with x marks on the numerical force-displacement and plastic strain-displacement curves.
For both blank geometries numerical determined fracture strains are consistent with experimental calculated ones in section 5.5. The comparison of the experimentally and numerically
determined fracture strains is shown in Table 6.10.

Table 6.10: Comparison of the numerically determined fracture strains to the experimentally
determined ones for the Nakazima tests with blank width of 70mm and 90mm.
Width=70mm

Width=90mm

Experimental fracture strain

0.57

0.88

Numerical fracture strain

0.55

0.84

The evolution stress triaxiality and Lode angle parameter with respect to plastic strain
p at critical location are shown in Fig. 6.21. For the test with blank with of 90mm, stress
state is exactly equi-biaxial ( = 2/3 and = 1) through the loading to fracture initiation.
The numerical results with the blank width of 70mm diers from the equi-biaxial stress state
especially in the Lode angle parameter value and not in the stress triaxiality value (w = 0.64
and w = 0.74). From this point of view, it is clear that the signicant dierence in fracture
strains between two tests indicates the inuence of Lode angle parameter.

6.6 Numerical simulations

75

(a) Width=70mm

(b) Width=90mm
0.50

0.9

0.45

0.8

0.40

0.7

Plastic strain

0.35

0.6

0.30

0.5

0.25

0.4

0.20

0.3

0.15

0.2

0.10

0.1

0.05

(c) Width=70mm

(d) Width=90mm

Stress triaxiality

0.7

Lode angle parameter

(e) Width=70mm

0.7

0.6

0.6

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

1.0

(f) Width=90mm

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

0.0

0.0

-0.2

-0.2

-0.4

-0.4

-0.6

-0.6

-0.8

-0.8

-1.0

-1.0

Figure 6.19: Plastic strain ((a), (b)), stress triaxiality ((c), (d)) and Lode angle parameter
((e), (f)) distribution for the Nakazima tests with blank width of 70mm and 90mm at fracture
formation.

6.6.5

Butterfly tests

The numerical simulations of buttery tests are investigated for the loading angles 10 compression , 0 shear, 10 tension, 20 tension and 60 tension. In experimental setup the force
is measured from the machine and displacement is measured optically. By considering the
dimensions of the test bench, the machine stiness is assumed to be suciently large and
machine is not modeled in numerical simulations. The boundary conditions at the grips are

76

6 Determination of Fracture Strain and Stress State


(a) Width=70mm

(b) Width=90mm

100

0.8

60

0.6

40

0.4

20

0.2

Force (kN)

80

120

10
15
20
Displacement (mm)

F-d Exp.-1,2,3
F-d Sim.
p -d Sim.

80

1
0.8

60

0.6

40

0.4

20

0.2

0
0

1.2

25

F-d Exp.-1,2,3
F-d Sim.
p -d Sim.

100

1.2

Force (kN)

120

0
0

10
15
20
Displacement (mm)

25

Figure 6.20: Comparison of the force-displacement curves obtained from numerical simulations to the experimental ones and evolution of plastic strain at critical location with respect
to displacement for Nakazima test blank geometries with width of (a) 70mm and (b) 90mm.
1

Stress triaxiality or
Lode angle parameter

0.75
0.5
0.25
0
-0.25

w=70mm
w=70mm

w w=70mm
w w=70mm

w=90mm
w=90mm

w w=90mm
w w=90mm

-0.5
-0.75
-1
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Plastic strain

Figure 6.21: The weighted stress state values and evolution of the stress triaxiality and Lode
angle parameter with respect to the plastic strain for the Nakazima specimens with blank
width of 70mm and 90mm.
applied to shoulders of the buttery specimen as in the test bench.
At the critical location a mesh size of 0.075mm is used, which is consistent with previous
numerical simulations explained in this section. Due to symmetry conditions only 1/2 of
the specimen is modeled with applied symmetry conditions. The 1/2 FE model has 79.000
8-node fully integrated brick elements. The nite element discretization of the central region
of specimen is shown in Fig. 6.22 in detail.
Further assessment of mesh size inuence is carried out by developing a FE model with
mesh size of 0.0375mm in the gauge section. The dierence between two models in terms

6.6 Numerical simulations

77

Figure 6.22: A view of nite element model discretization of the central region of the buttery
specimen.
of plastic strain and stress state parameters at fracture initiation is around 0-3% for ve
loading angles. Therefore, in the current thesis the results for the mesh size of 0.075mm are
presented.
For all loading angles the plastic strain distribution on specimens just before fracture
initiation are shown in Fig. 6.23.
10 compression

0 shear

10 tension
0.8

1.1

1.1

1.0

1.0

0.7

0.9

0.9

0.6

0.8

0.8

0.5

0.7

0.7

0.4

0.6

0.6

0.3

0.5

0.5

0.2

0.4

0.4

0.1

0.3

0.3

0.2

0.2

0.1

0.1

20 tension

60 tension

0.45

0.7

0.40

0.6

0.35

0.5

0.30

0.4

0.25

0.3

0.20

0.2

0.15

0.1

0.10
0.05
0

Figure 6.23: Plastic strain distribution contours for the buttery specimens just before the
fracture initiation for all loading angles.
Due to the unique geometry of the buttery specimen for all loading cases, the deformation is highly concentrated at the center of the specimen. Thus, it is reasonable to assume
that the fracture formation occurs at the center of the specimens, however usually it is not

78

6 Determination of Fracture Strain and Stress State

possible to observe the exact location of the fracture formation through the thickness of the
specimen. Therefore, the fracture formation location is determined from numerical simulations by checking the elements located at the center of specimen beginning from the surface
to the middle of thickness. The quantities, plastic strain and stress triaxiality are used for the
assessment of fracture formation location through the thickness. For the loading angles 10
compression , 0 shear and 10 tension the dierence in the stress triaxiality on the surface
and middle of thickness is not signicant. For these loading angles, the locations with higher
plastic strain are used for the calibration. For the loading angle 20 tension the stress triaxiality is slightly higher at the middle of thickness, whereas the plastic strain is higher on the
surface at the center of specimens. For this loading angle fracture is assumed to initiate at
the surface. For the loading angle 60 tension, the location with the highest stress triaxiality
and plastic strain is the middle of thickness. The evolution of the stress triaxiality and plastic
strain on the surface and middle of thickness for the loading angles 10 compression and 60
tension are shown in Fig. 6.24. The shown experimental normalized force-displacement curve
is the one, which lies in the middle of three curves.
(a) 10 compression

(b) 60 tension

0.8
0.6

F/A0 Exp.

0.4

p S
p M

0.2

0
-0.2

M
0

0.1

0.2
L/L0

0.3

0.4

P or or F/A0 (Gpa)

P or or F/A0 (Gpa)

1
0.8
0.6
0.4
0.2
0

0.02

0.04

0.06

0.08

0.1

L/L0

Figure 6.24: Stress triaxiality and plastic strain evolution on the surface and middle of thickness of buttery specimens for the loading angles (a) 10 compression and (b) 60 tension.
Determined fracture initiation locations in the main plane and in the thickness direction
for all loading angles are listed in Table 6.11.
Numerical normalized force-displacement curves are compared with experimental ones in
Fig. 6.25.
The results obtained from numerical simulations are in a very good accordance with experimental ones. The maximum dierence is observed for the loading angle 10 compression
as around 3% at normalized displacement 0.05. Since the buttery tests cover a wide range on
the stress triaxiality and Lode angle parameter plane, it may be concluded that the investigated material does not exhibit stress triaxiality and Lode angle parameter dependent plastic

6.6 Numerical simulations

79

Table 6.11: Fracture initiation locations in the main plane and thickness directions of buttery
specimens for dierent loading angles.
Loading angle

Location in the main plane

Location in the thickness direction

10 compression

Center

Surface

shear

Center

Surface

10 tension

Center

Surface

20

tension

Center

Surface

60 tension

Center

Middle of thickness

ow in the stress triaxiality range of [-0.1:0.8]. The evolution of plastic strain at determined
critical locations are also shown in Fig. 6.25. Tests with loading angles 10 compression, 0
shear and 10 tension exhibit globally slightly softening before fracture. For these loading
angles the fracture is assumed to initiate just after maximum of the global normalized force.
Due to scatter in duplicate tests, the mean of three tests are used in determination of displacement at fracture initiation. The normalized global displacement at fracture initiation
and corresponding plastic strains are depicted with x marks in Fig. 6.25.
Also correlation between experiments and simulations is good in terms of deformation.
The deformations of the gauge section and edge curves in experiments are captured in numerical simulations. The deformation comparison for the loading angle 0 shear is shown in
Fig. 6.26.
The evolutions of the stress triaxiality with respect to plastic strain at critical location
for all loading angles are demonstrated in Fig. 6.27. It should be emphasized that the stress
triaxiality has a relatively constant trend with increasing plastic strain. The weighted stress
triaxiality values (squares in Fig. 6.27) are calculated by assuming the damage exponent n=2.
The Lode angle parameter vs. plastic strain at critical locations are shown in Fig. 6.28.
Unlike stress triaxiality evolution, the Lode angle parameter evolution is not constant with
respect to plastic strain especially for the loading angles 10 compression and 0 shear.
In Fig. 6.29 the histories for ve loading cases are shown in the space of Lode angle
parameter and stress triaxiality. It should be noted that stress state is plane stress from the
beginning of loading to fracture for the tests with loading angles 10 compression, 0 shear,
10 and 20 tension. Thus tests with mentioned loading angles also give information that can
be applied in 2-dimensional applications.
The determined fracture strains and weighted stress triaxiality and Lode angle parameter
values are listed in Table 6.12

6 Determination of Fracture Strain and Stress State


(b) 0 shear
0.8

1.2

0.6

0.9

0.6

0.9

0.4

0.6

0.4

0.6

0.2

0.3

0.2

0.3

0
0.1

0.2

0.3

0
0.4

0
0

0.05

0.1

0.15

L/L0

0.25

0
0.3

L/L0

(c) 10 tension

(d) 20 tension
1.2

0.8

1.2

0.6

0.9

0.6

0.9

0.4

0.6

0.4

0.6

0.2

0.3

0.2

0.3

0
0

0.05

0.1

0.15

0.2

F/A0 (Gpa)

0.8

F/A0 (Gpa)

0.2

0
0.25

0
0

0.05

0.1

L/L0

0.15

0.2

F/A0 (Gpa)

1.2

F/A0 (Gpa)

(a) 10 compression
0.8

80

0
0.25

L/L0

p - Sim.

0.8

0.6

0.6

0.4

0.4

0.2

0.2

0
0

0.03

0.06

0.09
L/L0

0.12

F/A0 - Exp.-1,2,3
F/A0 - Sim.

