Vous êtes sur la page 1sur 237

DESIGN FOR UNCERTAINTIES OF SHEET METAL

FORMING PROCESS

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

Wenfeng Zhang, M.S.

****

The Ohio State University

2007

Dissertation Committee: Approved By

Dr. Rajiv Shivpuri, Adviser

Dr. Allen Yi Adviser


Graduate Program in
Dr. Tunc Aldemir Industrial and Systems Engineering
ABSTRACT

The exclusion of inherent process variations in the current deterministic design methods

for sheet metal forming can lead to very unreliable result that may cause high scrap rate,

frequent rework, machine shut down and thus huge loss of profit. Extensive research has

been done in exploring the deterministic effect of each factor on the part, but the impact

of their variations on the fluctuation of output quality for the stamping process is seldom

addressed and quantified. The inclusion of uncertainty in the design and optimization

cycle should lead to better understanding of the impact of uncertainty associated with

system input on the system output. This understanding can then be applied for managing

such uncertainties. To date, there are very few reports on incorporating uncertainties and

variations in the design of stamping processes.

In this research we propose three different probabilistic design approaches. It uses sheet

metal forming finite element method (FEM) simulation as the fundamental tool. When

the system meta-model is not complex, the design of experiments (DOE) technique and

response surface method (RSM) are integrated with FEM to build an explicit function to

connect process input to process performance output. With the quantification of

uncertainties of input variables, by Monte Carlo Simulation (MCS) or other reliability

analysis methods, the probability that the product conforms to its specification can

ii
therefore be assessed. Through the right formulation of probabilistic optimization the

robust optimal design configuration can be found. In the case where the number of

process inputs is too many so that the design of experiment cannot handle and the system

meta-model cannot be approximated, the alternate approach that integrates FEM,

multi-objective genetic algorithm and reliability assessment is illustrated.

By the proposed probabilistic design approaches, deeper understanding about the

relationship between process variables uncertainties with the part qualities is achieved.

Ultimately, the process output variability could be reduced, defect rates could be

minimized and product quality would be improved.

iii
DEDICATION

Dedicated to my parents: Zhongping Zhang/Xiaohui Shu and my wife Ji Li.

iv
ACKNOWLEDGMENTS

First and foremost I would like to thank Dr. Rajiv Shivpuri, my adviser, for his support,

guidance and above all, encouragement throughout the years of my PhD program. I have

learned knowledge and the way to obtain knowledge under his instructions that has

prepared me to make more contributions in the future.

I thank other members of my dissertation committee, Dr. Allen Yi and Dr. Tunc Aldemir

for their scientific inputs and advice, as well as the member of my general exam

committee, Dr. Steve MacEachern and Dr. Tedd Allen for their valuable comments and

suggestions.

I also own thanks to our group members, Lin Yang, Yijun Zhu, Yongning Mao, Xiaomin

Cheng, Yuanjie Wu, Meixing Ji and Chun Liu, for their academic help and friendship.

Thanks also go to Dr. Jiang Hua, Dr. Ziqiang Sheng, and Dr. Satish Kini for their kindly

supports and helpful suggestions.

Finally, I want to express my sincere gratitude to my wife, Ji Li for her longtime support

and unconditional love! She always makes me smile even facing the greatest difficulty. I

am so glad that she will get her PhD degree the same time with me next month! We are

going to be a proud and happy dual-PhD family!


v
VITA

May 16, 1978 Born - Wuhan, P.R. China

2000 B.E., Mechanical Engineering, Huazhong Univ. of


Science and Technology, Wuhan, P.R. China

2001 – 2003 M.S., Mechanical Engineering, the Ohio State


University, Columbus, Ohio, U.S.A.

2001 – 2007 Graduate Research Associate, Department of


Industrial, Welding and Systems Engineering, the
Ohio State University

PUBLICATIONS

Zhang, W., Sheng, Z., Shivpuri, R., Probabilistic Design of Aluminum Sheet Drawing for
Reduced Risk of Wrinkling and Fracture, Numisheet (Detroit), 2005, pp.247-252.

Zhang, W., Shivpuri, R., Investigating Reliability of Variable Blank Holder Force Control
in Sheet Drawing under Process Uncertainties, ASME, Journal of Manufacturing Science
and Engineering, accepted, 2006.

Zhang, W., Shivpuri, R., A new Discrete Friction Concept in Sheet Metal Deep Drawing
and Its Process Optimization by Multi-objective Optimization, ASME, Journal of
Manufacturing Science and Engineering, accepted, 2006.

vi
FIELDS OF STUDY

Major Field: Industrial & Systems Engineering

Major Area: Manufacturing

Minor I: Reliability Engineering

Minor II: Statistical Quality Control

vii
TABLE OF CONTENTS

Page

ABSTRACT....................................................................................................................... ii
DEDICATION.................................................................................................................. iv
ACKNOWLEDGMENTS ................................................................................................ v
VITA.................................................................................................................................. vi
LIST OF TABLES........................................................................................................... xii
LIST OF FIGURES ....................................................................................................... xiii

CHAPTER
1. INTRODUCTION......................................................................................................... 1
1.1 Traditional Approach to Process Design .................................................................. 4
1.2 New Approach for Robust Process Design............................................................... 6
1.3 Objective of Research ............................................................................................... 8
1.4 Research Significance and Benefits.......................................................................... 9
1.5 What is My Contribution? ...................................................................................... 10
1.6 Dissertation Outline ................................................................................................ 11

2. BACKGROUND AND LITERATURE REVIEW ................................................... 13


2.1 Finite Element Method and Its Application in Sheet Metal Forming..................... 15
2.1.1 Introduction to Sheet Metal Forming............................................................... 15
2.1.2 Basis of Finite Element Method ...................................................................... 17
2.1.3 Application of Finite Element Method in Sheet Metal Forming ..................... 18
2.2 Design of Experiments Techniques ........................................................................ 21
2.2.1 Full-factorial Design ........................................................................................ 21

viii
2.2.2 Orthogonal Arrays ........................................................................................... 21
2.2.3 Latin Hypercube Design .................................................................................. 22
2.2.4 Central Composite Design ............................................................................... 25
2.2.5 Box-Behnken Design ....................................................................................... 25
2.2.6 Computer Aided Designs................................................................................. 27
2.3 Approximating Methods ......................................................................................... 27
2.3.1 Response Surface Method................................................................................ 28
2.3.2 Kriging Meta Models....................................................................................... 30
2.3.3 Neural Networks .............................................................................................. 31
2.4 Literature Review of Deterministic Sheet Metal Forming Process Optimization
Employing FEM, DOE and Approximation Methods .................................................. 32

3. RESEARCH METHODOLOGY .............................................................................. 38


3. 1 Taguchi Robust Design.......................................................................................... 38
3.1.1 Introduction to Taguchi Robust Design........................................................... 38
3.1.2 General Taguchi Robust Design Optimization Formulation ........................... 43
3.1.3 Disadvantages of Taguchi Robust Design method .......................................... 44
3.2 Reliability Based Optimization............................................................................... 45
3.2.1 Reliability Analysis.......................................................................................... 45
3.2.2 Reliability Based Optimization Formulation ................................................... 51
3.3 Proposed Probabilistic Design approach ................................................................ 53

4. EXPERIMENT OBSERVATION ON THE PART QUALITY VARIATIONS ..... 56


4.1 Experiment Objective ............................................................................................. 56
4.2 Experimental Design............................................................................................... 58
4.2.1 Criteria to Select Right Part to Draw ............................................................... 58
4.2.2 Drawing Die Design ........................................................................................ 59
ix
4.2.3 The Measurement of Quality Characteristics .................................................. 63
4.2.4 The Experiment Procedure .............................................................................. 65
4.3 The Experiment Result ........................................................................................... 66
4.3.1 Wrinkling Measurement Result ....................................................................... 67
4.3.2 Fracture Measurement Result .......................................................................... 71
4.3.3 Springback Measurement Result ..................................................................... 74
4.4 Experiment Conclusion........................................................................................... 76

5. PROBABILISTIC DESIGN OF ALUMINUM SHEET DRAWING FOR


REDUCED RISK OF WRINKLING AND FRATURE............................................... 77
5.1 Introduction............................................................................................................. 78
5.2 General Approach for Probabilistic Design............................................................ 82
5.3 Quality Index to Measure Risk ............................................................................... 83
5.4 The Numerical Simulation Model .......................................................................... 85
5.5 Selection of Input Variables.................................................................................... 86
5.6 Prediction of Wrinkling and Fracture ..................................................................... 87
5.7 DOE and RSM ........................................................................................................ 91
5.8 Probabilistic Assessment for Wrinkling and Fracture ............................................ 95
5.9 Optimal Design Approach ...................................................................................... 96
5.10 Effect of Variation on The Optimum Design ..................................................... 103
5.11 Conclusion .......................................................................................................... 105

6. INVESTIGATING RELIABILITY OF TEMPORAL VARIABLE BLANK


HOLDER FORCE CONTROL IN SHEET DRAWING UNDER PROCESS
UNCERTAINTIES........................................................................................................ 106
6.1 Introduction........................................................................................................... 107
6.2 General Approach for Probabilistic Design Optimization.................................... 110
x
6.3 Parameterization of The Variable Blank Holder Force ........................................ 114
6.4 The numerical Simulation Model ......................................................................... 117
6.5 Selection of Input Variables and Output Variables .............................................. 120
6.6 Deterministic Design vs. Probabilistic Design ..................................................... 122
6.7 Conclusion ............................................................................................................ 134

7. SPATIALLY VARYING CONSTRAINTS AND PROBABILISTIC DESIGN BY


MULTI-OBJECTIVE GENETIC ALGORITHM..................................................... 138
7.1 Introduction........................................................................................................... 140
7.2 Setup of Spatially Varying Constraints in Hishida Part Drawing......................... 149
7.3 Multi-Objective Genetic Algorithm: NSGA-II..................................................... 153
7.3.1 Pareto optimal and Pareto front ..................................................................... 155
7.3.2 Non-dominated Sorting Genetic Algorithm-II............................................... 157
7.4 Design Optimization Model.................................................................................. 161
7.4.1 Design Objectives .......................................................................................... 161
7.4.2 Design Variables............................................................................................ 165
7.4.3 Integration of NSGA-II Optimization and Pam-Quickstamp Full FEA ........ 165
7.4.4 Hishida Part Drawing Process Heuristics ...................................................... 166
7.5 Optimization Result and Analysis ........................................................................ 167
7.6 The Probabilistic Design Search........................................................................... 174
7.7 Conclusion ............................................................................................................ 180

8. CONCLUSIONS AND FUTURE WORK .............................................................. 182


LIST OF REFERENCES ............................................................................................. 187
APPENDIX A ................................................................................................................ 193
APPENDIX B ................................................................................................................ 218

xi
LIST OF TABLES

Page

Table 4.1. The die/punch design and test setup ................................................................ 62

Table 5.1. Variation of material properties [28][29]......................................................... 81

Table 5.2, The variation and DOE levels for selected parameters.................................... 90

Table 5.3. Deterministic design result ............................................................................ 101

Table 5.4. Probabilistic design result.............................................................................. 102

Table 6.1. Circular blank dimensions, material properties, and friction coefficients used
in the simulation of drawing the conical cup from AKDQ steel. ........................... 119

Table 6.2. The sensitivity of initial selected process parameter. .................................... 121

Table 6.3. The deterministic and probabilistic design result with constraint reliability. 130

Table 7.1. Material properties used in the simulation..................................................... 164

Table 8.1. The summary of prob. & determ. design strategies....................................... 184

xii
LIST OF FIGURES

Page

Figure 2.1. Latin Hypercube design [11] .......................................................................... 24

Figure 2.2. Comparison of CCD design and Box-Behnken design. ................................. 26

Figure 3.1. Taguchi robust design matrix [11]. ................................................................ 41

Figure 3.2. Taguchi loss function [11].............................................................................. 42

Figure 3.3. Reliability based analysis [11]........................................................................ 46

Figure 3.4. FORM reliability analysis method [11].......................................................... 47

Figure 4.1. Illustration of process output variations caused by input variations for sheet
drawing process. ....................................................................................................... 57

Figure 4.2. The cylindrical cup drawing die design.......................................................... 60

Figure 4.3. The hydraulic press used in the drawing test.................................................. 61

Figure 4.4. Illustration of measurement of three quality characteristics of drawn cups... 64

Figure 4.5. Parts drawn without and with lubricants at two levels of BHFs. ................... 68

Figure 4.6. The parts after drawing and the defective parts.............................................. 69

Figure 4.7. The measurement result for number of wrinkling along the flange. .............. 70

Figure 4.8. The result for the maximum height of wrinkling mark along cup sidewall. .. 72

Figure 4.9. The fracture measurement result. ................................................................... 73

Figure 4.10. The springback angle measurement result. .................................................. 75

Figure 5.1. General probabilistic design approach. .......................................................... 88

xiii
Figure 5.2. Simulation model of the Hishida forming process. ........................................ 89

Figure 5.3. The measurement of sidewall wrinkling. ....................................................... 93

Figure 5.4. Sensitivity plot of wrinkling and thinning...................................................... 94

Figure 5.5. Probabilistic design of BHF. .......................................................................... 99

Figure 5.6. Probabilistic design of mean of Lub1........................................................... 100

Figure 5.7. Effect of variation of random variables on the quality index....................... 104

Figure 6.1. The blank holder control profile from [44] and the fitted Gaussian
approximation for sheet drawing. ........................................................................... 116

Figure 6.2. The FEM simulation model of the sheet drawing process and the final drawn
part. ......................................................................................................................... 118

Figure 6.3. Maximum thinning and sidewall wrinkling (y-axis) at different s of BHF
(x-axis). ................................................................................................................... 125

Figure 6.4. The design solution of PI control, deterministic design and probabilistic
design for csw = 0.21 . ............................................................................................. 128

Figure 6.5. The design solution of PI control, deterministic design and probabilistic
design for csw = 0.23 .............................................................................................. 129

Figure 6.6. Histograms of sidewall wrinkling and maximum thinning for the deterministic
designs..................................................................................................................... 131

Figure 6.7. Histograms of sidewall wrinkling and maximum thinning for the probabilistic
designs..................................................................................................................... 132

Figure 6.8. Evaluation of probabilistic design vs. deterministic design ......................... 135

Figure 7.1. Example of one setup of drawbeads on the Hishida part drawing [51]........ 142

xiv
Figure 7.2. The segment-elastic binder used in the Hishida part drawing [51]. ............. 144

Figure 7.3. Micro-texture to alter the friction condition [53]. ........................................ 147

Figure 7.4. Topological effect at the roughness level [55]. ............................................ 148

Figure 7.5. The Hishida Part. .......................................................................................... 150

Figure 7.6. Spatial distribution of discrete friction zones............................................... 152

Figure 7.7. The Pareto optimal and Pareto front............................................................. 156

Figure 7.8. The procedure of NSGA-II [61]. .................................................................. 159

Figure 7.9. The simulation model of discrete friction drawing of Hishida part. ............ 163

Figure 7.10. The optimization flow chart. ...................................................................... 168

Figure 7.11. The Pareto front of discrete friction NSGA-II optimization. ..................... 169

Figure 7.12. Uniform friction design and strain distribution after drawing.................... 171

Figure 7.13. Discrete friction design and strain distribution after drawing. ................... 172

Figure 7.14. Strain distribution at different locations of the part after drawing. ............ 173

Figure 7.15. Reliability analysis of the Pareto front points. ........................................... 176

Figure 7.16. Reliability of feasible points in the last Pareto front when BHF COV=0.05,
friction COV=0.1 .................................................................................................... 177

Figure 7.17. Reliability of feasible points in the last Pareto front when BHF COV=0.05,
friction COV=0.2 .................................................................................................... 178

Figure 7.18. Reliability of feasible points in the last Pareto front when BHF COV=0.05,
friction COV=0.4 .................................................................................................... 179

xv
CHAPTER 1

1. INTRODUCTION

At present, metal stampings are used in almost every mass-produced product. Consider

the number of consumer and industrial products that include sheet metal parts:

automobile and truck bodies, airplanes, railway cars, farm and construction equipment,

appliances, office furniture, computers, and more. Although these examples are

conspicuous because they gave sheet-metal exteriors, many of their internal components

are also made of sheet. According to a survey in the US, some 100,000 metal stampings

could be found in the average American home in the 1980s [1]. The commercial

importance of sheet metalworking is significant. The three major categories of sheet

metal processes are cutting, bending, and drawing. Cutting is used to separate large

sheets into smaller pieces, to cut out a part perimeter, or to make holes in a part. Bending

and drawing are used to form sheet metal parts into their required shapes. In this research,

we focus mainly on the drawing process.

