You are on page 1of 172

MODELING EVOLUTION OF ANISOTROPY AND HARDENING

FOR SHEET METALS

DISSERTATION

Presented in Partial Fulfillment of the Requirements for

the Degree Doctor of Philosophy in the Graduate

School of The Ohio State University

By

Yangwook Choi, M.S.

*****

The Ohio State University


2003

Dissertation Committee: Approved by

Professor June K. Lee, Adviser


Adviser

Professor Robert H. Wagoner, Co-adviser


Co-adviser

Professor Mark E. Walter, Co-adviser


Co-adviser

Professor Brian D. Harper


Graduate program in Engineering Mechanics
ABSTRACT

Modeling the material behavior accurately in large strain and large rotation is

essential to sheet metal forming simulation and springback prediction, especially when

the material points experience multi-axial/-path and/or cyclic loadings.

Important features of the material behavior in complex loading are reorientation

of the texture and the modification of dislocation structures. It is viewed that the

reorientation of texture and the modification of dislocation structures influence the

evolution of anisotropy and the transient behavior of hardening, respectively.

The plastic spin theory is introduced to define the rotation of texture and the

rotational aspect of hardening. The combined Anisotropic Nonlinear Kinematic (ANK)

hardening model proposed by Chun et al. (2002) is modified to model the transient

behavior more realistically by including the rotation of yield loci.

A new hardening model, called the Rotational-Isotropic-Kinematic (RIK) model,

is proposed to account for the rotation as well as the expansion (isotropic) and translation

(kinematic) of the yield loci in terms of the plastic spin theory and the anisotropic

hardening model. A precipitate hardened material is investigated for another kind of

evolution of anisotropy. Existing models show improper hardening in 45° orientation

when the rotation of anisotropy axes is not considered. The effects of reorientation of the

ii
backstress in alloys containing hard precipitates appear to be the dominant factor for the

stress hardening response.

The capability of the RIK model is demonstrated by showing two characteristic

effects of multi-path loadings: crossing and Bauschinger effects. The proposed model

reproduces the hardening in an arbitrary orientation as well for a wide range of strain.

Only two material parameters are added to describe the evolution of anisotropy

and related backstress, which can be identified from tensile tests in three orientations and

curve fittings. Three initial R-values can be obtained from the tensile tests. The other six

parameters for defining isotropic and combined kinematic hardening are identified by the

three-point-bend test and the inverse method developed by Zhao and Lee (2000),

modified by Chun et al. for the ANK model. Relatively straightforward material

parameter identification is demonstrated to be effective for the proposed RIK model

under cyclic and multi-path loadings.

The RIK model is applied to sheet metal forming and springback predictions.

Cylindrical cup and square cup drawings are simulated and compared with available

experimental data. The RIK model shows good trends in earing and draw-ins, compared

with other models. The draw bead test simulation shows a significant difference in the

springback shape when compared with other hardening models. The RIK model corrects

the unexpected twisting mode which is computed for the specimen cut in an arbitrary

orientation away from the RD when using conventional isotropic and combined

kinematic hardening models.

As a conclusion of this investigation, the proposed RIK model significantly

improves the representation of anisotropy behaviors of rolled sheet metals in the

phenomenological framework for the continuum approach of forming simulation and

prediction of springback.
iii
Dedicated to my parents and family

iv
ACKNOWLEDGMENTS

I would like to express my heartfelt thanks to my adviser, Professor J.K. Lee, for

his continuing encouragement and intellectual discussions. I appreciated the help of my

co-advisers: Professor R.H. Wagoner shared his expertise in the sheet metal forming

research. Professor M.E. Walter led the experimental observations and taught me how to

treat the data statistically.

Professor B.D. Harper is acknowledged for serving on my dissertation committee

and making suggestions to improve this dissertation.

I appreciate Dr. C.-S. Han for sharing his ideas and mentoring this research and

algorithmic expression of the model.

I would also like to thank all of the members of ComPro Lab: Dr. B.K. Chun, J.

Choi, H.S. Kim and Hyungjun Kim for their interactive and fruitful discussions and help

with the experiments.

Partial financial supports by the following organizations are gratefully

acknowledged:

- OSU – CAMMAC/ TREP

- NIST through ERIM on The Springback Predictability Project.

Finally but not at the least, I would like to express my sincere gratitude to my

parents for cheering me, and my wife and children for being with me throughout the long

period of this endeavor.

v
VITA

March 31, 1963 ..............................................Born – Seoul, Korea

February, 1985 ...............................................B.S., Mechanical Engineering,


Korea University, Korea

August, 1991 ..................................................M.S., Engineering Mechanics,


The Ohio State University

1985 – 1998....................................................Research engineer,


Hyundai Motor Company, Korea

1998 – present ................................................Graduate Research Associate,


The Ohio State University

PUBLICATIONS

Research Publication:

1. Choi, Y., Han, C.-S., Lee, J.K., and Wagoner, R.H., Effect of Anisotropy Axes
Rotation on Spring Back, in Numisheet 2002, Jeju Island, Korea, p. 325-330.

2. Han, C.-S., Choi, Y., Lee, J.K., and Wagoner, R.H., Modeling of Anisotropic Work
Hardening, in Numisheet 2002, Jeju Island, Korea, p. 103-108.

vi
3. Han, C.-S., Choi, Y., Lee, J.K., and Wagoner, R.H., A Fe Formulation for Elasto-
Plastic Materials with Planar Anisotropic Yield Functions and Plastic Spin,
International Journal of Solids and Structures, 39, 2002, p. 5153-5141

4. Choi. Y., Han, C.-S., Lee, J.K., Wagoner, R.H., Modeling Multi-Axial Elasto-Plastic
Deformation of Planar Anisotropic Material, Part I: Theory, to be submitted to
International Journal of Plasticity.

5. Choi. Y., Han, C.-S., Lee, J.K., Wagoner, R.H., Modeling Multi-Axial Elasto-Plastic
Deformation of Planar Anisotropic Material, Part II: Applications, to be submitted to
International Journal of Plasticity

6. Choi, Y., Walter, M.E., Lee, J.K., and Han, C.-S., Observations of Anisotropy
Evolution and Identification of the Plastic Spin Parameters, to be submitted to
International Journal of Solids and Structures

FIELDS OF STUDY

Major Field: Engineering Mechanics

Studied in Plasticity, Numerical Computation, Sheet metal forming

vii
TABLE OF CONTENTS

Page

Abstract ............................................................................................................................... ii

Acknowledgments............................................................................................................... v

Vita .................................................................................................................................... vi

List of tables....................................................................................................................... xi

List of figures.................................................................................................................... xii

Chapters:

1. Introduction..................................................................................................................... 1

2. Summary of finite strain plasticity in intermediate configuration and rotational effect


of planar anisotropy .......................................................................................................... 10

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


2.2 Material formulation ......................................................................................... 11
2.3 Algorithmic treatment....................................................................................... 14
2.4 Numerical simulation........................................................................................ 18

3. Modeling multi-axial elasto-plastic deformation of planar anisotropic material, Part I:


Theory ............................................................................................................................... 29

3.1 Introduction....................................................................................................... 30
3.2 Material formulation ......................................................................................... 33
3.2.1 General model for isotropic, kinematic and rotational hardening ............ 33
3.2.2 Rotation of the symmetry axes of anisotropy (Rotational hardening)...... 36
3.2.3 Armstrong-Frederick type backstress components................................... 37
3.2.4 Nonlinear kinematic hardening with permanent softening....................... 38
viii
3.2.5 Correction of flow stress with kinematic hardening associated with the
anisotropy axes rotation ............................................................................................ 39
3.3 Application on plane stress problem for sheet metal forming. ......................... 41
3.4 Numerical investigation .................................................................................... 43
3.4.1 Material properties:................................................................................... 43
3.5 Conclusions....................................................................................................... 46

4. Experimental observation of anisotropy evolution in sheet metal and identifying the


material properties ............................................................................................................ 59

4.1 Introduction....................................................................................................... 60
4.2 Definition of R-value ........................................................................................ 63
4.3 Experimental set up........................................................................................... 64
4.4 Hardening models ............................................................................................. 67
4.5 Experimental results and material parameters .................................................. 68
4.5.1 DDQ (Experiment with DIC method)....................................................... 69
4.5.2 DDQ (NUMISHEET 2002) ...................................................................... 70
4.5.3 DQ............................................................................................................. 70
4.5.4 HSS ........................................................................................................... 71
4.6 Discussion ......................................................................................................... 71
4.7 Concluding remarks .......................................................................................... 72

5. Modeling multi-axial elasto-plastic deformations on planar anisotropic materials, part


II: Applications ................................................................................................................. 93

5.1 Introduction....................................................................................................... 94
5.2 RIK hardening model........................................................................................ 95
5.3 Material properties ............................................................................................ 97
5.4 Applications ...................................................................................................... 98
5.4.1 Cylindrical cup drawing............................................................................ 98
5.4.2 Square cup drawing................................................................................. 100
5.4.3 Pulling the strip through a draw bead toolset and springback ................ 101
5.4.4 Hydroforming ......................................................................................... 102
5.5 Concluding remakrs ........................................................................................ 104

6. Work hardening in precipitate hardened materials ..................................................... 123

6.1 Introduction..................................................................................................... 124


6.2 Material modeling........................................................................................... 125
6.3 Algorithmic treatment for the rotation............................................................ 128
6.4 Numerical investigations for the kinematic hardening models....................... 128
6.5 Discussion of the model.................................................................................. 129
6.6 Numerical application..................................................................................... 129
6.7 Conclusion ...................................................................................................... 130

ix
7. Conclusion .................................................................................................................. 139

BIBLIOGRAPHY........................................................................................................... 141

Appendices:

A. Explicit formulation ................................................................................................... 149

B. Algorithmic treatment for plane stress problem (IMPLICIT).................................... 151

x
LIST OF TABLES

Table Page

Table 2.1: Comparison of continuum setting in the intermediate configuration and the
( ε e ) ) is used. .... 20
current configuration. Small elastic strain assumption ( Ve = I +O

Table 2.2: Material parameters of Kim and Yin [4]. ....................................................... 21

Table 2.3: Algorithmic step (implicit) ............................................................................. 22

Table 3.1: Capability of reproducing multi-axial/-path loadings characteristics of


conventional hardening models. ............................................................................... 48

Table 3.2: Constitutive equations of RIK hardening model ............................................. 48

Table 3.3: Material properties for DDQ. .......................................................................... 49

Table 4.1: Material properties of RIK hardening model for DDQ, DQ and HSS. .......... 74

Table 5.1: Material properties for RIK hardening model. ............................................. 105

Table 5.2: Clamping force and drawing force for different orientation and different
hardening for draw bead test................................................................................... 105

Table 6.1: YLD96 parameters......................................................................................... 131

Table 6.2: Material properties for Numerical analysis. ................................................. 131

xi
LIST OF FIGURES

Figure Page

Figure 1.1: Macro-mechanical test: a) A big specimen is cut from a rolled sheet with an
orientation Ψ to the RD and is pre-stretched by ε*, b) smaller specimens are cut
from the pre-stretched specimen in every 15° from the RD and are stretched to
measure yield stress .................................................................................................... 5

Figure 1.2: Definition of angles used in this research. The symmetry axes ( eiφ ), initial
orientation to the RD ( ψ ), angle of the symmetry axes rotation ( η ), and the angle
between the symmetry axes and the straining directioin ( ϑ ). Symmetry axes of
anisotropy rotates toward pre-stretched direction (ε*) from the RD and TD
whichever closer. ........................................................................................................ 6

Figure 1.3: Rotation of the symmetry axes as the pre-strain increases: measured yield
stress (0.2% offset, ●) and curve fitting to Hill’s 48 yield function [6]. ................... 7

Figure 1.4: Rotation of the symmetry axes observed by using pole figures for 45° initial
orientation. The symmetry axis is rotating gradually from TD counterclockwise to
45° as strain increases [5]. .......................................................................................... 7

Figure 1.5: Treatment of anisotropy evolution in polycrystalline plasticity. Texture


orientation and substructures are considered to be the key features of anisotropy
which are related through slip processes. [29]........................................................... 8

Figure 1.6: Treatment of anisotropy evolution in phenomenological plasticity. Motivated


from the polycrystalline plasticity, yield function rotation and hardening evolution
are formulated to describe the anisotropy evolution in the RIK hardening model..... 8

Figure 1.7: Schematic diagrams of hardening models: a) Isotropic hardening – expand


the yield surface uniformly, b) Kinematic hardening – move the yield surface to the
xii
loading direction, c) Rotational hardening – rotate the yield surface with the rotation
of the symmetry axes of anisotropy, d) RIK hardening – combine all three hardening
definitions; the yield surface can grow, move, and rotate. ......................................... 9

Figure 2.1: The configurations and relationship between the configurations: reference
( B0 ), intermediate ( B, B ), and current ( B ) configurations. In infinite number of
intermediate configuration, a specific intermediate configuration is chosen by
leaving only stretch in the elastic deformation and all relative rotations are contained
in the plastic deformation, F = Ve R*U p . .................................................................. 23

Figure 2.2: Rotation angle of the symmetry axes using analytical solution and FE
analysis: (a) Ψ=30°, (b) Ψ=45°, (c) Ψ=60°............................................................... 24

Figure 2.3: Rotation angle of the symmetry axes using modified plastic spin scalar (a = -
350): (a) Ψ=30°, (b) Ψ=45°, (c) Ψ=60...................................................................... 26

Figure 2.4: Flow stresses for the RD, the TD, and 45°. The 45° flow stress approaches
the flow stress of the TD when the rotation of symmetry axes is used with an
isotropic hardening.................................................................................................... 28

Figure 3.1: Schematic stress-strain behavior of an IF steel during a monotonic simple


shear test (case a), a tension-shear test (case b) and a reverse-shear test (case c).
Characteristic crossing (∆c) and Bauschinger effect (∆b) are shown according to the
loading path............................................................................................................... 50

Figure 3.2: Multi-path loading response of conventional hardening models: (a) isotropic
hardening, (b) anisotropic nonlinear kinematic hardening model. Both models
cannot reproduce the ‘crossing’ effect while the Bauschinger effect is caputred in (b).
................................................................................................................................... 51

Figure 3.3: Rotational evolution of yield function: The yield function is expanded to the
stress state C when only the isotropic hardening is considered. The yield surface is
rotate by using rotational hardening and the iterated stress state is B. The yield
surface is moved toward C with the amount of backstress by the kinematic
hardening................................................................................................................... 52

Figure 3.4: Flow stress in the RD, 45°, and the TD orientation with a) ANK, b) RIK.
ANK predicts higher flow stress for 45° while RIK correlates the measurement data.
................................................................................................................................... 53

Figure 3.5: Flow curves for the 45° specimen: isotropic+kinematic hardening model
(ANK) produces rather higher stress ( ( R 45 )avg , ( R 45 )init ), rotation+isotropic

xiii
hardening model produces lower stress, while the RIK follows the measured data
rather closely............................................................................................................. 54

Figure 3.6: Rotation angle (degree) of the symmetry axes for 45° oriented specimen. .. 55

Figure 3.7: Stress offset tests for loading-unloading-reloading in different path: (a)
loading condition – give different loading path to the pre-stretched specimen, (b)
stress offset with respect to the plastic spin parameter. The stress is normalized with
respect to the offset of 180º ( ∆τϕ / ∆τ180o )................................................................. 56

Figure 3.8: Flow stress response when using the rotational hardening with isotropic
hardening scheme. Stress jump (crossing) is shown but Bauschinger effect is not
captured..................................................................................................................... 57

Figure 3.9: Flow stress result of the multi-path loading when the RIK hardening model is
used. Both ‘cross’ and ‘Bauschinger’ effects are captured...................................... 57

Figure 3.10: Multi-path loading responses show the flow stress crossing after various
pre-strains; more pre-strain yields higher jump in stress. ........................................ 58

Figure 4.1: Flow stresses of tensile loading at various orientations compared to


measuement data of DDQ from Numisheet 2002 [61]. ABAQUS standard was used
to compute the flow stresses. .................................................................................... 75

Figure 4.2: Transverse strain versus longitudinal strain of DDQ. The slope of each curve
is used to compute the R-values. (Numisheet 2002 [61])......................................... 76

Figure 4.3: Evolution of R-values: R0 and R90 are not evolving compared to the R45. . 77

Figure 4.4: R-value measurements for rolled copper in compression experiments.


Particularly R45 shows evolution.[59]....................................................................... 78

Figure 4.5: FEM validation of specimen dimensions to avoid clamping effect: a) strain
measurement locations, b) computed strains. The strains are closer enought to show
that the measured area is safe from the clamping effect........................................... 79

Figure 4.6: Schematic diagram of the experimental set up using DIC(Digital Image
Correlation). .............................................................................................................. 80

Figure 4.7: The random pattern sprayed on the specimen and captured by CCD camera: a)
before and b) after deformation. The small rectangles highlight a 100 x 100 pixels
subset of the image that has been deformed. ............................................................ 81

xiv
Figure 4.8: a) Comparison of stress-strain measurement by DIC(symbols) and strain
gauge (lines) for a small strain, b) comparison of strain measurement by DIC
(symbols) and NUMISHEET 2002 data (lines) for large strain. .............................. 82

Figure 4.9: DDQ experiments:a) Transverse-strain versus Longitudinal strain, b) flow


stress curves for each orientation. Stress prediction of 45° orientation is
overestimated at large strain over 5% strain. ............................................................ 83

Figure 4.10: Simulation results of averaged R value for 45° orientation comparing with
the experimental results: a) Transverse strain versus longitudinal strain, b) flow
stresses. Prediction of 45° orientation is underestimated. ....................................... 84

Figure 4.11: Sensitivity of the plastic spin parameter: a) R45 evolution, b) flow stresses of
45° orientation according to the plastic spin parameter. The gap between the
measurement and computed flow stress is the amount of backstress evolution. ...... 85

Figure 4.12: Simulation with the RIK hardening model and comparison with
experimental results: a) Transverse-strain versus longitudinal strain, b) flow stresses.
Prediction for 45° orientation captures the experiment well. ................................... 86

Figure 4.13: DDQ- data provided by NUMISHEET 2002. Computed by RIK hardening
model comparing with experimental results: a) Transverse strain versus longitudinal
strain, b) flow stresses, c) R45 evolution. .................................................................. 87

Figure 4.14: DQ – experiment. Computed by RIK hardening model comparing with


experimental results: a) Transverse strain versus longitudinal strain, b) Flow stresses,
c) R45 evolution. ........................................................................................................ 89

Figure 4.15: HSS – experiment. Computed by RIK hardening model comparing with
experimental results: a) Transverse strain versus longitudinal strain, b) Flow stresses.
................................................................................................................................... 91

Figure 4.16: Linear relationship between the plastic spin parameter (a) and the planar
anisotropy (∆R). This relationship can apply for mild steel which has B.C.C. crystal
structure..................................................................................................................... 92

Figure 5.1: Cylindrical cup drawing: a) Dimension of the tool set. (R1: punch
radius=50mm, R2: die radius=51.25mm, R3: punch fillet=9.5mm, R4: die fillet
radius=7.0mm, R0: blank radius=105mm<DDQ>), b) initial mesh....................... 106

Figure 5.2: Response in cylindrical cup drawing: a) Punch force-displacement, b) earing


patterns with respect to the hardening models. The difference in punch force is not
severe while the earing pattern shows a lot of difference with respect to the
hardening models. Higher friction coefficient gives better results. ....................... 107

xv
Figure 5.3: Cylindrical cup drawing: Residual von Mises stress distribution with respect
to (a) isotropic hardening, (b) RIK hardening ........................................................ 108

Figure 5.4: Rotation angle (in degree) of the symmetry axes of anisotropy................... 109

Figure 5.5: Square cup drawing - Forming tool for the deep drawing of a square cup (a)
and the FE model (b)............................................................................................... 110

Figure 5.6: Square cup drawing – Show the difference of a) Major and minor strain
distribution along the diagonal direction, b) thickness distribution along the diagonal
direction with respect to hardening models.. .......................................................... 111

Figure 5.7: Square cup drawing – Punch force-travel shows different history according to
the hardening models. ............................................................................................. 112

Figure 5.8: Square cup drawing – Draw-in with respect to hardening models compared to
the averaged experimental results of NUMISHEET ’93. ....................................... 113

Figure 5.9: Dimension of the draw bead experiment tools............................................. 114

Figure 5.10: Stress history during the process for the 30° oriented specimens .............. 115

Figure 5.11: Springback shape of the 30° oriented specimen: (a) rear view, (b) side view
................................................................................................................................. 116

Figure 5.12: Von Mises stress distribution before springback: a) isotropic, b) ANK, c)
RIK hardening model. RIK shows symmetric distribution of stress while the other
two show anti-symmetric distributions which causes unexpected twisting model
after springback....................................................................................................... 117

Figure 5.13: Hydroforming – Dimensions of the tool and initial tube and loading
conditions of the pressure and displacements at both ends of the tube. ................. 118

Figure 5.14: Residual equivalent stress distribution: a) isotropic hardening, b) ANK


hardening, c) RIK hardening. More stress at curved zone with RIK hardening model
(marked area). ......................................................................................................... 119

Figure 5.15: Equivalent plastic strain distribution: a) isotropic hardening, b) ANK


hardening, c) RIK hardening. The arrow points severely different area. .............. 120

Figure 5.16: Rotation angle of the symmetry axes of anisotropy. Rotational evolution of
the anisotropy is concentrated at the highest stress area......................................... 121

Figure 5.17: Reaction force at both edges when feed the material axially toward center in
the rate of displacement shown in Figure 5.13b. .................................................... 122
xvi
Figure 6.1: Stress flow measurements of each direction for Al-3%Cu precipitate
hardened material. Hardening of 45° shows more anisotropy than the RD and TD.
................................................................................................................................. 132

Figure 6.2: Neither conventional isotropic hardening (a) nor kinematic hardening (b)
with Barlat YLD96 anisotropic yield function does trace the measurements. ....... 133

Figure 6.3: Flow stresses of 45° for different backstress formulations. Those
formulations cannot trace the trend of anisotropic hardening shown in the
measurement. .......................................................................................................... 134

Figure 6.4: Flow stresses using the fourth-order tensor for precipitate inclusion without
rotation (a) and with rotation (b). The anisotropic hardening for 45° is captured by
rotation of backstress. ............................................................................................. 135

Figure 6.5: Earing pattern of cylindrical cup according to the hardening models: a)
Isotropic hardening only, b) Isotorpic + kinematic hardening, c) Isotropic +
Anisotropic kinematic hardeningm, d) Isotropic + Anisotropic kinematic hardening
with rotation. ........................................................................................................... 136

Figure 6.6: Residual stress distributions with respect to the hardening model: (a)
isotropic hardening, (b) anisotropic hardening with the fourth-order tensor, but no
rotation, (c) anisotropic hardening with the fourth-order tensor and rotation. ....... 137

Figure 6.7: Thickness distribution in 45° direction: a) Isotropic hardening, b) Isotropic +


Kinematic hardening, c) Isotropic + Anisotropic kinematic hardening, d) Isotropic +
Anisotropic kinematic hardening with rotation. ..................................................... 138

xvii
CHAPTER 1

INTRODUCTION

This research is focused on describing the anisotropic material behavior where the

texture development is significant. There are many sources for developing texture, such

as dislocation structures, lattice bending, and fragmentations [1]. The texture

development in the polycrystalline aggregate can be described by the rotation of the

preferred orientation which can be defined from the orientation distribution function

(ODF) [2, 3]. A preferred orientation is developed in certain crystallographic planes that

tend to rotate in a preferred manner according to the direction of maximum strain.