F/A0 (Gpa)

(e) 60 tension
0.8

0
0.15

Figure 6.25: Normalized force-displacement curves and evolution of the plastic strain at critical location for buttery test simulations with loading angles (a) 0 compression, (b) 0 shear,
(c) 10 tension, (d) 20 tension and 60 tension. The gauge length L0 for the loading angle
60 tension is 5mm, for the rest L0 =6.5mm.
Table 6.12: List of fracture strain and weighted stress state parameters for buttery tests.
10 comp.

0 sh.

10 ten.

20 ten.

60 ten.

Fracture strain

1.02

1.08

0.70

0.83

0.44

Stress triaxiality

-0.06

0.07

0.18

0.52

0.82

Lode angle parameter

-0.25

0.27

0.70

0.99

0.28

6.6 Numerical simulations

81

Experiment

Simulation

Figure 6.26: Comparison of the deformation just before fracture initiation under loading angle
0 shear.

1
10 compression

0.8

0 shear

Stress triaxiality

10 tension
0.6

20 tension
60 tension

0.4

w 10 compression
w 0 shear

0.2

w 10 tension
w 20 tension

w 60 tension

-0.2
0

0.2

0.4

0.6
Plastic strain

0.8

1.2

Figure 6.27: Evolution of the stress triaxiality with respect to plastic strain and weighted
stress triaxiality values for simulations of buttery tests with loading angles 10 compression,
0 shear, 10 , 20 and 60 tension. x marks denote the fracture strain for corresponding
loading angle.

82

6 Determination of Fracture Strain and Stress State

Lode angle parameter

1
0.8

10 compression

0.6

0 shear

0.4

10 tension
20 tension

0.2

60 tension

w 10 compression

-0.2

w 0 shear

-0.4

w 10 tension

-0.6

w 20 tension
w 60 tension

-0.8
-1
0

0.2

0.4

0.6
Plastic strain

0.8

1.2

Figure 6.28: Evolution of the Lode angle parameter with respect to plastic strain and weighted
Lode angle parameter values for simulations of buttery tests with loading angles 0 compression, 0 shear, 10 , 20 and 60 tension. x marks denote the fracture strain for corresponding
loading angle.

Lode angle parameter

1
0.8

Plane stress

0.6

10 compression

0.4

0 shear
10 tension

0.2

20 tension

60 tension
10 comp. weighted

-0.2

0 shear weighted

(
)
= 27
2 13
2

-0.4

10 tension weighted

-0.6

20 tension weighted

-0.8

60 tension weighted

-1
-0.4

-0.2

0.2

0.4

0.6

0.8

Stress triaxiality

Figure 6.29: Representation of the histories and weighted stress state values obtained from
numerical simulations in the plane of stress triaxiality and Lode angle parameter for the
buttery tests.

6.7 Discussion

6.7

83

Discussion

The inuence of the two stress state parameters, stress triaxiality and Lode angle parameter
can be separated by keeping one of them constant in the investigations. As discussed in
previous section, the Lode angle parameter has a constant value through deformation to
fracture for the set of axisymmetric notched round specimens ( = 1) and for the set of
grooved at specimens it converges quickly to = 0 at the beginning of the deformation.
Thus, the inuence of stress triaxiality may be examined for two dierent constant Lode
angle parameter values. The JC equation can be applied to the two sets separately by mean
square error method with a good correlation (Fig. 6.30). However, it should be noted the
1.5

Fracture strain

1.2

0.9

Fitted JCGrooved flat


0.33 + 12.35 exp (3.81)

Fitted JCNotched round


w Grooved flat

0.6

w Notched round

0.56 + 1072.12 exp (10.27)

0.3

0
0.4

0.6

0.8

1.2

Stress triaxiality

Figure 6.30: Fracture strain with respect to weighted stress triaxiality and calibrated JC
equations for test sets of axisymmetric notched round specimens and grooved at specimens.
range of stress triaxiality for the calibration is relatively small.
For all tests, the fracture strain with respect to weighted stress triaxiality values is plotted
in Fig. 6.31. It is clear that the material ductility of the examined material exhibits Lode angle
parameter sensitivity, as some tests with similar stress triaxiality show signicant dierence
in fracture strain. For example, the two Nakazima tests and axisymmetric notched round
specimen with notch radius R=4mm have similar stress triaxiality values; however fracture
strain dierence is signicant (%118 between the lowest and highest fracture strain).
The calibrated curve obtained at high stress triaxiality range is unlikely to t for the low
stress triaxiality values. The calibrated curve obtained from the set of axisymmetric notched
round specimens mismatches the buttery test with loading angle 20 tension signicantly,
which has a similar Lode angle parameter value (w = 0.96). The similar trend is also observed
for the tting curve obtained from grooved at specimens. The fracture strain of buttery
tests with loading angles 10 compression (w =-0.06) and 10 shear (w =0.07) are lower than

84

6 Determination of Fracture Strain and Stress State


1.5
Flat specimens
R=2mm

Smooth

Fracture strain

1.2

R=4mm

Grooved flat specimens


R=0.5mm

0.9

R=1mm

R=2mm

R=4mm

Notched round specimens


R=0.5mm

0.6

R=1mm

R=2mm

R=4mm

Butterfly tests
10 compression

10 tension

20 tension

0.3

0 shear
60 tension

Nakazima tests
w=70mm

w=90mm

0
-0.3

0.3

0.6

0.9

1.2

Stress triaxiality

Figure 6.31: Fracture strain versus stress triaxiality for all tests.
the tting curve of grooved at specimens predicts. Therefore all points may be used to
generate a tting curve, which will be an optimum curve rather than matching every single
point.
The fracture strain of tests are shown in stress triaxiality and Lode angle parameter space
in Fig. 6.32. Usually the at specimen tests are used to determine the material properties
1
Flat specimens

Lode angle parameter

0.75

R=2mm

Smooth

R=4mm

Grooved flat specimens

0.5

R=0.5mm

R=1mm

R=2mm

R=4mm

0.25
Notched round specimens
0

R=0.5mm

R=1mm

R=2mm

R=4mm

Butterfly tests

-0.25
-0.5
-0.75

10 compression

0 shear

20 tension

60 tension

10 tension

Nakazima tests
w=70mm

w=90mm

-1
-0.3

0.3
0.6
Stress triaxiality

0.9

1.2

Figure 6.32: Representation of the weighted stress state values in the plane of stress triaxiality
and Lode angle parameter.
under plane stress state. However, it is seen that none of the at specimens are on the plane
stress curve in Fig. 6.32. It can be concluded that the ratio of width with respect to thickness

6.7 Discussion

85

of specimen is a determining factor. In the current study a ratio of 2.5 is applied and the stress
state deviates from the plane stress. As discussed previously (subsection 6.6.5), buttery tests
with four loading angles and two Nakazima tests have plane stress at the critical locations.
For the current research, it is shown that there are not any experiments that represent
the negative stress triaxiality range less than = 0.06. Many researchers pointed out
that the fracture strain is signicantly higher at high stress triaxialities compare to low
stress triaxialities ([144]). The material is unlikely to fail under negative stress triaxialities
in practical applications. However the negative stress triaxialities may be investigated with
uniaxial compression tests and biaxial compression tests further. The covered stress state
regions by uniaxial and biaxial compression tests are shown as blue boxes in Fig. 6.33.
Lode angle
parameter
1

Axisymmetric notched
round specimens

Bu

Plane stress

Flat specimens

tter
fly

0.5

tes

ts

Biaxial
compression
tests

0
-1

-0.5

0.5

Grooved flat
specimens
1

Stress triaxiality
1.5

Plane strain
-0.5

Uniaxial
compression -1
tests

Nakazima
tests

Figure 6.33: Illustration of the possible additional test types (depicted as blue boxes) in
order to cover negative stress triaxialities on the plane of stress triaxiality and Lode angle
parameter.

86

6 Determination of Fracture Strain and Stress State

Chapter 7
Determination of Fracture Locus
Abstract: In this chapter, the fracture strain is dened as the third dimension over
the plane of stress triaxiality and Lode angle parameter. Two dierent fracture strain
surfaces are generated as analytical and mathematical fracture strain surface. The numerical simulations are carried out with generated fracture strain denition. The parameters of analytical fracture locus derived in section 4.3 are calibrated by considering
all data points (fracture strain and weighted stress state values) obtained in numerical simulations of tests. As the second approach, mathematical fracture strain surface
based on biharmonic spline method is generated. In this approach a continuous dierentiable fracture strain surface is placed exactly on the data points in fracture strain,
stress triaxiality and Lode angle parameter space. The numerical results obtained with
fracture strain denitions with damage exponent n=2 are presented. Predictions of two
fracture surface denitions are compared with experimental results.

7.1
7.1.1

Simulation with analytical fracture strain definition


Calibration of parameters

The nine-parameter (D1+ ...D3 ) fracture strain denition derived in section 4.3 is calibrated
according to data points (fracture strain, weighted stress triaxiality and Lode angle parameter) obtained in numerical simulations of tests in section 6.6. A MATLAB code is used in
order to minimize the dierence between the fracture strains obtained from numerical simulations of tests f ,i and analytical fracture strain f ,i denitions. The optimization is done
with the method of nonlinear least squares. The nine damage parameters are held in the
vector p and minimum of error function is dened through
M in Error (w , w ; p) = M in

i=1

[f ,i (w , w ) f ,i (w , w ; p)]2 .

