You are on page 1of 230

IMPROVING THE FORMABILITY LIMITS OF LIGHTWEIGHT METAL ALLOY

SHEET USING ADVANCED PROCESSES


- FINITE ELEMENT MODELING AND EXPERIMENTAL VALIDATION-


DISSERTATION


Presented in Partial Fulfillment of the Requirements for
the Degree of Doctor of Philosophy in the Graduate
School of The Ohio State University

By

Serhat Kaya, M.S.
* * * * *

The Ohio State University
2008
Dissertation Committee: Approved by
Professor Taylan Altan, Adviser
Associate Professor Jerald Brevick ______________________
Assistant Professor Allen Yi Adviser
Industrial and Systems Engineering
Graduate Program


ii
ABSTRACT


Weight reduction is one of the major goals in the automotive, appliance
and electronics industries. One way of achieving this goal is to use lightweight
alloys such as aluminum and magnesium that have high strength to weight
ratios. However, due to their limited formability at room temperature, advanced
forming processes are needed. Room temperature and elevated temperature
hydraulic bulge tests (using a submerged tool) were conducted for Al 5754-O
and Mg AZ31-O to determine their mechanical properties. Experiments were
conducted between room temperature and 225 C, at various approximate true
strain rates. Strain values up to 0.7 were obtained under equi-biaxial state of
stress at elevated temperatures. Flow stress curves were calculated using the
membrane theory.
Deep drawability of aluminum and magnesium alloys is investigated
through experiments and process simulation at room temperature (using solid
dies), against liquid pressure (hydroforming) and at elevated temperatures
(warm forming). Limiting Draw Ratio (LDR) of Al 5754-O is increased from 2.1
(room temperature) to 2.4 when hydroforming is used as the drawing process.
This value is increased to 2.9 when warm forming is used. Formability of Mg

iii
AZ31-O is found to be limited at room temperature while LDR up to 3.2 is
obtained at elevated temperatures. Warm forming experiments were conducted
using a servo motor driven press and a heated tool set. The in-die dwelling
concept is developed by using the flexibility of the servo press kinematics and
blanks were heated in the tool set prior to forming. Temperature time
measurements were made at various blank holder interface pressures in order
to determine the required dwell time to heat the blank to the forming
temperature. Several lubricants for elevated temperature forming were
evaluated using the deep draw test and a PTFE based film was selected as a
lubricant at elevated temperatures. Deep drawing tests were conducted to
determine the process window (max. punch velocity as functions of blank size
and temperature) for Al 5754-O and Mg AZ31-O. Maximum punch velocities of
35 mm/s and 300 mm/s were obtained for the Al and Mg alloys, respectively.
Comparisons for the Mg alloy sheets from two different suppliers were made
and significant differences in formability were found. Additional experiments
were conducted in order to understand the effect of constant and variable
punch velocity and the temperature on the mechanics of deformation. Variable
punch velocity is found to improve the thickness distribution of the formed
part and provide 60 % reduction in the drawing time. By calculating heat
transfer coefficients using inverse optimization, computational models are

iv
developed and experimental results are used to validate the predictions from
the computational model.




v







Dedicated to my family













vi

ACKNOWLEDGEMENTS


I wish to thank my adviser, Taylan Altan, for intellectual support,
encouragement and enthusiasm which made this thesis possible, and for his
patience in correcting both my stylistic and scientific errors.
I thank my candidacy and dissertation committee members Professors
Jerald Brevick, Ted Allen and Allen Yi for discussions and support.
Most of this research was supported by a grant from the National
Science Foundation.
Finally, I sincerely thank to my mother Neriman Kaya, my father Ahmet
Kaya, my sister Neslihan Kaya and my brother Ferhat Kaya for their unending
support, encouragement and patience.

vii
VITA
July 07, 1973 Born Istanbul, Turkey
1998... Die Design Engineer, Turkey
1999 M.S., Mechanical Eng.
Yildiz Technical University, Istanbul

2001 M.S., Industrial and Systems Eng.
The Ohio State University
2002 present... Graduate Research Associate,
The Ohio State University

PUBLICATIONS
Research Publication
Book Chapter
1. William Thomas, Taylan Altan and Serhat Kaya, (2003) Handbook of
Aluminum, Volume I, Physical Metallurgy and Processes, Chapter 18, pp.837-
880, Marcel Dekker, Inc., New York, ISBN: 0-8247-0494-0

Journal Papers
2. Taylan Altan, Serhat Kaya, Yingyot Aue-u-lan, (2007), Forming Al and
Mg Alloy Sheet and Tube at Elevated Temperatures, Key Engineering
Materials, vol.344, pp.317-323

3. Serhat Kaya, Taylan Altan, Peter Groche, Christian Kloepsch, (2006),
Determination of Flow Stress of Magnesium AZ31-O Sheet At Elevated
Temperatures Using the Hydraulic Bulge Test, (accepted/in print - Special
Issue of International Journal of Machine Tools and Manufacture)

viii
Conference Papers
4. Taylan Altan, Serhat Kaya, Yingyot Aue-u-lan, (2007), Forming Al and
Mg alloy Sheet and Tube at Elevated Temperatures, Shemet 07, 12th
International Conference on Sheet Metal, April 1-4, Palermo, ITALY

5. Hariharasudhan Palaniswamy, Ajay Yadav, Serhat Kaya, Taylan Altan,
(2007), New Technologies to Form Light Weight Automotive Components,
4th International Conference and Exhibition on Design and Production of
Machines and Dies/Molds, DIEMOLD 2007, June 21-23, Cesme, TURKEY

6. Serhat Kaya, and Taylan Altan, (2006), Forming Limits of AZ31-O
Magnesium Alloy Sheet at Elevated Temperatures, 2006 NSF Design, Service
and Manufacturing Grantees and Research Conference, July 24-27, 2006, St.
Louis, Missouri, USA

7. Taylan Altan, Yingyot Aue-u-lan, Hariharasudhan Palaniswamy, Serhat
Kaya, (2005), State of the art-visions and priorities in research and
development in metal forming, Proceedings of the 8th International
Conference on Technology of Plasticity, ICTP 2005, ISBN 88-87331-74-X, pp.3 -
24, October 9-13 2005, Verona, ITALY (Keynote Paper)

8. Taylan Altan, Yingyot Aue-u-lan, Serhat Kaya, (2004), Tube and Sheet
Hydroforming-New Developments in Equipment, Tooling and Process
Simulation, Society of Manufacturing Engineers, SME, 2nd Annual North
American Hydroforming Conference, September 27-29, 2004, Ontario,
CANADA


FIELDS OF STUDY
Major Field: Industrial and Systems Engineering
Minor: Design / Mechanics of Materials
Minor: Design of Experiments, DOE

ix
TABLE OF CONTENTS
Abstract..ii
Dedication..v
Acknowledgements.vi
Vita...vii
List of Tables..xii
List of Figuresxiv

Chapters:
1. Introduction....................................................................................................... 1
2. Objectives........................................................................................................... 4
3. Mechanical properties of aluminum and magnesium alloys ........................... 6
3.1 Forming behavior of Mg alloy sheet ...................................................... 8
3.2 Strength asymmetry in Mg alloy sheets............................................... 11
4. Determination of the flow stress of lightweight alloy sheet under equi-biaxial
state of stress........................................................................................................ 18
4.1 Introduction.......................................................................................... 18
4.2 Experimental Setup.............................................................................. 20
4.3 Calculation of stresses and strains under equi-biaxial state of stress. 22
4.4 Membrane Theory................................................................................ 23
4.5 Calculation of the radius of the bulge (dome) ..................................... 24
4.6 Calculation of the Thickness at the Top of the Dome......................... 25
4.7 Stress Strain relationship at elevated temperature for magnesium
alloy sheet ............................................................................................................ 26
4.8 Experimental investigation.................................................................. 29
4.8.1 Observations.............................................................................. 29
4.8.2 Test conditions and experimental matrix ................................ 30
4.8.3 Measurement of initial sheet thickness.................................... 32

x
4.8.4 Sheet Draw-in During Bulging.................................................. 33
4.8.5 Pressure bulge height curves .................................................... 34
4.9 Analysis of the dome geometry............................................................ 35
4.9.1 Deviation from the original bulge shape .................................. 39
4.10 Thickness distributions along the dome curvature........................... 42
4.11 Conclusions ........................................................................................ 48
5. Deep drawing process at room temperature .................................................. 50
5.1 Drawability Criteria.............................................................................. 52
5.2 Critical Stroke in deep drawing........................................................ 54
5.3 Deep drawing of aluminum and magnesium at room temperature.... 55
6. Hydroforming of sheet at room temperature .................................................. 56
6.1 Mechanics of SHF-P Process................................................................ 60
6.2 Prediction of the Initial Pressure Value............................................... 63
6.3 SHF-P of a 90 mm round cup .............................................................. 67
7. Deep drawing at elevated temperature ........................................................... 70
7.1 Introduction.......................................................................................... 70
7.2 Experimental Setup.............................................................................. 73
7.2.1 Servo press and its kinematics ................................................. 73
7.2.2 Design of the tooling ................................................................. 75
7.3 Issues at the interface in forming at elevated temperature................. 78
7.4 Heating the blank ................................................................................. 79
7.4.1 Effect of the interface pressure on the hardness and the surface
roughness ............................................................................................................. 82
7.4.1.1 Hardness Measurements................................................. 83
7.4.1.2 Surface roughness measurements .................................. 85
7.4.1.3 Dome formation in the sheet under pressure................. 86
7.5 Lubricant evaluation for warm deep drawing process........................ 88
7.6 Preliminary Experiments ..................................................................... 92
7.7 Process Optimization / Windows......................................................... 94
7.7.1 Effect of constant forming velocity and temperature on
deformation.......................................................................................................... 97
7.7.1.1 Results for Al 5754-O...................................................... 98
7.7.1.2 Results for Mg AZ31-O (Supplier A) ............................ 103

xi
7.7.1.3 Results for Mg AZ31-O (Supplier B)............................. 114
7.8 Effect of Variable Forming Velocity................................................... 115
7.9 Conclusions ........................................................................................ 118
8. Modeling of non-isothermal deep drawing process ..................................... 121
8.1 Determination of the heat transfer coefficients................................. 122
8.1.1 Inverse Analysis...................................................................... 123
8.1.2 Setup of the inverse analysis problem................................... 124
8.2 Non-isothermal deep drawing of Mg AZ31-O................................... 125
8.2.1 Mg AZ31-O flow stress in literature....................................... 125
8.2.2 FE Modeling in LS-Dyna3D.................................................... 127
8.2.2.1 Plastic-thermal material model available in LS-Dyna.. 128
8.2.2.2 Thermal contact definition ........................................... 131
8.2.3 Simulation matrix ................................................................... 133
8.2.4 Comparison of FE predictions and experimental results....... 135
8.3 Non-isothermal deep drawing of Al 5754-O ..................................... 140
8.3.1 Yield criterion adopted in the FE model for the Al 5754-O... 142
8.3.2 Flow stress curves for aluminum........................................... 143
8.3.3 Comparison of numerical predictions with experimental
measurements .................................................................................................... 145
8.4 Non-isothermal modeling of SHF-P................................................... 148
8.4.1 Conclusions............................................................................. 152
9. Determination of drawability using fracture criterion ................................. 153
9.1 Cockcroft & Latham ductile fracture criterion................................... 153
9.2 Approach............................................................................................ 154
9.3 Setup of the FE model for tensile test and deep drawing ................. 157
9.4 Results and discussion....................................................................... 158
9.4.1 Determining the Critical Damage Value................................. 158
9.4.2 Deep drawing analyses ........................................................... 160
9.5 Conclusions........................................................................................ 163
10. Summary and conclusions .......................................................................... 165
Bibliography....................................................................................................... 169

xii
LIST OF TABLES

Table
3.1 List of wrought Mg alloys and their product range [Avedesian et al. 1999] .. 9

4.1 Hensel coefficients obtained from elevated temperature forming of
magnesium sheets in tensile test [Droeder, 1999] ........................................... 29

4.2 Experimental matrix for high formability sheets (strain rate values are
approximate)........................................................................................................... 31

4.3 Experimental matrix for low formability sheets (strain rate values are
approximate)........................................................................................................... 32

4.4 Thickness measurements for high and low formability sheets...................... 33

4.5 Comparison of radius values (R
1
& R
2
) obtained by using 5 and 3 points.... 39

4.6 Measured and calculated thickness values at the apex ................................... 44

6.1 Curvilinear length and percentage stretch of the sheet in the sheet radius
zone according to Figure 6.7 (pressure curve: P4)........................................... 67

6.2 Friction coefficients used in the FEA.................................................................. 68

7.1 Hardness (Brinell) measurements of Al 5052-H32 ........................................... 84

7.2 Hardness (Brinell) measurements of Mg AZ31-O (Supplier A)...................... 84

7.3 Hardness (Brinell) measurements of Mg AZ31-O (Supplier B) ...................... 85

7.4 List of lubricants and experimental conditions................................................. 88

7.5 Summary of the preliminary screening experiments for Al 5754-O, Al 5052-
H32 and MgAZ31-O.............................................................................................. 93

7.6 Experimental results for Al 5754-O..................................................................... 98


xiii
7.7 Experimental results for Mg AZ31-O (Supplier A)......................................... 104

7.8 Experiments at 225
o
C.......................................................................................... 114

7.9 Experiments at 250
o
C and lower velocities..................................................... 115

7.10 Significant savings in drawing time is obtained through the use of variable
forming velocity................................................................................................... 118

8.1 AZ31B-O simulation matrix................................................................................ 134

8.2 Summary of input data used for AZ31B-O simulations ................................ 135

8.3 List of input parameters to the FE model (Al 5754-O) ................................... 141

9.1 Simulation matrix for the deep drawing analysis........................................... 158

10.1 Summary of obtained drawability through the use of different
manufacturing processes.................................................................................... 167

A.1 Calculated flow stress values and related process parameters........180

A.2 Press characteristics.182

B.1 Mesh Density Windows (same values used in S1, S2 and S3).194

B.2 Flow stress data of St14...195

C.1 Stainless steel blank dimensions used for the drawability investigation (D
B
= blank diameter, D
P
= punch diameter)....200

C.2 Blank dimensions and process conditions...201


xiv

LIST OF FIGURES
Figure
3.1 a) Uniform and post-uniform strains vs. temperature for different strain
rates and b) variation of total elongation with temperature and strain rate
for alloy 5182-O [Ayres, 1977] .............................................................................. 8

3.2 Temperature dependent flow stress of Mg AZ31B alloy determined by
tensile test [Doege et al. 2001]............................................................................. 11

3.3 Flow stress at different temperatures for Mg AZ31B sheet with different
temper [Doege et al. 2001].................................................................................... 11

3.4 Stress-strain data for AZ31B plate deformed in in-plane tension (IPT) and
compression (IPC) and through-thickness compression (TTC) [Agnew,
et.al., 2002] ............................................................................................................. 13

3.5 Compression flow curves as a function of testing temperature. The
transverse (open symbol) samples are stronger than the rolling (closed
symbols) at all temperatures................................................................................ 14

3.6 A comparison of experimental (symbols) and simulated (curves)
compressive flow behavior at a) RT, b) 150C, c) 175C and d) 200C........ 15

3.7 Measured (symbol) and predicted tensile r-values.......................................... 16

3.8 Plot of the variation in r-values after compression (strain ~0.11) at different
temperatures. (symbols are experiments curves are predictions) [Jain, 2005]
................................................................................................................................... 17

4.1 Initial (left) and the deformed sheet (right) in the hydraulic bulge test ...... 19

4.2 Geometrical and process related parameters .................................................... 19

4.3 Elevated temperature Hydraulic Bulge Tooling (PtU, TU Darmstadt)......... 21

4.4 Hoop (
1
) and Transverse (
2
) Stresses and Dome Radii in a Membrane . 22

xv

4.5 Comparison of Hensel Model and Experimental Data [Droeder, 1999]....... 28

4.6 Thickness measurement along Rolling Direction (RD), and Transverse
Direction (TD) (dimensions in mm) ................................................................... 32

4.7 Pressure-bulge height curves at the approximate strain rate of 0.25s
-1
........ 34

4.8 Pressure-bulge height curves at the approximate strain rate of 0.025 s
-1
.... 35

4.9 Measurement points on the bulged sheet.......................................................... 37

4.10 Comparison of calculated and measured bulge radius values .................... 38

4.11 Radius difference using 5 and 3 points ........................................................... 39

4.12 Residual plots for sample at h
d
=14.7 mm....................................................... 40

4.13 Residual plots for sample at h
d
=21.7 mm....................................................... 41

4.14 Residual plots for samples at h
d
=33.9 mm..................................................... 41

4.15 Thickness distribution along the RD (bulge height (h)=~21mm and h33
mm) .......................................................................................................................... 43

4.16 Schematic representation of the strain rate and flow stress gradients in the
bulged sample ........................................................................................................ 45

4.17 True stress and true strain curves at 0.25 s
-1
.................................................. 46

4.18 True stress and true strain curves at 0.025 s
-1
................................................ 47

4.19 True stress true strain curve with possible errors at 225 C (0.025 s
-1
).... 48

5.1 Axisymmetric ......................................................................................................... 51

5.2 Top view of drawn sheet ...................................................................................... 51

5.3 Stress on an............................................................................................................. 51

5.4 Thickening of the sheet in the flange area ........................................................ 51

xvi

5.5 Definition of cup height (h) for a cup with a flange and without a flange.. 53

5.6 Deep drawn cup with ears.................................................................................... 53

5.7 Schematic view of the critical stroke (S
cr
) before (left) and after (right)...... 55

5.8 RT drawing of Al 5754-O ..................................................................................... 55

5.9 RT drawing of Mg AZ31-O................................................................................... 55

6.1 SHF-P process a) without leakage; b) with leakage ......................................... 58

6.2 Important tool parameters that influence the SHF-P process........................ 60

6.3 During a hydromechanical deep drawing process the sheet loses contact
with the die............................................................................................................. 61

6.4 Maximum thinning around the cup bottom radius at two different values of
the punch stroke.................................................................................................... 62

6.5 Position of the slab in the sheet radius zone (a) and equilibrium of the
forces in the slab (b).............................................................................................. 63

6.6 Pressure-stroke curves used in the critical stroke............................................ 66

6.7 Initial (a) and final (b) length of stretch in the sheet used to evaluate the
material stretching in the sheet radius zone .................................................... 67

6.8 Fluid pressure determined based on the critical stroke ................................. 69

6.9 Thinning distribution along the 45 cup wall .................................................. 69

7.1 Open (left) and closed (right) condition of the tool......................................... 72

7.2 AIDA servo-mechanical press drive mechanism............................................. 74

7.3 Ram motion of the servo press (TDC: top dead center, BDC: Bottom dead
center) ...................................................................................................................... 74

7.4 Schematic view and the dimensions of the tool (dimensions are in mm) .. 76

xvii

7.5 Warm forming with in-die dwelling process sequence .................................. 76

7.6 Open warm forming tooling (top die set on the left, bottom die set on the
right) [Kaya, et.al., 2006] ...................................................................................... 77

7.7 Assembled tooling on the Aida servo press...................................................... 77

7.8 110 ton Aida Servo Press ..................................................................................... 78

7.9 Top view of the fixture used to determine the dwell time necessary to heat
the blank (thickness 3mm) .................................................................................. 80

7.10 Schematic view of the experimental setup with the affected interfaces... 80

7.11 Temperature-time curves obtained with the test fixture for different
interface (blank holder) pressures (tool temperature=300
o
C) ..................... 81

7.12 Punch and knockout pin temperatures at tool temperatures of 250
o
C, 275
o
C and 300
o
C.......................................................................................................... 82

7.13 R
a
values for 5754-O and 5052-H32 ................................................................. 85

7.14 R
a
values for Mg AZ31-O (Supplier A) and Mg AZ31-O (Supplier B) ....... 86

7.15 BHP dome height at T=300 C........................................................................ 87

7.16 Change in the sheet diameter with respect to the BHP (Sheet Diameter:
100 mm) .................................................................................................................. 87

7.17 Thickness distributions for different lubricants under same process
conditions [Kaya, et.al., 2006]............................................................................. 90

7.18 Cup with Lube A after drawing [Kaya, et.al., 2006] ...................................... 90

7.19 Al cups formed with Lube C, Lube B and Lube A (from left to right) ....... 91

7.20 Mg cups formed with Lube C, Lube B and Lube A (from left to right) ...... 91

7.21 Adopted methodology for preliminary experimentation ............................. 92


xviii
7.22 Variation of LDR with punch velocity (Mg AZ31-O, T=250
o
C) ................ 95

7.23 Variation of LDR with punch velocity (Mg AZ31-O, T=275
o
C) ................ 95

7.24 Variation of LDR with punch velocity (Mg AZ31-O, T=300
o
C) ................ 96

7.25 Process window for the Mg AZ31-O alloy ...................................................... 96

7.26 Process window for the Al 5754-O alloy......................................................... 97

7.27 Effect of temperature on thickness distribution (5 mm/s, DR 2.5) ............. 99

7.28 Effect of temperature on thickness distribution (15 mm/s, DR 2.5) ......... 100

7.29 Punch load stroke curves at 5mm/s for different temperatures ............. 101

7.30 Punch load stroke curves at 15mm/s for different temperatures........... 101

7.31 Change in cup bottom temperature at 5 mm/s (Punch temp ~55
o
C, 65
o
C,
73
o
C)...................................................................................................................... 102

7.32 Change in cup bottom temperature at 15 mm/s (Punch temp: ~55
o
C, 65
o
C) ........................................................................................................................... 102

7.33 Effect of temperature on thickness distribution (5 mm/s, DR 2.5) ........... 105

7.34 Effect of temperature on thickness distribution (15 mm/s, DR: 2.5) ........ 105

7.35 Effect of temperature on thickness distribution (50 mm/s, DR: 2.5) ........ 106

7.36 Effect of forming velocity on thickness distribution (250
o
C, DR: 2.5) .... 106

7.37 Effect of forming velocity on thickness distribution (275
o
C, DR: 2.5) .... 107

7.38 Effect of forming velocity on thickness distribution (300
o
C, DR: 2.5) .... 107

7.39 Punch load stroke curves at 5mm/s for different temperatures ............. 108

7.40 Punch load stroke curves at 15mm/s for different temperatures........... 109

7.41 Punch load stroke curves at 50mm/s for different temperatures........... 109

xix

7.42 Change in cup bottom temperature at 5 mm/s............................................. 110

7.43 Change in cup bottom temperature at 15 mm/s........................................... 111

7.44 Change in cup bottom temperature at 50 mm/s........................................... 111

7.45 Pictures of the formed cups from Al 5754-O (left) and Mg AZ31-O
(Supp.A) (right) .................................................................................................... 112

7.46 Effect of blank temperature and punch velocity upon wall thinning at the
bottom corner of the drawn cup for the Al 5754-O alloy............................. 113

7.47 Effect of blank temperature and punch velocity on thinning at the bottom
corner of the drawn cup for the Mg AZ31-O alloy........................................ 113

7.48 Effect of variable forming speed on the thickness distribution of the
drawn Al cups...................................................................................................... 117

7.49 Effect of variable forming speed on the thickness distribution of the
drawn Mg cups..................................................................................................... 117

8.1 Experimental and calculated temperature-time curves................................ 123

8.2 Schematic view of the FE model....................................................................... 124

8.3 Modeling of the measurement fixture with the temperature measurement
point ....................................................................................................................... 124

8.4 Calculated heat transfer coefficients ................................................................ 125

8.5 Graphical representation of the database for Magnesium AZ31 Tensile
tests conducted at different strain rates and temperatures [Sivakumar, et al,
2006]. ..................................................................................................................... 127

8.6 FE model for non-isothermal simulations of deep drawing of round cups
................................................................................................................................. 128

8.7 Contact-Forming card used to define mechanical and thermal contact
parameters. ........................................................................................................... 133


xx
8.8 Experimental and computed punch load vs. stroke curves (left) and cup
thinning distributions (right). (RD=rolling direction; TD=transverse
direction) ............................................................................................................... 136

8.9 Experimental and computed punch load vs. stroke curves (left) and cup
thinning distributions (right). (RD=rolling direction; TD=transverse
direction) ............................................................................................................... 137

8.10 Experimental and computed punch load vs. stroke curves (left) and cup
thinning distributions (right). (RD=rolling direction; TD=transverse
direction) ............................................................................................................... 137

8.11 Experimental and computed punch load vs. stroke curves (left) and cup
thinning distributions (right). (RD=rolling direction; TD=transverse
direction) ............................................................................................................... 138