Stamping product qualities have always been one of the most important concerns of the

industries. Any quality issues can be very costly to the manufacturer, creating difficulties

in assembly, causing rework or repair in the production floor or the field, and resulting in

1
customer dissatisfaction. Traditionally, the products quality is assured by inspecting parts

in full or fractional after they have been manufactured against the specifications and

standards such as dimensions, visual characteristics, mechanical and electrical properties.

The inspection can only be done after the part is made. If the part is out of specification,

action will be taken based on experience about whether to stop the production line and

investigate the root cause. It is usual that before an out of specification signal is sent out,

the production has been kept running for a long time and produced a large amount of

scrap parts.

The practice of inspecting products after they are made is now being replaced rapidly by

the online quality control methods which were pioneered primarily by Demming,

Taguchi and Juran. Among them the statistical processes control (SPC) and related

control charts are the most famous and widely used tools in industries. The key concept

here is the usage of control limits instead of specification limits denoted in the design

drawing or manufacturing instruction card. The control limits, which should be much

narrower than the specification limits, are developed statistically based on the natural

process variation in a stable status or a good status. The goal is to keep the process under

this stable status so that all the parts will meet the spec.

Besides setting up the control limits and charts for the products, some important process

parameters could also be monitored, for example, the furnace temperature in hot forging

process. By this much tighter control limit, system abnormality such as mean/variation

2
shift could be detected much faster than the traditional inspection method and sources of

quality problems could be rapidly identified.

However, these online quality control methods only concentrate on the manufacturing

stage and cannot compensate for poor design quality. It is commonly known that by the

20-80 rule, nearly 80 percent of parts quality issues are due to improper product or

process design. Improper product design sometime means the product is not well

designed for the manufacturing. When a process is not well designed, it is highly likely

that the product quality is very sensitive to the process variation. If the variation is

uncontrollable, prohibitively costly process control schemes may be required to improve

the process capability, and they cannot guarantee a product robust to deterioration and

variability due to uncontrollable process factors.

More specifically, for sheet metal stamping process, a well designed process is much

more important than online SPC control. The reason is that when the traditional SPC

charts are used to monitor the process, and the out-of-control signals indicating the

process mean shift are encountered, a sheet metal stamping process does not have the

necessary adjustability in its process variable input settings to allow adjusting the mean

response in an out-of-control condition. Hence, the signals often go ignored. This means

that additional expense might be incurred due to service costs under warranty and, more

importantly, due to the loss of market share because of customer dissatisfaction.

3
Therefore, if the quality concept is moved further upstream to the design process, these

costs can be avoided. The need for costly process control, mass inspection, and service

costs are minimized if one optimizes product and process design to ensure product

robustness.

1.1 Traditional Approach to Process Design

Drawing is a sheet metal forming operation used to make cup-shaped, box-shaped, or

other complex-curves, hollow-shaped parts. It is performed by placing a piece of sheet

metal over a die cavity and then pushing the metal into the opening with a punch. The

blank must usually be held down flat against the die by a blank holder.

A lot of efforts have been put into the design phase in order to produce a defect-free part.

Given material and part shape, the optimum deep drawing design, in general, focuses on

three main aspects: the blank design, tooling design and process design. The blank design

includes optimizing blank geometries and thickness. The tooling design is to determine

the optimal punch and die radii, the punch and die clearance, the drawbead shape and

location. The process design tends to find the best setting of process parameters such as

the friction (lubricant type and procedure), the punch speed and the blank holder force.

4
However, most design processes mentioned above are based largely on designer’s

experiences and nowadays widely used deterministic finite element method (FEM)

simulations. Designer could check the drawability through the numerical prediction of

fracture and final sheet thickness, wrinkling, surface defects, springback and residual

stresses. If simulation reveals any potential failure or defect, designer would modify the

process according to the specific defect and his previous experiences. A new simulation

will be conducted again to verify the new design. If defect still exists, a second round of

design modification will be taken. This iteration is basically the traditional engineering

trial-and-error process using the computer simulation instead of real drawing experiment.

It is noticed that during this process, no process uncertainty or variation is considered. All

the simulation input parameters such as material properties are a sure and fixed number.

Therefore, we say the traditional design process is deterministic in nature.

After the design is put in the production, however, the process variations are actually

taken care of by the online quality control methods like statistical process control (SPC).

It is apparent that a gap exists between the process design and the process control. The

deterministic design does not consider any process uncertainty/variation while the

corresponding process control statistically utilizes them in a fundamental level.

Majeske [2] has researched several leading automobile manufacturers to identify sources

of variation in sheet metal stamping. His result shows that within the same batch, which

means same die and process setup, the part-to-part variation is around 30 percent of the

5
total part variation in the long run. This is equal to say that no matter how good your

deterministic design is, the inevitable process variation may make the design output large

variation and thus bad quality.

1.2 New Approach for Robust Process Design

As illustrated in the previous section, a design process considering the uncertainties and

variations should be adopted to improve the products quality. Robust design is such a

design philosophy. We call it a philosophy because its principle is straightforward:

design a process that is insensitive to the noise factors, but its implement is not trivial. In

fact, different robust design methods are still being developed nowadays.

The Taguchi robust design is believed to be the most widely used and recognized design

method. However, its shortcomings are obvious. First, the optimal design can only be at

the experimental points. Basically, the method calculates the signal/noise ratio for each

combination of control variables. The configuration with the largest ratio is selected as

the optimal design. No point between them could be evaluated. Second, the way Taguchi

design the experiment matrix arbitrarily assumes independence between control factors

and noise factors. Thus the highly possible coupling of them is ignored, which is not

justifiable. Third, the relationship between those variables is still not clear and effect of

variation of each noise factor on the part quality could not be quantified.

6
Comparing to Taguchi robust design, the reliability-based optimization is more

systematic and quantitative. It seeks to identify design solutions that not only optimize

performance (minimize or maximize one or more objectives), but also satisfy constraints

on the minimum reliability (or maximum probability of failure). Consequently, a

deterministic optimization problem is modified in creating a reliability-based

optimization problem by adding (and defining) random variables, and modifying

deterministic constraints to become probabilistic reliability constraints. However, the

biggest issue of this approach is that the objective function is evaluated at the mean value

of design variables and hence it is deterministic in nature instead of stochastic. Most

optimization schemes developed for this method could not handle it if we change the

objective function to include uncertainties and variation.

In this research, we propose three different probabilistic design approaches. It uses sheet

metal forming finite element method (FEM) simulation as the fundamental tool. When

the system meta-model is not complex, the design of experiments (DOE) technique and

response surface method (RSM) are integrated with FEM to build an explicit function to

connect process input to process performance output. With the quantification of

uncertainties of input variables, by Monte Carlo Simulation (MCS) or other reliability

analysis methods, the probability that the product conforms to its specification would

therefore be assessed. Through the right probabilistic optimization formulation that we

will talk about in later chapters, the robust optimal design could be found. In this

dissertation, we will also explain the method to deal with the case where the number of

7
process inputs is too many so that the design of experiment cannot handle and the system

meta-model cannot be approximated. The integration of FEM, multi-objective genetic

algorithm and reliability assessment approach will be illustrated in chapter 7. Its main

aim is to reduce process variability, reduce defect rates and improve process capability by

robust process design.

1.3 Objective of Research

The specific objectives of this research are as follows:

y To model an existing sheet metal forming process for:

à Understanding the impact of variation of process input on the variation of process

output.

à Designing the best input parameter settings for robust process design

à Improvement of the process capability and reduction of defect rate for a better

product quality.

The proposed methodology could be applied to most stamping process and other typical

manufacturing processes such as forging, rolling, injection molding and die casting.

Moreover, the analysis from the probabilistic design of the sheet drawing process would

reveal very important facts in the stochastic prospective and give us more understanding

about the process.

8
1.4 Research Significance and Benefits

Although extensive research has been done in exploring the deterministic effect of each

factor on the part, the impact of their variations on the fluctuation of output quality for

the stamping process is seldom addressed. The inclusion of uncertainty in the design and

optimization cycle should lead to better understanding of the impact of uncertainty

associated with system input on the system output. This understanding can then be

applied for managing such uncertainties.

The proposed probabilistic design approach, which incorporates process uncertainty, is

capable of reducing process performance variation/defect rate and improving process

capability/quality. Our approach fully integrates finite element method (FEM), design of

experiments (DOE), response surface methodology (RSM), and performance variability

assessment methods like Monte Carlo Simulation (MCS)/Sensitivity-Based Variability

Estimation to deal with relatively simple case. For some really complex situations, an

integration of FEM, multi-objective genetic algorithm and reliability assessment is

illustrated. Though numerical methods have been in practical use for process design and

control, few studies in the past have addressed integration of FEM, system modeling and

approximation such as RSM, and probabilistic design. Our research will have great utility

in the design and evaluation of production processes.

9
The proposed systematical approach would significantly contribute to:

y Designing a production process without undergoing costly trial and error

procedures.

y Finding the robust production process design in terms of reduced defects,

improved quality and hence less costs.

y Understanding quantitatively the impact of process input variation on the

variation of process performance output so that the production system is not a

black-box any more.

1.5 What is My Contribution?

There are three aspects in this research that are my contributions.

1. Very few researches have been done to systematically and quantitatively include the

process uncertainties into the process design and analysis cycle in the sheet metal

forming area. There have some Taguchi robust design works done before. But as

mentioned previously, the Taguchi method has many shortcomings.

2. The proposed integrated probabilistic design approach is very new in the sheet metal

forming area and even in other metal forming areas like forging and rolling. It

combines sheet forming numerical simulation, process modeling/approximation,

10
uncertainties modeling and probabilistic design optimization formulation and leads to

a robust setting. Hence, the traditional trial-and-error process design efforts, which

add to the cost of manufacturing, will be reduced.

3. This research also proposes a new method to deal with complex system robust design

where the number of process input is so large that the system can not be modeled by

the DOE and RSM approach. The integration of FEM, multi-objective genetic

algorithm, and reliability assessment provides a good solution.

1.6 Dissertation Outline

Chapter 1 outlines the motivation behind this work along with the objectives and research

approach. Chapter 2 presents a brief overview of FEM, DOE and system approximation

techniques. Chapter 3 gives introduction to current used Taguchi robust design, reliability

based optimization and our proposed probabilistic design. Chapter 4 presents a simple

cylindrical cup drawing experiment showing that at the fixed level of process setting the

part quality characteristics can have a very large variation. This experiment serves as a

justification to conduct the probabilistic design for the sheet metal forming process. The

simple probabilistic design methodology using the quality index as objective is illustrated

in chapter 5. By considering the material properties variation and process condition

variation, an optimal deterministic blank holder force and stochastic friction coefficients

11
means are obtained to minimize the defect rate. A more formal probabilistic design

formulation is explained in chapter 6, where the temporal varying blank holder force

itself is treated as random variable like what we see in the real world. By incorporating

process variations, both the reliability constraints and Taguchi type objective are

considered in the design optimization formulation. The probabilistic design is also

compared with the deterministic design and the PI design which is usually adopted to

find the variable blank holder force profile. In chapter 7, spatial varying constraints on

the sheet are studied with the introduction of the very new discrete friction concept. The

finite element simulation is integrated with the multi-objective genetic algorithm and

drawing process heuristics to find the optimal deterministic design configuration. Then

the reliability analysis method is used to find the probabilistic design optimum. Chapter 8

concludes our research work and address the future work.

12
CHAPTER 2

2. BACKGROUND AND LITERATURE REVIEW

To understand the effect of process uncertainties on the product quality, we need to have

a system model which connects the input to output. A basic tool in knowing output to a

certain input for a manufacturing process would be the numerical simulation based on

finite element method. Currently, for most manufacturing processes like forging, rolling,

stamping, injection modeling, die casting and machining, proved and effective FEM

commercial software have been widely adopted in the industries. For forging/rolling

simulations, DEFORM and FORGE3 are two major FEM packages. For stamping,

PAM-STAMP and LS-DYNA are dominant. For injection molding, we have

MOLDFLOW. For die casting, we have PROCAST. For machining, DEFORM has

developed a dedicated module to simulate the chip formulation and breakage.

The widely used FEM manufacturing process simulation software not only provide an

effective tool in developing process for a much shorter time cycle, but also offer us an

economic way to investigate the impact of process uncertainty on the product quality. We

can image the case where we need to understand how the variation of the die radius

would affect the drawing quality. By real experiments, it will be so expensive and
13
even prohibitive to fabricate different configurations of dies, because each set of die can

easily cost a couple of thousands dollars. However, by using PAM-STAMP, the industry

proved simulation software, we only need several hours of computation to identify the

relation between die radius and part wrinkling/fracture.