(7.1)

88

7 Determination of Fracture Locus

In previous chapter numerically obtained fracture strains and weighted stress state parameters are summarized in (Table 7.1).
Table 7.1: List of fracture strain and weighted stress state parameters obtained from numerical
simulations of all tests.
Stress triaxiality

Lode angle parameter

Fracture strain

smooth

0.61

0.73

1.41

R=2mm

0.67

0.36

0.78

R=4mm

0.63

0.59

0.81

R=0.5mm

1.03

0.01

0.59

R=1mm

0.88

0.01

0.70

R=2mm

0.80

0.01

0.84

R=4mm

0.76

0.01

0.98

R=0.5mm

0.95

1.00

0.67

R=1mm

0.82

1.00

0.88

R=2mm

0.74

1.00

1.08

R=4mm

0.70

1.00

1.19

w=70mm

0.64

-0.74

0.55

w=90mm

0.66

-1.00

0.84

10 compression

-0.06

-0.25

1.02

0.07

0.27

1.08

Flat specimens

Grooved flat specimens

Notched round specimens

Nakazima tests

Butterfly tests
shear

10

tension

0.18

0.70

0.70

20

tension

0.34

0.96

0.95

60 tension

0.82

0.28

0.44

As the initial values for the optimization process, the parameters obtained for the sets of
axisymmetric notched round specimens and grooved at specimens (section 6.7) are used for
the bound curves with = 1 and = 0, respectively. The parameters obtained for the set
of axisymmetric notched round specimens are also used as initial values for the remaining
bound curve = 1.
The list of optimized nine parameters is given in Table 7.2.
The calibrated fracture locus is represented in Fig. 7.1. In the stress triaxiality range of

7.1 Simulation with analytical fracture strain denition

89

Table 7.2: A list of calibrated parameters of the analytical fracture strain surface.
D2+

D3+

D10

D20

D30

D1

D2

D3

0.391

0.953

0.404

0.534

0.706

0.018

1.298

0.725

D1+
0.783

(a)
Flat specimens

1.5

R=2mm

Smooth

R=4mm

Grooved flat specimens

R=0.5mm

R=1mm

R=2mm

R=4mm

Notched round specimens

0.5

R=0.5mm

R=1mm

R=2mm

R=4mm

Butterfly tests

10 compression

0 shear

10 tension

20 tension

60 tension

0.4

0.8
1.2

-0.5

-1

0.5

Nakazima tests
w=70mm

w=90mm

(b)

(c)
1.5

1.5

1
f

f
0.5

0.5

0
0

0.4

0.8

1.2

-1

-0.5

0.5

Figure 7.1: Illustration of the generated analytical fracture surface from (a) perspective (b)
f plane and (c) f plane view. The vertical lines on the data points denote the range
of fracture strain obtained from the duplicate tests.

[-0.3-1.2], the calibrated fracture strain surface predicts the minimum dierence in fracture
strain through the Lode angle parameter direction at = 0.3 (25%) and maximum dierence
at = 1.2 (47%). It is clear that, the generated fracture surface does not match perfectly for
the all data points and it is an optimum solution for the all data points.

90

7 Determination of Fracture Locus

7.1.2

Numerical simulations

The numerical simulations of the tests are carried out with J2 -plasticity and GISSMO damage
model1 is activated. The plasticity and damage indicator are not coupled, which means plastic
ow is not aected by damage. The material damage is modeled by eroding elements as the
damage indicator reaches the unity D = 1.
The fracture locus is implemented as a table denition into commercial code LS-DYNA
under *MAT ADD EROSION material card. Table denition consists of stress triaxiality
dependent fracture strain curves for dierent Lode angle parameters. In LS-DYNA the fracture strain value for a specic combination of a stress state parameters is calculated by linear
interpolation of the nearest curves. In order to obtain a high resolution, the calibrated analytical fracture locus is generated for the stress triaxiality and Lode angle parameter increments
of 0.01.
Flat specimens
The numerical simulations were run for the un-notched and notched at specimens with notch
radii R=4mm and 2mm. The numerical normalized force-displacement curves are compared
with experimental ones in Fig. 7.2. It is observed that for un-notched at specimen the
(b) Notched: R=2mm and R=4mm
0.8

0.6

0.6
F/A0 (Gpa)

F/A0 (Gpa)

(a) Un-notched
0.8

0.4
Experiments
Simulation

0.2

0.4
Exp. R=4mm
Sim. R=4mm
Exp. R=2mm
Sim. R=2mm

0.2

0
0

0.1

0.2

0.3
L/L0

0.4

0.5

0.01

0.02

0.03
0.04
L/L0

0.05

0.06

Figure 7.2: Comparison of the normalized force-displacement curves from numerical simulations and experiments for the (a) un-notched (L0 =10mm) and (b) notched at specimens
with notch radii equal to 2mm and 4mm (L0 =30mm).
normalized displacement at fracture is 9% lower than the experimental reference curve. On
the other hand, plastic strain at the center of the specimen is 41% lower than the fracture
strain for the specimen determined in subsection 6.6.1. The high dierence indicates the
severe strain localization at high plastic strains for the specimen, which is already mentioned
1

GISSMO damage model is implemented in *MAT ADD EROSION in finite element code LS-DYNA.

7.1 Simulation with analytical fracture strain denition

91

in subsection 6.6.1. For notched specimens, the normalized displacements at fracture obtained
from numerical simulations have good agreement with experimental ones. For the specimen
with notch radius R=2mm, the displacement at fracture initiation is within the range of
experimental data. For the notch radius R=4mm, the displacement at fracture initiation
is slightly higher (2%) then the duplicate experiment curve with maximum displacement
at fracture. In fact the results are consistent with the calibrated fracture strain surface in
previous section.
The distribution of the damage indicator D at fracture initiation is illustrated in Fig. 7.3.
It can be seen that, damage is highly localized at the center of specimens, which is also
(a) Un-notched

(b) R=2mm

(c) R=4mm

0.9

0.9

0.9

0.8

0.8

0.8

0.7

0.7

0.7

0.6

0.6

0.6

0.5

0.5

0.5

0.4

0.4

0.4

0.3

0.3

0.3

0.2

0.2

0.2

0.1

0.1

0.1

Figure 7.3: Contour plot of the damage indicator D at fracture initiation for the at specimens; (a) un-notched, notched with (b) R=2mm and (c) R=4mm.
consistent with the assumption of occurrence of fracture initiation at the material point with
highest stress triaxiality and plastic strain.
Grooved flat specimens
Numerical simulations of the tests of grooved at specimens were run. The normalized forcedisplacement curves obtained in numerical simulations are compared to experimental ones in
Fig. 7.4. A good correlation is obtained between numerical simulations and tests in terms of
global displacement at fracture for the specimens with groove radii R=0.5mm and 1mm. On
the other hand, the numerical displacement at fracture is lower than the experiments for the
groove radii R=2mm and especially for R=4mm. The dierence between the curves of the
numerical simulation and the duplicate test with lower displacement at fracture is %11 for
R=2mm and %15 for R=4mm, even though the plastic strain value at the center of specimens
diers %20 and %27, respectively. This situation indicates the localization phenomena.
The contour plot of damage indicator at fracture initiation deformation is shown in
Fig. 7.5.
The damage is highest at the center of the specimens, where maximum stress triaxiality
and plastic strain are observed and in the numerical simulations the rst failed element is

92

7 Determination of Fracture Locus


(b) R=1mm and R=2mm
1

0.8

0.8
F/A0 (Gpa)

F/A0 (Gpa)

(a) R=0.5mm and R=1mm


1

0.6
Exp. R=0.5mm
Sim. R=0.5mm
Exp. R=1mm
Sim. R=1mm

0.4
0.2

0.6
Exp. R=2mm
Sim. R=2mm
Exp. R=4mm
Sim. R=4mm

0.4
0.2

0
0

0.005

0.01
L/L0

0.015

0.02

0.005

0.01

0.015 0.02
L/L0

0.025

0.03

Figure 7.4: Comparison of the normalized force-displacement response (L0 =20mm) of grooved
at specimens between the experiments and numerical simulations for the groove radii (a)
R=0.5mm, 1mm, (b) 2mm and 4mm.
(a) R=0.5mm

(b) R=1mm
1

0.9

0.9

0.8

0.8

0.7

0.7

0.6

0.6

0.5

(c) R=2mm

0.4

0.5

(d) R=4mm

0.4

0.3

0.3

0.2

0.2

0.1

0.1

Figure 7.5: Contour plot of the damage indicator D at fracture initiation for the grooved at
specimens for groove radii (a) R=0.5mm, (b) R=1mm, (c) R=2mm and (d) R=4mm.
exactly located at the center of the specimens.
Axisymmetric notched round specimens
The numerical simulations were run for the specimens with notch radii R=0.5mm, 1mm,
2mm and 4mm. The comparison of numerical and experimental force-displacement responses
are presented in Fig. 7.6. For the specimen geometry with R=0.5mm the numerical normalized displacement at fracture is signicantly higher (19%) than the experimental curve with
higher fracture point; on the contrary for the specimen geometry with R=4mm the numerical
prediction is conservative (10%) compared to experiments. The numerical displacements at
fracture are slightly outside of the range obtained in experiments for the specimens with notch
radii R=1mm and 2mm. The numerical results are consistent with the calibrated fracture
locus (see Fig. 7.1 the bound curve with red color).

7.1 Simulation with analytical fracture strain denition


(b) R=2mm and R=4mm

0.8

0.8

0.6

0.6

F/A0 (Gpa)

F/A0 (Gpa)

(a) R=0.5mm and R=1mm

0.4
Exp. R=0.5mm
Sim. R=0.5mm
Exp. R=1mm
Sim. R=1mm

0.2

93

0.4
Exp. R=2mm
Sim. R=2mm
Exp. R=4mm
Sim. R=4mm

0.2

0
0

0.01

0.02
0.03
L/L0

0.04

0.02

0.04
L/L0

0.06

Figure 7.6: Comparison of the normalized force-displacement responses (L0 =10mm) of axisymmetric notched round specimens between the experiments and numerical simulations for
the notch radii (a) R=0.5mm, 1mm, (b) 2mm and 4mm.
The contour plot of damage indicator is shown in Fig. 7.7. As expected, the damage
(a) R=0.5mm

(b) R=1mm
1

0.9

0.9

0.8

0.8

0.7

0.7

0.6

0.6

0.5

(c) R=2mm

0.4

0.5

(d) R=4mm

0.4

0.3

0.3

0.2

0.2

0.1

0.1

Figure 7.7: Contour plot of the damage indicator D at fracture initiation for the axisymmetric
notched round specimens for notch radii (a) R=0.5mm, (b) R=1mm, (c) R=2mm and (d)
R=4mm.
localization occurs at the center of specimens and fracture initiates exactly at the center of
the specimen, which corresponds to the location with highest stress triaxiality and fracture
strain.
Nakazima tests
The numerical simulations are carried out for the Nakazima tests with blank width of 70mm
and 90mm. The comparison between numerical and experimental results in terms of punch
force-displacement curves are illustrated in Fig. 7.8 The correlation between numerical sim-

94

7 Determination of Fracture Locus


(a) Width=70mm

(b) Width=90mm

120

120
100
Experiments
Simulation

80

Force (kN)

Force (kN)

100

60

60

40

40

20

20

Experiments
Simulation

80

0
0

10
15
20
Displacement (mm)

25

10
15
20
Displacement (mm)

25

Figure 7.8: A comparison of the force-displacement curves obtained from experiments and
numerical simulations of Nakazima tests.
ulation and experiments in terms of displacement at fracture is satisfactory for the test with
blank width of w=90m. On the other hand the displacement at fracture is overestimated for
the blank width w=70mm, as expected from the calibrated fracture locus (see Fig. 7.1). The
fracture strain prediction of the calibrated analytical fracture surface is about 40% higher
than the determined fracture strain in previous chapter. However, due of the nonlinear behavior of plastic strain at critical region with respect to global displacement (see Fig. 6.20),
the dierence is 15% in terms of global displacement at fracture initiation.
The distribution of the damage indicator D is presented in Fig. 7.9. It is observed that,
(a) Width=70mm

(b) Width=90mm
1

0.9

0.9

0.8

0.8

0.7

0.7

0.6

0.6

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

Figure 7.9: Contour plot of the damage indicator D at fracture initiation for the Nakazima
tests with blank width of (a) 70mm and (b) 90mm.
the damage is not localized in a small region, but spans a larger area compared to other sets
of tests. Thus there is as a relatively wide region, where the fracture initiation can occur. The
exact location of fracture initiation can be also inuenced by the local material irregularities

7.1 Simulation with analytical fracture strain denition

95

in the experiments. The post-mortem state of the blanks is compared between the numerical
simulations and experiments in Fig. 7.10.
(a) w=70mm

Experiment

(b) w=90mm

Experiment

Figure 7.10: Comparison of the post-mortem state of Nakazima tests between experiments
and numerical simulations for the tests with blank width of (a) w=70mm and (b) w=90mm.