The preferred direction can be related to the symmetry axes of anisotropy and the

reorientation of the symmetry axes have been determined experimentally by both micro-

mechanical and macro-mechanical experiments. Micro-mechanical experiments have

been performed with the texture analysis using the X-ray pole figures to see the rotation

of the symmetry axes directly, while macro-mechanical experiments have been done by

fitting the yield stress of the pre-strained specimen to the symmetric anisotropy function.

The macro-mechanical experiment requires approximately 400 tensile tests for a material

[4].

The motivating experiments have been done by Boehler [5] and Kim and Yin [4].

Boehler performed both micro-mechanical and macro-mechanical experiments on a mild

steel. For the macro-mechanical experiments, specimen in each orientation to the RD as

1
shown in Figure 1.1(a) is prepared. The specimens are pre-strained up to 6%, and then

small specimens are cut at every 15° to the RD from the pre-strained specimen as shown

in Figure 1.1(b). They are stretched to measure the yield stresses and then fit the data of

the yield stresses to the symmetric yield function such as Hill’s ’48 [6]. The processes

for pre-strains of 6%, 14% and 36% are repeated. The angles used in here are defined

with the pre-straining direction and the orthogonal anisotropy axes in Figure 1.2. The

measured yield stresses are then fit to the curve shown in Figure 1.3, and the symmetry

axes of the curves can be set as shown with the blue lines in Figure 1.3. From the fitted

curve, one can observe that the symmetry axes are rotating toward the initial orientation

of 45°.

The micro-mechanical pole figures are shown in Figure 1.4. Interpreting the

figures, the symmetry axes originally located in a vertical direction rotate gradually, and

reach 45° at around 36% of strain. The rotation of the symmetry axes of anisotropy is

proved by these experiments.

In polycrystalline plasticity, the evolution of the crystallographic texture preferred

orientation and development of dislocation structures are known as the key features of

anisotropy evolution [29], as shown in Figure 1.5. These are different in scale but they

influence each other through slip processes. The preferred orientation is decided by the

slip process, and the slip process is defined by active slip system which lies closer to the

preferred orientation. The slip process influence the resistance against dislocation, again

the dislocation decides the slip process. In order to mimic the micro structural behavior

in phenomenological plasticity, the two key features should be modeled as in Figure 1.6.

The reorientation of the texture is considered to be the rotation of yield function by

assuming that the anisotropy axes are laid on the orthogonal axes of the yield function.

And the development of substructure is modeled by the evolution of hardening. The


2
relationship between the rotation of the yield function and the evolution of hardening is

defined with respect to the amount of the rotation.

The new phenomenological hardening model might include isotropic hardening

(to expand the yield surface), kinematic hardening (to move the yield surface), and

rotational hardening (to rotate the yield surface) as in Figure 1.7. The isotropic and

kinematic hardening descriptions have been developed by many researchers and can be

found in the literature. The rotational hardening is named for the reorientation of the

symmetry axes of anisotropy as assuming that the anisotropy axes are the basis of the

yield function, and the symmetry axes are rotating according to the reorientation of the

preferred direction.

The continuum setting for finite strain plasticity in intermediate configuration and

transformation of the formulation to the current configuration are summarized in Chapter

2. The setting is used basically in this research. The reorientation of the symmetry axes

using the plastic spin theory is also explained. The spin theory is combined with the

kinematics of the shell element to describe the evolution of planar anisotropy.

The latest isotropic combined anisotropic non-linear kinematic (ANK) hardening

model, suggested by Chun et al. [7, 8], is combined with the rotational hardening theory.

The resulting RIK hardening model is introduced in Chapter 3. The relation of the

symmetry axes rotation with the loading path change is discussed. The capability of the

RIK hardening model is demonstrated for multi-axial/-path loading cases.

Experimental approach for the anisotropy evolution and the identification of the

material properties are introduced in Chapter 4. Mild steel (DDQ, DQ) and high strength

steel (HSS) have been tested. Mild steel has a high rate of rotation, while HSS does not.

Questions about defining ‘R-values’ are raised and a method to determine the rotation of

the anisotropy using simple tension tests are discussed. A linear relationship between the
3
plastic spin parameter and the difference of plane anisotropy along orientation, ∆R, is

found. A method of identifying the plastic spin parameter from the R-values is proposed

for the materials which have the B.C.C. crystal structure such as mild steel.

Numerical applications in sheet metal forming and spring back are compared with

some experimental results in Chapter 5. The material properties identified in Chapter 4

are used. The RIK hardening model provided more reliable results than other

conventional hardening models. Deep drawing problems demonstrate the capability of

the RIK model in forming processes. The draw bead simulations in an arbitrary

orientation demonstrate the capability of the model for springback prediction.

For a precipitate hardened material, Al-3%Cu, the anisotropy cannot be estimated

properly by using existing models [9]. The preferred orientation of the precipitate

hardened material is determined by the orientation of the precipitate. As the preferred

direction of texture is decided by the crystallographic planes, the preferred direction of a

precipitate hardened material is decided by the precipitate. This can be treated in the

same manner as the rotation of the symmetry axes. With the hint, the anisotropic work

hardening for the aluminum alloy is investigated in Chapter 6. The material properties in

Barlat and Liu [10] are used, including Barlat 96 anisotropic yield function [11] and 3D

solid elements. The rotation of the symmetry axes is not limited to in-plane. As a

numerical application, a cylindrical cup drawing is demonstrated for the effect of the

model in the forming process.

Concluding remarks and future works are summarized in Chapter 7.

4
(a) (b)

Figure 1.1: Macro-mechanical test: a) A big specimen is cut from a rolled sheet with an
orientation Ψ to the RD and is pre-stretched by ε*, b) smaller specimens are cut from the
pre-stretched specimen in every 15° from the RD and are stretched to measure yield
stress

5
Figure 1.2: Definition of angles used in this research. The symmetry axes ( eiφ ), initial
orientation to the RD ( ψ ), angle of the symmetry axes rotation ( η), and the angle
between the symmetry axes and the straining directioin ( ϑ ). Symmetry axes of
anisotropy rotates toward pre-stretched direction (ε*) from the RD and TD whichever
closer.

6
Figure 1.3: Rotation of the symmetry axes as the pre-strain increases: measured yield
stress (0.2% offset, ●) and curve fitting to Hill’s 48 yield function [6].

{1,0,0} Ψ= 45°

Figure 1.4: Rotation of the symmetry axes observed by using pole figures for 45° initial
orientation. The symmetry axis is rotating gradually from TD counterclockwise to 45° as
strain increases [5].

7
Figure 1.5: Treatment of anisotropy evolution in polycrystalline plasticity. Texture
orientation and substructures are considered to be the key features of anisotropy which
are related through slip processes. [29]

Figure 1.6: Treatment of anisotropy evolution in phenomenological plasticity. Motivated


from the polycrystalline plasticity, yield function rotation and hardening evolution are
formulated to describe the anisotropy evolution in the RIK hardening model.

8
Figure 1.7: Schematic diagrams of hardening models: a) Isotropic hardening – expand
the yield surface uniformly, b) Kinematic hardening – move the yield surface to the
loading direction, c) Rotational hardening – rotate the yield surface with the rotation of
the symmetry axes of anisotropy, d) RIK hardening – combine all three hardening
definitions; the yield surface can grow, move, and rotate.

9
CHAPTER 2

SUMMARY OF FINITE STRAIN PLASTICITY IN INTERMEDIATE


CONFIGURATION AND THE ROTATIONAL EFFECT OF PLANAR ANISOTROPY

2.1 Introduction

In conventional elasto-plastic material models for anisotropic sheet materials, the

anisotropy axes are embedded in the material and thus rotate only if the material itself is

rotated. Induced deformations without material rotation do not change the direction of

the anisotropy axes, i.e., the deformation gradient is equal to the right or left stretch

tensor F = U = V , where the anisotropy axes are laid on symmetry planes of the yield

surface. Such a behavior can be verified experimentally in the rolling and transversal

directions of sheet metals. However, in uni-axial tension tests in other directions, e.g.,

30° to the RD, rotations of the anisotropy axes have been observed by several authors

using mechanical experiments and texture analysis, as discussed in Chapter 1. For mild

steels, rotations of up to 45° at 10% strain have been observed in these experiments,

related to the rotation of symmetry planes of the polycrystalline texture evolving with the

plastic deformation [5]. To model such a behavior, material formulations with a plastic

spin have been suggested by several investigators [12-14]. The attempt to relate the

material model with corotational shell elements is proposed by Han et al. [15], using an

isotropic hardening. The rotation rate of the anisotropy axes is supposed to be dependent

on materials because some aluminum sheets investigated in Bunge and Nielsen [3] and

Truong Qui and Lippmann [16] have much lower rotation rates of 5° at 20 percent strain.
10
The goal of this chapter is to summarize the continuum setting of the reorientation

of the anisotropy axes in intermediate configuration, which is generally used to formulate

finite strain. Also, the effect of the reorientation of the anisotropic axes in the plane

stress condition and detailed implicit algorithmic treatment for the rotation is discussed.

This model assumes that the anisotropy incurred in rolling processes is much larger than

the anisotropy due to subsequent strains by further forming. This assumption is used to

maintain the shape of the yield function that is defined as initial so the yield function can

only be rotated and expanded. The equations in the intermediate configuration are then

pushed forward into the current configuration using a small elastic deformation

assumption. It is generally accepted that elastic strains are negligible compare to plastic

strains at a large deformation for most metallic materials. An implicit algorithmic

treatment is presented for Kirchhoff-type shell formulation incorporating large material

rotations, finite strains, and assuming plane stress through the thickness direction.

The reorientations of the anisotropy axes with isotropic hardening are studied by

comparing the numerical results with the experimental data presented by Kim and Yin [4]

for mild steel.

2.2 Material formulation

The material formulation for finite strain is generally done in the stress-free

intermediated configuration, which is shown in Figure 2.1. Theoretically, there are

infinite intermediate configurations due to the rigid body rotation:

F = Fe Fp = FeQ T QFp = Fe Fp , (2.1)

where Q is an orthogonal tensor to represent the rigid body rotation. B0 and B are the

reference and current configuration, respectively. B denotes an arbitrary configuration

11
and B does the specific intermediate configuration which is used in this research. The

specific intermediate configuration is chosen by assuming that the elastic deformation

contains only small stretch ( Ve ) and all relative rotations are put in one rotation R * ,

as Fp = R e Fp = R *U p . In order to define the specific intermediate configuration, an

additional constitutive relationship is required to decide R * . The formulations in the

intermediate configuration are based on the multiplicative decomposition of the

deformation gradient

F = Fe Fp = Ve Fp . (2.2)

The velocity gradient is decomposed as

l = FF −1 = Ve Ve−1 + Ve Fp Fp−1Ve−1 , (2.3)

where the elastic and plastic parts in the current and the intermediate configuration are

defined by l e = Ve Ve−1 and L p = Fp Fp−1 , respectively. The plastic part in the intermediate

configuration can be split into the symmetric and the antisymmetric parts:

L p = Dp + Wp . (2.4)

The antisymmetric part of (2.4) is

(
Wp = ( L p ) = R *R*−1 ) + (R U U
* p
−1
p R *−1 ) . (2.5)
A A A

Equation (2.5) can be rewritten as [14]

Wp = Θ + Ω p , (2.6)

12
where Θ is the constitutive spin for defining R * which is considered to be the anisotropy

axes rotation and Ω p is the plastic spin, defined by Dafalias [17]

Ω p = µ φ ( PDp − Dp P ) , (2.7)

where P = Ce S denotes the Mandel stress tensor defined by the elastic Cauchy-Green

tensor Ce = FeT Fe and the second Piola-Kirchhoff tensor S , defined in the intermediate

configuration.

The symmetric part of L p is formulated using the associated flow rule and

consistency relation

∂φ
Dp = γ . (2.8)
∂P

The assumption of a small elastic strain is used to transform those formulae to the

current configuration. The transformations between the configurations are summarized

in Table 2.1.

The material spin in the current configuration is given as w = θ + ω p with the

plastic spin [14, 18]

ω p = µ φ ( τd p − d p τ ) , (2.9)

where τ is Kirchhoff stress. The anisotropy axes remain unchanged, i.e., eiφ = eiφ , for

small elastic strains as shown in Figure 2.1. The constitutive spin θ and the anisotropy

axes e iφ are assumed to be related to each other through a co-rotational rate

eiφ = eiφ − θeiφ = 0 . The scalar µ φ was defined by Dafalias [14] as


o


µφ = , (2.10)
φ
13
where cφ is material parameter and φ the yield flow stress. However, the definition is

insufficient to model different rates of rotation during deformation and different initial

orientations [14, 15]. The analytical solutions of Dafalias [14] and the FE model using

(2.10) are shown in Figure 2.2. They are not correlated with the experiments using a

single material parameter. Observing the experimental curve of the anisotropy axes

rotation, the rate of rotation is initially higher and it decreases gradually as the anisotropy

axes are rotated toward the straining direction. With this hint, a modified material

parameter definition was suggested by Han et al. [15]

a
µφ = tan ( ϑ ) , (2.11)
φ

where a is material parameter and ϑ ∈ [0 o ,45 o ] describes the angle between the

anisotropy axes and the pre-stretched direction ε∗ as shown in Figure 1.2.

2.3 Algorithmic treatment

The implicit algorithmic incremental treatment is given here for Kirchhoff-type

shell formulations applying the plane stress assumption and assuming that the transversal

shear strains are small. The deformation tensor F is treated in a specific way for shell

formulations since the transversal shear components of F are selectively integrated and

do not enter the elasto-plastic material description (see Han et al. [15]).

The plastic arc length s is introduced with

s 2 ∂φ 2
γ= , where β = and s = dp , (2.12)
β 3 ∂τ 3

which is used for the description of the isotropic hardening and evolution of the yield

stress

14
iso
τ y = c iso (s o + s p ) n . (2.13)

.
The Oldroyd rates for stresses are given with (.) = (.)− l T (.) − (.)l which is identical to the

∂ 
Lie-derivative Lv (.) = F  (.)  F T [19] and can be written in incremental form as
 ∂t 

n +1
τ = F n τ F T + n +1∆τ , (2.14)

where the left superscript denotes the loading step: n is for current and n+1 is for next

loading step.

With ∆ε = d∆t the stress increments are given as

∆τ = Γ e (∆ε − ∆ε p ) , (2.15)

where ∆ε can be identified with the incremental Almansi strain tensor and Γ e is the

elasticity tensor.

The constitutive equations are performed with the plane stress condition in the

tangential coordinate system of the shell. In order to remain in a two-dimensional

coordinate system, (2.14) is formulated relative to the tangential referential coordinate

system written as τˆ = U n τ U T + n +1∆τ . The stress tensor in the current configuration is

ˆ T , where the tensor R is defined as R = FU −1 . However, the


then recovered by τ = RτR

coefficients of τ do not change, if the corotational shell element is used.

Correspondingly, ∆εˆ = R T ∆εR = 1


2 (1 − c ) is obtained for the strain increments.
−1
For the

trial stresses, the plastic strain increments are assumed to be zero, yielding
trial
τ = U n τ U T + Γ e ∆ε . The Hill’s anisotropy yield function is applied and defined by

Φ = τ ⋅ K τ − τ y2 with the fourth-order anisotropic tensor K whose components are defined

with the orthogonal axes as K αβγδeαφ ⊗ eβφ ⊗ eφγ ⊗ eδφ . It is convenient to convert the

15
symmetric tensor K into a matrix form P so that the plastic strain can be related to the

stress in the matrix form,

 ε11
p
  P11 P12 P13   τ11 
 p    
 ε 22  =  P21 P22 P23  τ22  . (2.16)
 p P33   τ12 
2ε12   P31 P32

While the components of the fourth-order tensor K relates,

 ε11p
  K1111 K1112 K1121 K1122   τ11 
 p  
 ε12   K1211 K1212 K1221 K1222   τ12 
 p =   (2.17)
 ε 21   K 2111 K 2112 K 2121 K 2122   τ21 
 
ε 22   K 2211 K 2222  τ22 
p
K 2212 K 2221

By comparing (2.16) and (2.17), the matrix P can be described with the components of

the tensor K and vice versa:

 K1111 K1122 2K1112 


P =  K 2211 K 2222 2K 2212  and (2.18)
 2K1211 2K1222 4K1212 

 P11 P13 / 2 P12 


P13 / 2
P / 2 P33 / 4 P33 / 4 P32 / 2 
K =  31 . (2.19)
 P31 / 2 P33 / 4 P33 / 4 P32 / 2 
 
 P21 P23 / 2 P23 / 2 P22 

The initial value of P is defined at the initial loading step as

 1 −β12 0 
0
P =  −β12 β22 0  (2.20)
 0 0 β66 

16
where the three parameters (β12, β22, and β66) are defined in terms of the R-values:

R0 R (1 + R 90 ) ( R + R 90 )(1 + 2R 45 ) .
β12 = , β22 = 0 , β66 = 0 (2.21)
1+ R0 R 90 (1 + R 0 ) R 90 (1 + R 90 )

According to (2.20) and (2.19), the initial components of K is defined as

 1 0 0 −β12 
 0 β66 / 4 β66 / 4 0 
0
K mnpq = (2.22)
 0 β66 / 4 β66 / 4 0 
 
 −β12 0 0 β22 

Using such a vector notation, the algorithmic procedure from load step n to n+1 is

outlined in Table 2.3.

The trial orientation for the anisotropic yield function is computed with
trial φ
e = R n eφα . The directional evolution of eφα in plastic deformation is described as
α

n +1 φ
e = R*φ n eφα . The incremental orthogonal transformation R *φ is corresponding to the
α

constitutive spin ∆θφ which is defined ∆w φ − ∆ω φp . With the trigonometric additive

relations, the rotation can be described in the form

R *φ = R Tωp R . (2.23)

Therefore the anisotropy axes direction in next step is directly related to the rotation

which correspond to the plastic spin as

eφα i +1 = R ωT p trial φ
e .α (2.24)

The rotation tensor R ωp is defined in relating to the incremental plastic spin

( )
∆ω φp i +1 = µφ τ i +1∆ε p − ∆ε p τ i +1 with

17
 cos ∆ωφp12i +1 sin ∆ωφp12i +1 
R ωp =  φ i +1 . (2.25)
 − sin ∆ωp12 cos ∆ωφp12i +1 

Updating the fourth-order anisotropic tensor K i +1 with the evolution of the symmetry

axes eφα i +1 in (2.24)

+1
K iαβγδ = ∑∑∑∑
m =1,2 n =1,2 p =1,2 q =1,2
0
( i +1

)( i +1

)( i +1
K mnpq eαφ ⋅ 0e m eβφ ⋅ 0e n eφγ ⋅ 0e p eφδ ⋅ 0eq , )( i +1

) (2.26)

where the superscript 0 means the initial loading step and initial direction of the

anisotropy axes. Therefore, e0m is (1,0) and (0,1) when e10 is along the RD. The matrix

notation P i +1 is updated with respect to the Ki+1 by (2.19),

i +1
 K1111 K1122 2K1112 
P =  K 2211
i +1
K 2222 2K 2212  (2.27)
 2K1211 2K1222 4K1212 

2.4 Numerical simulation

To assess the performance of the derived stress-update algorithm, the uniaxial

stretch tests investigated by Dafalias [14] and Kim and Yin [4] are considered.

Rectangular samples are cut from a larger sheet in 30°, 45°, 60° angles from the rolling

direction (RD) and stretched along in these directions. An analytic solution for rigid-

plastic material, incorporating plastic spin and Hill’s yield function without hardening is

presented in Dafalias [14]. The material parameters used for the simulation with the

proposed algorithm are taken from the pre-stretched mild steel described in Kim and Yin

[4], and are given in Table 2.2. The results of this approach for the tensile stretch test in

the directions of 30°, 45° and 60° angles are presented in Figure 2.3 using a single

material parameter, a = - 350.

