You are on page 1of 187

University of New Hampshire

University of New Hampshire Scholars' Repository

Doctoral Dissertations Student Scholarship

Fall 2011

Analytical and numerical prediction of localized necking in sheet


metal forming processes
Raed Zuhair Hasan
University of New Hampshire, Durham

Follow this and additional works at: https://scholars.unh.edu/dissertation

Recommended Citation
Hasan, Raed Zuhair, "Analytical and numerical prediction of localized necking in sheet metal forming
processes" (2011). Doctoral Dissertations. 625.
https://scholars.unh.edu/dissertation/625

This Dissertation is brought to you for free and open access by the Student Scholarship at University of New
Hampshire Scholars' Repository. It has been accepted for inclusion in Doctoral Dissertations by an authorized
administrator of University of New Hampshire Scholars' Repository. For more information, please contact
Scholarly.Communication@unh.edu.
ANALYTICAL AND NUMERICAL PREDICTION OF LOCALIZED

NECKING IN SHEET METAL FORMING PROCESSES

BY

RAED ZUHAIR HASAN

Master of Science, University of Technology-Baghdad, 2001

Ph.D. Dissertation

Submitted to the University of New Hampshire

in Partial Fulfillment of

The Requirements For The Degree of

Doctor of Philosophy

in

Mechanical Engineering

September, 2011
UMI Number: 3488791

All rights reserved

INFORMATION TO ALL USERS


The quality of this reproduction is dependent upon the quality of the copy submitted.

In the unlikely event that the author did not send a complete manuscript
and there are missing pages, these will be noted. Also, if material had to be removed,
a note will indicate the deletion.

UMI
Dissertation Publishing

UMI 3488791
Copyright 2011 by ProQuest LLC.
All rights reserved. This edition of the work is protected against
unauthorized copying under Title 17, United States Code.

ProQuest LLC
789 East Eisenhower Parkway
P.O. Box 1346
Ann Arbor, Ml 48106-1346
This dissertation has been examined and approved.

Dissertation Director, Brad Lee Kinsey,


Associate Professor of Mechanical Engineering

'or [jutti (PC?


Dissertatio/i Co-Director, Igor Tsukrov,
Professor of Mechanical Engineering

Robert Jerard,
Professor of Mechanical Engineering

Yajmis Korkons,
Assistant Professor of Mechanical Engineering

Edmund Chu,
Technical Leader/Manager at Alcoa

T/t'/n
Date
DEDICATION

This work is dedicated to my parents.

"£A+Ji\ p\&\ £i*j k a i l '

L?-&"

"Heaven lies under the feet of mothers."

Hadith Nabawi

m
ACKNOWLEDGEMENTS

Support from NSF (grant #0654124) and from Mechanical Engineering Department at

the University of New Hampshire is gratefully acknowledged.

I would like to thank my advisors Dr. Brad L. Kinsey and Dr. Igor Tsukrov for having

introduced me to the topic of sheet metal forming and numerical simulations, their

insistence on excellence, solid work ethics and endless patience enabled the successful

completion of this work.

Appreciation and gratitude are extended to Dr. Robert Jerard, Tracey Harvey and Dr.

Todd Gross the chair of the Mechanical Engineering Department for providing help and

support during my study at the University of New Hampshire.

Also, I would like to thank Dr. Edmund Chu from Alcoa for agreeing to be a member

of the examiners committee. Special thanks to Dr. Yannis Korkolis and my brother Dr.

Bessam Z. Hasan for the useful suggestions and discussions.

The last but not least thanks are reserved for my parents, my cousin, Inam and my

wife, Wijdan for the patience, encouragement and support. Thank you for all you have

done for me.


TABLE OF CONTENTS

LIST OF TABLES viii

LIST OF FIGURES ix

ABSTRACT xvii

NOMENCLATURE xix

CHAPTER PAGE

1. INTRODUCTION 1

1 -1. Background Information 1

1-2. Definition of Failure In Sheet Metal Forming 2

1-3. Motivation and Problem Statement 7

1-4. Research Goals and Contributions 12

2. ANALYTICAL MODELING OF FAILURE IN SHEET METAL FORMING

14

2-1. Review of Analytical Models Based on Plastic Instability 14

2.1.1. Empirical Equations 14

2.1.2. Swift-Hill's Model 15

2.1.3. Storen-Rice Model 17

2.1.4. M-K Model and H-N Extension of M-K Model 18

2.2. Comparison of Models to Past Experimental Data 24

2.3. Stress-Based Models 34

2.3.1. Non-Incremental Effective Stress Ratio Model 34

2.3.2. Modified Effective Stress Ratio Model 37


2.3.3. Comparison Between The M-K and The Effective Stress Ratio Models 40

2.3.4. Incremental Effective Stress Ratio Model 46

2.4. Non-Incremental, Major Principal Strain Ratio Model 48

2-5. Conclusions 58

3. NUMERICAL AND EXPERIMENTAL INVESTIGATIONS OF KEY

ASSUMPTIONS IN ANALYTICAL FAILURE MODELS 59

3.1 Introduction 59

3.2 Methodology 60

3.2.1 Experimental Procedure 60

3.2.2 Numerical Simulations Details 66

3.3 Results 68

3.3.1 Results for M-K Model Assumption 70

3.3.2 Results for Effective Stress Ratio Model Assumption 83

3.3.3 Results for Major Strain Ratio Model Assumption 90

3.5 Conclusions 100

4. EFFECT OF ELEMENT TYPES ON FAILURE PREDICTION USING A

STRESS-BASED FORMING LIMIT CURVE 102

4-1. Introduction 102

4-2. Marciniak Test Background 103

4-3. Numerical Simulations Details 106

4-4. Element Selection Guidelines 107

4-5. Comparison Between Element Types 109

4-6. Discussion 118

VI
4-7. Conclusion 119

5. CONCLUSIONS 120

6. FUTURE WORK 123

6.1 Analytical Modeling Improvements 123

6.2 Numerical Modeling Improvements 123

6.3 Experimental Verifications 124

REFERENCES 126

APPENDIXES 136

Appendix A: M-K Model Algorithm 137

Appendix B: M-K Predictions of The Materials Presented In Table 2.1 153

Appendix C: Modified Effective Stress Ratio Model Algorithm 161

Appendix D: Strain To Stress Conversion Methodology 164

vn
LIST OF TABLES

Table 2.1: Data of materials used in analyses

Table 2.2: Material models, f0 values and F-parameters determined

Table 3.1: Mechanical properties of AISI1018 used

Table 4.1: Seven specimen and washer dimensions used

Table 4.2: Number of nodes and elements in FEA models

Table 4.3: True stress and strain components values reported at crossing

paths with the experimental stress-based FLC

vm
LIST OF FIGURES

Figure 1.1: Examples of automotive industry applications 1

Figure 1.2: Strain-based forming limit diagram (FLD) 3

Figure 1.3: Forming Limit Curves (FLCs) for various sheet metals 3

Figure 1.4: Forming and fracture limit curves for AA5154-0 5

Figure 1.5: Fracture types during sheet forming 6

Figure 1.6: Formability window 7

Figure 1.7: Forming limit diagram with different stain paths 8

Figure 1.8: Strain-based FLCs for various uniaxial (U), equi-biaxial (E), and plane strain

(P) prestrains induced longitudinal (L) and transverse (T) to the major strain direction 8

Figure 1.9: Eleven strain-based FLCs from Graf and Hosford degenerate to one curve in

stress space when analytically converted using Barlat's 1989 yield criterion 10

Figure 1.2: Schematic of M-K model with reduced thickness in defect region (B)

compared to safe region (A) 19

Figure 2.2: Dependence of strain ratio on the defect orientation during localized necking

20

Figure 2.3: Schematic of M-K model with varying orientation of the defect (9) 21

Figure 2.4: Effect of changing 6 on FLC predictions for AA2008-T4 22

Figure 2.5: Predictions of FLC for HC220YD Steel (1) using different analytical models

26

Figure 2.6: Predictions of FLC for AA2008-T4 using different analytical models 27

ix
Figure 2.7: Example of strain-based FLC predicted using M-K model for AA2008-T4 28

Figure 2.8: Effect of changing / „ values on the FLC prediction of AA2008-T4 29

Figure 2.9: f0 values from M-K model that provide the best fit for the experimental data

29

Figure 2.10: Example of stress-based FLC predicted using M-K model for AA2008-T4

31

Figure 2.11: Average F- for p > 0 obtained from M-K model that provides the best fit

for the experimental data 32

Figure 2.12: Effect of changing strain path and the yield criteria on the F- parameter

using the modified M-K model and Al 2008-T4 alloy data 32

Figure 2.13: Effect of changing strain path on theF- parameter for different materials

33

Figure 2.14: F- parameter versus normalized depth at failure obtained from M-K model

33

Figure 2.15: Effective stress ratio model diagram 34

Figure 2.16: Methodology to increment OIA guess to determine a failure point on the

stress-based FLC 37

Figure 2.17: Effective stress ratio model results for Al 2008-T4 alloy with two yield

criteria 38

Figure 2.18: Effect of adding a concentrated stress region orientation {9 ) on the FLC

generated by the effective stress ratio model 39


Figure 2.19: Effect of adding a concentrated stress region orientation (0 ) on the stress-

based FLC generated by the effective stress ratio model 39

Figure 2.20: Comparison of perpendicular defect orientation (i.e., 0 = 0), Hill's equation

and au maximum method on the negative minor strain side of the FLC 40

Figure 2.21: Schematic of analysis methodologies in M-K and effective stress ratio

models 41

Figure 2.22: Comparison between the strain FLC prediction from M-K and the effective

stress ratio models for Al 2008-T4 43

Figure 2.23: Comparison between the stress FLC prediction from M-K and the effective

stress ratio models for Al 2008-T4 44

Figure 2.24: Strain paths from the incremental effective stress ratio model without

assuming compatibility 48

Figure 2.25: Least square slope ofF^j values obtained from M-K model that provides

the best fit for the experimental data 49

Figure 2.26: Linear fit of major principal strain data in the safe (A) and defect (B)

regions obtained from the M-K model for HC220YD Steel (1) 50

Figure 2.27: Linear fit of major principal strain data in the safe (A) and defect (B)

regions obtained from the M-K model for AA2008-T4 50

Figure 2.28: Comparison between forming limit stress diagrams obtained from M-K

model and the major principal strain ratio (MPSR) model for AA2008-T4 53

Figure 2.29: Comparison between FLC obtained from M-K model and the major

principal strain ratio (MPSR) model with varying yield function for AA2008-T4 54

XI
Figure 2.30: Comparison between FLC obtained from M-K model and the major

principal strain ratio (MPSR) model with varying 8 for AA2008-T4 56

Figure 2.31: Comparison between FLC obtained from M-K model and the major

principal strain ratio (MPSR) model with varying imperfection factor f0 for AA2008-T4

57

Figure 2.32: Comparison between FLC obtained from M-K model and the major

principal strain ratio (MPSR) model with varying the major strain concentration factor

Fs] for AA2008-T4 57

Figure 3.1: Representation of main groups of geometries for the Raghavan modification

to Marciniak test 60

Figure 3.2: Contour plot for locating X- and Y- axis on the specimen 61

Figure 3.3: Example of a grid of extracted nodes on the specimen 62

Figure 3.4: Strain paths varying from uniaxial to balanced biaxial cases for the various

specimen types 62

Figure 3.5: Tensile test data for AISI1018 used 65

Figure 3.6: AISI1018 hardening curves 67

Figure 3.7: Numerical simulated stress paths varying from uniaxial to balanced biaxial

cases for the various specimen types 69

ds
Figure 3.8: Initial Strain Path ( p = — - ) versus Forming Depth at Failure 69
dsx

Figure 3.9: Numerical and experimental contour plots of major true strain for Type 1-1

specimens 71

xn
Figure 3.10: Numerical and experimental contour plots of major true strain for Type II-2

specimens 72

Figure 3.11: Numerical and experimental contour plots of major true strain for Type II-3

specimens 73

Figure 3.12: Numerical and experimental contour plots of major true strain for Type IV

specimens 74

Figure 3.13: Strain path curves for the defect node and different locations away in the X-

direction for Type 1-1 specimens 76

Figure 3.14: Strain path curves for the defect node and different locations away in the X-

direction for Type II-2 specimens 77

Figure 3.15: Strain path curves for the defect node and different locations away in the X-

direction for Type II-3 specimens 78

Figure 3.16: Strain path curves for the defect node and different locations away in the X-

direction for Type IV specimens 79

Figure 3.17: Incremental major strain ratios versus forming depth for Type 1-1 specimen

81

Figure 3.18: Incremental major strain ratios versus forming depth for Type II-2 specimen

81

Figure 3.19: Incremental major strain ratios versus forming depth for Type II-3 specimen

82

Figure 3.20: Incremental major strain ratios versus forming depth for Type IV specimen

82

xni
Figure 3.21: Initial strain path versus incremental major strain ratio directly before

failure 83

Figure 3.22: Critical stress concentration factor (i.e.,F- parameter) versus the X-

direction location for Type 1-1 specimens at various distances from failure 85

Figure 3.23: Critical stress concentration factor (i.e.,F- parameter) versus the X-

direction location for Type II-2 specimens at various distances from failure 86

Figure 3.24: Critical stress concentration factor (i.e., F- parameter) versus the X-

direction location for Type II-3 specimens at various distances from failure 87

Figure 3.25: Critical stress concentration factor (i.e.,/7- parameter) versus the X-

direction location for Type IV specimens at various distances from failure 88

Figure 3.26: F- parameter versus distance from the defect node in the X-direction for

different specimen types 89

Figure 3.27: Critical stress concentration factor versus initial strain path for both

experimental data (Exp) and numerical simulation (FEA) 90

Figure 3.28: Critical major strain concentration factor (i.e., Fc parameter) versus the X-

direction location for Type 1-1 specimens at various distances from failure 92

Figure 3.29: Critical major strain concentration factor (i.e., Fe parameter) versus the X-

direction location for Type II-2 specimens at various distances from failure 93

Figure 3.30: Critical major strain concentration factor (i.e., F£ parameter) versus the X-

direction location for Type II-3 specimens at various distances from failure 94
Figure 3.31: Critical major strain concentration factor (i.e., Fe parameter) versus the X-

direction location for Type IV specimens at various distances from failure 95

Figure 3.32: Fe parameter versus distance from the defect node in the X-direction for

different specimen types 96

Figure 3.33: Critical major strain concentration factor versus initial strain path for both

experimental data (Exp) and numerical simulation (FEA) 97

Figure 4.1: Modified Marciniak test tooling 104

Figure 4.2: Specimen and washer geometry types 105

Figure 4.3: FEA model showing the punch, die, locking ring and specimen 107

Figure 4.4: Boundary Conditions used in the FEA model for Type I specimen 108

Figure 4.5: Stain paths generated using the three element types with the experimental

Forming Limit Curve 113

Figure 4.6: Stress paths generated using the three element types with the experimental

stress-based FLC 114

Figure 4.7: FLC generated using the three element types with the experimental FLC 114

Figure 4.8: Punch displacements at failure versus the strain ratio for seven specimen

types 115

Figure 4.9: Location of the imperfection added to the mesh for specimen Type III-2 115

Figure 4.10: Effect of imperfection thickness (shell element) for specimen Type III-2

117

Figure 4.11: Location of failure for specimen Type III-2 117

Figure 6.1: Biaxial testing apparatus 125

Figure A.1: Original M-K model flowchart for a single deformation path 144-145

xv
Figure A.2: The modified M-K model with the transverse direction t parallel to the

defect and the normal direction n perpendicular to the defect 147

Figure B.l: M-K prediction of FLC for HC220YD Steel (1) 153

Figure B.2: M-K prediction of stresses at failure for HC220YD Steel (1) 153

Figure B.3: M-K prediction of FLC for HC220YD Steel (2) 154

Figure B.4: M-K prediction of stresses at failure for HC220YD Steel (2) 154

Figure B.5: M-K prediction of FLC for AK Steel (2) 155

Figure B.6: M-K prediction of stresses at failure for AK Steel (2) 155

Figure B.7: M-K prediction of FLC for AISI 304 SS 156

Figure B.8: M-K prediction of stresses at failure for AISI 304 SS 156

Figure B.9: M-K prediction of FLC for AA2008-T4 157

Figure B.10: M-K prediction of stresses at failure for AA2008-T4 157

Figure B. 11: M-K prediction of FLC for AA5182 158

Figure B.12: M-K prediction of stresses at failure for AA5182 158

Figure B.13: M-K prediction of FLC for AA6016-T4 159

Figure B.14: M-K prediction of stresses at failure for AA6016-T4 159

Figure B.15: M-K prediction of FLC for AA6111-T4 160

Figure B.16: M-K prediction of stresses at failure for AA6111-T4 160

xvi
ABSTRACT

ANALYTICAL AND NUMERICAL PREDICTION OF LOCALIZED

NECKING IN SHEET METAL FORMING PROCESSES

By

Raed Z. Hasan

University of New Hampshire, September, 2011

The goal of this research is to investigate, develop and validate analytical and

numerical tools that can accurately predict failure in sheet metal forming operations. The

strain-based forming limit curve (FLC) is one of the tools to predict the maximum

permissible strains of thin metallic sheets which are loaded in the plane of the sheet to

different states of stress. It may be used to assess forming operations in the press shops as

well as unintentional deformations, such as vehicle, aircraft, or train crashes. The

accurate prediction of failure in sheet metal stamping can shorten product lead times,

decrease tooling costs, and allow for overall more rigorous and robust designs. In the

present work, three main analytical models are discussed: the modified Marciniak and

Kuczynski (M-K) model which included a varying defect orientation with respect to the

principal stress directions, the effective stress ratio model and the major strain ratio

model. The M-K predictions of the FLC are demonstrated with several yield criteria for

eight different materials. Furthermore, the stress-based FLC theory of the effective stress

ratio model is described, and its extension to include varying defect orientation is

xvii
presented. Also the original non-incremental FLC theory of the major strain ratio model

is presented, and its extension to include varying defect orientation is described. In

addition, numerical and experimental data are used to investigate the key assumption of

the three analytical models and the results show that the parameters investigated to

predict failure (i.e., the incremental strain ratio, critical stress concentration factor and

critical strain concentration factor) were not constant for the various strain paths for three

analytical models considered. Finally, finite element analysis (FEA) is used to predict the

FLC with a stress-based failure criterion and a comparison between three different

element types (shell, solid and solid-shell) is investigated in detail. The numerical results

show that despite the differences in stress distribution assumptions, shell, solid and solid-

shell elements would not provide differences in failure prediction for the uniform, in-

plane stretching states examined when a stress-based failure criterion is used.

xvm
NOMENCLATURE

ava2,(7^ Principal Stresses (MPa)

eve2,e3 Principal Strains

m Strain Rate Sensitivity Exponent

n Strain Hardening Exponent

K Strength Coefficient (MPa)

& Effective Stress (MPa)

£ Effective Strain

s Strain Rate

g Effective Strain Rate

ds. Ratio of Incremental Minor Strain to Incremental Major


dex
Strain

<T 2 Ratio of Principal Stresses


a =
a

t
l
Initial Thickness of The Safe Region (mm)
0A

t Initial Thickness of Defect (mm)

f _ OB
Initial Thickness Ratio
Jo —~
*0A

tB Instantaneous Thickness Ratio


tA

Higher Order Yield Criterion Index, (For BCC Materials

a=6 and For FCC Materials a=8)

xix
Plastic Anisotropy Ratio

Plastic Anisotropic Ratio in Rolling Direction

Plastic Anisotropic Ratio Transverse to Rolling

Direction

Average Plastic Anisotropic Ratio

Ratio of Effective Stress to Major Stress

Ratio of Incremental Effective Strain to Incremental

Major Strain

Barlat's Yield Function Parameters

Defect Orientation

Critical Stress Concentration Factor

Critical Major Strain Concentration Factor

Linear Tolerance In Major Strain Ratio Model

xx
CHAPTER 1

INTRODUCTION

1-1. BACKGROUND INFORMATION

Sheet metal forming operations are among the most useful mass production operations

in metal forming due to low costs and the existence of sheet materials with excellent

strength-to-weight ratios. These processes account for a sizable proportion of the

manufactured goods made in industrialized countries each year. For example, American

can manufacturers produce about 130 billion aluminum beverage cans annually,

equivalent to one can per American per day [1]. In automotive applications, deep drawing

and bending are among the most frequently employed operations. They are widely

utilized for the forming of outer panel and inner structural members (see Figure 1.1).

Figure 1.1: Examples of automotive applications [2]

1
To increase the performance of the manufactured products and address environmental

concerns, more and more light weight and high strength materials have been used as a

substitute for conventional steel. These materials usually have limited formability. Thus,

a thorough understanding of the deformation mechanisms and the factors limiting the

forming of defect-free parts is essential, not only from a scientific or engineering

viewpoint, but also from an economic viewpoint.

In sheet metal forming operations, the amount of useful deformation is limited by the

occurrence of necking or plastic instability which takes the form of diffuse or localized

necking. Traditionally, necking in sheet metal forming is predicted to occur if the in-

plane principal strains fall above the associated strain-based Forming Limit Curve (FLC)

when plotted in strain space [3]. The FLC is defined as the connection of all points which

mark the onset of instability for a sequence of strain paths. Figure 1.2 shows a typical

strain-based FLC with "Fail" and "Safe" regions labeled. This Forming Limit Diagram

(FLD) was generated experimentally after testing of several specimens. The strains at a

given location in the material would be compared to the FLC to determine if failure is a

concern. Different sheet metals have different FLCs, for example see Figure 1.3 [1].

1-2. DEFINITION OF FAILURE IN SHEET METAL FORMING

There are three major failure modes in sheet metal forming processes, i.e., localized

necking, wrinkling and rupture/fracture [2, 4, 5]. Localized necking results from

excessive in-plane stretching which leads to plastic instability. The necking area aligns

with the direction of zero extension in the plane of the sheet and would then proceed to

tearing. Necking can be predicted by comparing the principal in-plane strain components

2
with strain-based FLCs. If the imposed strain components are above FLC, then the

material is expected to fail.

1 On
o safety
09- T crack or neck
T
v
08- •
fUjk"'

07- e° \ -\ j \
c 9>\ T : Fail Region „o^
're § o\ T ] ^~15v
•fe ° 6 '
d) ^^ \ 1. jT

S 05-

! O-
•§• 0 4 -
Y Safe Region
03-
02-

01-

n A_
-0 5 -04 -03 -02 -01 0 0 01 02 03 04 05
Minor true strain

Figure 1.2: Strain-based Forming Limit Diagram (FLD) [2]

Plane strain

Equal (balanced)
biaxial

Low-carbon
Failure steel
zone^-^-ss^-^Brass

High-strength
40

20
V steel
/Aluminum alloy

60
. Simple \ x \
tension
0 (forJ?= 1)
40 20
HiZ
0
* /

20
j _

40
Safe zone
_L
60 80
Minor strain (%)

Figure 1.3: Forming Limit Curves (FLCs) for various sheet metals [1]
3
Wrinkling is the result of compressive loading and/or unsupported material. For

example in a typical drawing process, restraining forces are applied to the blank through

frictional forces between the tooling and the blank, by the binder ring and punch

geometries; and/or through a bending/unbending force imposed by drawbeads. Such

restraining forces determine how the material flows and consequently the stress states in

the sheet. When in-plane compressive stress exists in the sheet, wrinkling could initiate in

the formed wall region where the blank is free of a normal constraint or in the flange area

where a pressure is imposed onto the sheet by the binder. The wrinkling tendency is

affected by elastic modulus, sheet thickness, blankholder forces and tooling [2, 4].

Shear fracture often results from intense localization of strain on planes of maximum

shear in the root of the localized neck. The strain at fracture, ef , is a material property. It

is possible to measure these fracture strains and plot them on an in-plane strain diagram

similar to the FLD. If the fracture strain for the material exceeds the strain at which the

local neck reaches plane strain, it does not influence the limit strains. If, however, the

fracture occurs before plane strain-necking, it reduces the forming limit. This is

illustrated in Figure 1.4 from LeRoy and Embury [6], which shows that in biaxial

tension, the limit strain in an aluminum alloy is governed by fracture rather than by local

necking.