Butterfly tests
The numerical simulations are run for the loading angles 10 compression , 0 shear, 10
tension, 20 tension and 60 tension. Comparison of the numerical and experimental forcedisplacement responses are presented in Fig. 7.11. The agreement between numerical simulation and experiments in terms of global displacement at fracture initiation is satisfactory for
the loading angles 10 compression and 20 tension. On the other hand, numerical simulation
under predicts the global displacement at fracture initiation for the loading angle of 0 shear
(10% ), while the results are signicantly over predicted for the loading angles of 10 tension
(20%) and 60 tension (27%).
The contour plot of the damage indicator D at fracture initiation in numerical simulations
is shown in Fig. 7.12. It should be noted that damage concentration location is at the center
of the main plane. However location through the thickness diers depending on the loading
angle. For the loading angles 10 compression , 0 shear, 10 and 20 tension the damage is
localized on the surface and for the loading angle 60 tension the highest damage is observed
in the middle of the thickness. These results are consistent with the chosen locations for
the calibration in subsection 6.6.5. The critical region for the buttery specimens tends to
move from the surface to the middle of the thickness with increasing loading angle (stress
triaxiality).
The deformation comparison of the fractured specimens in numerical simulations and experiments are presented in Fig. 7.13. The deformation and fracture path of tests are captured
successfully in numerical simulations.

96

7 Determination of Fracture Locus


(b) 0 shear and 10 tension

0.7

0.7

0.6

0.6

0.5

0.5

F/A0 (Gpa)

F/A0 (Gpa)

(a) 10 compression

0.4
0.3
Experimens
Simulation

0.2
0.1
0

0.1

0.2
0.3
L/L0

0.4
0.3

Exp. 0 shear
Sim. 0 shear
Exp. 10 tension
Sim. 10 tension

0.2
0.1
0

0.4

0.05

0.1

0.7

0.7

0.6

0.6

0.5
0.4
0.3
Experimens
Simulation

0.2

0.3

0.5
0.4
0.3
Experimens
Simulation

0.2

0.1
0

0.25

(d) 60 tension

F/A0 (Gpa)

F/A0 (Gpa)

(c) 20 tension

0.15
0.2
L/L0

0.1
0

0.05

0.1

0.15
0.2
L/L0

0.25

0.3

0.05

0.1
L/L0

0.15

0.2

Figure 7.11: Comparison of the experimental and numerical force-displacement responses for
the buttery tests with loading angles (a) 10 compression, (b) 0 shear, 10 , (c) 20 and (d)
60 tension (L0 =5.0mm for the loading angle 60 tension, for the rest L0 =6.5mm).
(a) 10 compression

(b) 0 shear

0.9

0.9

0.9

0.8

0.8

0.8

0.7

0.7

0.7

0.6

0.6

0.6

0.5

0.5

0.5

(d)

20

shear

(c) 10 tension

0.4

0.4

0.4

0.3

0.3

0.2

0.2

0.2

0.1

0.1

0.1

0.3

(e)

60

tension

Figure 7.12: Damage indicator distribution at the gauge section of buttery tests for the
loading angles (a) 10 compression , (b) 0 shear, (c) 10 , (d) 20 and (e) 60 tension.

7.2 Simulation with mathematical fracture strain surface


(a) 10 compression
Simulation

(b) 0 shear

Experiment

Simulation

(c) 10 tension

Experiment

(d) 20 shear
Simulation

Experiment

97

Simulation

Experiment

(e) 60 tension
Simulation

Experiment

Figure 7.13: Comparison of the fractured specimens in numerical simulations and experiments.

7.2

Simulation with mathematical fracture strain surface

In previous section it is shown that the determined analytical fracture strain surface denition
can lead to good results in terms of global displacement at fracture in general. However the
dierence is signicant for some tests. As an alternative approach a mathematical fracture
strain surface is generated in order to enhance the prediction of the numerical simulations. The
mathematical surface is generated on the data points (w , w , f ) determined in chapter 6.
The used method MATLAB v4 [145] is based on biharmonic spline method [146]. The
generated surface is continuous and dierentiable, which always passes through the data
points. The discussion on interpolation methods is not in the scope of thesis and will not
discussed further. As in the previous section, the table is dened with stress triaxiality and
Lode angle parameter increments of 0.01 in order to obtain a good resolution.

7.2.1

Fracture strain surface determination

The data points obtained from all sets of experiments are used in the interpolation process
of fracture strain surface. In addition, the generated fracture surface is controlled by some
assumptions for the high and negative stress triaxialities, where no experimental data exists.

98

7 Determination of Fracture Locus

For the current research, there are not any experimental results for the negative stress
triaxiality range less than = 0.06. It has been shown by many researches that the fracture
strain values for negative stress triaxialities are signicantly higher than the values at high
stress triaxialities. Bao [144] dened a cut-o negative triaxiality at the stress triaxiality =1/3, below which material never fails. In the current section, material is assumed to fail at a
fracture strain value of 3 at stress triaxiality = 1/3 in all Lode angle parameter range.
The second assumption is done for the high stress triaxiality range, 1.2. For the Lode
angle parameters = 1 and = 0 the fracture strain is assumed to follow the exponential
trend obtained from the sets of axisymmetric notched round specimens and grooved at
specimens (see section 6.7), respectively. The fracture strain for the range of stress triaxiality
1.2 at = 1, = 0.5 and = 0.5 is also assumed to follow the exponential trend
obtained from the notched specimens.
The mathematical fracture strain surface generated over the plane of stress triaxiality
and Lode angle parameter is demonstrated in Fig. 7.14. Since the generated fracture surface
follows the data points, it has no certain behavior in the stress triaxiality and Lode angle
parameter direction as for the analytical fracture strain surface. Similar fracture strain surface
with local extrema was also introduced by Seid [5].

7.2.2

Numerical simulations

The numerical simulations with the denition of Mathematical Fracture Strain surface (MFS)
are investigated. Numerical force-displacement responses are compared to the experimental
ones. In order to assess the improvement, the numerical force-displacement results obtained
in previous section with Analytical Fracture Strain surface (AFS) are also shown in the
diagrams.
Flat specimens
The comparison of the force-displacement responses for the experiments and numerical simulations for the at specimens are shown in (Fig. 7.15). The agreement in terms of forcedisplacement responses between experiments and numerical simulations is slightly improved
with the denition of MFS, whereas for the notched at specimen with R=2mm the agreement with MFS is slightly worsened, which can be caused by the history path of stress state
variables through the loading.
Grooved flat specimens
The numerical simulations were run for the set of grooved at specimens with the MFS.
The force-displacement responses obtained from numerical simulations with AFS and MFS

7.2 Simulation with mathematical fracture strain surface

99

(a)

2.5
2
f

1.5
1
1
0.5

0.5
0

-0.5

0.4

0.8

(b)

1.2

-1

Flat specimens
R=2mm

Smooth

2.5

R=4mm

Grooved flat specimens

R=0.5mm

f 1.5

R=1mm

R=2mm

R=4mm

Notched round specimens


R=0.5mm

R=1mm

R=2mm

R=4mm

1
Butterfly tests
0.5
0

0.4

0.8

1.2

10 compression

0 shear

20 tension

60 tension

10 tension

Nakazima tests
w=70mm

w=90mm

Figure 7.14: Illustration of the generated mathematical fracture strain surface.


are compared to experimental ones for the groove radii R=0.5mm, 1mm, 2mm and 4mm
in Fig. 7.16. The correlation between experiments and numerical simulations with MFS in
terms of global displacement at fracture strain is good for all groove radii. The numerical
simulations with MFS show signicant improvement compared to numerical simulations with
AFS especially for the groove radii R=2mm and 4mm (see Fig. 7.16).
Axisymmetric notched round specimens
The numerical simulations of axisymmetric notched round specimens were run with MFS.
The force-displacement curves obtained from experiments and numerical simulations with
denitions of AFS and MFS are compared in Fig. 7.17. As for the grooved at specimens,

100

7 Determination of Fracture Locus


(b) Notched: R=2mm and R=4mm

0.8

0.8

0.6

0.6
F/A0 (Gpa)

F/A0 (Gpa)

(a) Un-notched

0.4
Experiments
Simulation MFS
Simulation AFS

0.2

Exp. R=4mm
Sim. R=4mm MFS
Sim. R=4mm AFS
Exp. R=2mm
Sim. R=2mm MFS
Sim. R=2mm AFS

0.4

0.2

0
0

0.1

0.2

0.3
L/L0

0.4

0.5

0.01

0.02

0.03
0.04
L/L0

0.05

0.06

Figure 7.15: Comparison of the normalized force-displacement curves from numerical simulations (with mathematical fracture surface (MFS) and analytical fracture surface (AFS)) and
experiments for the (a) un-notched (L0 =10mm) and (b) notched at specimens with notch
radii R=2mm and R=4mm (L0 =30mm).

(b) R=2mm and R=4mm


1

0.8

0.8
F/A0 (Gpa)

F/A0 (Gpa)

(a) R=0.5mm and R=1.0mm


1

0.6
Exp. R=0.5mm
Sim. R=0.5mm MFS
Sim. R=0.5mm AFS
Exp. R=1mm
Sim. R=1mm MFS
Sim. R=1mm AFS

0.4
0.2

0.6
Exp. R=2mm,,
Sim. R=2mm MFS
Sim. R=2mm AFS
Exp. R=4mm
Sim. R=4mm MFS
Sim. R=4mm AFS

0.4
0.2

0
0

0.005

0.01
L/L0

0.015

0.02

0.005

0.01

0.015 0.02
L/L0

0.025

0.03

Figure 7.16: Comparison of the normalized force-displacement response (L0 =20mm) of


grooved at specimens between the experiments and numerical simulations with analytical and mathematical fracture strain denitions for the groove radii (a) R=0.5mm, 1mm, (b)
2mm and 4mm.

the correlations between experiments and numerical simulations with MFS are good in terms
of global displacement at fracture initiation for all notch radii. Signicant improvement is
obtained in numerical simulations with MFS for the notch radii R=0.5mm and 4mm, for
which the numerical simulations with AFS over predict and under predict the experimental
results, respectively.