Although simulation is much cheaper than the real experiment in the monetary terms, the

time needed by numerical computation still pose a severe constraint in exploring the

manufacturing system behavior. Normally, a three dimensional simulation for a typical

industrial like forging case could take one day to finish. In some extreme case, for

example, the simulation of the chip formation in the machining process may consume up

to one month. Therefore, systematic way of running simulations by certain principles

instead of randomly searching the process design space is really important and necessary.

Design of Experiments (DOE), which includes several design techniques, could help us

figure out the economic number and design of simulation run so that a desired level of

system approximation model could be built.

Closely related to the DOE usage are the system approximation methods. DOE tells you

how to design a smart experiment to investigate a system, while approximation methods

are data analyzing/processing tool to build a compact mathematical system model which

captures and represents the system itself.

14
There are three major system approximation methods: the Response Surface Method

(RSM), the Kriging method, and Artificial Neural Network (ANN). In this chapter, we

will first give background information on finite element methods and its application in

sheet metal forming. Then we will talk about the common Design of Experiments

methods and the chosen design for our FEM simulations. Following DOE, the system

approximation methods will be addressed. All these three topics are the fundamentals to

understand later on how we incorporate process uncertainty to investigate/design the

process.

2.1 Finite Element Method and Its Application in Sheet Metal Forming

2.1.1 Introduction to Sheet Metal Forming

Sheet metal is simply metal formed into thin and flat pieces, which are usually less than

6mm. There are different metals that can be made into sheet metal. Aluminum, brass,

copper, cold rolled steel, mild steel, tin, nickel and titanium are just a few examples of

metal that can be made into sheet metal. Sheet metal has wide applications in car bodies,

airplane wings, medical tables, roofs for building and many other things.

Sheet metal forming refers to various processes used to convert sheet metal into different

shapes for a large variety of finished parts. The typical forming processes include
15
stretching, drawing, cutting, bending and flanging, punching and shearing, spinning,

press forming and roll forming. In this research we will primarily focus on the sheet

metal drawing process.

A sheet metal forming system comprises all the input variables relating the blank

(geometry and material), the tooling (geometry and material), the conditions at the

tool-material interface, the mechanics of plastic deformation, the equipment used, the

characteristics of the final product, and finally the plant environment in which the process

is being conducted. The design, control, and optimization of sheet forming processes

require analytical knowledge regarding metal flow, stresses as well as technological

information related to lubrication, material handling, die design and manufacturing, and

forming equipment. The mechanics of deformation provides the means for determining

how the metal flows, how the desired geometry can be obtained by plastic deformation,

and what the expected mechanical properties of the produced part are [3].

The state of deformation in a plastically deforming metal is fully described by the

displacements, velocities, strains, and strain-rates. The basic mechanisms in sheet metal

forming are stretching, drawing, and bending. Depending on the shape and the relative

dimensions of the blank and the tool, one or more basic mechanisms is predominantly

involved. The limits of sheet-metal forming are determined by the occurrence of defects,

such as wrinkles and fracture in the blank [3]. An important development in checking the

formability of sheet metals is the forming limit diagram. In this diagram the major and

16
minor surface strains at a critical site are plotted at the onset of visible, localized necking

in a deformed sheet, and the locus of strain combinations that will produce failures in an

actual forming operation can be drawn. Experimental methods are used to construct the

diagram.

2.1.2 Basis of Finite Element Method

The concept of the finite element procedure may be dated back to 1943 when Courant

approximated the warping function linearly in each of an assemblage of triangular

elements to the St. Venant torsion problem and proceeded to formulate the problem using

the principle of minimum potential energy. Similar ideas were used later by several

investigators to obtain the approximate solutions to certain boundary-value problems. It

was Clough who first introduced the term “finite elements” in the study of plane

elasticity problems. Since then numerous studies have been reported on the theory and

applications of the finite element method.

The basic concept of the finite element method is one of discretization. The finite

element model is constructed in the following manner. A number of finite points are

identified in the domain of the function, and the values of the function and its derivatives,

when appropriate, are specified at these points. The points are called nodal points. The

domain of the function is represented approximately by a finite collection of sub-domains

called finite elements. The domain is then an assemblage of elements connected


17
together appropriately on their boundaries. The function is approximated locally within

each element by continuous functions that are uniquely described in terms of the

nodal-point values associated with the particular element [4].

The path to the solution of a finite element problem consists of five specific steps: [4]

1. identification of the problem;

2. definition of the element;

3. establishment of the element equation;

4. assemblage of element equations; and

5. numerical solution of the global equations.

2.1.3 Application of Finite Element Method in Sheet Metal Forming

The use of finite element simulation technology for stamping applications is growing

rapidly these days. When looking at the recent history of virtual stamping, one can

distinguish several main time periods. The first period, before 1990, was in fact a

pre-industrialization period. In this period, people started using computers to understand

how to improve the forming process. At this stage the different attempts were mainly

focused on approaches such as expert system and knowledge based methods. An

important breakthrough occurred with the success of forming simulation applying finite

element method to stamping related problems.

18
Using the elastic-plastic approach, complete solutions of stretch-forming and

deep-drawing problems, taking into account the contact problem at the blank holder, die,

die profile, and punch head, were obtained by Wifi [5]. On the basis of the nonlinear

theory membrane shells, Wang and Budiansky [6] developed a procedure for calculating

the deformations in the stamping of sheet metal by arbitrarily shaped punches and dies.

Onate and Zienkiewicz [7] presented a finite-element formulation based on an extension

of the general viscoplastic flow theory for continuum problems to deal with thin shells.

Toh and Kobayashi [8,9] analyzed sheet-metal forming processes, axially symmetric and

non-symmetric, by the finite-element method based on the membrane theory. The

finite-element model takes into account the rigid-plastic material characteristics and

includes the normal anisotropy of the sheet metal as well as the finite deformation that

occurs during the sheet-forming process.

The main usage of stamping simulation software concentrated on strain predictions and

the introduction of stamping-related know-how. The user wanted to have answers to the

following questions: is this part feasible? Where does it fail? Where will wrinkling

happen? What does my forming limit curve look like? Over the years, stamping

simulation has helped to reduce the costs and lead-time of various components

considerably. One can identify the main parameters that are influencing the stamping

process: the part geometry and the die run-off design, the materials’ selection, the

manufacturing hardware and process and final quality control.

19
In our research, the FEM simulation package Pam-stamp is used intensively. Pam-stamp

2G is a calculation code that uses the finite element method (FEM). All the components

of a calculation (metal sheet, tools, drawbeads) are segmented as meshes, i.e. a discrete

representation of the geometry. For non-deformable tools, the mesh is only a

representation of the geometry, and the finite elements are only used for contact

description. On the other hand, for the blank or a deformable tool, the finite elements that

form this mesh represent small pieces of the material with a prescribed deformation

behavior. The mechanical phenomena that occur in a blank are faithfully reproduced

using a large number of these elements. The finer the mesh to be generated, the better

quality of the results, while the higher the number of elements, the longer the calculation

time.

Depending on the calculation type (implicit or explicit) the calculation is sub-divided into

increments or time-steps. Generally, implicit increments are large with respect to the

explicit time-steps. Positions, velocities, accelerations and forces are permanently

calculated at the nodes, which are points linked to the material. Within the elements,

strains are calculated from positions. Corresponding stresses are then obtained, which

result in forces on the nodes. This calculation is repeated over all the elements for the

entire duration of the calculation. Boundary conditions are used to remove degrees of

freedom, while velocities and forces further define the kinematic behavior of the finite

element model. To describe the actual deformation process, material properties and

thickness, can be assigned to an element [10].

20
2.2 Design of Experiments Techniques

Design of Experiments includes the design of all information-gathering exercises where

variation is present, usually under the full control of the experimenter. Often the

experimenter is interested in the effect of some process or intervention on some objects.

Design of experiments is a discipline that has very broad application. In the following

part, we will introduce the most frequently used DOE techniques.

2.2.1 Full-factorial Design

A full-factorial design is one in which all combinations of all factors at all levels are

evaluated. It is an old engineering practice to systematically evaluate a grid of points,

requiring n1 ∗ n2 ∗ n3 ∗ ...ni ( i is the number of factors, ni is the number of levels for

factor i ) design point evaluations. This practice provides extensive information for

accurate estimation of factor and interaction effects. However, it is often deemed

cost-prohibitive due to the number of analyses required [11].

2.2.2 Orthogonal Arrays

The use of orthogonal arrays can avoid a costly full-factorial experiment in which all

combinations of all factors at different levels are studied. A fractional factorial

21
experiment is a certain fractional subset (1/2, 1/4, 1/8, etc.) of the full factorial set of

experiments, carefully selected to maintain orthogonality (independence) among the

various factors and certain interactions. While the use of orthogonal arrays for fractional

factorial design suffers from reduced resolution in the analysis of results (i.e., factor

effects are aliased with interaction effects as more factors are added to a given array), the

significant reduction in the required number of experiments can often justify this loss in

resolution as long as some of the interaction effects are assumed negligible.

In fractional factorial designs, the number of columns in the design matrix is less than the

number necessary to represent every factor and all interactions of those factors. Instead,

columns are “shared” by these quantities, an occurrence known as confounding.

Confounding results in the dilemma of not being able to realize which quantity in a given

column produced the effect on the outputs attributed to that column. In such a case, the

designer must make an assumption as to which quantities can be considered insignificant

(typically the highest-order interactions) so that a single contributing quantity can be

identified [11].

2.2.3 Latin Hypercube Design

Another class of experimental design which efficiently samples large design spaces is

Latin Hypercube sampling. With this technique, the design space for each factor is

uniformly divided (the same number of divisions ( n ) for all factors). These levels are
22
then randomly combined to specify n points defining the design matrix (each level of a

factor is studied only once). For example, figure 2.1 illustrates a possible Latin

Hypercube configuration for two factors ( x1 , x2 ) in which five points are studied.

Although not as visually obvious, this concept easily extends to multiple dimensions.

An advantage of using Latin Hypercubes over Orthogonal Arrays is that more points and

more combinations can be studied for each factor. The Latin Hypercube technique allows

the designer total freedom in selecting the number of designs to run (as long as it is

greater than the number factors). While, the configurations are more restrictive using the

Orthogonal Arrays.

A drawback to the Latin Hypercubes is that, in general, they are not reproducible since

they are generated with random combinations. In addition, as the number of points

decreases, the chances of missing some regions of the design space increases [11].

23
Figure 2.1. Latin Hypercube design [11].

24
2.2.4 Central Composite Design

Central Composite Design (CCD) is a statistically based technique in which a 2-level

full-factorial experiment is augmented with a center point and two additional points for

each factor (star points). Thus, five levels are defined for each factor, and to study n

factors using Central Composite Design requires 2n + 2n + 1 design point evaluations.

The corner points are for the assessment of linear and 2-way interaction terms. Center

points are used to detect curvature and sometime replicated in experimental DOE to

estimate pure error. Star points are for the assessment of quadratic terms, see figure

2.2(a). Although Central Composite Design requires a significant number of design point

evaluations, it is a popular technique for compiling data for Response Surface Modeling

due to the expanse of design space covered, and higher order information obtained [11].

2.2.5 Box-Behnken Design

Box and Behnken developed a family of efficient three-level designs for fitting

second-order response surfaces. It exists only for 3-7 factors. Number of runs is very

close to CCD for the same number of factors. The Box-Behnken design doesn’t have any

corners and it is suitable for the situation when corners are not feasible (physical designs),

see figure 2.2(b).

25
(a) Central Composite Design

(b) Box-Behken Design

Figure 2.2. Comparison of CCD design and Box-Behnken design.

26
2.2.6 Computer Aided Designs

Other popular methods to select the design points are computer-aided designs. Computer

aided designs are generated based on a particular optimality criterion and are generally

optimal only for a specified model. The common types of optimality criteria include

D-optimality, A-optimality, G-optimality and V-optimality. Unlike the standard classical

designs such as factorials and fractional factorials, the computer-aided design matrices

are usually not orthogonal. These methods are particularly useful when standard designs

(e.g. factorial or CCD) cannot be simply implemented. Such situations might arise, for

example, when the design space is irregular due to feasibility constraints, or when there

are economic constraints on the size N of the experiment. In such cases, optimality

criteria and associated numerical techniques provide objective methods for selecting

design points.

2.3 Approximating Methods

Approximation concepts were introduced in structural design optimization in the late

1970s to do the following:

à Reduce the number of independent design variables through design variable

linking and reduced basis vectors concepts.

27
à Perform constraint deletion through truncation and regionalization schemes.

à Reduce the number of computer intensive, detailed analyses (or simulation code

evaluations) through the use of mathematical approximations of the design

optimization objective and constraint functions.

These approximations models can be used to reduce simulation codes or analyses that are

computation intensive. They can also help to eliminate the computational noise for

simulation codes in the case the outputs rapidly oscillate with gradual changes in the

values of input parameters. Computational noise has a strong adverse effect on

optimization by creating numerous local optima. Approximation models (Response

Surface Models in particular) naturally smooth out the response functions, and, in many

cases, help to converge to a global optimum faster. The usage of approximation is not

restricted to optimization. It also provides an efficient means of post-optimization or

sensitivity analysis. Their value is very high for computationally expensive engineering

methods, such as Monte Carlo Simulation, Reliability-Based Optimization, or

Probabilistic Design Optimization which we will conduct in this research [11].

2.3.1 Response Surface Method

Response surface method is a collection of statistical and mathematical techniques useful

for developing, improving, and optimizing processes. In some systems based on the

underlying engineering, chemical, or physical principles, the nature of the


28
relationship between y and x ’s might be know exactly. Then a model of the form

y = g ( x1 , x2 ,..., xk ) + e can be written. This type of relationship is often called a

mechanistic model. However, the more common situation would be that the underlying

mechanism is not fully understood, and the experimenter must approximate the unknown

function g with an appropriate empirical model y = f ( x1 , x2 ,..., xk ) + e . Usually the

function f is a first-order or second-order polynomial. This empirical model is called a

response surface model.

The model then can be used in optimization studies with a very small computational

expense, since evaluation only involves calculating the value of a polynomial for a given

set of design variables. Accuracy of the model is highly dependent on the amount of

information collected for its construction (number of exact analyses), shape of the exact

response function being approximated (like the order of polynomial), and volume of the

design space in which the model is constructed (the range covered by the RSM). In a

sufficiently small volume of the design space, any smooth function can be approximated

by a quadratic polynomial with good accuracy. For highly non-linear functions,

polynomials of 3rd or 4th order can be used. If the model is used outside of the design

space where it was constructed, its accuracy is impaired, and refining of the model is

required [11].

The response surface model relies on the fact that the set of designs on which it is based

is well chosen. Randomly chosen designs may cause an inaccurate surface to be

29
constructed or even prevent the ability to construct a surface at all. Because simulations

are often time-consuming or the experiments are expensive, the overall efficiency of the

design process relies heavily on the appropriate selection of a design set on which to base

the approximations. CCD design, Box-Behnken design, D-optimal design are the widely

used DOE methods to generate the design set for constructing a response surface model.