18
In Figure 2.4, the flow stresses obtained with the plastic spin are illustrated. The

flow stresses of 0°, 90° and 45° angles are illustrated as results of the FE analysis. The

flow stress of the 45° angle approaches the flow stress of the 90° angle as the symmetry

axes of anisotropy rotates. The phenomenon incurred by adopting the plastic spin will be

discussed and corrected by the kinematic hardening in Chapter 3.

19
Intermediate Current
Configuration Configuration

L = Le + L p ≅ L p l = Fe Fe−1 + Fe Fp Fp−1Fe−1
= l e + Fe L p Fe−1
= l e + ˆl p

D = ( L )S d = ( l )S

= ( Lp ) = ( Dp + Wp )
S S
( )
= ( l e )S + ˆl p
S

= Dp = ( l e )S + ( Fe Dp Fe−1 ) + ( Fe Wp Fe−1 )
S S
−1
= d e + Ve Dp V e

= d e + dˆ p ≅ d e + d p

W = ( L p ) = ( Fp Fp−1 )
A A
w = ( l )A = ( l e )A + ˆl p ( ) A

= ( R*R −1
* A ) + (R U U
* p
−1
p R )
−1
* A (
= 0 + Ve Dp V ) + (V W V )
e
−1
e p e
−1
A A

= Θp + Ωp (
= VeΘ p Ve−1 ) + (V Ω V )e p e
−1

= θp + ωp

E = Ee + E p e = ee + ep
1 1
=
1 T
2
( 1
) (
Fe Fe − 1 + 1 − Fp− T Fp−1
2
) =
2
( ) (
1 − Ve−1Ve−1 + Ve−1Ve−1 − F − T F −1
2
)

P = (1 + 2Ee ) S = CeS τ = Fe SFeT = Ve SVeT ≅ S


S = Fe−1τFe− T τ = Ce−1P = Ve−1Ve− T P ≅ P
∂Φ ∂Φ
Dp = γ dp = γ
∂P ∂τ
Ω p = µ ( PDp − Dp P )
φ
ω p = µ ( τd p − d p τ )
φ

Table 2.1: Comparison of continuum setting in the intermediate configuration and the
( ε e ) ) is used.
current configuration. Small elastic strain assumption ( Ve = I +O

20
Young’s modulus E = 206 GPa
Poisson’s ratio ν = 0.3
Initial yield stress τ0 = 107.06 MPa
Hill’s yield function β12 = 0.5837, β22 = 1.0092, β66 = 2.3550
Isotropic hardening ciso = 544 MPa, niso = 0.25

Table 2.2: Material parameters of Kim and Yin [4].

21
n
Input : F n +1 , τ n , s np , e φα
Compute: ∆ε , trial
τ , trial e φα , trial Φ = Φ ( trial τ, trial e φα , τ yn ) , trial
P
if Φ < tol then
n +1
τ n +1 = trial τ , s np +1 = s np , e φα = trial φ
e , Γ ep = Γ e
α

if Φ ≥ tol then
τ 0 = trial τ, s 0p = s pn , Φ 0 = trial Φ, P 0 = trial P
∆γ 0 = 0
do i = 0, imax
compute
• a = 2P τ , rτ = τ −
i i i
( trial
τ − ∆γ i Γ e a )
• (
Q = 1 + 2∆γ i Γ e P i )
Φ − aT Q −1rτ
• ∆∆γ =
aT Q −1Γ ea + 4H ' τ iT P i τ ( )
with H ′ = n iso c iso (s 0 + s p )
iso
n −1

• s ip+1 = s ip + ∆s p
with ∆s p = 2∆γ i τ iT P i τ i
• τ i +1 = τ i + ∆τ
with ∆τ = −Q −1 (rτ + ∆∆γΓ e a )
• ∆γ i +1 = ∆γ i + ∆∆γ
i +1

• eφα = R ωT p trial φ
e
α

if Φ i +1 < tol then


compute a , Q with τ i +1 , P i +1 , ∆γ i +1
(Q )( )
T
−1
Γ e a Q −1 Γ e a
Γ ep = Q −1Γ e −
(
aT Q −1Γ ea + 4H ' τ i +1 P i +1τ i +1
T

)
n +1 i +1
τ n + 1 = τ i + 1, s np + 1 = s ip+ 1 , e φα = e φα
exit to Output
end if
end do
end if
n +1
Output : Γ ep , τ n +1 , s np +1 , e φα

Table 2.3: Algorithmic step (implicit)

22
Figure 2.1: The configurations and relationship between the configurations: reference
( B0 ), intermediate ( B, B ), and current ( B ) configurations. A specific intermediate
configuration is chosen by leaving only Ve in the elastic deformation and all relative
rotations are put in R*, as F = Ve R*U p .

23
40
ο
Ψ = 30
35 φ
c = -300

30

ο
η
25 φ
c = -100
Rotation angle 20 φ
c = -200
15
Exp. (Kim & Yin)
10
FEM
5 Analytical Solution
(Dafalias)
0
0 2 4 6 8 10
Strain (%)
(a)

0
ο
Ψ =45

-10
φ
c = -100
ο
η

-20
Rotation angle

φ
c = -200
φ
-30 c = -300

-40

-50
0 2 4 6 8 10
Strain (%)
(b)

(continued)

Figure 2.2: Rotation angle of the symmetry axes using analytical solution and FE
analysis: (a) Ψ=30°, (b) Ψ=45°, (c) Ψ=60°.

24
(Figure 2.2 continued)

0
ο
Ψ = 60
-5

-10
φ
ο

c = -100
η

-15
Rotation angle

φ
-20 c = -200

-25

-30 φ
c = -300
-35

-40
0 2 4 6 8 10
Strain (%)

(c)

25
40
ο
Ψ = 30
35

30 FEM

ο
η
25

Rotation angle
20

15

10 EXP
(Kim & Yin)
5

0
0 2 4 6 8 10
Strain (%)
(a)

0
ο
Ψ = 45

-10
ο
η

-20
Rotation angle

-30 FEM

-40
EXP
(Kim & Yin)
-50
0 2 4 6 8 10
Strain (%)
(b)
(continued)
Figure 2.3: Rotation angle of the symmetry axes using modified plastic spin scalar (a = -
350): (a) Ψ=30°, (b) Ψ=45°, (c) Ψ=60

26
(Figure 2.3 continued)

0
ο
Ψ = 60
-5

-10
ο
η

-15
Rotation angle

-20 FEM

-25

-30
EXP
(Kim & Yin)
-35

-40
0 2 4 6 8 10
Strain (%)

(c)

27
400
o
0

o
300 45
Stress (MPa)

o
90
200

100

0
0 5 10 15 20
Strain (%)

Figure 2.4: Flow stresses for the RD, the TD, and 45°. The 45° flow stress approaches
the flow stress of the TD when the rotation of symmetry axes is used with an isotropic
hardening.

28
CHAPTER 3

MODELING MULTI-AXIAL ELASTO-PLASTIC DEFORMATION OF PLANAR


ANISOTROPIC MATERIAL, PART I: THEORY

The behaviors of the material points under multi-axial and multi-path loadings are

not modeled accurately by phenomenological models. With the help of the advances in

micro-mechanical investigations for the material, the behavior in multi-axial/-path

loadings has been being unveiled in polycrystalline plasticity. However, the formulation

requires a large set of parameters to describe the micro-mechanical evolution as well as

requiring a massive computation. Therefore, it is difficult to use in large scale industry

problems, even though it has accuracy in describing the material behavior. Motivated by

the micro-mechanical behaviors and polycrystalline approach, a new phenomenological

plastic hardening model named the Rotational-Isotropic-Kinematic (RIK), is suggested

herein. This model is developed to transform the major effects of micro-mechanical

behaviors to a phenomenological model for the computation. The proposed RIK model

allows the yield function to grow (isotropic hardening), move (kinematic hardening), and

even rotate (rotational hardening) as the deformation continues.

The RIK model has been formulated based on the Armstrong-Frederick type

kinematic hardening model and on the rotation of the symmetry axes of anisotropy with

the plastic spin theory. The model is accounting for the different backstress evolutions

for loading and reversal loading with respect to the difference of the loading direction for

29
the multi-axial loading and correcting the kinematic hardening with respect to the

reorientation of the symmetry axes. The ability of the model is demonstrated by the

multi-path loading of ‘tension-shear’ and ‘reverse-shear’ to reveal the ‘crossing’ and

‘Bauschinger’ effects. The effects of the RIK model in deformation and springback are

shown in later chapters.

3.1 Introduction.

The goal of this article is to describe the material behavior under multi-axial/-path

loadings in a phenomenological framework for planar anisotropic material.

Many investigations can be found in the literature for the multi-axial/-path

loadings. Tensile loading in the transverse direction (TD) after pre-straining in the

rolling direction (RD) was done by Doucet and Wagoner in plane stress condition [20,

21]. Simple shear test after various amounts of pre-strain was investigated by Rauch [22]

and shear test with various orientation angles to the RD, after pre-strain in the RD was

done by Thuillier and Rauch [23]. Hiwatashi et al. [24, 25] suggested a non-proportional

biaxial loading experiment and the material model for the loading. Takahasi et al. [26]

reported the experimental results of torsion-tension behaviors of aluminum tubes.

Teodosiu and Hu [27] suggested three simple tests: (case a), a simple shear test with the

shearing direction (SD) parallel to RD and the shear plane normal parallel to the TD;

(case b), tension-shear test, pre-tensile strain in the RD, followed by simple shear with

the SD also parallel to the RD; and (case c), a reverse-shear test, an amount of simple

shear in the RD, and then simple shear in the opposite direction. The schematic diagrams

of the tests are shown in Figure 3.1 with the corresponding flow stress results. The

experiments show two major effects of multi-path loading, ‘crossing’ and ‘Bauschinger’

effects, which occurs in tension-shear and reverse-shear loadings, respectively. The

30
abrupt change of the strain path causes the ‘crossing’ of the flow stress, and it gradually

approaches to the monotonic shear curve. The ‘Bauschinger’ effect and the permanent

offset are related to the kinematic hardening.

The advance in polycrystalline metal plasticity based on texture analysis makes it

possible to reproduce the complex material behavior in multi-axial/-path loadings [28-30].

However, these approaches remain impractical for use with large-scale boundary value

problems such as those encountered in metal forming or other structural problems. The

principal obstacle is the inability to represent the essential aspects of texture-based results

in a phenomenological form that can be implemented for efficient FE analysis.

Only a limited number of the phenomenological approach can be found in the

literature. Hiwatashi et al. [24] approached this strain-path change problem with the

phenomenological constitutive equation. However, he was not concerned about texture

development and the anisotropic yield function.

With the investigation of Peeters et al. [28, 29], the micro-structural developments

are mainly dislocation structures and reorientation of the preferred direction. The

development of the dislocation structures is considered to be the source of isotropic and

kinematic hardening, while the reorientation of the preferred direction can be considered

to be the rotation of the symmetry axes, which is related with the plastic spin [5, 15, 31-

33]. For the macroscopic approach to the Bauschinger effect, Armstrong-Frederick-type

non-linear kinematic hardening models [34] have been used to capture the nonlinear

hardening behavior and smooth transition from elastic to plastic deformation relatively

well. However, the reversal loading of the models approaches the monotonic hardening

curve and does not show the permanent softening characteristics. Chun et al. [7]

generalized the Armstrong-Frederick type model by introducing an independent

evolution of the bounding surface to reproduce the permanent offset, and also suggested
31
an advanced form of isotropic hardening to represent the uniaxial tensile experiment with

the kinematic hardening which is modified from Chaboche’s model.

Computation of ‘tension-shear’ and ‘reverse-shear’ loadings are performed with

the conventional isotropic, kinematic hardening and Chun’s ANK hardening model. The

results of isotropic hardening and the ANK hardening are shown in Figure 3.2. The

‘Bauschinger’ effect and permanent offset are captured by the ANK model, but the

‘crossing’ has not been captured with those hardening models. The capability of

reproducing the characteristics with respect to the hardening models is compared in Table

3.1. The existing models cannot reproduce the ‘crossing’ effect, even though the latest

ANK model can account for the directional effect between loading and reloading and can

capture the permanent offset. It is because the models do not have the ability to account

for the rotational hardening.

The texture development which is another polycrystalline aspect of the multi-

axial/-path loadings is considered to be the evolution of the anisotropy. It can be

formulated with the rotation of the symmetry axes of anisotropy and the formulation has

been introduced by the authors using the plastic spin theory [15]. The rotation of the

symmetry axes have been investigated experimentally by Boehler [5] and Kim and Yin

[4] for mild steel, and by Bunge [3] and Truong Qui [35] for aluminum alloy. Macro-

mechanical modeling of the plastic spin is done by several authors, e.g., Dafalias [14] and

Kuroda [13]. The reorientation of the anisotropy axes affects the yield function, and the

rotated yield function makes different evolutions in flow stress. Therefore, the anisotropy

evolution due to the rotation of the yield function can be called ‘rotational hardening’

along with isotropic and kinematic hardening [36].

The goal of this research is to develop a phenomenological plasticity model which

can handle the rotation, translation, and expansion of the yield function simultaneously as
32
motivated by the micro-mechanical behavior of the material. This research will also

define the relation of the backstress and the rotational hardening in correcting the flow

stress. The suggested hardening model is then called RIK hardening.

The results of the new RIK hardening model will be compared qualitatively with

the experimental results of Peeters [28]. The comparison will show that the suggested

RIK model is acceptable for reproducing the multi-path loading in a phenomenological

framework.

The model is derived in general form, and then formulated in planar anisotropic

and a time-independent plane stress problem specifically for the sheet metal forming

process. Demonstrating the flow curves in the RD, TD and 45° all follow the

experimental curve quite accurately in contrast to conventional models. The ‘crossing’

effect after various pre-strain also demonstrate the effect of pre-straining. The

formulation and numerical algorithms are done by the return mapping method and are

implemented on both ABAQUS/Implicit and ABAQUS/Explicit using user interface

material subroutines.

3.2 Material formulation

3.2.1 General model for isotropic, kinematic, and rotational hardening

The material formulation is based on the multiplicative decomposition of the

deformation gradient F = Fe Fp [37] and intermediate configuration in Figure 2.1, yielding

the velocity gradient as

l = FF −1 = Fe Fe−1 + Fe Fp Fp−1Fe−1 , (3.1)

where the elastic and plastic parts in the intermediate and current configuration are

defined by l e = Fe Fe−1 and L p = Fp Fp−1 , respectively. The upper bar denotes an


33
intermediate configuration and small letter denotes current configuration. With the small

elastic strain assumption, the symmetric part of l can be decomposed into an elastic and

plastic rate of deformation as

( l )S = d = de + d p . (3.2)

An yield function can be generalized by counting isotropic, kinematic hardening as well

as the rotation of the yield function and can be written as

Φ ( τ, y, q; eφα ) = φ ( τ − y; eφα ) + σ y , (3.3)

where τ is Kirchhoff stress tensor, y is backstress tensor, σ y is the isotropic hardening

part and eφα is the direction vector of the symmetry axes of the anisotropy which is

defined by the yield function φ . Considering the anisotropy, φ in (3.3) describes an

anisotropic yield function such as Hill’s [6] and Barlat’s [11], etc. The backstress y in

(3.3) can be considered to be the sum of multiple backstress component vectors

y = ∑qj . (3.4)
j

Those stress-like-internal variables are related to their work conjugated strain like

variables a j , s as

q j = −c ja j , σ y = σ y ( s ) (3.5)

where c j is a positive definite anisotropy tensor and it can be tensor, vector, and scalar

with respect to the order of the tensor.

Following Dafalias [14, 33], the anti-symmetric part can be decomposed into the

constitutive spin and the plastic spin

34
( l )A = w = θ + ω p , (3.6)

and the orientation of the anisotropic yield function is presumed to be defined by the

anisotropy axes eφα . The anisotropy axes eφα are related to the constitutive spin θ by a co-

rotational rate

o
eφα = eφα − θeφα = 0 . (3.7)

The anisotropy axes rotation does not evolve internal stress or strains, but the flow stress

can change the orientation of the yield function. Therefore, it can be called “rotational

hardening” [36].

The constitutive relation for plasticity is defined by the associated flow rule as

∂φ
dp = γ (3.8)
∂τ

s
and the consistent parameter γ can be described as γ = where s = 2
3 d p and
β
∂φ
β= 2
3 ∂τ .

Although it is classically assumed that the state variables of τ, q j , σ y , eφα evolve

independently, for simplicity, an interaction of these state variables in general can be

motivated by considerations on smaller scales. The anisotropy axes may affect the other

variables and the other state variables may affect the anisotropy axes.

τ = f τ ( d e , d p , eαφ )
q = f q ( d p , s, eφα )
(3.9)
σ y = f σy ( d e , d p , s, eφα )
eφα = f e ( τ, q, d p , σ y )

35
The relationship between the isotropic hardening and the kinematic hardening can

be found in the literature. Combining Armstrong-Frederick type kinematic hardening

with isotropic hardening was suggested by Chaboche [38, 39] and modified by Chun et al.

[7]. The relationship between the rotational hardening and the isotropic hardening was

investigated by Dafalias [14, 33], [13] and Han et al. [15]. Loret [40], Dafalias [17, 41,

42], Van der Gissen [43] suggest the plastic spin formulation for kinematic hardening.

Kuroda [13, 18, 44] and Dafalias [14] define a simple form of the plastic spin by

defining the plastic spin with the non-coaxiality of the stress and the plastic strain

increment direction. The relationship between the isotropic hardening, the kinematic

hardening, and the rotational hardening is suggested in this paper.

3.2.2 Rotation of the symmetry axes of anisotropy (rotational hardening).

Dafalias [14], Kuroda [13, 18, 44] and Zbib [45] define the plastic spin in the

current configuration as

ω p = µφ ( τd p − d p τ ) or (3.10)
ω p = µ φ {( τ − y ) d p − d p ( τ − y )} . (3.11)

From the observation of the experimental results of Kim and Yin [4], the rate of the

rotation is initially faster and then decreases as the symmetry axes rotate toward the

straining direction. Therefore, the rate factor (scaling parameter µ φ ) is considered to be a

function of the angle between the anisotropy axes direction and the straining direction [15,

46],

a
µφ = µφ ( ϑ) = tan ( ϑ ) , (3.12)
φ

where the angle ϑ is defined as

36
 φ 
−1 e α ⋅ n  π
ϑ = min cos φ
 , ϑ∈ 0,  (3.13)
α=1,2,3
 eα n   4

The vector n in (3.13) is the eigenvector of d p . For the plane stress setting such as sheet

metal, the angle is defined as in Figure 1.2.

Detailed formulation and the numerical algorithm are presented in Chapter 2 with

the isotropic hardening model and compared with Kim and Yin’s experimental results [4].

Choi et al. [46] showed the effect of the anisotropy axes rotation on springback.

3.2.3 Armstrong-Frederick type backstress components

From an analogy to the plastic strain rate, the evolution of the strain-like internal

variable for the kinematic hardening is written as

a ∆j = sb ja j − d p (3.14)

Following the derivative of the stress-like internal variable q j following Haupt and

Tsakmakis [47], the rate form of (3.5) is applied as

q∇j = −c ja ∆j (3.15)

Insert (3.14) to (3.15), then as a result,

q∇j = c jd p − sb jq j (3.16)

The evolution of the backstress in (3.16) would have variation according to the

material concerned. The variable c j in (3.16) can be a constant, scalar function for multi-

axial loading, and a tensor which describes the accommodation tensor of the material for

elastic inclusion of the precipitate [48] and also for texture development.

Thermodynamically consistent derivations and their algorithmic treatment of Armstrong-

37
Frederick type hardening from the intermediate configuration can be found in the

literature [47, 49, 50].

3.2.4 Nonlinear kinematic hardening with permanent softening.

In order to capture the Bauschinger effect with permanent offset in the planar

anisotropic time-independent plane stress problem, a different evolution scheme of the

backstress component has been introduced by Chun et al. [7] and Geng and Wagoner

[51]. The model proposed by Chun et al. has two backstress components and the first

component saturates to a certain amount, while the second component evolution is

dependent on the loading path.

The evolution equations of the first component are described in the Armstrong-

Frederick type nonlinear hardening, and the second component could be a simpler linear

form for springback prediction is suggested when a cyclic loading is expected during the

forming process [7]. The modified backstress components are:

q1∇ = c1d p − sb1q1 , q∇2 = k ( ϕ, ϑ ) c 2d p


 π
 1 + κϑ if ϕ ≤ 2 , initial loading . (3.17)
k ( ϕ, ϑ ) = 
 κϑ π
if ϕ > , reversal loading
 2

The parameter φ denotes the angle between the loading and reloading,

 d p ⋅ d*p 
ϕ = cos −1  , (3.18)
 d p d*p 
 

where, d p is for the current loading, while d*p is for the previous loading step.

The second backstress component is evolving with respect to the loading path

history. Once the loading path is in the reverse direction, the second backstress evolves
38
in a different scheme. The evolution of the second backstress is then considered as a

function of the angle between the previous and current loading paths. The angle can be

considered to be the difference between the directions of the symmetry axes and the

loading direction. Because the symmetry axes of anisotropy is rotating toward the

loading direction, the symmetry axes direction of the previous loading path is the

reference to measure the directional change of the current loading path. The evolution of

the second backstress is then formulated as in (3.17).