7.2 Simulation with mathematical fracture strain surface

101

(b) R=2mm and R=4mm

0.8

0.8

0.6

0.6

F/A0 (Gpa)

F/A0 (Gpa)

(a) R=0.5mm and R=1mm

Exp. R=0.5mm
Sim. R=0.5mm MFS
Sim. R=0.5mm AFS
Exp. R=1mm
Sim. R=1mm MFS
Sim. R=1mm AFS

0.4

0.2

Exp. R=2mm
Sim. R=2mm MFS
Sim. R=2mm AFS
Exp. R=4mm
Sim. R=4mm MFS
Sim. R=4mm AFS

0.4

0.2

0
0

0.01

0.02
0.03
L/L0

0.04

0.02

0.04
L/L0

0.06

Figure 7.17: Comparison of the normalized force-displacement curves (L0 =10mm) obtained
from experiments and numerical simulations with AFS and MFS for the axisymmetric
notched round specimens for the notch radii (a) R=0.5mm, 1mm, (b) 2mm and 4mm.
Nakazima tests
The numerical simulations are carried out for the Nakazima tests with blank width w=70mm
and 90mm. The force-displacement responses of experiments and numerical simulations with
MFS and AFS are illustrated in Fig. 7.18. The Nakazima blank with width of 90mm shows no
(a) Width=70mm

(b) Width=90mm

120

120
Experiments
Simulation MFS
Simulation AFS

80

Experiments
Simulation MFS
Simulation AFS

100
Force (kN)

Force (kN)

100

60

80
60

40

40

20

20

0
0

10
15
20
Displacement (mm)

25

10
15
20
Displacement (mm)

25

Figure 7.18: The force-displacement response comparison for the experiments and numerical
simulations with MFS and AFS for the Nakazima tests.
signicant dierence in terms of force-displacement response for the numerical simulations
with MFS and AFS. In fact for both fracture strain surface denitions, the data point of
the specimen is on the generated surfaces. For the blank geometry with width of 70mm
the displacement at fracture initiation is vastly improved in the numerical simulations with

102

7 Determination of Fracture Locus

MFS compare to AFS, since the MFS follows the fracture strain point of the test, which is
overestimated in the generated AFS.
Butterfly tests
The numerical simulations with MFS were run for the buttery tests. The normalized forcedisplacement curves acquired from numerical simulations with MFS and AFS and experiments are compared in Fig. 7.14. It is shown that the correlation between experiments and
(b) 0 shear and 10 tension
0.7

0.6

0.6

0.5

0.5

F/A0 (Gpa)

F/A0 (Gpa)

(a) 10 compression
0.7

0.4
0.3
Experimens
Simulation MFS
Simulation AFS

0.2
0.1

Sim. 0 shear AFS


Sim. 0 shear AFS

0.4
0.3
Exp. 0 shear
Sim. 0 shear MFS
Exp. 10 tension
Sim. 10 tension MFS

0.2
0.1

0
0

0.1

0.2
0.3
L/L0

0.4

0.05

0.7

0.7

0.6

0.6

0.5

0.5

0.4
0.3
Experimens
Simulation MFS
Simulation AFS

0.2
0.1

0.15
0.2
L/L0

0.25

0.3

(d) 60 tension

F/A0 (Gpa)

F/A0 (Gpa)

(c) 20 tension

0.1

0.4
0.3
Experimens
Simulation MFS
Simulation AFS

0.2
0.1

0
0

0.05

0.1

0.15
0.2
L/L0

0.25

0.3

0.05

0.1
L/L0

0.15

0.2

Figure 7.19: The force-displacement response comparison for the experiments and numerical
simulations with MFS and AFS for the buttery tests. The gauge length L0 for the loading
angle 60 tension is 5mm, for the rest of loading angles L0 =6.5mm.
numerical simulations in terms of global displacement at fracture is good with MFS for all
loading angles. The agreement between experiments and numerical simulations is enhanced
for the loading angles 0 shear, 10 tension (Fig. 7.19-(b)) and 60 tension (Fig. 7.19-(d))
with the MFS.

7.2 Simulation with mathematical fracture strain surface

103

Torsion tests
Torsion tests are not used in the calibration of the fracture strain surfaces (AFS and MFS).
The tests are used for the validation of the presented approach in this thesis. The numerical
simulations were run with the calibrated stress-strain curve obtained in section 6.3 and the
assumption of nonlinear damage accumulation n = 2 in order to be consistent with the other
numerical simulations in this chapter. The mesh size is chosen as 0.04mm at the critical location, which corresponds to the outer surface of the middle section. The numerical simulations
are run with both fracture strain surfaces AFS and MFS.
The geometry of the specimens and FE-model are shown in Fig. 7.20.

(a) Dimensions

(b) FE-modell

1.8

R2

Fixed

Figure 7.20: (a) Geometry and dimensions of the torsion specimens. (b) FE-model and applied
boundary conditions.

In the experiments the moment M and rotation angle of the specimens were recorded.
The maximum shear stress max on the specimens is calculated through

max =

2M
,
R3

(7.2)

where R is the radius at the center of the cross-section (In Fig. 7.20-(a) R=0.5mm) of the
circular specimen. The experimental and numerical max - responses are shown in Fig. 7.21.
A good correlation between experiments and numerical simulations in terms of shear stress
level is achieved. The deviation of rotation angle at fracture initiation between experiments
and numerical simulations with AFS and MFS are 11% and 19%, respectively.

104

7 Determination of Fracture Locus


0.8

max (Gpa)

0.6

0.4
Experiments 1-4
Simulation AFS
Simulation MFS

0.2

0.5

1.5

2.5

3.5

(rad)

Figure 7.21: Comparison of the max curves from experiments and numerical simulations
for the torsion specimens.

7.3

Discussion

Two dierent approaches are introduced to dene a stress triaxiality and Lode angle parameter dependent fracture strain surface. The analytical fracture strain surface has the exponential behavior in the stress triaxiality direction and second order polynomial in the Lode
angle parameter direction. In the second approach a dierentiable, continuous mathematical
fracture strain surface is generated on the data points obtained from numerical simulations
of tests.
It is clear that the analytical generated fracture strain surface is a compromise for all data
points acquired from tests. It is also observed that the error in local fracture strains is not
reected directly to the global response of the specimens, since the plastic strain evolution
at critical locations has a nonlinear behavior with respect to global displacement. The error
of local fracture strains is higher than the error in global displacement at fracture initiation,
which is the result of high material ductility.
It is shown that, with the mathematical surface denition for all specimens a good correlation is obtained between numerical simulations and experiments in terms of global displacements at fracture initiation. However this method requires many experiments and assumptions. In lack of experimental data the usage is limited. Besides the accordance between
the experiments and numerical simulations with analytical fracture surface in terms of global
displacement at fracture is also promising.

Chapter 8
Conclusions and Future Research
8.1

Conclusions of the present thesis

The aim of this thesis is the investigation of the inuence of stress state on damage modeling
with the focus on the Lode angle parameter. The stress state at a material point is dened
by two stress state parameters, stress triaxiality and Lode angle parameter uniquely. The
damage model GISSMO proposed by Neukamm et al. [1, 2] is extended to incorporate the
Lode angle dependence for the 3D-case. The subject damage modeling is divided into
two parts; damage plasticity and material ductility (fracture strain). The inuence of the
stress triaxiality and Lode angle parameter is veried through an experimental program. The
numerical simulations were run with ne mesh size. In the current thesis, DP600 from dual
phase steel groups is investigated. The important conclusions and ndings that fall out from
the prior chapters are summarized below.
Stress state dependence of plasticity model : The damage is usually the ultimate
result of the loss of load carrying capacity at local material points with high stress and
strain gradients. Therefore the accuracy of the plasticity model is the key factor in numerical
damage modeling. The experimental results taken from the literature indicate that stress
state sensitivity of the material plasticity diers depending on the material groups. It was
concluded that the plastic ow for the aluminum alloys may be especially sensitive to the
Lode angle parameter, whereas for steels the stress state sensitivity of plasticity is negligible.
For the investigated material, the stress state dependence of plastic ow is not signicant. It
has been shown that in numerical simulations and experiments a good correlation is obtained
in terms of force-displacement responses with the J2 -plasticity.
Coupling of damage with plasticity model : The stress-strain curve used in numerical simulations was calculated with the assumption that damaged material and undamaged
material matrix can not be separated. The damage does not inuence the plastic ow. It has
been shown that in numerical simulations with uncoupled damage and plasticity formulation
a good correlation in terms of force-displacement curves can be obtained for the tests with

106

8 Conclusions and Future Research

dierent stress states with this assumption.


Stress state dependence of material ductility : Stress triaxiality dependence of the
material ductility has been shown by many researchers, whereas the inuence of the Lode
angle parameter usually has been neglected. In recent years, the inuence was pointed out
by some researchers. In the current study material ductility (fracture strain) is modeled
as a function of two independent stress state parameters, stress triaxiality and Lode angle
parameter. It has been found out that for the investigated material DP600 both stress state
parameters inuence the material ductility.
Determination of fracture strain and stress state parameters : A hybrid approach
combining experimental and numerical results is presented. A wide range on the stress triaxiality and Lode angle parameter plane is covered with the proposed experimental program.
The stress triaxiality and plastic strain are assumed as controlling quantities of fracture. For
most of the tests the location with highest stress triaxiality and plastic strain is coincident
in the specimens. In order to obtain representative stress state parameters, weighting functions depending on the nonlinear damage increment rule in GISSMO damage model were
introduced.
Fracture locus determination : For the fracture locus determination two dierent
approaches are introduced. In the rst approach, a nine-parameter analytical fracture strain
denition, which is a function of stress triaxiality and Lode angle parameter, is proposed.
The proposed analytical fracture strain surface is calibrated based on the representative
stress state parameters and fracture strain obtained from numerical simulations of the tests.
A good correlation between experiments and numerical results is achieved. In the second
approach, the fracture strain is generated as a mathematical surface on the plane of stress
triaxiality and Lode angle parameter based on biharmonic spline method. The correlation
between experimental and numerical results is further increased with the second approach.
However, the second approach requires some assumptions and a relatively large number of
experiments. It is concluded that the analytical fracture strain surface denition is more
appropriate for engineering applications since it delivers good results and can be calibrated
with less numbers of experiments.
Butterfly specimens : The buttery specimens rstly used by Bai and Wierzbicki [8]
are optimized according to investigated material. It was shown that it is possible to investigate
the material ductility for dierent stress states with a single specimen geometry by changing
the loading angle. At low stress triaxialities the obtained stress state at critical location of
the specimen is plane stress, which is also relevant for the 2D applications such as thin sheet
metal components in crashworthiness simulations.