8.12 Experimental and computed punch load vs. stroke curves (left) and cup
thinning distributions (right). (RD=rolling direction; TD=transverse
direction) ............................................................................................................... 138

8.13 Experimental and computed punch load vs. stroke curves (left) and cup
thinning distributions (right). (RD=rolling direction; TD=transverse
direction) ............................................................................................................... 139

8.14 Experimental and computed punch load vs. stroke curves (left) and cup
thinning distributions (right). (RD=rolling direction; TD=transverse
direction) ............................................................................................................... 139

8.15 Experimental and computed punch load vs. stroke curves (left) and cup
thinning distributions (right). (RD=rolling direction; TD=transverse
direction ................................................................................................................ 140

8.16 Comparison of experimental yield loci and those predicted by the Von
Mises, Hill, Tresca, Logan-Hosford and Barlat criteria under biaxial stress
condition for Al 5083-O alloy sheet [Naka, et.al., 2003] .............................. 143

8.17 Extrapolated flow stress versus temperature for different strain
rates(original data from Boogard, 2001) .......................................................... 144

8.18 Elimination of the softening behavior from the stress-strain curve......... 144


xxi
8.19 Comparison of punch load predictions using various heat transfer
coefficients (HTC: kW/m
2
C) with experiment (5 mm/s at 250
o
C)............. 146

8.20 Predicted thickness distribution comparison with various heat transfer
coefficients (5 mm/s at 250
o
C)......................................................................... 146

8.21 Comparison of punch load prediction with experiment (5 mm/s at 275
o
C)
................................................................................................................................. 146

8.22 Predicted thickness distribution comparison with experiments (5 mm/s at
275
o
C).................................................................................................................... 146

8.23 Comparison of punch load prediction with experiment (5 mm/s at 300
o
C)
................................................................................................................................. 147

8.24 Predicted thickness distribution comparison with experiments (5 mm/s at
300
o
C).................................................................................................................... 147

8.25 Comparison of punch load prediction with experiment (15 mm/s at 275
o
C) ........................................................................................................................... 147

8.26 Predicted thickness distribution comparison with experiments (15 mm/s
at 275
o
C) ............................................................................................................... 147

8.27 Predicted thickness distribution comparison with experiments (5 mm/s at
300
o
C).................................................................................................................... 148

8.28 Predicted thickness distribution comparison with experiments (5 mm/s at
300
o
C).................................................................................................................... 148

8.29 Interfaces in elevated temperature SHF-P process ...................................... 149

8.30 Pressure and BHF used in the simulations................................................... 150

8.31 Temperature distribution of the sheet at the end of the stroke................. 150

8.32 Temperature distribution of a sheet formed using SHF-P ......................... 151

8.33 Temperature distribution of a deep drawn sheet ........................................ 151

9.1 Flow stress curve of Al 5754-O obtained from the bulge test...................... 155

xxii

9.2 Geometry of the tensile specimen (ASTM, E 8M-04).................................... 156

9.3 Flow stress curves used in the FEA.................................................................. 157

9.4 Point of instability in the load-stroke curve during the tensile test ........... 159

9.5 CDV (when necking starts during 3D tensile test simulation) for Al 5754-O
is 0.428................................................................................................................... 160

9.6 CDV (when necking starts during 3D tensile test simulation) for Al 6061-T6
is 0.148................................................................................................................... 160

9.7 Maximum CDV obtained for the successfully deep drawn Al 5754-O (LDR:
2.1) cup is 0.352................................................................................................... 161

9.8 Max CDV reaches 0.42 for the unsuccessfully deep drawn Al 5754-O (LDR:
2.2) cup.................................................................................................................. 161

9.9 Maximum CDV obtained for Al 6061-T6 (LDR: 1.6) is 0.175...................... 163

9.10 Maximum CDV obtained for Al 6061-T6 (LDR: 1.5) is 0.165.................... 163

A.1 Flow stress curves of the tested aluminum alloys.181

A.2 Thinning measurements are done in Direction 1-3 and Direction 2-
4.183

A.3 Comparison of the predicted and measured pressure values for 6061-
O.185

A.4 Thinning comparison in Direction 1-3; at corner 1 the difference is 14 (%)
while at corner 3 it is 11 (%)....185

A.5 Thinning comparison in Direction 2-4;at corner 2 the difference is 10 (%)
while in the corner 4 is 1.5 (%)....186

A.6 Pressure comparison; max pressure for simulation is 199 (Bar) while
measured pressure value from the experiments is 208 (Bar).....187


xxiii
A.7 Thinning comparison in Direction 1-3; at corner 1 the difference is 2 (%)
while at corner 3 it is 7 (%)...188

A.8 Thinning comparison in Direction 2-4; at corner 2 the difference is 3 (%)
while at corner 4 it is 2 (%)...189

A.9 Pressure comparison; max pressure for simulation is 192 (Bar) while
measured pressure value from the experiments is 218 (Bar).....190

A.10 Thinning comparison in Direction 1-3; at corner 1 the difference is 4 (%)
while at corner 3 it is 10 (%)....191

B.1 Mesh Density Windows for the blank holder.....194

B.2 Mesh Density Windows for the die-ring.........195

B.3 Predicted elastic deflection is shown with the dashed line.197

B.4 Predicted elastic deflection is shown with the dashed line [Kaya, et.al.,
2004]....197

C.1 Limiting DR - Die/BH temperature (forming velocity 2.5 mm/s)202

C.2 Geometrical parameters (height and external diameter) of the cups drawn
at different conditions...202

C.3 Thinning distribution along rolling and transverse direction for cup A and
cup B.203

C.4 Thinning distribution along rolling and transverse direction for cup C and
cup D.204

C.5 Limiting DR - Die/BH temperature (forming velocity 25 mm/s).205

C.6 Limiting DR - Die/BH temperature (forming velocity 50 mm/s).206


1

CHAPTER 1


INTRODUCTION




Weight reduction while maintaining functional requirements is one of
the major goals of engineering design and manufacturing so that materials,
energy, and costs are saved and environmental damage is reduced. Magnesium
(Mg) and Aluminum (Al) alloys offer great potential to reduce weight by
displacing the most commonly used materials, i.e. steel and polymers, because
of their high strength to weight ratio. Other important factors in selecting Mg
and Al alloys for engineering applications, compared to other engineering
materials include their thermal properties, damping capacity, fatigue
properties, dimensional stability and easy machinability [Avedesian et al.
1999, Naka et al. 2001]. Besides Al and Mg, it is important to mention weight
reduction can also be achieved by using thinner gauge steels (high strength
steel, stainless steel) and forming them using advanced forming processes.
The use of forming technology for Al and Mg alloy sheet is restricted
because of the low formability of these materials especially at room
temperature. Therefore, advanced forming methods are needed in order to
improve their formability. Some improvement is believed to be obtained when

2
drawing these materials against pressure (sheet hydroforming) and at elevated
temperatures (warm forming). Investigation of the formability limits of
aluminum and magnesium sheet is done by using these advanced forming
techniques.
This study focuses on 1) the design and development of forming process
with emphasis on a) material properties, b) deformation mechanics c) the
forming temperatures, d) the interface condition (friction, lubrication and heat
transfer), e) the tool temperature, f) tooling/ equipment design, and g) constant
and variable forming speed (or strain rate); 2) the influence of the forming
equipment (servo motor driven press) on the process and the final product and
3) computational modeling of the process.
Al and Mg alloys show increased formability especially at temperature
range of 200
o
C to 300
o
C. Currently formed Al alloy components find
applications only as shallow parts in automobile body panels and chassis
applications because of their low room temperature formability. Warm forming
technology for Al alloys has been investigated by Bolt et al., 2001a, Bolt et al.,
2001b, Boogard et al., 2001a, Boogard et al., 2001b, Wu et al. 2001, Taylor et
al. 1980, Naka et al. 2001, Shehta et al. 1978, Li et al., 2000, Groche 2001.
They have found that 5xxx series and 6xxx series Al alloys show increased
formability at the range of 250
o
C to 300
o
C. However, there is a lack of

3
sufficient knowledge on warm sheet forming processes which limited the
practical use of these alloys.

4
CHAPTER 2


OBJECTIVES

The overall objective of the proposed research is to increase the
formability (drawability) limits of aluminum and magnesium using advanced
forming methods.
Thus the major specific objectives of this study are;
1. Determine elevated temperature mechanical properties of magnesium
sheet under equi-biaxial state of stress using hydraulic bulge test
2. Determine the influence of the process parameters in hydroforming of Al
and Mg sheet (SHF-P / Sheet hydroforming with punch) and increase the
drawability
3. Investigate the use of a servo motor driven press for in-die-dwelling
using a warm deep drawing tool to improve warm drawability of
aluminum, magnesium and stainless steel alloys
4. Determine the effect of process parameters (punch velocity, lubrication,
temperature, interface pressure) and geometrical parameters (punch and
die radii, initial blank size) on metal flow in deep drawing round Al and
Mg alloy cups.

5
5. Develop FE models of the deep drawing and hydroforming process by
applying the appropriate pressure, temperature, interface and velocity
boundary conditions under isothermal and non-isothermal conditions,
and compare predictions with experimental data.
6. Develop a methodology to minimize experimentation to determine the
drawability of aluminum sheet.


6

CHAPTER 3


MECHANICAL PROPERTIES OF ALUMINUM AND MAGNESIUM ALLOYS

Most common testing methods to investigate the formability are uniaxial
tension, plane strain tension, bending, Limiting Dome Height (LDH) test and
drawing test. Tensile test provides information that characterizes materials
potential in formability and delivers qualitative information on the yield point
and the stress levels at the onset of necking in uniaxial state of stress. However,
most of the deformation in real forming operations occurs mostly in biaxial
even in triaxial state of stress.
It is possible to improve the formability of lightweight alloys through
forming operations at elevated temperatures. [Taylor et. al, 1976] conducted
their investigations using tensile test in the strain rate range of 7x10
-4
sec
-1
- 3
sec
-1
for alloys 5182-O, 5085-H11 and 5090-H34, and found a substantial
increase in the tensile elongation at a temperature of 473
o
K. In particular, their
results indicated that the work hardening rate and uniform elongation
decreased but the strain rate sensitivity and total elongation increased with
increasing temperature. In other words, the increase in the total elongation is
exclusively contributed by the post-uniform elongation. Figure 3.1 shows the
results obtained by [Ayres, 1977] for Al alloy 5182-O. It is clear from Figure

7
3.1a that, as temperature increases, the post-uniform strain increases while the
uniform strain decreases, and this trend is enhanced by decreasing strain rate.
Since the increase in the post-uniform strain dominates that in the uniform
strain, the total elongation increases with temperature at various strain rates
(Figure 3.1b) for temperatures of 373
o
K and higher.





8


Figure 3.1: a) Uniform and post-uniform strains vs. temperature for different
strain rates and b) variation of total elongation with temperature and strain rate
for alloy 5182-O [Ayres, 1977]
3.1 Forming behavior of Mg alloy sheet
Mg is the lightest of all commonly used structural metals with a density of
approximately 2/3 that of aluminum. Commercially available Mg alloy is
classified as zirconium containing alloys and zirconium free alloys. A
comprehensive list of Mg wrought alloys is shown in Table 3.1.

9

Forging
Extruded bars and
tubes
Sheet
Mg-Al-Mn-Zn
alloy system
M1A-F, AZ31B-F,
AZ61A-F, AZ80A-T5,
AZ80A-T6
M1A-F, AZ10A-F,
AZ31B-F, AZ61A-F,
AZ80A-T5
AZ31B
Mg-Zn-Cu
alloy system
ZM21-F ZC71-T6, ZM21-F ZM21
Zirconium
free alloys
Mg-Li alloy
system
Currently under
investigation
Currently under
investigation
Currently
under
investigation
Zirconium
containing
alloys
Mg-Zn-Zr
alloy system
ZK31-T5, ZK60A-T5,
ZK61-T5
ZK21A-F, ZK31-T5,
ZK60A-T5
.

Table 3.1: List of wrought Mg alloys and their product range [Avedesian et al.
1999]
The limited formability of Mg sheet at room temperature is due to the
hexagonal closed pack (hcp) lattice structure that has three slip systems
compared to the twelve slip systems in face centered cubic (fcc) and body
centered cubic (bcc) lattice structures. Therefore it is better to form Mg alloys at
elevated temperatures (200-300
o
C), which activates additional slip planes
thereby improving the material formability.
Mg alloy sheets, AZ31B and ZM21, which are commonly produced, are hot
rolled at temperatures of 315
o
C - 370
o
C to the final thickness. The amount of
work hardening remaining in the material and the amount of annealing that
takes place during and after final pass rolling is critical in forming applications.

10
[Doege et al., 2001] conducted exhaustive investigation on the plastic material
properties of Mg alloy sheet AZ31B at room temperature by conducting
uniaxial tensile tests. The Mg alloy showed an elongation of 17% and strain
hardening coefficient of 0.177. On the other hand, Mg alloy sheet (AZ31B)
shows remarkable increase in the formability in tensile test at temperatures
above 200
o
C as shown in Figure 3.2. The effect of rolling condition on the flow
stress decreases with increase in the temperature as shown in Figure 3.3.
[Takuda et.al., 2005] has conducted uniaxial tensile tests at constant crosshead
velocities (0.02 mm/s, 0.2 mm/s, 2 mm/s and 20 mm/s) between 150
o
C and 300
o
C and obtained strain values up to 0.7.
Most of the available literature on the formability of Mg is on the uniaxial
testing. Pneumatic bulge tests conducted at IFU of the University of Stuttgart
also showed significant formability increase in the Mg alloy sheet [Siegert,
2003] and [Siegert, 2004]. However, information on properties of Mg alloy
sheet obtained at elevated temperatures under equi-biaxial state of stress is very
limited. Also the geometrical evolution of the dome under equi-biaxial state of
stress, various temperatures and strain rates has not been studied.

11

Figure 3.2: Temperature dependent flow stress of Mg AZ31B alloy determined
by tensile test [Doege et al. 2001]

Figure 3.3: Flow stress at different temperatures for Mg AZ31B sheet with
different temper [Doege et al. 2001]
3.2 Strength asymmetry in Mg alloy sheets
Magnesium alloy sheets tend to exhibit asymmetric behavior in yielding (much
higher yield stress during in-plane tension than compression) due to different

12
metallurgical mechanisms, which operate under different loading conditions
[Agnew et al, 2002] (Figure 3.4). Asymmetric deformation behavior is a
characteristic of materials having a non-cubic crystal structure and does not
exist in polycrystalline cubic materials, due to their more symmetric structure.
The phenomenon is generally associated with mechanical twinning [Agnew, et
al, 2001]. [Klimanek, et.al., 2002] has investigated the microstructure evolution
under compressive plastic deformation of magnesium at different temperatures
and strain rates and provided stress-strain curves obtained under compression.
In the case of magnesium alloy sheet, a relatively soft deformation twinning
mechanism operates during in-plane compression, while only dislocation slip
(in poorly oriented grains) is possible during tension. Hence, the yield stress in
compression is much lower than in tension. It is important to note that
twinning is more common at high strain rates and low temperatures [Meyers et
al, 1999]. Early studies show that for Mg alloys (HM21A-T8, HK31A-H24 and
HM31A-T5) tensile yield strengths are considerably higher than compressive
yield strengths at room temperature. However, at 204 C, the differences
between tensile and compressive yield strengths decrease to an average value of
7.5 MPa. Above 315 C, the tensile and compressive yield strengths are equal
[Fenn, 1961].

13

Figure 3.4: Stress-strain data for AZ31B plate deformed in in-plane tension
(IPT) and compression (IPC) and through-thickness compression (TTC)
[Agnew, et.al., 2002]
Figure 3.5 shows the flow curves obtained from rolling direction (RD) and
transverse direction (TD) compression tests at different temperatures using Mg
alloy AZ31-H24. The samples were strained to strain of 1.0 or failure,
whichever occurred first. It is clear to see that the TD samples are stronger at
every temperature, as was observed in the tensile tests conducted by [Duygulu,
et.al., 2005]. At 250
o
C, the difference in the flow curves from RD and TD is
extremely small. Strong sigmoidal hardening is observed at room temperature
and mildly elevated temperature tests, while essentially no hardening is
observed at the higher temperatures (200
o
C & 250C). Significantly, there is no

14
softening in the initial yield stress with an increase in temperature up to 200
C, in fact there is a slight increase in strength.


Figure 3.5: Compression flow curves as a function of testing temperature. The
transverse (open symbol) samples are stronger than the rolling (closed symbols)
at all temperatures

Figure 3.6 shows the experimental flow behavior obtained at normal direction
(ND), RD and TD compression and TD and RD tension of Mg AZ31 alloy at
room temperature, 150
o
C, 175
o
C and 200
o
C.
Experimental findings can be summarized as;
i) The flow stress curve obtained from ND compression is different than
the TD and RD compression curves,

15
ii) At 200
o
C, the variation in the flow stress curves obtained at RD, TD
and ND is smaller than the ones obtained at 175
o
C.


(a)

(b) (c) (d)
Figure 3.6: A comparison of experimental (symbols) and simulated (curves)
compressive flow behavior at a) RT, b) 150C, c) 175C and d) 200C

16
[Duygulu, et.al., 2005] has determined the r-values in uniaxial tension along
the RD and TD using the Mg AZ31-H24 alloy. Figure 3.7 shows that the r-value
along the TD are higher (up to 5) than the ones in the RD (up to 2) and it
decreases with temperature. In contrast with these observations of high r-
values during in-plane tension, same samples compressed within the rolling
plane exhibit very low r-values (Figure 3.8).
Although the strengths along the three directions become quite similar at the
highest temperature T = 250C, the r-values remain distinct at all temperatures
[Jain, 2005].

Figure 3.7: Measured (symbol) and predicted tensile r-values


17

Figure 3.8: Plot of the variation in r-values after compression (strain ~0.11) at
different temperatures. (symbols are experiments curves are predictions) [Jain,
2005]

18
CHAPTER 4

DETERMINATION OF THE FLOW STRESS OF LIGHTWEIGHT ALLOY
SHEET UNDER EQUI-BIAXIAL STATE OF STRESS

4.1 Introduction
Properties of Mg alloys at elevated temperatures have been determined
by various researchers around the world. However, information on properties
obtained at elevated temperatures under equi-biaxial state of stress using
hydraulic bulge test is limited. Materials are often tested using the standard
uniaxial tensile test. Stress conditions in stamping, however, are not uniaxial as
they are in the tensile test. Therefore, it is necessary to obtain material
properties under biaxial deformation conditions (Figure 4.1). In this test, the
sheet is clamped between the lower and the upper die. When the fluid in the
lower chamber is pressurized, the sheet is bulged into the cavity of the upper
die. The clamping force between the lower and upper die has to be high
enough to prevent sliding of the sheet between the dies. Often, a lockbead is
used to prevent the movement of the sheet in the clamped region. Thus, the
sheet will only be stretched and no draw-in will occur. When the deformation
of the material exceeds its formability limit, the bulged sheet will fracture. In
this test, the deformation is not affected by friction. Thus, the reproducibility of
the test results is good.

19
p
Lower Die
Upper Die
Pressurized
Fluid
Initial Sheet
Bulged Sheet

Figure 4.1: Initial (left) and the deformed sheet (right) in the hydraulic bulge
test
The main geometrical and process related parameters of the hydraulic bulge
test are shown in Figure 4.2.
p
h
d
d
c
t
d
t
0
R
c
R
d
F
c F
c

Figure 4.2: Geometrical and process related parameters
t
0
initial thickness of the sheet
t
d
thickness at the top of the dome
h
d
dome height
R
d
radius at the top of the dome
d
c
diameter of the cavity
R
c
radius of the fillet of the cavity
F
c
clamping force

20
p hydraulic pressure

The elevated temperature bulge tests were conducted in co-operation with the
PtU of the University of Darmstadt that has the appropriate apparatus available.
The specific objectives in conducting these tests were to;
a) gain experience and observe difficulties/advantages of using
hydraulic bulge test tooling submerged in heated heat transfer liquid
and check the applicability of the unique submerged testing concept
b) obtain the mechanical properties of Mg AZ31-O alloy at various
temperatures and approximate strain rates under equi0biaxial state of
stress.
4.2 Experimental setup
Figure 4.3 shows the elevated temperature hydraulic bulge tooling used
in this study. In this unique set-up, the die, the blankholder and the sheet are
submerged in the heated pressure medium. Thus the temperature variations in
the tool and the sheet are reduced. The pressure medium is heated via a)
cartridge heaters located at the bottom of the tool (Figure 4.3), b) cartridge
heaters in an outside tank and c) circulation pump equipped with heaters. A
potentiometer is used to record the bulge height while the medium pressure is
measured with a pressure transducer. A constant 90 bar blank holder pressure

21
(BHP) was applied to lock the sheet to prevent the draw-in of the sheet into the
die cavity. The die corner radius (R
c
) of the die was 4 mm. The heated pressure
medium used was Multidraw Hydrofluid HT 400 and it had a flash point of
280
o
C.
While the tool/liquid temperature was 275
o
C at the bottom of the tool,
the maximum sheet temperature was 225
o
C. Therefore, for safety reasons,
experiments were conducted up to sheet temperatures of 225
o
C.

Figure 4.3: Elevated temperature Hydraulic Bulge Tooling (PtU, TU Darmstadt)
In this test set-up the constant strain rate (estimated at the apex of the bulge)
could be reached only approximately, by controlling the appropriate flow rate
of the pressure medium [Groche et.al., 2002].

22
4.3 Calculation of stresses and strains under equi-biaxial state of stress
In hydraulic bulge test, equi-biaxial state of stress is achieved at the apex
during bulging. By using the membrane theory, which assumes that the
thickness stress is neglected, and by calculating the hoop and transverse
stresses, it is possible to obtain the effective stress values (Figure 4.4). The
parameters necessary for the calculation of the stresses are a) pressure
(obtained from experiment), b) instantaneous dome (bulge) radius (R
1
=R
2
) and
c) instantaneous thickness. Dome radius and thickness values are measured
from the bulged samples at various bulge heights. Since the capability of
measuring the dome radius and thickness measurements instantaneously did
not exist, analytical models were used to calculate the dome radius and
thickness. There is also a need to understand the applicability of these models
(membrane theory) at elevated temperature conditions.
1

R
1 R
2

Figure 4.4: Hoop (
1
) and Transverse (
2
) Stresses and Dome Radii in a
Membrane

23
4.4 Membrane theory
To determine the flow stress curve by using the hydraulic bulge test, the
most common theory used is the membrane theory [Panknin 1959], [Gologranc
1975]. The membrane theory neglects bending stresses. Thus, it is only
applicable for thin sheets. The following equation (1) represents this theory:
d
t
p
R R
= +
2
2
1
1


(1)
Where
1
and
2
are the principle stresses on the sheet surface, R
1
and R
2
are
the corresponding radii of the curved surface, p is the hydraulic pressure, and
t
d
is the sheet thickness. (Figure 4.2) For the axi-symmetric case of the
hydraulic bulge test,
1
=
2
and the radius at the top of the dome is
R
d
= R
1
= R
2
. Therefore, equation (1) can be simplified to:
d d
t
p
R
2
=


(2)
d
d
t 2
pR
=
(3)

24
In hydraulic bulge testing, pressure is applied on the internal sheet
surface. No normal forces acting on the outer sheet surface. Therefore, the
average stress in the sheet normal to the sheet surface is approximately:
2
p
2
0 p
n

=
+
=
(4)
The effective stress can be calculated by Trescas plastic flow criterion, since
Traesca and Von Mises criteria are the same at equi-biaxial state of stresses:
2
p
t 2
pR
d
d
min max

= =
(5)
|
|
.
|

\
|
+ = 1
t
R
2
p
d
d

(6)
The effective strain (thickness) is:
|
|
.
|

\
|
= =
0
d
t
t
t
ln
(7)
4.5 Calculation of the radius of the bulge (dome)
The radius at the top of the dome can be calculated by [Hill 1950]
assuming that the dome is spherical and that there is no fillet in the cavity of
the die:

25
d
2
d
2
c
d
h 8
h 4 d
R
+
=
(8)
Considering that there is a die radius R
c
, and assuming that the dome is
spherical, the radius of the dome can be calculated by [Hill 1950]:
d
d c
2
d
2
c
c
d
h 2
h R 2 h R
2
d
R
+
|
.
|

\
|
+
=
(9)
[Panknin 1959] investigated the hydraulic bulge test experimentally. He
measured the radius at the top of the dome of the deformed samples with radii
gages. He also calculated the radius at the top of the dome using the dome
height. He assumed that the dome is a part of a sphere and considered the fillet
in the cavity of the die. [Gologranc 1975] had similar results with his
experiments. Besides, he detected that for small dome heights the radius of the
dome is up to 10 % larger than the calculated ones.
Equation 9 is an analytical formula that can be used to calculate the bulge
radius at various bulge heights.
4.6 Calculation of the thickness at the top of the dome
[Hill 1950] developed analytical methods to describe the deformation in
the hydraulic bulge test. For his calculations, he assumed that the shape of the

26
bulge is spherical. With this assumption, the thickness at the top of the dome
can be calculated by the following equation:
2
2
c
d
0 d
d
h 2
1
1
t t
|
|
|
|
|
.
|

\
|
|
|
.
|

\
|
+
=
(10)
4.7 Stress Strain relationship at elevated temperature for magnesium
alloy sheet
For the description of flow behavior, it is necessary to express the flow
stress mathematically as a function of the relevant parameters true strain

,
true strain rate

&
and temperature T. For a given material and
microstructure, the flow stress can be expressed as:
( ) T f , ,
&
= (11)

At room temperature forming, Equation 11 can be described with the law of
Hollomon
( )
n
K = (12)
where K is a material specific constant factor, called strength coefficient.
Parameter n describes increasing hardening of the material with increasing
strain and is therefore called strain hardening exponent. In some cases, when

27
deformation takes place at higher temperatures (200
o
C-300
o
C) and at lower
strain rates, straining may continue with decreasing stress, i.e. work-softening
may occur. Using tensile tests conducted at elevated temperatures, several
approaches for describing both the softening effect and the influence of strain
rate were considered by [Brand 1998] in the following as;
( )
3 1 2 4
( )
m m T m m
T K T e e



=
&
(13)

This approach expresses the true strain dependent hardening behavior
with the additional term

4
m
e
, in which e is the base of natural logarithms. In
order to evaluate the accuracy of this approach tensile tests at elevated
temperatures were conducted at IFUM for magnesium sheet alloy AZ31B,
(thickness t
0
=1.3 mm) [Brand, 1998], [Droeder, 1999] and [Doege, 2001b] have
also used this model in their studies. Comparison of experimentally obtained
flow stress curves for AZ31-O with those of equation (13) are shown in Figure
4.5 for a temperature range of T= 150C to T= 235C.