3.2.5 Correction of flow stress with kinematic hardening associated with the anisotropy
axes rotation

The micro-mechanical motivation of the formulation is related with the

development of dislocation structures and its pile up at obstacles. When a polycrystalline

material is under plastic loading, each grain has micro-mechanical structures, such as slip

planes in its own direction. The directions are represented as the preferred direction of

the aggregates or anisotropy axes which can be found statistically by the orientation

distribution function (ODF). The mechanism of deciding the direction of slip planes of

each grain is beyond the phenomenological framework. However, the direction is

assumed to be maintained as long as the external loading direction is not changed. When

the loading path is changed, new slip planes of each grain develop in a different direction

according to the new loading direction. The mechanism is not well known and the

prediction of the dislocation direction is impossible; however, the preferred direction of

the aggregate will be rotated aligning the loading direction.

Now consider the interpretation of Peeters et al. [28,29]. Once the slip planes are

developed in a direction according to the previous loading path, the existing slip planes

act as obstacles to the newly developing dislocation according to the current loading.

The dislocation density would increase at the ‘old’ slip planes and the resistance would
39
increase. As the loading is continuous in the new direction, the micro-band occurs as the

resistance is over the theoretical shear stress of the ‘old’ slip planes in the region and they

cut the ‘old’ slip planes. More micro-band occurs as the loading continues, and gradually

the broken structures of the old slip planes are aligned to the new direction.

This procedure looks as if the slip planes are rotated according to the new loading

path. The resistance due to the pile up at the old slip planes decreases as the obstacles

rotate toward a new loading direction. The direction of each grain defines the preferred

direction of the aggregate in a statistical manner, so it is assumed that the resisting

behavior of the aggregate can be formulated with the rotation of the symmetry axes of the

anisotropy.

If the loading is in the RD, then the initial direction of the symmetry axes are

aligned with the loading path. Hence the angle ϑ is zero. Therefore, the second

backstress evolution in (3.17) will be c2. When the loading path is changed to the

opposite direction, it is reversal loading, but the angle between the loading direction and

the symmetry axes direction will still be zero. Therefore, the second backstress evolution

is zero. This is identical to the ANK hardening model as long as the loading and

reloading are in the RD. In contrast to the ANK model, the RIK model can account for

all directions between the previous and current loading paths.

As previous work on the anisotropy axes rotation in Chapter 2 indicates, the

rotation makes flow stress in any orientation approaching to the RD or the TD flow stress

rapidly and following the RD or the TD flow stress afterwards as shown in Figure 2.4.

However, the flow curves in the experiment do not approach the RD or the TD.

Therefore, it is necessary to correct the flow stress as long as the rotational hardening is

used. The mechanism of the hardening models with the rotational effect of the yield

surface is illustrated in Figure 3.3. Yield function will grow from A to C when only the
40
isotropic hardening is used. If the rotational hardening is added, the yield surface will

rotate and grow to a stable condition. And then the stress that satisfies the yield function

is iterated from A to B, not C. Once the yield surface is rotated and the direction of the

symmetry axes and the direction of loading are aligned, the yield surface is not rotated

any further. The hardening model in this article does not change the yield surface shape,

but only changes the size, location, and direction. This is the explanation for why all of

the flow curves are approaching the RD or the TD after the anisotropy axes rotate to the

loading direction, and it could be corrected with kinematic hardening. The distance of

the two stress states B and C would be recovered by the associated backstress, α , and it

needs to add the kinematic hardening in the model.

3.3 Application on plane stress problem for sheet metal forming.

The formulations are now applied to the plane stress problem with shell element

to apply the theory to sheet metal forming process. The quadratic yield function is

Φ = φ − σ 2y , (3.19)

where φ = ( τ − y ) ⋅ K ( τ − y ) for Hill’s 48 anisotropic yield function. Initially K is

represented in matrix P for plane stress problem as

 1 −β12 0   τ11   y11 


   
P =  −β12 β22 
0  , τ = τ22  , y =  y 22  (3.20)
 0 β66     
0  τ12   y12 

where the components are related to the R-values as [52] and [7]

R0 R (1 + R 90 ) ( R + R 90 )(1 + 2R 45 ) .
β12 = , β22 = 0 , β66 = 0 (3.21)
1+ R0 R 90 (1 + R 0 ) R 90 (1 + R 0 )

41
In order to avoid unwanted inconsistency for unidirectional loading with the

anisotropic yield function dp is commonly replaced with ( τ − y ) , the backstress evolution

equations in (3.17) are rewritten with (3.8) as stress formulation,

s
q1∇ = c1 ( τ − y ) − sb1q1
β
(3.22)
s
q∇2 = kc 2 ( τ − y )
β

The yield function is completed by defining the size of yield function q in (3.3),

which is modified from the Chaboche [39] to follow the simple tension test curve by

Chun et al. [7].

c1
(
σ y = σo + K 1 − e − Ns − ) b1
( )
1 − e − b1s − c 2s . (3.23)

The axes of anisotropy e φα in (3.7) can be updated via an incremental orthogonal

transformation R φ , corresponding to θ φ ∆t , where R φ has to satisfy R φ = θ φ R φ

and R φ (t ) = 1 . For an application within an algorithmic treatment, e iφ can be updated by

t +∆t
eiφ = t +∆t R φ t eiφ . (3.24)

In the case of a purely elastic increment, the plastic spin is identical to the zero

matrix ω φp = 0 , and hence, θ φ = w = RR T . Also R φ in (3.24) can be identified by


t + ∆t
R φ = t + ∆t R t R −1 in a total Lagrangian formulation, and simply by t + ∆t
R φ = t + ∆t R in an

updated Lagrangian formulation. In Hill’s anisotropic yield function, the fourth-order

tensor K would rotate along the basis, which is rotated with the symmetry axes of

anisotropy in the following form:

K ijkleiφ ⊗ eφj ⊗ eφk ⊗ eφl . (3.25)


42
The procedure to define the components of the tensor is described in Chapter 2.

The formulation and numerical algorithms are completed with a return mapping

method and implemented on both ABAQUS/Implicit and ABAQUS/Explicit using user

interface material subroutines. The constitutive equations used in this paper are

summarized in Table 3.2.

3.4 Numerical investigation

The developed model is applied for several loading cases to qualitatively show

that it can reproduce the multi-axial/-path loading responses. The model is formulated

and the algorithm is developed using the return mapping method. A detailed algorithmic

approach of the model is described in the appendix. In order to see the effect of the

model for controlled loading, uniaxial stretch and multi-axial loading problems are

investigated before applying the model to the multi-path loading cases.

3.4.1 Material properties:

The material concerned in this paper is mild steel. The material properties for

DDQ steel are determined from the series of experiments to identify the anisotropy and

plastic spin parameters. Using the RIK theory, the parameters can be limited to a for

plastic spin, c1, c2, and bl for kinematic hardening, and κ for defining the scale of the

function k ( ϑ ) in (3.17). The procedure of identifying the parameters is explained in

Chapter 4. Zhao et al. [53, 54] and Chun et al. [7] introduced the procedures to identify

the kinematic hardening parameters by an inverse method. The material parameters used

to demonstrate the theory are summarized in Table 3.3.

Uniaxial tension test: The uniaxial tension test is done to identify the material parameters

for the initial anisotropy and for isotropic hardening. The computed flow stresses along

43
the RD, TD and 45° are compared with measured data in Figure 3.4. The lines represent

the computed results and the symbols for measured data. In contrast to the RD and TD,

the stress evolves higher than that of the experiment in 45° orientation when the latest

kinematic hardening model (ANK) is used, as shown in Figure 3.4(a). Although it is not

shown in this diagram, the isotropic hardening model gives similar results. However, by

using the rotational hardening scheme the flow stress in 45° shows a good correlation

with measured data. The RD and TD curves are not affected with the rotational

hardening, because those directions lie in the symmetry direction of the anisotropy.

The anisotropy of the material is measured by R-values, which may change

because the anisotropy is evolving. The R-values may be different with respect to the

longitudinal straining of the specimen, so defining the R-value measured at low strain as

the initial R-value. In contrast to that, the R-value up to 25% strain is defined as the

averaged R-value. In order to see the effect of the hardening models and anisotropy, it is

necessary to compute the 45° oriented tensile experiment with various combinations of

hardening models and R-values, as shown in Figure 3.5.

The resulting flow stresses of the two R-values are rather higher than the

experimental results when use the ANK model. If the rotational hardening scheme is

applied, the R-value is changing according to the evolution of the anisotropy, and the

flow curve is decreased. Without the correction of the backstress associated with the

rotation of the symmetry axes, the flow curve is converging to the RD or TD, and in this

case it is converged to the TD (the TD flow curve is omitted here to compare with 45°

clearly.) The difference between the experimental curve and the computed curve by

using rotational scheme only is the amount of the backstress evolution, which is

associated with the rotational hardening. Therefore, the hardening curve is fitted with the

experimental result by the suggesting RIK hardening model.


44
The rotation angle of the symmetry axes is shown in Figure 3.6 for 45°, and the

symmetry axes are not rotating in the RD and TD as mentioned earlier.

Offset test: The purpose of this test is to see the effect of anisotropy evolution in a

different loading direction, and see the offset stress as we choose the plastic spin

parameter a. The offset amount of stress due to the Bauschinger effect is measured from

the numerical simulation for a single element. The tests are done in three steps. To make

a pre-strained condition, the element is stretched in the RD direction up to a 7%

equivalent plastic strain, and then release from the loading. The second loading is in a

different path which has a certain angle from the first loading. Measure the equivalent

stress at 13% equivalent plastic strain for comparison. In this multi-step loading

condition, the stress in the reloading stage is dependent on the different evolution of the

second backstress component. The loading sequence is illustrated in Figure 3.7(a), and

the response of the offset for the reloading stage with respect to the loading angle is

shown in Figure 3.7(b). The stress offsets are normalized with respect to the offset at

180° (full reversal loading). The stress offset is responses to the rate of rotation of the

symmetry axes.

Cross and Bauschinger effect: The loading conditions shown in Figure 3.1were used to

demonstrate the ability of the RIK hardening model in capturing the response of the

multi-path loading. It is already discussed in the introduction that neither the isotropic

hardening model nor the isotropic + kinematic hardening model can reproduce the key

effects of the multi-axial loading: ‘crossing’ and ‘Bauschinger’ effects simultaneously.

The same model is run with the rotational hardening scheme with and without the

45
kinematic hardening model. With these results, the anisotropy axes rotation is in charge

of the “cross” effect, and the kinematic hardening is of the “Bauschinger” effect.

The tension shear tests with different pre-strains are simulated for 3%, 7%, and

10% strain. The flow stress curves are illustrated in Figure 3.10. At every pre-strain, the

curve crosses and approaches the simple shear curve. As larger pre-strain, the amount of

stress “cross” gets bigger. Dislocation structures are developed more in the direction of

the dislocation structure which is previously developed due to the rolling process, and

then the effect of resistance to the dislocation structures in a different directional loading

gets bigger [29]. More texture develops in the tensile direction as the tensile strain

increases, which results in higher jumps in the stress.

3.5 Conclusions

By combining the rotational hardening with the isotropic hardening and with the

Armstrong-Frederick type non-linear kinematic hardening, material behavior under the

multi-path strain loading problem is modeled within the phenomenological continuum

framework.

With the model it is possible to

- capture the flow stress in 45° in correlating with the experimental results,

- evolve the backstress differently according to the loading angle change,

and

- couple the anisotropy axes rotation with the backstress evolution.

The results show that the suggested model can reproduce the “cross effect” and

the “Bauschinger effect” clearly. The anisotropy evolution with the anisotropy axes

rotation due to the plastic spin is responsible for the “cross effect.” Even if the rotational

hardening can change the axes of orientation, it cannot change the yield surface shape

46
itself. The yield surface shape in this model can grow, move, and rotate. The model

could be better if we considered the yield surface shape evolution by relating the yield

surface to the microscopic approach.

47
Model Cross Bauschinger Permanent Offset
Isotropic No No No
Kinematic No Yes No
ANK No Yes Yes

Table 3.1: Capability of reproducing multi-axial/-path loadings characteristics of


conventional hardening models.

∂φ
dp = γ
∂τ
s s
y = ∑ q j ; q1 = c1 ( τ − y ) − sb1q1 ; q 2 = kc2 ( τ − y )
j β β
c
( ) ( )
σ y = σo + K 1 − e − Ns − 1 1 − e − b1s − c 2s
b1
(
ωp = µφ τ d p − d p τ )
eφα = θeφα

Table 3.2: Constitutive equations of RIK hardening model

48
Young’s Modulus, E 210 Gpa
Poisson’s Ratio, ν 0.3
Initial yield stress, τ 0 152.0 Mpa
Isotropic Hardening parameters; K, N 235.81 Mpa, 9.23
Initial anisotropy; R0,R45,R90 2.64, 1.47, 2.17
Backstress parameter; c1,b1,c2,κ 3.0GPa, 300.0, 70.0, 1.8
Plastic spin parameter, µ φ -57.0

Table 3.3: Material properties for DDQ.

49
σ12
∆c
∆b

εp

Figure 3.1: Schematic stress-strain behavior of an IF steel during a monotonic simple


shear test (case a), a tension-shear test (case b) and a reverse-shear test (case c).
Characteristic crossing (∆c) and Bauschinger effect (∆b) are shown according to the
loading path.

50
300

250

Shear stress, MPa |σ 12|


200 Tension-shear Simple
shear
150

100

50
Reverse-shear
0
0 0.04 0.08 0.12 0.16 0.2
Eqv. plastic strain
(a)
300

250 Simple shear


Shear stress, MPa |σ |
12

200 Tension-shear

150

100

50
Reverse-shear
0
0 0.05 0.1 0.15 0.2
Eqv. plastic strain
(b)

Figure 3.2: Multi-path loading response of conventional hardening models: (a) isotropic
hardening, (b) anisotropic nonlinear kinematic hardening model. Both models cannot
reproduce the ‘crossing’ effect while the Bauschinger effect is caputred in (b).

51
moved yield surface

rotated yield surface

expanded yield surface C


B
A

initial yield surface

α
backstress

Figure 3.3: Rotational evolution of yield function: The yield function is expanded to the
stress state C when only the isotropic hardening is considered. The yield surface is rotate
by using rotational hardening and the iterated stress state is B. The yield surface is
moved toward C with the amount of backstress by the kinematic hardening.

52
500

400

True stress (MPa)


300

EXP ANK
200 RD
o
45
TD

100
0 0.05 0.1 0.15 0.2 0.25
Eqv. Plastic strain
(a)
500

400
True stress (MPa)

300

EXP RIK
200 RD
o
45
TD

100
0 0.05 0.1 0.15 0.2 0.25
Eqv. Plastic strain
(b)

Figure 3.4: Flow stress in the RD, 45°, and the TD orientation with a) ANK, b) RIK.
ANK predicts higher flow stress for 45° while RIK correlates the measurement data.

53
500
(R )
45 init
(R )
400 45 av g
True stress (MPa)

300 RIK

Rotation + Isotropic

200
EXP

100
0 0.05 0.1 0.15 0.2 0.25
Eqv. Plastic strain

Figure 3.5: Flow curves for the 45° specimen: isotropic+kinematic hardening model
(ANK) produces rather higher stress ( ( R 45 )avg , ( R 45 )init ), rotation+isotropic hardening
model produces lower stress, while the RIK follows the measured data rather closely.

54
0

-10
Rotation angle (η)

-20

-30

-40

-50
0 0.05 0.1 0.15 0.2 0.25
Eqv. Plastic strain

Figure 3.6: Rotation angle (degree) of the symmetry axes for 45° oriented specimen.

55
(a)

1.2

1
a=-100
0.8
Normalized stress

0.6

0.4
a=-350
0.2

-0.2
0 60 120 180 240 300 360
Loading angle
(b)

Figure 3.7: Stress offset tests for loading-unloading-reloading in different path: (a)
loading condition – give different loading path to the pre-stretched specimen, (b) stress
offset with respect to the plastic spin parameter. The stress is normalized with respect to
the offset of 180º ( ∆τϕ / ∆τ180o )

56
300

250

Shear stress (MPa) |σ 12|


Reverse-shear
Tension-shear
200

150
Simple shear

100

50

0
0 0.05 0.1 0.15 0.2
Accum. Plastic strain

Figure 3.8: Flow stress response when using the rotational hardening with isotropic
hardening scheme. Stress jump (crossing) is shown but Bauschinger effect is not captured.

300

250
Shear stress (MPa) |σ 12 |

Tension-shear
Simple shear
200

150

Reverse-shear
100

50

0
0 0.05 0.1 0.15 0.2
Accum. Plastic strain

Figure 3.9: Flow stress result of the multi-path loading when the RIK hardening model is
used. Both ‘cross’ and ‘Bauschinger’ effects are captured.

57
300

250
Shear stess (MPa)

200

150

100 3%
7%
10%
50

0
0 0.05 0.1 0.15 0.2 0.25
Accm. Eqv. Plastic strani

Figure 3.10: Multi-path loading responses show the flow stress crossing after various
pre-strains; more pre-strain yields higher jump in stress.

58
CHAPTER 4

EXPERIMENTAL OBSERVATION OF ANISOTROPY EVOLUTION IN SHEET


METAL AND IDENTIFYING THE MATERIAL PROPERTIES

Anisotropy of materials evolves during plastic deformation by micro-mechanical

development of texture and dislocation structures. Phenomenologically, the anisotropy is

characterized by different strain hardening responses for different specimen orientations.

Specimen orientation is specified with respect to the rolling direction (RD). The

relationship of the stress evolution in different orientations is measured by R-values. The

R-value is defined as the ratio of the transverse strain versus thickness strain at a certain

elongation, and the R-values are computed using the slope of the transverse strain versus

longitudinal strain curves. The R-values are measured from the experiment, but because

of the evolution of anisotropy, the hardening behavior may not be predictable for

orientations that are not in the RD or TD. With the rotational evolution of the anisotropy,

the slope of the strain-strain curve changes continuously. Anisotropy evolution for large

strains can be observed in simple uniaxial tension tests when measuring the transverse

and longitudinal strain continuously with the Digital Image Correlation (DIC) method.

To model the experimental response, a Rotational-Isotropic-Kinematic (RIK) hardening

model is investigated. With the RIK hardening model and its ability to represent the

rotational evolution of the anisotropy, the hardening behavior is predictable for non-RD

59
and non-TD directions. Methods to identify the plastic spin parameter and kinematic

hardening correctors are also discussed.

4.1 Introduction

In sheet metal forming and springback processes, incorporating evolution of

material anisotropy is important for making accurate predictions. Cold rolled sheet

metals have initial anisotropy which is incurred during the rolling process. The

anisotropy causes different flow stress with respect to the orientation angle that is

measured from the rolling direction (RD). The different flow stresses have significant

affects on the forming process (e.g., different earing of edges, punch force-displacements,

different springback, etc.). In order to address material anisotropy, anisotropic yield

functions such as those of Hill [6, 55] and Barlat [11, 56] have been developed. Hill’s

quadratic yield function [6] is known to be appropriate for BCC materials such as mild

steel, while Barlat’s yield function [11] is good for FCC materials such as aluminum

alloys. However, these yield functions do not account for the evolution of material

anisotropy during the deformation.

The evolution of the anisotropy of materials has been observed by both micro-

mechanical and macro-mechanical experiments [4, 5, 28, 32]. At a micro-mechanics

level, the anisotropy evolution is generally considered to be the result of texture

development which changes the preferred orientation of a grain aggregate. Using micro-

mechanical approaches, the evolution of anisotropy is observed and modeled by Agnew

[57], Asaro [31], Beaudoin [58], Kocks [59] and Nakamachi [60]. On the other hand,

Boehler [5], Kim and Yin [4], Kuroda [13], Dafalias [14] and Han et al. [15] have

observed and modeled anisotropy evolution using macro-mechanical approaches that rely

on phenomenological descriptions.

60
For sheet metals, R-values are provided to represent the planar anisotropy and to

compare the anisotropy for different materials [2]. The R-value is defined as the ratio of

transverse strain to thickness strain at a certain longitudinal strain. Usually R-values are

provided with the longitudinal strain at which the transverse strains are measured because

the measured R-values are different with respect to the longitudinal strain.

Even if the R-values are used to define the anisotropic relationship between a

certain direction and the RD and the experimental hardening curve for the RD is used

with Hill’s 48 anisotropic yield function, the computed hardening curve may not properly

model the experiment results for orientations not in the RD or TD. The computed

hardening curves in various directions are compared with the experimental results in

Figure 4.1. The curves of the RD and TD correlate quite well with the experimental

results. However, the computed 45° orientation curve overestimates the experimentally

observed result. The computed curves were generated by ABAQUS standard using the

Hill’s quadratic yield function and isotropic hardening model. The experimental results

are provided from the Numisheet 2002 benchmark problem set [61]. The material was

deep drawing quality (DDQ) mild steel.

The convenience of using an in-plane stress model is evident; however, Hill’s 48

yield criterion has major drawbacks in reproducing flow stress curves that depend on the

loading path. In other words, the dependence of the yield stress on direction is poorly

predicted by the theory [2]. These shortcomings are usually interpreted to result from the

evolution of anisotropy during the deformation. To address the anisotropy evolution, a

rotational hardening scheme is introduced. In order to measure anisotropy evolution,

tensile tests for RD, TD and 45° specimens were repeated on different materials. The

digital image correlation (DIC) method was used to measure the strain continuously up to

large strains.
61
The discussion will focus on the hardening model which can properly describe the

flow stresses. The R-values are effective for characterizing the RD and TD, however, it

is not capable of predictive modeling for the 45° orientation. It is assumed that the

unexpected higher estimation of the stress evolution in 45° in Figure 4.1 is caused by the

evolution of the anisotropy and that the R values can be used to determine the ‘initial’

anisotropy of the material.