8.2 Future research

8.2

107

Future research

Regularization : The current research has done with very ne discretization (0.050.2mm). On the other hand, the mesh size used in the current crashworthiness simulations are
relatively limited (2-10mm). The mesh size has a signicant inuence for post-critical response
especially for the coarse mesh sizes which are used in current crashworthiness simulations.
Therefore further eort for the subject regularization must be spent.
Strain rate influence: The presented thesis considers the quasi-static loading case. In
crashworthiness simulations the components are subjected to strain rates ranging from quasistatic case to strain rates up to 150 (1/s). Thus numerical modeling of strain rate inuence
on the fracture strain is another topic for future investigation.
Determination of damage exponent: Some methods used in damage exponent determination and their diculties are pointed out. In the present research, nonlinear damage
accumulation (n=2) with respect to equivalent plastic strain is considered. Well designed
multistep experiments with constant evolution of stress state parameters at crack initiation
points or new experimental techniques to determine damage exponent are of great interest.

108

8 Conclusions and Future Research

Appendix A
Derivation of stress state dependent
fracture function
The stress triaxiality and Lode angle parameter dependent fracture locus is derived. Three
bound curves (Fig. A.1) axisymmetric deviatoric tension = 1, plane strain = 0 and
axisymmetric deviatoric compression = 1 are dened as reduced form of the JC equation
(Eq. 3.3).

f = D1 + D2 exp(D3 ),

(A.1)

0f = D10 + D20 exp(D30 ),

(A.2)

+
+
+
+
f = D1 + D2 exp(D3 ),

(A.3)

where D1 , D2 , D3 are JC parameteres for = 1, D10 , D20 , D30 are JC parameteres for = 0
and D1+ , D2+ and D3+ are JC parameteres for = 1.
The dependence of the Lode angle parameter is described with a second order polynomial
functional and the fracture locus is dened as
2

f (, ) = a() + b() + c().

(A.4)

The functional coecents a, b and c are functions of stress triaxiality and can be specied
+
0
with three limiting bounds
f , f and f . = 0 yields that
f (, = 0) = c = 0f .

(A.5)

By considering other two bound curves f (, = 1) = +


f and f (, = 1) = f , other two

functional coeecens a and b are found


a=

)
1( +
0
f +
f f ,
2

(A.6)

110

A Derivation of stress state dependent fracture function

+
+
+
+
f = D1 + D2 exp(D3 )

Fracture strain
0f = D10 + D20 exp(D30 )

f = D1 + D2 exp(D3 )

-1

Figure A.1: Proposed fracture strain locus and bound curve formulations.
)
1( +
f
(A.7)
f .
2
Substituting the Eq. A.6, Eq. A.7 and Eq. A.5 into Eq. A.4, one gets the fracture surface
denition with 9 parameters
[
]
)
)
1(
1(
+
0

+
+
0
0
f (, ) =
D1 + D1 D1 +
D2 exp(D3 ) + D2 exp(D3 ) D2 exp(D3 ) 2
2
2
[
]
) 1( +
)
1( +

+
D1 D1 +
D2 exp(D3 ) D2 exp(D3 )
2
2
+ D10 + D20 exp(D30 ).
(A.8)
b=

Bibliography
[1] F. Neukamm, M. Feucht, and A. Haufe. Considering damage history in crashworthiness
simulations. In The 7th European LS-DYNA Conference, Salzburg, Austria, 2009.
[2] F. Neukamm, M. Feucht, and A. Haufe. Consistent damage modelling in the process chain of forming to crashworthiness simulations. In The 7th German LS-DYNA
Conference, Bamberg, Germany, volume 30.
[3] F. Neukamm, M. Feucht, and M. Bischo. On the application of continuum damage
models to sheet metal forming simulations. In Proceedings, X International Conference
on Computational Plasticity. CIMNE, Barcelona, Spain, 2009.
[4] F. Neukamm. Lokalisierung und Versagen von Blechwerkstoffen. PhD thesis, University
of Stuttgart, in preperation.
[5] J.D. Seidt. Plastic deformation and ductile fracture of 2024-T351 aluminum under
various loading conditions. PhD thesis, Ohio State University, 2010.
[6] X. Gao, T. Zhang, M. Hayden, and C. Roe. Eects of the stress state on plasticity and
ductile failure of an aluminum 5083 alloy. Int. J. Plast., 25(12):23662382, 2009.
[7] T. Wierzbicki and L. Xue. On the eect of the third invariant of the stress deviator
on ductile fracture. Technical report, Impact and Crashworthiness Laboratory, Massachusetts Institute of Technology, Cambridge, MA (USA), 2005.
[8] Y. Bai and T. Wierzbicki. A new model of metal plasticity and fracture with pressure
and lode dependence. Int. J. Plast., 24(6):1071 1096, 2008.
[9] W.H. Tai. Plastic damage and ductile fracture in mild steels. Eng. Fract. Mech.,
37(4):853 880, 1990.
[10] C.C. Tasan. Micro-mechanical characterization of ductile damage in sheet metal. PhD
thesis, Eindhoven University of Technology, 2010.
[11] J. Lemaitre. A continuous damage mechanics model for ductile fracture. J. Eng. Mater.
Technol., 107(1):8389, 1985.

112

BIBLIOGRAPHY

[12] S.P. Timoshenko. History of Strength of Materials: with a Brief Account of the History
of Theory of Elasticity and Theory of Structures. Dover Pubns, 1983.
[13] R.O. Davis and A.P.S. Selvadurai. Plasticity and Geomechanics. Cambridge Univ Pr,
2002.
[14] L. Xue. Ductile fracture modeling - theory, experimental investigation and numerical
verification. PhD thesis, Massachusetts Institute of Technology, 2007.
[15] X. Gao, G. Zhang, and C. Roe. A study on the eect of the stress state on ductile
fracture. Int. J. Damage Mech., 19(1):75, 2010.
[16] W.F. Chen, D.J. Han, and DJ Han. Plasticity for Structural Engineers. J Ross Pub,
2007.
[17] N.S. Ottosen and M. Ristinmaa. The Mechanics of Constitutive Modeling. Elsevier
Science Ltd, 2005.
[18] M. Yu. Generalized Plasticity. Springer Verlag, 2006.
[19] Y. Bai. Effect of loading history in necking and fracture. PhD thesis, Massachusetts
Institute of Technology, 2008.
[20] F.A. McClintock. A criterion for ductile fracture by the growth of holes. J. appl. Mech,
35(2):363371, 1968.
[21] J.R. Rice and D.M. Tracey. On the ductile enlargement of voids in triaxial stress elds.
J. Mech. Phys. Solids, 17(3):201 217, 1969.
[22] J.W. Hancock and A.C. Mackenzie. On the mechanisms of ductile failure in highstrength steels subjected to multi-axial stress-states. J. Mech. Phys. Solids, 24(2-3):147
160, 1976.
[23] A.C. Mackenzie, J.W. Hancock, and D.K. Brown. On the inuence of state of stress on
ductile failure initiation in high strength steels. Eng. Fract. Mech., 9(1):167168, 1977.
[24] G.R. Johnson and W.H. Cook. Fracture characteristics of three metals subjected to
various strains, strain rates, temperatures and pressures. Eng. Fract. Mech., 21(1):31
48, 1985.
[25] Y. Bao and T. Wierzbicki. A comparative study on various ductile crack formation
criteria. J. Eng. Mater. Technol., 126:314, 2004.

BIBLIOGRAPHY

113

[26] Y. Bai and T. Wierzbicki. Forming severity concept for predicting sheet necking under
complex loading histories. Int. J. Mech. Sci., 50(6):1012 1022, 2008.
[27] M. Luo and T. Wierzbicki. Numerical failure analysis of a stretch-bending test on
dual-phase steel sheets using a phenomenological fracture model. Int. J. Solids Struct.,
47(22-23):3084 3102, 2010.
[28] T. Pardoen and J.W. Hutchinson. An extended model for void growth and coalescence.
J. Mech. Phys. Solids, 48(12):2467 2512, 2000.
[29] I. Barsoum and J. Faleskog. Rupture mechanisms in combined tension and shear:
Experiments. Int. J. Solids Struct., 44(6):1768 1786, 2007.
[30] F. Bron and J. Besson. Simulation of the ductile tearing for two grades of 2024 aluminum alloy thin sheets. Eng. Fract. Mech., 73(11):1531 1552, 2006.
[31] F. Beremin. Cavity formation from inclusions in ductile fracture of A508 steel. Metall.
Mater. Trans. A, 12:723731, 1981.
[32] M. Berveiller and A. Zaoui. An extension of the self-consistent scheme to plasticallyowing polycrystals. J. Mech. Phys. Solids, 26(5-6):325 344, 1978.
[33] A. Needleman. A continuum model for void nucleation by inclusion debonding. J.
Appl. Mech., 54(3):525531, 1987.
[34] D. Steglich and W. Brocks. Micromechanical modelling of the behaviour of ductile
materials including particles. Comp. Mater. Sci., 9(1-2):7 17, 1997.
[35] M. Shabrov, C. Briant, A. Needleman, S. Kim, E. Sylven, D. Sherman, and L. Chuzhoy.
Void nucleation by inclusion cracking. Metall. Mater. Trans. A, 35:17451755, 2004.
[36] A.L. Gurson. Continuum theory of ductile rupture by void nucleation and growth,
1. Yield criteria and ow rules for porous ductile media. J. Eng. Mater-T. Asme,
99(1):215, 1977.
[37] P.F. Thomason. A three-dimensional model for ductile fracture by the growth and
coalescence of microvoids. Acta Metall., 33(6):1087 1095, 1985.
[38] A.A. Benzerga. Micromechanics of coalescence in ductile fracture. J. Mech. Phys.
Solids, 50(6):1331 1362, 2002.
[39] L.M. Brown and J.D. Embury. Initiation and growth of voids at second-phase particles. In Proc. Conf. on Microstructure and Design of Alloys, volume 1, pages 164169.
Institute of Metals and Iron and Steel Insitute, London, 1973.

