You are on page 1of 143

DEVELOPMENT OF PLASTICITY AND DUCTILE FRACTURE MODELS

INVOLVING THREE STRESS INVARIANTS

A Dissertation
Presented to
The Graduate Faculty of The University of Akron

In Partial Fulfillment
of the Requirements for the Degree
Doctor of Philosophy

Tingting Zhang
May, 2012

DEVELOPMENT OF PLASTICITY AND DUCTILE FRACTURE MODELS


INVOLVING THREE STRESS INVARIANTS

Tingting Zhang
Dissertation

Approved:

Accepted:

__________________________
Advisor
Xiaosheng Gao

____________________________
Department Chair
Celal Batur

__________________________
Committee Member
Fred Choy

____________________________
Dean of the College
George K. Haritos

__________________________
Committee Member
Gregory Morscher

____________________________
Dean of the Graduate School
George R. Newkome

__________________________
Committee Member
Ernian Pan

____________________________
Date

__________________________
Committee Member
Kevin Kreider

ii

ABSTRACT

It has been shown that the plastic response of many materials, including some
metallic alloys, depends on the stress state. Based on plasticity analysis of three metal
alloys, a series of new plasticity models with stress state effect is proposed. The effect of
stress state on plasticity and the general forms of the yield function and flow potential for
isotropic materials are assumed to be functions of the first invariant of the stress tensor
(I1 ) and the second and third invariants of the deviatoric stress tensor (J2 and J3 ). Finite
element implementation, including integration of the constitutive equations using the
backward Euler method and formulation of the consistent tangent moduli, are presented
in this thesis.
A 5083 aluminum alloy, Nitronic 40 (a stainless steel), and Zircaloy-4 (a
zirconium alloy) were tested under tension, compression, torsion, combined torsiontension and combined torsion-compression at room temperature to demonstrate the
applicability of

proposed I1 -J2 -J3 dependent models. It has shown that the output

produced by the proposed model have better agreement with experimental data than those
produced by the classical J2 plasticity theory for the tested loading conditions and
materials.
Furthermore, the Gurson-Tvergaard-Needleman porous plasticity model, which is
widely used to simulate the void growth process of ductile fracture, is extended to include
iii

the effects of hydrostatic stress and the third invariant of stress deviator o n the matrix
material.
The experimental and numerical work presented in this thesis reveals that the
stress state also has strong effects on the ductile fracture behavior of an aluminum 5083
alloy. For the ductile fracture analysis, The Goluganu-Leblond-Devaux (GLD) model is
employed to describe the porous plasticity behavior of aluminum 5083. The effect of
stress triaxiality and Lode angle is analyzed and fracture locus is calibrated as a criterion
for void coalescence. The GLD model combined with the fracture locus can be applied to
predict the failure of aluminum 5083 specimens with experiencing a large range of stress
triaxiality and Lode angle. The numerical analyses agree with the experimental data very
well.

iv

ACKNOWLEDGEMENTS

I would like to take this opportunity to express my sincere gratitude to my advisor


Dr. Xiaosheng Gao for his encouragements, consistent help and, most important, his
valuable instructions in the academic area. The accomplishment of this degree would be
impossible without his support.
I would also like to show my thanks to my committee members: Dr. Fred Choy,
Dr. Gregory Morscher from Mechanical Engineering Department, Dr. Ernian Pan from
Civil Engineering Department, and Dr. Kevin Kreider from Mathematics Department.
Special appreciation is given to them for their time on reading and evaluating my thesis.
Their comments and suggestions have improved the quality of my research work.
I would also like to thank my colleagues, including Jun Zhou, Sunil Prakash,
Yifei Gao and others. Discussions on the academic issues with them always enlightened
me to approach the research objectives. Special Thanks are given to Stacy and Christina,
the secretaries of Mechanical Engineering Department, for taking care of my
administrative issues during the last five years.
Last but not least, I want to thank my parents for their endless love and support.
During the time when I was pursuing my Ph.D. degree, my maternal grandfather passed
away. If there were no consolation or encouragement from my parents on the overseas
v

calls, I would be struck down by sorrow and sadness. I love them more and more as time
goes by.

vi

TABLE OF CONTENTS

Page
LIST OF TABLES ............................................................................................................... x
LIST OF FIGURES ........................................................................................................... xii
CHAPTER
I. INTRODUCTION .............................................................................................................2

1.1 Development of plasticity models............................................................................. 2


1.2 Research objectives on plasticity modeling .............................................................. 3
1.3 Stress state effects on ductile fracture ....................................................................... 4
1.4 Research objectives on the ductile fracture study ..................................................... 6
1.5 Outline of the Thesis ................................................................................................. 7
II. LITERATURE REVIEW .................................................................................................9

2.1 Basic Concepts and Definitions ................................................................................ 9


2.2 Review of Plasticity Models.................................................................................... 13
2.3 Micro Mechanics Models ........................................................................................ 18
2.4 Non-associated Flow Rule ...................................................................................... 24
III. MODELING OF PLASTICITY RESPONSES:

INVOLING THREE STRESS

INVARIANTS ...................................................................................................................26
3.1 Introduction ............................................................................................................. 26
vii

3.2 Plasticity Modeling ................................................................................................. 28


3.2.1 Influence of I1 and J3 ........................................................................................ 28
3.2.2 Yield Function and Flow Potential for 5083 Aluminum Alloy and Zircaloy-4 32
3.2.3 Yield Function and Flow Potential for Nitronic 40 Stainless Steel .................. 33
3.3 Numerical Implementation of the I1 -J2 -J3 plasticity model .................................... 34
3.3.1 Stress Update .................................................................................................... 34
3.3.2 The Consistent Tangent Moduli ....................................................................... 36
3.4 Applications of the stress-state dependent plasticity models .................................. 39
3.4.1 Designing of torsion-tension specimen ............................................................ 39
3.4.2 Materials, Specimens and Experiments ............................................................ 52
3.4.3 Finite Element Procedure ................................................................................. 58
3.4.4 Plasticity response of an aluminum 5083 ......................................................... 61
3.4.5 Plasticity response of a Nitronic 40 .................................................................. 68
3.4.6 Plasticity response of a Zircaloy 4 .................................................................... 72
3.5 Concluding Remarks ............................................................................................... 78
IV. MODIFIED POUROUS GURSON MODEL ...............................................................80

4.1 Introduction ............................................................................................................. 80


4.2 Modified Gurson Model with I1 , J3 Effects............................................................. 81
4.2.1 Modified Gurson Model Theory....................................................................... 81
4.2.2 Numerical Implementation ............................................................................... 82
4.3 Numerical examples using the Gurson- Tvegaard-Needleman model..................... 83
4.4 Conclusion............................................................................................................... 88
V. PLASTICTY AND DUCTILE FRACTURE ANALYSIS FOR ALUMINUM 5083 ..89
viii

5.1 Introduction ............................................................................................................. 89


5.2 GLD model.............................................................................................................. 90
5.3 Specimen geometries and finite element procedures .............................................. 93
5.4 Experimental and numerical results .........................................................................97
5.4.1 Effect of the stress state on the materials plastic response .............................. 97
5.4.2 Effect of the stress state on the ductile failure strain ....................................... 101
5.5 Concluding remarks .............................................................................................. 108
VI. CONCLUDING REMARKS AND SUGGESTIONS FOR FUTURE WORK ......... 110

6.1 Conclusions and future works on plasticity modeling involving three stress
invariants ..................................................................................................................... 110
6.2 Conclusions and future works on stress state effects on Ductile Fracture ............ 111
BIBLIOGRAPHY ............................................................................................................ 114
APPENDICES ................................................................................................................. 124
APPENDIX A. EXAMPLES OF FIRST ORDER HOMOGENEOUS FUNCTIONS OF
STRESSES ....................................................................................................................... 125
APPENDIX B: GLD ........................................................................................................ 127

ix

LIST OF TABLES

Table

Page

3.1 Chemical composition (in weight percent) ................................................................53


3.2 Elastic properties........................................................................................................54
3.3 Notch radii of the notched round bars........................................................................55
3.4 Groove radii of the plane strain specimens ................................................................56
3.5 Ratios of the applied tensile displacement and applied twist angle used in the
tension-torsion tests ..........................................................................................................56

LIST OF FIGURES

Figure

page

2. 1 (a) The Haigh-Westergaard stress space, and (b) the deviatoric plane ...................... 11
3. 1 Variation of T I 1 /(3 3J 21 / 2 ) and 3 3J 3 /( 2 J 23 / 2 ) with plastic deformation in
the center element of the smooth round bar. ..................................................................... 41
3. 2 Variation of T I 1 /(3 3J 21 / 2 ) and 3 3J 3 /( 2 J 23 / 2 ) with plastic deformation in
the center element of the E-notch specimen. .................................................................... 41
3. 3 Variation of T I 1 /(3 3J 21 / 2 ) and 3 3J 3 /( 2 J 23 / 2 ) with plastic deformation in
the center element of the G-groove specimen................................................................... 42
3. 4 Dimensions of NT specimen. All dimensions in mm. ............................................. 43
3. 5 Dimensions of Lindholm specimen. All dimensions in mm. ................................... 44
3. 6 Comparison of gage sections in NT and Lindholm specimens (to scale). ................ 44
3. 7 Finite element meshes for the, (a) NT specimen, and the, (b) modified Lindholm
specimen............................................................................................................................ 46
3. 8 Through-thickness distribution of T, and p at the mid-section of the specimen for

p 0.37 and pure torsion loading, (a) NT specimen, (b) Lindholm specimen. ............. 48
3. 9 Through-thickness distribution of T, and p at the mid-section of the specimen for
effective plastic strain of 0.02, (a) NT specimen, (b) Lindholm specimen. ..................... 49
3. 10. Through-thickness distribution of T, and p at the mid-section of the specimen
for effective plastic strain of 0.20, (a) NT specimen, (b) Lindholm specimen. ............... 49
3. 11 Through-thickness distribution of T, and p at the mid-section of the specimen
for effective plastic strain of 0.44, (a) NT specimen, (b) Lindholm specimen. ................ 50
xi

3. 12 Through-thickness distribution of T, and p at the mid-section of the specimen


when p reaches 0.14, (a) NT specimen, (b) Lindholm specimen .................................... 51
3. 13 Evolution of T, Lode parameter , and with increasing effective plastic strain, (a)
NT, (b) Lindholm specimens. ........................................................................................... 52
3. 14 Sketches of a round tensile bar, a notched round bar specimen, a compression
specimen with L/D = 0.75, a grooved plan strain specimen and a torsion specimen. ...... 55
3. 15 Typical finite element meshes for (a) a notched round bar specimen, (b) a grooved
plane strain specimen, (c) a compression specimen with L/D = 0.75, and (d) a torsioncompression specimen with a central pin. ........................................................................ 60
3. 16 Comparisons of load vs. displacement and/or torque vs. twist angle responses
between the experimental data and the J2 model prediction for aluminum 5083: (a) the
round tensile specimen, (b) the compression specimen with L/D = 0.75, and (c) the
torsion specimen (Experimental data were from [111]). .................................................. 62
3. 17 (a) Projection of the yield surface of aluminum 5083 on the -plane; (b) Plot of the
flow potential; (c) The equivalent stress vs. equivalent plastic strain curve describing the
strain hardening behavior of the material. ........................................................................ 64
3. 18 Comparisons of the predicted load vs. displacement and/or toque vs. twist angle
responses using the calibrated I1 -J2-J3 plasticity model with experimental records for
aluminum 5083: (a) the round tensile specimen; (b) the compression specimen with
L/D=0.75; (c) torque vs. twist angle response of the pure torsio n specimen; (d) axial force
vs. axial displacement response of the torsion-tension specimen (TT-15); (e) torque vs.
twist angle response of the torsion-tension specimen (TT-15) (Experimental data were
from [111]) . ...................................................................................................................... 65
3. 19 Comparisons of the numerical predictions and experimental records: (a) notched
round bar (E-Notch); (b) plane strain specimen (G-Groove); (c) torque vs. twist angle
response for torsion-tension test (TT-16); (d) axial force vs. axial displacement response
for torsion-tension test (TT-16) (Experimental data were from [111]). ........................... 67
3. 20 Comparisons of the numerical predictions using the classical J2 -flow theory and the
proposed I1 -J2-J3 model for the tension-torsion test TT16 with experimental results: (a)
torque vs. twist angle; (b) axial force vs. axial displacement (Experimental data were
from [111]). ....................................................................................................................... 68
3. 21 Comparisons of load vs. displacement and/or torque vs. twist angle responses
between the experimental data and the J2 plasticity theory predictions for Nitronic 40: (a)
the tensile specimen, (b) compression specimen with L/D = 1.5, (c) axial force vs. axial
displacement response for the torsion-compression specimen, and (d) toque vs. twist
angle response for the torsion-compression specimen (Experimental data were
from[112]). ........................................................................................................................ 69
xii

3. 22 (a) Projection of the yield surface of Nitronic 40 on the -plane, and (b) the
equivalent stress vs. equivalent plastic strain curve describing the strain hardening
behavior of the material. ................................................................................................... 70
3. 23 Comparisons of the predicted load vs. displacement and/or toque vs. twist angle
responses using the calibrated I1 -J2-J3 plasticity model with experimental records for
Nitronic 40: (a) the tensile specimen, (b) the compression specimen with L/D = 1.5, (c)
axial force vs. axial displacement response for the torsion-compression specimen, and (d)
torque vs. twist angle response for the torsion-compression specimen (Experimental data
were from[112]). ............................................................................................................... 71
3. 24 Comparisons of numerical predictions using the calibrated I1 -J2-J3 plasticity model
with experimental data for Nitronic 40: (a) load vs. displacement response of the
compression specimen with L/D=0.75; (b) torque vs. twist angle response of the pure
torsion specimen (Experimental data were from[112]). ................................................... 72
3. 25 Comparisons of the experimental data and the J2 model predictions for Zircaloy: (a)
load vs. displacement response of the tensile specimen, (b) load vs. displacement
response of the compression specimen with L/D = 1.5, and (c) torque vs. twist angle
response between experimental data and the I1 -J2 model prediction for the torsion
specimen (Experimental data were from[112]) ................................................................ 74
3. 26 (a) Yield surface of Zircaloy, (b) Equivalent stress vs. equivalent plastic strain
curve describing the strain hardening behavior of the material ........................................ 76
3. 27 Comparisons of the predicted load vs. displacement and/or torque vs. twist angle
responses using the calibrated I1 -J2 -J3 plasticity model with experimental records for (a)
the tensile specimen, (b) the compression specimen with L/D = 1.5, (c) axial force vs.
axial displacement response for the torsion-compression specimen, and (d) torque vs.
twist angle response for the torsion-compression specimen (Experimental data were
from[112]). ........................................................................................................................ 77
3. 28 Comparisons of the experimental data with the numerical results computed using
the calibrated I1 -J2 -J3 plasticity model for Zircaloy: (a) load vs. displacement responses
of the compression specimen with L/D = 0.75, (b) torque vs. twist angle response of the
pure torsion specimen (Experimental data were from[112]). ........................................... 78
4. 1 A cubic element. ........................................................................................................ 83
4. 2 Comparison of the 22 / 0 vs. u 2 / D0 response and f vs. u 2 / D0 response predicted
using a1 = a2 = b1 = b2 = 0 (dotted lines) and a1 = a2 = 610-4 and b1 = b2 = 0 (solid lines).
........................................................................................................................................... 85
4. 3 Comparison of the 22 / 0 vs. u 2 / D0 response and f vs. u 2 / D0 response predicted
using a1 = a2 = b1 = b2 = 0 (dotted lines) and a1 = a2 = 0 and b1 = b2 = -60.75 (solid lines).
........................................................................................................................................... 85
xiii

4. 4 Comparison of the 22 / 0 vs. u 2 / D0 response and f vs. u 2 / D0 response predicted


using a1 = a2 = 610-4 and b1 = b2 = 0 (associated flow rule; dotted lines) and a1 = 610-4
and a2 = b1 = b2 = 0 (non-associated flow rule; solid lines).............................................. 86
4. 5 Comparison of the 22 / 0 vs. u 2 / D0 response and f vs. u 2 / D0 response predicted
using b1 = b2 = -60.75 (dotted lines) and b1 =-60.75 and b2 = 0 (solid lines), where a1 and
a2 are taken as zero............................................................................................................ 87
4. 6 Comparison of the 22 / 0 vs. u2 / D0 response and f vs. u2 / D0 response when 1 =
0.268 and 2 = 0.634 (dotted lines) with those when 1 = 0.4 and 2 = 0.4 (solid lines),
where a1 = a2 = 610-4 and b1 = b2 = -60.75..................................................................... 88
5. 1 Sketches of a smooth round bar, a notched round bar, a grooved plane strain
specimen and a torsion specimen...................................................................................... 94
5. 2 Dimensions of the specimens (unit: mm): (a) notched round bar, (b) plane strain
specimen and (c) torsion specimen. .................................................................................. 95
5. 3 Typical finite element meshes: (a) an axi-symmetric model for a notched round bar,
(b) a 1/8-symmetric model for a grooved plane strain specimen. ..................................... 97
5. 4 (a) Measured, uniaxial, engineering stress-strain curve, (b) measured shear stress vs.
shear strain curve. ............................................................................................................. 98
5. 5 Comparison of the true stress vs. true plastic strain curves (power-law) obtained
using the smooth tensile bar data and the torsion test data. ............................................ 100
5. 6 Comparison of the numerical and experimental load vs. displacement curves for (a)
the smooth tensile specimen and (b) the torsion specimen (Experimental da ta were from
[118])............................................................................................................................... 101
5. 7 (a) Variation of the critical failure strain with stress triaxiality for =1; (b) (d)
Comparison of the predicted and measured load-displacement responses for the notched
round tensile specimens having three different notch radii respectively (Experimental
data were from [118]). .................................................................................................... 103
5. 8 (a) Undeformed mesh for the E- notch specimen, (b) deformed mesh just before
failure occurs, and (c) comparison of the predicted and measured load vs. diametral
contraction response (Experimental data were from[118])............................................. 105
5. 9 (a) Variation of the critical failure strain with stress triaxiality for =0; (b) (d)
Comparison of the predicted and measured load-displacement responses for three
grooved plane strain specimens respectively (Experimental data were from[118]). ...... 106
5. 10 Variation of the ductile failure strain with the stress triaxiality for the aluminum
5083 alloy........................................................................................................................ 107
xiv

CHAPTER I
INTRODUCTION

In recent years, rapid developments in computational mechanics have enabled


engineers to be capable of analyzing complex structural components, assessing structural
reliability and optimizing structural designs. Consequently, the need for more accurate
material models become increasingly evident, particularly when minimizing design
margins becomes the approach for weight optimization or life-extension efforts. The
predictions of a numerical model can vary widely depending on the material models
employed and the numerical prediction is useless unless a proper constitutive model to
accurately describe the material behavior is provided. In this chapter, the development of
the plasticity models and ductile fracture models are briefly introduced, and the research
purpose of this thesis is illustrated. The structure of the thesis is outlined at the end of this
chapter.

1.1 Development of plasticity models


The scientific study of plasticity may justly be regarded as beginning in 1864
when Tresca published his results on punching and extrusion experiments and formulated
his famous yield criterion [1] . This yield criterion was then used by Saint-Venant [2] and
Levy [3] in their development of the theory for rigid-perfectly plastic solid. Another well
1

known yield criterion was proposed by von Mises [4] on the basis of purely mathematical
considerations. Later von Mises criterion was interpreted by Hencky [5] as plastic
yielding occurs when the elastic shear-strain energy reaches a critical value. Von Mises
also independently proposed the equations similar to Levys yield criterion for the rigidperfectly plastic materials. Other important contributions in the early development of the
plasticity theory include the works by Prandtl [6], Reuss [7], among others. Subsequently,
within the scope of elastic-plastic materials under small deformation, the notation of yield
in the stress space formulation was generalized to cover work-hardening materials and a
unified theory of plasticity began to emerge after World War II [8, 9].
To date, an overwhelming majority of the structural analyses employ the classical
J2 plasticity theory to describe the plastic response of metallic alloys. Although the J2
plasticity theory has shown great success in various applications, it has been found that
the classical J2 plasticity theory does not lead to satisfactory predictions for some
materials. It is the geomechanics community that has long recognized the so-called
pressure sensitive and Lode dependent yielding of many geomaterials and
incorporated the hydrostatic stress (pressure) and/or the Lode angle (related to the third
invariant of the stress deviator) into the yield functions of various plasticity models [1016] .
Experiments also showed that many other materials such as certain polymers,
ceramics, metallic glasses and metallic alloys exhibit pressure sensitive yielding and
plastic dilatancy [17-24]. With these experimental findings, a large amount of studies
have been concentrating on building the plasticity models with the hydrostatic pressure
2

and/or the Lode parameter effects [10, 25-32] in order to provide a better description of
the plastic responses of these materials. Recently, Chaboche [33] provided an extensive
literature review of the plasticity and viscoplasticity constitutive theories for metal alloys.
The literature review shown in the chapter II also provides the detail formulations of
some well known plasticity models that include the hydrostatic stress and/or Lode
parameter effects.