Ductile fractures are another common fracture type in sheet metal forming processes

(see Figure 1.5) [2, 4, 5]. Ductile fractures are characterized by high deformation before

fracture. Stresses above the yield strength will in ductile materials result in plastic

deformation due to atomic lattice gliding and mobility/generation of dislocations. The

first phase of the plastic deformation will occur for the most part uniformly without

4
localized necking. Microvoids will be created and joined as the deformation process

progresses. Eventually the loading will be such that a localized neck will develop due to a

defect at a specific location in the material. Finally, triaxial stresses develop and result in

a cleavage fracture.

Figure 1.4: Forming and fracture limit curves for AA5154-0 [6]

There are two types of ductile fractures on a macroscopic scale:

• Flat-face tensile fracture

• Shear-face tensile fracture

Flat-face tensile fracture is a fracture normal to the largest tensile stress and with a

shear lip to the free surface. Shear-face tensile fracture is a fracture 45° to the surface.

5
Face-centered cubic (FCC) metals like austenitic stainless steels (e.g., AISI 316) and pure

aluminum are ductile due to the significant number of slip systems (e.g., 12). Body-

centered cubic (BCC) metals like ferritic steels also have a significant number of slip

systems (e.g., 12). Hexagonal close-packed (HCP) metals like magnesium have only

three slip systems and normally exhibit a brittle behavior [2].

Figure 1.5: Fracture types during sheet forming [2]

Based on the modes of failure mentioned above, the concept of a formability window

was first proposed and developed by Marciniak and Duncan [5]. A robust stamping

process is one in which the strains in the part lie well within the formability window as

shown in Figure 1.6. The nature of the window and the influence of material behavior on

its shape can be predicted [2, 5]. In this dissertation, the focus will be on the prediction

of localized necking and plastic instability.

6
Fracture Tearing
Fracture
f&
/ Forming
window
Wrinkling/* WL
•c£—xm— ~qpr,»
1
->f 2
Compression

Figure 1.6: Formability window [5]

1-3. MOTIVATION AND PROBLEM STATEMENT

The experimental determination of a complete FLD as the one shown in Figure 1.2

can require several tests to be performed with various specimen geometries to vary the

strain paths (ratio of minor strain to major strain) from uniaxial to biaxial stretching as

shown in Figure 1.7. In-plane strains must be measured during each experiment using

Circle Grid Analysis (CGA) or Digital Image Correlation DIC systems. Strain

measurements by CGA are error prone and take a long time to be completed successfully.

While strain measurements by DIC require calibrations and require an expensive system.

A major concern with the strain-based failure approach is that the FLCs exhibit

significant strain path dependence. During forming of complex geometries, sheet

materials often undergo a series of operations before the final part is produced. Each step

in the process changes the strain path induced in the material at a given location, thus

7
shifting and changing the shape of the strain-based FLC. Figure 1.8 shows the significant

effect that prestrain has on the FLCs of AA 2008-T4 [7].

%=•-«

<0
S
^o 0 Forming
Limit
Curve
E, = %

o
/*Y^V. \lfc
5ft
^ \ >J%
\J ^n* &
n
8?

Ax \

w <4r
« +
Minor Strata, 6j

Figure 1.7: Forming limit diagram with different stain paths [2]

04

03-
c
s NO PRESTRAIN
55
PT
| 02
l-

01 -

-02 -0 1 0 01 02 03 04
Minor True Strain

Figure 1.8: Strain-based FLCs for various uniaxial (U), equi-biaxial (E), and plane strain

(P) prestrains induced longitudinal (L) and transverse (T) to the major strain direction [7]
Marciniak and Kuczynski [8] developed an analytical model to predict these diagrams

without the use of significant experiments. By imposing conditions on strain along with

an imposed material defect, a strain-based FLC is obtained. This model has been shown

to match experimental data for the positive minor strain side of FLD with reasonable

agreement. The curves created by this analytical approach are not surprisingly affected by

changes in material parameters and plasticity models [8]. The M-K model also accurately

captures the effects of prestrain on the location and shape of the FLC [9].

Other research has shown that the strain-based FLD can be used in FEM analyses to

predict tearing concerns in numerical simulations when multiple strain paths are present

[10]. However, computation time increased significantly as the strain paths for each

element must be considered and M-K analyses performed in order to accurately predict

failure. These methods prove to be very complex when designing a forming process with

multiple operations.

An alternative to the strain-based FLD is one that is based on the stresses in the

material. Stoughton [11] analytically converted the curves from Graf and Hosford [7] to

stress space and showed that a stress-based FLD is less sensitive to the deformation path

(see Figure 1.9). Note that the 12% equi-biaxial case is an outlier presumably due to the

high prestrain value or experimental error. This conversion from strain to stress requires

the assumption of yield criterion and hardening models. While most curves degenerate to

a single stress-based FLC, the shape of the stress-based FLC varies depending on the

material models used as will be shown in Chapter 2.

Numerical simulations that incorporate a stress-based FLD have been shown to be less

computationally taxing and still achieved results that match experimental data [12].

9
These methods however still would require knowledge of the existing strain-based FLC

of a given material.

12% equi-biaxial case


400 -
L \ _ ^
^ ™ — 5 ^ — B ^ ^ " *
8 300 •
&
C/3

K 200 •
o

100 -

C) 100 200 300 400 500


Minor True Stress (MPa)

Figure 1.9: Eleven strain-based FLCs from Graf and Hosford degenerate to a single

curve in stress space when analytically converted using Barlat's 1989 yield criterion [7]

In addition, one major criticism of the stress-based FLC is the fact that it has not yet

been experimentally validated. The traditional FLD experimental procedures have no

direct method to measure the amount of stresses applied to the material. In-situ stresses

measurements are possible using a X-ray diffraction head placed over the failure location,

but this technique is still far from being applied outside specific laboratories. An

experimental setup with the ability to measure the applied loads in the principal

directions is necessary to obtain information about the stress in the material at failure.

Derov et al. [13] developed an analytical model to predict a stress-based FLD

referenced here as the effective stress ratio model. This model predicts both stress and

strain-based FLCs by assuming a constitutive relationship, yield criterion, and plastic

10
anisotropy ratio (R ) for the material. Similar to the M-K model, the effective stress ratio

model assumes a defect region in the material. However, instead of the decreased

thickness region as in the M-K model, a region of concentrated stress in the material is

assumed for the defect region. The critical stress concentration factor (i.e., F-

parameter), which is the ratio of the effective stress in the safe region of the material to

the effective stress in the defect region and determines the deformation state at necking

failure, is a key assumption in this model. As stress appears to be less deformation path

dependent, a model based on the effective stress seems reasonable. Also, the notion that

necking and tearing failures are related to the stress in the material is not unreasonable

considering that other material behaviors, such as plastic yielding and buckling, are

dependent on this parameter. The effective stress ratio model has been shown to

accurately predict the effects of prestrain on the location and shape of the FLC [13, 14].

In this dissertation, analytical and numerical procedures for the prediction of FLC are

proposed. Three analytical models are used in this dissertation: the strain-based M-K

model, a modified version of the effective stress ratio model where the concentrated

stress region can be oriented in any direction with respect to the principal stress directions

and a new major strain ratio model. A comparison between these three models is given

and factors affecting the accuracy of FLC prediction for all three models are discussed in

detail. Also the key assumptions of the three analytical models are investigated

experimentally and numerically. Finally, numerical investigations of three different

element types are presented to determine the best element type that can be used with the

stress-based failure criterion.

11
1-4. RESEARCH GOALS AND CONTRIBUTIONS

The main goal of this research is to develop analytical and numerical tools that can

accurately predict failure in sheet metal forming operations and may be used as a valid

alternative to the current experimental procedures required to generate a complete FLC.

For example, instead of performing several experimental tests to generate the FLD shown

in Figure 1.2, the analytical models proposed in this research can be used to generate

similar FLD with much fewer experimental tests for validation. In particular, a new

stress-based approach to predict failure, which is significantly less path dependent, is

investigated. Also, a new major strain ratio model is presented that can be more effective

during finite element simulations

Contributions to the literature and knowledge include the following:

• The effective stress ratio model is further developed by including the orientation

change in the concentrated stress region and adding more advanced yield criteria

to this model. The varying defect orientation improves the analytical prediction

on the negative minor strain side of the FLC, while the more advanced yield

criteria improves prediction on the positive minor strain side of the FLC.

• A new major strain ratio model is presented. It has significant computational

advantages over the traditional M-K model since it is a non-incremental approach

• The key assumptions of the M-K, the effective stress ratio and the major strain

ratio models were investigated experimentally and numerically.

• The element types which are appropriate for a stress-based failure criterion were

determined.

12
In Chapter 2 of this dissertation, the analytical framework of the M-K, the effective

stress ratio and the major strain ratio models are introduced and their results are

presented. In Chapter 3, the key assumptions in the analytical models are investigated

numerically and compared against previous experimental work. In Chapter 4, a

comparison between different element types used in numerical models that predict plastic

instability using a stress-based failure criterion is given. Finally, the conclusions and

future work are presented in Chapters 5 and 6 respectively.

13
CHAPTER 2

ANALYTICAL MODELING OF FAILURE IN SHEET METAL

FORMING

2-1. REVIEW OF ANALYTICAL MODELS BASED ON PLASTIC INSTABILITY

After the introduction of the FLDs concept by Keeler and Backofen in 1964 [3],

research focused on the development of mathematical models for theoretical

determination of FLDs was initiated. Here some of the well-known models are presented

briefly.

2.1.1. Empirical equations

Keeler and Brazier [15] proposed an empirical relationship for FLC calculation based

on statistical data collected for deep drawing quality (DQ) steels. The true limit major

strain for plane strain loading, FLCQ (see Figure 1.4), can be found from:

FLC0=\n -JL-(0.2325 + 0.1413f0) + ll (2.1)


VO.zllo J
where n is the strain hardening exponent from Ludwik-Hollomon strain-hardening law

(a = K(s)"), and t0is the thickness of the sheet (t0<3.175mm). The positive and

negative minor strain sides of the FLC can be calculated respectively using:

ff, = FLC0 + s2 (0.784854 - 0.008565 s2) (2.2)

ex =FLC0 +e2 (0.027254 s2 -1.1965 ) (2.3)

14
Keeler and Brazier's (K-B) equations are still used in industry to predict the FLC of

steel and the results obtained are in good agreement even with the experimental data for

new grades of steel alloys (Advanced High-Strength Steel (AHSS), Dual Phase (DP)

Steel, Transformation Induced Plasticity (TRIP) Steel etc.) [16].

The North American Deep Drawing Research Group (NADDRG) introduced an

empirical equation for predicting the FLD for several grades of steel alloys [16].

According to this model, the FLD is composed of two lines through the point FLC0 in the

plane strain state given in Eq. (2.1). The slopes of the lines located respectively on the

negative and positive minor strain sides of the FLC are 45° and 20°.

2.1.2. Swift-Hill's model

Swift [17] determined the conditions for plastic instability under plane stress for a

given material. Assuming von Mises yield condition and a Ludwik-Hollomon strain-

hardening law, a sheet element loaded along two perpendicular directions was analyzed

by applying the Considere criterion (e =n) for each direction. When^ 2 > 0 , the limit

strains are:

+ a.
\daU Kda2j ydaxj
(2.4)
2
^
cr. + <j-
yd<rxj yda2j j

'df^ 'df^
+ cr,
8a
\ U ydaxJ yd(T2j
£
2 = (2.5)
2
^
cr. + <T,
\d(JU Kd(j2j

15
where / is the yield function.

By using different yield functions, it is possible to evaluate Swift's limit strains as

functions of the ratio of principal stresses a = — and the material parameters (e.g.,

strain hardening exponent n and plastic anisotropy ratio R ) . As an example, if Hill's

1948 yield criterion is used, the limit strains are:

r _ ( TR ^
l + R(\-a) 1 =a + a2
X i+* JJ n (2.6)
l + 4i?+2J?-
(i+i?Xl+«) a +a
J)

({l + R)a-R{\-^la + a^
JJ
(2.7)
1 + 4R+2R'
(\ + R\l + a) 1 - a +a
(1 + Rf
W JJ
The expressions of the limit strains associated with other yield criteria are presented in

[18].

Hill [19] was the first who proposed a general criterion for localized necking in thin

sheets under plane stress states in the negative minor strain region (i.e., e2 <0). Hill [19]

showed that the neck angle is coincident with the direction where zero-elongation along

the neck occurs. Thus the straining perpendicular to the necking direction is due only to

the sheet thinning and is defined by:

KdaXJ
*i = (2.8)
^ ' ^
\d(7U yda2j

16
yda2/l
£2 =
f n (2.9)
df^ ^
+
\\d(7u \d(7u

If Hill's 1948 yield criterion is used, Hill's limit strains are:

l + RJl-af
*i = (2.10)
(l + «) J

a-R{l-a)
£2 = (2.11)

The FLC is obtained by computing the values of e} and £2 for different loading ratios

a (i.e., a=0 to a=\ for uniaxial tension to balanced biaxial tension). Swift criterion is

used for the right side of the FLC and Hill's criterion is used for the plane strain and left

side of the FLC.

2.1.3. Storen-Rice model

In order to create a bifurcation that allows a neck to form in plane strain for

proportional loading cases, Storen and Rice [20] introduced a vertex on the yield surface

whose discontinuity grows until a plane strain plastic increment is possible. They derived

the following relation for the major strain at the instant of instability using von Mises

yield function and Ludwik-Hollomon power law:

c 3p2+n(2 + pf
for n2=0 (2.12)
2{2 + pXl + p + p2)

3 + n{l + 2pf
£i = for n i=0 (2.13)
2(l + 2p\\ + p + p2)

17
wherep = — is the strain ratio(-0.5 <p<\) and ni and n2 are the components of the
*i

unit normal to the neck. The expressions of the limit strains associated with other yield

criteria are presented in [21].

2.1.4. M-K model and H-N extension of M-K model

Marciniak and Kuczynski [8] proposed in 1967 an incremental model for theoretical

determination of FLDs that included an infinite sheet metal containing a region of local

imperfection where heterogeneous plastic flow develops and localizes (see Figure 2.1).

The M-K analysis is based on assuming that necking is initiated by a geometrical or

structural non-homogeneity of the material represented by a variation in sheet thickness.

This thickness imperfection is assumed to be in the form of a groove perpendicular to the

principal strain directions as shown in Figure 2.1. The sheet is composed of the nominal

or "safe" area and weakened or "defect" area, which are denoted by A and B,

respectively. The initial imperfection factor of the defect f0 is defined as the thickness

ratio/ 0 = — , where t denotes the thickness and subscript o denotes the initial state
'CM

before the deformation. The strain and stress states in the two regions are calculated by

ds
imposing a strain path pA = ——, specifying the principal strain d£XB and guessing the
d£XA

principal strain dsXA . Compatibility (i.e., de2B = d£2A), the work-energy relationship and

constitutive material models for yielding and work hardening are used to calculate all the

stress and strain values in the two regions. To determine if the guessed value of d£XA is

correct, force equilibrium is verified, (see Appendix A for the set of equations and the

18
complete M-K algorithm used). As the deformation progresses, the strain paths in

regions A and B diverge (i.e., the ratio —— increases). The deformation of the specimen
d£XA


is assumed to be a localized neck in region B when —— is approximately 10 according
d£XA

to [8, 22, 23]. The first and second principal strains in region A (eXA and £2A) at the onset


of plastic instability define a point on the FLC, and by varying the strain ratio pA = ——
d£XA

different points on the FLC are obtained [22].

The assumption that the imperfection is oriented perpendicular to the principal

direction of loading is not correct for the negative minor strain region. According to Hill's

observations and the theory of localized necking [19], failure occurs at a certain angle 9

between (0° to 35°) depending on the loading path and material when sheet metal is

subjected to tensile stresses (see Figure 2.2).

Figure 2.1: Schematic of M-K model with reduced thickness in defect region (B)

compared to safe region (A) [8]

19
40
tan(6>) = J -P
35

30
t 25
ye Ml

1 #» 20 g

f 15

Localized nee k[5]


l 10

t t i i 1 s
'
-0.55 -0.5 -0.45 -0.4 -0.35 -0.3 -0.25 -0.2 -0.15 -0.1 -0.05 (3
P

Figure 2.2: Dependence of defect orientation on the strain ratio during localized necking

The M-K theory was extended by Hutchinson and Neale [23] by introducing an

orientation of the defect with respect to the principal stress and strain directions (i.e., 9

in Figure 2.3). They found that for each proportional strain path on the negative minor

strain side of the FLD, an imperfection orientation angle 6 exists which gives the

minimum limiting major strain. In this modified M-K analysis, the angle 0 between the

imperfection and one of the principal directions is updated from an initial value at each

increment of the plastic deformation as:

1 + d£,,
tm(0 + d0) = tan(0) (2.14)
l + fl?£\

20
'J

M
V^ f\^Jf t
Ol^py/ A
/ A
^ "^11 / fe
# B / * 1
/ 1 /
~K 9 y l a
i
0<
' £ .—_»___J5^^

Figure 2.3: Schematic of M-K model with varying orientation of the defect (#) [9]

where d£XA and <fe2/l are the major and minor principal strain increments in the plane of

the sheet of the nominal area, respectively. By varying 0 between 0° and 90° and

performing the analysis, a minimum value of £XA is found. This minimum value and its

corresponding minor strain are defined as the necking failure point on the FLC as shown

in Figure 2.4.

There are modifications to incorporate the varied defect angle into the basic equations

for the M-K analysis as presented in Appendix A. The geometric compatibility equation

is expressed as:

d£llA=d£llB (2.15)

and the force equilibrium equations across the imperfection groove are:

F„„A = FnnB (2.16)

21
where subscripts n and t denote the normal and tangential directions of the groove (see

Figure 2.3), respectively, and F is the force per unit width in the specified direction.

K = 537 MPa, n = 0.2845, R0 = 0.78, R45 = 0.485, R^ = 0.58


0.8
\
\\
0.7
\
\\ 0.6
\
\ -S
- o.5 a
v CO
\ <w
A \ - 0.4 u
. \ Hu
S \
A \ O
•"AA - 0.3 'j?
• Experimental data %±
A i 0.2
—•—fo=0.9805, von Mises-Swift, theta=0
- A - fo=0.9805, von Mises-Swift, theta = Hill's equation 0.1
- • - fo=0.9805, von Mises-Swift, theta varies
0
-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 ()
Minor True Strain

Figure 2.4: Effect of changing 6 on FLC predictions for AA2008-T4 [7]

Coordinate transformations are used to determine the forces, stresses and strains in the

safe and the defect regions from the principal stresses applied (see Appendix A). Using

the rotation matrix, T, the stress tensor is transformed to the defect orientation by:

U cos(#) sin(#)"
a"' = T^T7 = <*** m
where T = (2.17)
Vnt a
n -sin(0) cos(#)

The failure criterion in the M- K analysis can be defined in many ways. In the work of

Marciniak and Kuczynski [8], necking was considered to occur when the localization of

deformation reached a critical value (i.e., the plane stain case). Several researchers e.g.,

22
[24-26] have associated the necking in the M-K analysis to the ratio of the defective

strain increment in the groove to that in the nominal area —— . In their calculations,
KdelAJ

necking was assumed to occur when this ratio was greater than 10. This failure criterion

is also used in this dissertation.

The implementation of different yield criteria and material constitutive relationships in

the M-K model has been tested by several authors, e.g., Graf and Hosford [22], Barata

Da Rocha et al. [24, 25, 27], Ahmadi et al. [28], Assempour et al. [29, 30], Cao et al. [9,

31] and Banabic et al. [32]. Their research show that the predicted limit strains strongly

depend on the constitutive law, yield criterion and the anisotropic parameters

incorporated in the analysis. The use of a suitable yield function and material model that

describe accurately the plastic behavior of metals produces a better prediction of limit

strains with respect to the shape and location of the FLC. Sheet metal materials are

generally anisotropic, and it was assumed in this dissertation that the principal axes of

anisotropy are coincident with the principal stress directions in the sheet. Hill's 1979 non-

quadratic yield function [33] is one of the yield criteria used in this dissertation to

account for the sheet anisotropy. For plane stress conditions, this function can be written

as:

2(l + R)aa = (o-, + CJ2)a + (l + 2R\ax -<x2)" (2.18)

where a is the higher order yield criterion index (for BCC materials a = 6 and for FCC

materials a = 8) and R is the average anisotropy ratio determined from uniaxial tensile

tests conducted in three orientations with respect to the rolling direction of the sheet (0°,

45° and 90°):

23
- = R0+2R45+R90
4

Based on polycrystal experimental results, Lian and Barlat [34] proposed a yield

function (i.e., Barlat '89) which is able to describe the flow surface for aluminum alloy

sheets well and in some cases BCC materials as well. The Barlat '89 yield function has

the form:

a=(axa(b + baa +{2-bXl-hcc)aJa (2.20)

where a = 8, b and h are the parameters that account for anisotropic effects:

6 = 2 - 2 Us-^S^,h= L^L_l±^i (2.21)

Two hardening laws, namely Swift and Voce, are considered in this dissertation to

describe the materials behavior during work hardening. These laws are:

o(s) = K(E0 + ef (EY (Swift) (2.22)

a(s)=A-Bexp(- Ce) (Voce)

where a is the effective true stress, s is the effective true strain, K is the strength

coefficient, n is the strain hardening exponent, m is the strain rate sensitivity exponent, £0

is the prestrain, A, B and C are material constants. The Voce model is used here to

describe the behavior of aluminum alloys in particular.

2.2. COMPARISON OF MODELS TO PAST EXPERIMENTAL DATA

Published experimental data for eight different materials was used to determine M-K

model parameters which provided reasonable predictions of FLCs. See Table 2.1 and

Appendix B for details of the considered steel and aluminum alloy cases, and Figures

24
2.5 and 2.6 for representative examples of the experimental data and different model

results. In these figures, sixteen strain paths were used for the M-K model predictions

( - 0 . 5 < / ? < l i n 0.1 increments). From these figures, it is clear that the M-K model

provides the best prediction compared to this experimental data if the correct yield

function and constitutive relationship are used [40].

Swift Model Voce Model Anisotropic


Initial
Parameters Parameters Parameters
Material Thickness Ref.
K A B | C
t« (mm) E0 ; n m Ro Rts R90
(MPa) (MPa) <MPa)|
:
HC220YD Steel (1) 0.80 661.0 0.005 0.222 0.015 : " - 1.28 i 2.08 1.86 [35,36]

HC220YD Steel (2) L60" "(512.1"' 0.005 0.2280.015 - - - 1.68 2.12 I 2.54 [35,36]

AKSteel - 426.2" 0.010 67228 0.018 i" - 1.70 1.27 2.13 " [37]

AISI304 SS 0.70 1527.0 - 0.470 0.012 - - - 0.92 1.30 0.78 [38]

AA2008-T4 1.70 537.0 0.005 0.285 -0.003; 0.78 0.49 0.58 m


426 234.0 5.130

AA5182 1.10 507.7 0.005 0.280 0.91 0.68 0.83 ![35,36]


380 249.7 i 7.895

AA6016-T4 417.9 0.010 0.2451 - 0.80 0.43 0.61 [25]


318 1191.1 8.706

[
AA6111-T4 1.05 554.0 10.010 0.257 -0.004 0.79 ! 0.63 0.67 [39]
440 275.2j7.015

Table 2.1: Data of materials used in analyses

Figure 2.7 shows the M-K predictions of FLC for AA2008-T4 material using four

different yield functions (von Mises, Hill's 1948, Hill's 1979 and Barlat' 89). As expected

both von Mises and Hill's 1948 yield functions give poor predictions for the positive

25
minor strain side when used to describe the behavior of the AA2008-T4 material. Figure

2.8 shows how the /„ value affects the predicted FLC.

0.8 experimental data


<C-B equation
0.7 Modified M-K model, Hill's 79, a=6 - Swift
NADDRG
0.6 • Swift-Hill, Hill's 1948
storen-Rice n2=0, von Mises
e Storen-Rice n2=0, Hill's 1948 o
« 0.5
• * *

CO
o
2 0.4
H
v.