114

BIBLIOGRAPHY

[40] V. Tvergaard. Inuence of voids on shear band instabilities under plane strain conditions. Int. J. Fract., 17:389407, 1981.
[41] V. Tvergaard. On localization in ductile materials containing spherical voids. Int. J.
Fract., 18(4):237252, 1982.
[42] V. Tvergaard and A. Needleman. Analysis of the cup-cone fracture in a round tensile
bar. Acta Metall., 32(1):157 169, 1984.
[43] C.C. Chu and A. Needleman. Void nucleation eects in biaxially stretched sheets. J.
Eng. Mater. Technol., 102(3):249256, 1980.
[44] L. Xue. Constitutive modeling of void shearing eect in ductile fracture of porous
materials. Eng. Fract. Mech., 75(11):3343 3366, 2008.
[45] K. Nahshon and J.W. Hutchinson. Modication of the Gurson model for shear failure.
Eur. J. Mech. A. Solids, 27(1):1 17, 2008.
[46] M. Gologanu, J.-B. Leblond, and J.Devaux. Approximate models for ductile metals
containing non-spherical voids - case of axisymmetric prolate ellipsoidal cavities. J.
Mech. Phys. Solids, 41(11):1723 1754, 1993.
[47] D. Lassance, D. Fabregue, F. Delannay, and T. Pardoen. Micromechanics of room and
high temperature fracture in 6xxx al alloys. Prog. Mater Sci., 52(1):62 129, 2007.
[48] M.L. Wilkins, R.D. Streit, and J.E. Reaugh. Cumulative-strain-damage model of ductile
fracture: simulation and prediction of engineering fracture tests. Technical report,
UCRL-53058, Lawrence Livermore National Lab., CA (USA); Science Applications,
Inc., San Leandro, CA (USA), 1980.
[49] M. Kachanov. Rupture time under creep conditions. Int. J. Fract., 97:1118, 1999.
[50] Y.N. Rabotnov. On the equation of state of creep. In ARCHIVE: Proceedings of the
Institution of Mechanical Engineers, Conference Proceedings 1964-1970 (vols 178-184),
Various titles labelled Volumes A to S, volume 178, pages 117122. Prof Eng Publishing,
1963.
[51] N. Bonora. A nonlinear CDM model for ductile failure. Eng. Fract. Mech., 58(1-2):11
28, 1997.
[52] T. Brvik, O.S. Hopperstad, S. Dey, E.V. Pizzinato, M. Langseth, and C. Albertini.
Strength and ductility of Weldox 460 E steel at high strain rates, elevated temperatures
and various stress triaxialities. Eng. Fract. Mech., 72(7):1071 1087, 2005.

BIBLIOGRAPHY

115

[53] L. Xue and T. Wierzbicki. Ductile fracture initiation and propagation modeling using
damage plasticity theory. Eng. Fract. Mech., 75(11):3276 3293, 2008.
[54] O. Mohr. Abhandlungen aus dem Gebiete der technischen Mechanik; mit zahlreichen
Textabbildungen. Wilhelm Ernst & Sohn, 1906.
[55] D.C. Drucker and W. Prager. Soil mechanics and plastic analysis or limit design. Q.
Appl. Math., 10(2):157165, 1952.
[56] S.S. Hsieh, E.C. Ting, and W.F. Chen. A plasticity-fracture model for concrete. Int.
J. Solids Struct., 18(3):181197, 1982.
[57] K.J. Willam and E.P. Warnke. Constitutive model for the triaxial behavior of concrete. In Proceedings, International Association for Bridge and Structural Engineering,
volume 19, pages 130. ISMES, Bergamo, Italy, 1975.
[58] N.S. Ottosen. Constitutive model for short-time loading of concrete. J. Eng. Mech.
Div., 105(1):127141, 1979.
[59] R. Hill. On discontinuous plastic states, with special reference to localized necking in
thin sheets. J. Mech. Phys. Solids, 1(1):19 30, 1952.
[60] F. Barlat and K. Lian. Plastic behavior and stretchability of sheet metals. Part I:
A yield function for orthotropic sheets under plane stress conditions. Int. J. Plast.,
5(1):51 66, 1989.
[61] A.P. Karallis and M.C. Boyce. A general anisotropic yield criterion using bounds and
a transformation weighting tensor. J. Mech. Phys. Solids, 41(12):1859 1886, 1993.
[62] H. Vegter and A.H. van den Boogaard. A plane stress yield function for anisotropic
sheet material by interpolation of biaxial stress states. Int. J. Plast., 22(3):557 580,
2006.
[63] H. Ziegler. A modication of Pragers hardening rule. Quart. Appl. Math, 17(1):5565,
1959.
[64] J.L. Chaboche and G. Rousselier. On the plastic and viscoplastic constitutive equations
- Part I: Rules developed with internal variable concept. J. Pressure Vessel Technol.,
105(2):153158, 1983.
[65] K. Chung, M.-G. Lee, D. Kim, C. Kim, M.L. Wenner, and F. Barlat. Spring-back
evaluation of automotive sheets based on isotropic-kinematic hardening laws and nonquadratic anisotropic yield functions: Part I: Theory and formulation. Int. J. Plast.,
21(5):861 882, 2005.

116

BIBLIOGRAPHY

[66] L. Geng and R.H. Wagoner. Role of plastic anisotropy and its evolution on springback.
Int. J. Mech. Sci., 44(1):123 148, 2002.
[67] P.W. Bridgman. Studies in Large Plastic Flow and Fracture. McGrraw-Hill, New York,
1952.
[68] J.J. Lewandowski and P. Lowhaphandu. Eects of hydrostatic pressure on mechanical
behaviour and deformation processing of materials. Int. Mater. Rev., 43(4):145187,
1998.
[69] I.E. French and P.F. Weinrich. The inuence of hydrostatic pressure on the tensile
deformation and fracture of copper. Metall. Mater. Trans. A, 6(4):785790, 1975.
[70] P.F. Weinrich and I.E. French. The inuence of hydrostatic pressure on the fracture
mechanisms of sheet tensile specimens of copper and brass. Acta Metall., 24(4):317
322, 1976.
[71] A. Brownrigg, W.A. Spitzig, O. Richmond, D. Teirlinck, and J.D. Embury. The inuence of hydrostatic pressure on the ow stress and ductility of a spherodized 1045 steel.
Acta Metall., 31(8):11411150, 1983.
[72] W.A. Spitzig and O. Richmond. The eect of pressure on the ow stress of metals.
Acta Metall., 32(3):457463, 1984.
[73] C.D. Wilson. A critical reexamination of classical metal plasticity. J. Appl. Mech.,
69:63, 2002.
[74] Y. Bao and T. Wierzbicki. On fracture locus in the equivalent strain and stress triaxiality space. Int. J. Mech. Sci., 46(1):8198, 2004.
[75] Y. Bai, X. Teng, and T. Wierzbicki. On the application of stress triaxiality formula for
plane strain fracture testing. J. Eng. Mater. Technol., 131(2):021002, 2009.
[76] T. Brvik, O.S. Hopperstad, and T. Berstad. On the inuence of stress triaxiality and
strain rate on the behaviour of a structural steel. Part II. Numerical study. Eur. J.
Mech. A. Solids, 22(1):1532, 2003.
[77] X. Teng and T. Wierzbicki. Evaluation of six fracture models in high velocity perforation. Eng. Fract. Mech., 73(12):16531678, 2006.
[78] T. Brvik, O.S. Hopperstad, T. Berstad, and M. Langseth. A computational model
of viscoplasticity and ductile damage growth of for impact and penetration. Eur. J.
Mech. A. Solids, 20(5):685 712, 2001.

BIBLIOGRAPHY

117

[79] I. Barsoum. The effect of stress state in ductile failure. PhD thesis, Royal Institute of
Technology, 2008.
[80] T. Coppola, L. Cortese, and P. Folgarait. The eect of stress invariants on ductile
fracture limit in steels. Eng. Fract. Mech., 76(9):1288 1302, 2009.
[81] H. Mae, X. Teng, Y. Bai, and T. Wierzbiki. Correlation between tensile/ shear fracture strains and pore sizes in a cast aluminum alloyalloy. The Japanese Society for
Experimental Mechanics, 9(2):129135, 2009.
[82] X. Teng, H. Mae, Y. Bai, and T. Wierzbicki. Statistical analysis of ductile fracture
properties of an aluminum casting. Eng. Fract. Mech., 75(15):46104625, 2008.
[83] L. Xue. Stress based fracture envelope for damage plastic solids. Eng. Fract. Mech.,
76(3):419438, 2009.
[84] H. Huang and L. Xue. Prediction of slant ductile fracture using damage plasticity
theory. Int. J. Press. Vessels Pip., 86(5):319328, 2009.
[85] A. Kamoulakos, P. Culiere, and T. Araki. Prediction of ductile metal rupture with
the EW model in PAM-CRASH. In Proceedings of the International Body Engineering
Conference, Chiba, 2003.
[86] I.E. French and P.F. Weinrich. The eect of hydrostatic pressure on the tensile fracture
of [alpha]-brass. Acta Metall., 21(11):1533 1537, 1973.
[87] I.E. French and P.F. Weinrich. The unusual variation of the tensile fracture strain of
brasses at high pressures. Scr. Metall., 8(1):7 9, 1974.
[88] A. Brownrigg, W.A. Spitzig, O. Richmond, D. Teirlinck, and J.D. Embury. The inuence of hydrostatic pressure on the ow stress and ductility of a spherodized 1045 steel.
Acta Metall., 31(8):1141 1150, 1983.
[89] A.S. Kao, H. Kuhn, O. Richmond, and W.A. Spitzig. Tensile fracture and fractographic
analysis of 1045 spheroidized steel under hydrostatic pressure. J. Mater. Res., 5(1):83
91, 1990.
[90] W.A. Spitzig. Eect of hydrostatic pressure on deformation, damage evolution, and
fracture of iron with various initial porosities. Acta Metall. Mater., 38(8):1445 1453,
1990.
[91] I. Barsoum and J. Faleskog. Rupture mechanisms in combined tension and shear:
Micromechanics. Int. J. Solids Struct., 44(17):5481 5498, 2007.