28


Figure 4.5: Comparison of Brand model and experimental data [Droeder, 1999]

Coefficients m1, m2, m3 and m4 were derived from experimental data
using a regression analysis. (Table 4.1) Exponents m1 and m2 appeared to be
constant while m1 provides a measure for dependence of flow stress on
forming temperature. Constant m2 shows the dependence of flow stress on
strain rate. As in the power law, exponent m3 is the same as the hardening
coefficient n. In addition, exponent m4 indicates an additional dependence of
the flow stress in function of strain. As it is seen from Table 4.1, m4 is negative
when softening takes place. At T= 150C (m4=0) therefore no softening is
True strain
Y
i
e
l
d

S
t
r
e
s
s

[
M
P
a
]


29
present and an almost constant strain hardening occurs. Dependence of the K
value on temperature is also seen from Table 4.1.
Temperature [C] 150 200 235
Constant K [MPa s] 190 141 112
Exponent m
1
[-] 18x10
-4
18x10
-4
18x10
-4

Exponent m
2
[-] 10
-4
10
-4
10
-4

Exponent m
3
[-] 0.12 0.103 0.072
Exponent m
4
[-] 0 -0.52 -0.62
Table 4.1: Brand coefficients obtained from elevated temperature forming of
magnesium sheets in tensile test [Droeder, 1999]
4.8 Experimental investigation
4.8.1 Observations
Approximately 1.2 mm thick (5% thickness tolerance) square sheet
samples (260mm 260mm) were provided by Salzgitter A.G., Germany. Rolling
directions were marked on each specimen. Thickness measurements were
made on specimens to check the uniformity of thickness. The location of each
sample (260mm 260mm) in the original rolled sheet before cutting was not
known. Following observations were made before and during the experiments:
a) Specimens were difficult to etch with grids and some of the specimens
exhibited corrosion after the tests.

30
b) Before the tests, the specimens were not all flat and they exhibited
approximately 5 % variation in thickness (nominal thickness was 1.2
mm)
c) The tests showed considerable differences in formability (as measured
by bulge height at bursting) among the samples that were obtained from
the same coil. Furthermore, samples were separated as low and high in
formability. They exhibited color difference in surface.
d) In between tests, it was necessary to compensate for temperature losses
by using a circulation pump to circulate the pressure medium.
e) The temperature of the sheet was measured manually by a thermocouple
before the test. It was not possible to monitor the temperatures during
the test.
f) When the sample fractured, in some selected tests, due to impact at
bursting, the fixture holding the potentiometer hit the apex of the bulged
sheet. This impact created a dent at the apex, which is not very
important since the sheet is already burst. However, the impact might
damage the potentiometer or reduce its life.
4.8.2 Test conditions and experimental matrix
Due to a) the difficulty in pressure/temperature control and long test
cycles; b) unexpected material property variations and waste of samples c) time

31
delay because of excessive smoke generation at elevated temperatures,
experimentation is quite slow. Results of experiments will be explained in the
upcoming sections. Table 4.2 and Table 4.3 give a summary of the conducted
experiments for the high and the low formability sheets, respectively. It is
seen that, the test temperatures vary, because it is difficult to control the sheet
temperature precisely. Three samples were used for most of the test conditions.

Temperature (
o
C) Strain Rate (s
-1
)
20 0.25
145 0.25
180 0.25
205 0.25
217 0.25

20 0.025
165 0.025
190 0.025
210 0.025
225 0.025
Table 4.2 : Experimental matrix for high formability sheets (strain rate values
are approximate)







32

Temperature (
o
C) Strain Rate (s
-1
)
179 0.25
206 0.25

191 0.025
206 0.025
Table 4.3: Experimental matrix for low formability sheets (strain rate values are
approximate)
4.8.3 Measurement of initial sheet thickness
Figure 4.6 shows the measurement directions for the as-rolled sheets.
Thickness measurements were made for randomly selected three high
formability and three low formability sheets. Measurements are made at
every 20 mm increments along A-B and C-D using a micrometer and are shown
in Table 4.4.

Figure 4.6: Thickness measurement along Rolling Direction (RD), and
Transverse Direction (TD) (dimensions in mm)

33

High Formability Sheets Low Formability Sheets
Sheet1 Sheet 2 Sheet3 Sheet 4 Sheet 5 Sheet 6
A-B C-D A-B C-D A-B C-D A-B C-D A-B C-D A-B C-D
1 1.1938 1.1938 1.1684 1.1811 1.143 1.1684 1.16841.18111.1684 1.1938 1.1684 1.1938
2 1.1938 1.1938 1.1684 1.1684 1.143 1.1684 1.16841.16841.1684 1.1811 1.1684 1.1938
3 1.1938 1.1938 1.1684 1.1684 1.143 1.143 1.18111.16841.1684 1.1811 1.1684 1.1938
4 1.2065 1.1938 1.1684 1.1684 1.143 1.143 1.16841.16841.1684 1.1684 1.1684 1.1938
5 1.1938 1.1938 1.1684 1.1684 1.143 1.143 1.16841.16841.1684 1.1684 1.1684 1.1938
O-center 1.1938 1.1938 1.1684 1.1684 1.143 1.143 1.16841.16841.1684 1.1684 1.1938 1.1938
6 1.1938 1.1938 1.1684 1.1684 1.1557 1.143 1.16841.16841.1811 1.1684 1.1938 1.1938
7 1.1811 1.1938 1.1684 1.1684 1.1684 1.143 1.19381.16841.1938 1.1684 1.1938 1.1938
8 1.1938 1.1938 1.1684 1.1684 1.143 1.143 1.19381.19381.1938 1.1684 1.1938 1.1938
9 1.1938 1.1811 1.1684 1.1557 1.1557 1.143 1.19381.19381.1938 1.1684 1.1938 1.1938
10 1.1938 1.1684 1.1684 1.1557 1.143 1.143 1.20651.19381.1938 1.1684 1.1938 1.1938
Avg. 1.1938 1.1903 1.1684 1.1672 1.1476 1.1476 1.17991.17651.1788 1.173 1.1823 1.1938
Table 4.4: Thickness measurements for high and low formability sheets
It is seen from Table 4.4 that maximum and minimum thickness values
measured from the high formability sheets are 1.19 mm (sheet 1) and 1.14 mm
(sheet3), which corresponds to almost a 5% difference. When the average
thickness values are compared, variation can easily be seen between sheet 1,
sheet 2 and sheet 3. Thickness values are pretty close in A-B and C-D directions
for almost every sheet.
4.8.4 Sheet draw-in during bulging
During hydraulic bulging, the sheet is held in order to prevent draw-in
by using either a lockbead or applying higher clamping forces. In the set-up
that was used, there was no lockbead in the tool due to lower forming loads at
elevated temperatures. Therefore, all the square sheets (260 mm 260 mm)

34
side length was measured to check whether draw-in occurred during
deformation. As a result of measurements, it is seen that no draw-in to the die
cavity happened.
4.8.5 Pressure bulge height curves
Figure 4.7 shows the pressure-bulge height curves at various
temperatures for the approximate strain rate of 0.25 s
-1
. It is seen that at
temperatures above 180 C, bulge height increases with decreasing pressure.
This is possibly a sign of thermal/work softening effect. Three samples were
bulged at the same condition for understanding the experimental repeatability.
The variation between these curves was within 8% of the measured data.

Figure 4.7: Pressure-bulge height curves at the approximate strain rate of 0.25s
-1

35
Figure 4.8 shows the pressure-bulge height curves at various
temperatures for the approximate strain rate of 0.025 s
-1
. The effect of possible
thermal/work softening is clearly seen above 165 C. The difference between
Figure 4.7 and Figure 4.8 is that, samples in Figure 4.8 show higher bulge
heights, lower pressure values and extended pressure drops (possible softening
behavior).

Figure 4.8: Pressure-bulge height curves at the approximate strain rate of 0.025
s
-1

4.9 Analysis of the dome geometry
The bulge shape obtained in equi-biaxial bulging is actually not a perfect
sphere, although to simplify the analysis, most researchers assume that the
experimental bulge is a perfect sphere. In order to investigate this in elevated

36
temperature forming of Mg AZ31-O alloys, several pressurized sheets (without
fracture) were selected for determining the bulge geometry.
The bulge geometries were measured using a Coordinate Measurement Machine
with a 0.98 mm probe tip. Measurements were made along the rolling direction
(RD) and transverse direction (TD) by obtaining 21 points (10 points on each
side of the apex). It is entirely possible that there was some small amount of
springback, especially at the lower range of the forming temperatures that were
investigated. However, in this study, it was not possible to determine the
magnitude of springback which can be expected to be small at elevated
temperature forming.
The radius of the bulge was calculated by using 5 points (near the apex) and 21
points for the entire bulge. 5 points were used because, it is believed that a 3
point fit (at apex) might lead to erroneous radius values due to possible local
imperfections at the apex of the bulge while 5 points around the apex will
provide more realistic data. Least Squares Circle Fit (LSCF) was conducted to
fit the 5 and 21 points to the best circle by minimizing the residuals through an
optimization routine. Figure 4.9 shows the schematic of the measurement
points.

37

Figure 4.9: Measurement points on the bulged sheet
Figure 4.10 shows the obtained radius values along RD using a) 5 point
LSCF b) 21 point LSCF and c) analytical solution with Eq.9 for samples having
bulge heights of 14.7mm, 21.7mm and 33.9 mm, respectively. It is seen that
radius values calculated analytically are closer to the radius values obtained
using 5 points at the apex, rather than those obtained using 21 points.
Differences between predictions made with Eq.9 and the radii obtained using 5
point LSCF for bulge heights of 14.7 mm, 21.7 mm and 33.9 mm, are 1.5 %, 3 %
and 6 %, respectively. However, as shown in Figure 4.11, by selecting a 5 point
fit (points 2,3,4,5 & 6) we know that the fitted circle is not going to pass from
points (1 and 7) at the edges of the bulge. Therefore, a comparison between the

38
radius obtained from the 5 point (points 2,3,4,5 & 6) and a 3 point (points 1, 4 &
7) fit was done to determine the amount of change in the bulge radius.
For this purpose, two samples with different bulge heights (21.7 mm and 33.9
mm) were selected. Calculated R
1
& R
2
values (Figure 4.11) are given in Table
4.5.

Figure 4.10: Comparison of calculated and measured bulge radius values
In both cases, it is clear that R
1
is larger than R
2
. However, the difference is
limited to a maximum of 2.5 % for the sample with the higher bulge height. As
a conclusion, a 5-point Least Squares Circle Fit (LSCF) is assumed to be an
acceptable approximation for obtaining the radius at the apex of the bulge.

39
R
1
R
2
5 pt. LSCF
3 pt.
1
4
5
6
7
3
2

Figure 4.11: Radius difference using 5 and 3 points

Bulge
Height
(mm)
5-point
LSCF
RD (mm)
-R
1
-
3-point
LSCF
RD (mm)
- R
2
-

Difference
[%]
21.7 90.7 89.9 <1
33.9 65.3 63.6 2.5
Table 4.5: Comparison of radius values (R
1
& R
2
) obtained by using 5 and 3
points
4.9.1 Deviation from the original bulge shape
Figure 4.12, Figure 4.13, and Figure 4.14 show the difference (residuals)
between the original measured points and the fitted circles using 5 and 21
points for bulge heights of 14.7 mm, 21.7 mm and 33.9 mm, respectively. These
bulge heights were selected to demonstrate the evolution of change from the
lowest to the highest bulge height.

40

Figure 4.12: Residual plots for sample at h
d
=14.7 mm
In Figure 4.12 and Figure 4.13 symmetric distribution of residuals on
each side of the apex for h=14.7 and 21.7 mm are seen. While residuals around
the apex are very close to zero, they are larger near the edges of the bulges. The
residuals in Figure 4.13 are slightly higher compared to those in Figure 4.12,
but the maximum value is under 0.5 mm. For a larger bulge height, (Figure
4.14), the symmetric distribution is not seen and the value of residuals around
the edges increase up to 4 mm. It is also seen that the residuals are higher
around the apex when 21 points are used.

41

Figure 4.13: Residual plots for sample at h
d
=21.7 mm

Figure 4.14: Residual plots for samples at h
d
=33.9 mm
The residuals seen in Figure 4.12, Figure 4.13 and Figure 4.14 and other results
of measurements indicate that the bulge shape does not deviate from the

42
assumed sphere shape until the bulge height to diameter ratio h
d
/d
c
is
approximately larger than 0.2. Below this ratio, the maximum residuals
remained under 0.5 mm. It is clearly seen from Figure 4.14 that the residuals
are not symmetric, which possibly indicates that material becomes unstable at
that high deformation.
4.10 Thickness distributions along the dome curvature
Thickness measurements were conducted for samples that were
pressurized to approximately same bulge heights at different approximate
strain rates and temperatures. Measurements and predictions, made using
Eq.10 were compared.
Figure 4.15 shows the thickness distributions in the rolling direction along the
curvilinear lengths of the bulges formed to 21 mm height. As expected, due to
volume constancy, at approximately the same bulge height (~21 mm) there is
not a significant thickness difference for samples that were bulged at the same
strain rate (0.25 s
-1
) but at different temperature.
Table 4.6 shows the measured and the calculated thickness values at the apex
of the bulges shown in Figure 4.15. The error percentage using Eq.10 is up to
4% and 8 % for bulge heights around 21 mm and 33.9 mm, respectively.

43

Figure 4.15: Thickness distribution along the RD (bulge height (h)=~21mm
and h33 mm)
It is clear from Figure 4.15 that for samples formed at ~0.025 s
-1
,

the
lowest measured thickness is not at the apex. For samples formed at ~0.25 s
-1
,
at least three of the same thickness values are seen between curvilinear lengths
of 40 mm and 80 mm. In most samples at high temperature and low strain rate,
fracture was as big as the tip of a needle. Sheets formed to higher bulge heights
also clearly show the unusual thickness distribution (Figure 4.15). It was first
thought that the unexpected thickening at the apex may be due to the weight or
the temperature of the bulge height sensor. (The contact between the sensor
and the apex is a point contact) However, the weight of the bulge height sensor
is extremely low, and its tip is also submerged under the hot liquid. Therefore
it is not likely that the sensor could have caused the unexpected thickness
distribution.

44

Bulge
Height
(mm)

Temp.
(
o
C)
Strain
Rate
(s
-1
)
t
d
(measured
at apex)
(mm)
t
d

(Eq.4)
(mm)
21.7 180 0.25 0.96 0.92
21.6 207 0.25 0.95 0.92
21.2 194 0.025 0.92 0.93
33.9 214 0.025 0.73 0.67
Table 4.6: Measured and calculated thickness values at the apex

At room temperature hydraulic bulging of sheets, the minimum
thickness values are always observed at the apex. Furthermore, the fracture
takes place at the apex of the bulge. The thickness distribution plots, seen in
Figure 4.15, are clearly different from the thickness distributions observed at
room temperature hydraulic bulging of sheets. Most probably, this observation
is due to the strain rate effect during the deformation. During hydraulic bulging
at elevated temperatures, even though obtaining an approximate true constant
strain rate is possible at the apex by controlling the flow rate, there is a strain
rate gradient along the contour of the bulged sheet. In Figure 4.16, a schematic
representation of the strain rate distribution is shown. Consequently, at
isothermal conditions, there is also a flow stress gradient in the bulged sheet.
As a result, the sheet might fail at a location where a combination of low flow
stress and high strain (low thickness) is present.

45
Pneumatic bulging of Mg AZ31-O alloys conducted at elevated temperatures
show also that the fracture does not take place at the apex. [Siegert, 2001]
e e

&
,
d d

&
,
c c

&
,
b b

&
,
a a

&
,
e d c b a
e d c b a


> > > >
> > > >
& & & & &

Figure 4.16: Schematic representation of the strain rate and flow stress
gradients in the bulged sample
Figure 4.17 and Figure 4.18 show the calculated true stress (Eq.6) and
true strain (Eq.7) curves at approximate strain rate values of 0.25 s
-1
and 0.025s
-
1
at various temperatures. At 0.25s
-1
, a slight drop in stress takes place at 221 C.
However, at 0.025 s
-1
, the drop in stress is clearer at temperatures of 190 C, 214
C and 225 C. Maximum strain values of 0.45 and 0.7 are reached at strain rates
of 0.25 s
-1
and 0.025 s
-1
respectively, while the achievable strains using the
tensile test are approximately 0.3 at the same strain rate values. [Groche et al,
2002]

46
0
50
100
150
200
250
300
350
0 0.1 0.2 0.3 0.4 0.5
221
205
180
168
145
T:
o
C

T
r
u
e

S
t
r
e
s
s

(
M
P
a
)


True Strain
s / 1 25 . 0



Figure 4.17: True stress and true strain curves at 0.25 s
-1

As discussed earlier, there is an error in radius and thickness predictions when
membrane theory is used. This error was up to 6 % in radius and 8% in
thickness predictions for the highest obtained bulge height. For samples with
h
d
/d
c
ratios lower than 0.2, radius and thickness predictions were acceptable
since the shape of the bulge is nearly spherical, as assumed in analytical
calculations. In order to demonstrate the introduced error, the sample in Figure
4.18 that was formed at 225 C was selected. The same curve was plotted with
possible errors after (h
d
/d
c
)>0.2, which corresponds to a strain of ~0.4.
Therefore, the error plots are shown after 0.4 strain (Figure 4.19).

47
0
50
100
150
200
250
300
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
225
210
190
165
25
T:
o
C

T
r
u
e

S
t
r
e
s
s

(
M
P
a
)


True Strain
s / 1 025 . 0



Figure 4.18: True stress and true strain curves at 0.025 s
-1

In Figure 4.18, the flow stress curve of the sample formed at 225 C and
at the strain rate of 0.025 s
-1
show certain amount of crossing of the data. In
other words, the flow stress value for 225 C at a true strain rate of 0.025 s
-1
is
larger than the corresponding value at 210 C. This observation may be due to
the effect of recrystallization and microstructure changes or experimental
errors. These observations must be investigated in the future and the flow
stress data obtained in this study must be compared with data from other
investigations in order to clarify this seemingly perplexing observation. A
similar behavior is also seen in the elevated temperature tests conducted by
[Takuda et.al., 2005]. In that study, the flow stress curve at the highest
temperature 300 C, and at the lowest forming speed (0.2 mm/s), show the same
crossing behavior at lower strains.

48
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
100
105
110
115
120
125
130
135
140
145
150
155
160
T
r
u
e

S
t
r
e
s
s

(
M
P
a
)
True Strain
225 C/0.025 s
-1

Figure 4.19: True stress true strain curve with possible errors at 225 C (0.025
s
-1
)
4.11 Conclusions
1) Submerged hydraulic tool concept for conducting bulge test offers
homogeneous temperatures in the deformed sample. However,
repeatability of the experimental conditions is difficult to achieve.
Temperature control was one of the most difficult issues. It was nearly
impossible to have exact temperatures on the sheet when experiments
had to be repeated at the same temperature.
2) Variation of properties within the same batch of material is a serious
issue. Some sheets that were from the same coil unexpectedly showed
low formability. This might be due to varying processing conditions

49
during rolling of the sheet. Therefore, it is expected that data, obtained
for the same material but by different investigation (using different
material lots) may have variations.
3) It is observed that at elevated temperatures, depending on the strain rate
work softening behavior might take place along with work hardening.
4) By using elevated temperature hydraulic bulge system, strains up to 0.7
were obtained at a strain rate of 0.025 s
-1
at 225 C.
5) It was observed that almost up to 25 mm bulge height, (for the die
diameter of 115 mm used in this study) dome radius and thickness
predictions at the apex were acceptable. Above a certain bulge height,
(~25mm) it is clear that, the dome shape is not spherical and starts to
deviate from a sphere. The effect of strain rate is believed to have effect
on the dome geometry.
6) Minimum thickness values were found to be closer to the apex but not at
the apex. This is believed to be due to the strain rate and flow stress
gradient along the dome.


50

CHAPTER 5


DEEP DRAWING PROCESS AT ROOM TEMPERATURE

Before investigating the warm deep drawing of light alloys (Al, Mg) it is
appropriate to examine first the details of this process at room temperature.
Deep drawing is used for the mass production of parts used in automotive,
aerospace and appliance industries. In this process, a sheet metal is drawn into
a die cavity by a moving punch. Figure 5.1 shows a deep drawn sheet that is
divided into six regions. (A-B/ B-C/ C-D/ D-E/ E-F and F-G). At each region, the
sheet metal experiences different state of stress, therefore the sheet thickness is
not uniform. Figure 5.2 shows the top view of a drawn sheet, where A is the
edge of the deformed sheet and C is the point where sheet is entering the die
cavity. At point A, there is a free surface and the radial stress
r
=0, therefore
the dominant stress state is compressive, which causes the thickening of the
sheet. At point B, the stress state is both compressive and tensile and the sheet
thickness will remain unchanged. At point C, the radial stress will be very high
and the sheet will thin. Figure 5.3 shows the stresses on an element at a radius
r.

51

Figure 5.1: Axisymmetric
Deep Drawing


Figure 5.2: Top view of drawn
sheet


Figure 5.3: Stress on an
element at radius r [Marciniak, et al,
2002]


Figure 5.4: Thickening of the sheet
in the flange area

The metal at point A (Figure 5.4) is moving towards the points A, B and
C where the perimeter and surface area over the die are smaller. Since the
metal is incompressible, the material in region A thickens, as seen in Figure

52
5.4. In most analytical calculations the thickness is assumed to be uniform in
order to simplify the calculations [Lange, K., 1985]. However, in reality the
thickening of the sheet creates elastic stresses and deflections on the blank
holder and the die.
5.1 Drawability Criteria
In literature, deep drawability of an alloy is indicated by the Limiting
Draw Ratio (LDR). LDR is defined by the ratio of initial sheet diameter (D) to
the punch diameter (D
p
) (Figure 5.5). Maximum LDR refers to the cup that can
be drawn without fracture or other defects. The higher is the maximum LDR
that can be deep drawn for a material, the higher is the drawability of that
material. LDR is a function of many parameters such as the tool geometry,
lubricant, blank holder force which means it can change when any of these is
changed. Therefore, it is a criterion that heavily depends on the given
experimental setting. When LDR is mentioned, it is possible that the drawn cup
may or may not have a flange left. A deep drawn cup might be drawn with or
without a flange. If a cup is drawn with flange, using LDR as drawability
criterion does not reflect accurate forming severity. Another criterion and a
better parameter to indicate the deep drawability is the ratio of the height of
drawn cup (without defects) to the punch diameter is determined and called as
the Cup Height to Punch Diameter ratio (HDR) (h/D
p
). (Figure 5.5) It is believed

53
that this criterion represents the drawability better if the drawn cup has flange
left.