1.2 Research objectives on plasticity modeling


The main research objective on plasticity modeling area in this thesis is to build a
series of general plasticity models for isotropic materials, which are functions of the
second, third invariants of the stress deviator and the hydrostatic stress.
Then the finite element implementation of these proposed plasticity models
including integration of the constitutive equations using the backward Euler method and
the formulation of the consistent tangent moduli is presented. The derivation is based on
small-strain formulation. For finite strain plasticity, kinematic transformations are
performed first so that the constitutive equations governing finite deformation are
formulated using strains-stresses and their rates defined on an unrotated frame of
reference. Once the kinematic transformations eliminated the rotation effects on the rate
of the tensorial quantities, the stress updating procedure and the consistent tangent
stiffness formulation remain the same as those for small-strain formulation. Most
commercial finite element programs adopt this kind of treatment for finite strain plasticity
thus only small-strain formulation needs to be considered in development of a user
material subroutine.
3

As applications, the proposed plasticity models are calibrated and verified for a
5083 aluminum alloy, a Nitronic 40 and a Zircaloy 4. The detailed numerical and
experimental results are compared, and good agreement is achieved.
Furthermore, the Gurson-Tvergaard-Needleman porous plasticity model is
extended to include the effects of the hydrostatic pressure and Lode parameter in the
matrix material, and a few numerical examples are presented. The modified porous
plasticity model is expected to improve the accuracy in predicting ductile fracture process
of certain materials.

1.3 Stress state effects on ductile fracture


There is overwhelming evidence showing that ductile fracture has strong stress
state dependence. Research on the effect of the stress state on ductile fracture probably
can be traced back to the early work of Ludwik and Scheu [34], in which the authors
hypothesized that fracture of ductile metals was governed by a strength-strain curve.
They also recognized that the strength-strain curve could be obtained through testing
tensile specimens with circumferential notches of various depths and sharpness. The
exploratory work of Orowan [35] on notch brittleness was a pivotal point in the
development of the physics of ductile fracture, and since then the effect of plastic
constraint on ductile fracture has drawn much attention of the fracture mechanics
community.
The stress triaxiality parameter, defined as the ratio of the hydrostatic stress to the
equivalent stress, is often used to characterize the plastic constraint. High triaxial tension
4

subjected by the core of a plastically deformed circumferential notch explains why failure
starts at the center of the neck. The experimental work by Bridgman [36] showed the
strain to failure in a tension test could be greatly increased if the test was carried out
under pressure to reduce the stress triaxiality in the neck. Similarly, the influence of
superimposed hydrostatic pressure on the fracture mechanisms of copper, aluminum and
brass were studied by French and Weinrich [37-40]. In French and Weinrichs study, the
magnitude of the hydrostatic pressure increased the strength of the materials and changed
the fracture strain. Using axis-symmetric notched tensile specimens and flat- notched
plane strain tensile specimens, Hancock and Mackenzie [41] and Hancock and Brown
[42] demonstrated that the strain to initiate ductile fracture was a decaying function of the
stress triaxiality. A widely used ductile fracture criterion was provided by Johnson and
Cook [43] , in which a damage parameter was defined as a weighted integral with respect
to the effective strain and the integrand is the reciprocal of the effective failure strain as a
function of the stress triaxiality, strain rate and temperature. The experimental and
numerical studies of Mirza et al. [44] on the pure iron, mild steel and aluminum alloy
BS1474 over a wide range of strain rates and Bao and Wierzbicki [45] and Bao [46] on
aluminum alloy 2024 under quasi-static loading reaffirmed the strong dependence of the
equivalent strain to the crack formation on the level of stress triaxiality.
The common attribute of the above mentioned papers are that the materials
fracture is related to the stress triaxiality. What is missing is that the third invariant of
stress deviator, which is related to the Lode parameter, is not considered in the fracture
criterion. Studies by Kim et al. [47-49] and Gao et al.[50, 51] found that the Lode
parameter should be considered in order to distinguish the stress state with the same
5

triaxiality ratio, since the macroscopic stress strain response and the void growth and
coalescence behavior of a representative material volume significantly vary for each
stress state even though the triaxiality stays the same. Similarly, Wierzbiski and Xue [52]
proposed a macroscopic ductile failure criterion as a function of both the first and third
stress invariants, and the 3-D fracture locus based on this criterion was calibrated for an
aluminum alloy 2024-T351. Barsoum and Faleskog [53, 54] and Bai and Wierzbicki [55]
demonstrated that the stress triaxiality alone cannot sufficiently characterize the effect of
the stress state on the ductile fracture and the effect of the Lode angle needs to be taken
into account. More recently, Brunig et al. [56] proposed a stress state dependent damage
criterion for ductile materials based on a thermodynamically consistent continuum
damage model.

1.4 Research objectives on the ductile fracture study


In this thesis, the recent research efforts on modeling plasticity and ductile
fracture of an aluminum 5083 alloy are presented. The investigation reveals the strong
stress-state effects on the plastic response and the ductile fracture behavior. These stress
state effects can be described by using the stress triaxiality and the Lode parameter
(which is related to the third invariant of the stress deviator). In particular, it is found that
the stress triaxiality has relatively small effect on plasticity but significant effect on the
ductile failure strain. On the other hand, the effect of the Lode angle on ductile fracture is
negligible but its effect on plasticity is significant.
A porous micromechanics model, Gologanu-Leblond-Devaux (GLD) plasticity
model (Gologanu et al.,[57-59] ; Pardoen and Hutchinson [60]), and the computational
6

cell approach [61, 62] are used to perform the detailed finite element analyses of the
smooth and notched round tensile bars, grooved plane strain specimens and the
Lindholm- type torsion specimen [63]. Very good comparisons between the model
predictions and the experimental measurements are observed. The new findings of this
research challenge the classical J2 plasticity theory and provide a blueprint for
establishment of the stress-state dependent plasticity and ductile fracture models for
aluminum structural reliability assessments.

1.5 Outline of the Thesis


There are five chapters, two appendices and a bibliography of cited references
presented in the thesis.
Chapter 1 gives the history of plasticity theory and ductile fracture development,
provides the research objectivities of this thesis and presents the outline of thesis
structure.
Chapter 2 provides the detailed literature review about the stress state dependent
plasticity models, micromechanics models and the non associate flow rule.
Chapter 3 describes a series of new plasticity models, which is dependent on the
second and third invariants of the stress deviator as well as the hydrostatic stress.
Numerical implementations of the new plasticity models are also presented. As
applications, these plasticity models are calibrated and validated for three materials, a
5083 Aluminum, a Nitronic 40 and a Zircaloy 4.

Chapter 4 presents the extended Gurson-Tvergaard-Needleman porous plasticity


model to include the effects of the hydrostatic pressure and Lode parameter and gives a
few numerical examples.
Chapter 5 provides the research efforts on modeling ductile fracture of an
aluminum 5083 alloy by employing the GLD porous model.
Chapter 6 concludes the present thesis and gives the suggestions on the future
possible research work.

CHAPTER II
LITERATURE REVIEW

Great amount of theoretical work has been done on the plasticity and ductile
fracture areas. A literature review is provided in this chapter on development of the
plasticity theory with the hydrostatic pressure and Lode parameter effects. It is started by
a short introduction of some common used concepts and definitions in plasticity theories,
and proceeds with a detailed review about plasticity models with I1 and J3 effects,
micromechanics models and non-associated flow rule.

2.1 Basic Concepts and Definitions


Let ij be the Cauchy stress tensor with 1 , 2 and 3 being the maximum,
intermediate and minimum principal stresses respectively. The hydrostatic stress (or
mean stress) can be expressed as
1
3

1
3

1
3

h I 1 ii ( 1 2 3 )

(2. 1)

where I1 represents the first invariant of the stress tensor and the summation convention is
adopted for repeated indices. Let ij be the stress deviator with 1 , 2 and 3 being its
principal values, i.e.

ij ij h ij

(2. 2)

where ij represents the Kronecker delta. The first, second and third invariants of the
stress deviator are given as
1
( 1 2 3 ) 0
3
1
J 2 ij ji ( 1 2 2 3 3 1 )
2
1
( 1 2 ) 2 ( 2 3 ) 2 ( 3 1 ) 2
6
1
J 3 det( ij ) ij jk ki 1 2 3
3
J1

(2. 3)

The famous von Mises equivalent stress is expressed as

1
2

1 2 2 3 3 1
2

3J 2

(2. 4)

A particular stress state can be represented by a point in a Cartesian coordinate


system with axes 1 , 2 and 3, which is the so-called Haigh-Westergaard stress space as
shown in Figure 2.1(a). The diagonal of the stress space, denoted by a unit vector

n (1,1,1) / 3 , is called the hydrostatic axis. For any stress point P, a plane containing P

and perpendicular to the hydrostatic axis is called the deviatoric plane, which contains
line PN in Figure 2.1(a). The deviatoric plane passing through the origin is called the
plane. When it is viewed in the direction of hydrostatic axis, the projections of the 1 , 2
and 3 axes on the deviatoric plane are shown in Figure 2.1(b). It is often more
convenient to use the cylindrical Haigh-Westergaard coordinates given below to describe
the stress state
10

h 3

I1
3

ON

2 J 2 PN
cos3

(2. 5)

27 J 3
2 J 23 / 2

where is the Lode angle.

The Haigh-Westergaard coordinates, and therefore any

general stress state, are completely determined by the three stress invaria nts, I1 , J2 and J3 .
1

1
P(1,2,3)

ni

3
N

(a)

(b)

Figure 2. 1 (a) The Haigh-Westergaard stress space, and (b) the deviatoric plane
The stress triaxiality ratio is defined as the ratio of the hydrostatic stress (or mean
stress) over the effective stress,

T h / I1 /(3 ) 2 /(3 )

(2. 6)

where is the effective stress. It is very clear that, for a given stress triaxiality ratio T,
there exists an infinite number of stress states and each of them corresponds to a point on
the surface of a cone with ON as the axis. To distinguish various stress states having the
11

same triaxiality ratio, Lode angle is used as the second nondimensional parameter to
define a stress state.
There are many different definitions of the Lode angle (or Lode parameter) in the
literature. According to the work of Lode [64], the Lode angle , as the angle shown in
Figure 2.1(b), can be expressed as

tan

2 3 2 1
3 2 1

(2. 7)

Xue [65] defined the relative ratio of the principal deviatoric stresses as

2 3
1 3

(2. 8)

and an alternative Lode angle was expressed in the terms of ,

(2. 9)

Also this alternative Lode angle is related to the original Lode angle by

(2. 10)
Barsoum and Faleskog [53] defined another Lode parameter,, as

2 2 1 3
1 3

(2. 11)

This parameter can be easily connected to the original Lode angle by


12

tan

(2. 12)

Bai and Wierzbicki [55] introduced a normalized third deviatoric stress invariant,
B , which is related to Lode angle through

B cos 3

27 J 3
2 3

(2. 13)

All the parameters presented so far are based on the condition 1 2 3 . The Lode
parameter used in this thesis is defined as

cos 3

27 J 3

2 2 3

(2. 14)

2.2 Review of Plasticity Models


The most widely known theory of plasticity is the flow theory, which consists of a
yield criterion, a flow rule, a hardening law and the loading- unloading conditions. The
yield criterion determines the stress state when yielding occurs; the flow rule describes
the increment of plastic strain after yielding; the hardening law characterizes the
evolution of the flow stress with increased plastic deformation; and the loading- unloading
conditions determine if the stress path moves outward, inward or along the current yield
surface.
The most popular continuum plasticity model is the so-called J2 -flow theory. This
theory assumes hydrostatic stress as well as the third invariant of the stress deviator has
no effect on plastic yielding and the flow stress and it is widely employed to describe the
13

plastic response of metals. However, as shown in chapter I, numerous experimental


works have shown that this assumption is invalid for many types of materials. For this
reason, many criteria, such as Mohr-Coulomb [66, 67] and Drucker-Prager [10, 25]
among others, had been proposed to include the hydrostatic stress and third invariant of
the stress deviator effects.
The Mohr-Coulomb theory was named in honor of Charles-Augustin de Coulomb
[66] and Christian Otto Mohr [67]. The Mohr-Coulomb yield (failure) criterion is
expressed as

tan( ) c

(2. 15)

where is the shear strength, is the normal stress, c is the intercept of yield (failure)
envelope with the axis, and is the slope of the yield (failure) envelope with value
range from 0o to 90o . The Mohr-Colulomb criterion reduces to the Tresca criterion [1]
when =0o . On the other hand, if =90o , the MohrCoulomb model is equivalent to the
Rankine model [68]. This criterion is an extended form of the maximum shear stress
criterion, and it has made great success on modeling the failure in elastic range or small
strain plastic range. Besides, there were also several successful applications of this model
to predict ductile fracture. The distinctive characteristic of the Mohr-Coulomb criterion is
the explicit dependence on the Lode parameter, which is not considered in most ductile
fracture models. Bai and Wierzbicki [32] demonstrated the applicability of a extended
Mohr-Coulomb criterion to model ductile fracture of crack- free bodies.

14

The Drucker-Prager yield criterion [10] , which was introduced to model the
plasticity deformation of soils, is a pressure sensitive model to determine whether a
material has undergone plastic yielding. This yield criterion is expressed as
(2. 16)
where the I1 is the first invariant of the Cauchy stress tensor, J2 is the second invariant of
the deviatoric Cauchy stress tensor and A, B are the material constants determined by
experiments. If the yield stress, t , at uniaxial tension and the yield stress, c, at uniaxial
compression are known, the A and B can be expressed as

(2. 17)
With I1 term being considered, this yield criterion can describe the strength-differential
effect in tension and torsion. However, this form shows some insufficiency. For example,
if t approaches c, the value of B will be close to zero. At this situation, the hydrostatic
pressure effect cannot be described by this yield criterion anymore.
Different to the previous Drucker-Prager yield criterion [10], Drucker [25]
proposed a yield function that depends on the second and third invariants of the stress
deviator, by which the yield surface lies between the von Mises yield surface and the
Tresca yield surface. This Drucker criterion has the form
J 23 J 32 k 2

(2. 18)

where is a constant, which lies between -27/8 and 9/4 to make sure the yield surface
satisfy the convexity condition, and k is a material constant.
15

Inspired by the extensive experimental results reported by Spitzig et al. [69, 70]
on the behavior of high-strength metals undergoing uniaxial tension and compression,
Brunig [26] presented a formulation of a generalized I1 -J2 yield criterion to describe the
effect of the hydrostatic stress on the plastic flow properties. This I1 -J2 yield condition is
written in the form as

J 2 c(1 I1 ) 0

(2. 19)

where is a material constant, and the coefficients c are strain-dependent. It should be


noticed that this I1 -J2 yield criterion is an extended form of Drucker-Prager yield model.
But in Drucker-Prager yield model, c is a material constant and the work-hardening
effects are not considered. Later, this I1 -J2 plasticity model was applied by Brunig et al.
[56] to study the effect of stress triaxiality on the onset and evolution of damage in
ductile metals.
In Brunig et al. [27], the authors described an I1 -J2 -J3 flow theory in their
numerical simulation of the deformation and localization behavior of hydrostatic-stresssensitive metals. The new I1 -J2- J3 yield condition is shown as

J 2 c(1 I1 3 J 3 ) 0

(2. 20)

where and are material constants, and the coefficients c are strain-dependent.
Hu and Wang [29] proposed a stress-state dependent yield criterion for isotropic
ductile materials. This yield criterion can be indicated by the three stress invariants as

16

AI 1 J 2 B

J3
J2

(2. 21)

where A, B and C are three material constants, which can be determined through different
experiments.
In order to describe the constitutive response of a material that has pressure
sensitivity, Subramanya et al [71] employed an extended Drucker-Prager yield model to
study the roles of pressure sensitivity, plastic dilatancy and yield locus shape on the
interaction between the notch and a nearby void. The Extended Drucker-Prager model is
given as

q 1
1 r 3
tan
) c 0
1 (1 )( ) h tan (1
2 C
C q
3

(2. 22)

where h is the hydrostatic stress, q is the von Mises effective stress and r 3 27 J 3 / 2 .
Also, c is the true yield stress in a uniaxial compression test (with initial value of 0 )
and and C are materials parameters. This yield surface in principal stress space is a
conical surface with a vertex in hydrostatic tension axis. The minimum value of C is
0.778 to make sure the convexity condition of the yield surface. When C=1, this yield
criterion degenerated to original Drucker-Prager yield criterion, and when C=1 and =0,

is reduced to the famous von Mises yield criterion.


There are also other contributions with regard to the yielding criteria including the
I1 and J3 effects. Kuroda [28] presented a phenomenological plasticity model accounting
for hydrostatic stress-sensitivity and vertex-type effect. Cazacu and Barlat
17

[30] and

Soare et al. [31] extended Druckers J2 - J3 yield function [25] to include plastic
anisotropy and applied it to simulate sheet forming. A most recent study by Bai and
Wierzbicki [55] discussed a pressure and Lode dependent metal plasticity model and its
application in failure analysis. Yang et al. [72] conducted five types of tensile tests on a
2A12-T4 aluminum alloy and modified the von Mises yield criterion to include the Lode
dependence.

2.3 Micro Mechanics Models


The microstructure of materials is very complex and the fracture of ductile
materials usually follows a multistep failure process: void nucleation, growth and
coalescence. Void nucleation usually happens at a quite low stress level, so that material
can be assumed to have voids at the start point of loading. Two common methods are
used to describe this fracture mechanism. One is the straight- forward approach by
modeling individual voids explicitly using finite element mesh. Void growth performance
can be accurately simulated by this method. However, it would require a large number of
elements to model the voids in a structural component. A huge number of finite element
elements are too much for the computational capability at present. So another approach
called porous continuum method was developed to study the ductile fracture for a
practical alternative. Recently, Benzerga and Leblond [73] gave a very detailed literature
review about the ductile fracture by void growth to coalescence.
The innovative works of studying micromechanics were due to McClintock [74]
and Rice and Tracy [75] by considering the growth of isolated voids. McClintock
investigated an isolated cylindrical-shaped void that was subjected to uniform remote
18

stressing. Rice and Tracy considered the growth of a spherical void in non-hardening
material subjected to remote uniaxial tension strain rate field. Not only has the void
grown in the radial direction, but also the shape of the void has changed. Later on, based
on homogenization assumption, constitutive equations for porous materials were
developed. The most well-know porous constitutive model is proposed by Gurson [76], in
which the yield criteria were approximated through an upper bound theorem of plasticity.
The matrix materials were considered as rigid-perfectly plastic and the plastic behavior in
the matrix obeyed the Von Mises yield criterion. Associated flow rule is employed with
yield function served as flow potential. Equations of the Gurson model are formulated in
terms of the average macroscopic Cauchy stress ij , with corresponding stress deviator
Sij ij ij kk / 3 . The void shape was considered as a sphere. The yield criterion of

Gurson model was expressed as

2e

3
2 f cosh h (1 f 2 ) 0

2
2

(2. 23)

where e is the von Mises effective stress ( e 3Sij Sij / 2 ), is the yield stress of the
matrix material, f is the void volume fraction and h is the hydrostatic stress. The yield
surface given by equation (2.23) turns into von Mises yield criterion at f=0. Associative
flow rule is employed and the macroscopic plastic strain is

p
ij

ij

(2. 24)

19

where ijp are the rates of the plastic strain components, is a positive scalar called the
plastic multiplier and ij are the macroscopic stress components. The matrix plastic strain
is related to the macroscopic strain by enforcing the equivalence of plastic work, ie.

ijp : ij (1 f ) p

(2. 25)

In this model, the void volume fraction, f, is considered as an extra internal variable to
capture the growth of cavities.
For Gursons yield function, the complete loss of load carrying capacity occurs
when void volume fraction ratio equals one, which is greater than the experimental
observation. Tvergaard [77] introduced two more parameters, q1 and q2 , to improve this
situation based on their bifurcation study results. Moreover, Tvergaard and Needleman
[78] modified the Gurson model by introducing a void volume fraction function, f*, to
model the rapid loss of load carrying capacity after the coalescence occurs. The GursonTvergaard-Needleman (GTN) model was expressed as

e2

3q
2q1 f * cosh 2 h 1 ( q1 f * ) 2 0

2
2

(2. 26)

where e is the von Mises effective stress, is the yield stress of the matrix material, f *
is the void volume fraction function, h is the hydrostatic stress and q1 and q2 are model
parameters which have an effect on the shape of the yield surface. When q1 = q2 =1, this
model becomes the original Gurson model.
To account for the final material failure, the function f * (f) was specified by
20

f,

1/ q1 f c
f
f c f f ( f f c ),
f
c

f fc
fc f f f

(2. 27)

where f c and f f are the void volume fractions at coalescence and failure respectively. It is
noted that the macroscopic stress carrying capacity vanishes when the void volume
fraction function f * achieves the ultimate value f * 1/ q1 . According to the experiment
results from Brown and Embury [79] and the numerical model analysis from Andersson
[80], f c =0.15 and f f =0.25 are appropriate values.
The development of void volume fraction is partially caused by growth of existing
voids and partially caused by nucleation of new voids. So the evolution law for void
volume fraction is determined by:

f ( f ) growth ( f ) nucleation

(2. 28)

where ( f ) growth and ( f ) nucleation represent the increasing rate of void volume fraction due
to growth and nucleation respectively.
The matrix is assumed to be incompressible, but the macroscopic response of the
material is compressible because of the voids. So ( f ) growth is caused by the total volume
change

( f ) growth (1 f ) kkp

(2. 29)

For ductile material, void nucleation is caused by the cracking and decohesion of
inclusions or second-phase particles. Needleman and Rice [81] suggested that the
21

localization occurs at early stage of deformation. Due to complicated physical structure


of the inclusion and the microstructure, the void nucleation particles are not homogenous
in the solids. Various nucleation criteria [81-85] had been proposed. Chu and Needleman
[86] suggested a plastic strain controlled nucleation in which

fnucleation A ,

(2. 30)

where is the matrix plastic strain rate, and A as

1
fN
N
A
exp
2
s
s N 2

(2. 31)

where sN and N are the standard deviation and the mean value of the distribution of the
plastic strain, f N is the total void volume fraction of void nucleating particles.
Despite of the great success of GTN model, a distinct limitation of the GTN
model is the assumption that voids are spherical in materials and remain spherical in the
growth process. Actually many materials, such as rolled plates, have non-spherical void
shape. Even for materials having initially spherical voids, the void shape may change to
probate or oblate shape after deformation depending on the applied stress state. In order
to overcome this disadvantage, Goludanu, Leblond and Devaux [57-59] proposed the socalled GLD model, in which both void volume fraction and void shape evolve with
deformation.