2.3.2 Kriging Meta Models

Kriging (named after the South-Afican mining engineer Krige) is an interpolation method

that predicts unknown values. More precisely, a Kriging prediction is a weighted linear

combination of all output values already observed. These weights depend on the

distances between the new and the observed inputs. The closer the inputs, the bigger the

weights are. Kriging models are extremely flexible due to the wide range of correlation

functions which can be chosen for building the approximation model. Furthermore,

depending on the choice of the correlation function, the model either can provide an

exact interpolation of the data, or an inexact interpolation [11].

The most popular DOE for Kriging is Latin Hypercube Design. LHS offers flexible

design sizes n (number of scenarios simulated) for any value of k (number of simulation

inputs). Geometrically, many classic designs consist of corners of k-dimensional cubes,

so these designs imply simulation of extreme scenarios. LHS, however, has better space

filling properties.
30
2.3.3 Neural Networks

Artificial Neural Networks (ANN) has been studied for many years in the hope of

mimicking the human brain’s ability to solve problems that are ambiguous and require a

large amount of processing. Human brains accomplish this data processing by utilizing

massive parallelism, with millions of neurons working together to solve complicated

problems. Similarly, ANN models consists of many computational elements, called

“neurons” to correspond to their biological counter-parts, operating in parallel and

connected by links with variable weights. These weights are adapted during the training

process, most commonly through the back-propagation algorithm, by presenting the

neural network with examples of input-output pairs exhibiting the relationship the

network is attempting to learn. The most common applications of ANN involve

approximation and classification. Approximation models attempt to estimate input-output

transformation functions, while classification involves using the known inputs to

determine class membership. Until now, we found there is no much literature about the

optimal experimental design for neural networks or even verification of the effectiveness

of the traditional regression model based optimal design methods on the neural net.

However, we have conducted a comparative study and the result shows that for building

the approximation model by neural network, the Bayesian D-optimal design (one kind of

computer aided DOE) gives the better prediction accuracy than the other experimental

design methods such as Latin Hypercube design or D-optimal designs. The details are

shown in the appendix A.

31
2.4 Literature Review of Deterministic Sheet Metal Forming Process

Optimization Employing FEM, DOE and Approximation Methods

Ayed et al. [12] presented a FEM combined with RSM approach to optimize the blank

holder force considering the drawing of a front door panel. The numerical simulations

were performed using ABAQUS Explicit. The parameters of the finite element model

(mesh density, speed of punch) were set to achieve a good prediction with a minimum

simulation time. The objective function was defined to minimize the work of punch.

Three inequity constraints functions were defined to avoid necking and wrinkling. To

avoid necking, the major stress of the blank was limited to a value, which was

determined by using the modified maximum force criterion. To avoid wrinkling, under

the blank holder, the angle between the blank holder surface and an element of the blank

was limited to a value set by user, as proposed by Gelin and Labergere[13]. In the useful

part of the workpiece, the major stress was limited to a value. A central composite design

experiment was applied to generate response surface model. For n independent variables,

the central composite design requires 2n + 2n + 1 simulations: 2n factorial designs

augmented by 2n axial points and one center point. Thus for seven blank holder forces

143 numerical simulations were necessary. Then a SQP algorithm was used to find the

optimum blank holder force.

Kim and Huh [14] carried out optimization of the process parameters for process design

in sheet metal forming processes. The scheme incorporated rigid-plastic FEM for the

32
deformation analysis and RSM for the optimum searching of process parameters. The

algorithm developed was applied to design of the draw bead force and the die radius in

deep drawing processes of rectangular cups. The algorithm showed the capability of

designing process parameters which enabled the prevention of the part being weak or

fracture during stamping processes. Kim et al. [15] used rigid-plastic FEM with modified

membrane elements for an analysis tool and RSM for constructing the approximation

surface for searching the optimum draw bead force in the sheet metal forming process.

The algorithm developed was successfully applied to a design of the draw bead forces in

the deep drawing process.

Tezuka et al. [16] used rigid plastic FEM and RSM for process parameter determination

in the sheet metal forming process. Using this methodology, process parameters such as

the optimum bead force in the deep drawing process were effectively calculated.

Lepadatu, et al. [17] presented a sheet metal bending process optimization method for

springback minimization that combined finite element analysis, response surface method

and gradient optimization algorithm. In his work, the optimization computation was

carried out with a FORTRAN program using the gradient method. In the first phase,

polynomial models were generated with the available DOE data obtained by finite

element simulation. In the second phase, the optimizer used the objective function during

the search for the optimum until the final converged solution was obtained. Springback of

sheet parts during bending process was simulated using finite element model including

33
damage evolution effects within the sheet. This simulation was based on a constitutive

law of large elasto-plastic strains coupled with damage type Lemaitre. Die corner radius,

punch-die clearance and blank holder force were the three main variables considered.

Central composite experimental design (CCD) was selected to generate data for fitting

the response surface.

Wang, et al. [18] reported a strain path controlled forming process through adjusting the

blank holder force located in various flange areas to achieve a facture free product. Due

to the wide application and the high efficiency of the finite element method, it was the

means used instead of experiment to carry out the controlled forming process. Deep

drawing processes with two materials, SPCC and Al 5754, were simulated with yield

criteria Hill 48 and Hill 90, respectively. The empirical equation, built through response

surface method (RSM), was developed based on the finite element method (FEM)

simulation results. Central composite design (CCD) was used to guide the simulations.

The equation, working as system model, explored the relationship between the principal

strain of fracture risk elements and the process parameters, especially the blank holder

force located in various flange areas. The empirical equations were developed based on a

set of numerical simulation results within two elements which have the high fracture risk.

Three dimensional principal strain surfaces of one element were drawn which displayed

the obvious trend varied with the space variant blank holder force. The model adequacy

was checked and confirmed using ANOVA method. Four extra sets of numerical

simulation were carried out to compare with the value predicted using RSM. Good

34
consistency confirms the effectiveness of this empirical equation with the plasticity field.

With the assistance of the empirical model, it was feasible to acquire a fracture free

product.

Forsberg, et al. [19] investigated the accuracy of response surface and Kriging modeling.

For RSM, the true response is usually replaced with a low-order polynomial. In Kriging

the true response is replaced with a low-order polynomial and an error correcting

function. In both cases the D-optimality criterion has been used to distribute the design

points. Crashworthiness simulations were carried out at the design points. From the

investigation, they found that Kriging better than RSM resolved abrupt changes in the

response, e.g. due to buckling, contact or plastic deformation. However, as seen from the

derivation of the D-optimality criterion, it was not valid for the approximation using the

Kriging technique. Therefore, the conclusion about Kriging was better than RSM was not

fair enough. The space-filling design like the Latin hypercube should be included in the

investigation.

Huang, et al. [20] illustrated an efficient method to optimize the intermedial tool surfaces

in the multi-step sheet metal stamping process to obtain improved quality of a product at

the end of forming. The proposed method was based on a combination of finite element

method (FEM) and the response surface method (RSM). The objective of the

optimization was to minimize the thickness variation within the part at the final stage.

The optimal radius of the intermedial surface and fillet radius were found.

35
Yamazaki, et al. [21] tried to apply the response surface approximate method to develop

aluminum beverage can ends. Geometrical parameters of the end shell were selected as

design variables. The analysis points in the design space were assigned using an

orthogonal array in the design of experiment technique. Finite element analysis code was

used to simulate the deforming behavior and to calculate buckling strength and central

panel displacement of the end shell under internal pressure. On the basis of the numerical

analysis results, the response surface of the buckling strength and panel growth were

approximated in terms of the design variables. By using a numerical optimization

program, the weight of the end shell was minimized subject to constraints of the buckling

strength, panel growth suppression and other design requirements.

Lin, et al. [22] established an effective prediction model of the spring-back of material

during the processing of an L-shaped bend by artificial neural networks (ANN). FEM

simulation of an L-shaped bend was first carried out for various thickness of material,

punch-round-radii and die-round-radii. The results of spring-back from FEM simulation

were then input to a neural network to establish a model for the L-shaped bend variables.

He concluded that the neural networks approximating fitted the experiments and the

optimal design could be achieved by this model.

Ji, et al. [23] used finite element method and neural network to inversely design the

rolling process parameters. The neural network could accurately predict the seam

36
forming and grain size in the rolling process. Then an inverse neural network was

constructed to find the optimal process setting given the grain size and constraint of no

seams.

Hambli, et al. [24] presented a similar approach that combined finite element simulation

with neural network modeling of the leading blanking parameters in order to predict the

burr height of the parts for a variety of blanking conditions. The numerical results

obtained by finite element computation including damage and fracture modeling and tool

wear effects were utilized to train the developed simulation environment based on back

propagation neural network modeling. The comparative study between the results by

neural network and the experimental ones showed good match.

37
CHAPTER 3

3. RESEARCH METHODOLOGY

Before we start talking about our approach of the probabilistic design for sheet metal

forming, we need to give introduction to the current design methods that incorporate

uncertainties. The most widely used and famous method was created by the quality guru

Taguchi: the Taguchi robust design.

3. 1 Taguchi Robust Design

3.1.1 Introduction to Taguchi Robust Design

The motive of robust design is to improve the quality of a product or process by

achieving performance targets and minimizing performance variation. The variables are

usually classified as the following categories:

à Control Variables X . These variables can be designed and controlled in the

manufacturing process.

38
à Noise variables Z . They are either not controllable, or too difficult or expensive

to control in the manufacturing process. Noise variables can cause the variation of

responses Y and lead to quality loss.

à Response variables Y . They are dependent performance characteristics.

Responses are the system outputs, and are functions of the control and noise

variables.

The robust design is to seek the settings of the control variables to reduce the variation of

system performance responses caused by uncertainty of noise variables and achieve

performance targets. One note here is that the control variables may have variation too at

the manufacturing process and thus cause variation of response variable. In this case, we

are seeking the best mean values of the control variables.

Taguchi's robust design evaluates the mean performance and its variation by crossing two

arrays: an inner array, designed in the control variables, and an outer array, designed in

the noise variables. As shown in figure 3.1, a two level factorial design is adopted for

both the inner and outer array. For each row of the inner array, response values are

generated for each noise variables combination. For example, inner array row 1 with

outer column 1 leads to the response value y11 , inner row 1 with outer column 2 leads to

response value y12 , and so on. This design then leads to multiple response values for

each combination of control variables, from which a response mean, μ , and variance or

standard deviation, σ , can be computed.

39
Given the mean and variance for each inner array row, the experiments can be compared

to determine which set of control settings best achieves “mean on target” and “minimized

variation” performance goals. Taguchi uses the signal-to-noise ratio (shown as S / N yi

for each control experiment), and quality loss (measured using a loss function) to

combine the effects of mean performance and performance variation, which can then be

used to compare the set of designs represented by the control variables combinations.

The S / N ratio calculation depends on the particular response being investigated [11]:

μ2
à The S / N ratio is 10 log10 if the response is a desired value.
σ2

⎛1 n ⎞
à The S / N ratio is − 10 log10 ⎜ ∑ y i2 ⎟ if the response is desired to be a lower
⎝ n i =1 ⎠

value.

⎛1 n 1 ⎞
à The S / N ratio is − 10 log 10 ⎜⎜ ∑ 2 ⎟⎟ if the response is desired to be a higher
⎝ n i =1 y i ⎠

value.

The second performance characteristic used by Taguchi Robust Design Techniques, the

loss function, is generally used to measure the loss of quality associated with deviating

from a targeted performance value, as shown in figure 3.2.

40
Figure 3.1. Taguchi robust design matrix [11].

41
Figure 3.2. Taguchi loss function [11].

42
Conventionally, the acceptable quality range is defined by the lower and upper

specification limits. All values within these two limits are assumed to have no quality

loss, and all values outside the limits are defined as having 100% quality loss. The best

example of this concept is GD&T. If the part dimension is larger or smaller than the

limits, the part is rejected. However, in Taguchi robust design, quality loss is measured

by the deviation from the target. This means loss of quality occurs gradually when the

quality characteristic moves in either direction from the target value, rather than as a

sharp cutoff with the conventional approach. The standard form of the loss function

L( y ) is given as follows:

L( y ) = k ( y − T ) 2 (3.1)

In the equation (3.1), y is the quality characteristic, such as a dimension or

performance parameter, T is the target value for the quality characteristic, and k is

the loss constant, which is used to convert the deviation from the target to appropriate

measure of quality loss.

3.1.2 General Taguchi Robust Design Optimization Formulation

The S / N ratio, loss function, and μ , σ can be combined to use in the case of

multi-objectives. The general Taguchi Robust Design formulation is then to search

through all the control design combinations and find the setting that minimizes the

43
following objective:

Robust design objective = [− S / N yi ] + [ L yj ] + [(±) μ yk + σ y2k ] (3.2)

In this general form, y i , y j , y k represents all the responses considered in the Taguchi

robust design problem. The plus sign is used when lower y k is desired, and minus sign

is used when higher y k is desired. Weight and scale values can also be used with each

objective.

3.1.3 Disadvantages of Taguchi Robust Design method

Nevertheless, the shortcomings of Taguchi’s robust design are obvious. First, the optimal

design can only be at the experimental points. The configuration with the largest ratio is

selected as the optimal design. No point between them could be evaluated. Second, the

way Taguchi design the experiment matrix arbitrarily assumes independence between

control factors and noise factors. Thus the highly possible coupling of them is ignored,

which is not justifiable. Third, the relationship between those variables is still not clear

and effect of variation of each noise factor on the part quality could not be quantified.

Although the objective function in Taguchi robust design is formulated well, the lack of

system model to connect x , z and y limits its application range.

44
3.2 Reliability Based Optimization

Reliability based optimization was developed based on reliability analysis. So before

delving into the optimization formulation, we need to explain its foundations: the

reliability analysis.

3.2.1 Reliability Analysis

The reliability analysis was first developed and applied in the structure safety field. It

incorporates uncertainties associated with geometrical and material properties, loading

and boundary conditions, and operational environment into structural analysis and design

by defining these random variables, their associated probabilistic distribution functions,

and statistical properties [11]. Due to the variation of input variables, the performance of

the designed structure component or system will experience variation too, some of which

may violate the constraints. The structural reliability is then defined as the probability of

satisfying a constraint, and is equal to 1- probability of failure, under the input

uncertainties. The concepts of reliability and probability of failure are illustrated in figure

3.3.

There are many methods developed in recent years to estimate the probability of failure

or reliability (estimating the areas inside and outside the constraints). We will mainly

introduce the following three commonly used methods.