Figure 5.5: Definition of cup height (h) for a cup with a flange and without a
flange
Some drawn cups can experience earing due to the anisotropy that exists
in the incoming sheet material. For better accuracy, while determining the HDR
for cups with ears, cup height is decided to be the h
min
as shown in Figure 5.6.
Therefore, both LDR and HDR are used based on the definitions provided here.

Figure 5.6: Deep drawn cup with ears


54
5.2 Critical Stroke in deep drawing
Based on the conducted FEA and observations in the mechanics of
drawing, it is seen that the bending of the sheet takes place around the punch
and the die corner during the initial part of the drawing process. The sheet
thins during the initial bending. Therefore we can say that the cause of
thinning is due to a) continuous bending of the sheet b) stretching of the sheet
around punch corner. Bending of the sheet continues until the punch reaches a
critical stroke (S
cr
) which is defined as:
S
cr
=r
D
+r
P
+t
where, r
D
: Die corner radius, r
p
: Punch corner radius, t: Sheet thickness.
When the punch is exactly at the end of the critical stroke, centers of the die
ring radius and the punch radius are aligned (Figure 5.7).
When the punch travels within the critical stroke, the sheet is tangent
both to the blank holder and the punch corner radius. Both tangency point
positions change until the punch travels the whole critical stroke. After the
punch completes traveling the critical stroke, the sheet is mostly in contact
with the punch wall and tangent to the blank holder. The critical stroke
concept will be used later on with the hydroforming and the warm deep
drawing processes.

55

Figure 5.7: Schematic view of the critical stroke (S
cr
) before (left) and after
(right)
5.3 Deep drawing of aluminum and magnesium at room temperature
Drawability limits of Al 5754-O and Mg AZ31-O at room temperature is
determined experimentally. Figure 5.8 and Figure 5.9 show that LDR of 2.1
(HDR:0.84) is obtained for the Al 5754-O alloy at room temperature. It is clearly
seen that Mg AZ31-O alloy experienced early fracture and was not drawn
successfully. A nylon film was used as a lubricant.


Figure 5.8: RT drawing of Al 5754-O Figure 5.9: RT drawing of Mg AZ31-O

56
CHAPTER 6


HYDROFORMING OF SHEET AT ROOM TEMPERATURE

A variety of innovative metal forming processes are being developed for
forming lightweight materials. Among these processes, hydroforming of sheet is
quite significant. Sheet hydroforming can be done with a stationary die (SHF-
D) or with a moving punch (SHF-P). SHF-D is pressurization of a sheet to a die
that has certain geometry. Therefore, it is a combination of less drawing and
more of a stretching process. Research on SHF-D has shown that excessive
thinning is seen in the deformed sheets, especially at the die corner areas, and
therefore it is not very beneficial. Summary of some results for SHF-D process
is given in Appendix A. However, sheet hydroforming with punch (SHF-P)
provides better thickness distribution and offers great potential for low and
medium volume production. In this process, instead of using two hard dies as
done in conventional drawing, a single die or even no die is used. A viscous or
liquid medium applies pressure to the sheet metal that is deformed and pressed
against the punch that has the part geometry (Figure 6.1). Increased
formability, high surface finish and reduction in the number of tools are the
main advantages of SHF. However, compared to conventional deep drawing,
control of liquid medium pressure and the blank holder force (in time and in

57
space) become critical in order to prevent leaking and wrinkling. Also higher
machine capacity is required due to the forming that takes place against
pressure. A fundamental understanding of the mechanics of sheet
hydroforming process, including the effect of elastic deformation of tools that
affect the process robustness, can lead to wide application of this technology
for cost effective production of lightweight components. Appendix B shows
some results of elastic deflection analysis conducted using FEA in deep
drawing operation.
If the pressure is high, a reaction force (F
LIFT
) is generated that acts
against the BHF (F
BH
) (Figure 6.1). If F
LIFT
> F
BH
, then leakage takes place as it is
shown in (Figure 6.1b). If F
LIFT
< F
BH
then perfect sealing takes place, like it is
shown in (Figure 6.1a) By using SHF-P, higher LDRs can be obtained due to a)
eliminated frictional forces at the sheet-die corner interface b) reduced drawing
stresses at the cup wall due to pressure and increased frictional forces between
sheet and punch. The fluid pressure profile is an important factor and it
influences the sheet formability.

58

a) b)
Figure 6.1: SHF-P process a) without leakage; b) with leakage

Experimental results for SHF-P process of soft 70/30 brass round cups
with a LRD equal to 3.44 are presented by [Sebaie et al., 1973]. Yossifon and
Tirosh studied and developed an analytical model that provided a pressure
path relative to the punch stroke that took wrinkling and fracture into account
[Yossifon et.al., 1985a] and [Yossifon et.al., 1985b]. Their formulation is based
on the classical theory of plasticity (with simple Power-Law hardening
) (
n
K = and Mises-Hill normal anisotropic yielding) assuming plane strain
tensile failure. This study concluded that in hydroforming, wrinkling occurs at
the unsupported cup wall sheet close to the die corner [Yossifon et.al., 1984]
and [Yossifon et.al., 1988]. The authors defined a Maximum Drawing Ratio

59
(MDR) as the highest drawing ratio before wrinkling and fracture takes place.
They concluded that the MDR is increased as the anisotropy parameter (R) and
wall thickness (t) are increased for a given strain-hardening exponent (n). They
determined that MDR reaches its minimum between the strain hardening
exponent n=0.1 0.2, but increases moderately beyond this range [Yossifon,
1990]. Most of the research and analysis has been conducted for SHF-P using
interchangeable dies. SHF-P with uniform pressure on the flange is an
alternative to hydroforming with dies. In this process, the die is eliminated and
the sheet metal is placed in the pressure pot. A brief state of the art of
hydromechanical deep drawing has been summarized by [Thiruvarudchelvan
et.al. 2003]. The effect of anisotropy and pre-bulging on mild steels has also
been investigated by [Zhang et.al., 2003].
Main objective is to determine an approximate pressure profile, which provides
a defect-free (no wrinkling or early fracture) cup with a high LDR while
preventing leakage throughout the process. The pressure profile curve is a
function of both the sheet material and the tool dimensions (Figure 6.2).
The punch corner radius (r
p
), the die ring corner radius (r
D
) and the
clearance between punch and blank holder (c) affect the SHF-P process. A too
small value of the punch corner radius and the die ring corner radius could tear
the sheet. Bulging of the sheet between the punch and the blank holder could

60
happen if a too large clearance (c) between punch and blank holder exist.
These aspects are analyzed for a 90 mm round cup tool geometry provided by
Schnupp Hydraulik of Bogen, Germany.
Sheet
Die
Blank
holder
Punch
Rp
rp
rD
c

Figure 6.2: Important tool parameters that influence the SHF-P process
6.1. Mechanics of SHF-P process
In SHF-P, the cause of thinning is due to a) continuous bending of the
sheet b) stretching of the sheet around punch corner due to counter pressure.
When the punch travels within the critical stroke, the sheet is tangent both to
the blank holder and the punch corner radius. Both tangency point positions
change until the punch travels the whole critical stroke.
After the punch travels the critical stroke, the sheet is always in contact with
the punch wall and tangent to the blank holder. If the fluid pressure doesnt
change, the tangency point between sheet and blank holder remains the same.

61
Therefore, after the critical stroke, the sheet radius is constant if the fluid
pressure is constant. Higher pressure causes the tangent point (distance H in
Figure 5.7) to shift in the direction of +x. (Figure 5.7) When this point shifts in
the direction of +x, it causes more sheet area to be in contact with the blank
holder. In return, higher blank holder force is needed. Therefore, the higher the
fluid pressure, the smaller the sheet radius and the larger the BHF required to
avoid leakage.
It is seen from the previous FE Analyses that, it is very important to
determine the pressure curve. As it is seen in Figure 6.3, if the pressure is high
enough at the beginning of the process than the sheet loses contact with the die
corner.

Figure 6.3: During a hydromechanical deep drawing process the sheet loses
contact with the die

62
Max thinning happens at an early stage of the drawing at the punch
corner and stays almost constant until the total punch stroke is completed
(Figure 6.4). In order to demonstrate this phenomenon, the mechanical
deformation of the sheet has been observed from FEA. It is seen from Figure 6.4
that, at punch strokes of 20 mm and 70 mm, for the same given conditions
(pressure, BHF, etc..) maximum thinning percentages are approximately 15 %.
This shows that once the max thinning percentage is reached at initial stroke
levels, it remains the same the rest of the process.
-20%
-15%
-10%
-5%
0%
5%
10%
15%
20%
0 15 30 45 60 75 90 105 120 135 150
Curvilinear Ditance [mm]
T
h
i
n
n
i
n
g

p
e
r
c
e
n
t
a
Thinning distribution (Punch Stroke = 20 mm)
Thinning distribution (Punch Stroke = 70 mm)


Figure 6.4: Maximum thinning around the cup bottom radius at two different
values of the punch stroke
A
C
C
A
B

63

6.2. Prediction of the initial pressure value
Since it is beneficial that the sheet has to be lifted from the die corner to
eliminate the frictional forces, the issue is to find the first approximate pressure
value that we need to have in the pressure pot that will lift the sheet from the
die corner. Therefore, in order to determine this pressure value, slab analysis
was conducted to find an approximate relationship between the pressure, sheet
corner radius, thickness of the sheet, circumferential and radial stress
components.
BLANK HOLDER
DIE
Rp
FBH
PUNCH
Y
X
rs
rd
R


Figure 6.5: Position of the slab in the sheet radius zone (a) and equilibrium of
the forces in the slab (b)

Equation (1) shows the obtained relationship from the global equilibrium
of the slab. This equation provides an initial guess for the pressure value to
(b)
(a)

64
start with if the circumferential and the radial stresses are assumed to be the
yield stress of the material. The calculated pressure later was slightly altered
using FEA to optimize the movement of the sheet around the die corner.
|
|
.
|

\
|
+ =
R r
t p
s
r

(1)
where:
p: pressure in the pressure pot;

r
: radial stress;

: circumferential stress;
r
s
: instant sheet radius;
R: distance from the centerline of the punch to the slab.
The pressure inside the die pot (p) is a function of the yield stress (
y
) and the
sheet geometry during the process (instant radius of the sheet (r
s
) and distance
from the center of the punch to the slab (R)) (Figure 6.5). By using this
equation, the starting pressure that needs to be in the pressure pot
approximately calculated to be around 100 bars for low carbon steel. Low
carbon steel was selected due to the previously availability of experimental
data to validate the FE model [Kaya et.al., 2004] [Contri et.al., 2004]. From
trials with FEA, it is observed that this pressure cannot remain constant within
the critical stroke and needs to increase as the process proceeds. A linear

65
increase until the critical stroke is generally found to be satisfactory. The need
for the increase in the pressure is due to the strain hardening in the sheet.
Effect of various pressure curves (P1, P2, P3 and P4) on thinning behavior was
investigated using FEA (Figure 6.6). Following conclusions can be drawn from
this investigation;
a) When P1 was used, 20% thinning was obtained at the end of the critical
stroke due to excessive stretching.
b) After observing the deformation in P1, it was decided to start with a
lower pressure value and increase gradually (P2). The thinning
percentage at the end of the critical stroke was around 17%.
c) In order to quantify the effect of P1 on thinning, a constant 100 bar
pressure was input. (P3) After the punch traveled half of the critical
stroke, it was observed that the sheet started rubbing the die corner.
Therefore, the pressure curve needed to be updated to eliminate this
problem
d) Finally, when P4 was used, the sheet was never in contact with the die
corner and it was kept at a minimum distance from the die corner in
order to eliminate more stretching. By coupling the FEA with the slab
analysis P4 was later modified to increase linearly (the initial flat part of

66
the curve until 6 mm stroke is eliminated based on observation in FEA)
within the critical stroke while lifting the sheet off the die corner.
0
50
100
150
200
250
300
350
400
450
0 3 6 9 12 15 18 21 24 27 30
Punch Stroke [mm]
P
r
e
s
s
u
r
e

[
b
a
r
]
P1 P2 P3 P4

Figure 6.6: Pressure-stroke curves used in the critical stroke

It is of interest to know whether the material is stretching during the
critical stroke, since it is directly related to thinning. Therefore, the amount of
stretching in the sheet radius zone has been evaluated (Figure 6.7). As Table
6.1 shows, the stretching of material increases with the punch stroke. Material
stretching at the end of the critical stroke is around 15%.

P2
P1
P4
P3

67
Punch
Blank
holder
Die
Sheet
A B

Sheet
Die
Blank
holder
Punch
A
B

Figure 6.7: Initial (a) and final (b) length of stretch in the sheet used to evaluate
the material stretching in the sheet radius zone

Punch Stroke
[mm]
Curvilinear length
[mm]
Stretching
percentage [%]
0 25.86 -
2.58 26.07 0.81
3.76 26.23 1.43
4.95 26.39 2.05
7.35 26.76 3.48
8.55 27.04 4.56
9.75 27.38 5.88
10.95 27.69 7.08
12.15 28.11 8.70
13.35 28.51 10.25
14.55 29.54 14.23
15.75 30.03 16.13
18.15 30.47 17.83
Table 6.1: Curvilinear length and percentage stretch of the sheet in the sheet
radius zone according to Figure 6.7 (pressure curve: P4)
6.3. SHF-P of a 90 mm round cup
Experienced obtained using steel is applied to investigate the attainable
LDR for Al 5754-O. In this section, results of computer simulations are
(a) (b)

68
presented for Al 5754-O. For 90 mm diameter round cup with flat die and
blank holder surfaces was analyzed. Material properties are taken from
[Boogard 2001] and friction coefficients are taken based on previous modeling
and experimental experience [Table 6.2] [Contri, et al., 2004]. Friction
coefficients were validated with available experimental data from Schnupp
Hydraulik. In order to reduce computation time, quarter of each die set was
modeled.
Friction coefficients, m
Sheet/Punch (
sp
) 0.12
Sheet/Die (
sd
) 0.08
Sheet/Blank holder
(
sb
)
0.08
Table 6.2: Friction coefficients used in the FEA
Figure 6.8 shows the pressure curve. Initial pressure value of 60 bars is
determined from the slab analysis. A linear increase in the pressure curve is
found to be acceptable until 300 bars, since above this pressure value the sheet
is bent severely and excessive thinning was observed. Results showed that an
LDR of 2.4 is obtained with a maximum thinning percentage of 12 % for the Al
5754-O alloy (Figure 6.9).
It is already known that bendability of Mg at room temperature is limited.
Therefore, modeling of Mg with SHF-P process is not conducted.

69

Figure 6.8: Fluid pressure determined based on the critical stroke
-40%
-30%
-20%
-10%
0%
10%
20%
0 20 40 60 80 100 120 140 160
Curvilinear Distance [mm]
T
h
i
n
n
i
n
g

p
e
r
c
e
n
t
a

Figure 6.9: Thinning distribution along the 45 cup wall
A
C
D
E
F
B

70
CHAPTER 7


DEEP DRAWING AT ELEVATED TEMPERATURE

7.1 Introduction
The use of conventional forming technologies for Al and Mg alloy sheet
is restricted because of the low formability of these alloys at room temperature.
Increasing demand in weight reduction requires the development of innovative
forming techniques for forming difficult-to-form alloys. Studies in the past have
shown that the formability of Al & Mg alloys increases with temperature.
The rather complex mechanical deformation that takes place in deep
drawing process becomes more complicated when the effect of temperature and
strain rate is introduced. In elevated temperature deep drawing, the die and the
blank holder are generally heated, however the punch is cooled. When the
sheet first touches the cooled punch, its temperature decreases and therefore
part of the sheet in contact with the punch will be cooler than the rest of the
sheet. This means that the sheet closer to the punch corner can withstand more
stress and it will stretch less. Since the deformation is continuous, the portion
of the drawn cup in contact with the punch continues to cool. However, the
sheet that is under the heated dies has a higher temperature which helps the
material flow easier.

71
In order to develop a rational R&D strategy for warm forming of Al and
Mg alloys and to establish research priorities, it is important to consider the
warm sheet forming process as a system. The incoming material shape and
properties, the forming temperature, the interface condition (friction and heat
transfer), the tool temperature, forming speed (or strain rate), and the forming
equipment influence the final product shape and properties as well as the
economics of the process. A fundamental understanding of the relationship
between the input and output variables of the system is essential for developing
a robust, productive and economical manufacturing process [Altan et al. 1983].
Thus, the development of warm forming process for sheet metal requires
critical consideration of: (a) material flow behavior of Mg and Al alloy sheet at
elevated temperature (b) lubrication system at the tool/blank interface (c) tool
design with temperature control, and (d) warm forming process design,
experimentation and numerical modeling using finite element method (FEM).
A warm deep drawing tool was designed and Figure 7.1 shows the schematic of
the designed tool in open and closed condition.

72

Figure 7.1: Open (left) and closed (right) condition of the tool

In this study, the focus was on a) the effect of interface pressure on the
dwell time and on the surface roughness, b) selection of the best performing
lubricant among the available lubricants manufactured for elevated
temperature forming conditions c) the effect of forming temperature and
velocity (constant and variable) on the deformation mechanics d) process
optimization of the warm forming process.

73
7.2 Experimental setup
7.2.1 Servo press and its kinematics
The most obvious benefit of the servo press is the flexibility it offers.
This flexibility comes from the ability to program the speed and motion of the
slide in an infinite number of ways, with a constant load available throughout
the stroke at any speed. In warm deep drawing process, the servo press allows
the workpiece to be heated in the tool, through the ability to dwell. Therefore,
it eliminates the extra equipment necessary for heating and transferring the
sheet to the die.
The schematic drive system of the 110 ton AIDA Servo press, used in
this study, was shown in Figure 7.2. The key feature of this design is the
companys proprietary high torque, low RPM servomotor. This motor is
mounted directly to the driveshaft, eliminating the need for linkage systems. In
this design, the need for mechanism to multiply the torque and therefore the
flywheel, clutch, and drive motor are all eliminated. The drive shaft drives the
main gear, which in turn, is connected to an eccentric drive mechanism.

74

Figure 7.2: AIDA servo-mechanical press drive mechanism
The eccentric drive mechanism facilitates the movement of the top ram.
The capacitor stores power in the non-working portion of the stroke, helping to
reduce the power consumption to levels comparable to standard mechanical
presses. The press is equipped with an air operated blank holding system to
provide the blank holder force (BHF).

1-2 Fast approach
2-3 Slower approach reduces impact and
vibrations. Both tools are in contact at
3.
3-4 Dwell (heating of the blank)
4-5 Slower punch velocity for forming
sharp corner radii.
5-6 Higher velocity for faster forming
6-7 Slower exit from the tool
7-8 Faster return to TDC

Figure 7.3: Ram motion of the servo press (TDC: top dead center, BDC: Bottom
dead center)

75
7.2.2 Design of the tooling
Figure 7.4 gives a schematic view of the warm forming tooling. The
operation of the designed tool set is illustrated in Figure 7.5. In the proposed
process sequence, initially the blank is placed on the bottom die/blank holder.
The top ram moves down till it touches the blank and dwells. During this dwell
time, the blank is heated to the required temperature by the heated die and
blank holder. After the dwelling period, the top ram moves further down
against the stationary punch and the sheet is formed. The servomotor driven
press allows infinite freedom in programming the velocity characteristics i.e.
initially move rapidly to touch the blank, dwell for the desired time and finally
form the sheet at desired forming velocity. A stroke vs. time profile that can be
used for the warm deep drawing process is shown in Figure 7.3.
The die and the blank holder are heated with cartridge heaters (up to 310
o
C) while the punch is cooled approximately to room temperature (up to 65
o
C)
with water circulation. The die and blank holder temperatures are measured
and controlled. Glass fiber insulators are located on top and bottom, between
the tool plates and the die ring holder. The tool set is equipped with a load cell
and a displacement transducer to measure the punch force and ram
displacement during the process. Figure 7.6 shows the tool in open condition.
Figure 7.7 and Figure 7.8 show the tool in the press and the whole view of the

76
servo press with the temperature controller and the punch cooling unit,
respectively.
R
6
R
4
22,3
PUNCH
DIE
20

Figure 7.4: Schematic view and the dimensions of the tool (dimensions are in
mm)

Figure 7.5: Warm forming with in-die dwelling process sequence


77

Figure 7.6: Open warm forming tooling (top die set on the left, bottom die set
on the right) [Kaya, et.al., 2007]


Figure 7.7: Assembled tooling on the Aida servo press

Top die
Bottom die
Bottom springs

78

Figure 7.8: 110 ton Aida Servo Press

7.3 Issues at the interface in forming at elevated temperature
It is well known that heat transfer at the tool-blank interface is affected
by the surface roughness of the blank/tool, lubricant and the amount of
interface pressure. Lubrication is more important in warm forming than in cold
forming of Al and Mg alloys because the likelihood of galling increases with the
increase in temperature [Avedesian et al. 1999]. Lubricant performance
depends primarily on the forming temperature. For temperatures up to 120
o
C
(250 F), oil, grease, tallow, soap and wax are generally used in warm forming of
metals. A soap solution is acceptable for temperatures of up to 230
o
C. When
Programming
interface
Temperature
controller
Punch cooling
water tank

79
the forming temperature exceeds 230
o
C the choice of lubrication is restricted to
molybdenum disulphide, colloidal solution of graphite, and Teflon. These
lubricants should be cleaned from Mg parts as soon as possible to avoid
corrosion [Avedesian et al. 1999], [Dow, 1962]. Although a limited number of
lubricants exist that can be used for warm forming of metals (up to 300
o
C), no
extensive lubricant evaluation has been done for warm forming processes.
Therefore, investigation on the performance of these lubricants in Al and Mg
alloy sheet forming is of utmost importance. To evaluate the performance of
lubricants closer to production conditions, the deep drawing test at room
temperature was used by [Meiler, M., et al, 2005], [Meiler, M., et al, 2004],
[Pfestorf, M., 2002], [Wagner, S., et al, 2002]. [Semiatin et.al.,1987] have
conducted experiments to measure the change in temperature under different
pressures in forging processes. However, the effect of interface pressure on heat
transfer in rolled sheet metal has not been studied.
7.4 Heating the blank
In hot forming, heat transfer is affected by the interface pressure,
[Semiatin et.al.,1987]. In order to determine the necessary dwell time and the
effect of interface pressure on temperature increase, experiments were
conducted by using a designed aluminum fixture (Figure 7.9). Four holes (each
2 mm in diameter) vertical to the thickness direction were drilled towards the

80
center of the fixture. Depths of these holes were designed to be 20 mm, 25 mm,
30 mm and 40 mm, with respect to the tool dimensions. As the top view of the
fixture is seen in Figure 7.9, 40 mm hole touches the center of the fixture.
In this type of experiment, one of the difficulties is to have the sheet
touch the heated upper and the lower dies at the same time. In order to
minimize this, the servo press was programmed to operate at the highest
velocity between points 1 and 3. (Figure 7.3) Therefore, once sheet/fixture was
located on the lower die, approximately after one second the upper and lower
tools were both in contact with the sheet/fixture before the dwell stage started.

Figure 7.9: Top view of the fixture
used to determine the dwell time
necessary to heat the blank
(thickness 3mm)

q q
q q
P
P
DIE
BLANK
HOLDER
Thermocouple
PUNCH
Interface A
Interface B
Interface C

Figure 7.10: Schematic view of the
experimental setup with the affected
interfaces

Experiments were conducted by inserting an E-type thermocouple into the 40
mm hole, in order to make temperature measurements at the center of the

81
fixture, while the fixture was under different interface pressures. Figure 7.10
shows the schematic view of the attached thermocouple and the affected
interfaces. Figure 7.11 shows the experimentally measured temperature-time
curves at various interface pressures (Blank Holder Pressure, BHP) when the
tool temperature is at 300
o
C. BHP levels were selected in such a way that the
range of 1.5 MPa 4 MPa would fall into the range that was required in the
deep drawing experiments. As it is seen from Figure 7.11, when the fixture was
under 26 MPa of BHP, it reached to 250
o
C 7 seconds before than the test that
was conducted with 1.5 MPa of BHP.

Figure 7.11: Temperature-time curves obtained with the test fixture for
different interface (blank holder) pressures (tool temperature=300
o
C)
During these measurements the variation in the temperatures of the
punch and the knock-out pin (used to take the drawn cup out of the tool) were

82
also recorded. Figure 7.12 shows the temperature range plots at 250
o
C, 275
o
C
and 300
o
C.