Since non-spherical voids were considered in this model, a preferred

material orientation existed and the plastic behavior became anisotropic. The GologanuLeblong-Devaux (GLD) models yield function is expressed as

22

2

2
' h X 2q g 1 g f cosh h g 1 q 2 g f

(2. 32)

0
where ij are the macroscopic stress components, f is the void volume fraction, is the
yield stress of the matrix material,

denotes the von Mises norm, represents the

~
deviatoric stress tensor, h is the generalized hydrostatic stress defined by
~
h 2 xx zz 1 2 yy

(2. 33)

and X is defined as
X 2 / 3e y e y 1 / 3e x e x 1 / 3e z e z

(2. 34)

where (ex, ey, ez) is an orthogonal basis with ey parallel to the axisymmetric axis of the
void and denotes tensor product. S is the void shape parameter. The parameters C, , g,
and 2 are functions of f and S, and the heuristic parameter q depends on initial void
volume fraction, strain hardening exponent of the matrix material, void shape parameter S
and the macroscopic stress triaxiality factor T. Detailed description of the GLD model
can be found in chapter V. The GLD model can turn into Gurson model by setting the
void shape as sphere. It can be also reduced to the von Mises yield criterion when f=0 at
prolate void condition. For oblate voids, f=0 corresponds to a material with a distribution
of penny-shaped cracks.

Kim and Gao [87] developed a generalized approach to formulate the consistent
tangent stiffness for complicated plasticity models. Using the approach developed by
23

Kim and Gao, the GLD model was implemented into ABAQUS via a user subroutine to
study stress state effects on ductile fracture of aluminum 5083 in chapter v.

2.4 Non-associated Flow Rule


A common approach in the metal plasticity theory is the adoption of the
associated flow rule or normality condition, which can be interpreted as that the normal
to the yield surface is the same as the direction of the plastic strain increment vector.
Normality condition requires the yield function and the flow potential to be identical.
The normality condition has been confirmed by a lot of experiments, based on the
observations that some metals do not have pressure sensitivity in yielding and volume
change after large deformation is negligible. However it is found to be seriously in
inaccurate for geological materials, where the associated flow rule overestimates the
plastic volume changes. For these materials, the non-associated flow rule must be
employed [88-90].
In metal plasticity, non-associated flow rule were developed after the
experimental observations by Spitzig and Richmond [19] for iron-based materials and
aluminum, in which they found that while yielding is pressure dependent, plastic
deformation is essentially incompressible. Therefore, their results suggests that normality
concept, which overestimated the plastic volume increase for their materials, may not
always be appropriate in metal plasticity. Even for the materials without the pressuresensitive effects, Lademo et al. [91] provided an evaluation of three proposed material
models, found that an associated flow rule was inadequate to describe the anisotropic
24

flow properties observed in uniaxial tension tests and suggested the non-associated flow
rule.
Many studies, such as those by Mroz [92], Nemat-Nasser and Shokooh [93],
Doris and Nemat-Nasser [94], Nemat-Nasser [95, 96], Runesson [97] , Brunig et al. [27] ,
Stoughton [98] and Stoughton and Yoon [99, 100] , also indicated that appropriate
constitutive description of many materials can be achieved by using the less restrictive
non-associated flow rule. Brunig [26] and Brunig et al. [27] demonstrated that pressure
sensitive yielding and non-associated flow rule remarkably influence the onset of
localization and the subsequent localization behavior. Stoughton [98] proposed a material
model based on non-associated flow rule for sheet metals to describe the directional
dependence of uniaxial tension data. In order to account for the strength differential
effect, Stoughton and Yoon [99] proposed a non-associated flow rule based on a pressure
sensitive yield criterion. To model the ductile fracture process in solids, Ma and
Kishimoto [101] proposed a non-associated flow rule to characterize the yield and plastic
deformation of void-containing materials, where the yield function took the form of the
Gurson-Tvergaard-Needleman porous plasticity model [76-78, 102] while the flow
potential took a slightly different form. More recently, Cvitanic et al. [103] presented a
plasticity model based on non-associated flow rule and detailed finite element
formulations for sheet metal forming. Taherizadeh el al. [104] presented an anisotropic
material model based on non-associated flow rule and mixed isotropic-kinematic
hardening for simulation of sheet metal forming.

25

CHAPTER III
MODELING OF PLASTICITY RESPONSES: INVOLING THREE STRESS
INVARIANTS

3.1 Introduction
Plasticity describes the deformation of a material undergoing non-reversible
changes of shape in response to applied forces. Study of materials plasticity is an
indispensable part of the structural deformation research. In recent days, structural
analysis and integrity/risk assessments of high performance engineering components
often demand constitutive models that can accurately describe a material's plasticity
behavior. A complete plasticity model consists of a yield criterion, a flow rule and a
hardening law. The most popular continuum plasticity model is the so-called J2 -flow
theory.

In this theory, the second invariant of the deviatoric stress tensor controls

yielding and plastic flow. The hydrostatic stress (first invariant of the stress tensor) as
well as the third invariant of the stress deviator (Lode parameter) is assumed to have no
effect. The J2 flow theory has been widely employed to describe the plastic response of
various materials. For a material that obeys J2 -flow plasticity, its plasticity behavior is
characterized by the (von Mises) equivalent stress-strain curve which can be obtained by
conducting a uniaxial tension test, a compression test, or a torsion test. The stress-strain
curve obtained from either one of these tests is then used to predict the materials p lastic
response under various states of stress.
26

However, increasing experimental evidence shows that a fundamental assumption


made in von Mises theory, that plasticity is only characterized by J2 , contains varying
levels of inaccuracies over a wide range of materials. Hydrostatic pressure and Lode
parameter also play very important roles in the plasticity behavior for some certain
materials. As we discussed in chapter I and chapter II, there have been a lot of
contributions made to include the hydrostatic pressure and Lode parameter effects into
the plasticity models.
In this chapter, a series of new plasticity models is introduced to include the
hydrostatic pressure and Lode parameter (related to the third invariant of stress deviator)
effects as well as the second invariant of stress deviator. Based on the procedure
proposed by Avaras [105] and Kim [87], the finite element implementation including
integration of the constitutive equations using the backward Euler method and the
formulation of the consistent tangent modules is presented. These plasticity models are
implemented in to the finite element software ABAQUS via UMAT.
These new plasticity models are calibrated and verified for an Aluminum 5083
alloy, a Zircaloy-4 and a Nitronic 40 to simulate the plasticity behaviors under different
stress-state tests. In order to provide large range of triaxiality and Lode parameter values,
different types of testing specimens were employed. A torsion-tension specimen is
designed to provide stress states with low triaxiality values. Test matrix includes round
bar tensile specimens, notched round bar tensile specimens, compression specimens,
grooved plains train specimens, modified Lindholm torsion specimens, modified

27

Lindholm torsion-tension and torsion-compression specimens. The numerical plasticity


predictions using this new proposed plasticity models agree with experiments very well.

3.2 Plasticity Modeling


In this section, detail formulation of the new I1 -J2-J3 plasticity model is presented.
The influence of I1 and J3 are also shown, and the appropriate yield function and flow
potential forms are listed for the three materials studied in this thesis. The yield function
and flow potential forms for these three materials are determined by comparing numerical
simulation results by using different yield function and flow potential forms with the
experimental data (three sets of experimental data: round bar tensile test, compression
tests and torsion test) to see which set of the yield function and flow potential forms leads
to best match.

3.2.1 Influence of I1 and J3


Let ij be the Cauchy stress tensor and 1 , 2 and 3 be the three principal stress
values. The hydrostatic stress (or mean stress) can be expressed as
1
3

1
3

1
3

h I 1 ii ( 1 2 3 )

(3. 1)

where I1 represents the first invariant of the stress tensor and the summation convention is
adopted for repeated indices. Let ij be the stress deviator tensor and 1 , 2 and 3 be
its principal values, i.e.

28

ij ij h ij

(3. 2)

where ij represents the Kronecker delta. It is obvious that the first invariant of the stress
deviator tensor is zero. The second and third invariants of the stress deviator tensor are
defined in Eq. (3.3) as
J1 0
1
J 2 ij ji ( 1 2 2 3 3 1)
2
1
( 1 2 ) 2 ( 2 3 ) 2 ( 3 1 ) 2
6
1
J 3 det( ij ) ij jk ki 1 2 3
3

(3. 3)

For isotropic materials, the plastic behavior is often described by the stress
invariants I1 , J2 and J3 and consequently, the general forms of the yield function (F) and
the flow potential (G) are expressed as functions of I1 , J2 and J3 . Eq. (3.4) describes the
yield condition

F ( I1 , J 2 , J 3 ) 0

(3. 4)

where is the hardening parameter.


When material deforms plastically, the flow rule is used to define the inelastic
part of the deformation [106]

ijp

G ( I 1 , J 2 , J 3 )
ij

(3. 5)

29

where ijp are the rates of the plastic strain components, is the plastic multiplier which
is a non- negative scalar, and G is a function called the flow potential [107]. Eq. (3.5)
suggests that the directions and magnitudes of the rates o f the plastic strain components
are determined by the plastic flow potential and the plastic multiplier respectively. In
most cases of metal plasticity, the flow potential and the yield function are assumed to be
identical, i.e., F = G, which is known as the associated flow rule or normality rule. When
F and G are not identical for a material, this material is said to follow a non-associated
flow rule.
Various forms of F and G functions can result in different plasticity models.
While there is not a single functional form F and G must satisfy, here they are taken as
first order homogeneous functions of stress. The choice of F as a first order homogeneous
function of stress leads to a straightforward definition of the equivalent stress, e F .
Appendix A gives six examples of first order homogeneous functions of stress
that are considered in this work to represent the yield function, F, and the flow potential,
G.

These functions are generalized from existing plasticity models, e.g. t he Drucker

model [25] and the Drucker-Prager model [10]. These functions depend on material
constants a1 , b1 , a2 , and b2 . The a1 and b1 values control the hydrostatic pressure and
Lode parameter effects on yielding respectively. The ranges of the parameters involved in
these functions are also given in Appendix A to ensure the convexity requirement is
satisfied.

30

For all six functional forms listed in Appendix A, if a1 = a2 and b1 = b2 the


material follows the associated flow rule. If a1 = b1 = a2 = b2 = 0, the plasticity model
reduces to the formulation of the classical J2 -flow theory and e defined by e F
becomes the von Mises equivalent stress.
The hardening parameter depends on the strain history. By enforcing the
equivalence of plastic work, i.e.,

p ij ijp

(3. 6)

the equivalent plastic strain increment can be defined as

p ij ijp /

(3. 7)

For strain hardening materials, the hardening behavior can be described by a


stress vs. plastic strain relation ( p ) , where

p dt . If a material follows the J2

plasticity theory, the stress strain curve obtained from a uniaxial tensile test defines

( p ) . For materials with I1-dependent plasticity, the input curve must be modified as
discussed later in this chapter. Since the flow potential is taken to be a first order
homogeneous function of stress, Eulers homogeneous function theorem results in

ij ijp ij

G
G
ij

(3. 8)

From (3. 6) and (3. 8), the plastic multiplier and the equivalent plastic strain rate
can be related through

31

F
G

(3. 9)

If the material follows the associate flow rule (F = G), it is obvious that and p
are equal. For materials in which the non-associated flow rule applies, the equivalent
plastic strain rate defined by Eq. (3. 9) differs from by a factor F / G .
Because of the I1 term in the flow potential G, the plastic response becomes
dilatant, with the rate of volume change given by

kkp

G
F
6
6
(3c2 a2 I15 / G 5 ) 3c2 a2 I15 p 6
kk
G

(3. 10)

For materials considered in this study, the 5083 aluminum alloy, the Nitronic 40
stainless steel and the Zircaloy 4, the following forms of F and G functions are found
appropriate by comparing model output with test data and manually manipulating the a1 ,
a2 , b1 , and b2 parameters to produce what is believed to be the best match based on visual
inspection of the experimental data.

3.2.2 Yield Function and Flow Potential for 5083 Aluminum Alloy and Zircaloy-4
For aluminum 5083 and Zircaloy-4, the following first order homogeneous
function of stress is found proper for defining the yield function
F c1[a1 I1 (27 J 23 b1 J 32 )1/ 6 ]

(3. 11)

32

where a1 , b1 are material constants and c1 is determined by substituting the uniaxial


condition into Eq. (3.11) so that the equivalent stress defined by e F equals to the
applied stress. This results in

c1 1 / a1 (4b1 / 729 1)1 / 6

(3. 12)

The flow potential takes a similar form, i.e.,


G c2 [a2 I1 (27 J 23 b2 J 32 )1 / 6 ]

(3. 13)

where c2 1 /( a2 (4b2 / 729 1)1 / 6 ) .


If a1 = a2 =0, the plasticity model reduces to the form of the Drucker model [25]
and if b1 = b2 =0, it degenerates to the form of the Drucker-Prager model [10].

3.2.3 Yield Function and Flow Potential for Nitronic 40 Stainless Steel
For Nitronic 40, the following first order homogeneous function of stress is
proposed for the yield function
F c1 [a1 I1 (3 3J 23 / 2 b1 J 3 )1/ 3 ]

(3. 14)

where a1 , b1 and c1 are material constants. Again, the constant c1 can be found by
substituting the uniaxial condition into Eq. (3.14), which leads to

c1 1 / a1 (2b1 / 27 1)1/ 3

(3. 15)

When the material is subjected to a uniaxial stress , the value of c1 given by Eq. (3.15)
ensures F .
33

The flow potential takes a similar form, i.e.,


G c2 [a2 I1 (3 3J 23 / 2 b2 J 3 )1/ 3 ]

(3. 16)

where c2 1 / a2 (2b2 / 27 1)1/ 3 .


Again, If a1 = a2 =0, the plasticity model reduces to the form of the Drucker
model [25] and if b1 = b2 =0, it degenerates to the form of the Drucker-Prager model [10].

3.3 Numerical Implementation of the I1 -J2 -J3 plasticity model


Following the procedures of Aravas [105] and Kim and Gao [87] , we implement
the I1 -J2-J3 plasticity model outlined in the previous section into ABAQUS [108] via a
user defined subroutine. Since ABAQUS/Standard is employed to carry all the numerical
analysis, stress update and consistent tangent moduli are needed to be provided in this
user subroutine (UMAT).

3.3.1 Stress Update


We adopt the backward Euler method to integrate the rate constitutive equations.
For the strain driven integration algorithm, the stresses and state variables are known at
the start of each increment and their values need to be updated at the end of the increment
corresponding to the total strain increment. The elasticity equations give


ij

t t

e
Cijkl
kle

t t

e
Cijkl
kle

where

34

kl

e
klp ijT Cijkl
klp

(3. 17)

e
kle kl
ijT Cijkl
t

(3. 18)

is the elastic predictor, t represents the time at the start of the increment, t+t represents
the time at the end of the increment, and the superscripts e and p denote elastic and
plastic components respectively. The total strain increment kl is known, and if the
e
linear elastic behavior is isotropic, the elastic moduli Cijkl
can be expressed as

e
Cijkl
G ik jl il jk K G ij kl
3

(3. 19)

where G and K are the elastic shear and bulk moduli respectively. Therefore, to update
stresses, the plastic strain increments need to be determined. The following outlines the
procedure for computing ijp .
The yield condition and the flow rule are written as
F ( I1t t , J 2t t , J 3t t ) ( t p t ) 0

(3. 20)

and

p
ij

G ( I 1 , J 2 , J 3 )

ij

t t

(3. 21)

Equations (3.20) and (3.21), when considered together with Eq. (3.17), result in
p
10 equations for and nine components of ij , among which 7 equations are
p
independent due to the symmetry of ij . If the state variable p is updated and thus

35

( t p t ) is known, these equations can be solved iteratively for ijp and using the
Newton-Raphson method. The iterative process follows these steps: 1) assume initial

ijp 0 and use Eq. (3.17) to estimate ijt t ; 2) use Eqs. (3.20) and (3.21) to solve for
ijp ; 3) update stresses using Eq. (3.17); 4) repeat steps 2) 3) until convergence
conditions are satisfied.
To update p , consider the hardening law and the evolution equation for

( t p t ) ( t p p )

(3. 22)

p ijt t ijp

(3. 23)

and

At each iteration of ijp , p and ( t p t ) can be obtained by solving equations (3.22)


and (3.23) iteratively using the Newton-Raphson method.

3.3.2 The Consistent Tangent Moduli


Simo and Taylor [109] showed that use of the consistent tangent moduli
significantly improves the convergence characteristics of the overall equilibrium
iterations. The consistent tangent stiffness corresponding to the backward Euler
integration can be obtained by linearization of equation (3.17).

J ijkl

ij

kl

t t

(3. 24)

36

Since all quantities in calculating Jijkl are referred to time t+t, the superscript t+t will
be dropped from hereafter.
Equation (3.17) can be rewritten as

e
e
e
ij Cijkl
kle Cijkl
kl klp Cijkl
kl klp klp

(3. 25)

Differentiating (3.25) results in


e
e
ij Cijkl
kl Cijkl
( klp )

(3. 26)

The relationship between ij and ( klp ) can be found as follows.


Let h be the hardening modulus, i.e., h / p , then Eq. (3.23) can be
rewritten as

h ij ijp

(3. 27)

By differentiating (3.27) and simplifying the resulted relatio n, the following


equation can be obtained

h ijp

h mn
2

p
mn

ij

h ij
p
h mn mn
2

ijp

(3. 28)

Differentiating Eq. (3.20) and combining the result with Eq. (3.28) give

h ij
p
2 h mn mn

h ijp
F ( I 1 , J 2 , J 3 )
( ) (
2
) ij
p
ij
h mn mn
p
ij

37

(3. 29)

Eliminating from the nine equations given by (3.21) results in eight equations,
such as
G ( I 1 , J 2 , J 3 ) p G ( I 1 , J 2 , J 3 ) p

11
22 0

22
11

G ( I 1 , J 2 , J 3 ) p G ( I 1 , J 2 , J 3 ) p
22 0

21
22
21

(3. 30)

Differentiating (3.30) leads to

G
G
2G
2G

11p
22p 22p
11p

11 ij
22 ij
22
11

G
G

21p
22p 22p

22
21

ij

2G
2 G
21p
ij
21 ij
22 ij

(3. 31)

Eqs. (3.29) and (3.31) provide nine equations between ( ijp ) and ij , which
can be summarized as

Kp D

(3. 32)

p
P
, 11, 21, 31,, 33 , K is the coefficient
, 31p ,, 33
where p 11p , 21
T

matrix of p and D is the coefficient matrix of .


From Eqs. (3.24) and (3.30), we can obtain

p K DCe

DCe

(3. 33)

38

e
where Ce is a 99 matrix representing the elasticity tensor Cijkl
. Finally, the consistent

tangent matrix can be obtained by Substituting (3.33) into (3.26)

J / Ce Ce K DCe

DCe

(3. 34)

3.4 Applications of the stress-state dependent plasticity models


In this section, the proposed I1 -J2-J3 plasticity models are employed to study the
plastic response of Aluminum 5083, Nitronic 40 and Zircaloy-4. Designing of the
torsion-tension specimen is also shown here. These proposed I1 -J2 -J3 plasticity models
are calibrated and validated for these three materials and good comparison between the
numerical predictions and experimental data have been achieved.