I 0.3
If^O"*^^1*^
0.2

0.1
K =661 MPa, n =0.2222, R0 1.28, R45 = 2.08, R90 = 1.86, m=0.015, fo=0.997

Figure 2.5: Predictions of FLC for HC220YD Steel (1) using different analytical models

26
0.8 .
K == 537 MPa, n = 0.2845, R0 = 078, R45 = 0.485, Rgo = 0.58, m=-0.003, fo=0.9805
0.7 o - Experimental data

Modified M-K model, Hills 79, a=8- Swift


0.6
> . . . . . . Swift-Hill, Hill's 1948
e
' 3 0.5 Storen-Rice n2=0, von Mises
•**

GO
0> ". « » _ Storen-Rice n2=0, Hill's 1948
5 0.4
H
J? 0.3

0.2

0.1

nU J, r i i T * r i

-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 0.6
Minor True Strain

Figure 2.6: Predictions of FLC for AA2008-T4 using different analytical models

Note that the/ 0 values shown in Table 2.2 and Figure 2.9 are material model

independent. /„ is the same to match the experimental data regardless of material model

used. That is the / 0 values for steel with Hill-Swift and the f0 values for aluminum with

Barlat-Voce which provide the most accurate prediction of the experimental data are

fairly consistent (i.e., above 0.99). Note however that if Hill-Swift is used for aluminum

cases, different f0 values (i.e., 0.98-0.987) are obtained. Furthermore, a significantly

different /„ value was found for the AISI 304 stainless steel case considered (i.e., 0.963).

This was also observed by several others researchers [22, 27, 41, 42].

27
Yield Hardening F s from Eq.
Material a /. F5 for p > 0 F el forp>0
Criterion Model (2.43) at 0 A = 1

HC220YD Steel (1) Hills'79 6 Swift 0.997 o".931 0.941 0.767

HC220YD Steel (2) Hills'79 6 Swift 0.998 0~929 0.943 0.783

AK Steel Hills 79 6 Swift 0.999 0.930 0.940 0.813

AISI304 SS Hills 79 6 Swift 0.963 0.830 0.852 0.611

Hills 79 ! Swift o.m 0.920 0.929 0.658


AA2008-T4 !8
Barlat '89 i Voce 0.990 0.962 - 0.738

Hills 79 1 6 Swift 0.987 0^937 0.944 0.714


AA5182
Barlat '89; 8 Voce 0.999 0.993 - 0.852

I Hills '79 6 Swift 0.980 6525 0.939 0.660


AA6€16-T4
iBarlat '89 8 Voce 0.996 0.987 - 0.799

Hills 79 6 Swift 0.980 0.928 0.937 0.666


AA6111-T4
Barlat '89 8 Voce 0.993 0.976 - 0.778

Table 2.2: Material models, /„ values and F-parameters determined from M-K model

0.7
K=537 MPa, n=(|.2845, Ro=0.78, R45=0.485, R90=0.58, A=426 MPa, B=234.0
MPa,C=5.130
0.6

.5 °-5
'8
VI
u 0.4
s
u o Experimental data
H
s- 0.3
O ?<j*+++ ® - fo=0.9805, von Mises- Swift
0.2 •••••• fo=0.9805, Hill's 1948- Swift
-»— fo=0.9805, Hill's 79, a=6- Swift
0.1 -fo=0.9805, Hill's 79, a=8- Swift
- • - fo=0.99 Barlat '89-Voce
0 r—
-0.3 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Minor True Strain

Figure 2.7: Example of strain-based FLC predicted using M-K model for AA2008-T4

28
0.4
K = 537 MPa, n = 3.2845, R0 = 0.78, R45 = 0.485, Rg0 = 0.58, m=-0.003
0.35

0.3 A w °

« 0.25

s 0.2
1.
H
i.
o
0.15
o Experimental data
0.1
-»• fo=0.97, Hills 79, a=8-Swift

0.05 -A—fo=0.9805, Hills 79, a=8- Swift


• - fo=0.99, Hills 79, a=8- Swift

-0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Minor True Strain

Figure 2.8: Effect of changing f0 values on the FLC prediction of AA2008-T4 [7]

i Hill-Swift HBarlat-Voce

Figure 2.9: /„ values from M-K model that provide the best fit for the experimental data

29
In addition to the strain data generated from the M-K model, stress data is also

determined as shown in Figure 2.10. As with the strain-based FLC, the yield criterion

affects the stress-based FLC. Recall from Chapter 1 that the stress-based failure criterion

is less path dependent than a strain-based approach. Therefore, stress parameters from the

M-K model were assessed with the goal of creating a stress-based model to predict

necking concerns. The critical stress concentration factor (i.e., F- parameter) [13, 14]

which is the ratio of the effective stress in the safe region of the material to the effective

stress in the defect region was calculated at failure from M-K model results. Table 2.2

and Figure 2.11 show average F~ values with p > 0 for the eight material cases

investigated. There is a clear distinction between values of F- for steel and aluminum

cases (i.e., approximately 0.93 and 0.962-0.993 respectively). Note that F~ is not a

constant for all loading paths, which is reasonable as the stress concentration would vary

depending on the minor stress induced as shown in Figures 2.12 and 2.13. Also, the F-

parameter is not constant during the deformation process as shown in Figure 2.14 where

normalized depth was plotted to show the consistent trend of decreasing F- parameter

values as deformation progresses.

The F- values presented in Table 2.2 and Figures 2.11 and 2.13 are the ones at


failure (i.e., —— > 10) for the best fitting M-K analyses. There appears to be two values
d£XA

for the F- parameter at failure for positive and negative minor strain values as shown in

Figures 2.12 and 2.13. The average values ofF- for the positive minor strains are

reported in Table 2.2. Only positive minor strains were included so the orientation of the

30
defect (i.e., 9 in Figure 2.2) does not have to be considered. Recall that /„ values were

relatively constant for steel and aluminum cases (above 0.98). Thus this F- parameter

appears to be material dependent. Again in contrast to the /„ values, the F- values are

consistent for a given material model (i.e., values for Hill-Swift are between 0.920-0.937

for both steel and aluminum cases). As this F- parameter appeared to be promising, an

analytical stress-based model to predict necking failure was developed.

550
K=537 MPa, n=0.2845, R o =0.78, R 45 =0.485, R ^ O . 5 8 , A=426 MPa,
B=234.0MPa, C=5.130 ^^®
500
•* •

450

400

.Q^^F
350

9i
21 300
•*->

Qi
9 250 -
u
H
u
©
200 — •— von Mises, a=2 - Swift

150 ••<!•• Hill's 1949, a=2- Swift

— • • Hill's 1979, a=6 - Swift


100

—*—Hill's 1979, a=8 - Swift


50
—•— Barlat 89, a=8 - Voce
0 t -'-'r" ? 1 f r ' 1 i * 1 s

-50 0 50 100 150 200 250 300 350 400 450 500 550
Minor True Stress (MPa)

Figure 2.10: Stress-based FLCs predicted using M-K model for AA2008-T4 with various

yield criteria

31
• Hill-Swift a Barlat-Voce
1 -
0.98 -
0.96 - r-
1
0.94 -J
0.92 -
9
1 ,
F °" " "i

° 0.88 -
0.86 -
II .

- .*..
0.84 -
IS k
0.82 -
J
IS
SSl .
1 ' ^ :ti
0.8 m 'hJi

AISI304 SS
HC220YD

HC220YD

AA2008 T4
"3 •*
Steel(1)

H
Steel(2) OO
• * *

CO
1—1
NO H
< I—1
o I-H
< NO
NO

3 3
Figure 2.11: Average Fff for /? > 0 obtained from M-K model that provides the best fit

for the experimental data

1 -
K=537 MPa, n=0.2845, Ro=0.78, R45=0.485, Ro0=0.58,
A=426 MPa, B=234.0 MPa, C=5.130
0.975 -

i.
at
• * *

E
« 0.925 ~
CS
0.
Mr*^
ir — ®— von Mises
•••••• Hills' 1948
0.875 - —•> • Hill's 1979, a=6
<-#—Hill's 1979, a=8
—•—Barlat 89, a=8 -Voce
! I i i• i t
! %
-0.6 -0.4 -0.2 () 0.2 0.4 0.6 0.8 1 1.2
P

Figure 2.12: Effect of varying the yield criteria on the F- parameter using M-K model

and AA 2008-T4 data

32
1.05

-•—HC220YD Steel (1), Swift-Hill's


•••••• HC220YD Steel (2),Swift-Hill's
• - AK Steel, Swift-Hill's
-*—AISI 304 SS, Swift-Hill's
0.7 -A- -A12008-T4, Voce-Barlat
•••••• AA5182, Voce-Barlat
0.65 + - AA6016-T4, Voce-Barlat
A— AA6111-T4, Voce-Barlat
0.6 n r~ r r™ "i r~

-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 2.13: Effect of changing strain path on the F- parameter for different materials

K = 537 MPa, n = 0.2845, R =0.58, fo=0.9805, Hills 1979

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9


Normalized Depth at Failure

Figure 2.14: F- parameter versus normalized depth at failure obtained from M-K model

33
2.3. STRESS-BASED MODELS

2.3.1. Non-incremental effective stress ratio model

A non-incremental stress-based analytical model to predict stress-based and strain-

based FLC using the F- parameter was developed by Derov et al. [13, 14]. Other inputs

into this model include a constitutive relationship, yield criterion and anisotropic material

parameters. Like the M-K model, it assumes that an initial defect region is present in the

material, but instead of a thickness difference, a stress concentration as denoted by F- is

used (see Figure 2.15). The model also assumes that material parameters inside and

outside the defect region are the same.

The goal of the effective stress ratio model is to find the sets of <7XA and <72A that

satisfy compatibility, force equilibrium, work-energy relationships and the specified F-

value, which corresponds to failure (see Appendix C for equations). The stress u2A is

specified, while the stress <7XA is initially chosen. Thus, the stress path can be defined

o2A
(Ot ). The strains outside the defect can then be computed. First the effective
a
\A

stress is calculated:

| Defect Region
a2
1; ] Safe Region
/

°"i
A B A

Figure 2.15: Effective stress ratio model diagram [13]


34
^A=f(^A^2A) (2-23)

The relationship between the stress ratio (X and the strain ratio p in region A is

defined:

P = f(a) (2.24)

From the Ludwik-Hollomon power hardening law for example, the effective strain can

be defined:

£A = K (2-25)

The work-energy principle can then be used to obtain the major and minor strains in

region A (with proportional loading assumed):

°A£A
£lA (2.26)
°IA$+<*A PA)

_VA£A-VIA£IA
S
2A (2.27)
G
1A

All stresses and strains outside of the defect have now been obtained. To begin

computations inside the defect, the stress concentration factor, F-, is used to solve for aB

as:

B
~ F- (2-28)

Using Ludwik-Hollomon power hardening law and assuming the material parameters

are the same inside and outside the defect, £„ can also be calculated:

S
- A
S
B ~ —T (2.29)
te)-

35
Similar to the M-K model, it is assumed that compatibility is satisfied and £2 inside

and outside the defect are equal (£2A = ^ 2 B ) - Note that this model assumes proportional

loading in both regions A and B. With the knowledge of the effective strain and minor

strain within the defect, the major strain in the defect can be solved:

£
W=/{£B^2B) (2-30)

Note again that the incremental strain form of this equation is not used as proportional

loading is assumed in both the safe and defect regions. Thus this model is a non-

incremental approach. The strain ratio pB and stress ratio aB inside the defect can be

computed from similar relationships to Eq. (2.24), but for the region B.

Equations (2.26) and (2.27) can then be used inside the defect to obtain aXB and <J2B .

All parameters are now known, however these calculations can be performed for any

specified stresses (°'IA^(J2A) outside the defect. To determine if this stress state is

consistent with the physical condition, a force balance in the 1-direction (see Figure

2.15) is used to compare the assumed uXA to the one calculated (<ylA°) based on the stress

and strain values computed in the model.

^°=^iB^{eiB-^A) (2.31)

If these major stresses (crXA and <rM°) are not within a small error, the guess of <JXA is

modified and the process is repeated. When the values coincide, the conditions for failure

have been met and the values in the safe region are recorded as the failure stresses and

strains. The minor stress (<J2A) is then incremented to continue calculating additional

failure points on the stress-based FLC (see Figure 2.16).

36
The limit stress and strain values calculated depend on the choice of the yield criterion

used as shown in Figure 2.17. The key assumption in the effective stress ratio model is

the selection of the critical stress concentration factor (F- parameter).

A limitation of the original effective stress ratio model by Derov et al. [13, 14] was

that the defect orientation on the negative minor strain side of the FLC was not allowed to

vary. Furthermore, the model could be enhanced by incorporating higher order yield

criteria. These modifications to the effective stress ratio model were added during this

dissertation work and are discussed in detail in the next section.

o-i i
0"l A (failed)

1
1
1

y 0"lA(guess)

/ 1
0"2A a2

Figure 2.16: Methodology to increment G\A guess to determine a failure point on the

stress-based FLC [14]

2.3.2. Modified effective stress ratio model

As with the M-K model, the assumption of a perpendicular defect to the major

principal stress direction is not valid for the negative minor strain side of the FLC. Thus

there is a need to incorporate defect orientation into the effective stress ratio model. The

37
concentrated stress region orientation 0 is added in a similar manner to the modified M-

K model as described by Eqs. (2.15) - (2.17) (see Appendix C).

Compatibility, force equilibrium and work-energy balance equations are again

satisfied during this numerical solution. The analysis is repeated for different initial

orientations [0 ) of the defect in the range between 0° and 35° as shown in Figures 2.18

and 2.19.

K = 537 MPa, n = 0.2845, R« = 0.78, R45 = 0.485, R90 = 0.58, /^=0.85

0.75 0- Experimental Data 600 - —•- von Mises


—»e«• von Mises
550 - - - a - - H i l l ' s 1948
• Hill's 1948
0.6
500 •

450 -
.5 °- 45 \
U
+^
CO
St e 400 - •• •
3

H 0.3
u \ °\L vi 350 -
0 <u
ce °o\^ i l l l l l T *"°"r 3

s 0.15 3 ^ 300 -
0

^ 250 -

S i

-0.3 -0.15 0 0.15 0.3


200 250 300 350 400 450
Minor True Strain
Minor True Stress (MPa)

Figure 2.17: Effective stress ratio model results for AA 2008-T4 with two yield criteria

[13, 14]

38
von Mises, K = 537 MPa, n = 0.2845, 0.6 ~

Fff=09
0.5

0.4 ,S
'3s -
• * *

VI
0.3 «
— — 10 \ \ Sr
s-
02
••••••-20
- — 28
"• v

\
'I
§
- • -30
0.1
_ =.33
— . 35
0
-0.4 -0.3 -0.2 -0.1 0
Minor True Strain

Figure 2.18: Effect of adding a concentrated stress region orientation \0 )on the FLC

generated by the effective stress ratio model

600 - von Mises, K = 537 MPa, n = 0.2845,


£=0.9
500

400 -

•0
^300
VI
10
a>
s 20
£ 200
s- 28
o •30
"5* •33
gioo
35

100 200 300 400 500


Minor True Stress (MPa)

Figure 2.19: Effect of adding a concentrated stress region orientation \0 ) on the stress-

based FLC generated by the effective stress ratio model

39
The band orientation can be obtained either by using Hill's equation (tail \6) - -J— p )

(see Figure 2.2) or by the maximizing of <7XA value versus 9 . Note that 9 should be

added only to the negative minor strain side of the FLC and should remain zero on the

positive minor strain side of the FLC. Figure 2.20 shows (9 critical) calculated based on

Hill's equation and the one for the o~u maximum method. Finally, the prediction of the

FLC in the effective stress ratio model was improved by applying Hill's 1979 non-

quadratic yield function Eq. (2.18).

von Mises, K = 537 MPa, n = 0.2845, 0.6

0.5

o2*
0.4 .5
^ es
-*-
1_
C/3
0.3 «u
H
0.2 e

- - Theta Critical (Hill's)


s
0.1
—• • Theta Critical (Max ol A)
1 "!' i " i 0
-0.4 -0.3 -0.2 -0.1 ()
Minor True Strain

Figure 2.20: Comparison of perpendicular defect orientation (i.e., 9 = 0), Hill's equation

and <7U maximum method on the negative minor strain side of the FLC

2.3.3. Comparison between M-K and effective stress ratio models

In order to compare the M-K and the effective stress ratio models, the analysis

methodologies in both models should be reviewed first. In the M-K analysis, a groove

defect is added to weaken the material, create a strain concentration and to propagate the

40
failure. The strain paths in both the safe A and defect B regions are nearly identical

initially but begin to eventually diverge (point C in Figure 2.21). The safe region

continues deforming along the linear strain path pA, while the strain in the defect

accelerates along a nonlinear path pB. When the strain state in the defect reaches plane

strain (e.g., when dsie/deiA is greater than 10 at point E), it is assumed that localized

necking has occurred (i.e., no further deformation in the safe region occurs at point D).
* *
The strain values £1A and£2i4 , outside the neck, are defined as the limit strains.

E
n
*4
£x / m
/ •»
Region B strain path b / «f
(3) / «!
R p g i n n R strain f • * •
path (M-K) ' ! 1
/ *• *;
* * u*
'/ *• l
M
C & IB

1
' *
• Jr
yi*
' # r Jr *§
i *y
/ jj^ *
* *
/ y \4
#
t
/ M
y *
/ y 1 #
*
/ yp #
;y R egion A strain ;
jy path (M-K andF9)\ ,' 1
o .1
' k.

£2 w

<
'2 4.«
.J

Figure 2.21: Schematic of analysis methodologies in M-K and effective stress ratio

models

41
In the effective stress ratio model, a stress concentration defect is assumed in the B

region (i.e., oB>oA since ^ V - " * ^ ) - The strain paths in regions A and B diverge

immediately (as shown in Figure 2.21) since this model is non-incremental and assumes

proportional loading in both regions (i.e., from points (O to D) and (O to E) respectively).

Thus there is no failure propagation in the effective stress ratio model. The force balance

in the 1-direction is used to check for failure. When the major stresses (crXA and <JXA") are

within a small error (i.e., 1%), it is assumed that localized necking has occurred and the

values outside the defect are recorded as the failure stresses and strains.

The main difference between the two models is present in the condition of

proportional loading for the defect region B. For example in the M-K model, the work-

energy equation is satisfied in region B as:

d£ a = d£x ax + d£2 a2 (2.32)

While in the effective stress ratio model, the work-energy equation is satisfied as:

£(J=£]GX+£2G2 (2.33)

Equation (2.33) is true if and only if proportional loading holds. Thus, proportional

loading is assumed for the effective stress ratio model and a non-incremental approach is

used.

Obviously, the two models define the reaching to localized necking and failure in two

different ways. Figures 2.22 and 2.23 show strain and stress FLC of both models when

Hill's 1979 yield criterion is used. Two of the F- parameters found directly from the M-K

model were used (F-=0.929 for the equi-biaxial case and F- =0.898 for the plane strain

case). Despite this, significant differences in the FLC results were obtained indicating the

42
effect of the proportional loading assumption in the defect region and the use of a

constant F- parameter despite varying values with deformation (see Figure 2.14). Thus,

the effective stress ratio model is not able to accurately predict failure due to this

assumption.

0.5 K == 537 MPa, n = 0.2845, R=0.58 - o Experimental data


- - M-K, fo=0.9805
0.45 -
...... Derov, Fa=0.929
**••
0.4 — • Derov, Fa=0.898
\ \
- * -Derov, Fa=0.79
0.35 * V\
OH \ \
tE 0.3
SB
x \> \ \
X 5
^ 8

8 X * \ ''''•••.
IZJO.25
v
\ " ' \
s
£ 0.2 V
s
'a? *~~*> Q ^ ^ *

\ s %

0.15 ^
%

0.1

0.05

n , t i s i s i t r : !

-0.25 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 0.25 0.3
Minor True Stress (MPa)

Figure 2.22: Comparison between the strain FLC prediction from M-K and effective

stress ratio models for AA 2008-T4 [7]

43
500 K = 537 MPa, n = 0.2845, R=0.58
450
s e fe *
400 -

350 -

g 300
s-
35 250
o>
£200 • M-K, fo=0.9805
i-
•|l50
• Derov, Fo=0.929
100 H • Derov, Fo=0.898
50
Derov, Fo=0.79
0
0 50 100 150 200 250 300 350 400 450
Minor True Stress (MPa)

Figure 2.23: Comparison between the stress FLC prediction from M-K and effective

stress ratio models for AA 2008-T4

Note that a direct mathematical relationship between the M-K and the effective stress

ratio models can be found starting from force equilibrium:

FlA=F1B (2.34)

or with compatibility (i.e., d£2A -d£2B ):

a
\A %4 - a\B h (2.35)

By rearranging:

V (2.36)
\fAj

Defining the instantaneous thickness ratio \f):

44
f = -T=f0exp(e3B-e3A) (2.37)

Recall that f0 is the initial thickness ratio fo Eq. (2.36) can be rewritten as:
l
V 0A

°lA=°wf0VQ(SlB-eiA) (2.38)

From the definition of the ratio of the effective stress to the major principal stress:

l+ aa+R(l-a)a
<h = — = (2.39)
R+\

for Hills '79 non-quadratic yield function.

Equation (2.38) can be rewritten as:

(2.40)
<PA h

Force equilibrium equations in the effective stress ratio model are identical to that in

the M-K model, (i.e., Eqs. (2.34) - (2.36)):

^A=^W^M£3B-£3A) (2-41)

Note that Eqs. (2.41) and (2.38) are identical for f0 =1 (i.e., there is no initial

thickness difference) in the effective stress ratio model.

From the definition of the F- parameter (i.e., aA =F- aB) and Eq. (2.40):

^=fo^p(e3B-e3A)^- (2.42)
<PB

Equation (2.42) is a direct mathematical relationship between the key variables in the

two models (f0 and F-). For example, for the non-quadratic Hills '79 yield function:

45
1
f
\ + aA +R~(l-aA)a^
F
S = ~ = fo eX
P (£1B - £3A ) (2.43)
l + < +^(!-«fi)a

where a=6 for BCC materials and a=8 for FCC materials. Table 2.2 shows the F-

parameter values obtained using Eq. (2.43) when ccA=\.

2.3.4. Incremental effective stress ratio model

Since the non-incremental effective stress ratio model does not provide reasonable

results due to the embedded proportional loading in the defect area assumption (which is

not physically correct), an incremental approach was considered.

There are six inter-related variables in regions A and B regions

\d£x,d£2,d£, <Tj, <T2 and a). These variables can be calculated in region A by applying

the stress u1A and guessing the stress <JXA . As with the non-incremental effective stress

ratio model, the stresses and incremental strains outside the defect can then be computed

using Eqs. (2.23)- (2.27) presented earlier.

In essence for both the M-K and effective stress ratio models, the following equations

are applied to region B. In the M-K analysis, the computations start by knowing d£2B

from compatibility. There are six unknowns and six equations:

d£2B = d£2A Compatibility (2.44)

G
\A
&\B - T x Force equilibrium (2.45)

exp {e3B-e3A)

ag=/(ffg) Hardening function (2.46)

=
°B /(criB'cr2s) Yield function (2.47)
46
d£B(JB=d£XBoXB+d£2BG2B Work-energy (2.48)

and

,/ x= f( x Flow rule (2.49)

Equations (2.47)-(2.49) can also be written in terms of Ct and p .


In an incremental effective stress ratio model, one more condition is specified (i.e.,

aA = F-aB) which over-determines the system and makes an exact solution impossible

unless one of the above equations is not applied. For example, if the compatibility

condition can be removed (}.e.,d£2A ^d£2B), in theory, then the above set of five

~A
Equations (2.45)-(2.49) with the additional condition of c s - — could be solved. Figure
F
s

2.24 shows the result of this approach with an equi-biaxial strain path \pA = l) for say

the first step in the incremental approach with a step size of 0.1% strain. In this

approach, it was assumed that d£XB > d£XA and thus d£2B > d£2A since aB > aA . Therefore,

the strain path in the safe region is less than in the defect region (i.e., pB < pA ). This

would continue to persist with increasing deformation which is not correct physically.

Thus, an incremental approach to the effective stress ratio model that can satisfy all

the above mentioned equations does not provide a realistic solution. So far all attempts to

utilize the F- parameter in a generalized model that can predict FLC were not successful.

Therefore, an alternative F parameter is investigated as described in the next section.