Figure 7.12: Punch and knockout pin temperatures at tool temperatures of 250
o
C, 275
o
C and 300
o
C
7.4.1 Effect of the interface pressure on the hardness and the surface
roughness
Surface roughness of the material affects the heat transfer, which would
at the same time affect the dwell time necessary for heating the sheet in warm
deep drawing. Along with the surface roughness, change in the hardness of the
material after operating at elevated temperatures, is also of interest for further
practical applications. Therefore, in this experimental analysis, hardness and
surface roughness values were measured for each sample (Al 5052-H32, Mg
AZ31-O (Supplier A), Mg AZ31-O (Supplier B)) before and after the
experiments at two locations. Al 5754-O is not used because it is already at the
O heat treatment condition. (One point in the center and the other point ~10

83
mm off the edge of the sheet). The reason for conducting the measurements at
two points is because of the slight variation of the sheet temperature between
the center and the edge of the sheet. The center of the sheet is relatively cooler
compared to the side of the sheet, because it is open to air. Tool temperature
was set to 300
o
C and the dwell time was selected to be 90 seconds to make
sure that the whole sheet reaches approximately to the same temperature.
Diameter of the samples was 100 mm.
7.4.1.1 Hardness Measurements
Hardness measurements in Brinell scale (500 kg weight) were made using a
1/16 ball. Table 7.1, Table 7.2 and Table 7.3 show the hardness values of the
sheets before and after the experiments. Results indicate that there is a decrease
in the hardness values of Al 5052-H32 sheet both at the center and at the side.
For Mg materials, there is a slight increase in the hardness. The reason for this
increase might be related to the decrease in the surface roughness. Local plastic
deformations on the surfaces might have hardened the surface of these
materials.
As it is seen from the tables, hardness measurements for the highest BHP
values could not be made at the center. Because, during the experiments, with
increasing BHP, axisymmetric domes were observed at the centers of the sheets
(open to air from both sides). Due to the formation of these domes, flatness is

84
lost at the center of the sheets; therefore hardness measurements were not
possible.

Al 5052-H32
BHP (psi) Before Exp. After Exp
Center Side Center Side
90 58 58 54 54
270 58 58 56 56
750 59 59 - 55
1740 60 60 - 55
Table 7.1: Hardness (Brinell) measurements of Al 5052-H32




Mg AZ31-O (Supplier A)
BHP (psi) Before After
Center Side Center Side
90 58 58.5 60 59.5
270 59 60 60 59.5
750 57.5 59 - 61
1740 59 59 - 61
Table 7.2: Hardness (Brinell) measurements of Mg AZ31-O (Supplier A)










85
Mg AZ31-O (Supplier B)
BHP (psi) Before After
Center Side Center Side
90 57 57 58 58
270 57 56.5 57.5 57.5
750 57 55.5 - 56
1740 56.5 56 - 56.5
Table 7.3: Hardness (Brinell) measurements of Mg AZ31-O (Supplier B)
7.4.1.2 Surface roughness measurements
Figure 7.13 and Figure 7.14 show the as received surface roughness conditions
of the sheets along the rolling and the transverse directions (RD and TD). It is
interesting to note that the surface roughness values for the Mg AZ31-O
(Supplier A) are the highest and they differ significantly from the Supplier B.

Figure 7.13: R
a
values for 5754-O and 5052-H32


86

Figure 7.14: R
a
values for Mg AZ31-O (Supplier A) and Mg AZ31-O (Supplier B)
7.4.1.3 Dome formation in the sheet under pressure
Figure 7.15 shows the dome heights at 90 psi, 270 psi, 750 psi and 1740
psi for tests conducted at 300
o
C, with a sheet diameter of 100 mm. Figure 7.16
shows the change in the diameter of the sheets. Diameters were measured
before and after the test. It is interesting to note that, as the dome height
increases, change in diameter (contraction towards center) also increases.


87

Figure 7.15: BHP dome height at T=300 C

Figure 7.16: Change in the sheet diameter with respect to the BHP (Sheet
Diameter: 100 mm)


88
7.5 Lubricant evaluation for warm deep drawing process
Selection of lubricants for elevated temperature deep drawing is one of
the most important issues. On the other hand, development of lubricants for
elevated temperatures is at its preliminary stage, therefore there are not many
lubricants available commercially. Four lubricants were evaluated for their
applicability in the warm deep drawing process. Table 7.4 gives the list of
lubricants and the experimental condition. Evaluation method is using the
existing deep drawing tool and keeping every process parameter constant and
changing the lubricant. Therefore, deep drawing tests were conducted under
the same temperature, draw depth, forming speed and using the same die
corner radius. Measurement criteria are; a) flange draw-in, b) thickness
distribution c) visual inspection of the blank surface.
Lube Description

Lube A
PTFE Film
600 % elongation
327
o
C melting temp.

Lube B
Forgeease 02
AIS Binder III
7.5% Boron Nitride

Lube C

Forgeease 02
AIS Binder III
Lube D Dry Lubricant
(info N/A)

Al 5754-O
Sheet Diameter: 112 mm
Punch velocity:5 mm/s
Dwell time is 30 sec.
T= 250
o
C
Environment temperature: 25
o
C
Relative Humidity: 55 %
BHF=4.4kN 8.8 kN (linear increase)
Lube C dries faster than Lube B (20 min.drying time)

Table 7.4: List of lubricants and experimental conditions


89
Figure 7.17 shows that Lube A (Vac-Pak HT-620, Richmond Aircraft Products)
provides a more uniform thickness distribution in the cup, compared to the
other lubricants. This is due to Lube As high stretchability. During drawing it
did not tear and therefore eliminated a metal to metal contact. Figure 7.18
shows the cup tested with Lube A. Visual analysis of the cup surfaces also
indicate that the scratches arise from drawing is not seen when Lube A is used.
(Figure 7.19 and Figure 7.20) After experimenting with Lube B (Fuchs) and
Lube C (Fuchs), dried (possibly burned) lubricant accumulation was seen on
the tools, which required some cleaning after consecutive tests. Fourth
lubricant (Lube D, Lubrizol), did not perform well and was eliminated at the
very first stage of tests. Lube A is used in the vacuum molding of composite
materials in aircraft applications where high temperatures are involved.
Therefore, a new use of a PTFE film (Lube A) was found as a lubricant at
elevated temperatures and successfully used first time. Therefore, in the rest of
the experimental study, Lube A was used as lubricant.
Figure 7.19 shows the pictures of the cups that were tested using the
lubricants. It was observed from the experiments that cups that were tested
with Lube B and Lube C needed to be cleaned right after the test (when they
were hot) in order to be able to take the residue from the lubricants. It was
difficult to take the lubricant residue off the cups once they cooled down.

90

Figure 7.17: Thickness distributions for different lubricants under same
process conditions [Kaya, et.al., 2007]


Figure 7.18: Cup with Lube A after drawing [Kaya, et.al., 2007]


91

Figure 7.19: Al cups formed with Lube C, Lube B and Lube A (from left to right)

Figure 7.20: Mg cups formed with Lube C, Lube B and Lube A (from left to right)

92
7.6 Preliminary Experiments
A set of preliminary screening experiments were conducted to
investigate the maximum attainable LDR for Al 5754-O, Al 5052-H32 and
MgAZ31-O using the lubricant A. As preliminary experiments, part of the
methodology shown in Figure 7.21 was followed to determine the maximum
LDR that can be reached for Al 5754-O at a particular die temperature. Similar
approach was used to find the maximum LDR that can be reached for
Aluminum 5052-H32 and Magnesium AZ31B-O. The objective is to find the
highest possible LDR by adjusting the process parameters and obtaining some
numerical values about the process.

Figure 7.21: Adopted methodology for preliminary experimentation

93
Table 7.5 gives the obtained LDR and HDR values As a result of the
experiments, LDR values of 2.9, 2.9 and 3.2 were obtained for Al5754-O,
Al5052-H32 and MgAZ31-O, respectively (Table 7.5). In order for the LDR to be
accurate, the cup has to be fully drawn. It is seen in Table 7.5 that all cups
from Al5754-O and Al5052-H32 are fully drawn. For Mg AZ31-O, LDR 2.6 is
fully drawn but LDR of 3.2 has some flange. In order to be more accurate we
also used the Cup Height/Punch Diameter Ratio (HDR), explained earlier
which represents the drawability better when there is a flange.
Al 5754-O Al 5052-H32 Mg AZ31-O



LDR

HDR

LDR

HDR

LDR

HDR
2.1 0.84 2.1 0.76 N/A N/A
2.6 1.33 2.6 1.34 2.6 1.21

Left
to
right
2.9 1.73

Left
to
right
2.9 1.68

Left
to
right
3.2 1.9
Table 7.5: Summary of the preliminary screening experiments for Al 5754-O,
Al 5052-H32 and MgAZ31-O


94
Through these preliminary experiments, the designed and assembled tool with
temperature controller, heating system, tool temperatures, thermocouples and
punch cooling system was also checked and verified for further experiments.
7.7 Process optimization / windows
Using the lubricant A and a dwell time of 90 sec. to heat the blanks, a
series of tests were conducted to determine the best blank temperature and
constant punch velocity for the Al 5754-O and the Mg AZ31-O (Supplier A)
alloy. Experiments were also conducted using SS304 since it is possible to
reduce weight using a thinner gauge material. Results of these experiments are
given in Appendix C. Figure 7.22 through Figure 7.24 show variations of LDR
with punch velocity at various blank temperatures, including when some of the
drawn samples were fractured for the Mg AZ31-O alloy sheet. Three to four
samples were formed under each condition. Figure 7.25 shows the process
window (temperatures and punch velocities for successful cup drawing) for Mg
AZ31-O alloy for the tool geometry, used in this study.
The numbers in boxes at each point in Figure 7.25 show the maximum
thinning (in percent) obtained in the drawn cups. Figure 7.26 shows the
corresponding data (formed and fractured cups) for the Al 5754-O alloy.

95

Figure 7.22: Variation of LDR with punch velocity (Mg AZ31-O, T=250
o
C)

Figure 7.23: Variation of LDR with punch velocity (Mg AZ31-O, T=275
o
C)


96

Figure 7.24: Variation of LDR with punch velocity (Mg AZ31-O, T=300
o
C)

Figure 7.25: Process window for the Mg AZ31-O alloy




97


Figure 7.26: Process window for the Al 5754-O alloy

Results show that punch velocities up to 300 mm/s are possible to obtain
for the Mg alloy sheet while maximum punch velocity remains at 35 mm/s for
the Al alloy. The effect of the blank size is obvious on the attainable punch
velocity.
7.7.1 Effect of constant forming velocity and temperature on deformation
Experiments are conducted using a 100 mm diameter sheet (draw ratio
of 2.5) for both Al 5754-O and Mg AZ31B-O. The objective is to understand the
effect of constant (5 mm/s, 15 mm/s and 50 mm/s) and variable punch velocity
and temperature (250
o
C, 275
o
C and 300
o
C) on the forming behavior of Al

98
5754-O and Mg AZ31-O cups. These velocity and temperature ranges were
obtained by the preliminary experiments that were conducted. During the
experiments, punch load and the temperature at the cop bottom are also
measured.
7.7.1.1 Results for Al 5754-O
Table 7.6 shows the summary of the conducted experiments. At various
temperatures and forming velocities, formed and fractured cups along with
their heights and flange diameters are given.
Draw Ratio Dwell time (sec)
Min. and max. blank holder
force/pressure (kN/MPa)
2.5 30 4.4 / 0.82 8 / 1.5
Punch Velocity (mm/s)
5 15 50

Cup Height / Flange Diameter of the drawn cup (mm)
2
5
0

47.5 / 53.1
Cup fractured

Cup fractured

2
7
5

48 / 52.5 48.8 / 54.4
Cup fractured

T
e
m
p
e
r
a
t
u
r
e

(

C
)

3
0
0

48.7 / 51 49 / 53.3
Cup fractured

Table 7.6: Experimental results for Al 5754-O


99
Figure 7.27 and Figure 7.28 show the thickness distributions for the
aluminum cups. Plots indicate that higher temperature helps reduce thinning
at the punch corner, which is the location of maximum thinning.

Figure 7.27: Effect of temperature on thickness distribution (5 mm/s, DR:
2.5)


100

Figure 7.28: Effect of temperature on thickness distribution (15 mm/s, DR: 2.5)
Figure 7.29 and Figure 7.30 show the punch load-stroke measurements
at various temperatures (250
o
C, 275
o
C and 300
o
C) and punch velocities (5
mm/s and 15 mm/s).

101

Figure 7.29: Punch load stroke curves
at 5mm/s for different temperatures

Figure 7.30: Punch load stroke
curves at 15mm/s for different
temperatures
It is important to obtain the temperature change in the cup as forming
process proceeds. This data could be helpful in order to inversely obtain more
accurate heat transfer coefficients at the sheet-punch interface through finite
element analysis. Therefore, design modifications were made in the tooling in
order to record the temperature change at the center of the sheet during the
dwelling and the forming stage. Figure 7.31 and Figure 7.32 show the cup
bottom temperature change curves for 5 mm/s and 15 mm/s punch velocities, at
various tool temperatures. (250
o
C, 275
o
C and 300
o
C) As it is seen, the sheet
starts cooling once it is in contact with the cold punch. The pattern of the
curves and the amount of temperature drop at the cup bottom is determined by
the forming velocity.

102

Figure 7.31: Change in cup bottom temperature at 5 mm/s (Punch temp: ~55
o
C, 65
o
C, 73
o
C)

Figure 7.32: Change in cup bottom temperature at 15 mm/s (Punch temp: ~55
o
C, 65
o
C)

It is seen from Figure 7.31 and Figure 7.32 that the sheet temperatures
are lower than the set tool temperatures. Possible reasons for this could be a)
the temperature gradient within the heated die block. Die surface temperatures

103
are usually ~15-20
o
C lower than the set temperature b) the center of the sheet
(~ 45 mm in diameter) is open to air from both sides, which causes some heat
loss.
7.7.1.2 Results for Mg AZ31-O (Supplier A)
Table 7.7 shows the summary of the conducted experiments and
provides cup heights and flange diameters. At various temperatures and
forming velocities, all cups are fully formed and there is no flange left.
Magnesium sheets were also cut to be 100 mm in diameter, which corresponds
to a draw ratio of 2.5. All the process conditions were kept exactly the same. It
is important to notice the difference in cup heights of aluminum (Table 7.6)
and magnesium (Table 7.7). Under same conditions aluminum cups are
generally 4 mm deeper than the magnesium cups. This shows that aluminum
cups are stretched more than the magnesium cups. This observation is also
backed by the lower thickness distribution of the aluminum cups.
Figure 7.33, Figure 7.34 and Figure 7.35 show the thickness distributions for
the magnesium cups at constant punch velocities (5 mm/s, 15 mm/s, 50 mm/s)
but at different temperatures. (250
o
C, 275
o
C, 300
o
C) Plots indicate that at
constant velocity, temperature does not change the thickness distribution
significantly. Figure 7.36, Figure 7.37 and Figure 7.38, show the thickness

104
distributions for the magnesium cups at constant temperatures but at different
velocities. The effect of velocity in thickness distribution is clearer in the plots.

Draw Ratio Dwell time (sec)
Min. and max. blank holder
force/pressure (kN/MPa)
2.5 30 4.4 / 0.82 8 / 1.5
Punch Velocity (mm/s)
5 15 50

Cup Height / Flange Diameter of the drawn cup (mm)
2
5
0

43.2 / 45.7 44.2 / 45.3 45.3 / 45.5
2
7
5

43.9 / 45.2 44 / 45.1 44.4 / 45.5
T
e
m
p
e
r
a
t
u
r
e

(

C
)

3
0
0

44.7 / 44.6 44.5 / 44.8 45.2 / 44.8
Table 7.7: Experimental results for Mg AZ31-O (Supplier A)





105

Figure 7.33: Effect of temperature on thickness distribution (5 mm/s, DR: 2.5)

Figure 7.34: Effect of temperature on thickness distribution (15 mm/s, DR: 2.5)

106

Figure 7.35: Effect of temperature on thickness distribution (50 mm/s, DR: 2.5)


Figure 7.36: Effect of forming velocity on thickness distribution (250
o
C, DR:
2.5)

107

Figure 7.37: Effect of forming velocity on thickness distribution (275
o
C, DR:
2.5)

Figure 7.38: Effect of forming velocity on thickness distribution (300
o
C, DR:
2.5)


108
Figure 7.39, Figure 7.40 and Figure 7.41 show the punch load-stroke
measurements at various temperatures and forming velocities for the Mg
(Supplier A). As expected, punch load decreases with temperature and
increases with velocity. Punch loads for magnesium are approximately 10 kN
lower compared to the ones in the aluminum experiments.


Figure 7.39: Punch load stroke curves at 5mm/s for different temperatures

109

Figure 7.40: Punch load stroke curves at 15mm/s for different temperatures

Figure 7.41: Punch load stroke curves at 50mm/s for different temperatures


110
Temperature change at the center of the sheet during the dwelling and
the forming stage was also recorded. Figure 7.42, Figure 7.43 and Figure 7.44
show the cup bottom temperature change curves for 5 mm/s, 15 mm/s and 50
mm/s forming velocities, at various tool temperatures. (250
o
C, 275
o
C and 300
o
C) Same temperature change patterns were observed for the magnesium cup,
and the patterns are determined by the punch velocity.

Figure 7.42: Change in cup bottom temperature at 5 mm/s

111

Figure 7.43: Change in cup bottom temperature at 15 mm/s

Figure 7.44: Change in cup bottom temperature at 50 mm/s


112

Figure 7.45: Pictures of the formed cups from Al 5754-O (left) and Mg AZ31-O
(Supp.A) (right)

Figure 7.45 show the pictures of the formed aluminum and magnesium
cups. Wrinkling was seen in the rolling direction in almost every formed
magnesium cup.
As a summary, Figure 7.46 and Figure 7.47 show the effects of
temperature and punch velocity on the maximum thinning values obtained at
the cup bottom corner (critical location for design) for the Al and Mg alloy
respectively.

113

Figure 7.46: Effect of blank temperature and punch velocity upon wall thinning
at the bottom corner of the drawn cup for the Al 5754-O alloy

Figure 7.47: Effect of blank temperature and punch velocity on thinning at the
bottom corner of the drawn cup for the Mg AZ31-O alloy

114
Plots indicate that the amount of thinning in Mg alloy is lower compared
to aluminum, which might be due to high anisotropy values seen in
magnesium sheets.
7.7.1.3 Results for Mg AZ31-O (Supplier B)
Experiments were conducted in order to investigate the difference in
material properties of magnesium AZ31-O sheet from a different supplier.
Same experimental conditions were repeated in Table 7.6 and Table 7.7. No
cups were formed successfully. Experiments were repeated with a lower
velocity and a lower temperature. The reason for this was, due to various
mechanical property tests available in literature that uses the AZ31-O from
supplier B shows less formability above 235
o
C. Therefore, tool temperature
was decreased to 225
o
C. This change also did not provide successful cups.
(Table 7.8)
Velocity (mm/s)
2.5 5
Temperature (225
o
C) Not Formed Not Formed
Table 7.8: Experiments at 225
o
C
Successful cups were obtained at 250
o
C and at 1 mm/s and 2.5 mm/s.
(Table 7.9) At the cup formed at 2.5 mm/s, cracks were seen in the area that
was bent around the punch. Since all the experiments were conducted under

115
same conditions, this result shows that there is significant difference in the
mechanical behavior of the material from different suppliers.
Velocity (mm/s)
1 2.5
Temperature (250
o
C) Formed Formed
Table 7.9: Experiments at 250
o
C and lower velocities
7.8 Effect of variable forming velocity
It was mentioned earlier that the most obvious benefit of the servo press
is the flexibility it offers. This flexibility comes from the ability to program the
speed and motion of the slide in an infinite number of ways. Figure 7.3
provided an explanation for a general ram motion that can be used in warm
deep drawing process. So far path (1-2-3-4-6-8) shown in Figure 7.3 is used,
which means that a constant velocity was set between points 4 and 6. It is
mentioned earlier that in deep drawing, severe deformation takes place around
the punch and the die corner. Therefore, as it is seen in Figure 7.3 , a variable
velocity could also be set between points 4-5 and 5-6. For example, a slower
velocity between 4 and 5 for forming around the punch and the die corner
radii, and a faster velocity between points 5 and 6 to complete the forming
stroke. The main unknown in the 4-5-6 path is the amount of stroke needed to
apply the slower velocity between points 4 and 5. In order to come up with an
approximate estimate the critical concept mentioned earlier is used.

116
In the existing tool, r
D
is 6 mm, r
P
is 4 mm and the sheet thickness, t, is
1.3 mm. Therefore the critical stroke is calculated to be ~12 mm. Based on this;
experiments were conducted to look at the effect of variable velocity on the
thickness distribution for Al 5754-O and Mg AZ31-O. Experiments were
repeated for two cups.
Figure 7.48 shows that while an aluminum cup cannot be formed at 40
mm/s, it can be formed by using a variable velocity profile (5/40 mm/s or 10/40
mm/s). In these graphs, for example; 5/40 mm/s means that the cup was
drawn at 5 mm/s during the first 12 mm of stroke and the rest of the stroke was
completed with 40 mm/s. Figure 7.49 shows the results for the Mg alloy. The
difference in the thickness of the cup formed at 40 mm/s is lower compared to
the 5 mm/s. However, there is significant similarity in the thickness
distributions of cups formed at 5 mm/s and 5/40 mm/s. Therefore, instead of
forming the cup at 5 mm/s, results show that it can also be formed using 5/40
mm/s. The benefit of this is shown in Table 7.10.

117

Figure 7.48: Effect of variable forming speed on the thickness distribution of
the drawn Al cups

Figure 7.49: Effect of variable forming speed on the thickness distribution of
the drawn Mg cups



118
1.1 sec 40
2.4 + 1=3.4
sec
5 / 40
9 5
Drawing
time
(s)
Punch
Velocity
(mm/s)
1.1 sec 40
2.4 + 1=3.4
sec
5 / 40
9 5
Drawing
time
(s)
Punch
Velocity
(mm/s)
5.6 sec
saving
per part

Table 7.10: Significant savings in drawing time is obtained through the use of
variable forming velocity
7.9 Conclusions
The following conclusions can be drawn from these deep drawing experiments
conducted at elevated temperatures:
a) The dwell time needed to heat blank is reduced with increasing interface
pressure, i.e. Blank Holder Pressure (BHP).
b) Surface roughness (R
a
) values decrease with increasing interface
pressures. The decrease was found to be 13 %, 30 % and 65 % for
Al5052-H32, MgAZ31-O (Supplier A), MgAZ31-O (Supplier B),
respectively.
c) Hardness measurements before and after the experiments revealed that,
hardness (Brinell scale) of A5052-H32 decreased from 60 to 55 in

119
average. However, both magnesium alloys showed a small increase in
hardness values.
d) During the dwell time when heating the blank, a slight dome was
formed in the free center of the blank. As the interface pressure
increased, the height of the dome also increased, for all investigated
materials. This might be due to the limitation of the radial thermal
expansion of the material because of higher interface pressures.
Therefore, expansion can happen at the center of the sheet since it is
open to air from both sides.
e) Three lubricants out of a total of four investigated were successful in
drawing a cup. However, only the PTFE based lubricant provided good
thinning distribution and best surface conditions. The other successful
lubricants left a residue on the formed cup and the tool.
f) Experiments showed that for Al5754-O, higher temperature helps to
reduce thinning at the punch corner.
g) Significant variations in the deformation of the same magnesium alloy
blank, from two suppliers, is observed.
h) At constant velocity but at different temperatures (250
o
C, 275
o
C and
300
o
C), there is no significant difference in wall thickness distribution

120
of the cups formed from Mg AZ31-O. However, at constant temperature,
slight increase in thinning is seen with increasing velocity.
i) The effect of variable punch velocity (slower at the initial stages of
drawing) was found to be significant. Instead of forming with a slower
constant velocity for the full stroke, a slower velocity within the critical
stroke and a faster velocity after this stroke can be used if the press can
provide this type of velocity control. In our experiments, 60 % reduction
in the drawing time was achieved.