Understanding the evolution of anisotropy is critical for predicting material

behavior in large strain deformation. Furthermore, it is essential to be capable of

representing the anisotropy evolution in multi-axial and multi-step loading. In order to

model the multi-axial elasto-plastic deformation of planar anisotropic material, an RIK

(Rotational-Isotropic-Kinematic) hardening model is suggested.

A simple method to measure the rotation of the symmetry axes is suggested with

the assumption that the anisotropy evolution is dominant to the rotation of the symmetry

axes of anisotropy. The method requires tensile experiments with continuous

measurement of R-values for specimens cut in 45° to the RD. According to the theory,

the symmetry axes do not rotate for the deformation along the RD or TD. To measure the

strain continuously up to large strains, the Digital Image Correlation (DIC) method is

used. The DIC method was developed by Sutton et al. [62] and improved by Vendroux

and Knauss [63], although it lacks accuracy at small strains, it can be used at large strains

and is simple to implement. Although Rao et al. [64] measured large strains using digital

images of grid marks on uniaxial sheet metal specimens, the grid deformation was not

used to determine R-values. The authors are not aware of any other investigations that

use the DIC method to determine R-values.

62
This paper presents a DIC-based experimental method for determining anisotropy

evolution and demonstrates how the material parameters for the RIK hardening model

can be determined from simple tensile experiments.

4.2 Definition of R-value

The definition of the R-value for a planar anisotropic material is the ratio of

transverse strain to thickness strain and is therefore, given as follows:

εw
R= . (4.1)
εt

Because the specimen thickness is generally very small, the thickness strain ε t for sheet

metal is very difficult to measure accurately. Therefore, the constant volume condition in

plastic deformation is used and the Eqn. (4.1) can be rewritten as

εw εw
R= =− , (4.2)
εt ( εl + ε w )

which is widely used to measure R-values. Then the R-values are defined by a linear

fitting of the slope of the ε w − εl (transverse strain versus longitudinal strain) curve in

Figure 4.2.

The R-value from the ε w − ε l for the RD is R90, for 45° is R45, and for the TD is

R0, respectively. The R-values are shown in Figure 4.3. The similar characteristics of

the R-values had been measured on a cube of copper with a rolling texture as shown in

Figure 4.4 by Stout and Kocks [59].

As assuming as the anisotropy axes are not rotating when the loading is along the

RD and TD, the R0 and R90 are computed by linear fit through whole strain. Anisotropy

63
evolution is significant in 45°, so the strain curve requires non-linear function to be fitted.

The function used in this research is

ε w = {a ln ( εl ) + b} εl . (4.3)

4.3 Experimental set up

Experiments were performed on an MTS Model 810 servo-hydraulic load frame

with an Instron 8500 digital controller. The controller was operated in displacement

control mode. Load was measured from a 25 MT load cell and was subsequently

converted to true stress. For some experiments, strain gages were used to assure the

accuracy of the strain measurement by the DIC. The strain gages were connected to a

Measurements Group 2100 strain gage conditioner. The strain conditioner output and the

load signal from the Instron controller were recorded by with National Instruments SCXI

hardware and LabView software at a rate of 20 points per second.

Specimens used in the experiments were cut from sheet metal with an angle to the

RD direction of 0°, 45°, and 90°. Specimens were 1 inch wide and 0.039 inch (1mm)

thick. Since the field of view of the CCD camera is fixed and only the lower grip of the

load frame moves, the region of interest has a significant downward displacement. The

size of the specimen must be carefully chosen to ensure that enough of the originally

undeformed image remains inside the camera’s field of view for all subsequent

deformation. For this reason, 1.5 inch gage length specimens were used. FEM

simulations were performed to confirm that the region where strains were measured were

far enough away from grips. The results of these FEM simulations are shown in Figure

4.5. By comparing strain-strain curves of four points, along the specimen centerline

64
(shown in Figure 4.5a), no significant differences were observed in the strains values

shown in Figure 4.5b. Hence, the clamping effect can be ignored.

Large strains were measured using digital image correlation (DIC). DIC is a

technique that compares digital images of a specimen surface before and after

deformation to deduce its two-dimensional surface displacement field and strains [63].

The images required for the DIC method were acquired by a Pulnix TMC-9701 digital

CCD camera with Nikon AF micro Nikkor 105mm lens and National Instruments PCI-

1424 frame grabber. TIFF images that were 768 by 472 pixels were written to a PC at a

continuous rate of 4 frames/sec. Because of the large amount of PC bus activity and the

large numbers of image files, separate computers are used for DIC image storage and

load and strain gage data acquisition. To minimize intensity change during the

experiment, directed light source (KL1500, SCHOTT) was used to illuminate the

specimen. A schematic of the experiment set up is shown in Figure 4.6.

The DIC method works best when the starting image contains a random pattern

that can carrys the specimen deformation exactly. To create such a random pattern on the

current specimens, spray paint was used. First white paint provided a bright background

and then black spray paint gave a high contrast speckle pattern. The black paint was

applied by spraying parallel to the specimen surface and letting paint droplets fall

randomly on the specimen. Examples of the random pattern before and after deformation

are shown in Figure 4.7.

The DIC program was set to measure incremental transverse and longitudinal

strain between subsequent images. The DIC program was first used in an interactive

mode to search good correlation parameters. For random patterns such as the one shown

in Figure 4.7, the best subset size was 100 x 100 pixels which corresponded to 4.2 x

4.2mm. The displacement was measured at 25 equally spaced points in the subset. The
65
maximum search range was set to 40 pixels in the longitudinal and transversal directions,

and the tolerance for optimization is set to 1.0x10-7. Subsequently, using the same

correlation parameters, a batch mode was used to process all the data for a particular

experiment. Since the subset used for correlation at the beginning of the loading has

undergone significant translation by the end the deformation, it was necessary to devise

an algorithm to track the approximate position of each subset used for correlation. Total

strain was obtained by adding the incremental strains. In order to reduce measurement

error, the correlation was performed on all 25 points in the subset, and then strains values

there were beyond one standard deviation were eliminated. The average strain was then

determined from the remaining strain values. The original DIC program was written by

Vendroux and Knauss in FORTRAN and C [63]. The modifications to track the overall

displacement of the subsets and to statistically treat the strain measurements were done

by the authors. The load is measured from the load cell, those strain data and the load

data are synchronized by the time.

The accuracy of the DIC method for measuring strain is demonstrated by

comparison with a strain gage for lower strains and with the data provided by

NUMISHEET 2002 for higher strains, and the results are shown in Figure 4.8. The strain

measurement with DIC has poor resolution at lower strains but shows quite good

correlation with the NUMISHEET 2002 data at large strain. Usually it is difficult to

measure large strains with a strain gauge, but DIC can measure 25% to 30% strain with

relative ease. The limitation at strains greater than 30% is related to the loss of adherence

of the applied speckle pattern to the specimen surface.

66
4.4 Hardening models

The hardening model used to fit the experimental curve in this paper is the RIK

(Rotational-Isotropic-Kinematic) hardening model. Isotropic hardening is defined

following Chun et al. [7] which is modified from the Chaboche model [39] in order to

better reproduce the tension test stress-strain curve. The yield function is defined as

follows:

c1
( )
σ y = σ0 + K 1 − e Ns −
b1
( )
1 − e b1s − c 2s , (4.4)

where σ0 is the initial yield stress, K, N are the fitting parameters of the hardening curve

measured in the RD, and c1, c2, b1 are the parameters for kinematic hardening, and s is the

effective plastic strain.

The corresponding Armstrong-Frederick type kinematic hardening [34] model

with permanent softening is defined with different backstress evolution equations as

c1
q1∇ = ( τ − y ) s − b1sq1 , (4.5)
β
c
q∇2 = k 2 ( τ − y ) s , and (4.6)
β
1 + κϑ for initial loading
k ( ϕ, ϑ) =  (4.7)
 κϑ for reversal loading

where ϑ is defined the angle of the direction of the symmetry axes and direction of

straining as shown in Figure 1.2.

The rotational hardening is defined by the rotation of the yield function φ whose

basis functions eiφ are assumed to lay on the symmetry axes of anisotropy. The rotation

of the symmetry axes is defined with the constitutive spin θφ which is defined from

material spin w p and the plastic spin ω φp as

67
θφ = w p − ω φp , and (4.8)
φ φ φ
e =θ e
i i (4.9)

The plastic spin in (4.8) is defined as follows and is due to the non-coaxiality of the stress

and the plastic strain increment directions ([14], [15], [46]):

ωφp = µφ ( τε p − ε p τ ) . (4.10)

The plastic spin parameter µφ in (4.10) is defined through the angle ϑ [15, 46],

a
µφ = tan ( ϑ ) , (4.11)
σy

and a is the only parameter to be defined for the plastic spin. The parameter can be

identified with the results of the transverse strain vs. longitudinal strain curve.

The non-linear kinematic hardening parameters c1,b1,c2 in Equations (4.5) and

(4.6) are identified using the inverse method which optimize the parameters from the

results of the three point bend experiments performed by Chun et al. [7]. If the rotational

hardening is ignored, a and κ are zero, then the RIK hardening model returns to ANK

model. In addition, the hardening model turns to Chaboche model if c2 becomes zero,

and the model becomes an isotropic hardening model if c1,b1 are set to zero.

4.5 Experimental results and material parameters

The experimental results of the materials are transverse strain versus longitudinal

strain curves to measure the anisotropy values (R-values) and flow stress curves in the

RD, TD and 45° directions. In order to show the dependence on material parameters, the

procedure to define parameters will be explained with DDQ steel data provided from

both the current experiments and NUMISHEET 2002. The experimental results of others

68
are shown with theoretically fitted curves and the parameters of the materials will be

summarized.

4.5.1 DDQ (Experiment with DIC method)

The sheet metal of DDQ used in the experiments is the same material used in

NUMISHEET 2002 benchmark problems [61]. The measured transverse strain versus

longitudinal strain curves for the RD, TD and 45° are shown in Figure 4.9(a), and the

corresponding flow curves are shown in Figure 4.9(b). Also shown in Figure 4.9 are

simulation results of initial anisotropy without rotational hardening. When the rotational

evolution of the anisotropy is not considered, the strain-strain relation for the 45°

orientation is higher than the experimental curve. The estimated flow curve for the 45°

orientation differs from the experimental results after approximately 5% strain. The ratio

of stress evolution between 0% and 5% strain correlates very well with the experiment.

For larger deformations, one can actually get better results when using the averaged R45

value which is computed from the linear fit of the strain-strain from 0% to 20% strain.

The comparison is in Figure 4.10. However, even though the stress level is much closer

to the experimental results than the estimation by using initial anisotropy values in Figure

4.9(b), the rate of stress evolution does not correlated with the experimental results for

the whole stain range.

It is hard to satisfy both strain-strain and flow stress relations simultaneously

because the anisotropy evolution in 45° orientation is not accounted in the conventional

hardening models. Following the assumption, the anisotropy evolution can be seen for

only 45° orientation, so consider R45 related to the flow stress. To model the R45

evolution, as shown in Figure 4.11(a), only the rotational hardening scheme can change

the evolution rate according to the deformation by using plastic spin theory. The
69
corresponding flow stresses for the 45° orientation then evolved with an abrupt change in

stress as the symmetry axes are rotate due to the plastic spin. The evolution occurs

because the basis of the yield function is laid on the symmetry axes of anisotropy. The

results for the flow stress are shown in Figure 4.11(b). The RD and TD curves are

omitted because there is no rotation in those two directions and the predictions in these

directions are identical to those without rotational hardening. The plastic parameter a = -

155 is selected as the computed R45 curve that best correlates to the experimental curve.

The difference in flow stress curves between the experimental and computed

results is then corrected by kinematic hardening (4.7). As shown in Figure 4.12(a) and

(b), kinematic hardening has corrected the flow stress curves but did not change the slope

of the strain-strain curves. The parameters for this modeling are summarized in Table 4.1.

4.5.2 DDQ (NUMISHEET 2002)

The flow stress curves are shown in Figure 4.1 and the transverse strain versus

longitudinal strain curves are shown in Figure 4.2. As was shown in Figure 4.1, the

hardening evolution for the 45° orientation is overestimated with average R value. The

simulation results with RIK hardening are shown in Figure 4.13 and are compared with

the experimental results. For plastic spin parameter, R45 evolving curve is shown in

Figure 4.13a by comparing the FE results to experiments.

4.5.3 DQ

The drawing quality (DQ) mild steel was also tested. Following the same

procedure that was described earlier for DDQ, the final fitted curves are shown in Figure

4.14.

70
4.5.4 HSS

High strength steel (HSS) was fitted to the RIK hardening model. In contrast to

mild steels, HSS does not show much rotation. Experimental results are compared with

the computational results in Figure 4.15.

As summarized in Table 4.1, when comparing R-values for all directions, only for

HSS the value of R45 is higher than other values. Even with high plastic spin factors, for

high R45 values, the anisotropy axes are not rotating toward the straining direction for the

45° orientation. In addition to that, the flow curves are accurately modeled with initial R

values chosen from the transverse strain versus longitudinal strain curve. Therefore, it

may be concluded that the rotational hardening in HSS is very small. The anisotropy of

this material is not effectively changed by the strains and deformation in these tests.

Therefore, it may be assumed that the texture of this material is not severely affected by

the deformation of these experiments. To verify this claim, the symmetry axes would

have to be determined by texture analysis with the orientation distribution function

(ODF).

4.6 Discussion

R-values are used to compare the anisotropy of materials in a relative manner.

The R-values measured in tensile experiments are not useful for predicting the rotational

evolution of anisotropy. On the other hand, R and ∆R which are defined as

R 0 + 2R 45 + R 90 R − 2R 45 + R 90
R= , ∆R = 0 , (4.12)
4 2

are used to judge the formability ( R ) and the earing pattern ( ∆R ) in circular cup

drawing. Formability increases as R increases. For mild steel with the rotational

hardening, the yield function shape evolution in σ1 , σ 2 space looks as if the R increases

71
with isotropic expansion of the yield locus. Therefore, it could be said that models with

rotational hardening can describe the formability better than non-rotational hardening

models.

The plastic spin in sheet metals defines the planar rotation of the symmetry axes

of anisotropy, and ∆R is known to be related to the planar anisotropy. Therefore, there

must be a relationship between the parameters of the plastic spin and the planar

anisotropy ∆R. The plastic spin and planar anisotropy values for the tested materials are

shown in the Table 4.1. As shown in Figure 4.16, the plastic spin is linearly related to

∆R. Detailed experimental determination of the rotation of the symmetry axes of

anisotropy is required to verify these results. Nonetheless, in Figure 4.16 the comparison

between the plastic spin parameter in Han et al. [15] using the measured data of the

rotation by Kim and Yin [4] shows excellent correlation. Therefore, the unknown plastic

spin parameter can be identified from the R-values of mild steel. However, more

experiments are required to verify this methodology for other alloys and crystal structures.

4.7 Concluding remarks

For two mild steels and high strength steel sheet metal, experimental results of

conventional flow stress curves and transverse versus longitudinal strain curves for

various orientations to the RD have been provided. Using the assumption that anisotropy

axes rotation is the dominant reason for the evolution of anisotropy, the material

parameters for plastic spin and for correcting kinematic hardening can be identified from

the simple tensile tests. With the help of rotational evolution of the anisotropy, the RIK

model can capture the material plastic hardening behavior with Hill’s quadratic yield

criteria.

72
Using the DIC method, it is relatively easy to measure large strains. The DIC

method was used to generate plots of transverse strain versus longitudinal strain for

simple tension experiments. The evolution of anisotropy is then correlated with the slope

of the transverse strain versus longitudinal strain curve. Only the plastic spin parameter

can change the slope of the computed transverse strain versus longitudinal strain curve.

The flow stress results are correctable with the kinematic hardening parameter. The

parameters for the RIK model are defined by three-point-bend-test [53, 65] and tensile

tests for three directions: RD, TD, and 45°. For mild steel, a linear relationship between

the planar anisotropy ∆R and the plastic spin parameter a has been found.

73
Material Elastic Anisotropy Isotropic Kinematic Rotational
R0 = 2.137 σ0 =152.22MPa c1 =3.0GPa
DDQ E=210GPa a = -155
R45 = 0.935 K =222.01MPa b1 =300
(Exp.) ν=0.3 κ = 50.0
R90 = 1.508 N =-7.87 c2 = 70.0MPa
R0 = 2.722 σ0 = 152.0MPa c1 = 3.0GPa
DDQ E=210GPa a = -57
R45 = 1.474 K = 235.81MPa b1 = 300
(NS2002) ν=0.3 κ = 1.8
R90 = 2.169 N = 9.23 c2 = 70.0MPa
R0 = 1.60 σ0 = 198.0MPa c1 = 3.3GPa
E=180GPa a = -100
DQ R45 = 1.010 K = 242.47MPa b1 = 220
ν=0.3 κ = 1.5
R90 = 1.46 N = 9.95 c2 = 103MPa
R0 = 0.832 σ0 = 332.22MPa c1 = 3.0GPa
E=210GPa a=0
HSS R45 = 1.185 K = 430.55MPa b1 = 150
ν=0.3 κ=0
R90 = 0.560 N = 1.7 c2 = 80MPa

Table 4.1: Material properties of RIK hardening model for DDQ, DQ and HSS.

74
400

350

300
True stress (MPa)

250

200
EXP. FEM
RD
150 45
o

TD
100
0 0.05 0.1 0.15 0.2 0.25
Eqv. Plastic strain

Figure 4.1: Flow stresses of tensile loading at various orientations compared to


measuement data of DDQ from Numisheet 2002 [61]. ABAQUS standard was used to
compute the flow stresses.

75
0

-0.05
o
45
Transverse strain

-0.1

-0.15 TD
RD

-0.2
0 0.05 0.1 0.15 0.2 0.25
Longitudinal strain

Figure 4.2: Transverse strain versus longitudinal strain of DDQ. The slope of each curve
is used to compute the R-values. (Numisheet 2002 [61])

76
3

2.5 R
0

2 R
90

1.5 R
R

45

0.5

0
0 0.05 0.1 0.15 0.2 0.25
Longitudinal strain

Figure 4.3: Evolution of R-values: R0 and R90 are not evolving compared to the R45.

77
Figure 4.4: R-value measurements for rolled copper in compression experiments.
Particularly R45 shows evolution.[59]

78
1 2 3 4

(a)

0.2
1
2
0.15
Transverse strain

3
0.1 4

0.05

0
0 0.05 0.1 0.15 0.2 0.25 0.3
Longitudinal strain
(b)

Figure 4.5: FEM validation of specimen dimensions to avoid clamping effect: a) strain
measurement locations, b) computed strains. The strains are closer enought to show that
the measured area is safe from the clamping effect.

79
LIGHT
SOURCE
MTS Controller

CCD
CAMERA

STRAIN GAUGE
CONDITIONER PC
STORE STRAIN
AND LOAD

PC
STORE IMAGE DATA A/D
CONVERTER

MTS

Figure 4.6: Schematic diagram of the experimental set up using DIC(Digital Image
Correlation).

80
(a)

(b)

Figure 4.7: The random pattern sprayed on the specimen and captured by CCD camera:
a) before and b) after deformation. The small rectangles highlight a 100 x 100 pixels
subset of the image that has been deformed.

81
300

250

200 Longitudinal

Stress (MPa) 150

100
Transverse

50 Comparison in
small strain
0
-0.01 -0.005 0 0.005 0.01 0.015
Strain
(a)

-0.05
Transverse strain

-0.1

-0.15

RD (DIC)
-0.2
RD (NS 2002)

-0.25
0 0.05 0.1 0.15 0.2 0.25 0.3
Longitudinal strain
(b)

Figure 4.8: a) Comparison of stress-strain measurement by DIC(symbols) and strain


gauge (lines) for a small strain, b) comparison of strain measurement by DIC (symbols)
and NUMISHEET 2002 data (lines) for large strain.

82
0

-0.05

Transverse strain
-0.1

EXP. FEM
-0.15
RD
o
45
TD
-0.2
0 0.05 0.1 0.15 0.2 0.25
Longitudinal strain
(a)
500

400
True stress (MPa)

300

EXP. FEM
200 RD
o
45
TD
100
0 0.05 0.1 0.15 0.2 0.25
Eqv. Plastic strain
(b)

Figure 4.9: DDQ experiments:a) Transverse-strain versus Longitudinal strain, b) flow


stress curves for each orientation. Stress prediction of 45° orientation is overestimated at
large strain over 5% strain.

83
0

-0.05

Transverse strain
-0.1

EXP. FEM
-0.15
RD
o
45
TD
-0.2
0 0.05 0.1 0.15 0.2 0.25
Longitudinal strain
(a)
500

400
True stress (MPa)

300

EXP. FEM
200
RD
o
45
TD
100
0 0.05 0.1 0.15 0.2 0.25
Eqv. Plastic strain
(b)

Figure 4.10: Simulation results of averaged R value for 45° orientation comparing with
the experimental results: a) Transverse strain versus longitudinal strain, b) flow stresses.
Prediction of 45° orientation is underestimated.

84
2
DDQ(DIC)

1.5
a = -250
a = -155
a = -50

45
1

R
a=0

EXP
0.5

0
0 0.05 0.1 0.15 0.2 0.25
Longitudinal strain
(a)
500

a=0
400 EXP
True stress (MPa)

a = -50
300

200
a = -155
a = -250
100

DDQ(DIC)
0
0 0.05 0.1 0.15 0.2 0.25
Eqv. plastic strain
(b)

Figure 4.11: Sensitivity of the plastic spin parameter: a) R45 evolution, b) flow stresses of
45° orientation according to the plastic spin parameter. The gap between the
measurement and computed flow stress is the amount of backstress evolution.