45
Figure 3.3. Reliability based analysis [11].

46
Figure 3.4. FORM reliability analysis method [11].

47
à First Order Reliability Method (FORM)

à Mean Value First Order (MVFO) Method

à Monte Carlo Simulation (MCS) Method

These methods are able to evaluate reliability for the current design point. Each method

is summarized in the following sections.

3.2.1.1 First Order Reliability Method

The idea of FORM is based on the desirable properties of the standard normal probability

distribution. Hasofer and Lind [25] defined the reliability index as the shortest distance

from the origin of the standard normal space ( U -space) to a point on the failure surface.

The reliability index β can be determined by a minimization problem with one equality

constraint (3.3):

β = min U
U (3.3)
−1
s.t. g ( X ) = g (T (U )) = g (U ) = 0

In this formulation, first the original random vector X is mapped to the standard,

uncorrelated normal vector U by a transformation T . Then a minimization searches

the closet point U * on the failure function g (U ) to the zero point. U * is called the

Most Probable Point (MPP). If the failure function g (U ) is linear in terms of the

48
normally distributed random variables U i , the failure probability is calculated as

Pf = Φ (− β ) , see figure 3.4. In this equation, Φ is the standard normal distribution

function. If the failure function is nonlinear, the equation above is still a good

approximation, provided that the curvature of the failure surface at the MPP is not too

large in magnitude.

3.2.1.2 Mean Value First Order Method

The MVFO reliability method utilizes the first order Taylor's series expansion of failure

functions g ( X ) at the mean values μ X . Then the variance of the g ( X ) is the sum of

the multiplication of variance of each dependent variable and its square of derivatives at

μ X . The mean-value reliability index, which is the same as β in the FORM method, is

then calculated by dividing the mean and standard deviation of g ( X ) [11]:

g (μ X ) μg
β= =
⎛ ∂g ⎞
2 σg (3.4)
∑ ⎜⎜ ∂X ⎟⎟ (σ X i ) 2
⎝ i ⎠

Similarly, the probability of failure can then again be calculated as Pf = Φ (− β ) . In

terms of the number of function evaluations, or simulation program executions needed to

calculate the reliability, MVFO is the most efficient reliability analysis method since it

49
requires only one time failure function evaluation to calculate the mean and sensitivity

evaluations to calculate the derivative. However, the mean-value reliability index is

accurate only for linear failure functions with normally distributed random variables.

3.2.1.3 Monte Carlo Simulation Method

Monte Carlo Simulation is a very straight forward method to calculate the probability of

failure. The reliability level R and probability of failure Pf can be calculated as in

equation (3.5) [11]:

⎛ # simulations in failure region ⎞ ⎛ # simulations in safe region ⎞


R = 1 − Pf = 1 − ⎜⎜ ⎟⎟ = ⎜⎜ ⎟⎟ (3.5)
⎝ total # system simulations ⎠ ⎝ total # system simulations ⎠

This calculation is done by the following steps:

1. Identify random variables by assuming appropriate distributions, and defining

properties for each (mean, standard deviation, or coefficient of variation).

2. Specify the number of simulations to be executed (often 1000 - 10,000

simulations are necessary for accurate prediction of response statistical

properties).

3. Generate uniformly distributed random numbers for each random variable.

50
4. Convert each uniform random number to a random variable value corresponding

to appropriate distribution.

5. Evaluate failure function(s) using random variable values, and determine whether

simulation point is a success ( g ( X ) > 0 ) or failure ( g ( X ) < 0 ) for each failure

function g ( X ) .

6. Repeat step 3 through step 5 for the number of simulations specified in step 2.

7. Compute reliability R for each failure function.

3.2.2 Reliability Based Optimization Formulation

While the reliability-based design analysis methods seek to determine the reliability or

probability of failure of the current design (act on a single design point), reliability-based

optimization seeks to identify design solutions that not only optimize performance

(minimize or maximize one or more objectives), but also satisfy constraints on the

minimum reliability (or maximum probability of failure) [11].

Therefore, we can formulate a reliability-based optimization problem by modifying a

deterministic optimization problem by adding random variables, and reliability

constraints. The reliability-based optimization problem is formulated as follows (3.6):

51
Find the set of design variables X that:

Minimize : F ( X ,Y )
Subject to : g idet ( X , Y ) ≤ 0
(3.6)
j ( X ,Y , β j ) ≤ 0
g rel
Xl ≤ X ≤ Xu

In this formula, Y is the set of random variables, gi is the ith deterministic constraint,

g j is the jth reliability constraint, and β j is the reliability index value of the jth

reliability constraint. The reliability index value is calculated from the standard normal

distribution function by the desired reliability level. By incorporating the reliability index

to the constraints, the probability of failure must be equal or smaller than the desired

reliability. For example, a reliability index value of 2.0 corresponds to a reliability of

97.725% and a probability of failure of 2.275%.

FORM is usually embedded in reliability-based optimization to evaluate the satisfaction

of the reliability constraints. Since it also requires the solution of an optimization

problem to calculate the reliability, using this reliability analysis method within a

reliability-based optimization approach is considered a “double-loop” method (FORM

optimization loop within outer reliability-based optimization loop) [11].

A “single-loop” method, as documented in [26], and is significantly more efficient than

double-loop methods. In this approach, the critical (failure) point for each reliability

constraint is calculated using derivatives of the constraints, with respect to the random

52
variables, and the desired reliability index. The reliability constraints are then evaluated

at the critical point, while the objectives are evaluated at the mean value point (defined

by the mean values of the random variables).

3.3 Proposed Probabilistic Design approach

The Probabilistic Design combines elements from each to create a complete formulation

for accessing and improving reliability and robustness. Each of the methods discussed in

this chapter has its specific focus with respect to incorporating uncertainty.

Monte Carlo simulation is an investigative tool; no constraints or objectives are

formulated. Reliability analysis is focused on constraints; any objective included is

evaluated only at the mean value point (the current design point, with any random

variables set to their mean values). Reliability-based optimization is again focused on the

constraint formulation of an optimization problem, in converting deterministic constraints

to probabilistic constraints; again the objective retains its deterministic formulation and is

evaluated at the mean value point. Finally, Taguchi Robust Design places attention on

desired response values through the formulation of an objective function that includes the

desired mean performance and the “minimize variation” element; constraints are not

explicitly formulated.

53
Probabilistic design optimization includes in its formulation uncertainty information

related to variables, constraints, and objectives. The focus then is not only to identify

solutions that are reliable or robust with respect to constraint satisfaction, but also to

reduce the variability associated with objective components. Further, by defining the

variance or standard deviation of uncertain input design parameters not as fixed, but as

design variables themselves, tolerance design/optimization can be implemented by

seeking standard deviation settings for these parameters that produce acceptable

performance variation (objective components, constraints).

In the following chapters, first, we will present a simple cylindrical cup drawing

experiment, which shows that at the fixed level of process setting the part quality

characteristics can have a very large variation. This experiment serves as a justification to

conduct the probabilistic design for the sheet metal forming process. Then the simple

probabilistic design methodology using the quality index as objective is illustrated in

chapter 5. By considering the material properties variation and process condition

variation, an optimal deterministic blank holder force and stochastic friction coefficients

means are obtained to minimize the defect rate.

A more formal probabilistic design formulation is explained in chapter 6, where the

temporal varying blank holder force itself is treated as random variable like what we see

in the real world. By incorporating process variations, both the reliability constraints and

54
Taguchi type objective are considered in the design optimization formulation. The

probabilistic design is also compared with the deterministic design and the PI design

which is usually adopted to find the variable blank holder force profile.

In chapter 7, spatial varying constraints on the sheet are studied with the introduction of

the very new discrete friction concept. The die surface is segmented into 10 discrete areas

and punch is divided to 4 areas. Since the number of design variables under research is

fifteen, the methodology used in chapter 5 and 6 is not applicable here. In this complex

case, two phases design method is developed. The first step is to integrate the finite

element simulation with the multi-objective genetic algorithm to find the optimal

deterministic design configuration for those fifteen variables. Then the reliability analysis

method: Mean Value First Order is used to assess the defect rate for all the points within

the deterministic constraints in the last Pareto fronts which are obtained through the

multi-objective genetic algorithm.

55
CHAPTER 4

4. EXPERIMENT OBSERVATION ON THE PART QUALITY


VARIATIONS

4.1 Experiment Objective

The basic assumption of conducting probabilistic design for sheet metal drawing is that

although the process seems fixed under the pre-designed setting, there could still exist

great variations from various resources: the sheet material properties, sheet thickness,

lubrication, or blank holder force, etc.. These input variations can then cause large

variations of drawn parts quality characteristics, as schematically shown in figure 4.1.

The objective of this experiment is to observe and study the variations of drawn parts

quality at the fixed level of process setting so that our basic assumption about the

probabilistic design can be tested. The information about these variations will not only

justify our probabilistic design approach, but also provide statistical data to conduct

probabilistic design.
56
Input variations Output variations

Material Quality

Process Quality

y y
y y
Other Quality

Figure 4.1. Illustration of process output variations caused by input variations for sheet

drawing process.

57
4.2 Experimental Design

4.2.1 Criteria to Select Right Part to Draw

In this experiment, we only want to see the impact of variation of process parameters on

the parts quality variation. These process parameters variations include: sheet material

properties variation such as strength hardening coefficient, sheet thickness variation,

lubrication variation and blank holder force variation. The quality characteristics include

three aspects: wrinkling, fracture and springback. The criteria of selecting the

experimental part to draw should be based on two concerns:

1) Only the process parameters abovementioned should be considered in the drawing

process. No complex die/punch/blank geometry should be involved because their

variations may be coupled with the process variations we want to study.

2) The measurement of the three quality characteristics should be easy. The benefit of

easy measurement is the higher measurement gage R&R. Thus the measurement result

will almost reflect the part quality variation instead of measurement variation.

After careful consideration, the simple cylindrical cup drawing is selected in this

experiment since it satisfies both two criteria.

58
4.2.2 Drawing Die Design

The die/punch components designed for the cylindrical cup drawing are illustrated in the

figure 4.2. The hydraulic press is shown in figure 4.3. The die/punch parameters are

listed in table 4.1. In this test, we designed four components for the die part. The first is

the bottom plate, which is used to connect the die block to the hydraulic press bottom

stage. The die block is positioned above the bottom plate and functions as the die ring

holder. After die ring is put onto the die block, a centering ring is placed on the top of the

die ring and is pressed against the die block by the four screw bolts. There are four extra

screw holes on the centering ring, and their function is to lift the centering ring up by

screwing the bolts through the holes. By this design, we can easily take the die ring out,

and put in different die ring for other experiments. Therefore, we could save a lot by

using the same die block and centering ring.

The punch is connected to the hydraulic press through a screw cap on the top. To avoid

the negative pressure between the punch and the blank during the drawing, which may

cause problem to separate them afterward, we drilled a small hole through the punch as

air vent. In this drawing test, we also designed the blank holder, which is connected to the

up plate by the screws. The up plate is then T-slot connected to the up platen of the

hydraulic press. The sheet used in this experiment is HSLA350. The sheet thickness is

1mm. The coupon is cut to the disk shape with diameter of 5.8 inches.

59
Figure 4.2. The cylindrical cup drawing die design.

60
Figure 4.3. The hydraulic press used in the drawing test.

61
Punch diameter 3.250in
Punch corner radius 0.600in
Die ring radius 1.675in
Die corner radius 0.157in
Blank diameter 5.8in
Blank thickness 1mm
Blank material HSLA350
Lubrication Drawing oil (mid-state lub M2C)
Table 4.1. The die/punch design and test setup

62
4.2.3 The Measurement of Quality Characteristics

The three quality characteristics usually concerned in the sheet metal drawing are the

wrinkling, fracture and springback. One of the reasons that we choose the cylindrical cup

drawing is that the measurement of these three aspects is not difficult. The figure 4.4

shows schematically the front view of a cut open part after drawing.

1) The wrinkling can be measured by two metrics. The first is the number of wrinkles

along the flange. The second is the maximum height of the wrinkling mark on the cup

outside sidewall (the figure 4.4 shows the inside).

2) The fracture tendency is represented by the minimum sheet thickness along the cut

open cup. In the figure 4.4, the thinning happens at the transition area from cup bottom

corner to the straight side wall.

3) The springback can be measured by the expansion angle from cup bottom to cup top

after drawing. In the figure 4.4, Dtop is the diameter measured at the intersection of

flange and the sidewall. Dbottom is the diameter measured at the intersection of cup

bottom corner and sidewall. H is the vertical distance between them. The springback

Dtop − Dbottom
angle can be calculated by a tan( ).
2H

63
Wrinkling

Dtop
h
H
Wrinkling mark

Dbottom
t

Figure 4.4. Illustration of measurement of three quality characteristics of drawn cups.

64
The height of wrinkling mark and the maximum thinning are measured by the vernier

caliper. The Dtop and Dbottom are measured by micrometer. The gage R&R for those

two measurement methods in this case are around 10%. Besides, for each metric, 3

measurements will be conducted to minimize the measurement error and variation.

Therefore at this level of accuracy, we could assume that the measurement result will

represent mostly the variation of the drawn part quality instead of measurement error.

4.2.4 The Experiment Procedure

After we select the right part, design the drawing die, and figure out the measurement

methods for the quality characteristics, next we need to plan the experiment procedures.

Step One: First, we need to run some initial trial to find the feasible region of cup

drawing. Since the sheet material we only have is HSLA350, which is known for its good

strength but relatively poor limit drawing ratio (LWR). We want to determine the process

setting, which includes the right blank holder force, the right amount of lubricants

applied on the blank and right blank size, so that a cup without any defect can be drawn.

Step Two: After we find this good setting, we will keep it at fixed level and then repeat

drawing one by one. This means we will adopt the same way to put the drawing oil on

the same spots of the blank, place the blank on the same location on the die ring, use the

65
same blank holder force, and apply the same hydraulic press operation sequences. By this

way, we can reduce as much variation as possible induced by the human factors.

Step Three: After drawing enough cups according to the sample size (in this case 30),

we then measure the three quality characteristics of those cups. The measurement should

be taken in two steps. The first is to count the number of wrinkling, measure the

maximum height of wrinkling mark and the springback angle. After this, the cups will be

cup open in half and the minimum sheet thickness could be measured.

Step Four: At last, we need to analyze the data. The histogram is a good tool to visualize

the data distribution. The mean and standard deviation should be calculated too.

Normality test is also necessary to test whether the quality data follows normal

distribution. Finally, the coefficient of variation should be calculated, which represents

the ratio of the variations of the quality characteristics relative to their means.

4.3 The Experiment Result

During the experiment, we first found that the part quality was more sensitive to

lubrication than to the blank holder force. To show the effect of lubrication, two blank

disks were drawn at 3000lbs and 4500lbs BHF without the lubrication. Another two disks

were drawn at 3000lbs and 4500lbs with drawing oil applied. Their shapes are shown in

the figure 4.5. From these pictures, we see that when no lubrication was applied, even

66
reducing the BHF from 4500lbs to 3000lbs could not make a good cup without fracture.