121
CHAPTER 8


MODELING OF NON-ISOTHERMAL DEEP DRAWING PROCESS

Traditionally, metal forming processes have been developed based on
expensive experimental trials. In recent years, finite element (FE) simulations
have been extensively used to reduce the amount of experiments and trial and
error involved in the process development. However, reliable results can be
expected only if the material properties and related process parameters input to
the FE simulations are accurate. Due to the highly non-linear nature and
coupling of variables in warm forming of Mg and Al, experimental
determination of variables such as material properties, friction, etc., are
difficult and expensive, but necessary for a reliable analysis.
To understand the complex interactions between material properties,
temperature and punch velocity (strain rate), it is necessary to model the
process using FEM. However, this modeling requires reliable input data, i.e. a)
material properties in function of strain, strain rate and temperature, b) heat
transfer coefficients at the die/blank interface and at the punch/cup bottom
interface, c) coefficient of friction in function of interface pressure and
temperature.

122
In non-isothermal sheet forming processes, the temperature of the sheet
during forming is not constant. In other words, different locations in the sheet
are subject to different temperatures during the operation. In the existing
experimental setup, the punch was cooled in order to increase the strength of
the material in the punch corner area while the flange was kept at higher
temperature to promote easy flow of the material. Having the right temperature
distribution in the part is then the key factor for the success of the process.
8.1. Determination of the heat transfer coefficients
In order to perform more accurate FE analysis of non-isothermal
processes, the user needs to approximately know the interface Heat Transfer
Coefficient. When conducting FEA, it is a common practice to assume a
constant value for the HTC. This simplification can reduce the accuracy of
FEA. Therefore instead of picking an average value from literature it was
decided to determine the heat transfer coefficients through the temperature
measurements conducted.
Various researchers, mostly in the forging area, have experimentally
demonstrated that the interface HTC is a function of interface pressure and
interface temperature [Semiatin et.al, 1987]. Pressure and temperature
conditions generate microscopic changes in the contact area conditions

123
(roughness and micro-welding) and significantly affect the heat transfer
phenomena.
8.1.1. Inverse Analysis
In order to determine the HTC values under different temperatures and
interface pressures an FEM based inverse optimization method [Kim, et.al.,
2001] has been implemented into Deform2D V9.0. In order to conduct this
analysis, temperature time curves given in Figure 7.11 were used as input.
Prior to the inverse analysis, heat transfer analysis was conducted by selecting
constant heat transfer coefficients and predicted temperature-time curves were
plotted against the experimental ones. Figure 8.1 shows the measured and
calculated temperature time curves.

Figure 8.1 Experimental and calculated temperature-time curves

124
8.1.2. Setup of the inverse analysis problem
Figure 8.2 shows the fixture in contact with the blank holder and the
die. Only the fixture was modeled and the areas where die and blank holder are
in touch with the fixture were selected and assigned as heat exchange areas
(Figure 8.3). The temperature measurement locations in the FE model of the
fixture are decided in correspondence with the location that the experimental
measurements were conducted. Experimentally measured temperature-time
curves (target functions for the optimization) are provided in tabular form for
the calculation.

Figure 8.2: Schematic view of the
FE model





Figure 8.3: Modeling of the measurement
fixture with the temperature
measurement point

The last step in the model setup is to assign an initial guess value to the HTC
and a certain number of temperature control points (a minimum of two control
points are required). The optimization program, after each iteration changes the
thermocouple
location
Areas of heat
exchange

125
value of the HTC in order to obtain a closer match with the target function. The
solution provides the variation of the HTC with respect to temperature.
Figure 8.4 shows the calculated heat transfer coefficients with changing
interface pressure. The calculated values were compared with the only
available data in [Groche et.al., 2002].

Figure 8.4: Calculated heat transfer coefficients
8.2. Non-isothermal deep drawing of Mg AZ31-O
8.2.1. Mg AZ31-O flow stress in literature
Among the commercially available magnesium alloys, AZ31B is, by far, the
most used for sheet forming applications. In the past, many researchers,

126
interested in sheet forming of lightweight materials, have studied the properties
of this alloy.
A comprehensive literature study to collect flow stress data obtained by
different researchers at various temperatures and strain rate conditions was
conducted. As a result of the study a flow stress database, for the AZ31 sheet
material was created [Sivakumar, et al, 2006]. Graphical representation of this
database is reported in Figure 8.5. Due to the observation in the property
variation of Mg AZ31-O in the bulging and drawing experiments, it was
decided not to use the literature data. Therefore, the flow stress obtained from
the hydraulic bulge test was used.

127

Behren
Xu
A, Jager
Chen Takuda
Doege
Chastel
Yi
Hecht S, Jager
Agnew
Doege
Takuda 05

Figure 8.5 Graphical representation of the database for Magnesium AZ31
Tensile tests conducted at different strain rates and temperatures [Sivakumar,
et al, 2006].
8.2.2. FE modeling in LS-Dyna3D
LS-Dyna is commonly used to model isothermal sheet metal forming
processes. There is almost no experience in conducting non-isothermal analysis
therefore it is more of a challenge to understand how to conduct a non-
isothermal forming process analysis. Once the first model was built, a series of
trial-and-error simulations were conducted to better understand the
significance of some input parameters and to make sure the model was working
properly.
The aim of the present work is to discuss two main issues encountered
during the setup of the FE simulation; namely:

128
flow stress data implementation,
definition of the thermal contact,

Figure 8.6 FE model for non-isothermal simulations of deep drawing of round
cups

8.2.2.1. Plastic-thermal material model available in LS-Dyna
For warm forming under non-isothermal conditions, material models are
needed where the equivalent plastic stress is defined in function of equivalent
strain, equivalent strain rate and temperature.
Only three plastic-thermal material models are available in LS Dyna. These are:
Mat 004 (*MAT_ELASTIC_PLASTIC_THERMAL)
Mat 015 (*MAT_JOHNSON_COOK)
Mat 106 (*MAT_ELASTIC_VISCO_PLASTIC_THERMAL)
Die
Blank holder
Sheet
Punch

129
Moreover, only the last two can include strain rate effects. They are discussed
briefly, below.
MAT 015 (*MAT_JOHNSON_COOK)
The Johnson-Cook material is strain and temperature sensitive plasticity; it is
usually used for problems where the strain rates vary over large ranges and
temperature increases due to plastic deformation.
Johnson and Cook express the flow stress as:
( ) ( )
1
0 2 1
, , 1 ln 1-
m
eq n
eq eq eq eq
T T
T A B c
T T


= + +

(
| | | |
(
| |
\ . \ . (

&
&
&


where:
A, B, c, n and m are input constants.
0

&
is set to 1 s
-1
by default.

MAT 106 (*MAT_ELASTIC_VISCO_PLASTIC_THERMAL)
Mat 106 was selected in our study to input flow stress data in LS-Dyna. This
model was preferred over the others since it gives the user more flexibility in
fitting the experimental data for different values of temperatures and strain
rates. The flow stress, in function of strain, strain rate, and temperature is
calculated using the formula below:
eq
&
eq
&

0

&

0
&

130
( ) ( ) ( )
( )
( )
1

P T
, , 1+
eq
eq eq eq eq
T MFS SF T
C T


(
| |
=
( |
|
(
\ .

&
&
Mater Flow
Stress Viscous component
Effect of
temperature
I II
III

I. The first component in the formula is the master flow stress, (MFS).
It is in the form of a table where stress and strain values are inserted as
data-points. It is suggested to use the flow stress of the material at room
temperature as the master flow stress.
II. The second component, SF, scales the masters flow stress values to
account for temperature effects. SF(T) is input in tabular form.
III. The last component is the viscous component, which includes strain
rate dependency. The viscous parameters C and P are used to take
into account strain rate effects. C(T) and P(T) are input in tabular form.
At the end, a total of four tables need to be input: MFS(
eq

), SF(T), C(T) and


P(T). These tables are used in the above formula to give flow stress value for
each value of strain, strain rate and temperature.
The fact that aluminum and magnesium material data can satisfactorily
be fit using material model 106, does not imply that every flow stress data can
by fitted using *MAT106. If the hardening of the material varies too much with
temperature and/or stain rate, the accuracy of the fitting could be very poor.
Effect of
strain
Effect of
temperatu

Effect of strain rate
eq
&
eq
&

131
Therefore, problems in fitting were encountered where there is slight work
softening in the flow stress. Through experience it is realized that it is better if
the flow stress data could be inserted entirely in data-point format.
8.2.2.2. Thermal contact definition
Thermal contact definition plays an essential role in the simulation of
non-isothermal warm forming process since it determines the amount of heat
that is transferred between the objects in contact. During forming, the heated
sheet comes in contact with the punch (at room temperature) and looses heat.
The region of the sheet metal in contact with the punch has lower temperature
compared to the sheet in contact with the die. The lower temperature in the
wall and higher temperature in the flange are essential in warm deep drawing
because, increase in the flow stress due to decrease of temperature enables the
cup wall at the punch corner to support more stress.
It must also be considered that, for given interface conditions (materials,
lubricants, surface finish), the heat transfer (or thermal contact) depends on
interface pressure that may vary within the entire contact area between sheet
and tool.


132
In LS-Dyna, the thermal contact can be activated from the same card used to
define the mechanical contact (see Figure 8.7). The algorithm accounts for the
possibility of having a gap of fluid between the sliding surfaces.
Since in our case there is no fluid between the objects, the only significant
parameters to input are:
HTC = heat transfer coefficient (h
cont
).
GCRIT = Critical gap (l
min
)
GMAX = no thermal contact if gap is grater than this value (l
max
)
LS-Dyna uses (h
cont
) for gaps in the range of 0<l
gap
<l
min
; l
gap
is recalculated
every step by the code based on deformation.
With the use of GMAX=GCRIT, and CF=FRAD=0 the assumption of no fluid
is modeled.

133

Figure 8.7 Contact-Forming card used to define mechanical and thermal
contact parameters.
Values of GCRIT of about one tenth of the sheet thickness have been used in
simulations. A reasonable value of this parameter was found by trial and error.
If lower values are used, the heat transfer between bodies could, depending of
the mesh size, become null.
The effect of the interface pressure on the heat transfer coefficient (HTC)
cannot be modeled due to the selected contact model.
8.2.3. Simulation matrix
Table 8.1 shows the matrix of the conducted simulations. Comparisons with
experiments were done for there different die temperatures and forming
mechanical
contact
definition
thermal
contact
definition

134
velocities from 2 mm/s to 35 mm/s. The drawing ratio of the cups varied from
2.7 to 3.0.
Simulations Matrix
Name
Drawing Ratio
(DR)
Die/BH
temperature
Forming
velocity

M1 2.7 10 mm/s
M2 2.8 10 mm/s
M3 3.0
250 C
2 mm/s

M4 2.8 25 mm/s
M5 3.0
275 C
10 mm/s

M6 2.7 35 mm/s
M7 2.8 12 mm/s
M8 3.0
300 C
5 mm/s
Table 8.1 AZ31B-O simulation matrix
A summary of the most important thermal and mechanical parameters used in
simulations is given in Table 8.2. Friction coefficient value of 0.04 was used.
This value is relatively lower than what is normally reported in literature (form
0.1 to 0.16) because of the low friction PTFE film lubricant used during
experiments. Our previous experience shows that the friction coefficient can be
in the range of 0.04 when PTFE film type lubricants are used.

135
Table 8.2 Summary of input data used for AZ31B-O simulations
8.2.4. Comparison of FE predictions and experimental results
Figure 8.8 to Figure 8.15 show the comparison between calculated and
experimental punch load vs. stroke curves and thinning distributions for the
eight selected cases.
It is noticeable that punch loads have always been overestimated. Depending
on the forming conditions, calculated loads are, in fact, 20% to 50% higher
than in experiments. This was not a surprising result. It is known that
magnesium behavior is different in flange (where compressive stresses are
dominant) than in tensile so the material flows much easily in reality than in
simulation. Another possible reason is also the different yield surface of
magnesium. It is also known that anisotropy of magnesium alloy sheet varies
Thermo-mechanical data used for simulations



thermal conductivity 77 W m
-1
C
-1

Fraction of
mechanical work
converted to heat
95 %
specific heat 1020 J kg
-1
C
-1

Interface heat transfer coefficient (HTC)

sheet-punch 5000 W m
-2
C
-1
Friction coefficient,
sheet-die 1000 W m
-2
C
-1
sheet-punch 0.04
sheet-binder 1000 W m
-2
C
-1
sheet-die 0.04
sheet-environment 50 W m
-2
C
-1


sheet-binder 0.04

136
between the rolling and the transverse direction under tensile and compressive
stresses.
Thinning distribution trends matched experimental measurements for forming
velocity up to 10 mm/s. In two cases where forming velocity was relatively
higher, 25 mm/s (Figure 8.11) and 35 mm/s (Figure 8.13), FE predictions
showed very high thinning compared with experiments.
Sim # :
M1
Drawing Ratio
(DR)
2.7
Die/BH Temperature 250 C
forming velocity 10 mm/s
0
5
10
15
20
25
30
0 10 20 30 40 50 60
EXP
SIM
Stroke [mm]
P
u
n
c
h

l
o
a
d

[
k
N
]

-0.7
-0.6
-0.5
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0 10 20 30 40 50 60 70
EXPTD
EXPRD
SIM
Curvilinear length [mm]
T
h
i
n
n
i
n
g
20 mm
60 mm

Figure 8.8 Experimental and computed punch load vs. stroke curves (left) and
cup thinning distributions (right). (RD=rolling direction; TD=transverse
direction)


SIM.


Exp

137

Sim # :
M2
Drawing
Ratio (DR)
2.8
Die/BH Temperature 250 C
forming velocity 10 mm/s
0
5
10
15
20
25
30
0 10 20 30 40 50 60
EXP
SIM
Stroke [mm]
P
u
n
c
h

l
o
a
d

[
k
N
]

-0.7
-0.6
-0.5
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0 10 20 30 40 50 60 70
EXPTD
EXPRD
SIM
Curvilinear length [mm]
T
h
i
n
n
i
n
g
20 mm
75 mm

Figure 8.9 Experimental and computed punch load vs. stroke curves (left) and
cup thinning distributions (right). (RD=rolling direction; TD=transverse
direction)
Sim # :
M3
Drawing Ratio
(DR)
3.0
Die/BH Temperature 250 C
forming velocity 2 mm/s
0
5
10
15
20
25
30
0 10 20 30 40 50 60
EXP
SIM
Stroke [mm]
P
u
n
c
h

l
o
a
d

[
k
N
]

-0.7
-0.6
-0.5
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0 10 20 30 40 50 60 70
EXPTD
EXPRD
SIM
Curvilinear length [mm]
T
h
i
n
n
i
n
g
20 mm
70 mm

Figure 8.10 Experimental and computed punch load vs. stroke curves (left) and
cup thinning distributions (right). (RD=rolling direction; TD=transverse
direction)
SIM.


Exp
SIM.


Exp

138
Sim # :
M4
Drawing
Ratio (DR)
2.8
Die/BH Temperature 275 C
forming velocity 25 mm/s
0
5
10
15
20
25
30
0 10 20 30 40 50 60
EXP
SIM
Stroke [mm]


-0.7
-0.6
-0.5
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0 10 20 30 40 50 60 70
EXPTD
EXPRD
SIM
Curvilinear length [mm]
T
h
i
n
n
i
n
g
20 mm
65 mm

Figure 8.11 Experimental and computed punch load vs. stroke curves (left) and
cup thinning distributions (right). (RD=rolling direction; TD=transverse
direction)
Sim # :
M5
Drawing
Ratio (DR)
3.0
Die/BH Temperature 275 C
forming velocity 10 mm/s
0
5
10
15
20
25
30
0 10 20 30 40 50 60 70
EXP
SIM
Stroke [mm]
P
u
n
c
h

l
o
a
d

[
k
N
]

-0.7
-0.6
-0.5
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0 10 20 30 40 50 60 70 80
EXP TD
EXP RD
SIM
Curvilinear length [mm]
T
h
i
n
n
i
n
g
20 mm
75 mm

Figure 8.12 Experimental and computed punch load vs. stroke curves (left) and
cup thinning distributions (right). (RD=rolling direction; TD=transverse
direction)

SIM.


Exp
SIM.


Exp

139
Sim # :
M6
Drawing Ratio
(DR)
2.7
Die/BH Temperature 300 C
forming velocity 35 mm/s
0
5
10
15
20
25
30
0 10 20 30 40 50 60
EXP
SIM
Stroke [mm]
P
u
n
c
h

l
o
a
d

[
k
N
]

-0.7
-0.6
-0.5
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0 10 20 30 40 50 60 70
EXPTD
EXPRD
SIM
Curvilinear length [mm]
T
h
i
n
n
i
n
g
20 mm
65 mm

Figure 8.13 Experimental and computed punch load vs. stroke curves (left) and
cup thinning distributions (right). (RD=rolling direction; TD=transverse
direction)
Sim # :
M7
Drawing
Ratio (DR)
2.8
Die/BH Temperature 300 C
forming velocity 12 mm/s
0
5
10
15
20
25
30
0 10 20 30 40 50 60
EXP
SIM
Stroke [mm]
P
u
n
c
h

l
o
a
d

[
k
N
]

-0.7
-0.6
-0.5
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0 10 20 30 40 50 60 70
EXPTD
EXPRD
SIM
Curvilinear length [mm]
T
h
i
n
n
i
n
g
20 mm
65 mm

Figure 8.14 Experimental and computed punch load vs. stroke curves (left) and
cup thinning distributions (right). (RD=rolling direction; TD=transverse
direction)

SIM.


Exp
SIM.


Exp

140
Sim # :
M8
Drawing
Ratio (DR)
3.0
Die/BH Temperature 300 C
forming velocity 5 mm/s
0
5
10
15
20
25
30
0 10 20 30 40 50 60 70
EXP
SIM
Stroke [mm]
P
u
n
c
h

l
o
a
d

[
k
N
]

-0.7
-0.6
-0.5
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0 10 20 30 40 50 60 70 80
EXPTD
EXPRD
SIM
Curvilinear length [mm]
T
h
i
n
n
i
n
g
20 mm
75 mm

Figure 8.15 Experimental and computed punch load vs. stroke curves (left) and
cup thinning distributions (right). (RD=rolling direction; TD=transverse
direction
8.3. Non-isothermal deep drawing of Al 5754-O
FE modeling of the nonisothermal deep drawing process of aluminum
is conducted and numerical predictions are compared with the experimental
measurements using DEFORM 2D. Due to the inconveniences encountered
using LS-Dyna, analysis was conducted with Deform 2D. Because Deform2D
allowed to input a) flow stress parameters to be input as data points, rather
than using a material model and b) heat transfer coefficients as function of
pressure. Over all Deform2D is designed for non-isothermal forging process
therefore it is believed to perform better in non-isothermal analysis. Table 8.3
provides the list of process parameters, thermal and mechanical properties of
SIM.


Exp

141
the Al alloy sheet. The flow stress data used in this study is obtained from
[Boogard, 2001]
Punch shoulder radius (mm) r
p
4
Punch diameter (mm) D
p
40
Die shoulder radius (mm) r
d


6

G
e
o
m
e
t
r
y

Die inner diameter (mm) D
d
44.6

Initial sheet thickness s
0
(mm) 1.3
Thermal conductivity (sheet) (N/(sec C))
140 (Incropera, et.al.,
(2002)
Heat capacity (sheet) (N/(mm
2
C))
2.40 (Incropera, et.al.,
2002)
Thermal conductivity (tool) (N/(sec C))
60.5 (Incropera, et.al.,
2002)
Heat capacity (tool) (N/(mm
2
C))
3.41 (Incropera, et.al.,
2002)
Interface heat transfer coefficient
(N/(secmmC))
6
Initial die / BH temperature (C) 250/275/300
Initial sheet temperature (C) 250/275/300
Initial punch temperature (C) 60
Youngs modulus, E (GPa) 68.9 (room temp)
Poissons ratio, 0.33
Flow stress curve [Boogard, 2001]
T
h
e
r
m
a
l

a
n
d

M
e
c
h
a
n
i
c
a
l

p
r
o
p
e
r
t
i
e
s

Friction coefficient, 0.04
Table 8.3: List of input parameters to the FE model (Al 5754-O)
In Table 8.3, the effect of temperature on the thermal conductivity and the
effect of radiation is neglected since the percentage change is approximately
2% for Al and steel alloys. The effect of temperature on the heat capacity is also

142
negligible within the studied temperature ranges. It is known that E-Modulus
also changes with temperature; however, in our FEA we have neglected this
effect since the deformation dominantly takes place in the plastic range.
Selection of the friction coefficient is based on a) previous experimental
experience on testing PTFE based lubricants b) through acceptable comparison
with the FE predictions.
8.3.1. Yield criterion adopted in the FE model for the Al 5754-O
Von Mises yield criterion is widely accepted in plasticity and it represents the
deformation of steels to a great accuracy. For Al alloys various yield surface
models have been proposed by researchers. [Barlat et.al., 2003] is widely
accepted and implemented in various FE codes. [Naka et.al., 2003] has
investigated the effects of temperature on the yield locus for 5083 aluminum
alloy sheet in forming operations. He suggests that Logan-Hosford and Barlat
criteria can well describe the yield loci and the equivalent stress-strain curves
for the Al-Mg alloy sheet at various temperatures. [Figure 8.16] shows the
experimentally determined yield loci for the aluminum alloy sheet at various
temperatures and those predicted by various suggested models. It is seen from
[Figure 8.16] that the Logan-Hosford and Barlat models fit better, however Von
Mises yield surface is also acceptable since the difference might be negligible
and Al 5754-O is isotropic. Therefore, in the present study, the Von Mises yield

143
locus is adopted to conduct the non-isothermal analysis of deep drawing
process. Thickness distributions and punch load predictions were compared
with the experimental measurements.

Figure 8.16: Comparison of experimental yield loci and those predicted by the
Von Mises, Hill, Tresca, Logan-Hosford and Barlat criteria under biaxial stress
condition for Al 5083-O alloy sheet [Naka, et.al., 2003]
8.3.2. Flow stress curves for aluminum
Flow stress data for the Al 5754-O alloy sheet, used in this study, was obtained
from [Boogard, 2001]. Data in the mentioned reference is obtained through
tensile tests up to 250
o
C with strain rates of 0.1 1/s, 0.02 1/s and 0.002 1/s.
Therefore the data for 275
o
C and 300
o
C were obtained by extrapolation as
shown in Figure 8.17.
In the definition of flow stress curves, softening behavior was not taken into
account. In other words, once the maximum stress point was reached in the

144
flow stress curve that maximum stress level was kept constant with respect to
the strain. [Figure 8.18]

y = 2.67764E-06x
3
- 3.69846E-03x
2
+ 3.85929E-01x + 2.51646E+02
y = -2.75832E-03x
2
+ 1.39462E-01x + 2.66395E+02
y = -7.68194E-08x
3
- 2.23240E-03x
2
+ 2.48930E-01x + 2.51066E+02
50
100
150
200
250
300
0 50 100 150 200 250 300 350
Temperatur C
F
l
o
w

S
t
r
e
s
s

M
P
a
Strain rate 0.002
Strain rate 0.02
Strain rate 0.1
Poly. ( Strain rate 0.02)
Poly. (Strain rate 0.002)
Poly. (Strain rate 0.1)

Figure 8.17: Extrapolated flow stress versus
temperature for different strain rates(original
data from Boogard, 2001)

T
r
u
e

S
t
r
e
s
s
True Strain
Max. stress points

Figure 8.18: Elimination of the
softening behavior from the stress-
strain curve

0.002
0.1
0.02

145

8.3.3. Comparison of numerical predictions with experimental
measurements
FEA of the experimental cases given in Table 7.6 was conducted. Initially
various constant heat transfer coefficients were used to investigate their effect
on the process. It was found that when constant HTC values of 2 kW/m
2
C and
3 kW/m
2
C were used the cup fractured at the first 15 mm of the stroke. Punch
load and thickness distribution curves are shown in Figure 8.19 through Figure
8.26, for various cases. It is seen that the predictions for 5 mm/s are acceptable
except the case with 15 mm/s. This is mostly due to the lack of flow stress data
at higher strain rates. Results indicate that the adopted Von Mises yield
criterion provides acceptable accuracy. The maximum difference in punch load
was obtained for 250
o
C with 5mm/s is around 10%. Differences in the other
cases are relatively lower.
Figure 8.27 and Figure 8.28 show the comparison of results for one case when
the heat transfer coefficient is input as function of interface pressure given in
Figure 8.4. There is a slight improvement in the thickness distribution.