47
0.00125 -
K = 537 MPa, n = 0.2845, Fs = 0.946, R = 1

0.001 -

a
y yy * ^*
r True Stra yy **
o
o
-o
d

Ul
o
yy **
y <»>
y *
yy +*
P

Ul
o
o
o
y *
yyy * * *
yy * *
0.00025 - y ^
/,+ - - - defect region B
s+y * _ safe region A
0 - J^r t • t ( i t i

0 0.00025 0.0005 0.00075 0.001 0.00125 0.0015


Minor True Strain

Figure 2.24: Strain paths from the incremental effective stress ratio model without

assuming compatibility

2.4. NON-INCREMENTAL, MAJOR PRINCIPAL STRAIN RATIO MODEL

A non-incremental approach to predicting necking failure is desirable to reduce the

computation time and complexity. To accomplish this, a critical major strain

£XA
concentration factor (i.e., FEX = — ) which is the ratio of the major true strain in the safe

region to the major true strain in the defect region at failure is used. Like theF- previously

investigated,F£] is calculated from the modified M-K model at failure. Table 2.2 and

Figure 2.25 show the least squares curve fit slope values of FeX for the positive minor

strain in the eight materials cases investigated. Only positive minor strains were

48
considered so the orientation of the defect (i.e., 9 in Figure 2.2) does not have to be

considered.

0.9 - • Hill-Swift H Barlat-Voce


0.85 -

0.8 -
;*;
4wA
0.75
1 •!:

l
pa
7
Fu °- I
0.65 - i
•i

0.6 •

0.55
I •

l
m m 1
n^ - I _ 31

AA6111 T4
AA6016 T4
AK Steel
HC220YD

HC220YD

AA2008 T4
AISI 304
Steel (1)

Steel (2)

SS

Figure 2.25: Least square slope of F£| values obtained from M-K model that provides the

best fit for the experimental data

FeX appears to be somewhat a material independent parameter as its values are

relatively constant (i.e., 0.74-0.85) for the steel and aluminum cases examined (Although

it is acknowledged that this is a large range of values). However, if Hill's 1979 yield

criterion is used, different values are obtained for the steel and aluminum cases (i.e.,

0.76-0.82 and 0.66-0.72 respectively). Figures 2.26 and 2.27 show the linear relationship

between the major true strain in the safe and defect regions for the right hand side of FLC

obtained from the M-K analysis. This linear relationship is the key assumption of the new

model. Note that there are intercept values as well for these linear curves, but these were

very small and insignificant, thus the intercept was ignored in this model for simplicity.

49
1 -

\ =0.8014x
< 0.8 -
,c
/S
Strain
as
/
V
S yy
xS^ > 0 767-lx
H 0.4 -
p
'S5
r
^ 0.2 - — Swift (Hills79 a=6)

— — Swift (von Mises a=2)


n-
() 0.2 0.4 0.6 0.8 1 1.2
Major True Strain in B

Figure 2.26: Linear fit of major principal strain data in the safe (A) and defect (B)

regions obtained from the M-K model for HC220YD Steel (1)

0.6 -i

-• Swift (Hill's 79 a=8)


• Voce (Barlat'89 a=8)
45
tc °" ^ . . . . . . Swift (von Mises a=2)
^y
A
)7202x

3 ir **
s-
-fc^
VI
% 0.3 J <y
u
y
H > - 0 738K jT
i-
^yy* y = 0.658lx
o ^0P"%&y
y ^
«
S 0.15 -

o,
c) 0.15 0.3 0.45 0.6 0.75
Major True Strain in B

Figure 2.27: Linear fit of major principal strain data in the safe (A) and defect (B)

regions obtained from the M-K model for AA2008-T4

50
The yield criterion and the hardening model affect the FeX values as shown in Figures

2.26 and 2.27. For example, in Figure 2.27 the linear fit provided poor matching to data

when higher order yield criteria were used.

The new non-incremental model predicts simultaneously stress-based and strain-based

FLCs using a constitutive relationship, yield criterion, anisotropic material parameters,

and the assumption of FeX. Like the M-K model, it assumes an initial defect region is

present in the material and represented in the form of a groove perpendicular to the

principal strain directions for positive minor strain values. The goal of the new analytical

model is to find the sets of <JXA and <J2A values, as well as £XA and £2A values, that result

in a major strain concentration factor, FeX, which corresponds to failure.

Again, there are six inter-related variables in regions A and B

\£x, £2, £, <Tj, o2 and a). Similar to the effective stress ratio model, these variables can

be calculated in region A by assuming the stress a2A and then guessing the stress <JXA.

The effective stress, the effective strain, the principal strains and the strain ratio pA

outside the defect can be computed using Eqs. (2.23) - (2.27) presented earlier.

All stresses and strains outside the defect have now been obtained. Similar to the M-K

model, it is assumed that compatibility is satisfied and d£2 inside and outside the defect

are equal (d£2A =d£2B). Using the same definition of failure from the M-K analysis

(d£XB=\0d£XA):

PB=^ (2-50)

51
the stress ratio in the defect region aB can now be calculated, e.g., for von Mises yield

criterion, from the strain ratio pB .

The work-energy principle can be used again to obtain the major and minor strains in

the defect region using:

£
B°'B ~ £\B®\B ~T~ £2B V^B^IB ) (2-51)

and the effective stress inside the defect cP can be determined from:

VB=VlB^-aB+al) (2.52)

for von Mises yield criterion. Thus Eq. (2.51) can be rewritten as:

eB = (2.53)
V(1-«B+«B)

Furthermore, Ludwik-Hollomon power hardening law can be used to relate the

effective stress and strain in region B (Eq. (2.25)), thus:

'<o i r
OxB^-aB+alB)
(2.54)
\Kj K

After substituting Eq. (2.54) into Eq. (2.53) and simplifying:

f V
£ £ a
\B + \ A p A B (2.55)
(*«) = * J_
2
{jl-aB+a B}|n+l J

Finally, from force equilibrium:

(<*U ) = (J\Bfo ™p (£3B ~ £ 3 A ) = Vwfo Q


*P (£IB ~ £U ) (2.56)

After substituting Eq. (2.56) into Eq. (2.55) and simplifying:

52
( \
(O £
\B +£
uPAaB
(2.57)
Kfoexp(£w-£u) 2
{jl-aB+a B} J

where £XB is the only unknown in this equation.

After solving Eq. (2.57) for£"1B, an ordered set oi{oXA,o~2A ) with the corresponding

ordered set of ( £XA, £2A ) is obtained. To determine the set of stress and strain values that

causes failure, the F£x predicted for the given (cr1A,(T2A ) values is compared to the Fe]

value shown in Figure 2.25 and Table 2.2. If this linear value is within a predefined

error, these values of ( £1A, £2A ) and the corresponding (<rXA, cr2A ) values are reported as

the limit stresses and strains respectively as shown in Figures 2.28 and 2.29.

700 -,
K = 537 MPa, n = 0.2845, R0 = 0.78, R45 = 0.48,
600 -
R90 = 0.58, m=0, fo=0.9805

500

u
<£ 400 ,....*••*
™@?^mh ^
s
u
H
u 300
o
•x-.. M-K model (Von Mises)
200
• - M-K (Hill's79 a=8)
100 -- »- -MPSRModel (von Mises, 5=0.99, FE 1=0.7202)

—• • MPSR Model (Hill's79 a=8, 8=0.99, FE1=0.6581)


0 100 200 300 400 500 600
Minor True Strain

Figure 2.28: Comparison between forming limit stress diagrams obtained from M-K

model and the major principal strain ratio (MPSR) model for AA2008-T4

53
Experimental data
•*••• M-K model (Von Mises)
• - M-K (HiU's79 a=8)
-*—MPSR model (von Mises, 5=0.999, FE 1=0.7202)
•* - MPSR model (von Mises, 5=0.99, FE 1=0.7202)
-*< • MPSR model (von Mises, 5=0.98, FE 1 =0.7202)
0.5
-»• MPSJR model (Hill's79 a=8, 5=0.99, FE 1=0.6581)

0.4

m^"
'3
s- 0.3
•**

VI
01
s
u
H
u
.©0.2
«

0.1

K = 537 MPa, n = 0.2845, R0 = 0.78, R45 = 0.48, R90 = 0.58, m=0, fo=0.9805

-0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 0.6


Minor True Strain

Figure 2.29: Comparison between FLC obtained from M-K model and the major

principal strain ratio (MPSR) model with varying yield function for AA2008-T4 [7]

Note that the linear fit hypothesis did not match well for higher order yield criteria and

the aluminum materials examined (see Figure 2.27). Thus, a linear tolerance (8) is

added to set a search region around the linear fit relationship between the major strains in

regions A and B. This 8 value is set as a tolerance or percentile to search around the

linear fit between the two strains for possible solution points (i.e., 100% for a perfect

zero-tolerance linear dependency). For example, if the linear tolerance o is set to 97%,

54
all the strain values deviating within 3% from the linear relationship will be considered as

a valid solution as shown in Figure 2.30.

The relationships between (p and (X) and ( a and Ct) are specific for each yield

criterion, so different results can be generated based on the yield criterion used. Hill's

1979 non-quadratic yield function (Eq. (2.18)) is applied in this model, which improved

the FLC prediction on the right hand side as shown in Figures 2.28 and 2.29. The FLC

prediction on the left hand side (shown in Figures 2.28 and 2.29) can be enhanced easily

by adding Hill's equation (tan(#) = yj— p ) to the major principal strain ratio model.

Another interesting feature in the major principal strain ratio model is that a solution

can be obtained even w h e n / , = 1 . For such a case the failure is physically caused by a

strain concentration rather than a difference in thickness or a groove. Note that the plane

strain case {pB = pA =0) is a discontinuity point for/ 0 = 1 , thus the major principal

strain ratio model algorithm does not necessarily converge to a solution near pA = 0 as

shown in Figure 2.31. Figure 2.32 demonstrates the sensitivity of new model results to

the change in FeX value. At plane strain case, as the value of FcX increases towards unity,

the major true strain value decreases towards zero.

55
All possible solutions points
...*... M-K model (Von Mises)

—*—MPSR model (von Mises, 5=0.999, FE 1=0.7202)

- «. -MPSR model (von Mises, 5=0.99, FEI =0.7202)


0.6
1_K . MPSR model (von Mises, 5=0.98, FEI =0.7202)

0.5

0.4
e
'3
u
vi

g 0.3
H
u
et
©
0.2

0.1
K = 537 MPa, n = 0.2845, R0 = 0.78, R45 = 0.48, Rgo = 0.58, m=0, fo=0.9805

-0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5


Minor True Strain

Figure 2.30: Comparison between FLC obtained from M-K model and the major

principal strain ratio (MPSR) model with varying 8 for AA2008-T4

56
0.6
K" = 537 MPa, n = 0.2845, 5=0.99, F E , = 0 . 7 2 0 2

0.5

a
*3 0.4
s.
-*•*

^-'\y^s
Vl
0) \
2 0.3
H
s-
§ 0.2 \S®%»*> M-K model (von Mises)
\^y _ _ MPSR Mo del (fo=0.97)
0.1 -
—• • MPSR Model (fo=0.9805)
• MPSR Model (fo=0.99)
A
+ MPSR Model (fo=l)
-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 0.6
Minor True Strain

Figure 2.31: Comparison between FLC obtained from M-K model and the major

principal strain ratio (MPSR) model with varying imperfection factor/ for AA2008-T4

0.7 .
••. K = 537 M 5 a, n = 0.2845, 5=0.99, fo=0.9805 .•
0.6 *•*

•S 0.5

|o.4
t'
s
^0.3 jBt*\ •••X"- M-K model (von Mises)
©
-Z ^j^f5>*'"* —•— MPSR Model (Fe 1=0.6)
g0.2 IjjgT _ * . MPSR Model (Fel=0.7202)
r —®—MPSR Model (Fsl =0.75)
0.1 - • - MPSR Model (Fel=0.8)
n —••• MPSR Model (Fel=0.85)
i i t t i i ! I

-0.5 -0.4 -0.3 -0.2 -0.1 C) 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
JVlinor True Strain

Figure 2.32: Comparison between FLC obtained from M-K model and the major

principal strain ratio (MPSR) model with varying the major strain concentration factor

FeX for AA2008-T4

57
2-5. CONCLUSIONS

• The models presented in this chapter (i.e., M-K, effective stress ratio and major

principal strain ratio models) are analytical models that can predict strain-based

FLC in sheet metal forming. But the key assumptions need to be verified for each

model.

• Results indicate that better fits are obtained if more advanced higher order yield

criteria are used with the three models.

• The proportional loading condition and the constant F- value in the effective

stress ratio model limit the application of this model in sheet metal forming

processes.

• An incremental approach to the effective stress ratio model that can satisfy the set

of seven equations with six unknowns was not found.

• The major principal strain ratio model has significant computational advantages

over the traditional M-K model since it is a non-incremental approach. Therefore

it can be programmed and integrated within a FEA package easily.

• Varying the defect angle can be included in the major principal strain ratio model.

• The key assumption of the major principal strain ratio model (i.e., major strain

concentration factorF£]) can be investigated experimentally, thus F£l tables can be

generated for different materials.

58
CHAPTER 3

NUMERICAL AND EXPERIMENTAL INVESTIGATIONS OF KEY

ASSUMPTIONS IN ANALYTICAL FAILURE MODELS

3.1 INTRODUCTION

In this chapter, the key assumptions in the M-K, effective stress ratio and major strain

ratio models are investigated for 1018 steel specimens with a thickness of 0.78 mm using

experimental and numerical data from Marciniak tests. In the experimental procedure,

Digital Imaging Correlation (DIC) is used to measure the major and minor in-plane

strains [44, 45]. The implementation of the DIC system provided the detection of the

strain concentrations during the experiments. The high strain concentration areas were

determined using a strain contour plot and a strain extraction grid. Strain components

were obtained on the grid points inside (i.e., the defect region) and adjacent (i.e., the safe

regions) to the high strain concentrations for four different strain paths (i.e., specimen

geometries). In the numerical analysis, FEA simulations with Marc Mentat were

performed with shell elements to investigate the four specimen geometries. The

incremental major strain ratio from M-K model, the critical stress concentration factor

from effective stress ratio model, and the critical major strain concentration factor from

major strain ratio model are investigated both numerically and experimentally. Thus, the

mechanics- and material-based failure phenomena in these three analytical models are

examined in this chapter to provide insight into the material behavior at failure.

59
3.2 METHODOLOGY

3.2.1 Experimental Procedure

Marciniak tests with the Raghavan modifications [46] were conducted on four

different geometries which produced varying strain paths from uniaxial to balanced

biaxial [44, 45]. These geometries are identified with Roman numerals (e.g., I for

uniaxial and IV for balanced biaxial). The four specimen geometries used were Type 1-1,

Type II-2, Type II-3 and Type IV (see Figures 3.1, 4.2 and Table 4.1 in Chapter 4 for

details). For biaxial cases, the failure location is controlled with a steel sacrificial washer.

As opposed to Nakajima tests where bending and friction are present, failure in the

central test region of Marciniak tests is dependent only on the material properties and

imperfections. Note that eight geometries were investigated in total but the other four

either produced redundant strain paths or the failure was not able to be controlled in the

central test area [46, 47].

Figure 3.1: Representation of main groups of geometries for the Raghavan modification

to Marciniak test [46]

60
The major and minor in-plane Lagrange strains were calculated from the correlation

analysis using VIC 3D software from Correlated Solutions Inc. (this is the DIC software

default tensor for the strain calculations). The area of highest concentrated major strain

was selected as the defect region. An "Origin and X-Axis" coordinate system

transformation was used to locate the origin at the point of highest concentrated major

strain for the digital image directly before a physical tearing failure was observed [44,

45]. The direction of the defect region was identified by changing the scale of the contour

plot and aligning the X-axis to the perpendicular direction to this region. From this

assumption, the Y-axis is assigned by the DIC system automatically (see Figure 3.2)

[45].

35 1
3
121

JF&
0 82

13 75 •R&- '•''
•ST'-f^-r
US' •»
-7 5 Est. ...-.-
Vlih * •

20 75
"V^. • *!

-50

-50 -22 5 32 5 60

Figure 3.2: Contour plot for locating X- and Y- axis on the specimen [45]

Following the coordinate transformation, strain data was obtained from the contour

plots for use in post-processing programs which distinguish between the safe and defect

regions in the material. Figure 3.3 demonstrates how the grid of the extracted nodes on

the specimen was constructed. Major and minor strain data were obtained from the DIC

software at a series of lmm-spaced nodes between ±25 mm in the X-direction. This

61
provided deformation data perpendicular to the defect as assumed in the three models. A

figure of the strain paths (major strain versus minor strain) for each node was created [44,

45] (see Figure 3.4).

i35
»* *
• * * . -

16 25

•2 5

•21 25

•40

-40 -2125 -2 5 16 25 35

Figure 3.3: Example of a grid of extracted nodes on the specimen [45]

K=488.5 M P a , n=0.215, R 0 =1.88, R4 5 =1.15, R 9 0 =2.23,


Type 1-1 a=2, fo=0.999

0.8 Type IV

'3
£ 0.6
(/}
v
s
i.
H
u .0.4
o

0.2 Experimental Data

FEA Data

M-K model, Hills 1948, Swift

-0.4 0 0.2 0.4 O.i 0.8 1


Minor True Strain

Figure 3.4: Strain paths varying from uniaxial to balanced biaxial cases for the various

specimen types

62
The major and minor Lagrangian strain values determined for the defect node (i.e., eis

and e2e) and the safe region, (i.e., eiA and e2A) were first converted to true strains (i.e.,

£IB, £2B and SIA, £2A) using:

£ = -]n{\ + 2eL) (3.1)

Note that Eq. (3.1) is used as an approximation for the true strain values. The "safe"

region refers to the area outside the defect node. From these values, the strain ratio of

d£2
minor true strain to major true strain (i.e., P = ~—) was calculated. The stress ratio
d£x

which is defined as the ratio of minor true stress to major true stress (i.e., a = ^±) was

calculated based on the p value assuming Hill's 1948 yield criterion:

1
1+- p+\
R
V QJ
a= (3.2)
( 1N
p+ 1+-
V ^90 J J

A strain ratio and stress ratio were calculated for every node at every instant of time

during the deformation [44, 45]. The critical stress concentration factor, F-parameter,

was then determined for all nodes. Assuming Hill's 1948 yield criterion and that the

stress normal to the sheet in the thickness direction, (cr 3 ), is negligible, the effective

stress and major true stress are related by:

l + R90 Res
= 11- a2- a (3.3)
1+
V -^90 J v *oy V1 +
*oy

This expression was then applied to both the defect and safe regions of the material,

and a ratio was created:

63
o\ 1+^90 R,
1+ a. a,
V ^90 J l + R,0 J 0 + R
oj (3.4)

a
H 1+
V
l + R90
^90 J
Rn
l + Roj
a. a.

By multiplying both sides of Eq. (3.4) by the ratio of the major true stresses in the safe

and defect regions, the definition of F- parameter, becomes:

l + R90 R, (2R± A
1+ «\-
V ^90 Jv
1+
^oy l + Roy a. f _ A
(3.5)
Rn 2*0 \aiB J
1+
V "^90 J l + R,oj a. v 1 + *oy
a.

The ratio of major true stresses in the safe and defect regions can be equated to the

respective true strains by considering force balance in the major direction (i.e., recall Eqs.

(2.34) and (2.35) presented earlier in Chapter 2)

A/1 ~~ MB (2.34)

°L4 \4 °"lB 'B (2.35)

and by assuming the same initial thickness exists prior to forming. Thus:

t = t0exp(-e3) = t0exp{-e2-el) (3.6)

By applying Eq. (3.6) to both the safe and defect regions and substituting it into Eq.

(2.35):

r exp(-s2B-s[B)
O =
exp(-e2A-£XA)
(3.7)

Finally, by substituting Eq. (3.7) into Eq. (3.5), the Fa parameter can be expressed as:

64
1+
1 + Rgo Y R0
aA-
{2R^
+ R
"A (
K, ^90 )\} o) exp \ £2B £
\B) (3.8)
Fs

i I + R90 Y R0 f20 Vexp(- £2A X


£
2 \A)J
1+ a B- aB
K *90 A 0, [i + Roj J

The material used in the experiments for both the specimen and the washer was AISI

1018. Mechanical properties for this material based on uniaxial tension tests are provided

in Table 3.1. The tensile tests were conducted in three orientations with respect to the

rolling direction of the sheet (0°, 45° and 90°) and each test was repeated twice as shown

in Figure 3.5. The plastic anisotropic ratios (R) of the sheet metal were measured in the

three directions (0°, 45° and 90°) using the ASTM E517 standard tensile test method (see

Table 3.1).

450 i

-0
-0'
-45
•• 45'
-90
-90'

0.15 0.2 0.25 0.35 0.4


True Strain

Figure 3.5: Tensile test data for AISI1018 used

65
Mechanical Properties (at 0°) Swift Model ! Anisotropic
Initial
Parameters (at 0°) i Parameters
Material | Thickness
UTS YS E Poisson's | K
t0 (mm) £o n m : Ro 1 R45; R90
(MPa) i (MPa) (GPa) Ratio i (MPa)
:
AISI 1018J 0.78 365 170 200 0.29 j 488.5 0.215!0.012J1.88; 1.15J2.23

Table 3.1: Mechanical properties of AISI1018 used

3.2.2 Numerical Simulations Details

The FEA simulations of the modified Marciniak test were conducted using MSC

MARC 2008 Rl. Taking advantage of symmetry, only one fourth of the model was

necessary for the simulations (as shown in Chapter 4, Figure 4.3). The four geometries

specified in Table 4.1 were modeled in Marc (i.e., Type 1-1, Type II-2, Type II-3 and

Type IV). The washer and specimen thicknesses were 0.78 mm. In these numerical

simulations, incompressible shell elements (Marc element type 139) were used with a

large strain analysis of the elasto-plastic material. For biaxial loading, the deformation

will be well beyond the strain limit observed in uniaxial tension. It is known that the

material behavior at these large strains will not simply follow the power hardening curve

from uniaxial tension [47, 48] and work hardening will be less. Therefore, an alternative

material behavior was sought. Figure 3.6 shows the hardening curve assumed in the

finite element simulations (i.e., the average curve). This average curve is the

mathematical average hardening effect between the power hardening law model

(i£=488.5 MPa, «=0.215) obtained from the tensile tests and the completely saturated

hardening material model (0 =377 MPa). This was a first guess approximation of the

material behavior for large strain values. Others have used similar techniques to

determine the hardening behavior at large strains [47, 48, 49 and 50]. The data points for

66
this average curve were entered into the numerical simulations to characterize the

material behavior. All numerical simulations were conducted assuming Hill's 1948 yield

criterion.

Figure 3.6: AISI1018 hardening curves

First order shell elements were used for all models with a uniform in-plane element

size of 0.933 mm x 0.933 mm for all specimens. Both the washer and the specimen were

treated as deformable bodies and the tooling was modeled as rigid bodies. The coefficient

of friction between all surfaces was set to u= 0.1 based on friction sensitivity analysis

tests [47, 50]. Three boundary conditions (BCs) were used in all the models, i.e., two

symmetry BCs on the symmetry edges and a fully constraining BC along the lock ring

edge (as shown in Chapter 4, Figure 4.4). This latter BC is used to prevent material

draw-in to the forming area as required for Marciniak tests. All nodes and elements

outside the lock ring were removed from the model to reduce the computational time

required.

67
3.3 RESULTS

The results of Marciniak tests for different geometries were investigated with respect

to the key assumptions in the M-K model (i.e., the incremental major strain ratio), the

effective stress ratio model (i.e., the critical stress concentration factor) and the major

strain ratio model (i.e., the critical major strain concentration factor). As expected, the

various specimen geometries produced a distribution of strain paths varying from

uniaxial to balanced biaxial (see Figure 3.4). As evident from this figure, deformation

paths from the numerical simulations were reasonably close to the experimental paths

thus validating the modeling effort. Also the strain-based and stress-based FLCs for

AISI1018 were generated using the modified M-K model and the average hardening data

(i.e., A=445 MPa, rc=0.17) as shown in Figures 3.4 and 3.7. In the numerical simulations

and by analytically converting the experimental strain paths to the stress space (see

Appendix D for the analytical conversion procedure), the various specimen geometries

produced also a distribution of stress paths varying from almost uniaxial to balanced

biaxial (see Figure 3.7). Thus, varying deformation states at failure could be assessed.

d£2
For example as shown in Figure 3.8, with increasing initial strain path (i.e., P — ——
d£x

which was taken at an approximate major strain of 0.1 (since the strain path is not

constant) and increases theoretically from -0.5 for uniaxial to 1 for balanced biaxial), the

trend is for the forming depth to reach failure to increase due to more material resisting

deformation. Recall that the sacrificial washer is not used for Type I specimens. The

sacrificial washer significantly increased the forming depth at failure for Type II

specimens compared to Type I specimens. In the experimental tests the specimens were

deformed till fracture, while the stopping criterion (i.e., necking) in the numerical

68
simulations was determined when the incremental major strain ratio reached the critical

value of 0.1 (except for Type IV specimen as shown later).