Once lubricant was used, at both BHFs, no fracture appeared. The only difference was

that under smaller BHF, the wrinkling magnitude was higher.

To keep a compromise between wrinkling and fracture, we fixed the blank holder force at

4500lbs, applied lubricant (drawing oil), and then drew 30 pieces of blanks to the depth

of 44mm. The part shape after drawing is shown in figure 4.6(a). Within those 30 cups,

we found 3 of them were defective due to fracture, which are shown in figure

4.6(b)(c)(d).

4.3.1 Wrinkling Measurement Result

First we examine the wrinkling measurement result. The numbers of wrinkles along the

flange for the thirty pieces of drawn cups are plotted against test number in figure 4.7(a).

The histogram of those thirty data is shown in figure 4.7(b). Beside the histogram we

could see the calculated mean and standard deviation. The ratio between them (standard

deviation divided by mean), called coefficient of variation (COV), gives the level of

wrinkling variation under fixed process setting (BHF=4500lbs, drawing depth=44mm,

same drawing oil applied by the same pattern for every cup drawing). From the figure,

the COV for number of wrinkling is 1.737/25.53=6.8%. The figure 4.7(c) shows the

normality test plot. The p-value is 0.087. Therefore at the 95% confidence level, we

could say the number of wrinkling follows roughly normal distribution.


67
(a) BHF=3000lbs, without lubrication (b) BHF=4500lbs, with lubrication

(c) BHF=3000lbs, with lubrication (d) BHF=4500lbs, with lubrication

Figure 4.5. Parts drawn without and with lubricants at two levels of BHFs.

68
(a) the part shape after drawing (b) defective part 1 with fracture

(c) defective part 2 with fracture (d) defective part 3 with fracture

Figure 4.6. The parts after drawing and the defective parts.

69
Plot of wrinkling number vs. test number

29

28

27

wrinkling number
26

25

24

23

22

3 6 9 12 15 18 21 24 27 30
Index

(a) plot of number of wrinkling vs. test number


Histogram of wrinkling number
Normal
9

8 Mean 25.53
7
StDev 1.737
N 30
6
Frequency

0
22 23 24 25 26 27 28 29
wrinkling number

(b) histogram of the number of wrinkling and the fitted normal distribution
Probability Plot of wrinkling number
Normal
99

95
90

80
70
Mean 25.53
Percent

60
50
40 StDev 1.737
30
N 30
20
AD 0.638
10
P-Value 0.087
5

1
21 22 23 24 25 26 27 28 29 30
wrinkling number

(c) normality test of the number of wrinkling


Figure 4.7. The measurement result for number of wrinkling along the flange.

70
Then we will check the variation of the maximum height of wrinkling mark along the cup

sidewall. The maximum heights of wrinkling mark along cup sidewall are plotted against

the test number, shown in figure 4.8(a). The histogram with normal distribution fit is

shown in figure 4.8(b) with the mean of 0.5308 and the standard deviation of 0.06093.

Therefore, the COV for the wrinkling height is 0.06093/0.5308=11.5%. The normality

test plot is given in figure 4.8(c). At 95% confidence level, we could say the wrinkling

height follows normal distribution since the p-value 0.11 is larger than 0.05. From the

results of number of wrinkling and height, we can see that the part quality in terms of

wrinkling has large variation. Assuming the drawing process is stable (plus/minus 3

sigma), the number of wrinkling from the production will range from 25.53-3*1.737

(around 20) to 25.53+3*1.737 (around 31). You could imagine how different the drawn

parts may looks like.

4.3.2 Fracture Measurement Result

The fracture tendency is represented by measuring the minimum sheet thickness along

the cut open cup sidewall. The plot of minimum thickness against the test number is

shown in figure 4.9(a). The histogram and the normality test are shown separately in

figure 4.9(b) and (c). From the figures, we see that the thickness does not follow normal

distribution since at 95% confidence level the p-value from normality test 0.018 is much

smaller than 0.05. However, we could still use COV to show its magnitude of variation,

which is equal to 0.002383/0.03418 (around 7%).


71
Plot of wrinkling length vs. test number
0.70

0.65

0.60

wrinkling length
0.55

0.50

0.45

0.40
3 6 9 12 15 18 21 24 27 30
Index

(a) plot of maximum height of wrinkling mark along cup sidewall vs. test number
Histogram of wrinkling length
Normal

9
Mean 0.5308
8
StDev 0.06093
7
N 30
6
Frequency

0
0.40 0.44 0.48 0.52 0.56 0.60 0.64 0.68
wrinkling length

(b) histogram of maximum height of wrinkling mark along cup sidewall with normal
distribution fit
Probability Plot of wrinkling length
Normal
99

95
90

80
70
Percent

60
50
Mean 0.5308
40 StDev 0.06093
30
20
N 30
10
AD 0.598
5 P-Value 0.110

1
0.40 0.45 0.50 0.55 0.60 0.65 0.70
wrinkling length

(c) normality test of maximum height of wrinkling mark along cup sidewall
Figure 4.8. The result for the maximum height of wrinkling mark along cup sidewall.

72
Plot of thinning vs. test number

0.038

0.037

0.036

0.035

thinning
0.034

0.033

0.032

0.031

0.030

0.029
3 6 9 12 15 18 21 24 27 30
Index

(a) plot of minimum thickness along cup sidewall vs. test number.

Histogram of thinning
Normal

6
Mean 0.03418
5 StDev 0.002383
N 30
4
Frequency

0
0.030 0.032 0.034 0.036 0.038
thinning

(b) plot of histogram of minimum thickness along cup sidewall and fitted normal
distribution
Probability Plot of thinning
Normal
99

95
90

80
70
Percent

60
50
40 Mean 0.03418
30
20
StDev 0.002383
10
N 30
5 AD 0.910
P-Value 0.018
1
0.0300 0.0325 0.0350 0.0375 0.0400
thinning

(c) normality test of minimum thickness along cup sidewall.

Figure 4.9. The fracture measurement result.

73
Again, if we assume the process is stable (plus/ minus 3 sigma), the minimum thickness

will range from 0.03418- 3*(0.002383) = 0.027031 to 0.03418+ 3*(0.002383) =

0.041329.

4.3.3 Springback Measurement Result

The springback is measured by the angle illustrated in figure 4.4. The plot of angle vs.

test number is shown in figure 4.10(a). The histogram of angles and fitted normal

distribution is in figure 4.10(b) with the mean of 0.7961, and the standard deviation of

0.2188. From the normality test plot in figure 4.10(c), we could conclude that the

springback angles follow normal distribution since at 95% confidence level, the p-value

0.298 is much larger than 0.05. The COV for the angles is equal to 0.2188/0.7961=27%.

One common known issue of HSLA sheet material is the high springback. From this

experiment, we also see that the variation of springback (27%) is very high comparing to

wrinkling (6.8%, 11.5%) or fracture variations (7%).

74
Plot of angle vs. test number

1.3

1.2

1.1

1.0

0.9

angle
0.8

0.7

0.6

0.5

0.4
3 6 9 12 15 18 21 24 27 30
Index

(a) plot of springback angle vs. the test number


Histogram of angle
Normal

7
Mean 0.7961
6 StDev 0.2188
N 30
5
Frequency

0
0.4 0.6 0.8 1.0 1.2
angle

(b) plot of histogram of springback angle with normal distribution fit


Probability Plot of angle
Normal
99

95

90

80
70
Percent

60
50 Mean 0.7961
40
30
StDev 0.2188
20 N 30
10 AD 0.424
5 P-Value 0.298

1
0.2 0.4 0.6 0.8 1.0 1.2 1.4
angle

(c) normality test of springback angle

Figure 4.10. The springback angle measurement result.

75
4.4 Experiment Conclusion

In this simple cylindrical cup drawing test, we first found the optimal process setting to

deep draw the cup. The lubricants must be used to prevent the fracture. Thirty cups were

drawn at blank holder force 4500lbs, which gave the best compromise between wrinkling

and thinning. Then we measured different metrics to represent the three aspects of the

drawn cup quality: the wrinkling, fracture and springback. The number of wrinkling

along the flange and the maximum height of wrinkling mark along the cup sidewall

represent the wrinkling tendency. The angle from the cup bottom to top represents the

springback tendency. We then cut those thirty cups open and then measured the minimum

sheet thickness along the cup sidewall. This represents the fracture tendency. After the

data are analyzed, we found all of these metrics have large variation. Their coefficients of

variation range from 6.8% to 27%. The fracture metric does not follow normal

distribution, while the other metrics seem following normal distribution.

Through this simple drawing test, we not only justified the necessity of the probabilistic

deign approach by the large product quality variations observed, but also get some

quantitative information about the magnitude of those variations.

76
CHAPTER 5

5. PROBABILISTIC DESIGN OF ALUMINUM SHEET

DRAWING FOR REDUCED RISK OF WRINKLING AND

FRATURE

The exclusion of inherent process variations in the current deterministic design methods

for sheet metal forming can lead to very unreliable result that may cause high scrap rate,

frequent rework, machine shut down and thus huge loss of profit. In this chapter, a

general approach is presented to quantify the uncertainties and to incorporate them into

RSM model so as to conduct probabilistic based optimization. As an application, deep

drawing process of Hishida part is analyzed. Given the blank shape and tooling, a

probabilistic design is successfully carried out to find the optimal combination of blank

holder force and friction coefficient under the presence of variation of material

properties.

The result shows that by the probabilistic design, the quality index (average defect rates

of wrinkling and fracture) improved (reduced) 42% over the traditional deterministic

77
design. It also shows that by further reducing the variation of friction coefficient to 2%,

the quality index will improve to 98.97%. In a mass production environment, this

achieved quality improvement is huge.

5.1 Introduction

Deep drawing is a process transforming flat sheets into cup or box shaped articles

without fracture or excessive localized thinning. Enormous efforts have been put into the

system design phase in order to produce a defect-free part. In the previous work, given

material and part, the optimum deep drawing design, in general, focuses on three main

aspects: blank design, tooling design and process design. The blank design includes

optimizing blank geometries and thickness [27]. The tooling design is to determine the

optimal punch and die radii, the punch and die clearance, the drawbead shape and

location [28] [29]. The process design tends to find the best setting of process parameters

such as the friction (lubricant type and lubrication procedure), the punch speed and the

blank holder force.

However, even if the drawing process is optimally designed, the part can have significant

scatter in dimensions and properties, as shown in the simple cup drawing test mentioned

in chapter 4. Majeske [2] analyzed data from several leading automobile manufacturers,

and highlighted the fact that the high scrap rate still remains as a predominant issue

78
especially in making complex shaped part for them. He reported that within the same

batch (same die and process setup) the part-to-part geometric variation can be as high as

30%. This high variation inevitably causes high scrap rate, frequent rework, machine shut

down and huge loss of profit.

Of course there are many factors such as inadequate or inaccurate modeling, unknown

failure mechanism, unpredictable human effect etc. causing the part fluctuation, but

variation that occur in the forming of each part is a significant and yet not well

recognized reason for the lack of process consistency.

Recently, a few researchers have investigated process variability. Gantar [30] identified

twelve most important influencing input parameters in the deep drawing process and

measured their variations. Karthik [31] investigated the variability of sheet material

properties and carried out measurements using more than 45 coils of same material. It

was shown that the strain hardening coefficient “ n ” can have coil-to-coil variation up to

14%. To give an idea about how much the material property can vary the result from

Gantar and Karthik is listed together in table 5.1. Cao [32] also pointed out that during

sheet bending process the variation of material strength “ K ” can be as high as 20%, the

strain hardening coefficient 16% and the friction coefficient 65%.

79
Extensive research has been done in exploring the deterministic effect of each factor on

the part, but the impact of their variations on the fluctuation of output quality for the

drawing process is seldom addressed and quantified. The inclusion of uncertainty in the

design and optimization cycle should lead to better understanding of the impact of

uncertainty associated with system input on the system output. This understanding can

then be applied for managing such uncertainties. Such designs that incorporate

uncertainty are more popularly referred to as probabilistic design [33]. To date, there is

no report on the probabilistic design of aluminum sheet deep drawing process.

80
ST-13[2] Stainless steel 409

Values Variation(%) Values Variation(%)

n 0.22 ± 0.036 16.3 0.2 ± 0.03 13.4

K 546 ± 50 9.1 751 ± 43 5.7

r 1.90 ± 0.20 10.5 1.5 ± 0.12 7.9

Table 5.1. Variation of material properties [28][29]

81
5.2 General Approach for Probabilistic Design

The concept of reliability based design was first introduced and developed in the

reliability calculations of structures. A limit state function, basically a failure criterion

with a deterministic model connecting the system input variables to the output variables,

will segment the multi-dimensional space spanned by the random variables into the

failure and safe domain. Given the joint probability density function of the random

variables, the probability of failure can be calculated by the portion of points in the

failure domain [34]. Each interested response will have its own limit state function. Then

the optimization is to search for the right means of random variables that minimize the

objective function while satisfying the reliability requirements.

Two problems need to be resolved when applying probabilistic design in the metal

forming process: construction of the limit state function and the joint probability. Since

metal forming is a highly nonlinear process influenced by many factors, there is no way

to analytically predict the system response. Nowadays, numerical simulations of sheet

metal forming processes based on finite element method have become a powerful tool to

predict the forming process. However, simulating each point according to the joint

probability to check its failure or safe is impossible. Even by some sampling technique,

this task is still too time-consuming. The only practical solution to overcome this

82
problem is to build the meta-model by the response surface method (RSM) to

approximate the limit state function. Through proper design of experiments (DOE), only

fraction of experiments or simulations is required. Then if the joint distribution of

random variables is known, statistical method like Monte Carlo Simulation (MCS) can

evaluate the probability of failure for the interested responses [35]. The second difficulty

is to determine the joint distribution. Since the experimental data available for statistical

modeling of the random variables in metal forming are limited, it is not possible to make

any clear identification of correlation between the variables. Therefore, the independence

between them is usually assumed and accordingly the joint probability density function is

directly established as a product of the marginal density functions.

5.3 Quality Index to Measure Risk

As mentioned before, drawing process can be affected by the selection of blank, tools and

process parameters for given material and part. In traditional design, those factors are

taken as certain variables, meaning they don’t have random features at all. However,

many of them are indeed random variables. For example, the sheet metal material

properties can vary coil by coil or stock by stock. The lubrication condition for one part

can also be quite different from another.

83
To illustrate the necessity of incorporating uncertainties into the traditional optimum

design, the deep drawing process (figure 4.1) is schematically described where the

variations of input parameters are linked somehow to the fluctuation of output quality.