146

Figure 8.19: Comparison of punch load
predictions using various heat transfer
coefficients (HTC: kW/m
2
C) with
experiment (5 mm/s at 250
o
C)

Figure 8.20: Predicted thickness
distribution comparison with various
heat transfer coefficients
(5 mm/s at 250
o
C)


Figure 8.21: Comparison of punch load
prediction with experiment (5 mm/s at
275
o
C)

Figure 8.22: Predicted thickness
distribution comparison with
experiments (5 mm/s at 275
o
C)


147

Figure 8.23: Comparison of punch load
prediction with experiment (5 mm/s at
300
o
C)

Figure 8.24: Predicted thickness
distribution comparison with
experiments (5 mm/s at 300
o
C)


Figure 8.25: Comparison of punch
load prediction with experiment (15
mm/s at 275
o
C)

Figure 8.26: Predicted thickness
distribution comparison with
experiments (15 mm/s at 275
o
C)


148

Figure 8.27: Predicted thickness
distribution comparison with
experiments (5 mm/s at 300
o
C)

Figure 8.28: Predicted thickness
distribution comparison with
experiments (5 mm/s at 300
o
C)

8.4. Non-isothermal modeling of SHF-P
Experience obtained from modeling the SHF-P (at room temperature) and the
deep drawing under non-isothermal conditions is combined for modeling the
non-isothermal SHF-P process. Heat transfer takes place in four interfaces in
elevated temperature SHF-P in which appropriate inputs to the model is
needed (Figure 8.29).

149
BLANK HOLDER
DIE
P
PUNCH
F
Liquid
q
q
q
q
q
Interface A
Interface B
Interface C
Interface D
V
q

Figure 8.29: Interfaces in elevated temperature SHF-P process

As an initial modeling exercise, some assumptions are made. These are;
a) Non-isothermal analysis is conducted (heat transfer at the sheet-die, sheet-
blank holder and sheet-punch interfaces)
b) Pressure is applied to the complete sheet (additional force generated due to
pressure is added to the blank holder force)
c) Initial sheet and liquid temperature are the same (275
o
C) Assumption: no
heat transfer between liquid and sheet. Figure 8.30 shows the pressure curve,
the blank holder force curve and the related process conditions used in the
analysis. Figure 8.31 shows the sheet temperature at the end of the drawing
process in the analysis. It is noticeable that there is no gap between the sheet
and the punch. Difference in sheet temperature using SHF-P and deep drawing
is clearly seen in Figure 8.32 and Figure 8.33, respectively.

150
Figure 8.30: Pressure and BHF used in the
simulations

Al 5754-O
Punch temperature: 60
o
C
Sheet temperature: 275
o
C
Die temperature: 275
o
C
Blank holder temperature: 275
o
C
Punch speed: 5 mm/s
Friction coefficient, = 0.04



Figure 8.31: Temperature distribution of the sheet at the end of the stroke


151

Figure 8.32: Temperature distribution of a sheet formed using SHF-P


Figure 8.33: Temperature distribution of a deep drawn sheet

152

8.4.1. Conclusions
a) Non-isothermal cup drawing simulations of magnesium AZ31-O have
been conducted using commercial FE software LS-Dyna. Difficulties in
the method of inputting the flow stress are encountered. At this point,
the need for inputting the flow stress as data points became obvious.
b) FEA of Al 5754-O cups were conducted using Deform 2D. Effect of the
heat transfer coefficient was found to be significant in the thickness
distribution of the cups. Heat transfer coefficients of ~11 kW/m
2
C and 6
kW/m
2
C were found to be better at the sheet-punch and sheet-tool
interfaces, respectively. An overall heat transfer coefficient of 6 kW/m
2
C
is also found to be satisfactory.

153

CHAPTER 9


DETERMINATION OF DRAWABILITY USING FRACTURE CRITERION

The LDR of a material is usually determined through experimentation,
which requires extensive resources. Therefore, there is a need for a quick and
approximate methodology to determine the LDR of a material by using the
available stress-strain curve and FEA. Thus, the objective is to predict an
approximate LDR (if possible) for aluminum (Al 6061-T6), with FEA by using
the flow stress (stress-strain curve) of this material and a fracture criterion.
9.1. Cockcroft & Latham ductile fracture criterion
The application of ductile fracture criteria for predicting the forming
limit of a sheet material has been studied by various researchers using
formability tests such as uniaxial tensile test, plane-strain tension test etc.
[Takuda, 1999 et al.] conducted FEA on deep drawing of various steel sheet by
using various ductile fracture criteria and compared the predictions with the
experimental findings. He concluded that Cockroft & Latham criterion could be
used effectively for various materials as a ductile fracture criterion.
Based on Takudas work, in this study, the Cockcroft & Latham criterion will be
used to determine the drawing limits of an aluminum alloy. By using this

154
criterion, the Critical Damage Values found using the Cockcroft & Latham
criterion, will be determined through FEA of uniaxial tension test and applied
in the deep drawing analysis. Cockcroft & Latham criterion is given in Eq.(1);

C d =

0
*
Eq. (1)
where,
*
is the maximum normal stress, is the equivalent stress, is the
fracture strain and C is a material constant called Critical Damage Value
(CDV).
9.2. Approach
For this purpose, hydraulic bulge tests were conducted to determine the
flow stress of the Al 5754-O at room temperature. The flow stress curve of Al
5754-O is given in Figure 9.1. Through additional drawing experiments, the
LDR of Al 5754-O was determined to be 2.1 using the designed deep drawing
tooling described in previous chapters. This available experimental data will be
used to obtain the Critical Damage Value (CDV) by FEA and will be applied in
the FEA of deep drawing Al 6061-T6 alloy. Same tool geometry will be used as
given in Figure 7.4.

155
Flow Stress Data from the Bulge test
0
50
100
150
200
250
300
350
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Strain
S
t
r
e
s
s

(
M
P
a
Bulge test

Figure 9.1: Flow stress curve of Al 5754-O obtained from the bulge test

The following steps will be carried out to validate the LDR/HDR for the Al
5754-O alloy and determine the LDR/HDR for the Al 6061-O alloy through FEA.
a) Conduct 3D FEA of uniaxial tension test using Al 5754-O to determine the
Critical Damage Value (CDV)
The CDV under uniaxial tensile state of stress, through modeling the
tensile test, will be determined by using the flow stress data given in Figure
9.1. The tensile specimen geometry used in the simulations is given in Figure
9.2.

156

A= 50 mm
(gage length)
W= 12.5 mm
R= 12.5 mm
L= 200 mm
C= 20 mm
Figure 9.2: Geometry of the tensile specimen (ASTM, E 8M-04)
b) Conduct FEA of deep drawing Al 5754-O
Through FEA it is necessary to show that a cup with LDR=2.1 can be
deep drawn for validation of the FE model. We will also check that in this FEA,
the normalized CDV obtained from step (a) corresponds to that determined in
this task. In other words, the objective in this task is to validate through FEA
the same experimentally determined LDR of 2.1.
c) Conduct 3D FEA of uniaxial tension test using Al 6061-T6 to determine the
CDV
The CDV will be determined under uniaxial state of stress through 3D
modeling of the tensile test by using the flow stress data given in Figure 9.3.
Flow stress data for the Al 6061-T6 is obtained from the literature
[Padmanabhan, 1997].
d) Conduct FEA to determine the LDR of Al 6061-T6 using the CDV from step
(c)

157
Various initial sheet diameters corresponding to various LDRs (such as
1.5 and 1.6) will be modeled to simulate the deep drawing operation for Al
6061-T6. In the simulations, drawn cup will be considered to fracture when the
normalized CDV calculated anywhere in the cup wall reaches the CDV
determined in step (c).
9.3. Setup of the FE model for tensile test and deep drawing
The tensile test specimen geometry is obtained from [ASTM, E 8M-04],
Figure 9.2. Flow stress curves given in Figure 9.3 are used both in modeling the
tensile test and the deep drawing analysis for Al 5754-O and the Al 6061-T6.

Figure 9.3: Flow stress curves used in the FEA

158
Both tensile test and deep drawing simulations are conducted at room
temperature (for Al 5754-O bulge test data and for Al 6061-T6 tensile data were
used). The friction coefficient in the deep drawing analysis is selected to be
0.04 based on previous experience. Due to the limitations in the blank holding
system of the press, blank holder force increased linearly from 4.4 kN to 8 kN
during the experiments. Therefore, in order to accurately conduct the analysis,
same linear increase is modeled in the FEA. Simulation matrix for Al 5754-O
and Al 6061-T6 is shown in Table 9.1.
Material LDR modeled in FEA
Al 5754-O 2.1 2.2
Al 6061-T6 1.6 1.5 1.4
Table 9.1: Simulation matrix for the deep drawing analysis

9.4. Results and discussion
9.4.1. Determining the Critical Damage Value
It is known that in uniaxial tension, load increases until the point when
necking starts and decreases post necking. Until necking, uniform elongation
takes place, therefore necking is usually considered to be the instability point.
This phenomenon is illustrated in Figure 9.4.

159
L
o
a
d

(
F
)
Stroke (s)
Uniform
elongation
Max. Load

Figure 9.4: Point of instability in the load-stroke curve during the tensile test
In the FEA of uniaxial tension, the CDV at the instability point (i.e. at the
point when necking starts) will be taken as the CDV for the analyzed material.
Based on this, Critical Damage Values of 0.428 and 0.148 are obtained for Al
5754-O and Al 6061-T6, respectively (Figure 9.5 and Figure 9.6). The length of
stroke before necking starts (i.e. at the end of uniform elongation) is found to be
18 mm for Al 5754-O and 5 mm Al 6061-T6. It is clearly seen that, the
deformation of Al 6061-T6 is noticeably low.

160

Figure 9.5: CDV (when necking starts
during 3D tensile test simulation) for
Al 5754-O is 0.428
Figure 9.6: CDV (when necking starts
during 3D tensile test simulation) for
Al 6061-T6 is 0.148

9.4.2. Deep drawing analyses
Initially, the deep drawing analysis is conducted for the Al 5754-O
(LDR=2.1) for checking whether the drawing will be successful and the
obtained CDV from the drawing analysis will be under 0.428. Figure 9.7 shows
that for Al 5754-O the draw is successful and the maximum CDV is 0.352 and it
is located at the cup bottom corner. Based on this result, a secondary deep

161
drawing analysis is conducted using a LDR of 2.2 for extra validation because it
was not possible experimentally to draw a LDR 2.2 cup. Figure 9.8 shows that
the obtained maximum CDV is 0.42 and the location is again at the cup bottom
corner. The cup reaches the CDV (for LDR of 2.2) at the early stage of the draw
(17 mm). Further drawing of this cup showed excessive thinning which
indicates fracture at the cup bottom corner. These results show that the
adopted methodology for determining the LDR works well, based on the
validation of Al 5754-O.


Figure 9.7: Maximum CDV obtained
for the successfully deep drawn Al
5754-O (LDR: 2.1) cup is 0.352
Figure 9.8: Max CDV reaches 0.42 for
the unsuccessfully deep drawn Al
5754-O (LDR: 2.2) cup

Same methodology used in the analysis of Al 5754-O is applied to Al 6061-T6.
Figure 9.9 and Figure 9.10 show that the cup reaches the Critical Damage
Values of 0.175 and 0.165 at around the 3 mm stroke for LDRs of 1.6 and 1.5

162
respectively and both values are higher than 0.148. LDR of 1.4 was also tried
and the CDV was found to be around 0.16. Results show that the sheet fractures
during the bending that takes place around the punch corner. Therefore, it is
concluded that for the existing tool dimensions and process conditions it is not
possible to draw Al 6061-T6. It is believed that this result makes sense because
for this material the maximum strain at fracture obtained from the tensile test is
only 0.1. On the other hand the maximum strain at fracture (in tensile test) for
Al 5754-O is around 0.3. It is also known that Al 6061-T6 has one of the lowest
formability due to its alloying and heat treatment conditions.
As a preliminary study with aluminum alloys, an idea about the
drawability of an alloy can be obtained using the methodology explained
above. Further experimentation will be beneficial with different materials and
tool geometries in order to establish the methodology.

163



Figure 9.9: Maximum CDV obtained for
Al 6061-T6 (LDR: 1.6) is 0.175
Figure 9.10: Maximum CDV
obtained for Al 6061-T6 (LDR: 1.5)
is 0.165

9.5. Conclusions
Following conclusions can be drawn from this study;
c) The adopted methodology for determining the LDR for a given material
by using the flow stress data and FEA is tested and found out to be
working satisfactorily.
d) With the existing tooling, it is not possible to draw Al 6061-T6 to any
degree, due to early fracture at the cup bottom corner. It is believed that
this result makes sense because for this material the maximum strain at

164
fracture obtained from the tensile test is only 0.1. On the other hand the
maximum strain at fracture (in tensile test) for Al 5754-O is around 0.3.

165
CHAPTER 10


SUMMARY AND CONCLUSIONS

Aluminum and magnesium components are being increasingly used by
major automotive and electronics companies. Current automotive magnesium
applications are mostly die castings, such as instrument panel beam, transfer
case, steering components and radiator support. Some aluminum sheet
applications can be seen in body panels of cars. The average magnesium
content in a typical 2003 model year family sedan built in North America was
only about 0.3% of the total vehicle weight, which compares to about 10% for
aluminum and 8% for polymers and polymer-based composites
[www.amm.com].
Aluminum and especially wrought magnesium alloys and their
manufacturing processes are receiving increasing attention. As especially
magnesium is expanding from interior components to more critical
applications in chassis and body areas, there is a great need for developing
wrought magnesium products and manufacturing processes to provide
improved mechanical and physical properties, crash performance and
corrosion resistance. In the present study, formability limits of aluminum and
magnesium alloy sheet are investigated through a) mechanical property

166
determination tests at room and at elevated temperatures (hydraulic bulging)
and b) deep drawing processes conducted at room temperature, against liquid
pressure (SHF-P) and at elevated temperatures. Major accomplishments can be
listed as;
a) Elevated temperature mechanical properties of Mg alloy sheet under equi-
biaxial state of stress were determined using the submerged hydraulic
bulge test. This concept is found to be best in obtaining close to uniform
temperature distribution in the sheet. True stress - true strain curves were
obtained using the membrane theory. Strain values up to 0.7 were obtained
under equi-biaxial state.
b) FE simulations illustrated that SHF-P process improved the drawability of
Al 5754-O from LDR of 2.1 (room temperature deep drawing) to LDR of 2.4.
This was obtained by the application of minimum pot pressure during the
critical stroke and increasing the pressure during the rest of the stroke.
c) Table 10.1 summarizes the improvement in the drawability of aluminum
and magnesium when warm forming process (with heated tools) is used.

167

Manufacturing Process
Material Deep Drawing (RT) SHF-P (RT) Warm Forming
Al 5754-O 2.1 2.4 2.9
Mg AZ31-O N/A N/A 3.2
Table 10.1: Summary of obtained drawability through the use
of different manufacturing processes

d) Due to the high thermal conductivity and low heat capacity of Al and
Mg, it is shown that through the use of a servo motor driven press, the
in-die-dwelling concept is applicable. With this method, the need for a
furnace and a transfer system is eliminated.
e) For obtaining a sound product, the maximum punch velocity was
decreased with increasing initial sheet diameter. Punch velocities of 35
mm/s and 300 mm/s were obtained for aluminum and magnesium
respectively. Up to 50 mm/s forming velocity, it is seen that temperature
improved the thinning behavior of aluminum significantly. However,
the same behavior was not seen in magnesium.
f) A foil type product was used as lubricant in warm forming and was
found to perform well. It provided scratch free cups, cleaner tools and a
smoke-free experimental environment.

168
g) Computational models of deep drawing and SHF-P at room and at
elevated temperatures (non-isothermal) were developed successfully.
Interface properties (heat transfer and friction coefficients) were
obtained through experimentation and inverse optimization methods.
Heat transfer coefficient was modeled as function of pressure and an
improvement in thickness distributions was obtained.
h) A methodology was used to determine the drawability limits of
aluminum alloys at room temperature through the use of Cockcroft &
Latham ductile fracture criterion. The validity of this approach was
demonstrated by computational modeling and experimental validation.
It was demonstrated that in some cases drawing experiments can be
eliminated and an approximate LDR value for a given material can be
estimated, using material properties and deep drawing conditions.

169

BIBLIOGRAPHY



[Agnew, et al, 2001] Agnew, S.R., M. H. Yoo, and C. N. Tom (2001)
Application of Texture Simulation to Understanding
Mechanical Behavior of Mg and Solid Solution Alloys
Containing Li or Y, Acta Mater, 49, 4277-4289.

[Agnew, et al, 2002] Agnew, S.R., D.W. Brown, S.C. Vogel, and T.M. Holden,
(2002) In-situ Measurement of Internal Strain Evolution
During Deformation Dominated by Mechanical
Twinning, in Proc. 6th Euro. Conf. Residual Stress,
(Trans Tech: Switzerland) pp. 747-754

[Altan et al 1983] Altan, T., Oh, S.I., Gegel, L.H., Metalforming -
Fundamentals and Applications, ASM International,
1983

[ASTM, E 8M-04] Standard test methods for tension testing of metallic
materials (metric), ASTM International, ASTM, E 8M-
04.

[Avedesian et.al.,
1999]
Avedesian, M., Buker, H., Magnesium and magnesium
alloys, ASM International, 1999.

[Ayres, 1977]
R.A. Ayres, Enhanced ductility in an Al-4%Mg alloy at
elevated temperature, Met. Trans., 8A (1977), 487-492.






170
[Barlat, et.al., 2003] F. Barlat, J.C. Brem, J.W. Yoon, K. Chung, R.E. Dick,
D.J. Lege, F. Pourboghrat, S.-H. Choi, E. Chu, Int. J.
Plast. 19 (2003) 1297.

[Bolt et al., 2001a] Bolt, P. J., Lamboo, N. A. P. M., Rozier, P. J. C. M.,
Feasibility of warm drawing of aluminum products,
Journal of materials processing and technology, Vol 115,
2001, pp 118-121.

[Bolt et al., 2001b] Bolt, P. J., Werkhoven, R. J., van de Boogaard, A. H.,
Effect of elevated temperatures on the drawability of
aluminum sheet components, Proceedings of
ESAFORM, 2001, pp 769-772.

[Boogaard ,2001] Boogard, A.H. van den, Thermally Enhanced Forming
of aluminum Sheet, Ph.D. Dissertation, University of
Twente, 2001.

[Boogaard et al.
2001a]
van de Boogaard, A. H., Bolt, P. J., Werkhoven, R. J.,
Aluminum sheet forming at elevated temperatures,
Simulation of Materials Processing: Theory, Methods
and Applications, 2001, pp. 819-824.

[Boogaard et al.
2001b]
van de Boogaard, A. H., Bolt, P. J., Werkhoven, R. J., A
material model for aluminum sheet forming at elevated
temperatures, Proceedings of ESAFORM, 2001, pp 309-
312.

[Contri et.al., 2004] Carlo Contri, Serhat Kaya, (2004), Prediction of Process
Parameters in Hydromechanical Deep Drawing using
Computer Simulation, ERC/NSM Report#:
SHF/ERC/NSM-04-R-29A, The Ohio State University

171

[Brand, 1998] Brand, A. J., (1998), Modellierung der
Gefuegeentwicklung bei der Warmumformung von
Aluminiumlegierungen mit Hilfe phaenomenologischer
und metallphysikalischer Ansaetze, Disseration, RWTH
Aachen, March 1998.

[Doege et al. 2001] Doege, E., Droder, K., Sheet metal formability of
magnesium wrought alloys-formability and process
technology, Journal of Materials Processing Technology
(115) 2001, pp 14-19.

[Doege, 2001b] Doege, E., Kurz, G., 2001, Development of a
Formulation to Describe the Work Softening Behavior of
Magnesium Sheets for Heated Deep Drawing Processes,
Annals of the CIRP, Vol. 50, 1, 2001

[Dow 1962] Forming Magnesium, The Dow Metal Products
Company, Division of The Dow Chemical Company,
Midland, Michigan

[Droeder, 1999] Droeder, K. G., (1999), Untersuchungen zum
Umformen von Feinblechen aus
Magnesiumknetlegierungen, Dr.-Ing. Dissertation,
Hannover, December 1999

[Duygulu et.al.
2005]
Duygulu, O, Agnew. S, 2005, Plastic Anisotropy and
Role of Non-basal Slip in Magnesium Alloy AZ31B,
International Journal of Plasticity, 21, pp.1161-1193



172
[Fenn, et al, 1961] Fenn, R.W., (1961), Evaluation of Test Variables in
Determination of Elevated-Temperature Compressive
Yield Strength of Magnesium Alloy Sheet, Symposium
on Elevated-Temperature Compression Testing of Sheet
Materials, ASTM Special Technical Publication No.303,
Sixty-fourth Annual Meeting, Atlantic City, N.J., June
29, 1961

[Gologranc, 1975] Gologranc, F., (1975), Beitrag zur Ermittlung von
Fliesskurven im kontinuierlichen hydraulischen
Tiefungsversuch, (Evaluation of the flow stress curve
with the continuous hydraulic bulge test), dissertation,
Institute for Metal Forming, University of Stuttgart,
Germany

[Groche, 2001] Groche, P. Director of the Institute for Production and
Forming, University of Darmstadt, Exchange of
technical information with Altan T.

[Groche, et al, 2002] Groche, P., Huber, R., Doerr, J., Schmoeckel, D., (2002),
Hydromechanical Deep-Drawing of Aluminum-Alloys
at Elevated Temperatures, CIRP Annals, 51/1, p.215

[Hill, 1950]

Hill, R., (1950) A Theory of Plastic Bulging of a Metal
Diaphragm by Lateral Pressure Phil.Mag. Ser. 7, 41,
pp.1133-1142

[Incropera, et.al.,
2002]
Incropera, F.P., Dewitt, P.D., (2002), Fundamentals of
Heat and Mass Transfer, 5th ed., John Wiley & Sons,
Inc., ISBN: 978-0-471-38650-6

[Jain, 2005] A. Jain, (2005), M.S. Thesis, University of Virginia

173

[Kaya et.al., 2004] Serhat Kaya, Florian Schnupp, Alexander Mattes and
Taylan Altan, (2004), Tool Design Strategies for
Hydromechanical Deep Drawing, ERC/NSM Report#:
SHF-ERC/NSM-04-R-06-B, The Ohio State University

[Kaya et.al., 2007] Serhat Kaya, Taylan Altan, (2007), Experimental and
Numerical Investigation of Non-Isothermal Deep
Drawing of Aluminum and Magnesium Sheet, CPF-
1.1/07/08, The Ohio State University, Columbus, Ohio

[Kim, et.al., 2001] Kim, H.K., Oh, S.I., (2001), Evaluation of heat transfer
coefficient during heat treatment by inverse analysis,
Journal of Materials Processing Technology, 112,
pp.157-165

[Klimanek, et al,
2002]
Klimanek, P., Potzsch, A., (2002), Microstructure
evolution under compressive plastic deformation of
magnesium at different temperatures and strain rates,
Materials Science and engineering A, vol.324, issue:1-2,
Feb 15 2002, pp.145-150

[Lange, K., 1985] [Lange, K., 1985, Handbook of Metal Forming, Society
of Manufacturing Engineers

[Li et al. 2000] Li, D., Ripson, P., Ghosh, A. K., warm forming of
aluminum alloys, DoE- LEMS Project Quarterly
Reports, 2000.

[Marciniak et.al.,
2002]
[Marciniak, Z., Duncan, J.L., Hu, S.J., 2002, Mechanics
of Sheet Metal Forming Butterworth-Heinemann

174

[Meiler, M., et al,
2004]
Meiler, M., Pfestorf, M., Merklein, M. and Geiger, M.
(2004), Tribolgogical Properties of Dry Film Lubricants
in Aluminum Sheet Metal Forming, Proceedings of the
2nd ICTMP, Lyngby, Denmark, 2004

[Meiler, M., et al,
2005]
Meiler, M. and Jaschke, H. (2003), Lubrication of
Aluminum Sheet Metal within the Automotive
Industry, Advanced Materials Research, Vol. 6-8, pp.
551-558

[Naka, et al., 2001] Naka, T., Torikai, G., Hino, R., Yoshida, F., The effects
of temperature and forming speed on the forming limit
diagram for type 5083 aluminum-magnesium alloy
sheet, Journal of materials processing and technology,
Vol. 113, 2001, pp 648-653.