500 i
ype IV
Type II-2
Type 1-1

• Experimental Data

* FEA Data

M-K model, Hills 1948, Swift

0 100 200 300 400 500 600


Minor True Stress (MPa)

Figure 3.7: Numerical simulated stress paths varying from uniaxial to balanced biaxial

cases for the various specimen types

35

s
S 30 Type IV

"3 25
JS
a.
• * *

«
Q
6X3 ,,„
C 20 • Experimental Data
s
©
1X4 » «B» - FEA Data
Type 1-1
15
-0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8
Initial Strain Ratio

d£2
Figure 3.8: Initial strain path ratio ( P = —— ) versus forming depth at failure
d£,

69
Figure 3.8 shows that all experimental specimens reached higher forming depths

than the numerical models predicted. This is due to the deformation not localizing as

quickly or being as concentrated in the experimental specimens. Figures 3.9 - 3.12 show

the contour strain plots for the four geometries investigated both experimentally and

numerically. During the experiments, Type II-3 specimen did not fail in the center as it

should. It failed instead near the hole edge location on the sacrificial washer. In the

numerical simulation of this specimen, the failure location was in the center. Also note

from the contour plots that experimental Type IV specimens had an abnormal shaped

area of concentrated strain (i.e., three "lobed" areas due to the balanced biaxial stretching

that was induced and multiple failure locations being initiated). The orientation of these

lobes (i.e., perpendicular and parallel to each other) is due to the normal principal and

maximum shear loading directions. These multiple failure locations were not observed in

the numerical simulations of Type IV specimens.

3.3.1 Results for M-K Model Assumption

The M-K model assumes a narrow defect region on the specimen in which the

thickness is assumed to be less than the safe region. The strain paths presented in Figures

3.13 - 3.16 are for the node (i.e., location) with the highest major strain value which was

defined as the defect Figures 3.9 - 3.12 show the defect region orientations and

perpendicular axis directions for all specimen types. To define the safe region in the

experimental data, the strain paths for the nodes at distances of 2, 5 and 10 mm

perpendicular to the defect in the X- direction are plotted for different specimen types as

shown in Figures 3.13 - 3.16. Alternatively, the safe region in the numerical models was

70
determined by plotting the strain and stress paths for the nodes at distances of 2, 3 and 5

mm perpendicular to the defect in the X- direction as shown in Figures 3.13 - 3.16. The

difference between the distances in the experiments and the FEA simulations is due to the

varying location of the deformation in the two cases.

a) Type I-1 FEA Data

'™ if i

b) Type 1-1 Experimental Data 0 94

"" 0 61

0 27

-0 07

-0 4

Figure 3.9: Numerical and experimental contour plots of major true strain for Type 1-1

specimens for a) numerical simulation and b) experimental data [44, 45]

71
a) Type 11-2 FEA Data

b) Type 11-2 Experimental Data


0.93

0.63

: 0.32

0.01

-0.3

Figure 3.10: Numerical and experimental contour plots of major true strain for Type II-2

specimens for a) numerical simulation and b) experimental data [44, 45]

72
a) Type II-3 FEA Data

b) Type II-3 Experimental Data


0.75

0.57

0.38

0.19

Figure 3.11: Numerical and experimental contour plots of major true strain for Type II-3

specimens for a) numerical simulation and b) experimental data [44, 45]


73
a) Type IV FEA Data

Lobe b) Type IV Experimental Data


0 83

0 70

: 0 57

0 44

03

Figure 3.12: Numerical and experimental contour plots of major true strain for Type IV

specimens for a) numerical simulation and b) experimental data [44, 45]

74
Conceptually, the divergence of the strain path curves for the defect location in the

M-K model should occur as observed in Figure 3.16 a) for Type IV specimens with the

safe region 5 mm away from the defect in the experimental data. That is, the defect and

safe regions follow the same strain path but then the defect region curve diverges just

prior to failure. Note that the divergence of the 10 mm away from the defect location

curves in the experimental data occurred early in the deformation process due to the

localization of the defect region Also, note that the location of safe region in numerical

models is related to the mesh size (i.e., the 2, 3 and 5 mm locations are approximated due

to the nodal locations at failure). For the experimental uniaxial cases (i.e., Figures 3.13

and 3.14), 10 mm away from the defect region, the divergence was not observed and the

defect and safe regions follow the same strain path throughout the process even up to

failure. For Figure 3.15 of Type II-3 specimens, the safe and defect regions do not follow

the same strain path determined experimentally even early in the process. However, for

the numerical simulations, the strain paths at various safe locations are the same initially

but then diverge.

75
a) Type I-l Experimental Data
-i 1
1
1
i 0.8
»

Strain
I

as
o
!
i s
I-
i H
0.4 u
\ ' p
- - Defect \ 'IT
. . . . . . . 2mm \ 0.2
_ . 5 m m \
— 10 mm \
0
-0.6 -0.4 -0.2 0
Minor True Strain

b) Type I-l FEA Data 1 1


1
i 0.8
i
1 a
1 '3
0.6 i
i VI
Major True

\
o

\
— — Defect \
i
. . . . . . . 2 mm \
o

— • 3 mm \
— — 5 mm \
t i « 0
-0.6 -0.4 -0.2 0
Minor True Strain

Figure 3.13: Strain path curves for the defect node and different locations away in the X-

direction for Type I-l specimens for a) experimental [44, 45] and b) numerical simulation

data
76
a) Type II-2" Experimental Data |
s !

i
i
i - 0.8

True Strain
as
o
0.4 o
— — Defect 1
. . . . . . . 2 mm 1
0.2
—™ • 5 mm 1
——— 10 mm \
! ! i 0
-0.6 -0.4 -0.2 0
Minor True Strain

1
b) Type II-2 FEA Data
1
1 . 0.8
1
i
'3
• 0.6^
4>
I s
i u
1 0.4 Ho
i '3°
— — Defect \
. . . . . . . 2mm \ - 0.2
spspsssra. ti T, r Y l t T l »

«=*«,_*„ 5 m m 1
i i r " i 0
-0.6 -0.4 -0.2 0
Minor True Strain

Figure 3.14: Strain path curves for the defect node and different locations away in the X-

direction for Type II-2 specimens for a) experimental [44, 45] and b) numerical

simulation data

77
1
a) Type II-3 Experimental Data
0.9

0.8
.1
0.7 •I
•l
0.6 1:1
0.5 Hi
"Jl
0.4
- Defect
0.3
•• 2 mm
0.2
• 5 mm
0.1
—10 mm
0
-0.3 -0.2 -0.1 0 0.1 0.2 0.3
Minor True Strain

1
b) Type II-:i FEA Data
0.9 -

0.8 - a
j
o
'-<t
Strain

-
OS
p

§0.5
u
H
u0.4

|0.3 II — — Defect
0.2 - i ••••••» 2 mm
f —» . 3 m m
0.1 -
1 ——5 mrn
A
! ! ? i i !

-0.3 -0.2 -0.1 0 0.1 0.2 0.3


Minor True Strain

Figure 3.15: Strain path curves for the defect node and different locations away in the X-

direction for Type II-3 specimens for a) experimental [44, 45] and b) numerical

simulation data

78
1 a) Type IV Experimental Data

/
0.8 /
a
'3
35 0.6 -
s
u
H
fe 0.4 -

Jf — — Defect
0.2 - y • • 2 mm
y —— • 5 mm
y ——». jo mm
0 -r « i t r i

() 0.2 0.4 0.6 0.8 1


Minor True Strain

1 -1 b) Type IV FEA Data

0.8 -
s f
et
Vt 0.6 -
s>-
H
£ 0.4 -

s y _ — Defect
0.2 - y ....... 2 mm
y — • 3 mm
0 -y —__5 mm
C 0.2 0.4 0.6 0.8 1
Minor True Strain

Figure 3.16: Strain path curves for the defect node and different locations away in the X-

direction for Type IV specimens for a) experimental [44, 45] and b) numerical simulation

data

79
In the M-K model, dsXB > \Ode1A defines the failure. Figures 3.17 - 3.20 show the

—L§_\ for the specimen types with respect to the defect


incremental major strain ratio (de
ds
K IA

node and the location 5 mm away from the defect location in the experimental data and 2

mm away in the numerical models. The difference in the defect location position is due to

the more concentrated deformation localization in the numerical simulations compared to

the experimental results (see Figures 3.9-3.12). In the experimental data, as the initial

strain path increases, the incremental major strain ratio for the image directly before the

tearing failure was observed decreased (e.g., approximately 12 for the uniaxial Type 1-1

case and 2.5 for Type IV balanced biaxial case) (see Figures 3.17 and 3.20). Thus, the

incremental major strain ratio is not a constant value which is often assumed when

implementing the M-K model. However, the line trajectory just prior to failure in these

figures is nearly vertical, thus indicating the localization of the deformation in the defect

region. While in the numerical simulation, all the models were able to reach to the

incremental major strain ratio of 10 except for the Type IV balanced biaxial case (see

Figure 3.20 and 3.21), the stopping criterion for this model was set to be the incidence of

direct contact between the punch and the test specimen due to the excessive sacrificial

washer hole expansion.

80
10 i
• FEA Data
• Experimental Data
et
«
fi
2 6
vi

B 4
a
a 2

10 11 12 13 14 15 16 17 19
Forming Depth (mm)

Figure 3.17: Incremental major strain ratios versus forming depth for Type I-l specimen

for experimental [44, 45] and numerical simulation data

10 - - - - FEA Data ^
T^ • j. 1 T-\ j.

L/Xpcnmcntal Data
8
i-
Incremental Strain
as

j
I
4^

I
i
k '
fO

C\ -
i i j

10 12 14 16 18 20 22 24 26 28 30
Forming Depth (mm)

Figure 3.18: Incremental major strain ratios versus forming depth for Type II-2 specimen

for experimental [44, 45] and numerical simulation data

81
10 FEA Data
• Experimental Data

20 22 24 26 28 30 32 34
Forming Depth (mm)

Figure 3.19: Incremental major strain ratios versus forming depth for Type II-3 specimen

for experimental [44, 45] and numerical simulation data

FEA Data
3.5
• Experimental Data
©
3
'"g
.5 2.5
et
u
VI „

a
*• 1.5

i .Ur^'V^H^^
0.5

10 12 14 16 18 20 22 24 26 28 30 32 34
Forming Depth (mm)

Figure 3.20: Incremental major strain ratios versus forming depth for Type IV specimen

for experimental [44, 45] and numerical simulation data

82
'3
u
vi
"3
*J
a
o
S
t
u
S
"3
B

-0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1 1.2


Initial Strain Ratio

Figure 3.21: Initial strain path versus incremental major strain ratio directly before

failure for experimental [44, 45] and numerical simulation data

3.3.2 Results for Effective Stress Ratio Model Assumption

The effective stress ratio model uses a key assumption that relates failure to a critical

stress concentration factor (i.e., F- parameter), which is a ratio of the effective stress in

the safe region to the effective stress in the defect region. Since the stress values at the

defect location are always expected to be higher than the other areas, the F- parameter is

always less than unity.

Experimental data was analyzed at four different images before failure (i.e., 0.75-3

seconds before failure for the Type I-l specimen and 1-4 seconds for the Type II and the

Type IV specimens) [44, 45] which corresponds to a given distance before failure. The

reason for choosing different time amounts was to capture the failure and strain effects

between the subsequent images (i.e., approximately 1% strain difference between

images). As with the M-K analysis, the highest major strain location was selected as the

defect node. The effective stress value at the defect node was divided into the effective

83
stress value at the given X-direction location as shown in Figures 3.9 - 3.12. Thus, at the

X-distance of zero (i.e., the defect node) the F- parameter was unity. Figures 3.22 a) -

3.25 a) show plots of the X-direction location versus F- parameter. The five curves are

labeled with respect to the punch location prior to failure (e.g., see Figures 3.22 where

the punch would travel 0.36 mm between images). Numerical data was also analyzed at

four different increments prior to failure and are included in Figures 3.22 b) - 3.25 b).

In Figure 3.22 a), for the experimental uniaxial case, the F- parameter becomes

relatively constant after a distance of 10 mm away from the defect location. This could

represent the safe region of the specimen. For other specimen geometries, the F~

parameter continues to vary beyond 10 mm; thus, the "safe" region is not as easily

defined. Note that for Figure 3.23 a), for the Type II-2 specimen, the F5 parameter

increases after 15 mm away from the defect. In Figure 3.25 a) (i.e., the balanced biaxial

geometry), the .F-parameter is very close to unity for -10 to 10 mm in the X-direction

from the defect. This is due to the large defect area as observed in Figure 3.12.

The numerical simulation data shows the same trends observed from the experiments

(except for the increase of Fd parameter in Type II-2 specimens). However, the safe

region of the specimens was closer to the defect location. This is consistent with the safe

region evaluation for the M-K analysis parameter in Figures 3.13-3.16. Also the value

that the Fs parameter leveled off to is different. For example for the Type I-l specimens,

the F- value leveled off at 0.7 for experiments while for numerical simulations, the value

was much lower (i.e., 0.5). This again is due to the more localized defect in the FEA

results.

84
Figure 3.22: Critical stress concentration factor (i.e., F- parameter) versus the X-

direction location for Type I-l specimens at various distances from failure for a)

experimental [44, 45] and b) numerical simulation data

85
Figure 3.23: Critical stress concentration factor (i.e., F~ parameter) versus the X-

direction location for Type II-2 specimens at various distances from failure for a)

experimental [44, 45] and b) numerical simulation data

86
a) Type II-3 Experimental Data

l.l

^^2»v

0.9 - ^l^«^
0.8 -
'• Fa -2.356 mm
0.7 - tfBt
~ Fa-1.767 mm
JS!

• Fa-1.178 mm
0.6 -
-Fa-0.589 mm
— Fa Fracture
—e^
-20 -15 -10 -5 0 5 10 15 20
Distance from defect node in X-Direction (mm)

b) Type II-2 FEA Data

Fa-1.023 mm
Fa -0.737 mm
0.2 - Fa -0.520 mm
•Fa-0.243 mm
• Fa Failure
-e-
-20 -15 -10 -5 0 5 10 15 20
Distance from defect (mm)

Figure 3.24: Critical stress concentration factor (i.e., F- parameter) versus the X-

direction location for Type II-3 specimens at various distances from failure for a)

experimental [44, 45] and b) numerical simulation data

87
Figure 3.25: Critical stress concentration factor (i.e., F- parameter) versus the X-

direction location for Type IV specimens at various distances from failure for a)

experimental [44, 45] and b) numerical simulation data

88
F- parameter values were also compared with respect to different specimen

geometries with strain paths varying from uniaxial to balanced biaxial. Figure 3.26

shows for both experimental and numerical data that as the initial strain path increases

(e.g., Type I-l for uniaxial and Type IV for balanced biaxial), the width of the defect

region increases. This effect is also physically reasonable. As the width area of the

specimen increases, the defect region for the specimen also increases.

a) Experimental Data
L2
1 ~-mw&^^ «_
^ ^ ^sn> ^ ^

0 8 -I ^
u
• * *

0.6 -
B -~^^ I-l
et
•u 0.4 - TypeII-2
et
OH
e — - TypeII-3
0.2 -
fa — —- Type IV
0 „
__
-20 -15 -10 -5 0 5 10 15 20
Distance from defect (mm)

•Type I-l
Type 11-2
Type II-3
Type IV

^_
-20 -15 -10 -5 0 5 10 15 20
Distance from defect (mm)

Figure 3.26: F- parameter versus distance from the defect node in the X-direction for

different specimen types for a) experimental [44, 45] and b) numerical simulation data
89
Figure 3.27 shows the F- parameter values with respect to the initial strain path in

the "safe" region at both 5 and 10 mm away from the defect location for the experimental

data (Exp) and 2 and 5 mm away for the numerical simulation data (FEA). As the initial

strain path increases in the specimen, the F- parameter value increases.

1.2 -,
Type II-3 Type IV
Type II-2

Type I-l y^^Z— • *"**


uv
•**

E
et
s-
et
C,
C
fa 5 mm away from the defect (Exp)
2 mm away from the defect (FEA)
10 mm away from the defect (Exp)
5 mm away from the defect (FEA)

-0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1 1.2


Initial Strain Path

Figure 3.27: Critical stress concentration factor versus initial strain path for both

experimental (Exp) [44, 45] and numerical simulation (FEA) data

3.3.3 Results for Major Strain Ratio Model Assumption

The major strain ratio model uses a key assumption that relates failure to a critical

major strain concentration factor (i.e., FeX parameter) which is the ratio of the major true

strain in the safe region to the major true strain in the defect region. Since the strain

values at the defect location are always expected to be higher than the other areas, the

FeX parameter is always less than unity.

90
Experimental data was analyzed at four different images before failure (i.e., 0.75-3

seconds before failure for the Type I-l specimen and 1-4 seconds for the Type II and type

IV specimens). To be consistent with the M-K and the effective stress ratio analyses, the

highest major strain location was selected as the defect node. For all specimens, the

defect node had the FeX parameter of unity. Figures 3.28 a) - 3.31 a) show plots of X-

direction location versus FeX parameter. The five curves are labeled with respect to the

punch location prior to failure (e.g., see Figures 3.28 a)). The punch would travel 0.36

mm between images. Numerical data was analyzed at four different increments prior to

failure. Figures 3.28 b) - 3.31 b) show the four curves labeled with respect to the punch

location prior to failure.

In Figure 3.28 a), for the experimental uniaxial case, the FeX parameter becomes

relatively constant although continues to decrease slightly after a distance of 10 mm away

from the defect location. This could represent the safe region of the specimen as was the

case in the effective stress ratio model. For other specimen geometries, the FsX parameter

continues to linearly decrease beyond 10 mm; thus, the "safe" region is not as easily

defined. Note that for Figure 3.31 b) (i.e., the numerical simulation of the balanced

biaxial geometry), the Fei parameter is very close to unity up to 12 mm in the X-

direction away from the defect. This is due to the large defect area as observed in Figure

3.12. The numerical simulation data shows the same trend observed from the experiments

but the safe region of the specimens is closer to the defect location in a similar trend to

the effective stress ratio model data. Again the values that the major strain ratio model

levels off at are different with lower values for the numerical simulations than the

experimental results.

91
Figure 3.28: Critical major strain concentration factor (i.e., FeX parameter) versus the X-

direction location for Type I-l specimens at various distances from failure for a)

experimental and b) numerical simulation data

92
Figure 3.29: Critical major strain concentration factor (i.e., F£X parameter) versus the X-

direction location for Type II-2 specimens at various distances from failure for a)

experimental and b) numerical simulation data

93
Figure 3.30: Critical major strain concentration factor (i.e., FeX parameter) versus the X-

direction location for Type II-3 specimens at various distances from failure for a)

experimental and b) numerical simulation data

94
a) Type IV Experimental Data

^ ^ * ^ T ° 5 L * 'M^t .as^

owe** 1 *
"J^-^^^^^ 0.8
1

• * *
1
<u
0.6
B
et
s-
e« . . . . . . . p E l -2.273 mm
&- 0.4
U
- - F E I -1.705 mm
fa I
1
— • F£l-1.137mm
0.2
— • — F E I -0.568 mm

— — F E I Fracture

-20 -15 -10 -5 0 5 10 15 20


Distance from defect (mm)

b) Type IV FEA Data


... .,....- .,. ,.. .4-.

ijiy 0.8 -
\ ^
u
o
0.6 -
B
et
u
et
OH . . . . . . . pel -0.876 mm
0.4 -
•H
W "=• ^ F E I -0.658 mm
fa
— • F E I -0.432 mm
0.2 -
— • - F E I -0.211 mm
1
' F E I Failure
is

-20 -15 -10 -5 0 5 10 15 20


Distance from defect (mm)

Figure 3.31: Critical major strain concentration factor (i.e., FeX parameter) versus the X-

direction location for Type IV specimens at various distances from failure for a)

experimental and b) numerical simulation data

95
Figure 3.32: F£] parameter versus distance from the defect node in the X-direction for

different specimen types for a) experimental and b) numerical simulation data

96
Fsl parameter values were also compared with respect to different specimen

geometries with strain paths varying from uniaxial to balanced biaxial. Figure 3.32

shows for both experimental and numerical data that as the initial strain path increases

(e.g., Type I-l for uniaxial and Type IV for balanced biaxial), the width of the defect

region increases.

Figure 3.33 shows the F£X parameter values with respect to the initial strain path in the

"safe" region at both 5 and 10 mm away from the defect location in the experimental data

(Exp) and 2 and 5 mm away in numerical simulation (FEA). As the initial strain path

increases in the specimen, the FEX parameter value increases in a similar trend to the

effective stress ratio model data. However, the Fs] parameter values seem more scatter.

Type IV

u
4t
S
es
;_
et

w
fa 5 mm away from the defect (Exp)

— • * 2 mm away from the defect (FEA)

D— 10 mm away from the defect (Exp)

O" - 5 mm away from the defect (FEA)

-0.6 0.2 0.4 0.6 1.2


Initial Strain Path

Figure 3.33: Critical major strain concentration factor versus initial strain path for both

experimental (Exp) and numerical simulation (FEA) data

97
3.4 DISCUSSION

The experimental results show that the key assumptions in the M-K, effective stress

ratio and major strain ratio models are not completely supported by the experimental data

of the tests described in [44, 45]. The incremental major strain ratio to indicate failure for

the M-K model is expected to be a constant value for every strain path. However, as

shown in Figure 3.21, the incremental major strain ratio decreases with increasing initial

strain path values in experimental data. Note, however the nearly vertical slope of the

curves just prior to failure in Figures 3.17 - 3.20. Thus, while the incremental major

strain ratio values were not a constant value (e.g., dsXB > 1 OdsXA ), only a small amount of

further deformation would have been necessary to achieve this value.

For the key assumption in effective stress ratio model, the critical stress concentration

factor was not constant at a given X-direction location for the various initial strain paths

either (see Figure 3.27). This is due to the saturation (i.e., leveling off) of the stress-strain

relationship at high strain values. Despite the higher strain values in the defect region as

shown in Figure 3.12, the calculated critical stress concentration factor did not decrease

significantly with the X-distance. This created near unity F- parameter in both the

experiments and numerical simulations for the balanced biaxial case. Furthermore, the

width of the defect region increased as the strain path varied from a uniaxial to balanced

biaxial case. For the Type IV case, the F- parameter was near unity over a 20-30 mm

range for the experimental results, and near unity over a 5-7 mm range for the numerical

simulation. This is much larger than a thin deformation zone assumed in all three

analytical models (see Figure 3.26).

98
For the key assumption in major strain ratio model, the critical major strain

concentration factor was not constant at a given X-direction location for various initial

strain paths either (see Figure 3.33). Since only two positive initial strain paths were

tested during the experimental and numerical tests, the hypothesis of a linear relationship

between the major stains in regions A and B for p > 0 cannot be completely verified and

more experiments are required to test this theory.

Thus, for none of the models considered (i.e., M-K, the effective stress ratio or the

major strain ratio models) did a constant parameter to predict failure exist. The M-K

model though had a more consistent location of the safe region (approximately 5 mm

away from the defect node in the experimental data and 2 mm away from the defect node

in the numerical simulations) and reasonable size of the defect region compared to the

effective stress ratio model which had a defect size of approximately 20-30 mm for the

experimental balanced biaxial case. Defect localization was also distinguishable in the

major strain ratio model as shown in Figure 3.32.

Note that for these comparisons of the M-K, the effective stress ratio and the major

strain ratio models assumptions, data was obtained to represent failure at the image

directly before a physical tearing failure was observed. However, plastic instability would

have occurred prior to this image. This, particularly, would have affected the results for

the M-K analysis as the deviation of the strain path in the defect region signifies plastic

instability. However, this would decrease the incremental major strain ratio at failure and

a typical value to predict failure is already lower than expected (i.e., <10) for all
d£XA

geometries except the uniaxial case.