To date, there is no literature found on probabilistic design application for aluminum

sheet deep drawing processes. However, Sahai [33] adopted a very similar strategy in

designing a sheet metal flanging process. In his paper, he treated the sheet thickness “t”

and gap “g” as random design variables, the die corner radius “r” as a deterministic

design variable and the Young’s modules “E” and Yield Stress “Y” as random parameters.

Then the objective was to find a combination of sheet metal and tooling configurations

that would minimize the difference of springback to the target under the probabilistic

constraint that 99.99 percent of maximum absolute strain of the flanged sheet metal

should not exceed a specified value.

The issue is that being the quality feature, the springback should have a scatter because of

the variation of the system. Therefore, the intuitive objective should be minimizing the

probability of sheet springback angle exceeding the tolerance, namely the defect rate

instead of the deterministic objective function. Nevertheless, changing the objective to be

probabilistic will invalidate the optimization method he introduced. Additionally, the

author simply assumed the variation of random variables and did not investigate the

effect of enlarging or reducing the variable variations on the defect rate, which will be

tackled in this chapter.

84
In this chapter, the wrinkling and fracture are selected as the quality features. A new

measure, quality index, is established as the weighted sum of probability of no wrinkling

and probability of no fracture. It is written as:

QI = weight * Pr[no wrinkling] + (1 − weight ) Pr[no fracture] (5.1)

Then the objective is naturally to maximize the quality index. The outline of used

strategy is drawn in figure 5.1. A non-symmetrical Hishida geometry, which has four

corners with different radii and four tapered walls with different sloped angles on each

side, is chosen to demonstrate the concept.

5.4 The Numerical Simulation Model

The industrial/simulation model for the Hishida parts forming process is shown in Figure

5.2. The aluminum car body sheet alloy Ecodal-608-T4 (AA6181A) with 1mm thickness

is used for this study. The Krupkowsky-law hardening equation: σ y = K (ε 0 + ε p ) n is

used for modeling the material behavior. The drawing process is simulated using

PAM-STAMP.

85
5.5 Selection of Input Variables

Siekirk [36] has identified more than 25 variables that influence sheet metal forming.

However, it is impossible to include all the variables in the model. Instead, it will assume

that the blank shape and tooling have been designed and produced. The interests are

hence the optimal process parameters. After checking the sensitivity analysis result from

Gantar [30] and Jaisingh [37], the input variables are selected and classified into three

categories which are listed in table 5.2. The deterministic design variable is the blank

holder force – BHF. It has no variation since it can be easily close-loop-controlled. The

random design variable is the friction coefficient between blank/die and blank/binder –

Lub1, and the friction coefficient between blank/punch – Lub2. A uniform lubrication

over the whole blank surface is presumed. Although it is by and large the lubricant type

and lubrication procedure that decide the friction coefficient, many other factors like the

blank or tool surface roughness, sliding velocity and contact pressure [38] also have

significant effects on it. Cao [32] estimated that in sheet metal forming, the variation of

friction coefficient can be as high as 65%. Gantar [30] used 10% in his stability

evaluation of deep drawing process. In this probabilistic design problem 10% is used

initially. The random parameters that cannot be designed (noise variables) are the strain

hardening coefficient n , the strength coefficient K , blank dimension variation along the

horizontal and vertical direction - pm1 , pm 2 .

86
5.6 Prediction of Wrinkling and Fracture

There are two forms of wrinkling, the sidewall wrinkle (SW) and flange wrinkle (FW).

Since the flange area is trimmed after drawing, only sidewall wrinkling is measured by

geometrical consideration of part cross-sections cut into planes perpendicular to the

forming direction at a 25% parts height [39], as shown in figure 5.3.

The criterion for wrinkle defect is:

{wrinkling : if normalized magnitude of wrinkle > 1} (5.2)

Fracture happens when the strain at the local region is concentrated and subsequent

deformation changes from a smooth and continuous one to a markedly non-uniform one.

In this study, the maximum sheet thinning is measured to indicate the risk of fracture. The

criterion for fracture defect is:

{ fracture : if normalized maximum sidewall thinning > 1} (5.3)

87
Figure 5.1. General probabilistic design approach.

88
Figure 5.2. Simulation model of the Hishida forming process.

89
Designable signal variables Un-designable noise
variables

BHF Lub1 Lub2 n K pm1 pm 2


Variation − 10% 10%
13% 6% 5% 5%

Low(DOE) 495 0.05 0.05


0.2375 446 0 0

High(DOE) 605 0.12 0.12


0.2625 493 1 1

Table 5.2, The variation and DOE levels for selected parameters

90
5.7 DOE and RSM

For the seven variables, a Box-Behnken DOE is selected. This design allows efficient

estimation of the first and second order coefficients. Also because Box-Behnken has

fewer design points, it is less expensive to run than other design methods such as central

composite designs under the same number of factors. After a series of finite element

simulations, the data of magnitude of SW wrinkling and maximum thinning are collected

and fitted by the second order polynomials. The R 2 for them are 94% and 98%

respectively. To check the accuracy of the response surface models, ten additional

simulations are run at different points. The responses obtained by the simulations with

PAM-STAMP are compared with the interpolation based on response surface models.

The results correlate very well. The deviations in the response surface predictions are

within a few percent. Hence, the response surface models are considered to be

sufficiently accurate for the subsequent probabilistic design of the aluminum sheet deep

drawing process.

In the table 5.1, the variation data are for steel material. In this DOE and research we

used Aluminum material. The reason is that it is more difficult to draw an Aluminum part

without defects. However, we have conduct a testing to see whether the RSM generated

from the Aluminum material DOE could be used to predict the behavior of the steel

91
material (ST-12), since the thinning and wrinkling are mostly determined by the n and K

if we keep other process parameters the same. After testing, we found the Aluminum

RSM could predict the steel material very well. This means that the response for thinning

and wrinkling are monotonous. As long as the ranges of n and K are not far away from

the DOE range, the prediction is acceptable. Please see Appendix B for details.

Before proceeding to the probabilistic assessment, the main effect plot is drawn to show

how sensitive each of the variables on the two responses, see figure 5.4(a) and (b). It is

observed that the friction coefficient between die/blank and binder/blank has the most

significant impact on the wrinkling and fracture. When it becomes large, the magnitude

of wrinkling decreases whereas the maximum thinning increases. The explanation is that

due to the increased friction coefficient, larger friction force under the same blank holder

force will resist the material flowing into the wall area, and therefore change the strain

distribution in the FLD to the area where wrinkling is less likely to occur. At the same

time, more punch force is needed to draw the material into the die. With less material

going to the wall the maximum thinning along the sidewall is inevitable. The blank

holder force, which is the second largest factor to the wrinkling and facture defects, can

be explained similarly. Other observations include: (1) the geometric variations along the

initial blank are not sensitive so that they can be neglected in the following analysis. (2)

The strain hardening coefficient and the strength coefficient have effect on the quality

features but are limited comparing to friction coefficient and BHF.

92
Figure 5.3. The measurement of sidewall wrinkling.

93
Sensitivity of parameters to the magnitude of wrinkling

Lub1

BHF K
n Lub2 pm1 pm2

(a), The sensitivity plots for the wrinkling.

Sensitivity of parameters to the maximum thinning

Lub1

BHF K Lub2
n pm1 pm2

(b), The sensitivity plots for the thinning.

Figure 5.4. Sensitivity plot of wrinkling and thinning.

94
5.8 Probabilistic Assessment for Wrinkling and Fracture

The response surface models built in section 5.4 are deterministic in nature. Given a

design vector, the prediction is one certain value. This is unrealistic since many variables

contained in the model, like material constants, friction coefficients and geometric

dimensions, etc. are known to have a certain degree of scatter around their nominal

values. The variation of system input variables should be incorporated into the

deterministic model.

To incorporate uncertainties, it is assumed that the vector of random variables

Ψ = [ X1 , X 2 ,..., X n , Z1 , Z 2 ,...Z m ] representing the aluminum sheet forming process random

nature. X refers to the designable random variable while Z refers to the un-designable

random parameters like n and K . The realization of Ψ is ψ = [ x1, x2 ,...xn , z1 , z2 , ...zm ]

representing a specific metal forming process and the joint probability density function

f Ψ (ψ ) measures the probability of that event. Suppose g wr ( Ψ ) is the response model for

the wrinkling and we know the criterion for wrinkling defect is

{wrinkling : if normalized magnitude of wrinkle > 1} ,then the probability of wrinkling can be

calculated by

Pfwr = P{Ψ ∈ Ω fwr } = P{Norm( g wr (ψ )) − 1 ≤ 0} = ∫Ω f (ψ ) dψ


fwr Ψ (5.4)

95
where Ω fwr is defined as Ω fwr = {ψ : Norm( g wr (ψ )) − 1 ≤ 0} . The evaluation of the

probability of fracture is conducted in the same way. For the probabilistic design, it is

assumed that n ~ N (0.245,0.032 2 ) and K ~ N (470,30 2 ) . The standard deviation for

friction coefficient is 0.01. All the random variables are normal distributed and there is no

correlation between them. Hence, in every iteration of the probabilistic optimization,

Monte Carlo Simulation (MCS) can be applied to evaluate the quality index for each

design vector.

5.9 Optimal Design Approach

In previous section it has already given the criteria of defining wrinkling and fracture, the

two major forms of defects in the manufacturing of Hishida parts. The optimal design

objective, therefore, is to find the right combination of BHF, Lub1 and Lub2 so as to

minimize the sum of two defects rates or maximize the quality index. There are two

approaches to achieve this objective.

The traditional way formulates the problem as

minimize {weight * Norm( Magnitude of wrinkling )


(5.5)
+ (1- weight ) * Norm( Maximum thinning )}

96
where weight is a coefficient between 0 and 1 that is decided by the significance of each

defect. Norm() is the function to normalize the magnitude of wrinkling and maximum

thinning so that they will have the same range during the optimization to avoid severe

bias.

The other approach that is adopted in this chapter captures the variations and expresses

the objective function in the form of probabilistic way as shown below:

maximize{weight * Prob[ no wrinkling ]


(5.6)
+ (1−weight )*Prob[ no fracture]}

where prob[] is the probability assessment function described in section 5.5. The first

optimal design approach is called “deterministic design” (DD) and the latter one

“probabilistic design” (PD).

For the PD two optimization techniques: the gradient based method and grid method are

tried. In the first method, for each iteration, several MCSs are conducted to calculate the

gradient at the current design point to determine the future search direction. It will

converge when the residual error is small enough. The second technique will segment the

design space into equal lattice and each intersection point will be evaluated and the best

design point who gives the largest quality index will be selected. It is found that the first

method is not very reliable since it is easy to be trapped in the local maxima. Although

the second method needs a much larger CPU time, for the sake of finding the global

97
maxima, the second method is used and all the results are reported by this method. To

illustrate the grid method, when the weight is 0.5, first the mean of Lub1 is fixed as 0.09

and mean of Lub2 as 0.12, and then the quality index is plotted against the different BHF,

see figure 5.5. In this case, the optimal BHF is easily determined at 605. In figure 5.6, the

mean of Lub2 is fixed as 0.12 and the BHF as 605, and then the quality index is plotted

against the different mean of Lub1. From the plot, the optimal mean of Lub1 is 0.09.

Based on the different value of weight, a series of optimization using both the DD and

PD operations is conducted. The results are tabulated in table 5.3 and 5.4.

First of all, it is seen that the deterministic design result is very different with

probabilistic design at the same weight. DD is prone to the extremes on the constraints

boundary. The BHF is either the minimum or the maximum. The friction coefficient also

has the similar pattern. The outcome is that either its probability of wrinkling is very

large or the probability of fracture is very large. In contrast, the PD makes a compromise

between the wrinkling and fracture very well. The weighted sum of the probabilities of

two defects is kept at a much lower level than the DD. When the weight equals 0.5, the

quality index for DD is 52.445% and the index for PD is 94.815%. The PD is much

superior to the DD. Secondly, from figure 5.5 and 5.6, it is found that the key to improve

the quality is the right design of lubrication between blank/die and blank/binder.

Considering the material properties’ variation, the quality index is more sensitive to

friction coefficient than the blank holder force.

98
Design of the optimal BHF
2

1.8
2*Quality index (weight=0.5)
1.6
Optimal BHF
1.4

1.2
Probability of no fracture
1

0.8

0.6 Probability of no wrinkling


0.4

0.2

0
480 500 520 540 560 580 600 620
BHF (mean of lub1=0.09, mean of lub2=0.12)

Figure 5.5. Probabilistic design of BHF.

99
Design of the optimal mean of Lub1
2
2*Quality index (weight=0.5)
1.8

1.6
Optimal mean of Lub1
1.4

1.2

1
Probability of no fracture
0.8

0.6

0.4
Probability of no wrinkling
0.2

0
0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12
Mean of Lub1(BHF=605, mean of Lub2=0.12)

Figure 5.6. Probabilistic design of mean of Lub1.

100
Weight wrinkle Weight fracture Lub1: binder/blank Lub2: punch/blank Pr[no Pr[no
BHF wrinkle]% fracture]%

0.1 0.9 0.05 0.12 495 0.22 100


0.2 0.8 0.05 0.12 605 4.89 100
0.3 0.7 0.05 0.12 605 4.89 100
0.4 0.6 0.05 0.12 605 4.89 100
0.5 0.5 0.05 0.12 605 4.89 100
0.6 0.4 0.09 0.12 605 93.78 95.85
0.7 0.3 0.12 0.12 605 99.98 42.63
0.8 0.2 0.12 0.12 605 99.98 42.63
0.9 0.1 0.12 0.12 605 99.98 42.67

Table 5.3. Deterministic design result

101
Weight Weight Lub1: binder/blank Lub2: punch/blank Pr[no Pr[no
wrinkle fracture BHF wrinkle]% fracture]%

0.1 0.9 0.08 0.12 605 77.06 99.03

0.2 0.8 0.09 0.12 590 91.16 96.80

0.3 0.7 0.09 0.12 605 93.78 95.85

0.4 0.6 0.09 0.12 605 93.78 95.85

0.5 0.5 0.09 0.12 605 93.78 95.85

0.6 0.4 0.1 0.12 565 96.73 91.48

0.7 0.3 0.1 0.12 585 98.20 89.17

0.8 0.2 0.1 0.12 595 98.52 88.23

0.9 0.1 0.1 0.12 605 98.93 86.50

Table 5.4. Probabilistic design result

102
5.10 Effect of Variation on The Optimum Design

In the previous probabilistic design, it has assumed that the variations of friction

coefficient, strain hardening coefficient and strength coefficient are 10%, 13% and 6%.

The real spread during manufacturing may be much larger. So in this section, the effect of

variations of random variables on the quality index to the optimum design will be

investigated.