3.4.1 Designing of torsion-tension specimen


In order to generate a wide range of stress states, special designed test specimens
are needed to be considered in our experimental and numerical analysis. In the study of
Gao et. al [110], round bar tensile specimen, notched round bar specimen, torsion
specimen and grooved plane strain specimens are employed to investigate the influence
of the hydrostatic stress and Lode angle on ductile failure of DH36 steel. With definitions
of triaxiality as T I 1 /(3 3J 21 / 2 ) and Lode parameter as 3 3J 3 /( 2 J 23 / 2 ) , round bar
tensile specimen has T value initially as 1/3 and increase subsequently after the onse t of
necking. Notched round bar specimens provide different triaxiality values according to
the notch radius. For both round bar tensile specimen and notched round bar tensile
specimens, the value is 1. Triaxiality T of torsion specimen is 0, and grooved plane
39

strain specimens have different triaxiality values according to different grooved radius.
For both torsion specimen and grooved plane strain specimens, Lode parameter is 0.
Aluminum 5083 is one of the materials studied in this thesis. In order to give a
clear view about how the stress states change with the loading history for different type
of specimens, detailed plots of triaxiality and Lode parameter for each type of specimens
are shown and analyzed here by taking aluminum 5083 as example. The geometries of
each specimen can be found in the later section of this chapter.
The stress state of a material point in the test specimens evolves as plastic
deformation increases. For aluminum 5083, Figures 3.1 3.3 show the variation of
triaxiality T and Lode parameter with loading history (measured by the equivalent
plastic strain, p ) in the element at the specimen center for the smooth round bar, the Enotch round bar (notch radius 6.35 mm ) and the G-groove plane strain specimen (groove
radius 5.08 mm). For the smooth round bar, remains at 1 during the entire loading
history while T increases from 1/3 to about 0.45 before specimen fractures. For the Enotch specimen, remains at about 1 and T increases from 0.71 to 0.8. For the G- groove
specimen, quickly decreases to 0 and remains at this level as plastic deformation
increases while T increases from 0.51 to 0.8 before failure occurs. For the notched and
grooved specimens, changing notch (groove) radius changes the level of T in the
specimen. Three different notched round bar specimens and three grooved plane strain
specimens are considered. The notch radii/groove radii for these specimens are shown
later in this chapter (Table 3-3 and table 3-4). Considering the entire loading history of
each specimen and the three different notch (groove) radii, the range of T experienced by
40

the center element is 0.71 T 1.6 for the notched round bar tests and 0.46 T 0.97
for the plane strain tests.
1.5

0.6

0.4

0.2

0.5

0
0

0.1

0.2

0.3

0.1

0.2

0.3

(b)

(a)

Figure 3. 1 Variation of T I 1 /(3 3J 21 / 2 ) and 3 3J 3 /( 2 J 23 / 2 ) with plastic


deformation in the center element of the smooth round bar.
1.5

1
0.8

0.6

0.4

0.5

0.2

0
0

0.05

0.1

0.15

0.2

0.05

0.1

0.15

0.2

p
(a)
(b)
1/ 2
Figure 3. 2 Variation of T I 1 /(3 3J 2 ) and 3 3J 3 /( 2 J 23 / 2 ) with plastic
deformation in the center element of the E-notch specimen.
p

41

0.6

0.8

0.3

0.6
0.4

0.2

-0.3

-0.6
0

0.05

0.1

0.15

0.2

0.05

0.1

0.15

0.2

(b)

(a)

Figure 3. 3 Variation of T I 1 /(3 3J 21 / 2 ) and 3 3J 3 /( 2 J 23 / 2 ) with plastic


deformation in the center element of the G- groove specimen.
As we can see from the all the figures above, the round tensile bars (smooth and
notched) and the grooved plane strain specimens cannot provide data points in the low
stress triaxiality range (T < 0.4). So a specimen is required to provide a low triaxiality
value and also with Lode parameter between 0 and 1. The low range of triaxiality can be
achieved using a simple notched tubular specimen tested in combined tension and torsion
by Faleskog et.al. [53, 54]. Here an alternative specimen design is developed, which can
be used for conducting combined tension-torsion tests to provide useful data points at this
range. This specimen is modified from the specimen that Lindholm [63] used for highrate torsion tests. The difference between this modified Lindholm specimen and the
original Lindholm specimen is that this modified Lindholm specimen is longer and the
ends are cylindrical rather than hexagonal.
Next, the investigation of the distributions of stress, triaxiality and Lode
parameter in the gage section of the modified Lindholm specimen is conducted. In order
to provide an effective comparison with established approaches, numerical analysis of the
42

modified Lindholm specimen is compared with the notched tube specimen used by
Barsoum and Faleskog [53, 54]. The dimensions of Faleskogs specimen (referred as NT
specimen) are shown in the figure 3.4. The dimensions of the modified Lindholm
specimen are shown in the figure 3.5. Even though the overall sizes of these two
specimens are close to the same, the gage sections are significantly different. The NT
specimen has both internal and external circular notches, while the modified Lindholm
specimen is notched only on the outside with a trapezoidal notch profile. The gage
section in the modified Lindholm is uniform over 2.54 mm (0.1 in), while the gage
section in the NT specimen is curved throughout.

The gage section width (wall

thickness) for the NT specimen is 1.2 mm at the narrowest point, which is about twice the
thickness of the wall throughout the gage section of the Lindholm specimen (0.74 mm).
The thin wall in the gage section acts to minimize the variation in stress through the wall
thickness. The very tight tolerance on concentricity shown in figure 3.5 is intended to
ensure axisymmetric deformation in the gage section.

Figure 3. 4 Dimensions of NT specimen. All dimensions in mm [53,54].


43

Figure 3. 5 Dimensions of Lindholm specimen. All dimensions in mm [53,54].

LINDHOLM

NT

Figure 3. 6 Comparison of gage sections in NT and Lindholm specimens (to scale).


A series of finite element stress analyses were conducted to compare the
distributions of stress and effective plastic strain in the gage section of the NT specimen
and the modified Lindholm specimen under different combinations of tension and
44

torsion. The stress state for both specimens will be characterized using the triaxiality
parameter, T, and the Lode parameter, [53, 54]. The Lode parameter can be gained by
re-arranged following the equation:

1 3

2
1 3

(3. 35)

For the comparison, J2 flow plasticity theory is employed to describe the material
plasticity response. The material we used here is Weldox 960, which is the same material
used by Barsoum and Faleskog [53, 54].

( )N
0
0

if
if

0
(3. 36)

where E = 208 GPa, = 0.3, 0 = 956 MPa, 0 = 0.0046 and N = 0.059. E, , 0 , 0 and N
represent the Youngs modulus, the Poissons ratio, the yield stress, the total strain at
yield and the strain hardening exponent respectively.
The finite element software ABAQUS is employed to run the numerical analysis.
Both specimens are modeled using four nodes reduce integration axisymmetric elements
with twist (CGAX4R). The element size in the gage section of both specimens is
approximately 0.05 x 0.05 mm. The finite element meshes for the two specimens are
shown in the following figure.

45

(b) LINDHOLM

(a) NT

Figure 3. 7 Finite element meshes for the, (a) NT specimen, and the, (b) modified
Lindholm specimen.
An axial displacement, u, and a twist angle, , are imposed on the top end of both
specimen while the bottom end is constrained against the axial or rotational motion. The
tension-torsion ratio based on displacement is defined as:

u
R

(3. 37)

In this equation R is the mean radius in the center of the gage section, which is 12 mm for
the NT specimen and 6.91 mm for the Lindholm specimen. During the loading process
the tension-torsion ratio, , is kept constant. The stress distributions presented in the
following are all along a radial line in the center of the gage section, which is on the
symmetry plane of the specimen.

46

Because the geometries of Lindholm specimen and NT specimen are different,


they can not achieve the same values of both T and at the same time. When the T value
is comparable for both specimens, values are different, while when the values are
similar, T values are different. For the Lindholm specimen, the two ends (outside the
gage section) can be consider as rigid bodies, while for the NT specimen, the two ends
have significant deformations relative to the gage section, and therefore can not be
consider as rigid bodies. Two different loading conditions are compared for these two
specimens: one is pure torsion and another is torsion-tension.

3.4.1.1 Pure torsion


Figure 3. 8 compares the through-thickness distribution of T, and p for pure
torsion loading ( = 0) at the time when the equivalent plastic strain, p , reaches a level
of about 0.37. For pure torsion, T and should both be zero. For the NT specimen, rm
represents the mid-radius in the gage section (12 mm) and t n is the half net thickness in
the narrowest part of the notch (0.6 mm). For the Lindholm specimen, rinner represents
the inner radius (6.54 mm) and t n is the wall thickness in the gage section (0.735 mm).
It is apparent that T and are very close to zero through the whole thickness for
both specimens. For the Lindholm specimen, the effective plastic strain remains constant
through the thickness while it shows a small variation in the NT specimen; it is higher at
the inner and outer surfaces and lower at the mid-radius.

47

0.5

-0.5
-1

0.5

-0.5

-1

-0.5

(r-rm)/tn

0.5

-1

0.5

(r-rinner)/tn

0.75

(b)

(a)

Figure 3. 8

0.25

Through-thickness distribution of T, and p at the mid-section of the

specimen for p 0.37 and pure torsion loading, (a) NT specimen, (b) Lindholm
specimen.

3.4.1.2 Torsion-tension
In an effort to provide a comparison under similar conditions, the tension-torsion
ratios applied to the two different specimens were chosen to produce similar values.
For the NT specimen, = 0.194, and for the Lindholm specimen, = 0.068.

These

values are chosen such that the two specimens produce similar values, which are
around -0.2.
Figure 3.9 compares the through-thickness distribution of T, and p at the time
when p in the element at the inner surface of the Lindholm specimen, and the center if
the NT specimen, reaches 0.02. Figure 3.10 compares the through-thickness distribution
of T, and p at a higher strain level of 0.20. The plastic strain at yield is 0.002, so these
strain levels are well beyond yield. Figure 3.11 compares the through-thickness
distribution of T, and p at the time when p reaches 0.44.
48

0.5

0.5

-0.5
-1

-0.5

-1

-0.5

0.5

(r-rm)/tn

-1

0.25

0.5

(r-rinner)/tn

0.75

(b)

(a)

Figure 3. 9 Through-thickness distribution of T, and p at the mid-section of the


specimen for effective plastic strain of 0.02, (a) NT specimen, (b) Lindholm specimen.

0.5

0.5

-0.5
-1

-0.5

-1

-0.5

(r-rm)/tn

0.5

-1

0.25

0.5

(r-rinner)/tn

0.75

(b)

(a)

Figure 3. 10. Through-thickness distribution of T, and p at the mid-section of the


specimen for effective plastic strain of 0.20, (a) NT specimen, (b) Lindholm specimen.

49

0.5

0.5

T
0

-0.5
-1

-0.5

-1

-0.5

(r-rm)/tn

0.5

-1

0.25

0.5

(r-rinner)/tn

0.75

(b)

(a)

Figure 3. 11 Through-thickness distribution of T, and p at the mid-section of the


specimen for effective plastic strain of 0.44, (a) NT specimen, (b) Lindholm specimen.
From the above figures, we can see that p remains constant through thickness
for the Lindholm specimen but shows thickness-variation in the NT specimen, with the
lowest point at about the mid-radius. Triaxiality in the NT specimen is lowest at the
inner/outer surfaces and higher at the center, while T in the Lindholm specimen is highest
at the inner surface and exhibits a small gradual decrease through the thickness. Lode
parameter shows a gradual increase from inner to outer surfaces in both specimens. In
general, the stress and strain distributions in the modified Lindholm specimen exhibit
good uniformity, thereby obviating the need to determine the exact failure location.
It is interesting to compare the two specimens at similar levels of triaxiality. In
this case, the Lode parameter will be different for these two specimens. In an effort to
achieve the same value for T, an analysis was run where 0.194 for the NT specimen
and 0.368 for the modified Lindholm specimen. These values were chosen such that the
two specimens produced similar T values, which were around 0.36.
50

Figure 3.12 compares the through-thickness distribution of T, and p at the time


when p in the element at the inner surface for Lindholm specimen, and center of the NT
specimen, reaches 0.14.
1

0.5

0.5

-0.5

-0.5

-1

-1

-1

-0.5

(r-rm)/tn

0.5

0.25

0.5

(r-rinner)/tn

0.75

(b)

(a)

Figure 3. 12 Through-thickness distribution of T, and p at the mid-section of the


specimen when p reaches 0.14, (a) NT specimen, (b) Lindholm specimen
We can see that once again p remains constant through thickness for the
Lindholm specimen but it shows thickness-variation in the NT specimen, with the lowest
point at about the mid-radius. T also exhibits more variation in the NT specimen,
although the Lindholm specimen does show a small decrease in going from the inner to
outer radius. For this analysis, the modified Lindholm specimen exhibits more variation
in Lode parameter than the NT specimen since the value for the modified Lindholm
specimen is approaching -1.
Uniformity is one desirable character of a calibration specimen, another one is to
have minimal variation in triaxiality, Lode angle and effective plastic strain as plastic
deformation evolves.

Ideally, these parameters would be stationary during plastic


51

deformation all the way to failure. This would eliminate the need to calculate average
values by integrating over strain. The evolution of triaxialty and Lode parameters and

with increasing effective plastic strain is shown in Figure 3. 13 at tension-torsion ratios


of = 0.194 for the NT specimen and = 0.068 for the modified Lindholm specimen.
0.5

0.5

0.25

0.25

T
0

-0.25

0.25

Effective Plastic Strain

-0.25

0.5

0.25

Effective Plastic Strain

0.5

(b)

(a)

Figure 3. 13 Evolution of T, Lode parameter , and with increasing effective plastic


strain, (a) NT, (b) Lindholm specimens.
It is apparent that the modified Lindholm specimen quickly approaches nearly
stationary values as plastic strain progresses.
In conclusion, the numerical analysis revealed that the modified Lindholm
specimen can provide relatively uniform stress and strain distribution in the gage section
comparing with the NT specimen. Also, triaxiality and Lode angle remain relatively
constant as plasticity deformation evolves up to failure.

3.4.2 Materials, Specimens and Experiments


All the experimental tests here were done by NSWCCD (Naval Surface Warfare
Center, Carderock Division). Three materials considered in this study are: a 5083
52

aluminum alloy, a Nitronic 40 stainless steel, and Zircaloy-4 (a zirconium alloy). The
chemical compositions of these alloys are given in Table 3-1.
The aluminum specimens were machined from a 25 mm thick plate, with axial
direction oriented transversely to the rolling direction. The Nitronic 40 specimens were
extracted in the radial direction from a forged disk and the Zircaloy specimens were
extracted from wrought material in the longitudinal direction. The Zircaloy was heattreated to produce a random texture on a macroscopic scale. The Youngs modulus and
Poissons ratio for these materials at room temperature are also provided in Table 3-2.
Table 3-1. Chemical composition (in weight percent)
5083
aluminum

Zircaloy

Nitronic 40

Si

0.20

0.03

Fe

0.21

Fe

0.35

Cr

20.2

Sn

1.53

Cu

0.052

Mn

8.8

Cr

0.11

Mn

0.69

0.34

0.13

Mg

4.41

Ni

6.8

Zr

balance

Cr

0.086

0.020

Zn

0.081

0.002

Ti

0.065

Si

0.67

Al

balance

Fe

balance

53

Table 3-2. Elastic properties


5083 aluminum

Nitronic 40

Zircaloy

68.4 GPa

194.5 GPa

99.6 GPa

0.3

0.288

0.34

The test matrix includes smooth round tensile bars, notched tensile round bar
specimens for the Aluminum 5083 only, cylindrical compression specimens with
different length/diameter (L/D) ratios, grooved plane strain specimens for the Aluminum
5083 only, the Lindholm- type thin- wall torsion specimens [63] loaded in pure torsion,
torsion-tension for the Aluminum 5083 specimens only, and torsion-compression for the
Nitronic 40 and Zircaloy-4 specimens only. The sketches for these specimens are shown
in Figure 3.14. All tests are performed at room temperature and are considered to be
quasi-static.
For the 5083 aluminum alloy, round bar tensile tests, notched round bar tests,
compression tests, torsion tests, grooved plane strain tests and torsion-tension tests are
conducted. The gage section diameter of the tensile specimen is 6.35 mm. All the notched
bars have the same diameters of 15.2 mm in the smooth sec tions and 7.6 mm at their
notched cross section. Three notch radii, 1.27 mm, 2.54 mm and 6.35 mm, are considered
for specimen D, B and E respectively (Table 3-3). The diameter of the compression
specimens is 8 mm with L/D = 0.75. The torsion specimen is a hollow cylinder having an
inner diameter of 13.1 mm and outer diameter of 23.8 mm, with the gage section having a
length of 2.54 mm and wall- thickness of 0.75 mm. The overall dimensions for the
circular-grooved plane strain specimens are 203.2 mm 31.8 mm 6.1 mm and the
54

thinnest cross section is 2 mm. Three different groove radii, 2.03 mm, 5.08 mm and 16.26
mm, are considered for specimens F, G and H respectively (Table 3-4). The dimensions
of the torsion-tension specimens are the same as the pure torsio n specimen except a
slightly larger outer diameter (25.4 mm) at both ends. Table 3-5 lists the ratio between
the applied tensile displacement and applied twist angle for each of the tension-torsion
specimens.

Figure 3. 14 Sketches of a round tensile bar, a notched round bar specimen, a


compression specimen with L/D = 0.75, a grooved plan strain specimen and a torsion
specimen.
Table 3.3 Notch radii of the notched round bars
Specimen

B-notch

D-notch

E-notch

Notch radius (mm)

2.54

1.27

6.35

55

Table 3.4 Groove radii of the plane strain specimens


Specimen

F-groove

G-groove

H-groove

Groove radius (mm)

2.03

5.08

16.26

Table 3.5 Ratios of the applied tensile displacement and applied twist angle used in the
tension-torsion tests
Specimen

TT-17

tensile displacement /
0.23
twist angle (mm/radian)

TT-19

TT-13

TT-16

TT-15

TT-14

0.69

0.92

1.16

1.38

2.54

For the Nitronic 40 stainless steel and Zircaloy, tensile tests, compression tests,
torsion tests and torsion-compression tests are conducted. The gage section diameter of
the tensile specimen is 12.7 mm. Two types of compression specimens, one with L/D =
1.5 and the other with L/D = 0.75, are tested. The diameter of both compression
specimens is 8 mm. The gage section length and wall- thickness of the pure torsion and
torsion-compression specimens are 2.54 mm and 0.7366 mm respectively. For the
torsion-compression tests, a solid central pin (sized within 0.02 mm of the specimen inner
diameter) is inserted in the central hole to provide stability against inward buckling.
The tensile tests are conducted at a nominal strain rate of 10 -3 /s in a computercontrolled test machine under crosshead displacement control. Specimens are
instrumented with a 25 mm extensometer and the elapsed time, crosshead displacement
(measured by a linear voltage differential transformer (LVDT)), force from the machine

56

load cell, and the axial displacements from the extensometer are recorded at
approximately 100 Hz.
The compression tests are performed on a uniaxial servo-hydraulic MTS load
frame with MTS 643.10A-03 compression platens. Each specimen is mounted between
two hardened steel parallel blocks with integral knife edges before being placed between
the machines platens. The parallel blocks have integral knife edges that accept an MTS
clip gage. The clip gage acts as an extensometer to locally measure the deformation of the
specimen, whereas the machines LVDT measures the deformation of the entire load
train. Specimens are compressed under LVDT displacement control with a constant
quasistatic cross- head rate of 0.12 mm/s for the L/D = 1.5 specimens and 0.06 mm/s for
the L/D = 0.75 specimens. Both actuator rates correspond to a nominal elastic strain rate
of 10-2 /s. Data are acquired at 20 Hz for the running time, LVDT displacement, clip gage
displacement, and load and stored in tab-delimited ASCII text files.
The pure torsion, torsion-tension and torsion-compression tests are conducted
using a two-axis MTS servo-hydraulic load frame that is capable of independent and
simultaneous application of tension or compression via an axial actuator and torsion via a
rotational actuator. In order to more accurately monitor and control the deformation in the
specimen gage section, a local transducer that measures axial displacement plus rotation
across the gage section is used. The transducer uses two capacitive displacement probes
to achieve non-contacting measurement of both contraction and rotation. All tests are
conducted under computer control, where axial displacements and rotation are ramped
with time to obtain a specific tension/torsion or compression/torsion ratios and strain
57

rates. The data file generated for each test contains running time, LVDT displacement,
axial probe displacement, axial force, RVDT (rotary voltage differential transformer)
angle, angle probe displacement, and torque.

3.4.3 Finite Element Procedure


The finite element software ABAQUS [108], which adopts an updated
Lagrangian formulation to solve large deformation elasto-plasticity problems, is used to
analyze all specimens. The plasticity models described in the previous section 3.2 are
implemented in ABAQUS via a user defined subroutine following the procedures of Kim
and Gao [87], as described in section 3.3. The finite element implementation includes
using the backward Euler method to integrate the constitutive equations and the
formulation of the consistent tangent moduli.
Both tension and compression specimens are axisymmetric and the four-node,
axisymmetric hybrid solid elements (CAX4H) are used in the finite element analyses. For
the grooved plane strain specimens, the 3D 8- node brick elements with hybrid integration
(C3D8H) are used. For the torsion specimen, torsion-tension specimen and torsioncompression specimen, an additional degree of freedom needs to be added to the
axisymmetric element to handle the twist and the CGAX4H element in ABAQUS is
designed to satisfy this purpose.
Usually the symmetry conditions allows for only 1/4 or 1/8 of the specimen being
modeled. Figure 3.15 (a) shows the finite element mesh for a notched round bar tensile
specimen, figure 3.15(b) gives the finite element mesh of grooved plane strain specimen.
58

Figure 3.15 (c) shows the finite element mesh of the compression specimen with
L/D=0.75. For compression tests, the compression blocks are modeled as two rigid
surfaces and the frictional surface contact is introduced to model the interaction between
the compression block and the specimen. Friction coefficients between 0.05 to 0.2 are
found to have little effect on the numerical predictions before the specimen deformation
reaches extremely high levels. Since the exact friction coefficient is unknown and
difficult to obtain, a value of 0.05 is used in the analyses presented in this paper. Figure
3.15 (d) provides a typical mesh of the torsion specimen with a central pin. The central
pin is simulated by a rigid circular surface right next to the inner surface of the specimen
mesh. The surface to surface contact is defined with the interaction prope rties as
frictionless for the tangential behavior and hard contact allowing separation after contact
for the normal behavior.
Two

nondimensional

parameters

defined

as

T h / I 1 /(3 )

and

27 J 3 /( 2 3 ) are used here to describe the stress state in test specimens, where T is
referred to the stress triaxiality ratio and is related to the Lode angle. It is worth noting
that these two parameters are based on the current equivalent stress, not the von Mises
effective stress anymore. For the smooth tensile specimen, T = 1/3 and 1 prior to the
onset of necking and for the compression specimen, T 1 / 3 and 1 prior to the
onset of barreling. In the thin-walled gage section of the torsion specimen, both T and
are zero while the values of T and vary with the applied torsion-tension or torsioncompression ratio for the torsion-tension or torsion-compression specimens.

59

(a)

(b)

(c)
(d)
Figure 3. 15 Typical finite element meshes for (a) a notched round bar specimen, (b) a
grooved plane strain specimen, (c) a compression specimen with L/D = 0.75, and (d) a
torsion-compression specimen with a central pin.
The experimental and numerical results presented in sections 3.4.3-3.4.5
demonstrate that the plasticity models defined in Section 3.2 better model the plastic
60

behavior of these three alloys than the classical J2 plasticity theory. The model
parameters are determined by adjusting their values until best matches between the
predicted and measured load vs. displacement and/or torque vs. twist angle curves of a
tension test, a compression test and a test including torsional load are obtained.
Comparisons between numerical predictions and experimental measurements for
additional specimens serve as verifications of the calibrated models.