99
3.5 CONCLUSIONS

In this chapter, the key assumptions in the M-K, effective stress ratio and major strain

ratio failure models were investigated by conducting Marciniak tests and numerical

simulations on AISI 1018 steel to determine the related parameters of interest. These

models are based on the assumptions regarding material behavior and the mechanics of

the deformation processes, and thus provide insight into failure. Strain was measured

experimentally in the process using Digital Imaging Correlation (DIC) and was converted

to a critical stress concentration factor (i.e., F- parameter) assuming Hill's 1948 yield

criterion, force equilibrium perpendicular to the defect region, and conservation of

volume. Strain paths generated from the Marciniak tests produced a reasonable

distribution between uniaxial and balanced biaxial cases.

For the M-K model, while a concentration of major strain existed in the defect region

compared to the safe region, the incremental major strain ratio at failure was not constant

but decreased as the initial strain path for the specimen increased. However, the trajectory

of the incremental major strain ratio curves was such that a plane strain state existed in

the defect region at failure. The safe region was shown to be approximately 5 mm away

from the defect region in the experimental data and 2 mm away from the defect region in

the numerical simulations. Note that the typical means to conceptualize failure in the M-

K model (i.e., the defect region strain path diverging from the safe region strain path)

does not occur for specimen geometries with uniaxial and plane strain paths.

For the effective stress ratio model, the F- parameter was not constant for the various

initial strain paths. Also, the size of the defect varied, which is expected due to the

varying specimen geometries. The saturation of effective stress for the stress-strain

100
relationship caused the F~ parameter for the balanced biaxial case to not decrease

significantly when considering locations perpendicular to the defect region. But the F-

parameter was able to characterize the localization of deformation, the size of the defect

region, and distinction between the safe and the defect regions.

For the major strain ratio model, the hypothesis of a linear relationship between the

major stains in both the safe and defect regions was not completely verified and more

experiments are required to verify this theory. But the FeX parameter was able to

characterize the localization of deformation, the size of the defect region, and distinction

between the safe and the defect regions in a similar trend to F- parameter.

101
CHAPTER 4

EFFECT OF ELEMENT TYPES ON FAILURE PREDICTION

USING A STRESS-BASED FORMING LIMIT CURVE

4-1. INTRODUCTION

Finite element analysis (FEA) is being used extensively in industry as another tool to

predict failure location during the forming process [49-56]. Literature review shows

extensive use of shell elements in sheet metal forming [55-60]. For example, Galbraith et.

al [56] compared the cost-benefits of seven four-noded shell elements included in LS-

DYNA3D element library for the Numisheet 93 square benchmark problem. Yoshida et

al. [58] performed a deformation analysis of dome test using shell elements. The limiting

cup height and rupture location were predicted accurately for mild steel, high-strength

steels, austenitic stainless steel and aluminum alloy.

Solid-shell elements have, also been used, for in sheet metal forming simulations [61-

68]. For example Cardoso et al. [62, 63] presented examples of the successful

implementation of solid-shell elements theory in sheet metal forming simulations,

including large deformation anisotropic and isotropic plasticity with friction, and

consistent numerical solutions compared to experimental results were provided.

On the other hand, 3D continuum elements were not so commonly utilized in the sheet

forming processes due to the large demand required in the computational time [69-70].

For example, Wriggers et al. [69] made a detailed comparison between modeling shells

by three-dimensional brick elements and simple shell elements. It has been shown that

102
the enhanced strain brick element yields accurate solutions and can be applied even to

very thin shells. Jung and Yang [70] presented a comparison of shell and solid elements

with experimental data for plane-strain conditions. Their comparisons indicated a close

agreement of all analyses with experiments in the pure stretching. Several comparisons

have been made between different element types [56, 59, 69, and 70] and the results, in

general, show that higher accuracy is achieved when using 3D continuum elements over

the shell elements.

This chapter presents a comparison between the strain, stress paths and the FLCs

generated using three element types: shell, solid and solid-shell elements. Through

numerical simulations and assuming a stress FLD, the results are compared to

experimental work to ensure that the numerical models are predicting the actual material

deformation and failure during the forming process. The results show that for the in-plane

stretching conditions examined; only minor differences exist in the FLDs generated with

each element type.

4-2. MARCINIAK TEST BACKGROUND

The Marciniak stretch test was designed to overcome the severe strain gradients

through the thickness that develop due to friction, normal punch pressure, and bending in

traditional dome tests [46, 47]. Without the influences of these parameters which are

difficult to quantify, Marciniak tests are more sensitive to material defects and can

provide more accurate and repeatable results that are useful for FEA simulations. In these

tests a sacrificial washer is inserted between the specimen and the punch to minimize the

strain localization under the punch edge. Also, a recessed punch surface is used in order

103
to minimize the friction in the central pole as shown in Figure 4.1. A lock ring is used to

prevent material flow into the forming area.

-PUNCH

100 mm (3 937") -

TOP BINDER
REMOVABLE
LOCK RING
-RECESSED
POCKET
n
-SACRIFICIAL
40 mm WASHER
RJ2 mm (0 472")
(1 575") L
R12mm(0 472'
~TT7\ -SPECIMEN
U
BOTTOM BINDER

110 mm (4 331")

Figure 4.1: Modified Marciniak test tooling

Raghavan developed and tested seventeen specimens configurations [46], and by

changing the specimen and the washer geometries he covered linear strain paths ranging

from uniaxial (i.e., ej = -2ei) to balanced biaxial stretching (i.e., £/ = ei). The width and

notch radius of the specimens as well as the hole diameter of the washers were varied to

ensure that strain localization and failure occurred within the in-plane straining region of

each sample. Seven specimens were selected from Raghavan's work for the numerical

simulations (see Figure 4.2 for the four major types of specimen geometries and Table

4.1 for the specimen details). All specimens and washers have a length of 177.8 mm. The

washer can be made from any stock steel or other metal with equal or better formability

104
than the specimen material. This assures that specimens fail before the edge of the

enlarged hole cracks or tears. Also, the washer thickness must be equal to or larger than

the specimen thickness if the same material is used for the specimen and washer;

otherwise fracture can occur in the washer.

Type I Type II Typem TypelV

177.8 * - W s •*. 177.8 177.8

177.8
Rs
c 177.8

Washer Geometry Dia W


I R« I DiaW
None 177.8
6 177.8 J ^ Q Q 177<8

Dia W

Figure 4.2: Specimen and washer geometry types

Specimen Effective Specimen Notch Radius Washer Hole Diameter


Type Width W s (mm) Rs (mm) W(mm)
Type I-l 25.4 76.2
No washer used
Type 1-2 50.8 63.5

Type II-l 120.6 40.64


No notch
Type II-2 139.7 38.1

Type IH-1 120.65 28.575 40.64

Type in-2 133.5 22.15 38.1

Type IV 177.8 No notch 30.25

Table 4.1: Seven specimen and washer dimensions used [46]

105
4-3. NUMERICAL SIMULATIONS DETAILS

The FEA simulations of the modified Marciniak test were conducted using MSC

MARC 2008 R l . Taking advantage of symmetry, only one fourth of the model was

necessary for the simulations as shown in Figure 4.3. The seven geometries specified in

Table 4.1 were modeled in Marc and twenty one simulations were conducted (seven with

shell element, seven with solid element and seven with solid-shell element). The washer

and specimen thicknesses are 0.8 mm. In all simulations, incompressible elements were

used with a large strain analysis of the elasto-plastic material. In this chapter, the material

used in the numerical simulation for both the specimen and the washer is hot dipped high

strength interstitial free steel HC220YD. Tensile tests and the experimental strain-based

FLD data were provided in the Numisheet 2008 benchmarking for this material [35]. The

mechanical properties of the HC220YD steel are £=200 GPa, v=0.29 with anisotropic

parameters of i?0=1.28, R45=2.08 and Rgo=l.$6. A power hardening law with parameters

of K= 678.3 MPa and rc=0.2258 was assumed. Numerical simulations were conducted

with Hill's 1948 yield criterion, and the yield loci for this material are (7.«/70=O.956 and

7p(/7o=1.005). The X-axis (see Figure 4.4) was considered parallel to the rolling

direction in all models.

First order elements were used for all models with a uniform initial in-plane element

size of 0.933 mm X 0.933 mm for all specimens. The number of elements and nodes are

provided in Table 4.2 for comparison between FEA models. Both the washer and the

specimen were treated as deformable bodies and the tooling was modeled as rigid bodies.

Since the area of interest in the Marciniak tests (i.e., the central region) has only in-plane

stresses and no bending, two elements per thickness were sufficient as the stress and

106
strain distributions through the thickness were uniform. Mid-plane nodes were used to

analyze the strain and stress values. The coefficient of friction between all surfaces was

set to ju= 0.1 based on friction sensitivity analysis tests performed by others [47, 50].

Three boundary conditions (BCs) were used in all the models, i.e., two symmetry BCs on

the symmetry edges and a fully constraining BC along the lock ring edge as shown in

Figure 4.4. This former BC is used to prevent material draw-in to the forming area as

required for Marciniak tests. All nodes and elements outside the lock ring were removed

from the model to reduce the computational time required.

Punch " * -•
*•

; ; A
*iJ^^^^^P''. .'••';^.t*WS F ^'
* \ \ - '''-.' •>.;• '—'•- $ ••.:•"• ^.*>

Bl"*ftjSSw

I'.'.V:.:^ t J-
t

flWBv*-s Lock ring


• "MX* 'i Type I specimen
4

/ 4>

Die

Figure 4.3: FEA model showing the punch, die, locking ring and specimen

4-4. ELEMENT SELECTION GUIDELINES

The shell element used in this chapter (element type 139 according to MSC MARC

notation) is a Kirchhoff-theory based 4-node bilinear thin-shell element with full

integration. The bilinear interpolation is used in this element for the coordinates,

displacements and rotations. The membrane strains are obtained from the displacement

field and the curvatures from the rotation field so that six degrees of freedom per node are

107
available. Integration through the shell thickness is performed numerically using

Simpson's rule and eleven integrations points though the thickness [71, 72].

Y-symmetry BCs

-X-symmetry
BCs

\
( 011M1 iininsi IK •>

(Kollni!! di'LLtion)

Figure 4.4: Boundary Conditions used in the FEA model for Type I specimen

The solid element used (Marc element type 7) is a 8-node three-dimensional

hexahedral continuum brick with full integration. This element uses trilinear (linear in all

three directions) interpolation functions; thus, the strains are constant throughout the

element. This results in a poor representation of shear or bending behavior. However, this

is not a concern in the central region of interest for Marciniak tests. The constant

dilatation integration procedure was used for this problem to prevent potential element

volumetric locking [71]. Higher order elements are not recommended for modeling large

plastic deformations because they are susceptible to volumetric locking when simulating

incompressible materials. Our test simulations support this observation. As a means to

demonstrate this, higher order quadratic elements were used and resulted in unstable

behavior of the model at strains higher than 30%. However at small strains, the

108
predictions of linear and high order elements models were identical. Also, higher order

elements should not be used in this type of problem due to the contact involved.

The solid-shell element used (element type 185) is a shell element of eight-node brick

topology with reduced integration. The element uses enhanced assumed strain

formulation for transverse normal component (i.e., E33) and transverse shear components.

The stiffness matrix of this element is formed using one integration point in the element

plane and a user defined number through the element thickness (with Marc's default of

seven layers selected). In this way the element can capture accurate material plasticity

under bending load. This element gives more accurate results than the classical shell

elements in applications that require double-sided contact [71, 72].

For all specimen types mentioned in Table 4.1 (except Type III), the element located

on the symmetry lines of the specimen experienced the highest strain and stress in the

model and thus was of interest to investigate for failure. For Type III specimens, the

element of interest was shifted by one row from the line of symmetry. The major and

minor stresses and strains that represent the deformation path for that element were then

plotted. Various strain paths from uniaxial to equibiaxial were obtained from the seven

geometries. All numerical simulations were conducted using Intel quad core 17- 950

processor (8MB of Cache, 3.06 MHz). Relative computational times are reported in

Table 4.2 for element type comparison.

4-5. COMPARISON BETWEEN ELEMENT TYPES

Figure 4.5 shows the experimental FLC from Numisheet 2008 [35] and the strain

paths obtained for the element in the mesh at the mid-plane which experienced the most

109
severe deformation. All the strain paths have crossed the experimental FLC. There are no

significant differences observed in the trend for the three element types when strain

values are below the FLC strain values. Strain paths for specimen Type II-2 shows some

minor deviation when e/=0.35 and Type III-2 shows some deviation throughout. More

obvious differences between the strain paths are observed above the FLC which indicates

strain localization and failure. Figure 4.6 shows the stress paths and the experimental

stress-based FLC (which was analytically converted from the experimental strain FLC,

see Appendix D for details regarding this analytical conversion). The stress paths show

more deviation than the strain paths with slight oscillation due to dynamic effects and

switching of contact conditions in the numerical simulations. In order to compare the

failure predicted by the stress-based FLD to the experimentally determined strain-based

FLDs, the stress-based FLD can be used to generate a strain-based FLD in the numerical

simulations. Crossing of the stress-based FLD determined the failure in the simulations

(as opposed to, for instance, a certain strain state, a certain thinning ratio, or any other

strain-based failure criterion). The strain state corresponding to failure in the stress state,

i.e., the point on the stress-based FLD obtained using linear interpolation, was plotted in

order to obtain a single point on the corresponding strain-based FLD. This was repeated

seven times for the seven different geometries for all three element types considered to

create seven points on the three strain-based FLCs. The strain-based FLC from the

numerical simulations was then generated by simply connecting the seven points with a

line. This numerically generated strain-based FLD was then compared to the

experimental strain-based FLD. Figure 4.6 and Table 4.3 show the values of true stress

110
components and the associated true strain components when the stress paths are crossing

the experimental stress-based FLC.

Row (1) Shell, (2) Solid, (3) Solid-Shell


Specimen Type
Number of nodes Number of elements Relative computational time

4193 4001 1

Type I-l 12579 8002 3.67

8002 4186 1.63

5793 5601 2.00

Type 1-2 17379 11202 6.16

17379 11202 4.12

12173 11856 17.45

Type II-l 22679 13311 21.41

30097 17745 20.96

14428 14088 23.95

Type II-2 34887 20590 68.90

26577 15622 27.83

9981 9662 23.37

Type III-l 23817 15336 51.90

23817 15336 37.59

10317 9998 22.87

Type III-2 24923 16068 51.26

S 24923 16068 35.73

10653 10328 22.51

Type IV 34088 19921 106.77

34088 19921 45.87

Tab le 4.2: Number oi "nodes and elements in FEA models

111
It is clear that good prediction is achieved except for specimen Type I-l, 1-2 and IV

(uniaxial and biaxial stretching). The maximum deviation in the major strain between the

element types occurred for the plane strain case (Type III-1) with a value of 0.058 and the

uniaxial tension case (Type I-l) with a value of 0.091. Figure 4.8 shows the punch

£2
displacement at failure versus the initial stain ratio (P - — ) . The solid-shell element
*i

shows a continuous trend of lagging behind both the shell and the solid elements for all

the geometries tested with a larger punch displacement required to indicate failure. Using

the reduced integration in the solid-shell element leads to a softer element and thus

volumetric locking in the plastic range has to be taken into account. This problem can be

solved by increasing the number of layers through the thickness of the solid-shell element

as shown in Figure 4.8.

For specimen Type III-2, an artificial geometrical imperfection (see Figure 4.9) was

introduced in the central four elements of all models (i.e., shell, solid and solid-shell) by

changing the element thickness from 0.8 mm to 0.76 mm (reduction by 5%). Adding this

imperfection was necessary so the stress path intersected with the experimental stress-

based FLC. Figure 4.10 shows the effect of reducing the thickness on the stress path for

the shell model. The shifting of the deformation path is due to the localization of the

deformation in that region. A similar effect was observed for all three element types. The

element thickness of 0.76 mm was selected since less reduction has little effect on the

stress path and thus failure predication. Note that the stress-based FLC was linearly

extended (see dotted portion of curve in Figure 4.5) so Type III-2 and IV specimen

intersected.

112
Failure location was changing depending on the specimen geometry. For example

specimen Types I-l, 1-2, II-1, II-2 and IV the failure locations were at the central element

along the lines of symmetry. However, for Type III-l and III-2, the failure locations were

shifted by one row of elements, as shown in Figure 4.11. This shift appears to be related

to the distribution of stresses for this particular geometry.

0.8

II

Figure 4.5: Stain paths generated using the three element types with the experimental

Forming Limit Curve

113
Figure 4.6: Stress paths generated using the three element types with the experimental

stress-based FLC

0.8 -

0.7 V -

0.6 Njt

0-5 "N^
V
w'0.4 \ - ,1-^y^^

-* j ^ ^
Vs
^**
0.3 ^—Experimental FLC
0.2 esi^OraB&p ^ h f * 1 1

0.1 - - — Solid
^^^^ Solid- Shell
U i i i- ?

-0.4 -0.3 -0.2 -0.1 () 0.1 0.2 0.3 0.4 0.5 0.6 0.7
£2

Figure 4.7: FLC generated using the three element types with the experimental FLC
114
55

50

0)
45
9
et
«M
• * *
4.0
et
• * *

eD
C
'••A
S 35 f-/
uet

30
M
t/t • shell
'6,'/
e>~- Solid
25 */ / • Solid-Shell 7 layers
Solid-Shell 11 layers
20
•Solid-Shell 15 layers

15
-0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.1 1

P =(E2/El)

Figure 4.8: Punch displacements at failure versus the strain ratio for seven specimen

types

~r - V ' 5*." > *" * i • I ~ - "at * "• '"**' < *\* y


»

•*• *- y y "X ^ y v >. v > < -

v x ' sc jy N ^ A ^ v * - > -f y
v
*>- ' . ' .*• '.•< ^ , * *

fi< Imperfection

Figure 4.9: Location of the imperfection added to the mesh for specimen Type III-2

115
Row (1) Shell (element 139), (2) Solid (element 7), (3) Solid-Shell (element 185)
Specimen Type
o, (MPa) C2(MPa) ; Ei | 62

751.5 199.3 0.715 -0.372

Type I-l 740.5 172.5 0.670 -0.352

735.9 162.5 0.624 -0.340

768.8 265.3 0.667 -0.310

Type 1-2 76778 !'" ~ " 252J 0.656 -0.306

772.8 236.3 0.639 0.310

769.3 341 0.556 -0.203

Type II-l 766.5 348.4 0.523 -0.189

770.8 323.4 0.528 0.197

426.8 750.4 0.424 -0.094

Type II-2 4071 ; 7544 0.441 -0.092

399.1 756.6 0.418 0.105

780.2 568.4 0.283 0.005

Type III-l 785.3 588.6 0.261 0.003

781.4 578.4 0.319 0.003

863.2 791.7 0.449 0.354

Type III-2 866.5 801.4 0.439 0.34

873.2 821 0.422 0.328

896.2 882.5 0.633 0.626

TypeFV 896.3 879.5 0.654 0.648

904.8 904.5 0.704 0.699

Table 4.3: True stress and strain components values reported at crossing of the stress

paths with the experimental stress-based FLC

116
1000

950

900 -

850 -

800

FLC
750 -
Element Thickness of 0.75 mm
Element Thickness of 0.76 mm
700 Element Thickness of 0.77 mm
Element Thickness of 0.78 mm
Element Thickness of 0.8 mm
650
650 700 750 800 850 900 950 1000
o7 MPa

ure 4.10: Effect of imperfection thickness (shell element) for specimen Type III-2

Figure 4.11: Location of failure for specimen Type III-2

117
4-6. DISCUSSION

In general, the strain paths generated utilizing the three different element types did not

show a significant difference below the strain-based FLC. The shell element showed the

ability to deform more than the solid and the solid-shell element, while the stress paths

had minor differences between the three elements. For example, stress paths generated

using the solid-shell element were more uniform and had less "zigzagging" than the other

elements indicating that solid-shell element is less sensitive to contact problems. The

FLCs generated using the three element types are in good agreement with the

experimental FLC and indicate that there is no preferred element for this problem. The

shell element-based FEA predictions of the forming depth at failure have been

experimentally validated for AISI1018 steel, as shown in Chapter 3.

However, usage of the shell element reduced the computational time considerably: by

474% and 204% (for the biaxial case) as compared to the solid and solid-shell elements,

respectively. Thus, because of this advantage in computational time, shell elements are

often used when modeling sheet metal forming operations.

Note that Marciniak tests produce a nearly linear strain path in the process. Kinsey et

al. [73] showed that the stress-based failure criterion could be used in numerical

simulations to predict failure for a square cup application. A restriking process was used

to create a sharper planar radius that could not be achieved in a single forming process.

Due to the path dependence of the strain-based FLD, an accurate prediction of the

forming severity could not be produced with a strain-based criterion. Thus, while only

linear strain paths were used in this research, similar results would be expected if non-

linear strain paths were induced as was the case in Kinsey et al. [73].

118
4-7. CONCLUSION

The effect of changing the element type has been investigated in numerical

simulations of the Marciniak test, and a stress-based failure prediction criterion was

successfully implemented.

The choice of an element for sheet-forming simulations must balance the

computational requirements against the desired accuracy of the results. Differences were

seen in the simulation time and performance of the three elements tested in MARC's

element library. If computational time is the dominant criterion for element selection,

then the shell element is a reasonable choice. More reliable predictions for FLD are

obtainable at a reasonable computational time with the solid-shell element. The stiffness

of this element can be controlled through the number of layers through the thickness and

15 layers shown to be reasonable for this problem.

119
CHAPTER 5

CONCLUSIONS

Sheet metal forming is a key manufacturing process for several industries including

automotive and aerospace. Critical to the successful implementation of sheet metal

forming processes is the accurate prediction of material failure through numerical and

analytical models. In this dissertation, research efforts in this area were described. In

Chapter 2, three main analytical models were discussed: the modified Marciniak and

Kuczynski (M-K) model which included a varying defect orientation with respect to the

principal stress directions, the effective stress ratio model and the major strain ratio

model. The M-K predictions of the FLC were demonstrated with several yield criteria for

eight different materials. The non-incremental stress-based FLC theory of the effective

stress ratio model was described, and its extensions to include varying defect orientation

and more advanced yield criteria were presented. It was clear that the embedded

proportional loading condition and the constant critical stress concentration factor value

in the effective stress ratio model limit the application of this model. An incremental

approach to the effective stress ratio model that can satisfy the set of all equations was

not found. Also, the original non-incremental FLC theory of the major strain ratio model

was presented, and this model was shown to reasonably match data from proportional

experiments in both stress and strain space. This model has significant computational

advantages over the traditional M-K model since it is a non-incremental approach.

In Chapter 3, a study was conducted using Marciniak tests, AISI1018 steel and the

shell element numerical models to numerically investigate the key assumption of the

120
three analytical models. The results show that strain paths generated from simulating

Marciniak tests produced a reasonable match with the experimental data. The parameters

investigated to predict failure (i.e., the incremental strain ratio, critical stress

concentration factor and critical strain concentration factor) were not constant for the

various strain paths for the three analytical models considered. For the M-K model, while

a concentration of major strain existed in the defect region compared to the safe region,

the incremental strain ratio at failure was not constant but decreased as the initial strain

path for the specimen increased. However, the trajectory of the incremental strain ratio

curves was such that a plane strain state existed in the defect region at failure. For the

effective stress ratio model, the Fd parameter was also not constant for the various initial

strain paths. Also, the size of the defect varied, which is expected due to the varying

specimen geometries. Despite the higher strain values in the defect region as observed

with the experimental data, the saturation of effective stress for the stress-strain

relationship caused the F- parameter for the balanced biaxial case to not decrease

significantly when considering locations perpendicular to the defect region. But the F-

parameter was able to characterize the localization of deformation, the size of the defect

region, and distinction between the safe and the defect regions. The results of this

numerical study were consistent with the past work [44, 45] except that the safe region

was shown to be approximately 2 mm away from the defect region for all specimen

geometries and "lobed" defect areas existed for the experimental balanced biaxial strain

path cases, which is likely due to defects in the material. For the major strain ratio

model, the hypothesis of a linear relationship between the major stains in both the safe

and defect regions was not completely verified and more experiments are required to

121
verify this theory. But the FeX parameter was able to characterize the localization of

deformation, the size of the defect region, and distinction between the safe and the defect

regions in a similar trend to F- parameter.

Finally in Chapter 4, finite element analysis (FEA) was used in the prediction of FLC,

and a comparison between three different element types (shell, solid and solid-shell) was

conducted in detail. The numerical results show that despite the differences in stress

distribution assumptions, shell, solid and solid-shell elements would not provide

differences in failure prediction when a stress-based failure criterion is used to simulate

the Marciniak stretch test. The choice of an element for simulating this test must balance

the computational requirements against the desired accuracy of the results. If

computational time is a dominate criterion for element selection, then the shell element is

a reasonable choice

The goal of this dissertation was to develop analytical and numerical tools that can

accurately predict failure in sheet metal forming operations. Hopefully the work

presented adds to the knowledge in this area and will be useful to the sheet metal forming

community.