85
0

-0.05

Transverse strain
-0.1

EXP. FEM
-0.15
RD
o
45
TD
-0.2
0 0.05 0.1 0.15 0.2 0.25
Longitudinal strain
(a)
500

400
True stress (MPa)

300

EXP. RIK
200
RD
o
45
TD
100
0 0.05 0.1 0.15 0.2 0.25
Eqv. Plastic strain
(b)

Figure 4.12: Simulation with the RIK hardening model and comparison with
experimental results: a) Transverse-strain versus longitudinal strain, b) flow stresses.
Prediction for 45° orientation captures the experiment well.

86
0

-0.05

Transverse strain
-0.1

EXP. FEM
-0.15 RD
o
45
TD
-0.2
0 0.05 0.1 0.15 0.2 0.25
Longitudinal strain
(a)
500

400
True stress (MPa)

300

EXP. RIK
200 RD
o
45
TD
100
0 0.05 0.1 0.15 0.2 0.25
Eqv. Plastic strain
(b)

(continued)
Figure 4.13: DDQ- data provided by NUMISHEET 2002. Computed by RIK hardening
model comparing with experimental results: a) Transverse strain versus longitudinal
strain, b) flow stresses, c) R45 evolution.

87
(Figure 4.13 continued)

1.5

1
45
R

Exp.
0.5
FEA

0
0 0.05 0.1 0.15 0.2 0.25 0.3
Longitudinal strain

88
0

Transverse strain
-0.05

-0.1
EXP. FEM
RD
o
45
TD
-0.15
0 0.05 0.1 0.15 0.2 0.25
Longitudinal strain
(a)
500

450
True stress (MPa)

400

350

300
EXP. RIK
RD
250 o
45
TD
200
0 0.05 0.1 0.15 0.2 0.25
Eqv. Plastic strain
(b)

(continued)
Figure 4.14: DQ – experiment. Computed by RIK hardening model comparing with
experimental results: a) Transverse strain versus longitudinal strain, b) Flow stresses, c)
R45 evolution.

89
(Figure 4.14 continued)

2
DQ (DIC)

1.5

1
45
R

0.5
Exp.
RIK

0
0 0.05 0.1 0.15 0.2 0.25
Longitudinal strain

90
0

-0.03

Transvers strain
-0.06

-0.09
EXP. FEM
RD
-0.12 o
45
TD
-0.15
0 0.05 0.1 0.15 0.2 0.25
Longitudinal strain
(a)
600

500
True stress (MPa)

400

EXP. RIK
300 RD
o
45
TD
200
0 0.05 0.1 0.15 0.2 0.25
Eqv. Plastic strain
(b)

Figure 4.15: HSS – experiment. Computed by RIK hardening model comparing with
experimental results: a) Transverse strain versus longitudinal strain, b) Flow stresses.

91
0
DDQ (NS)

-100
DQ (DIC)

DDQ (DIC)
-200
a

-300

Kim & Yin ('97)


-400
0 0.2 0.4 0.6 0.8 1 1.2
∆R

Figure 4.16: Linear relationship between the plastic spin parameter (a) and the planar
anisotropy (∆R). This relationship can apply for mild steel which has B.C.C. crystal
structure.

92
CHAPTER 5

MODELING MULTI-AXIAL ELASTO-PLASTIC DEFORMATIONS ON PLANAR


ANISOTROPIC MATERIALS, PART II: APPLICATIONS

In order to predict the deformations under multi-axial/-path loadings in a

phenomenological framework, a new RIK hardening model is suggested in Chapter 3.

The model combines isotropic, kinematic, and rotational hardening, which is motivated

from micro-mechanical behaviors such as developments of dislocation structures and

texture. Armstrong-Frederick type backstress components are used for the kinematic

hardening description. The plastic spin theory in literature is used to describe the rotation

of the symmetry axes of anisotropy, which is considered to be rotated along the texture

preferred direction.

The purpose of this chapter is to show the significance of the RIK hardening

model in the sheet metal forming process as well as springback prediction. With

rotational hardening and related kinematic hardening, the flow stress in each orientation

from the RD can be reproduced with a limited number of material parameters. Several

benchmark problems are considered to demonstrate the effect of the RIK hardening

model in comparison with other hardening models and experimental results. While the

deep drawing examples show the effect of the RIK model in the forming process,

simulation of pulling a sheet metal strip from a draw bead tool is used to demonstrate the

93
effect of the RIK model in springback. A typical form of hydroforming is explored using

various hardening models, and the differences are observed.

5.1 INTRODUCTION

With the help of polycrystalline plasticity theory and micro-mechanical

experiments, knowledge about the origin of the evolution of anisotropy has greatly

advanced. Some micro-mechanical developments in plastic deformations are dislocation

structures and texture [24, 29, 66-68]. The development of dislocation structures has

been considered to be related to isotropic and kinematic hardening, while the texture

development is related to the rotation of the symmetry axes of anisotropy [24, 36, 66].

Isotropic and kinematic hardening models have been developed so well that they are able

to capture the permanent offset of cyclic loading for some materials [7, 69, 70]. However,

the reorientation of the texture has not been incorporated into a phenomenological

framework. The reorientation of the preferred direction of textures is considered the

rotational evolution of anisotropy directions [4, 5, 14, 15, 67]. The RIK hardening model

is developed to translate the micro-mechanical behavior of texture reorientation and the

development of the dislocation structures into a macro-mechanical framework. The

formulation of the suggested RIK hardening model is presented in detail in Chapter 3.

Excluding the special experiments for measurement or academic purposes, the

material points are subjected to multi-axial or multi-path loading during deformation in

applied problems. The material behavior under those loadings has been described in

polycrystalline plasticity with complex material properties which are defined by many

experiments [25, 28, 29, 30, 71]. The polycrystalline approach requires massive

computation, so it has been hardly used in practical forming problems even though it may

yield a better accuracy than conventional phenomenological approaches. The RIK model

94
is capable of capturing multi-axial or multi-path loading by translating the key features of

the material behavior with a proper Bauschinger effect and rotational hardening, as

shown in [13].

In order to assess the capability of the RIK model in the application of industrial

problems, the results of some conventional sheet metal forming processes and spring

back problems will be presented and discussed in this paper.

5.2 RIK HARDENING MODEL

The suggested RIK hardening model is able to describe the growth of the yield

surface (isotropic hardening), its transition (kinematic hardening), and its rotation

(anisotropy axes rotation, via plastic spin). The yield function of the RIK hardening

model can be written as

( ) (
Φ τ, y, q;eiφ = φ ( τ − y ) ; eiφ + q , ) (5.1)

where q describes the isotropic hardening, y the backstress, and eiφ the orientation of the

( )
yield surface. Such a yield function φ τ, y; eiφ can represent Hill’s or Barlat’s yield

function for mild steels or aluminum alloys, respectively. Hill’s quadratic yield function

is considered here, which can be described as follows:

Φ = φ − σ 2y , (5.2)

where φ = (τ − y ) ⋅ K φ (τ − y ) . Therein K φ is the fourth-order tensor and initial values of

the matrix form P is

 1 −β12 0   τ11   y11 


P0 =  −β12 β22 0  , τ =  τ22  , y =  y 22  .
   (5.3)
 0 0 β66   τ12   y12 
95
The components of P0 can be related to the R-values by

R0 R (1 + R 90 ) ( R + R 90 )(1 + 2R 45 )
β12 = , β22 = 0 , β66 = 0 (5.4)
1+ R0 R 90 (1 + R 0 ) R 90 (1 + R 0 )

and the orientation of the tensor K φ is redefined in the deformation process by the basis

of eiφ which is defined with the plastic spin and the material spin. The reorientation can

be determined via

φ
K ijkl = ∑ K 0αβγδ ( eiφ ⋅ eα0 )( eφj ⋅ eβ0 )( eφk ⋅ e0γ )( eφl ⋅ eδ0 ) , (5.5)

where e0α are the initial anisotropy directions of the sheet.

The anisotropy axes eφα are related to the constitutive spin

eφα = θeφα (5.6)

and the constitutive spin is defined with material spin and plastic spin as

θ = w − ωp (5.7)

For the plastic spin the definition of Dafalias [14] and Kuroda [13, 18, 44] has been

modified as

a
ωp = tan ( ϑ ) ( τε p − ε p τ ) . (5.8)
φ

The isotropic hardening part of the yield function is described, as in Chun et al.

[7],

c1
(
σ y = σo + K 1 − e − Ns − ) b1
( )
1 − e − b1s − c 2s , (5.9)

96
with the equivalent plastic strain s, the initial yield stress σo and hardening coefficients K

and N, and other material parameters which are related to the kinematic hardening is

described below.

The Chaboche type kinematic hardening is defined with multiple backstress

components to capture the Bauschinger effect and permanent softening as

y = ∑qj , (5.10)
j

s
q1∇ = c1 ( τ − y ) − b1sq1 and (5.11)
β
s
q∇2 = kc 2 ( τ − y ) . (5.12)
β

The k in equation (5.12) is dependent on the loading direction and the angle of the

difference between the anisotropy axes direction and the plastic straining direction ( ϑ ),

respectively. The function is defined as

1 + κϑ initial loading
k= (5.13)
 κϑ reversal loading

5.3 MATERIAL PROPERTIES

Three material parameters are used in the RIK hardening model for the isotropic

hardening and three for kinematic hardening. In addition, the R values in three directions

are required to define the anisotropy yield function of Hill’s [6] for plane stress

conditions. The identification of these material parameters is clearly a non trivial problem.

A simple method to define the material properties is suggested in the experimental

observation and discussions in Chapter 4. The RIK hardening model can describe the

stress flow in an arbitrary angle to the RD, while a simple application of Hill’s quadratic

yield criterion is in general not capable to describe the stress evolution in the arbitrary

97
direction [2]. The advantage of the RIK hardening model is its capability to account for

the evolution of the anisotropy. The kinematic hardening parameters c1,c2 and b1 can be

identified from the three-point-bend test and an inverse method with a genetic algorithm

presented in Zhao and Lee [53, 54, 65] and modified by Chun et al. [7, 8].

If the rotational parameters (a, κ) are zeros, then the RIK hardening model

reduces to the non-linear anisotropic kinematic (ANK) hardening model proposed by

Chun et al. [7]. The computational investigations for the deep drawing and springback

processes were performed with respect to three hardening models: the isotropic, the ANK,

and the RIK hardening models. The materials used in this presentation were mild steel

DDQ, DQ and the material of NUMISHEET ’93. The material properties are

summarized in Table 1.

5.4 APPLICATIONS

5.4.1 Cylindrical cup drawing

In order to compare the simulation results with the experimental data, the

NUMISHEET 2002 benchmark problem [61] of a cylindrical cup drawing has been

considered to assess the behavior of the RIK hardening model on the sheet metal forming

process. The problem is commonly considered for an assessment of the planar anisotropy

via earing pattern along the outside of the flange after the forming process. The punch

force-displacement characteristics and the earing pattern are of specific interest in this

problem. The dimensions of the tool set are shown in Figure 5.1(a). The undeformed

mesh is shown in Figure 5.1(b) in which 900 reduced integrated Kirchhoff-type shell

elements are used, where the transversal shear strains are not explicitly contained in the

material formulation. A blank holding force of 70kN is applied to prevent wrinkling at

the flange. The considered material of the blank is a DDQ steel. Two friction

98
coefficients have been used in the simulations. A lower friction coefficient of 0.049 as

suggested in NUMISHEET benchmark problem is applied as well as a higher friction

coefficient of 0.12.

The simulation is performed with the ABAQUS/Explicit using explicit user

material subroutine VUMAT. A detailed algorithmic treatment can be found in Chapter

3. The punch forces are shown in Figure 5.2a for the isotropic, kinematic, and the RIK

hardening models. The differences in the punch forces are within the range of the

oscillations characteristic of the explicit integration scheme. However, the earing

patterns of the hardening models show many more differences, as can be seen in Figure

5.2b. The dots are the averaged values of the NUMISHEET benchmark experimental

results in both punch force-displacement and draw-in diagrams.

The earing pattern of the isotropic hardening and the ANK models are almost

identical and more draw-in is determined than that of the RIK hardening model. Also,

the differences of the earing pattern between the RD and TD are small in isotropic and

kinematic hardening in contrast to the big difference in the RIK model. In relation with

the experimental results, the RIK hardening model yields a better agreement than the

other two hardening models. The difference of the amount of the punch force and the

earing pattern are considered to be due to the friction coefficient which often have a

strong influence in the structural response of such problems. The punch force and the

earing patterns are closer to the results of the experiments. It is interesting to note that for

the friction effect on the earing pattern, the friction is more effective in the region of the

sheet of 0° to 45° to RD than in the region between 45° to 90° to RD. Therefore, a

smaller force and more draw-in are determined in the simulation than the experiment due

to the different friction coefficients.

99
Without the rotational hardening, the initial anisotropy is attached to the material

points and remains basically unchanged in this problem, while the orientation of the

anisotropy evolves with the deformation with rotational hardening. The stress

distribution plots are shown in Figure 5.3, showing the differences of the isotropic

hardening and the RIK hardening descriptions.

Severe differences in stress distribution at the diagonal direction and the rounded

corner between the side wall and flange can be found. It may be worth noting that there

are quite a few differences in stress distribution, although the punch force-displacement

are essentially the same. This can be explained by the rotation evolution of the symmetry

axes of anisotropy. The rotation angles of the anisotropy axes are shown in Figure 5.4,

and maximal rotations of about 45° are reached in the region of 43.8° and -40.8° to RD.

5.4.2 Square cup drawing

Another deep drawing example – the square cup of the NUMISHEET ’93

benchmark – is considered. The diagram of the tools is shown in Figure 5.5a and the FE

discretization is illustrated in Figure 5.5b. The tools were modeled with analytical rigid

surfaces and the blank sheet metal is descritized with 2500 reduced integrated shell

elements. A blank holding force 19.6kN and the friction coefficient of 0.15 are applied.

25 integration points in the thickness direction have been used. In order to see the effect

of multi-axial loading, major and minor strains and thickness distribution along the

diagonal directions are compared with respect to the hardening models in Figure 5.6. The

diagonal direction is chosen because a severe multi-axial loading is expected due to the

shape and complex material flow around the corner area. Significant differences of the

punch force-displacement are shown in Figure 5.7. The punch force prediction using the

RIK hardening model is 20% lower than the force with the isotropic hardening model.

100
As a result, the draw-in of the RD, TD, and diagonal direction are compared with the

averaged experimental results of the NUMISHEET ’93 proceedings as shown in Figure

5.8.

5.4.3 Pulling the strip through a draw bead toolset and springback

A draw bead simulation is discussed below to illustrate the effects of the

hardening models in springback deformation. The simulation is performed by an implicit

integration algorithm implemented in the standard ABAQUS user subroutine. The

geometry of the draw bead tool is shown in Figure 5.9. A strip is clamped between the

tools where a gap of twice the thickness between the dies, dg. The clamped strip is pulled

out for 165 mm while the tool remains fixed. In the last stage of this problem, the tools

are removed while the strip is fixed at the end.

The dimension of the specimen is 325mm in length, 25.5 mm in width and 0.9mm

in thickness. The elements are descritized 100 in length and 5 in width, respectively.

The width of the specimen is chosen that narrow in order to minimize the curvature

effects and to be as close to the plane stress condition as possible, to assess the twisting

mode after springback.

The results are shown in comparison to the isotropic hardening, the ANK

hardening, and the RIK hardening. Comparing the shapes of the 30° oriented specimens

after springback, differences in twisting modes as well as the heights can be seen. The

computations of the RD and the TD are ignored because there is no evolution of

anisotropy in the RD and the TD. The differences in the modes after springback are due

to the stress evolutions during the process. The ANK and RIK hardening models show

lower values in springback height, because the models have a permanent softening

scheme. The material experiences cyclic loading by bending and unbending while

101
passing through the draw bead tool. The stress histories of a material point are shown in

Figure 5.10 for the computations.

The shapes are different depending upon the hardening rules. The effect of the

hardening rules in the twisting mode is illustrated in Figure 5.11 for a 30° oriented

specimen.

The reason for developing different twisting modes is the stress distribution just

before springback. As shown in Figure 5.12, the stress distributions in the isotropic or

the ANK hardening model are anti-symmetric in both edges, while the RIK hardening

model develops symmetric stress distribution. The different stresses in both sides make

more twisting than symmetric stress distribution.

The clamping force and drawing force for each case are summarized in Table 5.2.

The forces for the initial clamping are increase when the rotational hardening scheme is

added. In contrast to the clamping force, the drawing force requires less force as the

rotational hardening scheme is used. With the numerical results, we can conclude the

rotational hardening in the RIK model has a key role in reproducing both the twisting

mode and permanent hardening effect.

5.4.4 Hydroforming

As one of the basic hydroforming processes, a T-tube expansion is frequently

discussed in the literature [72-74]. In the deformation process of the T-tube expansion,

material points in the expansion zone experience multi-axial loading situations because

material points have compression and tension in several directions as they move along

the tool surface. There are hoop and radial stresses due to the pressure inside the tube

and compression from both ends to feed the material. There have been many

investigations to optimize the pressure and the material feeding at both ends. Therefore,

102
the effect of the hardening models is discussed in this paper for the isotropic, the ANK,

and the RIK hardening models. The material used in this simulation is mild steel (DDQ).

The dimension of die set and loading conditions are shown in Figure 5.13. The tube is

modeled with 3500 shell elements. In order to see the difference of the expansion zone,

the top portion of the die is set to open. The pressure applied inside the tube is calculated

to slightly over the yield stress to make the tube expand. The movement of both ends to

feed the material at the center to make the expansion easier is chosen to prevent excessive

compressive stress at the center which may cause wrinkling.

The deformation height is not effective with the hardening model. The maximum

difference at the center of the dome is only 0.5% (45.54mm for isotropic hardening and

45.78mm for RIK hardening). However, the stress distribution shows much difference

with respect to the hardening models as shown in Figure 5.14. The highest stressed

region is shown at the connected area (marked with a circle) for isotropic and RIK

hardening. But the equivalent plastic strain is much larger in the ANK model, and the

region of the RIK model is smaller than other models (Figure 5.15). It is unique for the

RIK model that the plastic strain at the top of the dome is distributed in an elliptical shape

along the longitudinal axes (marked with an arrow). The rotational evolution of the

symmetry axes are shown in Figure 5.16 and the maximal rotation is 55° in accumulating

the rotation in loading and unloading. It should be noted that the region of severe change

in rotation is in the same area as the highest stress region.

The differences in evolving the stress yields significant differences in reaction

forces at both edges, while feeding the material toward the center of the tube in a certain

rate of displacement as shown in Figure 5.17.

103
5.5 CONCLUDING REMARKS

Three kinds of the forming process and one kind of the springback process are

investigated numerically. The rotational hardening scheme makes the material behavior

in hardening differently, and then the different behavior results in the springback. The

accuracy of the RIK hardening model in the forming process is shown at the draw-in of

the cylindrical cup. Only the RIK model can follow the hardening history around the

diagonal direction.

Severe deformation accompanying multi-axial/-path loading is simulated with the

square cup drawing problem. The corner of the cup is the area. The stress and strain

distributions in the area are different so much that the contour of the RIK model does not

resemble the other two hardening models. It is more useful to predict formability of the

hydroforming process, because the material points on the tube have experienced multi-

path/-axial strain and also large strain.

Those flow stress effects are also revealed in the springback. The springback of a

strip which is oriented a certain angle to the RD initially has a different shape with

respect to the amount of the symmetry axes rotation. The main effect is at the curvature

of the strip in the springback, and also the twisting mode which can be corrected only

with the rotational hardening.

104
Material Elastic Anisotropy Isotropic Kinematic Rotational
R0 = 2.137 σ0 =152.22MPa c1 =3.0GPa
DDQ E=210GPa a = -155
R45 = 0.935 K =222.01MPa b1 =300
(Exp.) ν=0.3 κ = 50.0
R90 =1.508 N =-7.87 c2 = 70.0MPa
R0 = 1.491 σ0 = 198.0MPa c1 = 3.3GPa
E=180GPa a = -100
DQ R45 = 1.01 K = 62.47MPa b1 = 220
ν=0.3 κ = 1.5
R90 = 1.180 N = 8.95 c2 = 103MPa
R0 = 1.790 σ0 = 154.0MPa c1 = 3.0GPa
E=206GPa a = -215
NS ‘93 R45 = 1.410 K = 253.1MPa b1 = 210
ν =0.3 κ = 5.0
R90 = 2.270 N = 9.78 c2 = 110MPa

Table 5.1: Material properties for RIK hardening model.

Clamping Force FC (N) Drawing Force FD (N)


Orientation 30° 45° 30° 45°
ISO 785.2/1006 807.9/1010.0 1173.0 1010.0
ANK 807.9/988.4 807.8/994.2 1130.0 994.2
RIK 886.0/1026.0 856.9/984.0 1105.0 984.0

Table 5.2: Clamping force and drawing force for different orientation and different
hardening for draw bead test.