3.4.4 Plasticity response of an aluminum 5083


If the material obeys the J2 flow plasticity theory, one should be able to use the
finite element analysis with the stress-strain curve obtained from the uniaxial tension test
to predict the materials behavior at any stress state. When the material experiences large
plastic deformation, the true stress-strain curve should be used in the finite element
analysis. Assuming plastic incompressibility, the true stress-strain curve can be obtained
from the measured engineering stress-strain curve according to true eng (1 eng ) and

true ln(1 eng ) for loads and displacements equal to or less than the point of plastic
instability at which necking is observed. Figure 3.16 compares the numerically predicted
and experimentally measured load-displacement and torque-twist angle curves of the
tensile specimen, the compression specimen with L/D = 0.75, and the torsion specimen,
respectively. The numerical results are obtained by using the classical J2 flow plasticity
theory with the stress-strain curve obtained from the uniaxial tension test. In Figure 3.16,
the thicker line marked as J2 represents the finite element result using the J2 flow
theory and the thinner lines represent the experimental measure ments. The good
agreement between the numerical result and the experimental data for the uniaxial tensile
61

shown in Figure 3.16 (a) is not surprising because the stress-strain curve used in the finite
element analysis was extracted from the measured load-displacement curve of the tensile
specimen. Agreement between numerical and experimental results for the compression
test shown in Figure 3.16 (b) implies that the plastic response of aluminum 5083 is not
dependent on hydrostatic stress. However, Figure 3.16(c) demonstrates that the J2
plasticity theory significantly over-predicts the torque vs. twist angle response of the pure
torsion specimen. The result shown in Figure 3.16 (c) indicates the plastic response of

12
10
8
6
4
2
0

-150
Load (kN)

Load (kN)

aluminum 5083 is J3 dependent.

EXP1
EXP2
EXP3
J2

-100

Experiments

J2

-50
0

Displacement (mm)

-1

-2

-3

-4

-5

Displacement (mm)

(a)

(b)

Torque (kNmm)

75
50
25

Experiments

J2

0
0

0.2

0.4

Twist Angle (radian)

(c)
Figure 3. 16 Comparisons of load vs. displacement and/or torque vs. twist angle
responses between the experimental data and the J2 model prediction for aluminum 5083:
(a) the round tensile specimen, (b) the compression specimen with L/D = 0.75, and (c) the
torsion specimen (Experimental data were from [111]).

62

It is found that the yield function and flow potential given by Eqs. (3.11) and
(3.13) can be used to describe the plasticity behavior of the 5083 aluminum alloy.
Because the plasticity response is independent of I1 , parameters a1 and a2 are zero. The
values of b1 and b2 are determined so that the model-predicted axial force vs. axial
displacement and torque vs. twist angle responses for the torsion-tension specimen match
the experimental records. The parameters so determined for this material are a1 = a2 = 0,
b1 = -60.75 and b2 = -50. As a result, the yield surface is still cylinder-like in the principal
stress space but the cross-section is no longer circular. Figure 3.17(a) shows the
projection of the yield surface on the -plane. Since the difference between b1 and b2 is
not significant, the plot of the flow potential as shown in Figure 3.17 (b) is almost the
same as the yield surface shown in Figure 3.17 (a). Figure 3.17 (c) show the equivalent
stress vs. equivalent plastic strain curve, ( p ) , describing the hardening behavior.
Using the plasticity model defined by Eqs. (3.11) and (3.13) with the calibrated
parameters and the stress-strain curve given in Figure 3.17 (c), the tension specimen,
compression specimen, torsion specimen and torsion-tension specimen are re-analyzed
and the numerically predicted and experimentally measured load vs. displacement and
toque vs. twist angle curves for these specimens are compared in Figure 3.18. The tensile
tests have seven specimens, compression tests have three specimens, the torsion tests
have ten specimens and the torsion-tension test has one specimen. The experimental data
are represented by thinner lines while the finite element result is represented by a thicker
line. In general, the numerical results show better agreement with experimental data than
J2 plasticity theory.
63

3 /

3 /

1 /

1 /

2 /
(a)

2 /
(b)

(MPa)

400
300
200
100

Aluminum 5083

0
0

0.05

0.1

0.15

p
(c)
Figure 3. 17 (a) Projection of the yield surface of aluminum 5083 on the -plane; (b) Plot
of the flow potential; (c) The equivalent stress vs. equivalent plastic strain curve
describing the strain hardening behavior of the material.

64

Load (kN)

Load (kN)

-150

12
10
8
6
4
2
0

Experiments

-100
-50

FEA
0

EXP1
EXP2
EXP3
FEA

Displacement (mm)

-1

-2

-3

-4

-5

Displacement (mm)

(a)

(b)

Torque (kNmm)

75
50
25

Experiments

FEA

0
0

0.1

0.2

0.3

0.4

Twist Angle (radian)

50

40

Torque (kNmm)

Load (kN)

(c)

4
3

TT-15
FEA

2
1

30
20

TT-15
FEA

10
0

0
0

0.1

0.2

Displacement (mm)

0.06

0.12

0.18

Twist Angle (radian)

(d)

(e)

Figure 3. 18 Comparisons of the predicted load vs. displacement and/or toque vs. twist
angle responses using the calibrated I1 -J2-J3 plasticity model with experimental records
for aluminum 5083: (a) the round tensile specimen; (b) the compression specimen with
L/D=0.75; (c) torque vs. twist angle response of the pure torsion specimen; (d) axial force
vs. axial displacement response of the torsion-tension specimen (TT-15); (e) torque vs.
twist angle response of the torsion-tension specimen (TT-15) (Experimental data were
from [111]) .
65

With the parameters of the plasticity model being calibrated, the next question
needs to be answered is whether this model correctly predicts the material response under
complex stress states. To this end, notched round bars with different notch radii, plane
strain specimens with different groove radii and modified Lindholm torsion specimens
subjected to different tension-torsion ratios are tested and analyzed and the numerical
predictions are compared with the experimental records. For notched round bars and
grooved plane strain specimens, the applied tensile force vs. extensometer gage
displacement response is monitored. For tension-torsion specimens, both applied axial
force vs. axial displacement and applied torque vs. twist angle responses are monitored.
Figure 3.19 shows typical comparisons again, model predictions agree with
experimental measurements very well. Similar comparisons between experimental results
and numerical predictions are observed for all other specimens (two notched round bar
specimens, two groove plane strain specimens and torsion-tension specimens with 4
different ratios of the applied tensile displacement and applied twist angle) in the test
matrix.

66

20

30

E-Notch
Load (kN)

Load (kN)

10

Experiments

FEA

20
15

0.2

0.4

Experiments

10

FEA

0
0

G-Groove

25

15

0.6

0.8

Displacement (mm)

0.1

0.2

0.3

0.4

Displacement (mm)

(b)

(a)
50

TT-16
Axial Force (kN)

TT-16

Torque (kNmm)

40
30
20

EXP

10

FEA

0
0

0.05

0.1

0.15

Twist Angle (radian)

EXP
FEA

0
0

0.05

0.1

0.15

Axial Displacement (mm)

(d)

(c)

Figure 3. 19 Comparisons of the numerical predictions and experimental records: (a)


notched round bar (E-Notch); (b) plane strain specimen (G-Groove); (c) torque vs. twist
angle response for torsion-tension test (TT-16); (d) axial force vs. axial displacement
response for torsion-tension test (TT-16) (Experimental data were from [111]).
To further justify the proposed plasticity model, Figures 3.20(a) and (b) compare
the numerical predictions using the classical J2 -flow theory and the proposed I1 -J2 -J3
model for the tension-torsion test TT-16 with experimental results. In these figures, we
also included the comparison between the non-associated flow rule (Non-AFR) and the
associated flow rule (AFR). As can be seen from the plots, only the proposed I1 -J2 -J3
model with the calibrated parameters (non-associated flow rule) leads to good agreement
67

with experimental records for both torque vs. twist angle and axial force vs. axial
displacement. The associated flow rule gives acceptable torque vs. twist angle prediction
but unsatisfying axial force vs. axial displacement prediction.
6

30
20
EXP
Non-AFR

10

Axial Force (kN)

Torque (kNmm)

40

EXP
Non-AFR
J2
AFR

J2
AFR

0
0

0.05

0.1

0
0

0.15

Twist Angle (radian)

0.05

0.1

0.15

Axial Displacement (mm)

(b)

(a)

Figure 3. 20 Comparisons of the numerical predictions using the classical J2 -flow theory
and the proposed I1 -J2-J3 model for the tension-torsion test TT16 with experimental
results: (a) torque vs. twist angle; (b) axial force vs. axial displacement (Experimental
data were from [111]).

3.4.5 Plasticity response of a Nitronic 40


To check if this material obeys the classical J2 flow plasticity theory, finite
element analyses of the compression specimen and the torsion-compression specimen are
performed using the J2 plasticity model with the stress-strain curve obtained from the
uniaxial tensile test. Figure 3.21 compares the numerical and experimental loaddisplacement and/or torque-twist angle curves for the tensile specimen, the compression
specimen with L/D = 1.5 and the torsion-compression specimen with the central pin.
Good agreement between the numerical and experimental results for the compression test
68

shown in Figure 3.21 (b) suggests that the I1 -dependence is negligible for Nitronic 40.
However, Figures 3.21 (c) and (d) indicate that the J2 plasticity theory significantly under
predicts the axial force vs. axial displacement and torque vs. twist angle responses of the
torsion-compression specimen. The results shown in Figures 3.21 (c) and (d) suggest that
the plastic response of this material is J3 dependent.

-300

80
60

Load (kN)

Load (kN)

100
RT-3
RT-4
RT-5
J2

40
20
0
0

10

20

C15-1
C15-2
C15-5
J2

-200
-100
0

30

40

Displacement (mm)

-2

-4

-8

-10

Displacement (mm)

(a)

(b)
250
Torque (kNmm)

-40
-30
Load (kN)

-6

-20

N40-t-4
N40-t-5
N40-t-8
J2

-10
0
0

-0.2

-0.4

-0.6

-0.8

200
150
N40-t-4
N40-t-5
N40-t-8
J2

100
50
0

-1

0.0

0.2

0.3

0.5

Twist Angle (radian)

Displacement (mm)

(c)

(d)

Figure 3. 21 Comparisons of load vs. displacement and/or torque vs. twist angle
responses between the experimental data and the J2 plasticity theory predictions for
Nitronic 40: (a) the tensile specimen, (b) compression specimen with L/D = 1.5, (c) axial
force vs. axial displacement response for the torsion-compression specimen, and (d)
toque vs. twist angle response for the torsion-compression specimen (Experimental data
were from[112]).

69

Based on visual comparison of the model and test results, the I1 -J2 -J3 dependent
plasticity model given by Eqs. (3.14) and (3.16) is found appropriate for this material.
The set of parameters, a1 = a2 = 0, b1 = b2 = 1.5, result in best matches between the
numerically predicted and experimentally measured load vs. displacement and/or toque
vs. twist angle responses for the tensile specimen, the compression specimen with L/D =
1.5, and the torsion-compression specimen respectively. Figure 3.22(a) shows the
projection of the yield surface on the -plane. Since this material follows the associated
flow rule, b1 = b2 , the flow potential is the same as the yield function. Figure 3.22(b)
shows the equivalent stress vs. equivalent plastic strain curve, ( p ) , describing the
hardening behavior. Figure 3.23 compares the numerical results using the calibrated I1 -J2J3 plasticity model with experimental records for the tensile specimen, the compression
specimen with L/D = 1.5, and the torsion-compression specimen with the central pin
respectively.

(MPa)

3 /

1500
1000
500

Nitronic 40

1 /

2 /
(a)

0.2

0.4

0.6

(b)

Figure 3. 22 (a) Projection of the yield surface of Nitronic 40 on the -plane, and (b) the
equivalent stress vs. equivalent plastic strain curve describing the strain hardening
behavior of the material.

70

-300

100
60

Load (kN)

Load (kN)

80
RT-3
RT-4
RT-5
FEA

40
20
0
0

10

20

C15-1
C15-2
C15-5
FEA

-200
-100
0

30

Displacement (mm)

40

-4

-6

Displacement (mm)

-8

-10

(b)

-50

350

-40

280

Torque (kNmm)

Load (kN)

(a)

-2

210

-30

N40-t-4
N40-t-5
N40-t-8
FEA

-20
-10
0
0

-0.5

-1

N40-t-4
N40-t-5
N40-t-8
FEA

140
70
0
0.0

-1.5

0.2

0.4

0.6

0.8

Twist Angle (radian)

Displacement (mm)

(c)

(d)

Figure 3. 23 Comparisons of the predicted load vs. displacement and/or toque vs. twist
angle responses using the calibrated I1 -J2-J3 plasticity model with experimental records
for Nitronic 40: (a) the tensile specimen, (b) the compression specimen with L/D = 1.5,
(c) axial force vs. axial displacement response for the torsion-compression specimen, and
(d) torque vs. twist angle response for the torsion-compression specimen (Experimental
data were from[112]).
This calibrated plasticity model is then applied to predict the responses of the
compression specimen with L/D = 0.75 and the pure torsion specimen. Figure 3.24
compares the numerical predictions with experimental measurements and it can be seen
that the numerical results match with the experimental data reasonably well and are a
better match than the J2 model shown in Figure 3.21.

71

As shown in Figure 3.23 (c) the model results deviate from the load/displacement
test data for the torsion-compression test. The deviations are hypothesized to result from
behavior at the transition between the specimen reduced section and ga ge section. This
transition region acts as a strain concentration in the torsion-compression test, which can
lead to complications in the finite element model. Even with this complication, the I1 -J2 J3 model produces better matches to test data (seen in Figure 3.23 and Figure 3.24) than
the J2 model.
200
C75-2
C75-3
C75-4
FEA

-300
-200

Torque (kNmm)

Load (kN)

-400

-100

150
T-1
T-2
T-3
FEA

100
50
0

0
0

-1

-2

-3

-4

0.0

-5

0.2

0.4

0.6

0.8

Displacement (mm)

Twist Angle (radian)

(a)

(b)

1.0

Figure 3. 24 Comparisons of numerical predictions using the calibrated I1 -J2-J3 plasticity


model with experimental data for Nitronic 40: (a) load vs. displacement response of the
compression specimen with L/D=0.75; (b) torque vs. twist angle response of the pure
torsion specimen (Experimental data were from[112]).

3.4.6 Plasticity response of a Zircaloy 4


Again, it is necessary to determine if I1 and/or J3 has any effect on the plastic
response of this material. The issue of pressure sensitivity (I1 dependency) is investigated
first. This can be done by checking if the finite element analysis using the J2 flow
plasticity theory and the stress-strain curve obtained from the uniaxial tensile test can
correctly predict the load vs. displacement response of the compression specimen.
72

Figures 3.25(a) and (b) compare the numerical and experimental load-displacement
curves of the tensile specimen and the L/D = 1.5 compression specimen. The J2 model
prediction for the tensile specimen is excellent because the stress-strain curve used in the
analysis was extracted from the tensile test data. But the J2 plasticity model significantly
under predicts the load required in the compression test, and over predicts the torque
versus twist data for the torsion test, suggesting that the plasticity behavior for Zircaloy-4
is hydrostatic pressure dependent. Using the plasticity model given by Eqs. (3.11) and
(3.13), we find a1 = a2 = 0.083 leads to a better match between the numerically predicted
and experimentally measured load vs. displacement response for the compression test
than for the J2 plasticity theory.

73

-100

C15-1
C15-2
C15-3
C15-4
J2

-80

80

Load (kN)

Load (kN)

100
60
RT-2
RT-3
RT-4
J2

40
20
0
0

10

Displacement (mm)

-60
-40
-20
0
0

15

(a)

-1

-2

-3

-4

Displacement (mm)

-5

(b)

Torque (kNmm)

100
80
60
40

T-1
T-2
J2

20
0
0.0

0.1

0.2

0.3

Twist Angle (radian)

(c)
Figure 3. 25 Comparisons of the experimental data and the J2 model predictions
Zircaloy: (a) load vs. displacement response of the tensile specimen, (b) load
displacement response of the compression specimen with L/D = 1.5, and (c) torque
twist angle response between experimental data and the I1 -J2 model prediction for
torsion specimen (Experimental data were from[112])

for
vs.
vs.
the

The pressure sensitive plasticity model with a1 = a2 = 0.083 gives excellent


predictions of the load-displacement responses of the tensile specimen and the L/D = 1.5
compression specimen (see Figure 3.27 (a) and (b)), but it over predicts the required
torque for the torsion specimen, Figure 3.25 (c). This indicates the plasticity behavior
also depends on J3 .

74

Therefore, both I1 and J3 terms should be included in the plasticity model, and it is
found, the yield function and flow potential given by Eqs. (3.11) and (3.13) with a set of
parameters, a1 = a2 = 0.083 b1 = -20, b2 = -40, result in best matches between the
numerically predicted and experimentally measured load vs. displacement and/or toque
vs. twist angle responses for the tensile specimen, the compression specimen with L/D =
1.5, and the torsion-compression specimen with the central pin respectively. Because of
the hydrostatic stress effect, the yield surface changes from a cylinder to a cone in the
principal stress space. Figure 3.26 (a) shows the yield surface in the 3D space of principal
stresses and Figure 3.26 (b) shows the comparison of the equivalent stress vs. equivalent
plastic strain curve, ( p ) , between the one used in the calibrated I1 -J2-J3 plasticity
model and the original stress-strain curve derived from the round bar tensile tests. The
thicker line (FEA) in the Figure 3.26 (b) represents the effective stress-strain curve used
for the calibrated I1 -J2-J3 plasticity model and the thinner line (J2 ) represents the stressstrain curve obtained from the uniaxial round bar tension test. Since the value difference
between b1 and b2 is not very large, the shape of the flow potential is similar to the yield
surface shown in Figure 3.26(a). It is worth noting that, because of the I1 effect, the
material is no longer plastically incompressible. The stress-strain curve used in the
calibrated I1 -J2 -J3 model is different from the one converted from the engineering stressstrain curve obtained from the tensile test.

75

(MPa)

3 /

1 /

2 /

(a)

(b)

Figure 3. 26 (a) Yield surface of Zircaloy, (b) Equivalent stress vs. equivalent plastic
strain curve describing the strain hardening behavior of the material
Figure 3.27 compares the predicted load vs. displacement and/or torque vs. twist
angle responses using the calibrated I1 -J2 -J3 plasticity model with the experimental
records for the tensile specimen, the compression specimen with L/D = 1.5, and the
torsion-compression specimen respectively. From Figure 3.27(c) it can be seen that the
model prediction does not provide a perfect match to the torsion-compression test data.
This is again believed to result from the strain concentration at the transition area
between the specimen reduced section and gage section. However, the I1 -J2 -J3 model
predictions show better agreement with the other loading conditions (comparing Figure
3.27 and 3.28 with Figure 3.25).

76

-100

80

-80

60

Load (kN)

Load (kN)

100

RT-2
RT-3
RT-4
FEA

40
20
0
0

C15-1
C15-2
C15-3
C15-4
FEA

-60
-40
-20

10

15

Displacement (mm)

-24

120

Torque (kNmm)

Load (kN)

150

FTOR-4
FTOR-5
FTOR-6
FEA

-6
0

-0.2

-0.4

-3

-4

-5

(b)

-30

-12

-2

Displacement (mm)

(a)

-18

-1

90
FTOR-4
FTOR-5
FTOR-6
FEA

60
30
0

-0.6

0.0

0.1

0.2

0.3

Displacement (mm)

Twist Angle (radian)

(c)

(d)

0.4

Figure 3. 27 Comparisons of the predicted load vs. displacement and/or torque vs. twist
angle responses using the calibrated I1 -J2-J3 plasticity model with experimental records
for (a) the tensile specimen, (b) the compression specimen with L/D = 1.5, (c) axial force
vs. axial displacement response for the torsion-compression specimen, and (d) torque vs.
twist angle response for the torsion-compression specimen (Experimental data were
from[112]).
This calibrated model is then applied to predict the plastic responses of the
compression specimen with L/D = 0.75 and the pure torsion specimen. Figure 3.26 (a)
compares the numerically predicted load vs. displacement response with experimental
measurements for the L/D = 0.75 compression specimen and Figure 3.26(b) compares the
predicted torque vs. twist angle response with experimental records for the pure torsion
77

specimen. The numerical results match with the experimental data very well and are an

-200

100

-160

80

Torque (kNmm)

Load (kN)

improved fit compared to the J2 model.

-120
C75-2
C75-3
C75-4
FEA

-80
-40
0
0

-1

-2

-3

60
40

T-1
T-2
FEA

20
0
0.0

-4

Displacement (mm)

0.1

0.2

0.3

Twist Angle (radian)

(a)

(b)

Figure 3. 28 Comparisons of the experimental data with the numerical results computed
using the calibrated I1 -J2-J3 plasticity model for Zircaloy: (a) load vs. displacement
responses of the compression specimen with L/D = 0.75, (b) torque vs. twist angle
response of the pure torsion specimen (Experimental data were from[112]).