122
CHAPTER 6

FUTURE WORK

6.1 ANALYTICAL MODELING IMPROVEMENTS

There are new advanced higher order yield criteria that are able to describe materials

behavior more accurately e.g., the behavior of aluminum alloys in the M-K analysis [27,

31, 32 and 40]. These higher order yield criteria would require an additional

transformation technique or iterative procedure to solve for the stress and strain ratios

which would cause the computations to become more intensive, but not impossible.

Work has already begun on the incorporation of Banabic BBC2000 and Barlat YLD2000

in the major strain ratio model.

6.2 NUMERICAL MODELING IMPROVEMENTS

As mentioned previously in Chapter 3 and Chapter 4, Hill's 1948 anisotropic yield

criterion was used in the numerical simulations of steel specimens. This yield criterion

does not accurately describe the behavior of aluminum alloys and there is a need to

incorporate other more advanced yield criteria (i.e., YLD2000 and BBC2000) if

aluminum alloys are going to be examined numerically.

Narasimhan and Wagoner [26] presented an interesting finite element model to

simulate in-plane forming limit diagrams of sheets containing finite defects without using

a contact problem (i.e., no punch or die were used). A specimen similar to Figure 2.1

was modeled on their study to investigate the key assumptions in the M-K model on the

right hand side of the FLC. Using such a FE model can reduce the simulation time

123
significantly and can be used to simulate the cruciform specimens in biaxial testing. A

similar finite element model for cruciform specimens is currently under development.

In addition, Abaqus EF commercial FEA package has a built-in M-K analysis that can

be used in the key assumptions verification and to make a comparison between the M-K,

effective stress ratio and major strain ratio models.

6.3 EXPERIMENTAL VERIFICATIONS

For this dissertation, only steel specimens were tested numerically for different

geometries. Aluminum specimens should also be tested for all strain paths in order to

comment on the analytical models compatibility for different materials.

In addition, the biaxial testing machine should be ready in near future (see Figure

6.1). This will allow testing the specimens under different principal stresses without using

different geometries. Also, the deformation path can be altered during the test to

investigate the path independence of the stress-based failure criterion. The major strain

ratio model will incorporate future findings from the experiments to be performed in near

future. The biaxial loading stage must be assembled and tested to validate the design's

integrity. Furthermore, a batch of cruciform specimens will be fabricated. Preliminary

tests using the biaxial loading stage coupled with DIC could be used to characterize the

behavior of the cruciform specimens prior to running experiments to evaluate their

performance under variable loading conditions, both linear and non-linear deformation

paths.

124
Figure 6.1: Biaxial testing apparatus [74]

The results of this research raise questions about the findings from the comparison

between analytically predicted stress and strain based FLCs and experimental data which

was presented in Chapter 3. In particular, Chapter 2 results suggest that the major strain

ratio model could be further developed to improve predictions overall. Thus, more

experimental testing using varied specimen geometries is needed to understand the

effects of strain path on the critical major strain concentration factor. Furthermore, the

results from future testing of varied specimen geometries coupled with DIC will produce

strain path data which could be used to further evaluate the key assumptions in the major

strain ratio model.

125
REFERENCES
1. Schmid S., Kalpakjian S., (2010), "Manufacturing Engineering and Technology" 6th

edition, Pearson Prentice Hall, ISBN: 978-0136081685.

2. "Metalworking: Sheet Forming", (2006), ASM Handbook, vol. 14B, ASM

International, ISBN: 978-0871707109.

3. Keeler S.P., Backofen W.A., (1964) "Plastic instability and fracture in sheets

stretched over rigid punches", ASM Trans. Quart. 56 25.

4. Hosford W. and Duncan J. (1999), "Sheet metal forming: A review", Journal of the

Minerals, Vol 51, pp. 39-44.

5. Marciniak Z., Duncan J., Hu S.J. (2002), "Mechanics of Sheet Metal Forming" 2nd

edition, Elsevier, ISBN: 978-0750653008.

6. LeRoy G. H. and Embury J. D., (1978), "Formability: Analysis, Modeling and

Experimentation", Ghosh A. K. and Hecker H. L., eds. (AIME, 1978), pp. 183-207

7. Graf, A., Hosford, W., (1993), "Effect of Changing Strain Paths of Forming Limit

Diagrams of Al 2008-T4," Metall. Trans. A, 24A, pp. 2503-2512.

8. Marciniak Z., Kuczynski K. (1967), "Limit strains in the processes of stretch-

forming sheet metal", Int. J. Mech. Sci. 9 609-620.

9. Yao H., Cao J. (2002) "Prediction of Forming Limit Curves Using and Anisotropic

Yield Function with Prestrain Induced Backstress", International Journal of

Plasticity, Vol. 18, pp. 1013-1038.

126
10. Geese H., Dell H., Kebler L., Hooputra H., Werner., (2001), "Continuous Failure

Prediction Model for Non Linear Load Paths in Successive Stamping and Crash

Processes", Society of Automotive Engineers.

11. Stoughton, T. B., (2001), "Stress-Based Forming Limits in Sheet-Metal Forming" J.

Eng. Mater. Technol., 123, pp. 417^122.

12. Sakash A., Moondra S., Kinsey B., (2006), "Effect of Yield Criterion on Numerical

Simulation Results Using a Stress-Based Failure Criterion" J. Eng. Mater. Technol.

vol. 128, Issue 3, pp. 436- 444.

13. Derov M. J., Kinsey B., Tsukrov I., (2008), "The effect of model parameters on

predicted stress based failure criterion for sheet metal", Proceedings of the

International Manufacturing Science and Engineering Conference MSEC- ASME,

October 7-10, 2008 Evanston, IL, USA.

14. Derov M. J., (2008), "Investigation of a stress based failure criterion for sheet

metal", Master's Thesis, University of New Hampshire.

15. Keeler SP, Brazier WG, (1975), "Relationship between laboratory material

characterization and press-shop formability". Proceedings of the Micro Alloying

Conference, New York, NY, pp. 21-32

16. S.B. Levy, (1996), "A comparison of empirical forming limit curves for low carbon

steel with theoretical forming limit curves of Ramaekers and Bongaerts", IDDRG

WG3, Ungarn, 13-14 June 1996.

127
17. Swift H.W., (1952), "Plastic instability under plane stress". Journal of the

Mechanics and Physics of Solids 1, pp. 1-18.

18. Banabic D., Dannenmannb E., (2001), "Prediction of the influence of yield locus on

the limit strains in sheet metals", Journal of Materials Processing Technology, vol.

109, Issues 1-2, pp 9-12.

19. Hill R., (1952) "On discontinuous plastic states, with special reference to localized

necking in thin sheets", J. Mech. Phys. Solids 1 pp. 19-30.

20. Storen, S., and Rice, J. R., (1975), "Localized Necking in Thin Sheets," J. Mech.

Phys. Solids, 23, pp. 421-441.

21. ZHU X., Klaus W., Abhijit C, (2001), "A unified bifurcation analysis of sheet metal

forming limits", Journal of Engineering Materials and Technology 2001, vol. 123,

no3, pp.329-333

22. Grafand A., Hosford W. F. (1990), "Calculations of forming limit diagrams", Met.

Trans. 21A, 87.

23. Hutchinson, J.W., Neale, K.W. (1978) "Sheet Necking-III. Strain-Rate Effects, in

Mechanics of Sheet Metal Forming", eds. D.P. Koistinen and N.M. Wang Plenum.

New York, pp. 269-283.

24. Barata Da Rocha, A., Barlat, F., Jalinier, J.M., (1984), "Prediction of forming limit

diagrams of anisotropic sheets in linear and nonlinear loading", Mat. Sci. Eng. pp.

68, 151.

128
25. Butuc M.C., Gracio J.J., Barata da Rocha A., (2003), "A theoretical study on

forming limit diagrams prediction", Journal of Materials Processing Technology

142,714-724.

26. Narasimhan, K., Wagoner, R.H., (1991), "Finite element modeling simulation of in

plane forming limit diagrams of sheets containing finite defects", Metall. Trans. A

22,2655.

27. Butuc M.C., Gracio J.J., Barata da Rocha A., (2002) "A more general model for

forming limit diagrams prediction", Journal of Materials Processing Technology

125-126,213-218.

28. Ahmadi S., Eivani A.R., Akbarzadeh A., (2009), "An experimental and theoretical

study on the prediction of forming limit diagrams using new BBC yield criteria and

M-K analysis", Computational Materials Science 44, 1272-1280.

29. Ganjiani M., Assempour A., (2007), "An improved analytical approach for

determination of forming limit diagrams considering the effects of yield functions",

Journal of Materials Processing Technology 182 ,598-607.

30. Ganjiani M., Assempour A., (2008), "Implementation of a Robust Algorithm for

Prediction of Forming Limit Diagrams", Journal of Materials Engineering and

Performance, vol. 17, no. l.pp 1-6.

31. Cao J., Yao H., (2000), "Prediction of localized thinning in sheet metal using a

general anisotropic yield criterion", International Journal of Plasticity 16, 1105-

1129.

129
32. Banabic D., Comsa S., Jurco P., Cosovici G., Paraianu L., Julean D., (2004),"FLD

theoretical model using a new anisotropic yield criterion", Journal of Materials

Processing Technology 157-158, 23-27.

33. Hill, R., (1979), "Theoretical plasticity of textured aggregates", Math. Proc. Camb.

Philos.Soc. 85, pp 179-191.

34. Lian J., Barlat F. and Baudelet B., (1989), "Plastic behaviour and stretchability of

sheet metals. Part II: Effect of yield surface shape on sheet forming limit",

International Journal of Plasticity, vol. 5, 2, pp. 131-147.

35. Volk W., Illig R., Kupfer H., Wahlen A., Hora P., Kessler L., Hotz, W. (2008),

"Benchmark 1 -Virtual forming limit curves. Part A: Physical tryout report",

Numisheet 2008, Part B: 3-9, Interlaken, Switzerland.

36. Volk W., Wahlen A., Hora P., Kessler L., Hotz, W. (2008), "Benchmark 1 - Virtual

forming limit curves. Part B: Benchmark analysis", Numisheet 2008, Part B: 11-19,

Interlaken, Switzerland.

37. S.S. Hecker, (1975), "Simple technique for determining forming limit curves", Sheet

Met. Ind., vol 52, pp 671-675.

38. Campos H. B., Butuc M. C, Rocha J., Duarte J. M., (2006), "Theoretical and

experimental determination of the forming limit diagram for the AISI 304 stainless

steel", Journal of Materials Processing Technology, 179, pp 56-60.

39. Graf, A., and Hosford, W., (1994), "The influence of strain-path changes on forming

limit diagrams of Al 6111 T4", Int. J. Mech. Sci. vol. 36, No. 10, pp. 897-910.

130
40. Butuc M.C., Banabic D., da Rocha A.B., Gracio J.J., Duarte J.F., Jurco P., and

Comsa D.S., (2002)," The Performance of Yld96 and BBC2000 Yield Functions in

Forming Limit Prediction", J. Mater. Process. Technol., 125-126, p 281-286.

41. Nurcheshmeh M., Green D. (2010), "Investigation on the strain-path dependency of

stress-based forming limit curves", International Journal of Material Forming, vol.

4, l,pp 25-37.

42. Assempour A., Hashemi R., Abrinia K., Ganjiani M., Masoumi E., (2009), "A

methodology for prediction of forming limit stress diagrams considering the strain

path effect", Computational Materials Science, vol. 45, Issue 2, pp.195-204.

43. Hosford W. F., Caddell R. M., (2007)," Metal Forming: Mechanics and

Metallurgy", 3rd edition, Cambridge University Press, ISBN: 978-0521881210.

44. Kasikci T., Wilson J., Kinsey B., (2011), "Experimental investigation of key

assumptions in analytical failure criteria for sheet metal forming", Proceedings of

NAMRI, Vol. 39.

45. Kasikci T., (2010), "Experimental investigation of key assumptions in analytical

failure criteria for sheet metal forming", Master's Thesis, University of New

Hampshire.

46. Raghavan K.S., (1995), "A simple technique to generate in-plane forming limit

curves and selected applications", Metallurgical and materials transactions A, vol.

26, No. 8, pp. 2075-2084.

131
47. Banabic D., (2010), " Sheet Metal Forming Processes: Constitutive Modelling and

Numerical Simulation", Springer-Verlag, New York, ISBN: 978-3540881124.

48. Li B., Nye T. J., Wu P. D., (2010), " Predicting the forming limit diagram of AA

5182-0", The Journal of Strain Analysis For Engineering Design, vol. 45, No. 4, pp.

255-273.

49. Firat M., (2008), "A numerical analysis of sheet metal formability for automotive

stamping applications" Computational Materials Science 43, pp. 802-811.

50. Reddy C. G., Reddy R. Reddy T.A., (2010), "Finite element analysis of the effect of

coefficient of friction on the drawability", Tribology International 43, pp. 1132-

1137.

51.Duan X., Jain M., Wilkinson D., (2006), "Development of a heterogeneous

microstructurally based finite element model for the prediction of forming limit

diagram for sheet material", Metallurgical and Materials Transactions A, vol. 37A,

pp. 2006-3489.

52. Simha C, Grantab R., Worswick M., (2008), "Application of an extended stress-

based forming limit curve to predict necking in stretch flange forming", Journal of

Manufacturing Science and Engineering, vol. 130.

53. Hongsheng L., Yuying Y.,Zhongqi Y., Zhenzhong S., Yongzhi W., (2009), "The

application of a ductile fracture criterion to the prediction of the forming limit of

sheet metals", Journal of Materials Processing Technology 209, pp. 5443-5447.

132
54. Uthaisangsuk V., Prahl U., Munstermann S., Bleck W., (2008), " Experimental and

numerical failure criterion for formability prediction in sheet metal forming",

Computational Materials Science 43, pp. 43-50.

55. Lee D., Majlessi S.A, (1993), " Deep drawing of square-shaped sheet metal parts,

part 1: finite element analysis", Journal of Engineering For Industry, 115, pp. 102-

109.

56. Galbraith P.C., Hallquist J.O., (1995), "Shell-element formulations in LS-

DYNA3D: their use in the modelling of sheet-metal forming", Journal of Materials

Processing Technology, 50, pp. 158-167.

57. Boudeau N., Gelin J.C, (1996), "Post-processing of finite element results and

prediction of the localized necking in sheet metal forming", Journal of Materials

Processing Technology, 60, pp. 325-330

58. Yoshida T., Katayama T., Usuda U., (1995), "Forming-limit analysis of

hemispherical-punch stretching using the three-dimensional finite-element method",

Journal of Materials Processing Technology, 50, pp. 226-237.

59. Zeng Q., Combescure A., Arnaudeau F., (2001), "An efficient plasticity algorithm

for shell elements application to metal forming simulation", Computers and

Structures, vol. 79, Issue 16, pp. 1525-1540.

60. Rank E., Duster A., Niibel V., Preusch K., Bruhns O.T., (2005), "High order finite

elements for shells", Computer Methods in Applied Mechanics and Engineering,

vol. 194, Issues 21-24, pp. 2494-2512.

133
61. Lu H., Ito K., Kazama K., Namura S., (2006), "Development of a new quadratic

shell element considering the normal stress in the thickness direction for simulating

sheet metal forming", Journal of Materials Processing Technology 171, pp. 341-347.

62. Parente M.P.L., Cardoso R.P.R., Alves de Sousa R.J., (2006), "Sheet metal forming

simulation using EAS solid-shell finite elements", Finite Elements in Analysis and

Design, 42, pp. 1137-1149.

63. R.J. Alves de Sousa, J.W. Yoon, R.P.R. Cardoso, R.A. Fontes Valente, J.J. Gracio,

(2007), " On the use of a reduced enhanced solid-shell (RESS) element for sheet

forming simulations", International Journal of Plasticity, 23, pp.490-515.

64. Doll S., Schweizerhof K., Hauptmann R., Freischlager C, (2000), "On volumetric

locking of low-order solid and solid-shell elements for finite elastoviscoplastic

deformations and selective reduced integration", Engineering Computations, vol. 17

no. 7, pp. 874-902.

65. Kulikov G.M., Plotnikova S.V., (2007), "Non-linear geometrically exact assumed

stress-strain four-node solid-shell element with high coarse-mesh accuracy", Finite

Elements in Analysis and Design 43, pp. 425-443.

66. Harnau M., Schweizerhof K., (2002), "About linear and quadratic "Solid-Shell"

elements at large deformations", Computers and Structures, 80, pp. 805-817.

67. Harnau M., Konyukhov A., Schweizerhof K., (2005), "Algorithmic aspects in large

deformation contact analysis using 'Solid-Shell' elements", Computers and

Structures 83, pp. 1804-1823.

134
68. Hauptmann R., Doll S., Harnau M., Schweizerhof K., (2001), "'Solid-shell'

elements with linear and quadratic shape functions at large deformations with nearly

incompressible materials", Computers and Structures, vol. 79, Issue 18,pp. 1671-

1685.

69. Wriggers P., Eberlein R., Reese S., (1995), "A comparison of three-dimensional

continuum and shell elements for finite plasticity", International Journal of Solids

and Structures, vol. 33, Issues 20-22, pp. 3309-3326.

70. Jung D.W., Yang K.B., (2000), "Comparative investigation into membrane, shell

and continuum elements for the rigid- plastic finite element analysis of two-

dimensional sheet metal forming problems", Journal of Materials Processing

Technology 104, pp. 185-190.

71. MSC Manual, Volume A, (2008), "Theory and User Information", MSC Software

Corporation.

72. MSC Manual, Volume B, (2008), "Element Library", MSC Software Corporation.

73. Kinsey B. L., Moondra S. and Sakash A., (2006), "Robust prediction of forming

severity using a stress based failure criterion for sheet metal", Transactions of the

North American Manufacturing Research Institute of SME, Vol. 34, pp. 491-498.

74. Mclaughlin K., (2009), "Investigation of a stress concentration factor for analytical

stress based failure criterion", Master's Thesis, University of New Hampshire.

135
APPENDICES

136
APPENDIX A
M-K MODEL ALGORITHM

A.1 ORIGINAL M-K MODEL

The M-K analysis is based on assuming that necking is initiated by a geometrical or

structural non-homogeneity of the material associated with a variation in sheet thickness.

This thickness imperfection is assumed to be in the form of a groove always oriented

perpendicular to the principal stress directions during the entire forming process (see

Figure 2.1). The sheet is composed of the nominal or "safe" area and weakened or

"defect" area, which are denoted by A and B, respectively. The initial imperfection factor

of the defect /„ is defined as the thickness ratio f0 — — where t denotes the thickness

HA

and subscript o denotes the initial state before the deformation. The strain and stress states

in the two regions are analyzed by imposing a strain path, specifying the principal strain

d£XB and principal strain d£XA. Compatibility and force equilibrium conditions are used

to calculate all the stress and strain values in the two regions and to determine if the
guessed value is correct [8, 23]. As the deformation progresses and the ratio —lS-

delA

becomes too high (approximately 10), the deformation of the specimen has localized in

region B. The first and second principal strains in region A (£XA&n&£2A) at the onset of
plastic instability defines a point on the FLC and by varying the strain ratio p = — L
dsx

different points on the FLC are obtained.

137
Before specifying the algorithm for the M-K model, the basic relationships for metal

deformation will be presented. These include the yield criterion (i.e., the transition for

elastic to plastic deformation) and the constitutive relationship (i.e., how stress is related

to strain in the material). As an example for Hosford higher order yield criterion (Hosford

1979) (assuming a plane stress condition):

R90(aiy+R0(a2y+R0R90(tT1-a2y=R90(R0 +\)aa (A.l)

<T = •(^M +R.M +R0R90{^ -<*iY) (A.2)


V^J o f a + 1 )

where R0 and R90 represent the anisotropic behavior of the material (i.e., the strain in the

width direction divided by the strain in the thickness direction during a uniaxial tension

test); <JX and G2 are the in-plane principal stresses, a is the higher order exponent

(typically 6 for BCC material and 8 for FCC material) and a is the effective stress.

The behavior of the material can be represented by Swift power hardening law:

<T(e) = K(eY{iY (A3)

where K is the strength coefficient, n is the strain hardening exponent, m is the strain rate

sensitivity exponent, e is the effective strain and £ is the effective strain rate.

From Eq. (A.2):

a (RQQ + RQ (a)a + R0RQO(1 ~ a)" ^


0=— = (A.4)
V ^90 fo+1)

where the ratio of the principal stress is defined as:

a =^ (A.5)

138
The flow rule which also relates stress to strain during plastic deformation is

expressed as:

dE dA^EL (A.6)

by applying the definition of dX from Eq. (A.6) in the principle of plastic work, the

principal strains and the effective stain can be related as:

d£x _ d£2
1 1 1
^oW +^o^9o(°"i - ^ r ^(^r -^ofo -^r1 (A 7)
_ - ds3 _ de
=
Rnfar+R0(*2y->= Unfa 4-ix^r

Thus:

^°^%w--'k+^('-ar) <A 8)
'

where d£x and d£2 are the incremental in-plane principal strains

From Eqs. (A.5) and (A.7), the ratio of the incremental principal strain is defined as:

_ ds2 = R0 \a)a - R0 i?qo (l - a)a (A.10)

dex R90+R0Rgo(l-ay-1

Using the volume constancy rule:

d£x + d£2 + d£3 = 0 (A.l 1)


From Eqs. (A.8), (A.9) and (A.l 1)

&
-\(v»r ( '-^ ( 1 -° ) i (A 12)
'
By applying the principle of plastic work, which is used to define the effective strain:

od.£ - axd£x + <J2d£2 , K , ON

139
ud£ = (J, d£x (l + ap) r A 14")

and

p=de_ = {\ + ap) (A15)


dex (j>

Now that the basic sheet metal forming relationships have been presented, equations

for the specific case of the original M-K model shown in Figure 2.1 are discussed. A

compatibility condition is assumed which relates the strain in A and B regions by:

d£2A =d£2B (A. 16)

From the assumption of imperfection factor:

(A i7)
u=Y -
and the definition of the strain through thickness:

*3, = l n - (A.18)
' h

f = foew(s3B-e3A)=tf (A.19)
'A

The equilibrium condition requires that the applied load remains constant along the

specimen (see Figure 2.1); therefore:


F
\A = Fw (A.20)

<7
\AtA =(7
\BtB (A.21)

From Eqs. (A.19) and (A.21):

°\A = fa\B (A.22)

From Eq. (A.4):

140
<*A = r^B (A.23)
<PA <t>B

From Eq. (A.3):

(eA + deA )" (eA )"' _ (eB + deB )" (eB f (A.24)
<I>A <t>B

From Eq. (A.19):

(sA+deAy{eAY f e™(r r )(£B + ^BY^B) (A.25)


Jo exp [£iB - S3A) :
0A <f>B

From Eqs. (A. 10), (A. 15) and (A. 16):

f r, \
— (£A+d£A)" £ = -rfoe*P (£3B - £3A )(£B + d£
B Y (A.26)
<PA \PAJ \PBJ

By substituting Eqs. (A.10), (A.12) and (A.15) in equilibrium Eq. (A.26), a governing

equation can be found and solved numerically for the error in [dsA).

Now after the presenting of fundamental relationships for plastic metal deformation

and those specific to the M-K model, the use of these in the M-K algorithm can be

presented (as shown in the flowchart of Figure A.1). The M-K methodology begins with

specifying a strain path \pA ) , material constants (K, n, R , m), order of yield criterion

(a), a finite increment of strain in region B \d£XB) and the imperfection factor f0. Then a,

<j> and /? values in region A can be found using Eqs. (A. 10), (A.5), and (A. 15), all the

strain components in region A can be defined after applying Eqs. (A. 10), (A.ll) and

(A.15). Equations (A.4), (A.10), (A.ll), (A.15), and (A.16) are used to calculate stress

and strain values in region B. Finally, force equilibrium (Eq. (A.26)) is used to check for

the error to determine if the stress and strain parameters in both regions A and B are

correct. If so, then the strain state in the material is updated and the procedure is applied

141

1 A? 1 A

again in order to determine a strain path for the deformation. When the ratio —— '\A > R) is
d£.
reached, the procedure is ended and a point on the FLC (i.e., the limit strains in region A)

is noted. This methodology is repeated for various strain paths in order to determine the

entire FLC for the material.