105
(a)

(b)

Figure 5.1: Cylindrical cup drawing: a) Dimension of the tool set. (R1: punch
radius=50mm, R2: die radius=51.25mm, R3: punch fillet=9.5mm, R4: die fillet
radius=7.0mm, R0: blank radius=105mm<DDQ>), b) initial mesh

106
120
Experiment
100

Punch force (kN)


80 Isotropic
ANK
60 RIK
Higher friction
40

20

0
0 10 20 30 40
Punch displacement (mm)
(a)
96

94
Distance to the center (mm)

RIK
High Friction ANK
92

Experiment
90
RIK
88

86
Isotropic
84
0 10 20 30 40 50 60 o 70 80 90
Angle to the RD ( )
(b)

Figure 5.2: Response in cylindrical cup drawing: a) Punch force-displacement, b) earing


patterns with respect to the hardening models. The difference in punch force is not
severe while the earing pattern shows a lot of difference with respect to the hardening
models. Higher friction coefficient gives better results.

107
(a)

(b)

Figure 5.3: Cylindrical cup drawing: Residual von Mises stress distribution with respect
to (a) isotropic hardening, (b) RIK hardening

108
Figure 5.4: Rotation angle (in degree) of the symmetry axes of anisotropy.

109
(a)

(b)

Figure 5.5: Square cup drawing - Forming tool for the deep drawing of a square cup (a)
and the FE model (b).

110
0.3

0.2 ε RIK
1
ISO
0.1

Strain
0

-0.1

-0.2 ε ISO
2
RIK
-0.3
0 20 40 60 80 100
Distance from center (mm)
(a)
0.9

0.85
Thickness (mm)

0.8
RIK

0.75
ISO

0.7

0.65
0 20 40 60 80 100
Distance from center (mm)
(b)

Figure 5.6: Square cup drawing – Show the difference of a) Major and minor strain
distribution along the diagonal direction, b) thickness distribution along the diagonal
direction with respect to hardening models..

111
80
ISO
ANK
60
Punch force (kN)

40

RIK

20

0
0 10 20 30 40
Punch travel (mm)

Figure 5.7: Square cup drawing – Punch force-travel shows different history according to
the hardening models.

112
30

25

20
Draw-in (mm

ISO
15 RIK
EXP
10

0
DX DD DY

Figure 5.8: Square cup drawing – Draw-in with respect to hardening models compared to
the averaged experimental results of NUMISHEET ’93.

113
FC

99
16.2

dg R=6.5
FD
R=6.5

unit : mm

Figure 5.9: Dimension of the draw bead experiment tools

114
3

Isotropic
2
RIK
Normalized stress

ANK
0

-1

Clamp Draw
-2
0 0.5 1 1.5 2
Loading history

Figure 5.10: Stress history during the process for the 30° oriented specimens

115
140

Isotropic
120
ANK
100 RIK

Z-coordinate (mm)
80

60

40

20

0
-40 -20 0 20 40 60
Y-coordinate (mm)
(a)
200

150 Isotropic

ANK
Z-coordinate (mm)

100

50 RIK

-50

-100
0 50 100 150 200 250 300
X-coordinate (mm)
(b)

Figure 5.11: Springback shape of the 30° oriented specimen: (a) rear view, (b) side view

116
(a)

(b)

(c)
Figure 5.12: Von Mises stress distribution before springback: a) isotropic, b) ANK, c)
RIK hardening model. RIK shows symmetric distribution of stress while the other two
show anti-symmetric distributions which causes unexpected twisting model after
springback.
117
(a)
1
Pressure (Max: 20Pa)
0.8
Rate of loading

0.6

0.4

Axial feeding
0.2
(Max: 8mm)

0
0 0.2 0.4 0.6 0.8 1
Loading history
(b)

Figure 5.13: Hydroforming – Dimensions of the tool and initial tube and loading
conditions of the pressure and displacements at both ends of the tube.

118
(a)

(b)

(c)

Figure 5.14: Residual equivalent stress distribution: a) isotropic hardening, b) ANK


hardening, c) RIK hardening. More stress at curved zone with RIK hardening model
(marked area).
119
(a)

(b)

(c)

Figure 5.15: Equivalent plastic strain distribution: a) isotropic hardening, b) ANK


hardening, c) RIK hardening. The arrow points severely different area.

120
Figure 5.16: Rotation angle of the symmetry axes of anisotropy. Rotational evolution of
the anisotropy is concentrated at the highest stress area.

121
32
Axial Force (kN)

28

RIK
24

ISO
20 ANK

16
0 2 4 6 8
Axial feed (mm)

Figure 5.17: Reaction force at both edges when feed the material axially toward center in
the rate of displacement shown in Figure 5.13b.

122
CHAPTER 6

WORK HARDENING IN PRECIPITATE HARDENED MATERIALS

Plastic anisotropy of materials is decided by preferred movement of dislocations

in plastic loading. If dislocations move with the same degree of resistance in every

direction, then the material shows isotropic characteristics. If dislocations move with a

different degree of resistance in accordance with directions, then the material shows

anisotropic characteristics. The sources of the resistance of dislocation movement are

considered to be grain boundaries, precipitate, and dislocation walls, etc. In the sources

of material anisotropy, the effect of precipitates is investigated in this chapter.

The rotation of texture and relating backstress evolution has been discussed in

earlier chapters for mild steels which have a BCC crystal structure. In contrast to the

BCC material, aluminum alloys are known for the small rotation in texture. However,

some aluminum alloys show strong anisotropy and anisotropy evolution in the orientation

away from the RD due to the non-shearable precipitate inclusion in the material matrix.

Usually a higher-order anisotropic yield criterion such as Barlat ’96 [11] is used for an

aluminum alloy, and the hardening behavior is often considered to be isotropic.

Therefore, the model cannot predict anisotropic hardening for an arbitrary angle to the

RD, which is usually used to measure the representative hardening.

The goal of this research is to incorporate the precipitate influence in anisotropic

hardening using a nonlinear Armstrong-Frederick type kinematic hardening with the

123
fourth-order tensor that Barlat and Liu [10] proposed for the precipitate hardened material,

e.g., Al-Cu3%.

6.1 Introduction

Though a successful prediction of the hardening in crystal plasticity, it is still too

costly to use in practical sheet forming applications. The phenomenological plasticity

models are used instead, because of their benefit in the cost and ease in implementing in a

finite element approach. However, for a certain problem, like springback analysis, these

models are not accurate, because they do not account for the anisotropy hardening in an

arbitrary orientation to the RD. Experiments on Al-3%Cu for three-directional

compression were performed by Barlat et al. [9] in Figure 6.1, which shows the flow

stress in 45° is evolving in an anisotropic manner at finite strain. The predictions of

hardening using conventional isotropic hardening and kinematic hardening are shown in

Figure 6.2. Neither isotropic hardening nor kinematic hardening can predict the

anisotropic hardening for 45° because they do not account for the anisotropy evolution.

In order to account for such anisotropic hardening, Barlat and Liu [10], Barlat et

al. [9] suggested a fourth-order tensor in backstress formulation to describe the

anisotropy hardening using a nonlinear Armstrong-Frederick type kinematic hardening

model. The coefficients of the fourth-order tensor are determined by micro-mechanical

considerations using the ellipsoidal precipitate inclusion model introduced by Mura [75].

From the research of Barlat and Liu [10], anisotropy evolution for the precipitate

hardened material is dominant to the rotation of backstress. From the rotation of yield

function using plastic spin theory for the texture spin [14, 15], it is hinted that the

backstress also can be rotated as the precipitate rotates along the metal matrix

surrounding the precipitates. The rotation is investigated by modeling the backstress

124
reorientation using a plastic spin similar to the one used in the rotation of the yield

function.

6.2 Material modeling

Because the process is for the finite strain, the material formulation is based on

the multiplicative decomposition of the deformation gradient for defining the

intermediate configuration and is pushed forward to the current configuration as

summarized in Chapter 1.

The assumptions used in the formulation are: (i) matrix and precipitates have

approximately the same elastic characteristics, (ii) precipitates deform only elastically

(non-shearable), (iii) precipitates do not interact with each other.

The yield criterion is denoted as

Φ = f ( τ − q ) − 2 τya (6.1)

where

a a a
f = α1 s 2 − s3 + α 2 s3 − s1 + α 3 s1 − s 2 (6.2)

with a=8 for fcc materials. The details of the coefficient αk can be found in Barlat et al.

[11].

The isotropic parts to define the yield function size are defined by

(
τy = τ0 + K 1 − e
− N εp
) (6.3)

The elastic strain in the precipitate is assumed [10]

ε ep = ( I − Λ ) ε p , (6.4)

125
where I and Λ are identity and a tensor for precipitate inclusion. The tensor is defined

by Mura [75] as

1
Λ = ∑
f ( p)
f (p) Λ(p) , (6.5)

where f(p) is the volume fraction of a precipitate (p), and Λ ( p ) is Eshelby tensor for a

precipitate (p).

The applied stress can be decomposed to the matrix part and precipitate part as

{ }
τ = (1 − f ) τ m + f τ p , (6.6)

where f is the volume fraction of the precipitate, {τ p } is the average over all precipitates,

and τ m is for matrix. The alloy containing non-shearable precipitates shows a strong

Bauschinger effect, therefore, the precipitates are considered to be modeled by kinematic

hardening effectively. By using the assumption of (i),

τ p = Γ e (1 − Λ ) ε p , (6.7)

where Γ e is the elasticity tensor. For the part of precipitate inclusion in (6.6), the

backstress for the precipitate inclusion is modeled by Barlat and Liu [10] as

q = fΓ e {I − Λ } ε p h ( ε ) , (6.8)

where ε is the effective strain in the matrix. With the hint of (6.8), backstress possible

formulations using the fourth-order tensor and Armstrong-Frederick type kinematic

hardening are

126
q∇ = cAd p − sbq
q∇ = cAd p − sbAq , (6.9)
q∇ = cAd p − sbA −1q

where A is the fourth-order tensor defined as fΓ e {I − Λ } and


A ijklnia ⊗ n aj ⊗ n ak ⊗ n al . (6.10)

The initial value of tensor A in matrix form M has been provided in Barlat and Liu [10]

as

 46900 9700 10100 


 9700 43600 13300 0 
 
10100 13300 43300 
M= . (6.11)
 14600 
 0 11300 
 
 11000 

With the plastic deformation, the texture are changed which in turn affects the

orientation of the precipitates. In addition, it has been observed that the precipitates

themselves can also rotate relative to the crystal orientation [76, 77]. Although

precipitates are often assumed to be embedded in the habit planes of the lattice,

geometrically necessary dislocations and lattice rotations in the neighborhood of the

precipitates enable rotations of the precipitates relative to the crystal orientation. The

orientation of the averaged precipitate can be rotated by a corotational rate

o
nia = nia − θnia = 0 , (6.12)

where the constitutive spin θ can be defined with material spin and the plastic spin as we

defined for the texture rotation as

127
θ = w − ωp . (6.13)

From the heuristic definitions of the plastic spin, the following definition is effective to

describe the backstress rotation.

a
ωp =
φ
{( τ − q ) dp − dp ( τ − q )} . (6.14)

6.3 Algorithmic treatment for the rotation.

The algorithm for the three-dimensional formulations is same as two dimensional

cases. In contrast to the plane stress problem, the spin is not limited within planar

rotation. Therefore, it is required to define the rotation in three-dimensional space. For

the three-dimensional spin in Cartesian coordinates, it is assumed that spins along each

axis are small enough to ignore the higher-order term. Spins along each axis can be

defined with this small spin assumption as

1 0 0   1 0 −dθ y   1 dθz 0
 
R x = 0 1 dθ x  , R y =  0 1 0  , R z =  −dθz 1 0  , (6.15)
0 −dθ x 1  dθ y 0 1   0 0 1 

and then the combined rotation is defined with

 1 dθz −dθ y 
 
R = R x R y R z =  −dθz 1 dθx  (6.16)
 dθ y −dθz 1 

6.4 Numerical investigations for the kinematic hardening models

The behaviors of anisotropic hardening using (6.9) are investigated numerically to

decide which form of the backstress evolution is the most efficient to follow the

measured anisotropic hardening. Flow stresses of 45° oriented specimen are examined
128
with and without the backstress rotation in Figure 6.3 to see the influence of backstress

reorientation in the formulation. When the fourth-order tensor is used in the non-linear

term, the stress flow shows an unexpected trend as the backstress rotates. By observing

the characteristics of each backstress model on the stress flow, the equation (6.9)a shows

the possibility to follow the trend of measurements.

6.5 Discussion of the model

The measurements are Al-3%Cu of Barlat et al. [9]. As expected, the isotropic

hardening model does not follow the experimental measurement as shown in Figure 6.2.

Even if the higher-order yield function is used, it is insufficient to predict the anisotropic

hardening in 45° orientation. By adding kinematic hardening in (6.9), it does not capture

the anisotropic hardening because there is no specific definition for an arbitrary

orientation. The two models fit the hardening in the RD and TD fairly well.

Using the precipitate inclusion model as a fourth-order tensor in backstress

evolution, the model gives under-estimate of the hardening for 45° as shown in Figure

6.4a. The backstress in 45° evolves very little. The model using the fourth-order tensor

and its reorientation with a plastic spin shows anisotropic evolution in the hardening and

gives a fairly good correlation to the experiment, as shown in Figure 6.4b. The

backstress for 45° evolves in continuous increments.

6.6 Numerical application

In order to see the effect of anisotropic evolution, a cylindrical cup drawing

simulation in Figure 5.1 is used. In contrast to the model used in Chapter 5, solid 3d

elements are used in the model. Two layers of elements are used in the thickness

direction. The material properties regarding hardening models are from the paper of

129
Barlat et al. [9] and summarized in Table 6.2. Comparisons between the hardening

models are the earing pattern and thickness distribution in 45°.

The results of earing patterns of each model are shown in Figure 6.5. The models

which do not account for the kinematic hardening develop a wide-banded earing pattern

while the models which account for the anisotropy for kinematic hardening develop a

smaller difference in the earing pattern. The difference of using the rotational scheme on

the kinematic hardening is not severe, as long as the anisotropy is accounted for the

kinematic hardening. The residual stress distributions are shown in Figure 6.6 with

respect to the hardening models.

The thickness distributions along the 45 degree line are compared in Figure 6.5.

As in the earing pattern, the difference between the hardening models is dependent on the

evolution anisotropy. It should be noted that the comparison is done only by numerical

analysis.

6.7 Conclusion

Even though there are insufficient validations of the model by comparing proper

experiments, the hardening curves show the effect of reorientation of the backstress to

describe the anisotropic hardening behavior of the precipitate hardened materials. This

could mean that in the precipitate hardened materials, anisotropy evolution is dominant to

the reorientation of the backstress, not to the reorientation of yield function. With the

introduction of the fourth-order tensor, it grafts micro-mechanical properties into a

phenomenological framework. Also, the model is required to verify the backstress

evolution experimentally in different orientations.

130
c1, c2, c3, c4, c5, c6 1.144, 1.076, 1.066, 1.126, 1.023, 1.075

ax0, ay0, az0 0.12, 0.80, 1.00

ax1, ay1, az1 0.12, 0.80, 0.06

a 8.0

Table 6.1: YLD96 parameters

Ciso Cm
Hardening Model biso b α M
(MPa) (GPa)

A. Isotropic only 165 22 0 0 0 -

B. Isotropic + Kinematic
110 15 3.0 80 0 -
hardening

C. Isotropic + Anisotropic
110 15 0.08 80 0 Defined
Kinematic hardening

D. Isotropic + Anisotropic
110 15 0.08 80 60 Defined
Rotational Kinematic hardening
* Properties for Elasticity and Barlat YLD96 are common.

Table 6.2: Material properties for Numerical analysis.

131
350
Al-3%Cu
300

250
Stress (MPa)

200

150

100 RD
o
45
50
TD
0
0 0.05 0.1 0.15 0.2 0.25 0.3
Compressive strain

Figure 6.1: Stress flow measurements of each direction for Al-3%Cu precipitate
hardened material. Hardening of 45° shows more anisotropy than the RD and TD.

132
300

250

200

Stress (MPa)
150

100
Exp.
50 Stress (FEM)

0
0 0.05 0.1 0.15 0.2 0.25 0.3
Compressive strain
(a)

300

250

200
Stress (MPa)

150

100
Exp.
50 Stress (FEM)

0
0 0.05 0.1 0.15 0.2 0.25 0.3
Compressive strain
(b)

Figure 6.2: Neither conventional isotropic hardening (a) nor kinematic hardening (b)
with Barlat YLD96 anisotropic yield function does trace the measurements.

133
350

300
No spin

250 Spin

Stress (MPa)
200

150

100

-1
50 q=c M-sb M
m m

0
0 0.05 0.1 0.15 0.2 0.25 0.3
Compressive strain
(a)

350
Spin
300
No spin
250
Stress (MPa)

200

150

100

50 q=c M-sb M
m m

0
0 0.05 0.1 0.15 0.2 0.25 0.3
Compressive strain
(b)

Figure 6.3: Flow stresses of 45° for different backstress formulations. Those
formulations cannot trace the trend of anisotropic hardening shown in the measurement.

134
300

250

200

Stress (MPa)
150
Exp.
Stress (FEM)
100
Backstress (FEM)

50

0
0 0.05 0.1 0.15 0.2 0.25 0.3
Compressive strain
(a)
300

250

200
Stress (MPa)

150
Exp.
Stress (FEM)
100
Backstress (FEM)

50

0
0 0.05 0.1 0.15 0.2 0.25 0.3
Compressive strain
(b)

Figure 6.4: Flow stresses using the fourth-order tensor for precipitate inclusion without
rotation (a) and with rotation (b). The anisotropic hardening for 45° is captured by
rotation of backstress.

135
95

(A)

92.5

(B)

90 (D)

(C)
87.5

85
0 10 20 30 40 50 60 70 80 90

Figure 6.5: Earing pattern of cylindrical cup according to the hardening models: a)
Isotropic hardening only, b) Isotorpic + kinematic hardening, c) Isotropic + Anisotropic
kinematic hardeningm, d) Isotropic + Anisotropic kinematic hardening with rotation.

136
(a)

(b)

(c)

Figure 6.6: Residual stress distributions with respect to the hardening model: (a)
isotropic hardening, (b) anisotropic hardening with the fourth-order tensor, but no
rotation, (c) anisotropic hardening with the fourth-order tensor and rotation.

137
1.1

1.05
Thickness (mm)

1
(C) (D)
0.95

0.9

0.85 (B)

0.8
(A)

0.75
0 35 70 105
Distance from center, undeformed (mm)

Figure 6.7: Thickness distribution in 45° direction: a) Isotropic hardening, b) Isotropic +


Kinematic hardening, c) Isotropic + Anisotropic kinematic hardening, d) Isotropic +
Anisotropic kinematic hardening with rotation.

138
CHAPTER 7

CONCLUSION

Through the research on the evolution of anisotropy a new hardening model –

Rotational-Isotropic-Kinematic (RIK) hardening model – is developed. The RIK

hardening model can predict multi-axial/-path loading and anisotropy hardening in an

arbitrary orientation other than the RD and the TD. The number of material properties to

express the rotational hardening is two. A simple method to identify the required

parameters is introduced.

The capability of the RIK hardening model is demonstrated by a deep drawing

process of the cylindrical cup and square cup of the NUMISHEET benchmark problems.

An isotropic hardening and the ANK non-linear kinematic hardening model are used to

compare the results with that of the RIK hardening model and the averaged experimental

results. The RIK model shows more accurate results in the prediction of earing or draw-

in as well as the stress and strain distribution.

For the springback prediction, draw bead simulation is done by all the three

hardening models. An arbitrary oriented specimen shows different springback shapes in

the height and twisting mode. With the rotational hardening scheme, the stress

distributions after the drawing process are quite a lot different than the hardening without

the rotational scheme. Different stress causes different springback.

139
Anisotropy of materials is evolving during deformation. For BCC mild steel, the

evolution is rotational and the effect is dominant to the yield function. For FCC materials,

such as aluminum alloy, the evolution appears more dominant to the backstress evolution.

Further research is required to assess the rotation of the symmetry axes using the

texture analysis. Also in order to generalize the model more experimental investigation is

needed for materials other than mild steel. It is necessary to model the yield function

evolution using the texture analysis directly into the yield function to predict material

anisotropic behavior more precisely.

140
BIBLIOGRAPHY

1. Dieter, G.E., Mechanical Metallurgy, McGraw-Hill, (1988).

2. Banadic, D., Bunge, H.-J., Pohlandt, K., and Tekkaya, A.E., Formability of
Metallic Materials, Springer, (2000).

3. Bunge, H.J. and Nielsen, I., Experimental Determination of Plastic Spin in


Polycrystalline Materials, International Journal of Plasticity, 13, p.435-446,
(1997).

4. Kim, K.H. and Yin, J.J., Evolution of Anisotropy under Plane Stress, Journal of
Mechanics and Physics of Solids, 45, p.841-851, (1997).

5. Boehler, J.P., Evolution of Anisotropy in Sheet-Steels Submitted to Off-Axes


Large Deformations, Applied plasticity, p.143-158, (1991).

6. Hill, R., A Theory of the Yielding and Plastic Flow of Anisotropic Metals, in Roy.
Soc. London A, p.281-297, (1948).

7. Chun, B.K., Jinn, J.T., and Lee, J.K., Modeling of the Bauschinger Effect for
Sheet Metals, Part I: Theory., International Journal of Plasticity, 18, p.571-595,
(2002).

8. Chun, B.K., Kim, H.Y., and Lee, J.K., Modeling the Bauschinger Effect for Sheet
Metals, Part II: Applications., International Journal of Plasticity, 18, p.597-616,
(2002).

9. Barlat, F., Liu, J., and C., B.J., Plastic Anisotropy in Precipitation-Strengthened
Aluminum Alloys Deformed in Uniaxial and Plane Strain Deformation,
Modelling Simul. Mater. Sci. Eng., p.435-443, (2000).