3.5 Concluding Remarks


In this chapter, we describe a series of plasticity models for isotropic materials, of
which the yield function and the flow potential are dependent on three stress invariants,
I1 , J2 and J3 . In particular, these functions are specified as first order homogeneous
function of stress and the ranges of the model constants are determined to satisfy the
convexity requirement. This I1 -J2-J3 plasticity model is implemented into the finite
element software, ABAQUS, via a user defined subroutine. Three materials, a 5083
aluminum alloy, Nitronic 40 stainless steel and Zircaloy-4, are investigated in this study
to illustrate the applications of the proposed plasticity model. Torsion-tension specimen is
designed to provide stress states with low triaxiality values. This modified Lindholm
78

torsion-tension specimens are compared with NT specimen to show more uniform


distribution of the stress state and strain along the gage section. A series of experiments,
including tension, compression, torsion, combined torsion-tension and combined torsioncompression, are conducted and analyzed. The function forms of the yield surface and
flow potential are determined, and the model parameters are calibrated for these three
materials respectively. For aluminum 5083, the effect of I1 on plasticity response is
negligible, the effect of J3 is significant, and the best matches to test data were obtained
using a non-associated flow rule. For Nitronic 40, the effect of I1 is also negligible, the
effect of J3 is significant, and the associated flow rule is appropriate. Fo r Zircaloy, both I1
and J3 have strong effects on the plastic response and the best matches to test data were
obtained with a non-associated flow rule. The calibrated models provide a better
prediction of the load vs. displacement and/or torque vs. twist a ngle responses for a range
of experiments than the J2 plasticity theory.

79

CHAPTER IV
MODIFIED POUROUS GURSON MODEL

4.1 Introduction
Ductile fracture of many structural materials is a result of void nucleation, growth
and coalescence. The constitutive description of this mechanism has received a great deal
of attention in the past thirty years, which leads to various forms of porous material
models being developed to describe void growth and the associated macroscopic
softening. One of the famous porous plasticity models was due to Gurson [76] with
modifications by Tvergaard and Needleman [77, 78, 102]. In the Gurson-TvergaardNeedleman model, an extra internal variable, the void volume fraction (f), is introduced
to capture the growth of cavities and its effect on material behavior. It is important to
notice that the Gurson-Tvergaard-Needleman model reduces to the J2 -flow theory of
plasticity with isotropic hardening in the absence of voids (f = 0).
In this chapter, the Gurson- Tvergaard-Needleman model is extended to include
the hydrostatic stress and Lode parameter effects. No longer employing the von Mises
plasticity criterion, the matrix material in this Gurson-Tvergaard-Needleman model obeys
the I1 -J2 -J3 plasticity criterion proposed in the chapter III.

This extended Gurson-

Tvergaard-Needleman plasticity model is implemented into ABAQUS [108] to

80

conduct a series of parametric study to show the effect of extended model on predicted
material response. Since only the plastic response of the material is studied, void growth
procedure is the only stage considered in our numerical implementation. Void nucleation
and void coalescence stages are not implemented into the ABAQUS/ UMAT.

4.2 Modified Gurson Model with I1 , J3 Effects


The Gurson- Tvergaard-Needleman (GTN) model is extended to include the
hydrostatic pressure and Lode parameter effect into the matrix material. Formulation and
implementation of this modified GTN model is listed in this section.

4.2.1 Modified Gurson Model Theory


Within the framework of the I1 -J2-J3 plasticity theory outlined in Section 3.2, the
yield function and flow potential of the Gurson-Tvergaard-Needleman model can be
modified as
2

q I
F
2q1 f cosh 2 1 1 q 12 f 2 0

2

(4. 1)

and
2

q I
G
2q1 f cosh 2 1 1 q 12 f 2 0

2

where

and

take

one

of

the

forms

(4. 2)

shown

in

Appendix

A,

i.e.,

F c1 (a1 I16 27 J 23 b1 J 32 )1 / 6 and G c2 (a2 I16 27 J 23 b2 J 32 )1 / 6 , I1 , J2 and J3 are three


invariants of the macroscopic stress, f is the current void volume fraction, is the
81

current yield stress of the matrix material, and q1 and q2 are parameters introduced by
Tvergaard [77, 102] to account for void interaction and matrix strain hardening. If a1 = b1
= a2 = b2 = 0, Eqs. (4.1) and (4.2) degenerate to the original Gurson-TvergaardNeedleman model.
Porous material models contain an additional state variable, f. Its worth noting
that matrix material in modified Gurson-Tvergaard-Needleman model is not following
von Mises yield criteria anymore, but following the proposed I1 -J2-J3 plasticity model.
With hydrostatic pressure under consideration, for this modified Gurson-TvergaardNeedleman model, the evolution equation for f can be obtained by considering the rate of
the net volume change

F
6
f (1 f )[kkp 3c2 a2 I15 p 6 ]
G

(4. 3)

By enforcing equality between the rates of macroscopic plastic work and the
matrix plastic dissipation, the matrix yield stress, , and the matrix plastic strain rate,
p , are coupled through

(1 f ) p ij ijp

(4. 4)

where the matrix material follows a prescribed hardening function ( p ) .

4.2.2 Numerical Implementation


Numerical implementation of the modified Gurson- Tvergaard-Needleman model
follows a similar procedure as outlined in Chapter III. Backward Euler method is
82

employed to integrate the constitutive equations and the consistent tangent moduli are
formulated according the numerical method provided by Kim and Gao (2005). Stress
update and formulation of the consistent tangent moduli are implemented in the
ABAQUS/UMAT.

4.3 Numerical examples using the Gurson- Tvegaard-Needleman model


In this section we conduct a series of parametric studies to illustrate the effect of
the modified Gurson-Tvergaard-Needleman model on predicted material response. Figure
4.1 shows a cubic element with dimensions of D0 D0 D0 . Displacement boundary
conditions are imposed on the element surfaces such that the macroscopic stress ratios, 1
= 11 / 22 and 2 = 33 / 22 , are kept constants during the entire deformation history
(which results in a constant stress triaxiality ratio and Lode angle). Faleskog et al. [113]
and Kim et al. [48] provide details of how to prescribe this kind of boundary cond itions.
The material parameters used in the analyses are E = 68.4 GPa, = 0.3, 0 = 207 MPa, N
= 0.125 and f 0 = 0.003, where E, , 0 , n and f0 represent the Youngs modulus, the
Poissons ratio, the yield stress, the strain hardening exponent and the initial void volume
fraction respectively. The q1 and q2 parameters in the Gurson-Tvergaard-Needleman
model are taken as q1 = 1.5 and q2 =1.

u 2 ; 22

u1 ; 11

u3 ; 33

Figure 4. 1 A cubic element.


83

For the first set of analyses, the boundary conditions are imposed such that 1 =
0.268 and 2 = 0.634, corresponding to a stress triaxiality ratio of 1 and Lode angle of 0.
When a1 = a2 = b1 = b2 = 0, the analysis results using our user subroutine are the same as
those obtained using the original Gurson-Tvergaard-Needleman model implemented in
ABAQUS, serving as a check of our numerical implementation.
Figure 4.2 illustrates the effect of I1 , where the dotted lines represent the
numerical results obtained using a1 = a2 = b1 = b2 = 0 and the solid lines represent the
numerical results obtained using a1 = a2 = 610-4 and b1 = b2 = 0. At the same applied
displacement in the x 2-direction (u2 ), both 22 and the void growth rate are lower when
the effect of I1 is taken into account.
Figure 4.3 demonstrates the effect of J3 by setting non-zero values for b1 and b2 .
Here the dotted lines show the results of a1 = a2 = b1 = b2 = 0 and the solid lines are
obtained by using a1 = a2 = 0 and b1 = b2 = -60.75. The analysis results show that
negative values of b1 and b2 lead to lower value of 22 and slower void growth rate.

84

22 / 0

0.03

0.02

a 1 = a 2 = b1 = b 2 = 0
a 1 = a 2 = 610-4
b1 = b2 = 0

0.01

0
0

0.1

0.2

0.3

u2 / D0

0.4

0.5

0.1

0.2

0.3

u2 / D0

(a)

0.4

0.5

(b)

Figure 4. 2 Comparison of the 22 / 0 vs. u 2 / D0 response and f vs. u 2 / D0 response


predicted using a1 = a2 = b1 = b2 = 0 (dotted lines) and a1 = a2 = 610-4 and b1 = b2 = 0
(solid lines).

22 / 0

0.02

0.01

a 1 = a 2 = b1 = b 2 = 0
a1 = a2 = 0
b 1 = b 2 = -60.75

0
0

0.1

0.2

0.3

u2 / D0
(a)

0.4

0.5

0.1

0.2

0.3

0.4

0.5

u2 / D0
(b)

Figure 4. 3 Comparison of the 22 / 0 vs. u 2 / D0 response and f vs. u 2 / D0 response


predicted using a1 = a2 = b1 = b2 = 0 (dotted lines) and a1 = a2 = 0 and b1 = b2 = -60.75
(solid lines).
85

The results shown in Figs. 4.2 and 4.3 are obtained using the associated flow rule.
Figures 4.4 and 4.5 compare the numerical results of the non-associated flow rule with
those of the associated flow rule. As shown in Figure 4.4, where the dotted lines are the
results of an associated flow rule (a1 = a2 = 610-4 and b1 = b2 = 0) and the solid lines are
the results of a non-associated flow rule (a1 = 610-4 and a2 = b1 = b2 = 0), increasing a2
from 0 to 610-4 leads to an increase of 22 and a decrease of void growth rate.
Figure 4.5 compares the results when b2 takes a different value than b1 . In these
analyses, a1 and a2 are taken as zero and the results using b1 = b2 = -60.75 (associated
flow rule) are denoted by the dotted lines while the results using b1 = -60.75 and b2 = 0
(non-associated flow rule) are denoted by the solid lines. Varying b2 from 0 to -60.75
results in negligible effect on 22 but a higher void growth rate.

22 / 0

0.03

0.02

a 1 = a 2 = 610-4
b1 = b2 = 0
a 1 = 610-4
a 2 = b 1 = b2 = 0

0.01

0
0

0.1

0.2

0.3

0.4

0.5

0.1

0.2

0.3

0.4

0.5

u2 / D0

u2 / D0
(a)

(b)

Figure 4. 4 Comparison of the 22 / 0 vs. u 2 / D0 response and f vs. u 2 / D0 response


predicted using a1 = a2 = 610-4 and b1 = b2 = 0 (associated flow rule; dotted lines) and a1
= 610-4 and a2 = b1 = b2 = 0 (non-associated flow rule; solid lines).
86

22 / 0

0.02

0.01

a1 = a2 = 0
b 1 = b 2 = -60.75
a1 = a2 = 0
b 1 = -60.75, b 2 = 0

0
0

0.1

0.2

0.3

u 2 / D0

0.4

0.5

0.1

0.2

0.3

0.4

0.5

u2 / D0

(a)

(b)

Figure 4. 5 Comparison of the 22 / 0 vs. u 2 / D0 response and f vs. u 2 / D0 response


predicted using b1 = b2 = -60.75 (dotted lines) and b1 =-60.75 and b2 = 0 (solid lines),
where a1 and a2 are taken as zero.

In the analyses presented above, the boundary conditions are imposed to keep 1 =
0.268 and 2 = 0.634. The stress state experienced by the material can be altered by
changing the boundary conditions so that different values of 1 and 2 are achieved. For
example, 1 = 0.4 and 2 = 0.4 corresponds to a stress triaxiality ratio of 1 and Lode angle
of -30. Figure 4.6 illustrates how this change of stress state (Lode angle changes from 0
to -30 while stress triaxiality ratio remains 1) affects 22 and the void growth rate, where
a1 = a2 = 610-4 and b1 = b2 = -60.75 are used in the calculations. In Figure 4.6, the dotted
lines represent the results of 1 = 0.268 and 2 = 0.634 (Lode angle equal to 0) and the
solid lines represent the results of 1 = 0.4 and 2 = 0.4 (Lode angle equal to -30). The

87

void growth rate is slightly higher and 22 becomes noticeably larger when the Lode
angle changes from 0 to -30.

22 / 0
3

0.02

2
0.01

1 = 0.268, 2 = 0.634

1 = 2 = 0.4

0.1

0.2

0.3

0.4

0.5

u2 / D0
(a)

0.1

0.2

0.3

0.4

0.5

u2 / D0
(b)

Figure 4. 6 Comparison of the 22 / 0 vs. u2 / D0 response and f vs. u2 / D0 response


when 1 = 0.268 and 2 = 0.634 (dotted lines) with those when 1 = 0.4 and 2 = 0.4 (solid
lines), where a1 = a2 = 610-4 and b1 = b2 = -60.75.

4.4 Conclusion
In this chapter, G-T-N model is extended to include the hydrostatic pressure and
Lode parameter effects in the matrix material. This modified G-T-N model is
implemented into ABAQUS to study the materials response under different parameter
values via a UMAT. In summary, the dependence of the matrix plasticity behavior on I1
and J3 results in significant changes in void growth and the macroscopic stress vs.
deformation response of porous materials. The modified Gurson-Tvergaard-Needleman
model provides a means to describe these changes.
88

CHAPTER V
PLASTICTY AND DUCTILE FRACTURE ANALYSIS FOR ALUMINUM 5083

5.1 Introduction
Ductile fracture usually means a type of fracture that has the void nucleation,
growth and coalescence stages. Coalescence of the growing voids is often considered as
the sign of the macroscopic crack initiation. A large deal of research has focused on
building the criterions for coalescence.
The most widely used coalescence criterion is the critical void volume fraction,
which assumes the coalescent stage occurs when the volume fraction of a cell reaches the
critical value, f c, no matter what the stress state condition is. But there is a distinct
disadvantage in this criterion. Since the void volume fraction, f, is related to the plastic
volume change, in the case of very low stress triaxiality state, which usually occurs in a
very thin plate or near the surface of a thick specimen, void volume fr action, f, in the
material does not increase much. Sometimes it never reaches the critical value. Therefore,
critical void volume fraction criterion cannot be applied in these kinds of problems. In
order to solve this issue, Nahshon and Hutchinson [114] proposed a modified Gurson
model to include a new effective damage accumulation law due to void distortion and
inter- void linking under shearing. Later, this model was further extended by Nahshon and
Xue [115] to incorporate the nucleation of voids, and the extended model is utilized to
89

model quasi-static punch-out experiments on DH36 steel. Another coalescence criterion


is based on the critical failure strain, which has been shown working very well at low
triaxiality loading [110].
The Gologanu-Leblond-Devaux (GLD) model provides a constitutive relation to
describe void growth and the associated macroscopic softening of material. It accounts
for the evolution of both void shape and void volume fraction. In our numerical analysis
here, critical failure strain, used as the coalescence criterion, is introduced into GLD
model. As indicated in the earlier literature review, both the stress triaxiality ratio and the
Lode parameter have significant effects on the ductile failure process for certain
materials. In this chapter, we will explore the stress state effects on plastic response and
ductile fracture for an Aluminum 5083 alloy. A function adopted by Xue and Wierzibicki
[52] will be employed to describe the triaxiality and Lode parameter effects on the failure
strain. The GLD porous model will be implemented into the commercial software
ABAQUS to illustrate the material damage evolution caused by the void growth and
stress state dependent failure strain criterion to predict void coalescence. The following
sections present the experimental and numerical work conducted by the author to
demonstrate the effects of the stress state on the plastic response and the ductile fracture
behavior of an aluminum 5083 alloy. These effects are characterized by the parameters T
(triaxiality) and (Lode parameter).

5.2 GLD model


The Gologanu-Leblong-Devaux models yield function is expressed as

90

2
2

' h X 2q g 1 g f cosh h g 1 q 2 g f 0

(5. 1)

where ij are the macroscopic stress components, f is the void volume fraction, S is the
void shape parameter, is the yield stress of the matrix material,

denotes the von

Mises norm, represents the deviatoric stress tensor, h is the generalized hydrostatic
stress

defined

by

h 2 xx zz 1 2 yy ,

and

is

defined

as

X 2 / 3 e y e y 1/ 3 ex ex 1/ 3 ez ez , where (ex, ey, ez) is an orthogonal basis


with ey parallel to the axisymmetric axis of the void and denotes tensor product. The
parameters C, , g, and 2 are functions of f and S and the heuristic parameter q depends
on initial void volume fraction, strain hardening exponent of the matrix material, S and
the macroscopic stress triaxiality factor T. Appendix B explains the parameters involved
in the GLD model in details.
With associative flow rule, the plastic flow function is identical with the yield
function. When material deforms plastically, the inelastic part of the deformation is
defined by :
ijp

G ( I 1 , J 2 , J 3 )
ij

(5. 2)

where ijp are the rates of the plastic strain components and is a positive scalar called
the plastic multiplier.

91

By enforcing equality between the rates of macroscopic plastic work and matrix
plastic dissipation, the matrix yield stress, , and the matrix plastic strain rate, p , are
coupled through

(1 f ) p ij ijp

(5. 3)

where the matrix material follows a prescribed hardening function ( p ) .


The evolution law for void volume fraction is determined by :

f ( f ) growth ( f ) nucleation

(5. 4)

where ( f ) growth and ( f )nucleation represent the rate of void volume fraction increase due to
void growth and void nucleation respectively. Void nucleation is not considered in our
current analysis, which leads to

f ( f ) growth (1 f ) kkp

(5. 5)

where kkp is the rate of plastic volume change.


As the macroscopic effective strain (Ee) of the element reaches the critical value
(Ec), void coalescence occurs and the element quickly loses its stress carrying capacity.
We adopt the f * function, introduced by Tvergaard and Needleman [78], to account for
the effects of rapid void coalescence at failure. It is worth noting that here f * is defined
differently from Tvergaard and Needleman [78] and it has been shown that this definition

92

of f * works better for low triaxiality situations [110]. After Ee reaches Ec, f is replaced
by f * in the Gologanu-Leblond-Devaux model, i.e.,

f,

f
Ee Ec
f c ( f u f c ) E ,
c

Ee Ec
Ee Ec

(5. 6)

In Eq. (5.6), f c is the void volume fraction at Ee = Ec, f u is the f * value at zero
stress, and is an adjustment parameter. For prolate voids, fu 1/ q , and for oblate
voids, fu (1 g gq) / q .
Kim and Gao [87] developed a generalized approach to formulate the consistent
tangent stiffness for complicated plasticity models. Using this approach, the GLD model
is implemented into ABAQUS via a user subroutine.

5.3 Specimen geometries and finite element procedures


All the tests were done by NSWCCD (Naval Surface Warfare Center, Carderock
Division). The material considered is an aluminum alloy 5083-H116, which was cold
worked to achieve its strength. All specimens are machined from a 25mm thick plate with
the tensile axis in the T-direction so that fracture occurs at the LS plane. The microstructural properties are obtained by analyzing the micrographs of polished and etched
orthogonal surfaces using an image processing program. It is fo und that the inclusion
shape can be approximated as oblate with an aspect ratio of approximately 0.6 and the
inclusion volume fraction and average spacing are about 3% and 20 m respectively (the
average long-axis of the inclusions is about 5 m). All the tests are performed at room
93

temperature and are considered to be quasi-static. The test matrix includes smooth and
notched round tensile bars, grooved plane strain specimens and the Lindholm-type
torsion specimen [63]. The specimen geometries and loading conditions are designed to
ensure failure occurs at different stress states. Figure 5.1 shows a smooth round bar, a
notched round bar, a grooved plane strain specimen and a torsion specimen. The diameter
of the gage section of the smooth round bar is 6.35 mm and the dimensions of the other
specimens are shown in Figure 5.2. Notched round bars with three different notch radii,
1.27 mm, 2.54 mm and 6.35 mm, and plane strain specimens with three different groove
radii, 2.03 mm, 5.08 mm and 16.26 mm, are tested in this study. Similar round notched
and plat grooved tensile specimens were employed by Bao and Wierzbicki [45] and Bai
and Wierzbicki [55] in their earlier work.

Figure 5. 1 Sketches of a smooth round bar, a notched round bar, a grooved plane strain
specimen and a torsion specimen.

94

29.2

17.4

7.6

15.2

50.8
101.6
(a)

31.8

203.2
2.0

6.1

(b)
32.1

2.54

13.1 14.6 22.1

300
66.7
(c)
Figure 5. 2 Dimensions of the specimens (unit: mm): (a) notched round bar, (b) plane
strain specimen and (c) torsion specimen.
95

The finite element software ABAQUS [108] is used to analyze the specimens. All
the specimens contain one layers of GLD elements located at the geometry center where
the failure most likely initiates. The element size of GLD elements is chosen to represent
the average inclusion spacing using the computational cell concept [61, 62] and the GLD
porous plasticity model described earlier is used to model the macroscopic behavior of
these GLD cell elements. Material of the rest of the elements is regarded as dense and
obeys the regular elastic-plastic constitutive law (the same as the matrix materials of the
cells). For notched round bars, axi-symmetric conditions are considered and the 4-node,
axisymmetric solid elements with reduced integration (CAX4R) are used. For the torsion
specimen, an additional degree of freedom needs to be added to the axisymmetric
element to handle the twist and the CGAX4R element in ABAQUS is developed for this
purpose. For the grooved plane strain specimens, the 3D 8- node brick elements with
reduced integration (C3D8R) are used. Usually the symmetry conditions allow for only
1/4 or 1/8 of the specimen being modeled. Figure 5.3 shows two typical finite element
meshes for a notched round bar specimen and a grooved plane strain specimen. A typical
axi-symmetric model has 700 elements and a typical 1/8-symmetric 3D model has 20,000
elements. Since reduced integration is used, its possible for the element to distort in such
a way that the energy calculated at the integration point is zero, which leads to an
uncontrolled distortion of the mesh. This is called hourglass mode [108], and the
hourglass control is employed by using *HOURGLASS STIFFNESS in the ABAQUS
input files to provide increased resistance to hourglassing.

96

GLD

GLD

(a)

(b)

Figure 5. 3 Typical finite element meshes: (a) an axi-symmetric model for a notched
round bar, (b) a 1/8-symmetric model for a grooved plane strain specimen.

5.4 Experimental and numerical results


In this section, the stress state effect on materials plastic response and ductile
failure response were studied. Numerical results were compared with experimental
results, good agreement had been achieved. The current analysis shows that Lode
parameter has strong effect on plastic response and negligible effect on ductile fracture
for Aluminum 5083.