The relationships of p and 0 are specific for each yield criterion, so different results

will be generated based on the yield criterion used. To incorporate other yield criteria, the

equation for ^ can be expressed in terms of a similar to Eq. (A.4) as:

<j> = 4\-a-\-a2

2R 2
=
Y~Y+~ia+ (A-27)

^ = (b + ba°+{2-b\l-hayfa)
for the von Mises, Hill '48, and Barlat '89 yield criteria respectively.

Where for Barlat'89:

2
% + Ri)\ + R9Q'h Vl + k *9o°' fl 8 (A 28)
'

and for Hill '48,

—=
R _ -"-0
R0 "*"
+ 2R
ZJ 45 + Rg0
MS ~*~ -^90 (A.29)
4

Also the equation for p can be expressed in terms « of similar to Eq. (A.10) as:

142
2a -I
2-a

+
p4 ^"-R (A.30)
(l + R)-Ra

_baa-1-h(2-b\l-hay i

P i
~ b+ {2-b\l-hay

for von Mises, Hill '48, and Barlat '89 yield criteria respectively. For Barlat '89 and

other higher order yield criterion, « i n terms of p is solved by an iterative process as

shown in the flowchart (Figure A.l).

Finally, the equation for the effective strain and effective stress can be expressed as:

d£ = J—[d£x2 + d£22 + d£xd£2 j - d£x j—(l + p2 + pj

rr———- SCTJ(\ + P2+P/L


= i°\ +v2 -vpz = —
2+p

R+l , i 2R
d£ = J-^=— (d£x + d£j + Rd£\iJ ): d£xJ\ + p2 +——p
\2R+V J2R+1 V R+l (A.31)

= (rf+cr2+R{<Jx-a2)2 =
y^hh+^p+p
V
\{ r
R+l
V ^ +1 R+l + Rp

d£x + ad.£2 1 + ap

for von Mises, Hill '48, and Barlat '89 yield criteria respectively.

Other yield criteria such as Yld96 [25] and BBC2000 [32, 40] have been successfully

used with the M-K analysis and have shown good fit to experimental data depending on

the material considered.

143
Start

Initialize
ElA-0; E 1 B =0; E 2 A =0; £ 2 B =0; E 3 A=0; E 3B =0; e Ae ff=0; EBeff=0

Input pA, material


constants (K,n,R,m), order of
yield criterion (a), value 01%,
=
JIB

No-

aA=(R+(l+R) pA)/(R+l+RpA)

when a>2 then solve (aA(a-1)-R*(l-aA)(a"1))/(l+R*(l-aA)(a"1))= PA

start with ocA zero and increase by 10"5 until this equation is satisfied to an Eq. (A.10)
error of (10'8)

* A = ( ( I W +R(1- a A )')/(R+l)r
Eq. (A.4)
1 '
pA=(l+aAa+R(l-aA)a)/ (l+RCl-aA)"'1 |<|)A|) Eq.(A. 15)

^'
-">IA "
« (
* \

AE2A=pA AE1A Eq. (A.10)


1r

A S 3 A = - A E 1 A - AE2A Eq. (A.ll)

Ae Aeff =AEi A *P A Eq.(A. 15)

Set AE2B=AE2A Eq. (A. 16)

FindpB=AEiB/AE 2 B Eq. (A.10)

ure A.1: Original M-K model flowchart for a single deformation path

144
Yes

a B =(R+( 1+R) P B ) / ( R + 1 + R P B )

when a>2 then solve (a B ( ' "-R*(l-a B ) < a ' V U + R H l W n


)= PB

start with a B zero and increase by 10 5 until this equation is satisfied to an Eq. (A.10)
error of (10 8 )

(M(l+a B a +R(l-a B n/(R+l)r Eq. (A.4)

PB^l+aB'+Ra-ctfi)')/(1+R(1-<*B)* ' |<M)


Eq.(A. 15)

AEBrff=AEiB* PB

AE 1B = - A E I B - AE 2B Eq. (A.ll)

From force equilibrium across defect, error= (£Aefr +A£Aetr)° /<|>A fo exp(£3B-£3A) (EBeff +AEBefr )7 i Eq. (A.26)
(this comes from C7IA= CTIB fo exp(E3B-E 3A ))

error changes sign (since the error is a


jjionotonic function with a single zero)
Increase AEIA = AE, A + ACIA™
-0
U p d a t e EIA, £ 2 A , EAeff, ElB,
£2B, E Beft , E 3 B , E 3 A a n d Store
El B , E2B, EIA, E2A

No • B

Figure A.l: Original M-K model flowchart for a single deformation path (continued)

145
As an example of the effect of the yield criterion on the M-K model results, Figure

2.7 demonstrates the need for incorporating higher order yield criteria to accurately

predict experimental data. The overestimated predictions on the positive minor strain side

of the FLC are obvious when a =2 (i.e., von Mises and Hill's 1948 yield criteria). Figure

2.4 shows the overprediction of the experimental data on the negative minor strain side

when the orientation of the defect is physically not aligned with the principle direction.

Thus the M-K model can be enhanced to improve FLC predictions on the negative minor

strain side by varying the defect orientation as shown in the next section.

A.2 MODIFIED M-K MODEL

According to Hill's observations and theory of localized necking [19, 43], failure

occurs at a certain angle between (0° to 35°) depending on the loading path and material

when thin sheets are subjected to tensile stresses (see Figure 2.2). Thus the assumption of

a perpendicular defect to the major principal stress direction is not valid for the negative

minor strain side of the FLC and the need to incorporate this into the M-K theory is

required.

In the modified M-K analysis, the defect can be oriented from 0-90° with respect to

the principal strain directions as shown in Figure 2.3. The computations in this procedure

are identical to the equations of the original M-K model. The strain compatibility and

force equilibrium equations must be applied at the interface between the safe and defect

region [29, 30]. The force equilibrium equation (previously Eq. (A.20)) can be rewritten

as:

146
w =w (A.32)
nnA nnB

•f/ilA ** MB

where subscripts n and t denote the normal and tangential directions of the groove,

respectively, and F is the force per unit width in the t direction.

The Instantaneous imperfection factor ( / ) has the same definition as in Eq. (A.19),

thus:

M nnB ^ nnA
(A.33)

fontB = V»L

The compatibility requirement (previously Eq. (A. 16)) assumes that the strain in the

defect direction is equal in both regions:

dellB — denA (A.34)

Finally, the work-energy relation (previously Eq. (A. 13)) can be rewritten as:

(d£nrPnn + &„<*„ + 2^„,0"„,)- d£<7 = 0 (A.35)

where a is calculated from the hardening rule.

Groove :one

Rolling
Direction (x)

Figure A.2: The modified M-K model with the transverse direction t parallel to the

defect and the normal direction n perpendicular to the defect [29]

147
To begin the algorithm, all strains are set to zero. As with the original M-K model, the

inputs are the strain path \pA ) , material constants (K, n, R , m), order of yield criterion

(a), a finite increment of strain in region B [d£XB)and the imperfection factor / „ . A

similar procedure to the original M-K is used to determine the strain components in

region A and B. Using the rotation matrix, T, the strain and stress tensors are transformed

to the defect orientation by:

° ™ <Fn, cos(0) sin(0)'


o-"'=ro-vrr = where T = (A.36)
°ni a
tr -sin(0) cos(f?)

After calculation of the stress and strain components in the safe region, the same

variables must be solved for in the defect region. For this purpose, another improvement

was incorporated in the modified M-K model by the use of Newton-Raphson (N-R)

method in the solution procedure to solve for the strain path data. The flowchart for the

original M-K model shown in Figure A.1 was closed form with an iteration to determine

the correct stress and strain values for force equilibrium. Using the N-R method, a

considerable reduction in the computation time is achieved.

As in the original M-K model, a localized necking in sheet material is assumed to


occur when the ratio —— > 10 is reached. At this stage, the computation is stopped and
d£XA

the corresponding limit strains (£iA,£1A) and limit stresses (GXA,cr2A) are stored. The

analysis is repeated for different initial orientations (9) of the defect in the range between

0° and 90° and the limit strain point is obtained after the minimization of £XA versus 9.

The unknown parameters in the defect region can be reduced to four:


B'(TnnB^CrnBand^ntB

148
To calculate these four parameters four equations must be solved numerically using

Newton-Raphson method [9, 25, 29, 30]:

G = [GX G2 G3 Gj (A.37)

where

r _ (dSnnlPnnB + dSnB°ttB + l d S
ntB°nts) 1_ A „, .
Gx - —— 1- u Work-energy equation
d£Ba

G2 = — 1=0 Compatibility equation

ttA

(A.38)

(J 3 = y —M— \ = Q Force equilibrium equation

G4 = / — 1=0 Force equilibrium equation


<*ntA

It should be noted that defect orientations (9) is being updated from an initial value at

each increment of the plastic deformation due to band rotation as shown in Eq. (2.14).

Here is the modified M-K algorithm:

1. Initialize £!A=0, £iB= 0, £2A=0, e2s= 0, £j^= 0, £3B= 0, £Aeff= 0, £Bejr= 0 assuming As1B

=0.0001

Input PA, material constants (K, n, R , m), order of yield criterion (a), value of f0

2. If PA^-0 then set # = 0 . Otherwise set a search region around theoretical limit of

# (search region could be from 0-^45)

3. Ifa=2then aA=(R+(1 + R) pA)/(R+1+R PA)

149
If a>2 then solve

(aA(a'])- R (l-a^'^J/Q + R (l-aA)(a'1))= pA that is start with aA zero and increase

by 10"5 until this equation is satisfied to an error of (10-5)

4. <f,A=((l+aAa+R (l-aA)a)/(R+l))1/a,/3A=(l+a/+R (l-aAf)/(l + R (l-aAy' l<f>A\)

5. Initialize A£\A=0

6. A£2A = pA A£]A

7. A£3A = - A£1A - A£2A

8. A£Aeff =A£1A pA

9. Transform strain increments to groove direction

A£nnA=A£iA cos 0+A£2A sin 0

A£ttA=A£iA sin20+A£2A cos20

A£MA =(-A£]A +A£2A) cos0sin0

10. <JAeff=k (£Aeff+A£Aeff) "

1 1. (TjA=(?Aeff/0A

12. <72A=aA (TlA

13. Transform stresses to groove direction

0~nnA = 0~lA COS 0+&2A SIH 0

&ttA=o-]A sin 0+a2A cos20

PntA =(-OlA +0-2A) COS0 SW0

14. Set A£ttB=A£ttA

15. Set initial guess for solver

GrmB =0~nnA/f; 0'nB=(TttA; A£Beff = A£Aeff, f=fa'/ GntB =<TntA/f;

150
16 find A£2B=(A£ttB-A£iB sin20)/cos20

17. Find pB =A£1B/A£2B and solve (aB(a'1J- R (l-aB)(a'l))/(l+R (l-aBf'1))= pB (start with

aB zero and increase by 10"5 until this equation is satisfied to an error of 10"5)

18. <t>B=((l+aB° + R (1- aBf)/( R +l)f°

19. pB=(l + aBa+ R (l-aB/)/ (l + R (l-aaf-1 l<t>B\)

20. A£Beff=A£iB J3B

21 Transform strain increments to groove direction

A£nnB=A£iB cos 0+A£2B sin 0


7 7

A£ttB=A£]B sin 0+A£2B cos 0

A£ntB =(-A£1B +A£2B) cos0sin0

22. crBefl=k (£Beff+A£Beff) "

23. CJiB=CrBeff/<t>B

24. a2B=aB cr1B

25. Transform stresses to groove direction

<TnnB=0'lB COS2 0+CT2B Sin2 0

0~ttB=0'lBSin 0+CT2BCOS 9

&ntB =(-0"lB +&2B) COS0Sin0

26. error,= annA- annBfo exp(£3B-£3A)

27. error2 = antA- crntBf0 exp(£3B-£3A)

28. error3 = aBeff- aByieid

29. error4 = A£ttB-A£ttA

151
30. From the Jacobian of the four error functions determine the direction that minimizes

the four error functions simultaneously. From which find AEUB, crntB, <j„nB and auB. Repeat

16-30 until the errors are minimized

31. Find the corrected A values from:

AEttA =A£ttB

A£]A= AsttA /(sir?0+pA cos20),

AE2A = PA A£JA

A£3A = - ASIA - AE2A

A£Aeff =A£]A PA

32. Update E1A, £2A, £Aeff

33. Update s1B, £2B, £Beff, £3B, £SA

34. Store eIB, £2B, £JA, ?2A

35. Repeat 5-34 until \A£IA\< = \A£iB\/10

36. If pA>0 stop otherwise continue iterating for different 0 by repeating 3-36 until the

value of minimum A£iA is obtained

152
APPENDIX B

M-K PREDICTIONS OF THE MATERIALS PRESENTED IN TABLE 2.1

0.8 K=660.965 MPa, n=0 2222, Ro=1.28,R45=2.08,


R90=1.86, m 0.015, f=0.997
0.7

0.6
s
et 0.5
\ s
u
• * *

w
V 0.4
s
Hs- 0.3
u
o
et
s 0.2

0.1 • Experimental data


-*—Hill's 79, a=6 - Swift
-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 0.6
Minor True Strain

Figure B.l: M-K prediction of FLC for HC220YD Steel (1) [35, 36]

0 100 200 300 400 500 600 700 800


Minor True Stress (MPa)

Figure B.2: M-K prediction of stresses at failure for HC220YD Steel (1)

153
0.9
K=612.1 MPa, n=0.3279, R^l.68, R45=2.12,
0.8 R90=2.54, m^0.015, f=0.9978

0.7

2 0.6
VI
0.5
3
0.4
©
0.3

0.2

0.1
Experimental data
-Hill's 1979, a=6-Swift
0 r—
-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 0.6
Minor True Strain

Figure B.3: M-K prediction of FLC for HC220YD Steel (2) [35, 36]

900 i

800

700 -
et
g 600 4 ^ ^ a—+~ * ,„„#

a> 500
u
•+*
r/5
Q> 400
9
300 -
p
'et 200 -

100

100 200 300 400 500 600 700 800 900


Minor True Stress (MPa)

Figure B.4: M-K prediction of stresses at failure for HC220YD Steel (2)

154
1A
K=426.2 MPa, n=0.228,lRo=1.7,R45=1.27,R90=2.13,
m=0.0 I8,f=0.999
1.2 •


c 1 .
2
™ 0.8 •
su
9 •
u
He? N*. *J*
0.4 \

0.2 • Experimental data


—*—Hill's 1979, a=6- Swift

-0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1


Minor True Strain

Figure B.5: M-K prediction of FLC for AK Steel (2) [37]

600

500

et ^p^f—#**"*—~*
&J 400
t/2
an
2 300
vi
9
£ 200 -
u
o

S 100

0 100 200 300 400 500 600 700


Minor True Stress (MPa)

Figure B.6: M-K prediction of stresses at failure for AK Steel (2)

155
0.8 K=1527 MPa, n=D.47, Ro=0.92, R45=1.3,
R90=0.|78,m=0.012
0.7

0.6
^e
'3
-fc 0.5
vi
<u
£ 0.4
H
.2,0.3
a

0.2
Experimental data
0.1
Hill's 1979, a=6-Swift
0 r-
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5
Minor True Strain

Figure B.7: M-K prediction of FLC for AISI 304 SS [38]

1600

1400 -

1200 i
et
OH
§ , 1000
wa
m
£ 800
Vl
| 600
H
.o 400
e?
200

200 400 600 800 1000 1200 1400 1600


Minor True Stress (MPa)

Figure B.8: M-K prediction of stresses at failure for AISI 304 SS

156
05
K=537 MPa, n=0.2845^ Ro=0.78, R45=0.485, R90=0.58,
A=426 MPa, E;=234.0MPa, C=5.130
0.4

a
'3
u #

v, 0.3 . #

9 a "C
© 0.2 *^
'eT

• Experimental data
0.1
—*-Hills 79, a=8-Swift
- « - Barlat'89, a=8 -Voce
-0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
Minor True Strain

Figure B.9: M-K prediction of FLC for AA2008-T4 [7]

500

450

et 400 H
OH

350 -
eft
300 -
t
VI 250
V
s
u
H 200 -
i-
os
'3 150

100 -
-Hill's 1979, a=8-Swift
50 -

_0_
•Barlat 89, a=8 - Voce
-50 0 50 100 150 200 250 300 350 400 450
Minor True Stress (MPa)

Figure B.10: M-K prediction of stresses at failure for AA2008-T4

157
0.5
K=507.7 MPa, n=0.22|, Ro=0.91, R45=0.68, R90=0.83
A=380 MPa, ^=249.7 MPa, C=7.895
0.4

et
u 0.3
to
9

^0.2

0.1 Experimental data


Hill's 1979, a=6 - Swift
Barlat 89, a=8 - Voce
-0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
Minor True Strain

Figure B.ll: M-K prediction of FLC for AA5182 [35, 36]

500

450
400
03
P. 3M)
S
Vj 300
VI
01
u 750
CO
•**
01
9 700
I.
H
la 150
©

s«« 100 •Hill's 1979, a=6- Swift


50
^Barlat 89, a=8- Voce
-50 0 50 100 150 200 250 300 350 400 450 500
Minor True Stress (MPa)

Figure B.12: M-K prediction of stresses at failure for AA5182

158
04
K=417.85 MPa, n=0.l45, R^O.8,R 4 5 =0.43, R90= 0.61,
A=318MPa, B=191.1MPa, C=8.706

0.3

1 \
| o , ^

'3 s
0.1
• Experimental data
-*~Hill's 1979, a=6 - Swift
- • - B a r l a t 89, a=8 - Voce
-0.2 -0.1 0 0.1 0.2 0.3
Minor True Strain

Figure B.13: M-K prediction of FLC for AA6016-T4 [25]

400 -

350 -
^ •v/f^y*^^
£ 300 —^f&sSrBHxab
*ga*

¥«25 ° -
u
•+*
<*> 200 -
Ol
sk.
H 150 -
s-
o
•™9

gioo -

50 - «*~Hills 1979, a=6 -Swift


—©'-Barlat 89, a=8 -Voce
0 ! 1 1 '" 1 1 1 '' ' i 'I

50 0 50 100 150 200 250 300 350 400


Minor True Stress (MPa)

Figure B.14: M-K prediction of stresses at failure for AA6016-T4

159
0.4 K=554 MPa, n="0 .2578, Ro=0.79, R 45 =0.63,
R 90 =0.67, A=440 MPa., B=275.2 MPa, C= 7.015

0.3
et
u
+*
CO

2 0.2
H
s-

'S5
0.1
Experimental data
Hill's 1979, a=6- Swift
Barlat 89, a=8 - Voce
o i—-
-0.2 -0.1 0 0.1 0.2 0.3
Minor True Strain

Figure B.15: M-K prediction of FLC for AA6111-T4 [39]

500 -j

450 -
^ ^ ^ ^
400

£350

S300
Vi

2 250
•*O
C -»
2 200
9
u _
*T 150
u
O
S'IOO -
S —•—Hill's 1979, a=6 - Swift
50 -
- • - B a r l a t 89, a=8 - Voce
0
-50 0 50 100 150 200 250 300 350 400 450 500
Minor True Stress (MPa)

Figure B.16: M-K prediction of stresses at failure for AA6111-T4

160
APPENDIX C
MODIFIED EFFECTIVE STRESS RATIO MODEL ALGORITHM

A is the safe region, B is the defect region, a is the higher order yield criterion index

Given o2A we iterate through o\A

O.A=G2A/GIA

pA = (aA(a-1}- R (l-aJ^yfl+RO-aJ^)

<pA=(l+ (aA)a+R (l-aA)a/(R+l)f/a)

oeffA= VIA 9A

eeffA= (GeffA/K)(,M)

e1A = ceffA aeffA/(a1A (l+aA p/))

£2A = (veffA eeffA-aiA £!A)/G2A

Now transform strains and stresses to groove direction


7 7
SnnA^ £lA COS 0+ E2A Sin 9
7 7

EttA— £1A Sin 0+ E2A cos 0

EntA= (~£lA+ £2A) COS0Sin0

<JnnA= OlA COS20+ G2A Sin"0

0~tlA = GIA Sin29+ G2A cos29

&ntA=(- OIA + G2A) cos 9 sin 9

161
Next we turn towards B region

EttB ~ £ttA

seffs =EeffA/y>/n)

GeffB= K*(£effB)"

The problem now is that given ettB, eeffB, GeffB we need to find e}B, GIB, G2B

We start with an initial guess for ennB and we evaluate the following set of auxiliary

equations
7 7 7 ?
£IB ~(E„„B cos 9-EttB sin 0)/( cos 0-sin 0)
7 7 7 7

E2B =( EttB cos 0-EnnB sin 0)/( cos 9-sin 9)

PB =£2B/£IB

Next find aB from pB by solving the following polynomial for real root

ay1 - R ( I +PB) (1-O.B y 1


-pB = 0

Given that G2B = «B GJB from which we see that we have two unknowns EIB, GIB and two

equations. The equations that applies on these variables are

GIB a + G2Ba+ R (GIB- G2B) " = (R +1) GeffB "

ElB GJB + £2B G2B= EeffB GeffB

We solve these two equations together with the auxiliary equations to find an update to

EnnB and get the value of GIB until the difference is very close to zero

162
Now having £mB and EUB (= SUA) we re-evaluate the auxiliary equations

EIB = ( EnnB cos2 0-£ttB sin 0)/(cos20-sin20)


7 7 7 7
E2B = (EttB COS 0- E„nB Sin 0)/( COS 0- SW 0)

PB =E2B/EIB

Next find aB from ps by solving the following polynomial for real root

ay1 - R (1+PB) (1 - aB)a'' - pB = 0

G2B = O-B GIB

Now transform stresses to groove direction

7 7
O"««B= GIB COS 0+ G2B sin 9

&ttB= GIB sin 9+ G2B COS 9

GMB=(- GIB + C>2B) cos9'sin9

Next find

E3A = -(E1A+ E2A)

E3B = -(EIB + E2B)

and perform a force balance test

GnnAcalc = CTnnB CXp(E3B - £3^

Force = (JnnA / CTnnAcalc

If 0.99<Force< 1.01 then record ajA, °~2A, £IA and E2A as failure point values otherwise

iterate with another value for <TiA

163
APPENDIX D

STRAIN-STRESS CONVERSION METHODOLOGY

I- Converting From Strain to Stress Space

Assuming a plane stress condition (cr3 =0) the ratio of minor true stress, o"2, to the

major true stress, <JX, is defined by the parameter:

a = ^- (D.l)
O",

The ratio of minor true strain, £2, to major true strain £x, is defined by the parameter:

p = £i (D.2)
£
\

Plasticity theory defines an effective stress, a , which is a function of stress tensor

components and a set of material parameters. For in-plane loading conditions, the

effective stress can be expressed in terms of principal stresses and material parameters

using Hill's 1948 yield criterion as follows:

*=x( V
\+a
. ™ ^
2R
-——a
*+l J
(D.3)

where the average anisotropic material parameter, R , is:

and the a-p relationship can be reduced to:

(l + j j O » - * (D.5)
(l + R)-Ra

Effective strain can be expressed in terms of £x,£2 ando; as:

164
(l + R) L 2 2R m „
£
= n LeJl + p2+—^P (D.6)
V(l + 2i?) V l+ R
The relationship between effective stress and effective strain is given by the power

hardening law:
G = K{s)" (D.7)

Major and minor stress can be calculated using the following equations:

a2 - aax (D.8)

Steps for converting strain-based FLC values to stress-based FLC values:

1. Calculate p from strain increments using (D.2)

2. Calculate a using (D.5)

3. Calculate effective strain using (D.6)

4. Calculate effective stress using (D.7)

5. Calculate o~, using (D.3)

6. Calculate minor stress using (D.8)

II- Converting From Stress to Strain Space

Given the pair (<JiA, G2A), the strain-based FLC can be obtained as follows:

1. Compute aA=CT2A /<JIA

2. Use aA and compute pA, (j>A and pA

3. Substitute <f>A and <TJA into the equation <JlA = GA I </>A to obtain 0A

4. With (TA , compute the effective strain £A through an inverse hardening curve equation

5. Using £A, compute E]A from £A = PA£lA

6. With the knowledge of EJA and pA, £2A is computed

165

You might also like