10. Barlat, F. and Liu, J., Precipitate-Induced Anisotropy in Binary Al-Cu Alloys,
Materials Science and Engineering, A257, p.47, (1998).

141
11. Barlat, F., Maedat, Y., Chung, K., Yanagawa, M., Brem, J.C., Hayashida, Y.,
Lege, D.J., Matsui, K., Murtha, S.J., Hattori, S., Becker, R.C., and Makosey, S.,
Yield Function Development for Aluminum Alloy Sheets., Journal of Mechanic
and Physics of Solids, (1997).

12. Kuroda, M., Interpretation of Behavior of Metals under Large Plastic Shear
Deformations: Comparison of Macroscopic Predictions to Physically Based
Predictions, Int. J. Plast., 15, p.1217-1236, (1999).

13. Kuroda, M., Interpretation of the Behavior of Metals under Large Plastic Shear
Deformations: A Macroscopic Approach, International Journal of Plasticity,
13(4), p.359-383, (1997).

14. Dafalias, Y.F., Orientational Evolution of Plastic Orthotropy in Sheet Metals,


Journal of Mechanics and Physics of Solids, 48, p.2231-2255, (2000).

15. Han, C.-S., Choi, Y., Lee, J.K., and Wagoner, R.H., A Fe Formulation for Elasto-
Plastic Materials with Planar Anisotropic Yield Functions and Plastic Spin,
International Journal of Solids and Structures, 39, p.5123-5141, (2002).

16. Truong Qui, K.H. and Lippmann, H., On the Impact of Local Rotation on the
Evolution of an Anisotropic Plastic Yield Condition, J. Mech. Phys. Solids, 49,
p.2577-2591, (2001).

17. Dafalias, Y.F., The Plastic Spin, Journal of Applied Mechanics, 52, p.865, (1985).

18. Kuroda, M., Plastic Spin Associated with Corner Theory of Plasticity, Int. J.
Plast., 11(5), p.547-570, (1995).

19. Simo, J.C. and Hughes, T.J.R., Computational Inelasticity, Springer, (1998).

20. Doucet, A.B. and Wagoner, R.H., Plane-Strain Work Hardening and Transient
Behavior of Interstitial-Free Steel, Metallurgical Transaction A, 18A, p.2139,
(1987).

21. Doucet, A.B. and Wagoner, R.H., Transient Tensile Behavior of Interstitial-Free
Steel and 70/30 Brass Following Plane-Strain Deformation, Metallurgical
Transaction A, 20A, p.1483, (1989).

22. Rauch, E.F., The Flow Law of Mild Steel under Monotonic or Complex Strain
Path, Solid State Phenomena, 23-24, p.317-334, (1992).

23. Thuillier, S. and Rauch, E.F., Development of Microbands in Mild Steel During
Cross Loading, Acta Metall. Mater., 42(6), p.1973-1983, (1994).

142
24. Hiwatashi, S., Van Bael, A., Van Houtte, P., and Teodosiu, C., Modelling of
Plastic Anisotropy Based on Texture and Dislocation Structure, Computational
Materials Science, 9, p.274-284, (1997).

25. Hiwatashi, S., Van Bael, A., Van Houtte, P., and Teodosiu, C., Prediction of
Forming Limit Strains under Strain-Path Changes: Application of an Anisotropic
Model Based on Texture and Dislocation Structure, Int. J. Plast., 14(7), p.647-669,
(1998).

26. Takahasi, H., Fujiwara, K., and Nakagawa, T., Multiple-Slip Work-Hardening
Model in Crystals with Application to Torsion-Tension Behavior of Aluminum
Tubes, International Journal of Plasticity, 14(6), p.489-509, (1998).

27. Teodosiu, C. and Hu, Z., Simulation of Materials Processing: Theory, Methods
and Applications, in Numiform 95, Balkema: Rotterdam, p.173, (1995).

28. Peeters, B., Seefeldt, M., Tedodosiu, C., Kalidindi, S. R., Van Houtte, P. and
Aernoudt, E., Work-Hardening/Softening Behavior of B.C.C. Polycrystals During
Changing Strain Paths: I. An Integrated Model Based on Substructure and Texture
Evolution, and Its Prediction of the Stress-Strain Behavior of an If Steel During
Two-Stage Strain Paths, Acta Materialaia, 49, p.1607-1629, (2001).

29. Peeters, B., Bacroix, B., Teodosiu, C., Van Houtte, P., and Aernoudt, E., Work-
Hardening/Softening Behavior of B.C.C. Polycrystals During Changing Strain
Paths: II. Tem Observations of Dislocation Sheets in an If Steel During Two-
Stage Strain Paths and Their Representation in Terms of Dislocations Densities,
Acta Materialia, 49, p.1621-1632, (2001).

30. Peeters, B., Kalidinidi, S.R., Teodosiu, C., Van Houtte, P., and Aernoudt, E.,
Modelling Microstructure-Based Anisotropy in Bcc Polycrystals During
Changing Strain Paths, Diffusion and Defect Data Pt. B: Solid State Phenomena,
87, p.163-168, (2002).

31. Asaro, R.J., Micromechanics of Crystals and Polycrystals, Advances in Applied


Mechanics, 23, p.1-115, (1983).

32. Bunge, H.J., Texture Analysis in Materials Science, Butterworth, (1982).

33. Dafalias, Y.F., Plastic Spin : Necessity or Redundancy?, International Journal of


Plasticity, 14(9), p.909-931, (1998).

34. Armstrong, P.J. and Frederick, C.O., A Mathematical Representation of the


Multiaxial Bauschinger Effect, G.E.G.B. p.731,(1966).

143
35. Truong Qui, K.H. and Lippmann, H., Plastic Spin and Evolution of an
Anisotropic Yield Condition, Int. J. Mech. Sci., 43, p.1969-1983, (2001).

36. Brocato, M., Tamagny, P., and Ehrlacher, A., Texture Evolution and Rotational
Hardening in Multiple Slip Plasticity: A Two Dimensional Study, Eur. J. Mech.
A/Solids, 20, p.345-365, (2001).

37. Lee, E.H., Elastic-Plastic Deformations at Finite Strains, Journal of Applied


Mechanics, 36, p.1, (1969).

38. Chaboche, J.L., Time-Independent Constitutive Theories for Cyclic Plasticity,


International Journal of Plasticity, 2, p.149-188, (1986).

39. Chaboche, J.L., Constitutive Equations for Cyclic Plasticity and Cyclic
Viscoplasticity, Int. J. Plast., 5, p.247-302, (1989).

40. Loret, B., On the Effects of Plastic Rotation in the Finite Deformation of
Anisotropic Elastoplastic Materials, Mech. Mater., 2, p.287-304, (1983).

41. Dafalias, Y.F., Issues on the Constructive Formulation at Large Elastoplastic


Deformations, Part 2: Kinetics, Acta Mechanica, 73, p.121-146, (1988).

42. Dafalias, Y.F. and Rashid, M.M., The Effect of Plastic Spin on Anisotropic
Material Behavior, International Journal of Plasticity, 5, p.227-246, (1989).

43. Van der Giessen, E., Micromechanical and Thermodynamical Aspects of the
Plastic Spin, International Journal of Plasticity, 7, p.365-386, (1991).

44. Kuroda, M., Roles of Plastic Spin in Shear Banding, International Journal of
Plasticity, 12(5), p.671-698, (1996).

45. Zbib, H.M. and Aifantis, E.C., On the Concept of Relative and Plastic Spins and
Its Implications to Large Deformation Theories, Part I: Hypoelasticity and
Vertex-Type Plasticity., Acta Mechanica, 75, p.15, (1988).

46. Choi, Y., Han, C.-S., Lee, J.K., and Wagoner, R.H., Effect of Anisotropy Axes
Rotation on Spring Back, Numisheet 2002 proceedings, (2002).

47. Haupt, P. and Tsakmakis, C., On Kinematic Hardening and Large Plastic
Deformations, International Journal of Plasticity, 2, p.279-293, (1986).

48. Han, C.-S., Choi, Y., Lee, J.K., and Wagoner, R.H., Modeling of Anisotropic
Work Hardening, Numisheet 2002 proceedings, (2002).

144
49. Tsakmakis, C., Kinematic Hardening Rules in Finite Plasticity, Part I: A
Constitutive Approach, Continuum Mechanics and Thermodyanamics, 8(215-231),
(1996).

50. Han, C.-S., Chung, K., Wagoner, R.H., and Oh, S.-I., A Multiplicative Finite
Elasto-Plastic Formulation with Anisotropic Yield Functions, Int. J. Plast., 19,
p.197-211, (2003).

51. Geng, L. and Wagoner, R.H., Springback Analysis with a Modified Nonlinear
Hardening Model, in SAE, Detroit, p.SAE2000-01-0410, (2000).

52. Valliappan, S., Boonlaulohr, P., and Lee, I.K., Non-Linear Analysis for
Anisotropic Material, International Journal of Numerical Method of Engineering,
10, p.597-606, (1976).

53. Zhao, K. and Lee, J.K., On Simulation of Bending/Reverse Bending of Sheet


Metals, ASME, 10, p.929-933, (1999).

54. Zhao, K. and Lee, J.K., Generation of Cyclic Stress-Strain Curves for Sheet
Metals, ASME MED, 11, p.667-674, (2000).

55. Hill, R., Constitutive Modelling of Orthotropic Plasticity in Sheet Metals, J. Mech.
Phys. Solids, 38, p.405-417, (1990).

56. Barlat, F., Lege, D.J., and Brem, J.C., A Six-Component Yield Function for
Anisotropic Materials, International Journal of Plasticity, 7, p.693-712, (1991).

57. Agnew, S.R. and Weertman, J.R., The Influence of Texture on the Elastic
Properties of Ultrafine-Grain Copper, Materials Science and Engineering, A242,
p.174-180, (1998).

58. Beaudoin, A.J., Dawson, P.R., Mathur, K.K., Kocks, U.F., and Korzekwa, D.A.,
Application of Polycrystal Plasticity to Sheet Forming, Computer Methods in
Applied Mechanics and Engineering, 117, p.49-70, (1994).

59. Kocks, U.F., Tome, C.N., and Wenk, H.-R., Texture and Anisotropy, Cambridge
University Press, (1998).

60. Nakamachi, E. and Doug, X., Study of Texture Effect on Sheet Failure in a Limit
Dome Height Test by Using Elastic/Crystalline Viscoplastic Finite Element
Analysis, Journal of Applied Mechanics, 64, p.519-524, (1997).

61. Yang, D.-Y., Oh, S.I., Huh, H., and Kim, Y.H., eds., Numisheet 2002 Proceedings
-Verification of Simulation with Experiment-. Proceedings of the 5th International

145
Conference and Workshop on Numerical Simulation of 3d Sheet Forming
Processes, ed. Series, Vol. 2: Jeju Island, (2002).

62. Sutton, M.A., Wolters, W.J., Peters, W.H., and Chao, Y.J., Determination of
Displacements Using an Improved Digital Correlation Method, Image Vision
Computing, 1(3), p.133-139, (1983).

63. Vendroux, G. and Knauss, W.G., Submicron Deformation Field Measurements:


Part 2. Improved Digital Image Correlation, Engineering Mechanics, 38(2), p.86-
92, (1998).

64. Rao, K.P. and Mohan, E.V.R., A Vision-Integrated Tension Test for Use in Sheet-
Metal Formability Studies, Journal of Materials Processing Tech., 118, p.238-245,
(2001).

65. Zhao, K. and Lee, J.K., Generation of Cyclic Stress-Strain Curves for Sheet
Metals, Journal of Engineering Materials and Technology, 124, p.391-397,
(2001).

66. Kubin, L., Canova, G.R., Condat, M., Devinere, B., Pontikis, V., and Brechet, Y.,
Dislocation Microstructure and Plastic Flow: A 3d Simulation, Solid State
Phenom., 23-24, p.455, (1992).

67. Peeters, B., Kalidinidi, S.R., Teodosiu, C., and Aernoudt, E., A Theoretical
Investigation of the Influence of Dislocation Sheets on Evolution of Yield
Surfaces in Single-Phase B.C.C. Polycrystals, Journal of Mechanics and Physics
of Solids, 50, p.783-807, (2002).

68. Shizawa, K. and Zbib, H.M., Thermodynamical Theory of Plastic Spin and
Internal Stress with Dislocation Density Tensor, J. Eng. Mater. Technol. Trans.
ASME, 121(2), p.247-253, (1999).

69. Chaboche, J.L., Cyclic Viscoplastic Constitutive Equations, Part I: A


Thermodynamically Consistent Formulation, J. Appl. Mech. Trans. ASME, 60(4),
p.813-821, (1993).

70. Geng, L. and Wagoner, R.H., Role of Plastic Anisotropy and Its Evolution on
Springback, Int. J. Mech. Sci., 44(1), p.123-148, (2002).

71. Schmitt, J.H., Shen, E.L., and Raphanel, J.L., A Parameter for Measuring the
Magnitude of a Change of Strain Path: Validation and Comparison with
Experiments on Low Carbon Steel, Int. J. Plast., 10(5), p.535-551, (1994).

72. Dohmann, F. and Hartl, C., Hydroforming - a Method to Manufacture Light-


Weight Parts, Journal of Materials Processing Tech., 60, p.669-676, (1996).
146
73. Asnafi, N., Analytical Modeling of Tube Hydroforming, Thin-Walled Structures,
34, p.295-330, (1999).

74. Zhang, S.H., Developing in Hydroforming, Journal of Materials Processing Tech.,


91, p.236-244, (1999).

75. Mura, T., Micromechanics of Defects in Solids, Martinus Nijhoff, (1987).

76. Ashby, M.F., The Deformation of Plastically Non-Homogeneous Materials, Phil.


Mag., 21, p.399-424, (1970).

77. Hosford, W.F. and Zeisloft, R.H., The Anisotropy of Age-Hardened Al-4%Cu
Single Crystals During Plane-Strain Compression, Metall. Trans. A, 3, p.113-121,
(1972).

78. Oldroyd, J.G., On the Formulation of Rheological Equations of State, in Roy. Soc.
Lond., p.523, (1950).

147
APPENDICES

148
APPENDIX A

EXPLICIT FORMULATION

For the iterative approach, Netwon-Raphson method is used to compute proper

value of ∆γ , the consistency parameter,

fi
∆γ i +1 = ∆γ i − (A.1)
( )
i
∂f
∂∆γ
i
 ∂f 
where superscript i is iterative step. The derivation of   is as follows.
 ∂∆γ 
The yield function f is

2
f = φ− τy (A.2)
3

where φ and τy are defined as Hill’s 48 anisotropic yield function and the size of

isotropic yield function as

2
φ= τ ⋅ Pτ
3 . (A.3)
τy = Ciso ( ε + ∆εp )
* Niso

149
Take a derivative from (A.2)

∂f ∂φ ∂τ 2 ∂τy ∂∆εp
= ⋅ − ⋅ (A.4)
∂∆γ ∂τ ∂∆γ 3 ∂∆εp ∂∆γ

where

∂φ
τ = τ 0 − Γ e ∆ε p = τ 0 − ∆γΓ e
, and (A.5)
∂τ
2 2 2
∆εp = ∆ε p = ∆γ τ ⋅ Pτ = ∆γ τ ⋅ Pτ = ∆γφ . (A.6)
3 3 3

From (A.5),

∂τ ∂φ
= −Γe + H.O.T. (A.7)
∂∆γ ∂τ
∂∆εp
= φ. (A.8)
∂∆γ

Insert (A.7) and (A.8) to (A.4) then,

∂f ∂φ ∂φ 2 ∂τy
= − ⋅ Γe − φ. (A.9)
∂∆γ ∂τ ∂τ 3 ∂∆εp

Using the value computed at (A.9), the Newton-Raphson iteration (A.1) is used to find

proper value for ∆γ .

150
APPENDIX B

ALGORITHMIC TREATMENT FOR PLANE STRESS PROBLEM (IMPLICIT)

From the step n to the step n+1, the strain increments ∆ε , and the variables
n
τ n , y n , q1n , q n2 , ε np , ε pn , e φα from the previous load step n are assumed to be given. For the

trial state with the trial elastic strain, the stress is

τ trial = Fτ n F T + ∆τ trial where ∆τ trial = Γ e ∆ε n +1 . (B.1)

i
which correlates to the Oldroyd derivative for stress like variables (.) = (.)− L (.) − (.) LT ,

identical to the Lie-derivative defined as Lv = F  ∂∂t (.)  F T in an updated Lagrangian

setting [19, 78]. Γ e in (B.1) denotes the elasticity tensor. It may be helpful to formulate

the equation (B.1) relative to the referential coordinate system eiφ , because formulation is

for the plane stress set which is defined in the tangential coordinate system of the shell as,

τ trial = Uτ n U T + ∆τ trial (B.2)

The backstress are

2
y trial = ∑ q trial
j , qj
trial
= Fq nj F T
j=1
(B.3)
151
The trial isotropic hardening is set to

σ try = σny (B.4)

The yield function with Hill’s quadratic yield function is

( )
2
Φ trial = φtrial − σ trial
y , (B.5)

( ) (
where φtrial = τ trial − y trial iK φtrial τ trial − y trial . )

Elastic loading

If the yield condition is fulfilled with these trial values

(
Φ trial = Φ τ trial , y trial , σ trial
y ; eα
φ trial
≤ TOL ) (B.6)

the load step n+1 is considered to be elastic and hence d p = ω φp = 0 , and

τ n +1 = τ trial , y n +1 = y trial , eφα n +1 = eφα trial , σ ny +1 = σ trial


y , (B.7)

then continue to the next load step.

Plastic loading

In the case of plastic loading, Φ trial ≥ 0 the consistency of the yield function can

be obtain from the derivation of the yield function

∆Φ = ∆φ − 2σ y ∆σ y (B.8)
∂φ
∆φ = ⋅ ∆σ, σ = τ − y . (B.9)
∂σ

where

152
∂φ
= 2K φσ = a (B.10)
∂σ

This projection is done with residual method.

Residual method:

The consistent parameter can be found in iteration with

∆γ i +1 = ∆γ i + ∆ 2 γ i (B.11)

With the residual method, the increment of the consistent parameter is formulated as

∆ γ =
2 i (
Φ i − ai ⋅ Q −1rτi − ryi ) (B.12)
(a ⋅ Q
i −1
Γ ea + A + A ikin
i i
iso )
∂ai
where Qi = 1 + ∆γ i Γ ebi , bi = = 2K iφ . (B.13)
∂σ i

Where the residual values used in equation (B.12) are for the stress residual,

(
rτi = τ i − τ trial − ∆γΓ eai , ) (B.14)

and the backstress residuals in (B.12) is

ryi = d1rqi1 + d 2rqi 2 + 2 ( d1c1 + d 2 kc 2 ) ∆γ i Q −1rτi , (B.15)

where d1 = 1/ (1 + 2c1∆γ + 2∆γσ y b1 ) and d 2 = 1/ (1 + 2kc 2 ∆γ ) , respectively.

The individual backstress residuals are

(
rqi1 = q1i − q1tr + 2c1∆γ i σ i − 2∆γ i σiy b1q1i , ) (B.16)
(
rqi 2 = qi2 − q 2tr + 2kc 2 ∆γ i qi2 ) (B.17)

The scalar values in denominator of (B.12) are

153
∂σ y
A iso = 4Hσ2y where H = and (B.18)
∂εp
A kin = 2a ⋅ {( d1c1 + d 2 kc 2 ) σ − d1σ y b1q1} (B.19)

Once the increment of consistent parameter is set with (B.12) then the increments

of state variables are

∆γ i +1 = ∆γ i + ∆2 γ i (B.20)
∆τ = −Q
i −1
(r
i
τ + ∆ γ Γ ea
2 i i
) (B.21)

(
∆q1i = −d1 rqi1 − 2c1∆ 2 γ iσ i + 2c1∆γ iQ −1rτi + 2∆ 2 γ i σiy b1q1i ) (B.22)
n
∆y i +1 = ∑ ∆qij+1 (B.24)
j=1

For the reorientation of anisotropy axis using the plastic spin is computed from

updated state variables. The incremental form of the plastic spin is

∆ω ip+1 = µ φ (τ∆ε p − ∆ε p τ ) .
i +1
(B.25)

The plastic spin is an anti-symmetric tensor, so the components in matrix form are

 0 ∆ω12 
∆ω ip+1 =  . (B.26)
− ∆ω12 0 

From the non-diagonal terms and the assumption of keeping orthogonal between

the two planar anisotropy axes the anisotropy direction evolution is driven.

cos ∆ω12 − sin ∆ω12 


e iφ+1 = R iω+1e 0φ where R iω+1 =  (B.27)
 sin ∆ω12 cos ∆ω12 

K is now re-written in the basis of e iφ+1 then,

154
(K )i +1
φ ijkl
= ∑ 0
K αβγδ ( )( )( )(
eii +1 ⋅ eα0 eij+1 ⋅ eβ0 eik+1 ⋅ e0γ eik+1 ⋅ eδ0 ) (B.28)
αβγδ=1,2

Check yield criterion with updated values,

( )
2
Φ i +1 = φi +1 − σiy+1 (B.29)

If it returns greater than tolerance, repeat the iteration. If it returns less than or

equal to the tolerance, then transfer the last iterative values to next step values and quit

the iteration loop.

2
τ n +1 = τ i , y n +1 = ∑ q nj +1 , q1n +1 = q1i , q n2 +1 = q i2 ε pn +1 = ε pi , e φn +1 = e iφ (B.30)
j=1

Tangent modulus for implicit algorithm would be written as

Γ ep = M −
( Ma ) ⋅ ( Ma ) , (B.31)
a ⋅ Ma + A iso + A kin

where M = Q −1Γe .

155

You might also like