5.4.1 Effect of the stress state on the materials plastic response


Figure 5.4 (a) shows the measured, uniaxial, engineering stress-strain curve from
the smooth round bar experimental data and Figure 5.4 (b) shows the measured shear
97

stress vs. shear strain response using the Lindholm- type torsion specimen. In the tension
test, the axial displacement was measured using an extensometer with an initial gage
length of 25 mm and the diametral contraction was monitored using a diametral gage.
The engineering stress and strain are calculated as eng applied load / initial crosssection area and eng axial displacement / initial gage length respectively. For the
torsion test, the shear stress is calculated as = /AR and shear strain is calculated as =
/L, where is the applied torque, A and R are the cross-section area and mean radius of
the specimen respectively, is the angular displacement measured with a transducer
mounted on the specimen and L is the length of the gage section. Details of the torsion
test setup can be found in Lindholm et al. [63].

Eng. Stress (MPa)

Shear Stress (MPa)

350

300
250
200

150
100
50

0
0

0.4

0.8

1.2

Shear Strain

Engineering Strain

(a)

(b)

Figure 5. 4 (a) Measured, uniaxial, engineering stress-strain curve, (b) measured shear
stress vs. shear strain curve.
Since ductile fracture initiation is preceded by large, local plastic deformation, the
true stress-strain curve is needed in the finite element analysis. Before the peak stress
point is reached, the uniaxial, engineering stress-strain curve can be converted to the true
98

stress-strain curve by using true eng (1 eng ) and true ln(1 eng ) . The plastic
portion of the true stress-strain curve is then fitted to a power- law function so that the
stress-strain curve can be extrapolated to cover the large strain region
1/ N


0
E 0

(5. 7)

where E = 70 GPa, 0 = 192 MPa and N = 0.17.


To convert the measured shear stress vs. shear strain curve to a true (equivalent)
stress-strain curve is not a simple task [116, 117]. Here an inverse, iterative method is
used to obtain the true (equivalent) stress-strain curve for torsion: (1) perform numerical
simulation of the torsion specimen using a given stress-strain curve (power- law); (2)
compare the predicted torque vs. angular displacement response with experimental
measurements; (3) adjust the stress-strain curve based on the error shown in step (2); (4)
repeat Steps (1) to (3) until the error between the numerical and experimental res ults
becomes acceptable. The calibrated 0 and N are 209 MPa and 0.125 respectively.
Figure 5.5 shows the true stress vs. true plastic strain curves (power- law) obtained
using the smooth tensile bar data and the torsion test data. The two curves clearly show
significant difference, suggesting that the stress state has a strong effect on the plastic
response of the material. As will be explained later, the effect of stress triaxiality on this
materials plastic response is insignificant and the difference shown in Figure 5.5 is
primarily attributed to the difference in the Lode angle (or parameter ) experienced by

99

the round tensile bar and the torsion specimen: for the round tensile bar, 1 , while for

Stress (MPa)

the torsion specimen, 0 .

Plastic Strain

Figure 5. 5 Comparison of the true stress vs. true plastic strain curves (power-law)
obtained using the smooth tensile bar data and the torsion test data.
Figure 5.6 compares the numerically predicted and the experimentally measured
load vs. displacement curves for the smooth tensile specimen and the torsion specimen
respectively. The tensile tests and the torsion tests each have seven specimens. The
tension and torsion stress vs. plastic strain curves shown in Figure 5.5 were used in the
finite element simulations of the respective specimens. In both cases, excellent
comparison between the numerical prediction and the experimental measurements is
observed.

100

FEA

Shear Stress (MPa)

Load (kN)

Experiments

Displacement (mm)

Experiments
FEA

Shear Strain

(a)

(b)

Figure 5. 6 Comparison of the numerical and experimental load vs. displacement curves
for (a) the smooth tensile specimen and (b) the torsion specimen (Experimental data were
from [118]).

5.4.2 Effect of the stress state on the ductile failure strain


To determine the critical failure strain as a function of the stress state, we need to
locate the cell element at which failure initiates and evaluate the stress and strain states of
this element. We first perform finite element analysis of the specimen without
considering material failure (void coalescence). Next, we compare the finite element
solution of the load-displacement curve with the experimental record and determine at
what load and deformation level failure occurs. Finally we locate the critical element
where failure starts, output the stress and strain components of this element and compute
the effective strain, stress triaxiality and Lode angle (or parameter ). The same procedure
to determine the stress state at failure strain was employed by Brunig et al. [56]. For
notched round bars, failure starts at the center. The -value of the element at the center of
all notched round bars is 1 and the T-value varies with the notch radius. The smaller the
notch radius is, the higher the stress triaxiality is. On the other hand, the -value of the
101

element at the region where failure starts in all plane strain specimens is approximately
zero while the T-value is different for specimens having different groove radius. For the
torsion specimen, the representative material volume undergoes pure shear and both and
T are zero.
As is demonstrated in the previous Section, the materials plastic response is
dependent on the stress state. Consequently, a stress-state dependent stress-strain relation
should be used in the finite element analyses of different types of the specimens. We find
that, comparing to the effect of the Lode angle, the effect of the stress triaxiality on the
plastic response is insignificant for this material. As indicated in chapter III, an I1 -J2 -J3
yield criterion, F c1[a1 I1 (27 J 23 b1 J 32 )1/ 6 ] with a1 =0 and b1 = -60.75, can describe
the Lode angle effect on the plastic response of this material reasonably well. But since
all the specimens in the present test matrix have a -value of either 1 or 0, we adopt a
simple approach to perform the finite element simulations. We use the stress-strain curve
obtained from the smooth tensile bar to analyze all the notched round bars ( =1) and use
the stress-strain curve obtained from the torsion specimen to analyze all the plane strain
specimens ( =0). It is worth mentioning that, since the porous plasticity model is used to
describe the macroscopic plastic response of the cell elements, (current yield stress of
the matrix material) rather than e (macroscopic effective stress) is used in calculating T
and for these elements, i.e., T h / and [27 J 3 / (2 3 )] .
Since all the round tensile specimens (smooth and notched) have the same - value
( =1), one should expect that the failure strain for these specimens exhibits a decaying
function of the stress triaxiality. Seven specimens were tested for each geometry and the
102

experimental data inevitably show some degree of scatter. Figure 5.7(a) shows the
variation of the failure strain (effective strain at which the critical cell eleme nt fails) with
the stress triaxiality for these =1 specimens. In this figure, the symbols represent the
average T and Ec values of the seven specimens and the error bars indicate the scatter.
The four (average) data points can be fitted into an exponential function, Ec 0.64e1.448T
, which is shown in Figure 5.7 (a) as a solid line.

0.7

B-Notch

0.6

=1
=1
Load (kN)

Ec

0.5
0.4
0.3

0.2

Experiments
FEA

0.1
0
0

0.5

1.5

2
Displacement (mm)

(a)

(b)

=1

=1

Experiments

Load (kN)

E-Notch

Load (kN)

D-Notch

Experiments
FEA

FEA

Displacement (mm)

Displacement (mm)

(c)

(d)

Figure 5. 7 (a) Variation of the critical failure strain with stress triaxiality for =1; (b)
(d) Comparison of the predicted and measured load-displacement responses for the
notched round tensile specimens having three different notch radii respectively
(Experimental data were from [118]).
103

The Ec vs. T relation obtained for this =1 case is then implemented in the
ABAQUS UMAT and the four round tensile specimens are re-analyzed with the failure
criterion (void coalescence) turned on. During the numerical analysis, at each time
increment the GLD elements are checked to see if the failure criterion is satisfied. If the
failure criterion is satisfied for a GLD element, the void coalescence will occur and this
element will rapidly lose the load carrying capacity. Figs 5.7 (b) (d) compare the
predicted load-displacement curves with experimental measurements for the three
notched round bars (the comparison for the smooth tensile bar was shown in Figure
5.6(a)). In these figures, the thicker lines represent the numerical predictions while the
thinner lines represent the experimental records. In general, the agreement between the
numerical predictions and the experimental measurements is very well. The small
discrepancies shown in some cases may be due to the negligence of the hydrostatic stress
effect on plasticity in the numerical analyses as well as errors in experimental
measurements.
To further validate the numerical results, the predicted and measured load vs.
diametral contraction records are also compared. Figure 5.8(a) shows the undeformed
mesh for the E-notch specimen and Figure 5.8(b) shows the deformed mesh just before
failure occurs. The finite element analysis takes 120 increments. Figure 5.8(c) shows
good agreement between the predicted and measured load vs. diametral contraction
records for the E-notch specimens. The comparisons for the B-notch and D- notch
specimens are similar to that of the E-notch.

104

Load (kN)

(a)

(b)

Experiments
FEA

Diametral Contraction (mm)


(c)

Figure 5. 8 (a) Undeformed mesh for the E-notch specimen, (b) deformed mesh just
before failure occurs, and (c) comparison of the predicted and measured load vs.
diametral contraction response (Experimental data were from[118]).

105

0.6

F-Groove

0.5

=0

=0
Load (kN)

Ec

0.4
0.3
0.2

Experiments
FEA

0.1
0
0

0.5

1.5

Displacement (mm)

(b)

G-Groove

H-Groove

=0

=0

Experiments

Load (kN)

Load (kN)

(a)

Experiments
FEA

FEA

Displacement (mm)

Displacement (mm)

(c)

(d)

Figure 5. 9 (a) Variation of the critical failure strain with stress triaxiality for =0; (b)
(d) Comparison of the predicted and measured load-displacement responses for three
grooved plane strain specimens respectively (Experimental data were from[118]).
Similar to the above analyses, the grooved plane strain specimens have the same
-value ( =0) as the torsion specimen and the variation of Ec with T for these specimens
is shown in Figure 5.9 (a). The four (average) data points (symbols) are fitted into an
exponential function, Ec 0.54e1.37T , which is represented by the solid line in Figure 5.9
(a). This relation is also implemented in the ABAQUS UMAT and the four specimens are
re-analyzed with the failure criterion turned on. Figs 5.6 (b) and 5.9(b) (d) show the
106

comparison of the predicted shear stress vs. shear strain or load vs. displacement response
with the experimental records. Again, good agreement between the numerical and
experimental results is observed.
If we combine the four data points shown in Figure 5.7(a) and the four data points
shown in Figure 5.9(a) and display them in Figure 5.10, we find that all data points can
be fitted into a single exponential function
Ec 0.55e 1.365T

(5. 8)

Figure 5.10 suggests that unlike the aluminum 2024 studied by Wierzbicki and Xue [52]
and the DH36 steel studied by Gao et al. [110], for which the Lode angle has strong
effect on the ductile failure strain, the effect of the Lode angle on the failure strain of the
aluminum 5083 alloy under this investigation is negligible.

Ec

0.6

0.5

=0

0.4

=1

0.3
0.2
0.1
0
0

0.5

1.5

Figure 5. 10 Variation of the ductile failure strain with the stress triaxiality for the
aluminum 5083 alloy.

107

In summary, the experimental and numerical results presented above suggest that
the plastic response of the aluminum 5083 plate considered in this study depends
sensitively on the Lode angle (or parameter ) and the ductile failure strain decays with
the stress triaxiality exponentially.

5.5 Concluding remarks


The experimental and numerical work represented in this chapter shows that the
stress state has significant effects on both plasticity and ductility of the 5083 aluminum
plate considered in this research. These effects can be described by using the stress
triaxiality and the Lode angle (or parameter ). In particular, it is found that the stress
triaxiality has relatively small effect on plasticity b ut significant effect on ductile failure
strain. On the other hand, the effect of the Lode angle on ductile fracture is negligible but
its effect on plasticity is significant.
Using the Gologanu-Leblond-Devaux porous plasticity

model and

the

computational cell approach, detailed finite element analyses of the smooth and notched
round tensile bars, grooved plane strain specimens and the Lindholm-type torsion
specimen specimens are performed and very good comparisons between the model
predictions and the experimental measurements are observed.
One of the important findings of this research is that, even for common
engineering materials, the assumptions made in the classical J 2 plasticity theory are not
always valid. A proper constitutive model is always a prerequisite for an accurate

108

numerical analysis. This study reconfirmed the need of establishing the general material
constitutive laws involving three stress invariants.

109

CHAPTER VI
CONCLUDING REMARKS AND SUGGESTIONS FOR FUTURE WORK

6.1 Conclusions and future works on plasticity modeling involving three stress invariants
In this thesis, a plasticity model involving hydrostatic pressure and Lode
parameter effects for isotropic materials is developed, which is a function of the second,
third invariants of the stress deviator and the first invariants of stress.
Finite element implementation of this I1 -J2 -J3 plasticity model including
integration of the elastic-plastic constitutive equations and the formulation of the
consistent tangent moduli is presented. Backward Euler procedure is employed on the
plasticity evolution.
Experiments including round and notched tensile round bar, grooved plane strain
specimen, compression specimen, modified Lindholm torsion specimen, torsion-tension
and torsion-compression specimen are performed. The proposed plasticity model is
calibrated and verified for a 5083 aluminum alloy, a Nitronic 40 and a Zircaloy 4. Good
agreement is achieved in the plastic responses of these three materials.
Furthermore the Gurson-Tvergaard-Needleman porous plasticity model is
extended to include the effects of the hydrostatic pressure and the Lode parameter on
plastic response of the matrix materials. A few numerical examples are analyzed to show
110

the effect of stress state on the material response, which illustrate the future numerical
analysis of using this modified GTN model.
For now we have succeeded on predicting the plastic response for 5083
Aluminum alloy, a Nitronic 40 and a Zircaloy 4. However, the plastic responses of
torsion-compression specimen are not perfectly matched for Nitronic 40 and Zircaloy 4,
which is due to the stress concentration effect at the transition area between the gage
section and the specimen reduced section. In order to provide a more effective
comparison, a new designed torsion-compression specimen, which have the ability to
reduce the stress concentration effect, is needed for the future analysis. For other future
research suggestions, a damage criterion can be incorporated into this stress state
dependent plasticity model to improve the prediction of the ductile failure of certain
materials. Also, this I1 -J2 -J3 plasticity model can be extended to include the effects of
strain-rate to study the rate-dependent behavior of some materials. Temperature effects
can also be integrated into this I1 -J2-J3 model to study the inelastic behaviors at different
temperature for temperature dependent materials.
For the modified porous plasticity model, a coalescence criterion can be
integrated into this model to improve the accuracy in predicting ductile fracture or crack
propagation process of certain materials.

6.2 Conclusions and future works on stress state effects on Ductile Fracture
In this thesis, our recent research efforts on modeling plasticity and ductile
fracture of an aluminum 5083 alloy are presented. The Gologanu- Leblond-Devaux
111

(GLD) porous plasticity model is implemented into ABAQUS via UMAT. Finite element
implementation of the GLD model follows the same procedure provided by Kim and Gao
[87].
Experiments for aluminum 5083 alloy include smooth round bar, notched round
bars, modified Lindholm torsion specimen, grooved plane strain specimen. Numerical
analysis was performed for these types of specimens, stress state dependent failure strain
coalescence criterion was calibrated for this model, and good comparisons between the
model predictions and the experimental measurements were observed.
Our investigation reveals a strong stress-state effect on the plastic response and
the ductile fracture behavior. This stress state effect is described by using the stress
triaxiality and the Lode angle (which is related to the third invariant of the stress
deviator). In particular, it is found that the stress triaxiality has relatively small effect on
plasticity but significant effect on ductile failure strain. On the other hand, the effect of
the Lode angle on ductile fracture is negligible but its effect on plasticity is significant.
With current calibrated fracture surface of 5083 Aluminum, further different
stress state experimental and numerical studies need to be planned to validate the
conclusions obtained in this study.
For current void evolution law in the GLD model, the void volume fraction is
completely tied to the plastic volume change. For high triaxiality loading, material
softening happens as the results of void growth, and for low triaxiality loading, material
softening is caused by shear localization of plastic strain of the inter-void ligaments due
to void rotation and distortion. This void evolution law works very well under high
112

triaxiality loading, but cannot describe the material softening under low triaxiality
loading. With this in consideration, the future research suggestion on fulfilling this GLD
model would be establishing a new void evolution law to include shear effect on material
softening, which can be expected to improve the prediction of material behaviors under
low triaxiality loading condition.

113

BIBLIOGRAPHY

[1]

H. Tresca, "Memoir on the flow of solid bodies under strong pressure," Comptesrendus de lacadmie des sciences, vol. 59, p. 754, 1864.

[2]

. Saint-Venant, "Sur l'tablissement des quations des mouvements intrieurs


oprs dans les corps solides ductiles au del des limites o l'lasticit pourrait les
ramener leur premier tat/B. De Saint-Venant," Comptes Rendus de l'Ac. des
Sciences, pp. 473-480, 1870.

[3]

M. Levy, "Mmoire sur les quations gnrales des mouvements intrieurs des
corps ductiles au del des limites en lasticit pourrait les ramener leur premier
tat," C. R. Acad. Sci, vol. 70, 1870.

[4]

R. von Mises, "Mechanik der festen Krpern im plastisch-deformablen Zustand


Nachrichten von der kniglichen Gesellschaft der Wissenschaften zu Gttingen,"
Mathematisch-Physikalische Klasse, 1913.

[5]

H. Hencky, "Zur Theorie plastischer Deformationen und der hierdurch im


Material hervorgerufenen Nachspannungen," ZAMMJournal of Applied
Mathematics and Mechanics/Zeitschrift fr Angewandte Mathematik und
Mechanik, vol. 4, pp. 323-334, 1924.

[6]

L. Prandtl, "Spannungsverteilung in plastischen Krpern," in In Proceedings of


the 1st Intrnational Congress on Applied Mechanics, Delft, Technische
Boekhandel en Druckerij, J. Waltman, Jr., 1924, pp. 43-54.

[7]

A. Reuss, "Bercksichtigung der elastischen Formnderung in der


Plastizittstheorie,"
ZAMMJournal
of
Applied
Mathematics
and
Mechanics/Zeitschrift fr Angewandte Mathematik und Mechanik, vol. 10, pp.
266-274, 1930.

[8]

R. Hill, The mathematical theory of plasticity. London: Oxford University Press,


1950.

[9]

A. Mendelson, Plasticity: theory and application. New York: Macmillan


Publishing Company, 1968.

[10]

D. C. Drucker and W. Prager, "Soil mechanics and plastic analysis of limit


design," Quarterly of applied mathematics, vol. 10, pp. 157-165, 1952.
114

[11]

J. Bardet, "Lode Dependences Pressure-Sensitive Materials," Journal of applied


mechanics, vol. 57, p. I499, 1990.

[12]

P. Menetrey and K. Willam, "Triaxial failure criterion for concrete and its
generalization," ACI Structural Journal, vol. 92, 1995.

[13]

D. Bigoni and A. Piccolroza, "A new yield function for geomaterials,"


Constitutive modelling and analysis of boundary value problems in geotechnical
engineering. Napoli:[sn], pp. 266-281, 2003.

[14]

Z. Yang and A. Elgamal, "Multi-surface cyclic plasticity sand model with Lode
angle effect," Geotechnical and Geological Engineering, vol. 26, pp. 335-348,
2008.

[15]

P. V. Lade and J. M. Duncan, "Elastoplastic stress-strain theory for cohesionless


soil," Journal of the Geotechnical Engineering Division, vol. 101, pp. 1037-1053,
1975.

[16]

T. Guo, et al., "Continuum modeling of a porous solid with pressure-sensitive


dilatant matrix," Journal of the Mechanics and Physics of Solids, vol. 56, pp.
2188-2212, 2008.

[17]

C. Pampillo and L. Davis, "Volume change during deformation and pressure


dependence of yield stress," Journal of Applied Physics, vol. 42, pp. 4674-4679,
1971.

[18]

W. Spitzig and O. Richmond, "Effect of hydrostatic pressure on the deformation


behavior of polyethylene and polycarbonate in tension and in compression,"
Polymer Engineering & Science, vol. 19, pp. 1129-1139, 1979.

[19]

W. Spitzig and O. Richmond, "The effect of pressure on the flow stress of


metals," Acta Metallurgica, vol. 32, pp. 457-463, 1984.

[20]

I. Chen and P. E. R. Morel, "Implications of Transformation Plasticity in


ZrO2Containing Ceramics: I, Shear and Dilatation Effects," Journal of the
American Ceramic Society, vol. 69, pp. 181-189, 1986.

[21]

J. Lu and G. Ravichandran, "Pressure-dependent flow behavior of Zr41. 2Ti13.


8Cu12. 5Ni10Be22. 5 bulk metallic glass," Journal of materials research, vol. 18,
pp. 2039-2049, 2003.

[22]

C. D. Wilson, "A critical reexamination of classical metal plasticity," Journal of


applied mechanics, vol. 69, p. 63, 2002.

[23]

X. Lei and C. J. Lissenden, "Pressure sensitive nonassociative plasticity model for


DRA composites," Journal of engineering materials and technology, vol. 129, p.
255, 2007.
115

[24]

W. Wang, et al., "Tensile tests and analyses of notched specimens fabricated from
high strength steels using a generalized yield model," Fatigue & Fracture of
Engineering Materials & Structures, vol. 33, pp. 310-319, 2010.

[25]

D. C. Drucker, "relations of experiments to mathematical theories of plasticity,"


vol. 16, pp. 349-357, 1949.

[26]

M. Brnig, "Numerical simulation of the large elasticplastic deformation


behavior of hydrostatic stress-sensitive solids," International Journal of Plasticity,
vol. 15, pp. 1237-1264, 1999.

[27]

M. Brnig, et al., "Numerical simulation of the localization behavior of


hydrostatic-stress-sensitive metals," International Journal of Mechanical Sciences,
vol. 42, pp. 2147-2166, 2000.