[Naka, et.al., 2003] Naka, T., Nakayama, Y., Uemori, T., Hino, R., Yoshida,
F., (2003), Effects of temperature on yield locus for
5083 aluminum alloy sheet, Journal of Materials
Processing Technology, 140, pp.494-499

[Padmanabhan,
1997]
Padmanabhan, M., (1997), Wrinkling and springback in
electromagnetic sheet metal forming and
electromagnetic ring compression, Thesis (M.Sc),
OCLC#: 36978823, The Ohio State University,
Columbus, Ohio

[Panknin, 1959] Panknin, W., (1959), Der Hydraulische
Tiefungsversuch und die Ermittlung von Fliesskurven,
(The hydraulic bulge test and the determination of the
flow stress curves), dissertation, Institute for Metal
Forming, University of Stuttgart, Germany

175

[Pfestorf, M., 2002] Pfestorf, M., (2002) Influence of Dry Film Lubricants
and Surface Structure on the Forming Behavior of
Aluminum Sheets BMW Group, Mnchen, Germany

[Sabaie et al., 1973] Sabaie, M.G., Mellor, P.B., Plastic Instability Conditions
when Deep-drawing into a High Pressure Medium,
International Journal of Mechanical Sciences 15 (1973),
485-501

[Semiatin, et.al.,
1987]
S.L., Collings, E.W., Wood, V.E., Altan, T., (1987),
Determination of the Interface Heat Transfer
Coefficient for Non-Isothermal Bulk-Forming Processes,
Journal of engineering for Industry, February 1987,
vol.109, pp.49-57

[Shehta, et al 1978] Shehta, F., Painter, M. J., Pearce, R., Warm forming of
aluminum/magnesium alloy sheet, Journal of
mechanical working technology, Vol 2, 1978, pp279-290

[Siegert, 2001] Siegert, K., Vulcan, M., (2001), Superplastic Forming of
Sheet Metals, Innovations in Processing and
Manufacturing of Sheet Materials, The TMS Materials
Processing and Manufacturing Division (MPMD), TMS
Annual Meeting, Feb.11-15 2001, New Orleans, Luisiana

[Siegert, et al, 2003] Siegert, K., Jaeger, S., (2003), Pneumatic Bulging of
Magnesium AZ31 Sheet Metals at Elevated
Temperatures, CIRP, vol.52. pp.241-244



176
[Siegert, et al, 2004] Siegert, K., Jaeger, S., (2004), Pneumatic Bulging of
Magnesium AZ31 Sheet Metal at Elevated
Temperatures, TMS, The Minerals, Metals & Materials
Society), Magnesium Technology 2004, pp.87-90

[Sivakumar et.al.,
2006]
Sivakumar, R., Aue-u-lan, Y., Kaya, S., Spampinato, G.,
Altan, T., (2006), Flow stress database for aluminum
and magnesium alloy for warm sheet forming, Report
No: ERC/NSM-06-R-10, Center for Precision Forming,
The Ohio State University, Columbus, Ohio

[Takuda, 1999 et al.] Takuda, H., Mori, K., Hatta, N., (1999), The application
of some criteria for ductile fracture to the prediction of
the forming limit of sheet metals, Journal of materials
processing technology, 95, pp.116-121

[Takuda et.al., 2005] Takuda, H., Morishita, T., Kinoshita, T., Shirakawa, N.,
(2005), Modelling of formula for flow stress of a
magnesium alloy AZ31 sheet at elevated temperatures,
Journal of Materials Processing Technology, 164-165,
pp.1258-1262

[Taylor et al. 1980] Taylor, B., Lanning, H.W., Warm forming of aluminum
production systems, 25th National SAMPE
symposium and Exhibition, 1980, pp 471-480.

[Taylor et. al, 1976]
B. Taylor, R. A. Heimbuch and S. G. Babcock, Warm
forming of aluminum, Proc. 2
nd
International
Conference on Mechanical Behavior of Materials,
Federation of Materials Societies (Boston, MA), 1976,
2004-2008.





177
[Thiruvarudchelvan
et al., 2003]
Thiruvarudchelvan S., Travis F.W., Hydraulic-Pressure-
Enhanced Cup-Drawing Processes-an Appraisal,
Journal of Materials Processing Technology 140 (2003)
70-75

[Wagner, S., et al,
2002]
Wagner, S., Kleinert, H. and Zimmermann, R., Dry Film
Lubricants for Sheet Metal Forming, International
Conference, New Developments in Sheet Forming,
University Stuttgart, Germany, 2002

[Wu et al. 2001] Wu, X., Liu, Y., Wang, S., Superplastic Deformation
and Forming of Commercial Magnesium and Aluminum
Alloys With Initial Coarse-Grains, Proceedings of the
2002 NSF Design, Service and Manufacturing Grantees
and Research Conference, pg 2072- 2080.

[Yossifon et.al.,
1984]
Yossifon, S., Tirosh, J., 1984, On Suppression of
Plastic Buckling in Hydroforming Processes Int.J.
Mech. Sci. Vol.26, pp.389-402

[Yossifon et.al.,
1985a]
Yossifon, S., Tirosh, J., 1985, Rupture Instability in
Hydroforming Deep Drawing Process, Int.J. Mech. Sci.
Vol.27. No.9, pp.559-570

[Yossifon et.al.,
1985b]
Yossifon, S., Tirosh, J., 1985, Buckling Prevention by
Lateral Fluid Pressure in Deep Drawing, Int.J. Mech.
Sci. Vol.27, pp.177-185

[Yossifon et.al.,
1988]
Yossifon, S., Tirosh, J., 1988, On the permissible
Fluid-Pressure Path in Hydroforming Deep Drawing
Processes, Journal of Engineering for Industry. Vol.37,
pp.225-239

178

[Zhang et al., 2003] Zhang, S.H., Jensen, M.R., Nielsen, K.B., Danckert, J.,
Lang, L.H., Kang, D.C., Effect of Anisotropy and
Prebulging on Hydromechanical Deep Drawing of Mild
Steel Cups, Journal of Materials Processing Technology
142 (2003), 544-550

www.amm.com American Metal Market, Metal Statistics, May 2004.




179






APPENDIX A

MODELING AND EXPERIMENTAL INVESTIGATION OF SHEET

HYDROFORMING WITH DIE (SHF-D)

180
The objectives of this study are to :
a) determine the flow stress of Al 3003-H14, 3003-O, 5052-H32, 5052-O and
6061-O alloy sheets at room temperature using the hydraulic bulge test
b) conduct SHF-D experiments and Finite Element Analysis with 3003-O, 5052-
H32 and 6061-O using an asymmetric die and compare the results to
investigate the feasibility of this process.
1) Flow stress results from the hydraulic bulge test
The equation used to describe the flow stress is the law of Hollomon given as;
n
K =

where, K is a material specific constant factor, called strength coefficient and
n describes increasing hardening of the material with increasing strain and is
therefore called strain hardening exponent. Table A.1 shows all the results,
obtained from the hydraulic bulge tests.
Material

Thickness
(mm)
Max.Bulge
Height (mm)
Burst
Pressure
(bar)
Max thinning

(%)

K
(MPa)
n

R
2

3003-H14 1.2 26.8 41 30 200 0.08
0.83
3003-O 1.2 36.1 37 31 208 0.31 0.96
5052-H32 1.2 28 63 33 322 0.15 0.95
5052-O 1.2 27 62 30 323 0.17 0.98
6061-O 2.23 31 79 36 213 0.27 0.97
Table A.1: Calculated flow stress values and related process parameters


181
Table A.1 shows that 3003-O gives the highest bulge height of 36.1 mm
compared to all other alloy sheets. [Explanation of R
2
: Lets assume that we
have a simple regression model, y=a*x
1
+b*x
2
, where y is a dependent variable
and x
1
and x
2
are explanatory variables. The usual interpretation of R
2

(coefficient of determination) is as the relative amount of variance of the
dependent variable y, explained or accounted for by the explanatory variables
x1, x2,.. For example, if R
2
= 0.83 we say that the explanatory variables (x1, x2)
"explains" 83 % of the variance of y]. Figure A.1 shows the calculated flow
stress curves for the tested alloys.

Figure A.1: Flow stress curves of the tested aluminum alloys


182
2. SHF-D experiments and numerical predictions

2.1 Press and tooling characteristics
The experimental press was a 160 ton hydraulic Minster press. The
characteristics of the press are listed in [Table A.2].
Type Minster Tranemo DPA-160-10
Slide Force 160 metric tons (176.4 tons)
Cushion Force 100 metric tons (110.3 tons)
Ejector Force 15 metric tons (16.5 tons)
Slide Speed 90 mm/s maximum (3.54 in/s)
Total Daylight 800 mm (31.496)
Slide Stroke 500 mm (19.685)
Cushion Stroke 190 mm (7.480)
Ejector Stroke 250 mm (9.843)
T-Slots 5 slots, 6 apart (152.4 mm)
Table A.2: Press characteristics
Aluminum alloys 6061-O, 3003-O and 5052-H32 were tested and the initial
sheet geometry was a square sheet with a 241 mm side length. Thickness
measurements are conducted along 1-3 and 2-4 directions (Figure A.2).

183

Figure A.2: Thinning measurements are done in Direction 1-3 and Direction 2-4

2.2. Comparison of the experiments and the FEA Predictions
Results of the FEA and the experiments are compared for 6061-O, 3003-O and
5052-H32 alloy sheets. The validation of the sheet hydroforming interface
Aquadraw in the commercially available finite element software Pam-Stamp
2K is made through comparison. It is important to note the following;
a) The comparison criteria between the FEA and the experiments were 1)
maximum pressure values and 2) the thinning distributions across the
part corners.
b) In the simulations the material was considered to be isotropic.
1
3
2
4

184
c) The lubricant that was used in the experiments was a highly viscous
lubricant. The friction coefficient for this lubricant was not available
experimentally. Therefore a friction coefficient of 0.1 was assumed in
the FEA.
2.2.1. Aluminum alloy 6061-O
The different behavior in pressure curves is well seen in Figure A.3. In the
experiment, the viscous medium used for pressurization has a transitory phase
(until 4 Sec) in which it is compressed and for this reason the pressure doesnt
increase. After this point (> 4 Sec) the pressure starts to increase which means
the medium has been compressed. The final value is consistent with the one
predicted by the FEA. The predicted and measured thinning percentage curves
along 1-3 and 2-4 are shown in Figure A.4 and Figure A.5.

185

Figure A.3: Comparison of the predicted and measured pressure values for
6061-O

Figure A.4: Thinning comparison in Direction 1-3; at corner 1 the difference is
14 (%) while at corner 3 it is 11 (%)


186

Figure A.5: Thinning comparison in Direction 2-4;at corner 2 the difference is
10 (%) while in the corner 4 is 1.5 (%).

187

2.2.2. Aluminum alloy 3003-O


Figure A.6: Pressure comparison; max pressure for simulation is 199 (Bar)
while measured pressure value from the experiments is 208 (Bar)


188

Figure A.7: Thinning comparison in Direction 1-3; at corner 1 the difference is
2 (%) while at corner 3 it is 7 (%)

189



Figure A.8: Thinning comparison in Direction 2-4; at corner 2 the difference is
3 (%) while at corner 4 it is 2 (%)


190

2.2.3. Aluminum alloy 5052-H32


Figure A.9: Pressure comparison; max pressure for simulation is 192 (Bar)
while measured pressure value from the experiments is 218 (Bar)


191

Figure A.10: Thinning comparison in Direction 1-3; at corner 1 the difference is
4 (%) while at corner 3 it is 10 (%)




192







APPENDIX B
MODELING OF ELASTIC DEFLECTION IN DEEP DRAWING


193
In computational modeling of deep drawing process, the tools (blank
holder, die and punch) are generally assumed to be rigid. This saves
computation time since the calculations are pretty acceptable. However, when
assumed rigid these tools do not adjust to the thickness changes that take place
in the sheet during a drawing operation. The blank holder and the die touch
the sheet metal where the thickness is a maximum and apply pressure only at
these points where uniform pressure on the sheet metal is desired.
Many studies have shown that the combined rigidity of the dies and the
press has a great influence on the accuracy in precision blanking and deep
drawing. Elastic deflections of press and tooling are difficult to predict and
control. Thus, they affect the quality of the parts and are undesirable.
An elastically deforming blank holder and die that adjusts itself to the
sheet thickness can provide a uniform distribution of the blank holder pressure
at the flange area. This will result in uniform metal flow during deep drawing
and ultimately in improved quality of the parts. However, the amount of elastic
deflection is always an unknown. In this study, computational modeling of the
deep drawing process is conducted in which the die and the blank holder is
modeled elastic instead of rigid bodies in order to determine the maximum
amount and location of elastic deflection.

194
1. Modeling of sheet, punch, die-ring and blank holder
To predict the elastic deflection in the blank holder and the die-ring over
the whole stroke, die-ring and blankholder were modeled as elastic objects. The
sheet is modeled as plastic and the punch as rigid body. In order to obtain
higher accuracy and less computation time, variable meshing densities were
used at different parts of the die-ring and the blank holder.
Table B.1 gives an overview of the mesh density windows, which are shown in
Figure B.1 and Figure B.2, respectively.
Number of
elements
Density
window
#1
Density
window
#2
Density
window
#3
Blank holder
(Binder)
500 5 0 -
Die-ring 500 5 0 15
Table B.1: Mesh Density Windows (same values used in S1, S2 and S3)


Figure B.1: Mesh Density Windows for the blank holder
1
2

195

Figure B.2: Mesh Density Windows for the die-ring
2. Sheet and tool material
St14 steel is used as the sheet material and tool material was determined
to be H13. The elastic tool material data is described by using Hookes Law. To
describe the plastic behavior of St14, Law of Hollomon is used.
n
K =
with K (strength coefficient) and n (strain hardening exponent). Table B.2
shows the material properties for St14.
Material K n
St14 747 0.1007
Table B.2: Flow stress data of St14
Since experimental data was not available, after conducting couple of
simulations, it was found that 15 kN as blank holder force is an adequate
constant blank holder force value. The friction conditions on the sheet-blank
holder, sheet-die and sheet-punch, were modeled using the Coulomb friction
3
2
1

196
model with a value of 0.08. This value was selected based on previous
experience.
3. Results and discussion
Figure B.3 shows the elastic deflection of the die and the blank holder in
conventional deep drawing schematically. In Figure B.3, continuous (ABCDE)
line is the initial flat die geometry and the dashed line (AB
1
C
1
D
1
E) is the
elastically deflected die geometry. It is seen that AB is deflected in the negative
Y direction and DE is deflected in the positive X direction. At the die corner
(BCD), maximum deflection happens at CC
1
. Therefore, we can see that the
elastic deflection varies from A thru E. It should also be noted that deflection
also changes with stroke, H. The blank holder (binder) deflects in the positive Y
direction. Figure B.4 shows the punch load and the die ring deflection plots
calculated from the numerical analyses.

197

Figure B.3: Predicted elastic deflection is shown with the dashed line
0.0
20.0
40.0
60.0
80.0
100.0
120.0
140.0
160.0
180.0
0.1 7.5 15.0 22.5 30.0 37.5 45.0 52.5 60.0 67.5
Stroke [mm]
L
o
a
d

[
k
N
]
0
1
2
3
4
5
6
D
e
f
l
e
c
t
i
o
n

[

m
]
Punch-Load Die-Corner-Deflection
20 mm 40 mm 60 mm

Figure B.4: Predicted elastic deflection is shown with the dashed line [Kaya,
et.al., 2004]

198








APPENDIX C
WARM FORMING OF STAINLESS STEEL (SS) 304



199

1. Warm forming of austenitic steel
Among austenitic stainless steels, Type 304 is superior in formability
and is most commonly used in parts for household appliances. However, the
austenitic phase is unstable and gets transformed into martensite during
forming. This transformation of austenite to martensite is a function of strain,
strain rate (punch velocity) and temperature. Martensite enhances the strain
hardening, thus delaying the onset of necking in sheet metal. Delayed necking
is desirable for high formability/drawability. However, martensitic phase raises
the forming loads, reduces formability and decreases corrosion resistivity. For
further deformation, an annealing operation is required. These intermediate
annealing operation(s) slow down production and increase cost. To eliminate
the intermediate annealing operations and avoid martensitic transformation,
warm forming process of stainless steels is explored.
2. Blank dimensions
The 304 stainless steel sheet material used for the investigations was
provided by Elkay Manufacturing Company. A total of 80 round samples (20 for
each draw ratio) were used for the tests. (Table C.1)



200












Table C.1: Stainless steel blank dimensions used for the drawability
investigation (D
B
= blank diameter, D
P
= punch diameter)

3. Experimental procedure
The mechanical servo press was programmed in order to run at the
constant velocity of 2.5 mm/s (this is the value Elkay Manufacturing requested
us to perform the investigation) and the following procedure has been used to
determine the limiting draw ratio (LDR) at various die temperatures.
step 1) a smaller blank has been drawn at room temperature (Note: during
experiments, this BHF value has been increased every time the formed
cup showed wrinkles)
step 2) the blank size has been increased until a fractured cup was obtained
consistently. The last cup that was formed successfully is used to
determine the LDR at room temperature.
Material: SS 304 Thickness 0.87 mm (0.035 in)
2.4 96
2.5 100
2.3 92
2.1 84
DR (D
B
/D
P
) Diameter, D
B
[mm]

201
step 3) the initial temperature of the dies has been increased and the same
procedure described in step 2 has been applied. Table C.2 shows the
blank dimensions and the experimental conditions.
90 dwell time [s]
10 Blank Holder Force [ kN]
2.5 forming velocity [mm/s]
90 dwell time [s]
10 Blank Holder Force [ kN]
2.5 forming velocity [mm/s]

2.1 25 A
2.3 70 B
2.3 100 C
2.5 150 D
LDR
Dies/BH
Temperature [C]
Forming
condition
2.1 25 A
2.3 70 B
2.3 100 C
2.5 150 D
LDR
Dies/BH
Temperature [C]
Forming
condition

Table C.2: Blank dimensions and process conditions

4. Experimental results
Results show that the formability of SS304 increases when the
temperature of the dies is increased (up to 150 C) [Figure C.1]. When the
temperature was set at 150 C, it was possible to form the cup with a 2.5 draw
ratio consistently. The remaining blanks were later used to repeat the same
study with higher forming velocities.

202
2
2.1
2.2
2.3
2.4
2.5
2.6
0 25 50 75 100 125 150 175
A
B
C
D
2
2.1
2.2
2.3
2.4
2.5
2.6
0 25 50 75 100 125 150 175
A
B
C
D
Die and BH temperature [C]
L
i
m
i
t
i
n
g

D
r
a
w
i
n
g

R
a
t
i
o

Figure C.1: Limiting DR - Die/BH temperature (forming velocity 2.5 mm/s)
Cups showed in Figure C.2 were cut and the thickness distribution along the
rolling direction (RD) and the transverse direction (TD) were measured.

Figure C.2: Geometrical parameters (height and external diameter) of the cups
drawn at different conditions
DR: 2.5 DR: 2.4 DR: 2.3 DR: 2.1
Height: 45 mm Height: 40 mm Height: 40 mm Height: 32 mm
Diameter: 45 mm Diameter: 45 mm Diameter: 45 mm Diameter: 45 mm
Cup D Cup C Cup B Cup A
DR: 2.5 DR: 2.4 DR: 2.3 DR: 2.1
Height: 45 mm Height: 40 mm Height: 40 mm Height: 32 mm
Diameter: 45 mm Diameter: 45 mm Diameter: 45 mm Diameter: 45 mm
Cup D Cup C Cup B Cup A

203






Cup A
25 Dies/BH Temperature [C]
2.5 forming velocity [mm/s]
2.1 Draw ratio
Cup A
25 Dies/BH Temperature [C]
2.5 forming velocity [mm/s]
2.1 Draw ratio

Cup B
70 Dies/BH Temperature [C]
2.5 forming velocity [mm/s]
2.3 Draw ratio
Cup B
70 Dies/BH Temperature [C]
2.5 forming velocity [mm/s]
2.3 Draw ratio

-0.50
-0.40
-0.30
-0.20
-0.10
0.00
0.10
0.20
0.30
0.40
0 10 20 30 40 50 60
RD
TD
20 mm
50 mm
Curvilinear length [mm]
T
h
i
n
n
i
n
g
21 %
-0.50
-0.40
-0.30
-0.20
-0.10
0.00
0.10
0.20
0.30
0.40
0 10 20 30 40 50 60
RD
TD
20 mm
50 mm
Curvilinear length [mm]
T
h
i
n
n
i
n
g
21 %

-0.50
-0.40
-0.30
-0.20
-0.10
0.00
0.10
0.20
0.30
0.40
0 10 20 30 40 50 60
RD
TD
20 mm
55 mm
T
h
i
n
n
i
n
g

Curvilinear length [mm]
24 %

Figure C.3: Thinning distribution along rolling and transverse direction for cup A and
cup B
LDR
LDR

204

Cup C
100 Dies/BH Temperature [C]
2.5 forming velocity [mm/s]
2.4 Draw ratio
Cup C
100 Dies/BH Temperature [C]
2.5 forming velocity [mm/s]
2.4 Draw ratio

Cup D
150 Dies/BH Temperature [C]
2.5 forming velocity [mm/s]
2.5 Draw ratio
Cup D
150 Dies/BH Temperature [C]
2.5 forming velocity [mm/s]
2.5 Draw ratio

-0.50
-0.40
-0.30
-0.20
-0.10
0.00
0.10
0.20
0.30
0.40
0 10 20 30 40 50 60 70
RD
TD
20 mm
60 mm
Curvilinear length [mm]
T
h
i
n
n
i
n
g
29 %
-0.50
-0.40
-0.30
-0.20
-0.10
0.00
0.10
0.20
0.30
0.40
0 10 20 30 40 50 60 70
RD
TD
20 mm
60 mm
Curvilinear length [mm]
T
h
i
n
n
i
n
g
29 %

-0.40
-0.30
-0.20
-0.10
0.00
0.10
0.20
0.30
0.40
0 10 20 30 40 50 60 70
RD
TD
20 mm
65 mm
T
h
i
n
n
i
n
g

Curvilinear length [mm]
39 %
-0.40
-0.30
-0.20
-0.10
0.00
0.10
0.20
0.30
0.40
0 10 20 30 40 50 60 70
RD
TD
20 mm
65 mm
T
h
i
n
n
i
n
g

Curvilinear length [mm]
39 %
Figure C.4: Thinning distribution along rolling and transverse direction for cup
C and cup D
From the graphs given in Figure C.3 and Figure C.4 it can be concluded that;
The thinning is always maximum at the cup bottom corner location and
this value increases when the draw ratio increases.
The 2.5 LDR cup showed a maximum thinning of almost 40%; (this
extremely high value of thinning, depending on the product applications,
could be considered unacceptable).
Remaining sheets have been used to investigate the formability limits of
the material at higher velocities, more close to the production levels. The
LDR LDR

205
forming velocities that were selected are 25 mm/s and 50 mm/s. The results of
this further investigation are summarized in Figure C.5 and Figure C.6.
2
2.1
2.2
2.3
2.4
2.5
2.6
0 25 50 75 100 125 150 175 200 225 250
Die and BH temperature [C]
L
i
m
i
t
i
n
g

D
r
a
w

R
a
t
i
o

Forming velocity: 25 mm/s
: formed : fractured
2
2.1
2.2
2.3
2.4
2.5
2.6
0 25 50 75 100 125 150 175 200 225 250
Die and BH temperature [C]
L
i
m
i
t
i
n
g

D
r
a
w

R
a
t
i
o

Forming velocity: 25 mm/s
: formed : fractured

Figure C.5: Limiting DR - Die/BH temperature (forming velocity 25 mm/s)

206
2
2.1
2.2
2.3
2.4
2.5
2.6
0 25 50 75 100 125 150 175 200 225 250
Die and BH temperature [C]
L
i
m
i
t
i
n
g

D
r
a
w

R
a
t
i
o

Forming velocity: 50 mm/s
: formed : fractured
2
2.1
2.2
2.3
2.4
2.5
2.6
0 25 50 75 100 125 150 175 200 225 250
Die and BH temperature [C]
L
i
m
i
t
i
n
g

D
r
a
w

R
a
t
i
o

Forming velocity: 50 mm/s
: formed : fractured

Figure C.6: Limiting DR - Die/BH temperature (forming velocity 50 mm/s)
5. Summary and conclusions
The effect of the forming temperature on the drawability of four different
blank sizes (DR 2.1, 2.3, 2.4 and 2.5) from SS 304 was studied at three
different forming velocities (2.5 mm/s, 25 mm/s and 50 mm/s).
For each velocity, the formability window as functions of LDR and
forming temperature was determined.
Higher forming temperatures (up to 150 C) and lower forming velocity
have increased the drawability of SS 304.
Forming temperature range for SS 304 is found to be lower (up to 150 C)
than the temperature ranges observed for Al and Mg alloy sheets.

207
Compared to the elevated temperature deep drawing experiments that
were conducted using Al and Mg alloy sheets, the max thinning was not
always observed at the bottom corner location of the cup, for SS304.
Maximum thinning is observed at the bottom corner of the cup.

You might also like