118

BIBLIOGRAPHY

[92] M. Alves and N. Jones. Inuence of hydrostatic stress on failure of axisymmetric


notched specimens. J. Mech. Phys. Solids, 47(3):643 667, 1999.
[93] G. Mirone. Role of stress triaxiality in elastoplastic characterization and ductile failure
prediction. Eng. Fract. Mech., 74(8):1203 1221, 2007.
[94] H. Li, M.W. Fu, J. Lu, and H. Yang. Ductile fracture: Experiments and computations.
Int. J. Plast., 27(2):147 180, 2011.
[95] Y. Bao. Dependence of ductile crack formation in tensile tests on stress triaxiality,
stress and strain ratios. Eng. Fract. Mech., 72(4):505 522, 2005.
[96] J.W. Hancock and D.K. Brown. On the role of strain and stress state in ductile failure.
J. Mech. Phys. Solids, 31(1):1 24, 1983.
[97] G. Mirone and D. Corallo. A local viewpoint for evaluating the inuence of stress
triaxiality and lode angle on ductile failure and hardening. Int. J. Plast., 26(3):348
371, 2010.
[98] D.J. Han and W.F. Chen. A nonuniform hardening plasticity model for concrete materials. Mech. Mater., 4(3-4):283 302, 1985.
[99] J.P. Bardet. Lode dependences for isotropic pressure-sensitive elastoplastic materials.
J. Appl. Mech., 57(3):498506, 1990.
[100] T. Nakai, H. Matsuoka, N. Okuno, and K. Tsuzuki. True triaxial tests on normally
consolidated clay and analysis of the observed shear behaviour using elastoplastic constitutive models. Soils Found., 26(4):6778, 1986.
[101] K.S. Zhang, J.B. Bai, and D. Franois. Numerical analysis of the inuence of the lode
parameter on void growth. Int. J. Solids Struct., 38(32-33):5847 5856, 2001.
[102] X. Gao and J. Kim. Modeling of ductile fracture: Signicance of void coalescence. Int.
J. Solids Struct., 43(20):6277 6293, 2006.
[103] J. Wan, Z. Lu, and Z. Yue. Growth of casting microcrack and micropore in singlecrystal superalloys analysed by three-dimensional unit cell. J. Mater. Sci. Technol,
22(2):183, 2006.
[104] S. Murakami and N. Ohno. A continuum theory of creep and creep damage. In Creep
in Structures, 3 rd IUTAM Symposium; Leicester, pages 422444. Springer-Verlag, Heidelberger Platz 3, D-1000 Berlin 33, 1980.

BIBLIOGRAPHY

119

[105] J.L. Chaboche. Anisotropic creep damage in the framework of continuum damage
mechanics. Nucl. Eng. Des., 79(3):309 319, 1984.
[106] M. Kachanov. Continuum model of medium with cracks.
106(5):10391051, 1980.

J. Eng. Mech. Div.,

[107] D. Krajcinovic. Constitutive equations for damaging materials.


50(2):355360, 1983.

J. Appl. Mech.,

[108] C.L. Chow and June Wang. An anisotropic theory of continuum damage mechanics for
ductile fracture. Eng. Fract. Mech., 27(5):547 558, 1987.
[109] S. Murakami. Mechanical modeling of material damage. J. Appl. Mech., 55(2):280286,
1988.
[110] Y. Hammi, D.J. Bammann, and M.F. Horstemeyer. Modeling of anisotropic damage for
ductile materials in metal forming processes. Int. J. Damage Mech., 13(2):123, 2004.
[111] M. Br
unig. Numerical analysis of anisotropic ductile continuum damage. Comput.
Meth. Appl. Mech. Eng., 192(26-27):2949 2976, 2003.
[112] A.M. Habraken, J.F. Charles, and S. Cescotto.

Calibration and validation of an

anisotropic elasto-plastic damage model for sheet metal forming. In Jiann-Wen


Woody Ju George Z. Voyiadjis and Jean-Louis Chaboche, editors, Damage Mechanics in Engineering Materials, volume 46 of Studies in Applied Mechanics, pages 401
420. Elsevier, 1998.
[113] J.-F. Charles, Y.Y. Zhu, A.-M. Habraken, S. Cescotto, and M. Traversin. A fully
coupled elasto-plastic damage theory for anisotropic materials. In M. Predeleanu and
P. Gilormini, editors, Advanced Methods in Materials Processing Defects, volume 45 of
Studies in Applied Mechanics, pages 33 42. Elsevier, 1997.
[114] J.L. Chaboche. Continuum damage mechanics: Part I - general concepts. J. Appl.
Mech., 55(1):5964, 1988.
[115] J. Lemaitre. How to use damage mechanics. Nucl. Eng. Des., 80(2):233 245, 1984.
[116] N. Bonora, A. Ruggiero, L. Esposito, and D. Gentile. CDM modeling of ductile failure
in ferritic steels: Assessment of the geometry transferability of model parameters. Int.
J. Plast., 22(11):2015 2047, 2006.
[117] J. Lemaitre and J.L. Chaboche. Mechanics of Solid Materials. Cambridge Univ Pr,
1994.

120

BIBLIOGRAPHY

[118] M. Alves, J. Yu, and N. Jones. On the elastic modulus degradation in continuum
damage mechanics. Comput. Struct., 76(6):703 712, 2000.
[119] N. Bonora, D. Gentile, A. Pirondi, and G. Newaz. Ductile damage evolution under
triaxial state of stress: Theory and experiments. Int. J. Plast., 21(5):981 1007, 2005.
[120] A. Weck, D.S. Wilkinson, E. Maire, and H. Toda. Visualization by X-ray tomography
of void growth and coalescence leading to fracture in model materials. Acta Mater.,
56(12):2919 2928, 2008.
[121] C.C. Tasan, J.P.M. Hoefnagels, and M.G.D. Geers. Indentation-based damage quantication revisited. Scr. Mater., 63(3):316 319, 2010.
[122] J.P.M. Hoefnagels, C.C. Tasan, M. Pradelle, and M.G.D. Geers. Brittle fracture-based
experimental methodology for microstructure analysis. Appl. Mech. Mater., 13:133
139, 2008.
[123] R. Chaouadi, P. Meester, and W. Vandermeulen. Damage work as ductile fracture
criterion. Int. J. Fract., 66:155164, 1994.
[124] M.S. Mirza, D.C. Barton, and P. Church. The eect of stress triaxiality and strain-rate
on the fracture characteristics of ductile metals. J. Mater. Sci., 31:453461, 1996.
[125] X. Teng, T. Wierzbicki, and M. Huang. Ballistic resistance of double-layered armor
plates. Int. J. Impact Eng., 35(8):870 884, 2008.
[126] Y. Bai, Y. Bao, and T. Wierzbicki. Fracture of prismatic aluminum tubes under reverse
straining. Int. J. Impact Eng., 32(5):671 701, 2006.
[127] A. Weck, D.S. Wilkinson, H. Toda, and E. Maire. 2D and 3D visualization of ductile
fracture. Adv. Eng. Mater., 8(6):469472, 2006.
[128] A. Weck and D.S. Wilkinson. Experimental investigation of void coalescence in metallic
sheets containing laser drilled holes. Acta Mater., 56(8):1774 1784, 2008.
[129] L. Xue. Damage accumulation and fracture initiation in uncracked ductile solids subject
to triaxial loading. Int. J. Solids Struct., 44(16):5163 5181, 2007.
[130] R. De Borst, L.J. Sluys, H.B. Muhlhaus, and J. Pamin. Fundamental issues in nite
element analyses of localization of deformation. Eng. Computation., 10(2):99121, 1993.
[131] J.O. Hallquist et al. LS-DYNA Keyword Manual 971. Livermore Software Technology
Corpration (LSTC), P. O. Box 712 Livermore, California 94551-0712, rev 5 edition,
May 2010.

BIBLIOGRAPHY

121

[132] International Iron and Steel Institute Committee on Automotive Applications. Advanced High Strength Steel (AHSS) Applicaiton Guidelines. World Steel association,
Brussels, rev 3 edition, 2006.
[133] A. Pineau and T. Pardoen. Failure of metals. In I. Milne, R. O. Ritchie, , and B. Karihaloo, editors, Comprehensive Structural Integrity, pages 684 797. Pergamon, Oxford,
2007.
[134] J. Besson, D. Steglich, and W. Brocks. Modeling of plane strain ductile rupture. Int.
J. Plast., 19(10):1517 1541, 2003.
[135] F.A. McClintock and Z.M. Zheng. Ductile fracture in sheets under transverse strain
gradients. Int. J. Fract., 64:321337, 1993.
[136] S. Sommer and D.-Z. Sun. Charakterisierung und Modellierung der Tragfahigkeit von
punktgeschweiten Stahblechlverbindungen unter Crashbelastung mit Hilfe von erweiterten Schadigungsmodellen. Technical report, Fraunhofer IWM, 2006.
[137] C.L. Walters. Development of a punching technique for ductile fracture testing over
a wide range of stress states and strain rates. PhD thesis, Massachusetts Institute of
Technology, 2009.
[138] D. Mohr and R. Treitler. Onset of fracture in high pressure die casting aluminum alloys.
Eng. Fract. Mech., 75(1):97 116, 2008.
[139] A. Pirondi and N. Bonora. Modeling ductile damage under fully reversed cycling.
Comp. Mater. Sci., 26:129 141, 2003.
[140] M. Considere. Memoire Sur Lemploi Du Fer et de lacier dans les constructions.
Dunod, 1885.
[141] I. Scheider, W. Brocks, and A. Cornec. Procedure for the determination of true stressstrain curves from tensile tests with rectangular cross-section specimens. J. Eng. Mater.
Technol., 126(1):7076, 2004.
[142] J. Choung and S. Cho. Study on true stress correction from tensile tests. J. Mech. Sci.
Technol., 22:10391051, 2008.
[143] P. Ludwik. Elemente der technischen Mechanik. Julius Springer, Pawtucket, RI, USA,
page 32, 1909.
[144] Y. Bao and T. Wierzbicki. On the cut-o value of negative triaxiality for fracture. Eng.
Fract. Mech., 72(7):1049 1069, 2005.

122

BIBLIOGRAPHY

[145] W. Schweizer. MATLAB kompakt. Oldenbourg Wissenschaftsverlag, 2009.


[146] D.T. Sandwell. Biharmonic spline interpolation of GEOS-3 and SEASAT altimeter
data. Geophys. Res. Lett., 14(2):139142, 1987.

Curriculum Vitae
Personal Data
Name
Date of Birth
Nationality

Merdan Basaran
January 02, 1981
Turkish

School Education
1987-1991
1991-1992
1992-1996
1996-1999

sbilen Primary School, Tekirdag


Ali Rza I
Hanife Sek Celep Primary School, Tekirdag
Mehmet Akif Ersoy Anatolian High School, Tekirdag
Atat
urk Science High School, Istanbul

University Education
2000-2004
2004-2008
2008-2011

Undergraduate Studies of Mechanical Engineering,


Istanbul Technical University, Istanbul
M.Sc. in Simulation Techniques in Mechanical Engineering,
RWTH Aachen University, Aachen
Doctoral Candidate, Institute of General Mechanics(IAM)
RWTH Aachen University, Aachen
DAIMLER AG, Sindelngen

Professional Carrier
11/2005-03/2007
since 06/2011

Student Assistent, FEV Motorentechnik GmbH, Aachen


CAE Enginner, DAIMLER AG, Sindelngen

Você também pode gostar