[28]

M. Kuroda, "A phenomenological plasticity model accounting for hydrostatic


stress-sensitivity and vertex-type of effect," Mechanics of materials, vol. 36, pp.
285-297, 2004.

[29]

W. Hu and Z. Wang, "Multiple- factor dependence of the yielding behavior to


isotropic ductile materials," Computational materials science, vol. 32, pp. 31-46,
2005.

[30]

O. Cazacu and F. Barlat, "Application of the theory of representation to describe


yielding of anisotropic aluminum alloys," International journal of engineering
science, vol. 41, pp. 1367-1385, 2003.

[31]

S. Soare, et al., "Applications of a Recently Proposed Anisotropic Yield Function


to Sheet Forming," Advanced Methods in Material Forming, Ed. Banabic D.,
Springer, Heidelberg-Berlin, pp. 131-149, 2007.

[32]

Y. Bai and T. Wierzbicki, "Application of extended MohrCoulomb criterion to


ductile fracture," International Journal of Fracture, vol. 161, pp. 1-20, 2010.

[33]

J. Chaboche, "A review of some plasticity and viscoplasticity constitutive


theories," International Journal of Plasticity, vol. 24, pp. 1642-1693, 2008.

[34]

P. Ludwik, Scheu, R, "Ueber Kerbwirkungen bei Flusseisen," Stahl U. Eisen, vol.


43, 1923.

[35]

E. Orowan, Notch, Brittleness and the Strength of Metals: Institution of Engineers


and Shipbuilders in Scotland, 1945.

[36]

P. W. Bridgman, Studies in large plastic flow and fracture: Mc. Graw-Hill, 1952.

[37]

I. French and P. Weinrich, "The effect of hydrostatic pressure on the tensile


fracture of [alpha]-brass," Acta Metallurgica, vol. 21, pp. 1533-1537, 1973.
116

[38]

I. French and P. Weinrich, "The effects of hydrostatic pressure on the mechanism


of tensile fracture of aluminum," Metallurgical and Materials Transactions A, vol.
6, pp. 1165-1169, 1975.

[39]

I. E. French and P. F. Weinrich, "The influence of hydrostatic pressure on the


tensile deformation and fracture of copper," Metallurgical and Materials
Transactions A, vol. 6, pp. 785-790, 1975.

[40]

P. Weinrich and I. French, "The influence of hydrostatic pressure on the fracture


mechanisms of sheet tensile specimens of copper and brass," Acta Metallurgica,
vol. 24, pp. 317-322, 1976.

[41]

J. Hancock and A. Mackenzie, "On the mechanisms of ductile failure in highstrength steels subjected to multi-axial stress-states," Journal of the Mechanics
and Physics of Solids, vol. 24, pp. 147-160, 1976.

[42]

J. Hancock and D. Brown, "On the role of strain and stress state in ductile
failure," Journal of the Mechanics and Physics of Solids, vol. 31, pp. 1-24, 1983.

[43]

G. R. Johnson and W. H. Cook, "Fracture characteristics of three metals subjected


to various strains, strain rates, temperatures and pressures," Engineering fracture
mechanics, vol. 21, pp. 31-48, 1985.

[44]

M. Mirza, et al., "The effect of stress triaxiality and strain-rate on the fracture
characteristics of ductile metals," Journal of materials science, vol. 31, pp. 453461, 1996.

[45]

Y. Bao and T. Wierzbicki, "On fracture locus in the equivalent strain and stress
triaxiality space," International Journal of Mechanical Sciences, vol. 46, pp. 8198, 2004.

[46]

Y. Bao, "Dependence of ductile crack formation in tensile tests on stress


triaxiality, stress and strain ratios," Engineering fracture mechanics, vol. 72, pp.
505-522, 2005.

[47]

J. Kim, et al., "Modeling of crack growth in ductile solids: a three-dimensional


analysis," International journal of solids and structures, vol. 40, pp. 7357-7374,
2003.

[48]

J. Kim, et al., "Modeling of void growth in ductile solids: effects of stress


triaxiality and initial porosity," Engineering fracture mechanics, vol. 71, pp. 379400, 2004.

[49]

J. Kim, et al., "Modeling of ductile fracture: Application of the mechanism-based


concepts," International journal of solids and structures, vol. 44, pp. 1844-1862,
2007.
117

[50]

X. Gao, et al., "On ductile fracture initiation toughness: effects of void volume
fraction, void shape and void distribution," International journal of solids and
structures, vol. 42, pp. 5097-5117, 2005.

[51]

X. Gao and J. Kim, "Modeling of ductile fracture: Significance of void


coalescence," International journal of solids and structures, vol. 43, pp. 62776293, 2006.

[52]

T. Wierzbicki and L. Xue, "On the effect of the third invariant of the stress
deviator on ductile fracture," Technical Report, Impact and Crashworthiness Lab,
MIT, 2005.

[53]

I. Barsoum and J. Faleskog, "Rupture mechanisms in combined tension and shear-Experiments," International journal of solids and structures, vol. 44, pp. 17681786, 2007.

[54]

I. Barsoum and J. Faleskog, "Rupture mechanisms in combined tension and shear-Micromechanics," International journal of solids and structures, vol. 44, pp.
5481-5498, 2007.

[55]

Y. Bai and T. Wierzbicki, "A new model of metal plasticity and fracture with
pressure and Lode dependence," International Journal of Plasticity, vol. 24, pp.
1071-1096, 2008.

[56]

M. Brnig, et al., "A ductile damage criterion at various stress triaxialities,"


International Journal of Plasticity, vol. 24, pp. 1731-1755, 2008.

[57]

M. Gologanu, et al., "Approximate models for ductile metals containing nonspherical voids--Case of axisymmetric prolate ellipsoidal cavities," Journal of the
Mechanics and Physics of Solids, vol. 41, pp. 1723-1754, 1993.

[58]

M. Gologanu, et al., "Approximate models for ductile metals containing


nonspherical voidscase of axisymmetric oblate ellipsoidal cavities," Journal of
engineering materials and technology, vol. 116, p. 290, 1994.

[59]

M. Gologanu, et al., "Recent extensions of Gurson's model for porous ductile


metals," 1995, pp. 134-203.

[60]

T. Pardoen and J. Hutchinson, "An extended model for void growth and
coalescence," Journal of the Mechanics and Physics of Solids, vol. 48, pp. 24672512, 2000.

[61]

L. Xia, et al., "A computational approach to ductile crack growth under large
scale yielding conditions," Journal of the Mechanics and Physics of Solids, vol.
43, pp. 389-413, 1995.

118

[62]

X. Gao, et al., "Ductile tearing in part-through cracks: experiments and cell- model
predictions," Engineering fracture mechanics, vol. 59, pp. 761-777, 1998.

[63]

U. Lindholm, et al., "Large strain, high strain rate testing of copper," Journal of
engineering materials and technology, vol. 102, p. 376, 1980.

[64]

W. Lode, "Versuche ber den Einflu der mittleren Hauptspannung auf das
Flieen der Metalle Eisen, Kupfer und Nickel," Zeitschrift fr Physik A Hadrons
and Nuclei, vol. 36, pp. 913-939, 1926.

[65]

L. Xue, "Damage accumulation and fracture initiation in uncracked ductile solids


subject to triaxial loading," International journal o f solids and structures, vol. 44,
pp. 5163-5181, 2007.

[66]

C. A. Coulomb, Essai sur une application des rgles de maximis & minimis
quelques problmes de statique, relatifs l'architecture: Mem Acad Roy des Sci,
1776.

[67]

O. Mohr, "Abhandlungen aus dem Gebiete der Technischen Mechanik (2nd ed).
Ernst, Berlin," 1914.

[68]

W. J. M. Rankine, "On the stability of loose earth," Philosophical Transactions of


the Royal Society of London, pp. 9-27, 1857.

[69]

W. A. Spitzig, et al., "Pressure dependence of yielding and associated volume


expansion in tempered martensite," Acta Metallurgica, vol. 23, pp. 885-893, 1975.

[70]

W. A. Spitzig, et al., "The effect of hydrostatic pressure on the deformation


behavior of maraging and HY-80 steels and its implications for plasticity theory,"
Metallurgical and Materials Transactions A, vol. 7, pp. 1703-1710, 1976.

[71]

H. Subramanya, et al., "Influence of crack tip constraint on void growth in


pressure sensitive plastic solids-I: 2D analysis," Engineering fracture mechanics,
vol. 75, pp. 1045-1063, 2008.

[72]

F. Yang, et al., "Yield criterions of metal plasticity in different stress states," Acta
Metallurgica Sinica (English Letters), vol. 20, pp. 123-130, 2009.

[73]

A. Benzerga and J. B. Leblond, "Ductile fracture by void growth to coalescence,"


Advances in Applied Mechanics, vol. 44, pp. 169-305, 2010.

[74]

F. A. McClintock, "A criterion for ductile fracture by the growth of holes,"


Journal of applied mechanics, vol. 35, p. 363, 1968.

[75]

J. Rice and D. M. Tracey, "On the ductile enlargement of voids in triaxial stress
fields*," Journal of the Mechanics and Physics of Solids, vol. 17, pp. 201-217,
1969.
119

[76]

A. Gurson, "Continuum theory of ductile rupture by void nucleation and growth:


Part I-Yield criteria and flow rules for porous ductile media," Journal of
engineering materials and technology, vol. 99, pp. 2-15, 1977.

[77]

V. Tvergaard, "On localization in ductile materials containing spherical voids,"


International Journal of Fracture, vol. 18, pp. 237-252, 1982.

[78]

V. Tvergaard and A. Needleman, "Analysis of the cup-cone fracture in a round


tensile bar," Acta Metallurgica, vol. 32, pp. 157-169, 1984.

[79]

L. Brown and J. Embury, "Initiation and growth of voids at second-phase


particles," 1973.

[80]

H. Andersson, "Analysis of a model for void growth and coalescence ahead of a


moving crack tip," Journal of the Mechanics and Physics of Solids, vol. 25, pp.
217-233, 1977.

[81]

A. Needleman and J. Rice, "Limits to ductility set by plastic flow localization,"


Mechanics of Sheet Metal Forming, pp. 237-265, 1978.

[82]

A. L. Gurson, "Plastic flow and fracture behavior of ductile materials


incorporating void nucleation, growth, and interaction[Ph. D. Thesis]," 1975.

[83]

A. Gurson, "Porous rigid-plastic materials containing rigid inclusions: yield


function, plastic potential, and void nucleation," Brown Univ., Providence, RI
(USA). Div. of Engineering1976.

[84]

A. Argon, et al., "Distribution of plastic strain and negative pressure in necked


steel and copper bars," Metallurgical and Materials Transactions A, vol. 6, pp.
815-824, 1975.

[85]

A. Argon, et al., "Cavity formation from inclusions in ductile fracture,"


Metallurgical and Materials Transactions A, vol. 6, pp. 825-837, 1975.

[86]

C. Chu and A. Needleman, "Void nucleation effects in biaxially stretched sheets,"


Journal of engineering materials and technology, vol. 102, p. 249, 1980.

[87]

J. Kim and X. Gao, "A generalized approach to formulate the consistent tangent
stiffness in plasticity with application to the GLD porous material model,"
International journal of solids and structures, vol. 42, pp. 103-122, 2005.

[88]

D. C. Drucker and M. Li, "Non-associated plastic deformation and genuine


instability," Acta Mech. suppl, vol. 3, pp. 131-171, 1992.

[89]

Lade P V, et al., "Instability of granular materials with nonassociated flow,"


Journal of engineering mechanics, vol. 114, p. 2173, 1988.
120

[90]

J. W. Rudnicki and J. Rice, "Conditions for the localization of deformation in


pressure-sensitive dilatant materials," Journal of the Mechanics and Physics of
Solids, vol. 23, pp. 371-394, 1975.

[91]

O. G. Lademo, et al., "An evaluation of yield criteria and flow rules for
aluminium alloys," International Journal of Plasticity, vol. 15, pp. 191-208, 1999.

[92]

Z. Mroz, "Non-associated flow laws in plasticity,". J. Mc. Tho. Appl, vol. 2, pp.
21-42, 1963.

[93]

S. Nemat-Nasser and A. Shokooh, "On finite plastic flows of compressible


materials with internal friction," International journal of solids and structures, vol.
16, pp. 495-514, 1980.

[94]

J. Dorris and S. Nemat-Nasser, "A plasticity model for flow of granular materials
under triaxial stress states," International journal of solids and structures, vol. 18,
pp. 497-531, 1982.

[95]

S. Nemat-Nasser, "On finite plastic flow of crystalline solids and geomaterials,"


Journal of applied mechanics, vol. 50, p. 1114, 1983.

[96]

S. Nemat-Nasser, "Phenomenological theories of elasto-plasticity and strain


localization at high strain rates," Appl. Mech. Rev, vol. 45, pp. s19-s45, 1992.

[97]

K. Runesson and Z. Mroz, "A note on nonassociated plastic flow rules,"


International Journal of Plasticity, vol. 5, pp. 639-658, 1989.

[98]

T. B. Stoughton, "A non-associated flow rule for sheet metal forming,"


International Journal of Plasticity, vol. 18, pp. 687-714, 2002.

[99]

T. B. Stoughton and J. W. Yoon, "A pressure-sensitive yield criterion under a


non-associated flow rule for sheet metal forming," International Journal of
Plasticity, vol. 20, pp. 705-731, 2004.

[100] T. B. Stoughton and J. W. Yoon, "Review of Druckers postulate and the issue of
plastic stability in metal forming," International Journal of Plasticity, vol. 22, pp.
391-433, 2006.
[101] F. Ma and K. Kishimoto, "On yielding and deformation of porous plastic
materials," Mechanics of materials, vol. 30, pp. 55-68, 1998.
[102] V. Tvergaard, "Influence of voids on shear band instabilities under plane strain
conditions," International Journal of Fracture, vol. 17, pp. 389-407, 1981.
[103] V. Cvitanic, et al., "A finite element formulation based on non-associated
plasticity for sheet metal forming," International Journal of Plasticity, vol. 24, pp.
646-687, 2008.
121

[104] A. Taherizadeh, et al., "A non-associated constitutive model with mixed isokinematic hardening for finite element simulation of sheet meta l forming,"
International Journal of Plasticity, vol. 26, pp. 288-309, 2010.
[105] N. Aravas, "On the numerical integration of a class of pressuredependent
plasticity models," International Journal for Numerical Methods in Engineering,
vol. 24, pp. 1395-1416, 1987.
[106] J. Lubliner, "1990, Plasticity Theory. Macmillan Publishing Company, New
York."
[107] L. E. Malvern, "Introduction to the Mechanics of a Continuous Medium, PrenticeHall, Inc., Englewood Cliffs, New Jersey," 1969.
[108] "ABAQUS/Stanard User's Manual (version 6.9), 2008," ed: SIMULIA,
Providence, RI.
[109] J. C. Simo and R. L. Taylor, "Consistent tangent operators for rate-independent
elastoplasticity," Computer methods in applied mechanics and engineering, vol.
48, pp. 101-118, 1985.
[110] X. Gao, et al., "A Study on the effect of the stress state on ductile fracture,"
International Journal of Damage Mechanics, vol. 19, pp. 75-94, 2010.
[111] X. Gao, et al., "On Stress-State Dependent Plasticity Modeling: Significance of
the Hydrostatic stress, the Third Invariant of Stress Deviator and the NonAssociated Flow Rule," International Journal of Plasticity, vol. 27, pp. 217-231,
2011.
[112] T. Zhang, et al., "Application of the Plasticity Models that Involve Three Stress
Invariants," International Journal of apllied mechanics, vol. Vol. 4 No. 2, 2012.
[113] J. Faleskog, et al., "Cell model for nonlinear fracture analysisI. Micromechanics
calibration," International Journal of Fracture, vol. 89, pp. 355-373, 1998.
[114] K. Nahshon and J. Hutchinson, "Modification of the Gurson model for shear
failure," European Journal of Mechanics-A/Solids, vol. 27, pp. 1-17, 2008.
[115] K. Nahshon and Z. Xue, "A modified Gurson model and its application to punchout experiments," Engineering fracture mechanics, vol. 76, pp. 997-1009, 2009.
[116] H. C. Wu, Continuum mechanics and plasticity vol. 3: CRC Press, 2005.
[117] J. Kang, et al., "Constitutive Behavior of AA5754 Sheet Materials at Large
Strains," Journal of engineering materials and technology, vol. 130, p. 031004,
2008.
122

[118] X. Gao, et al., "Effects of the stress state on plasticity and ductile failure of an
aluminum 5083 alloy," International Journal of Plasticity, vol. 25, pp. 2366-2382,
2009.

123

APPENDICES

124

APPENDIX A. EXAMPLES OF FIRST ORDER HOMOGENEOUS FUNCTIONS OF


STRESSES

As stated in the text, there is no unique set of F and G functions that can be used to
describe the plastic behavior.

Listed below are six examples of the first order

homogeneous functions of the stress that can be used as the yield function and plastic
flow potential. The ranges of the model parameters are also given to ensure that the
convexity requirement is satisfied. The famous Drucker-Prager model (Drucker and
Prager, 1952) and Drucker model (Drucker, 1949) are just special cases of these
functions.
(1) -60.75 b1 , b2 91.125
2

F c1 (a1 I16 27 J 23 b1 J 3 )1 / 6
2

G c2 (a2 I16 27 J 23 b2 J 3 )1 / 6

c1 1 / a1 4b1 / 729 1

1/ 6

c2 1 / a2 4b2 / 729 1

1/ 6

(2) -60.75 b1 , b2 91.125


F c1[a1 I1 (27 J 23 b1 J 32 )1/ 6 ]

c1 1 /[ a1 (4b1 / 729 1)1 / 6 ]

G c2 [a2 I1 (27 J 23 b2 J 32 )1/ 6 ]

c2 1 /[ a2 (4b2 / 729 1)1 / 6 ]

(3) -2.25 b1 , b2 2.531


F c1[a1 I12 27 J 2 b1 ( J 3 / J 2 ) 2 ]1/ 2

c1 1 / a1 4b1 / 81 1

G c2 [a2 I12 27 J 2 b2 ( J 3 / J 2 ) 2 ]1/ 2

c2 1 / a2 4b2 / 81 1

1/ 2

1/ 2

125

(4) -2.25 b1 , b2 2.531


F c1 [a1 I1 (3J 2 b1 ( J 3 / J 2 ) 2 )1/ 2 ]

c1 1 /[a1 (4b1 / 81 1)1/ 2 ]

G c2 [a2 I1 (3J 2 b2 ( J 3 / J 2 ) 2 )1/ 2 ]

c2 1 /[a2 (4b2 / 81 1)1/ 2 ]

(5) -6.75 b1 , b2 6.75

F c1 (a1 I13 3 3J 23 / 2 b1 J 3 )1/ 3

c1 1 /( a1 2b1 / 27 1)1/ 3

G c2 (a2 I13 3 3J 23 / 2 b2 J 3 )1/ 3

c2 1 /( a2 2b2 / 27 1)1/ 3

(6) -6.75 b1 , b2 6.75

F c1[a1 I1 (3 3J 23 / 2 b1 J 3 )1/ 3 ]

c1 1 / a1 (2b1 / 27 1)1/ 3

G c2 [a2 I1 (3 3J 23 / 2 b2 J 3 )1/ 3 ]

c2 1 / a2 (2b2 / 27 1)1/ 3

126

APPENDIX B: GLD

Rx2

Rx2

Ry2

Ry2

Ry1

Ry1

Rx1
Rx1

(a)

(b)

Figure B1. Geometrical representation of a representative material volume: (a) prolate


void shape, (b) oblate void shape.
The GLD model [57-60] is derived from a material volume of spheroidal shape
containing a confocal spheroidal void (Figure A1). The void can be either prolate
(referred to by the symbol P) or oblate (referred to by the symbol O) and it is
axisymmetric about the y-axis with an aspect ratio of W = Ry1 /Rx1 . Because of
confocality, there exists a relationship

Rx21 R y21

Rx22 R y22 c , where c is the focal

distance.

the

and

The

eccentricities

of

inner

127

outer

spheroidal

shapes

are e1 c R y1 (P) or c Rx1 (O) , e2 c R y 2 (P) or c Rx 2 (O). These eccentricities are


related to the void volume fraction f and the shape parameter S ( S ln W ) by

1 e12 exp( 2 S ),

1 e22 1 e12
3
f
e23
e1

2
2
f 1 e2 1 e1

e23
e13

(P),
(B.1)

(O).

The g, , 2 , and C are expressed by

0
3

g e2
1 e2
2

(P),
(B.2)

(O),

e1 1 3 e12 2 3 e14 3
1

ln
3 2 ln ln
2
4
3
e2
3
3 ln( f )
3 e2 2 3 e2

-1
2
2 5 2
5 2 4
5 2
5 2
2 3 g f g1 5 g f g1 3 g f g1
g
,
g
,

f
g
g
ln(
/
)
3
g f
f
1

4
3 e1
4
3 e2

g1

g
g 1

(P),

(O),
(B.3)

1 e22

3 e24
2
2
2
1 e2 1 2e2
3 6e22 4e24

(P),

(B.4)

(O),

g 1g f sh
,
Q H

(B.5)

where

sh sinhH , ch coshH , H 21 2 , Q 1 f ,

128

(B.6)

1 1 e12
tanh -1 e1
(P),
2
3
e
e
2
2
1
1
1
2
2
1 e1 1 e1 sin -1 e (O).
1
2e12
2e13

(B.7)

The 1 and 2 are given by

9
2

1 1 T 1 f
2

1 3
1

'
1

(B.8)

1 3 1
3 1 1,
f

(B.9)

where
1
3 e 2
1
1'
2
1

e
1

2
3 e1

(P),

(B.10)
(O),

T 1 0.5T ,
T

xx

yy

3 yy xx

(B.11)

(B.12)

129

You might also like