The mean values and type of probability density function for the random variables were

kept constant, while the percent of variation (and accordingly the standard deviations)

were increased at different levels from 2% to 30%. Using MCS, the quality index at each

level can be evaluated. The relative importance of the various random variables is plotted

in figure 5.7. It is observed that the different level of variation of material properties does

not have effect on the quality index (also the defect rates). However, the variation of

lubrication can cause the quality index drop from 98.97% to 77.41% if its variation

increases from 2% to 30%.

103
Quality index vs. variation of random variables
1

0.95

Friction coefficient
0.9 strain hardening n
strength K

0.85

0.8

0.75
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Percent of Variation [% ]

Figure 5.7. Effect of variation of random variables on the quality index.

104
5.11 Conclusion

In this chapter, deep drawing process of Hishida part is analyzed. Given the blank shape

and tooling, a probabilistic design is successfully conducted to find the optimal

combination of blank holder force and friction coefficient under the presence of variation

of material properties. The result shows that by the probabilistic design, the quality index

(average defect rates) improved 42% over the traditional deterministic design. It is also

noticed that by further reducing the variation of friction coefficient to 2%, the quality

index will improve to 98.97%. In a mass production environment, the achieved quality

improvement is huge.

105
CHAPTER 6

6. INVESTIGATING RELIABILITY OF TEMPORAL

VARIABLE BLANK HOLDER FORCE CONTROL IN

SHEET DRAWING UNDER PROCESS UNCERTAINTIES

The blank holder force, which regulates the amount of metal drawn into the die cavity,

stands to be a very effective measure for the success of deep drawing. It has been shown

that properly designed temporal varying BHF profile can make a part with fewer defects.

Extensive research has been carried out in the determination of this optimum profile

usually in a deterministic way that does not consider the inherent process variations. This

exclusion of variations, however, leads to unreliable design. In this chapter, a

probabilistic design approach is presented to incorporate these variations. As a

demonstration, the cylindrical cup drawing process is analyzed. Under the presence of the

variation of sheet thickness and frictions between sheet/die/binder/punch, the

probabilistic design successfully finds the optimal variable BHF. The result shows that by

the probabilistic design, the yield (probability of good parts) improved to 99.98% from

48.04% by the traditional deterministic design. In mass production environment, the

achieved process robustness is huge.


106
6.1 Introduction

Deep drawing is a process transforming flat sheets into cup or box shaped articles

without fracture or excessive localized thinning. It is one of the most common processing

techniques utilized in mass production. Nowadays, more and more complex parts are

being deep drawn. A proper designed drawing process is crucial in order to produce a

defect-free part. In the previous work, given material and part, the optimum deep drawing

design, in general, focuses on three main aspects: blank design, tooling design and

process design. The blank design includes optimizing blank geometries and thickness.

The tooling design is to determine the optimal punch and die radii, the punch and die

clearance, the drawbead shape and location. The process design tends to find the best

setting of process parameters such as the friction (lubricant type and lubrication

procedure), the punch speed and the blank holder force.

Within all aforementioned design variables, the blank holder force, which regulates the

amount of metal drawn into the die cavity, stands to be a very effective measure for the

success of deep drawing. During drawing, the BHF could be kept constant or varied with

time. The advantage of constant BHF control is its simplicity, but practices do prove that

properly designed variable BHF profile, i.e. variation of BHF with punch stroke, can

make a part with less defects and higher quality. This improvement is due to the fact that

it allows the BHF to vary according to the status of stress state in the sheet so that the

optimal process conditions are maintained at each time step.

107
To determine a good variable BHF profile, an experimental or computational approach is

usually used. Hirose [40] examined the blank holder profile experimentally while

forming full scale automobile panels. He found the profiles that did succeed were the

ones starting at lower binder forces and then increasing during the latter stage of the

process. Hardt [41], instead of figuring out the profiles directly, developed a PI controller

to adjust the blank holder force in-process to ensure a previously determined optimal

forming-punch force trajectory or a normalized average thickness trajectory was followed

and replicated. However, to obtain that target punch force profile, a constant BHF was

applied in his experiment. These approaches, although giving acceptable results, require

extensive trial and error, and consume too much material and time. Therefore, the

computational approach is preferred. Sim and Boyce [42] first predicted the optimal BHF

profile for a round cup drawing through a closed-loop controlled simulation system using

the same idea as in [41]. The issue is that since the sidewall wrinkling was not considered

in the controller, the maximum cup height might have been overestimated. Cao and

Boyce [43], nevertheless, included both the wrinkling and fracture criteria in their work.

By applying a similar PI control strategy and using the major principal strains and

amplitudes of wrinkles at the location of the die radius as state variables, a variable BHF

profile for conical cup drawing was predicted and the failure-free drawing depth was

increased over that achieved with a constant BHF. Sheng [44] further improved the

optimal BHF profile prediction by taking the maximum thinning and flange wrinkling, as

well as sidewall wrinkling amplitudes as control indices. His adaptive simulation method

108
can give optimum variable BHF profile for drawing conical cups with acclaimed

increased failure-free drawing depth and uniformity of the wall thinning distribution.

Although extensive research has been done in the exploration of the optimum variable

BHF in the deep drawing process, researchers have seldom included the existing sheet

and process variations in their models and discussed the impact on the part quality.

In fact, it is observed that even if the drawing process is optimally designed, the produced

part can still have significant scatter in dimensions and properties. Majeske [2] analyzed

data from several leading automobile manufacturers and highlighted the fact that the high

scrap rate still remains as a predominant issue especially in making complex shaped part.

He reported that within the same batch (same die and process setup) the part-to-part

geometric variation can be as high as 21%.

To date, most predictions of variable BHF profiles by the computational approach

basically exclude these possible sheet and process variations and assume them at constant

values. This kind of design without uncertainties or variations is characterized as

deterministic design.

Although many authors have verified their BHF profile through experiments and

concluded that the design improved the drawing quality, however, these experiments

were conducted at the well controlled laboratory conditions. There may be high risk in

109
real production. The reason is that this deterministic optimization tends to push the

design towards one or more constraints boundaries until the constraints are active, thus

leaving the designer with a design for which even slight uncertainties in the problem

formulation or changes in the operating environment could produce failed products.

Therefore, the inclusion of unavoidable variations in the design of optimum variable

BHF should lead to better understanding of the impact of uncertainty associated with

system input on the system output. This understanding can then be applied for managing

such uncertainties. Such designs that incorporate uncertainty are more popularly referred

to as probabilistic design [33].

6.2 General Approach for Probabilistic Design Optimization

The concept of including uncertainties into design was mainly developed from the

communities of structural reliability, which focuses on accessing the probability of failure

of a design with respect to the structural constraints and evaluates the variation of these

constraints. A limit state function, basically a failure criterion with a deterministic model

connecting the system input variables to the output variables, will segment the

multi-dimensional space spanned by the random variables into the failure and safe

domain. Given the joint probability density function of the random variables, the

probability of failure can be calculated by the portion of points in the failure domain.

110
Thus a deterministic constraint can be converted to probabilistic constraint by this

approach.

Reliability-based optimization is then searching for the right means of random design

variables that minimize the objective function while satisfying the reliability

requirements [34]. However, the objective function is evaluated at the mean value point.

The reliability refers to the constraints only and the objective itself is regarded as

deterministic. That is why it is named reliability-based optimization. Its focus is on

constraints, shifting constraint distributions away from the constraint boundaries, but not

on the spread of the objective response distributions and the possible variation reduction.

On the other hand, the robust design, first developed by Taguchi, aims at driving the

objective mean performance towards a target and minimizing the variance of objective

performance. The Taguchi method uses the “crossing design matrix”, “Signal-to-noise

ratio” and “Loss function” to evaluate potential designs and select the best alternative

from among those evaluated. However, the constraints are not formulated as typically

done with optimization formulations and the optimization is only performed at the

discrete design points. Reducing the objective response variation, balancing “mean on

target” and “minimize variation” are its interests. The term “robustness” in the

engineering design context is defined as the sensitivity of performance parameters to

fluctuations in design parameters, particularly uncertain design parameters.

111
The probabilistic design, however, combines the features of both the reliability-based

optimization and the robust design by incorporating input constraints (bounds of random

design variables), output constraints (reliability constraints) and objective robustness

(minimize mean and variation). Buranathiti [45] was the first to conduct a probabilistic

design of a wheelhouse stamping process in sheet metal forming by maximizing the total

mean values of margins to failure and minimizing the variance of the margin. It

efficiently took the process uncertainties into account and created a system-level robust

probabilistic design model. Weighted three-point-based method was used to estimate the

means and standard deviations for the quality features. Li [46] also proposed a similar

approach in the sheet metal forming design. However it utilized the dual response surface

models for those means and standard deviations.

The general formulation for the probabilistic design is given as follows:

Find the set of design variables X that:


k wμ wσ
Minimize : F( μ y ( X ), σ y ( X )) = ∑ [ i ( μ yi − Ti ) 2 + i σ y2i ]
i =1 sμi sσ i (6.1)
Subject to : gi ( μ y ( X ), σ y ( X )) ≤ 0
X L + nσ X ≤ μ X ≤ X U − nσ X

Here yi , i = 1,..k are the output responses that designer cares about. wμi and wσ i are the

weights and sμi and sσ i are the scale factors for mean and variation of performance

response yi which has the target Ti . Like the probability constraint for input vector X ,

gi is the probability constraint for performance response yi , which is used to


112
confine its distribution within the upper and lower specification limit. For the case in

which the mean performance is to be minimized rather than directed towards a target, Ti

will be put equal to zero.

The key to implement the probabilistic optimization mentioned above is to estimate the

mean and variance for all the performance responses yi at each evaluated design point

during the optimization iterations. Monte Carlo Simulation (MCS) is the most

straight-forward method to acquire the knowledge of the probability distribution of

responses of uncertain systems given distributions of inputs with high accuracy. The

procedure is to first sample the values of the random variables by the given probabilistic

distributions and then run the system simulations according to the sample points.

However, to get a good estimate of the mean and variance, researchers have shown that at

least one hundred sample points should be evaluated even with variance reduction

sampling techniques. Therefore, for instance, if our optimization requires 10 iterations,

the total number of metal forming simulations would be at least 1000. Considering the

computation and time effort for each nonlinear metal forming simulation, this method is

obviously not good. The practical solution to overcome this problem is to build a meta

model Y ( x) by the response surface method (RSM) to approximate the FEM simulation.

Through proper design of experiments (DOE), only fraction of simulations is required.

Then by the Propagation of Error, the mean of this performance response can be

calculated by setting the uncertain design parameters to their mean value,

113
1 m d 2Y 2
μ y = Y ( μ x ) + ∑ 2 σ xi (6.2)
2 i dxi

And the standard deviation of Y ( x) is given by the second order Taylor’s expansion:

2
1 m m ⎛ ∂ 2Y ⎞
2
m
⎛ ∂Y ⎞
σ y = ∑ ⎜ ⎟ (σ xi ) + ∑∑ ⎜
2
⎟⎟ (σ xi ) (σ x j )
2 2
(6.3)
i =1 ⎝ ∂xi ⎠ 2 i j ⎜⎝ ∂xi ∂x j ⎠

where σ xi is the standard deviation of the ith parameter and m is the number of

uncertain parameters which include designable random variables and un-designable

random variables.

6.3 Parameterization of The Variable Blank Holder Force

The probabilistic design of variable blank holder force cannot start from nowhere. In

reference [44], the author predicted its optimum profile for the drawing of conical cup in

a deterministic manner. In this chapter, however, this case is taken as an example to

investigate its performance under the presence of sheet and process uncertainties. To

illustrate the probabilistic design approach the same simulation model is adopted and it is

assumed that even when variations exist, the basic shape of the robust profile is the same

as the one determined in the reference [44]. Based on the features of the optimized

114
variable BHF in the referenced [44], the profile is characterized by two connected

Gaussian functions in the form of

x − μ1 x − μ2
−( )2 −( )2
σ1 σ2 (6.4)
y = I A1e + (1 − I) A2 e

where I is an indicator variable following the rule

⎧1 if x ≤ x*
I=⎨ (6.5)
⎩0 otherwise

where x* is the intersection point of the two Gaussian functions. After fitting the data, the

BHF curve can be approximated as

x −11.79 2 x − 27.7 2
−( ) −( )
y = I( x ≤ 16.3)348.9e 6.751
+ (1 − I( x ≤ 16.3))473e 13.12 (6.6)

The fitted curve is very close to the original optimized BHF in the reference [44], as

shown in the figure 6.1. The objective of parameterizing the predicted profile is to extract

the simple control parameters which can be used later in the design optimization.

115
Figure 6.1. The blank holder control profile from [44] and the fitted Gaussian

approximation for sheet drawing.

116
6.4 The numerical Simulation Model

The simulation model built in PAM-STAMP is shown in figure 6.2 and the model inputs

in table 6.1. This figure also includes the part drawn at the last step. The stroke is 47mm,

the same as in the reference [44]. The BHF profile in the simulation is inputted from the

fitted Gaussian curve. Due to the symmetry, only quarter of the tooling and part are

modeled. An elastic-plastic material model with isotropic hardening is considered. The

other inputs are kept the same as in the reference [44] and listed in table 6.1.

117
Binder Punch

Die

Figure 6.2. The FEM simulation model of the sheet drawing process and the final drawn

part.

118
Material AKDQ steel

Blank diameter 248mm/9.7in.

Blank thickness 0.86mm/0.034in.

σ = 795(0.0052 + ε )
0.2
Flow stress (MPa)

Friction coefficient Punch/blank,0.25;binder/blank,0.15;die/blank,0.15

Table 6.1. Circular blank dimensions, material properties, and friction coefficients used in
the simulation of drawing the conical cup from AKDQ steel.

119
6.5 Selection of Input Variables and Output Variables

Siekirk [36] has identified more than 25 variables that influence sheet metal forming.

However, it is impossible to include all the variables in the model. Instead, it is assumed

that the blank shape and tooling have been designed and produced. The interest is to find

the optimum process parameters, mostly referred to BHF, so that the drawing process is

desensitized to the unavoidable condition variations. After checking the result from

Gantar [30], Jaisingh [37] and Zhang [47], the process parameters initially selected for

sensitivity analysis are the punch speed: ps , friction between die/blank and binder/blank:

fd , friction between punch/blank: fp , and sheet thickness: t . To calculate the

sensitivity of each parameter, it is varied at three levels: low, medium and high while

keeping all other parameters at the medium level. There are three output responses of

interest: the maximum thinning y fr which represents the tendency to fracture, the sidewall

wrinkling ysw and the flange wrinkling y fw . The sensitivity is calculated by

⎡ yhigh
i
− ymed
i i
ymed − ylow
i ⎤
⎢ i i

1 3 ⎢ ymed ymed ⎥

6 i =1 ⎢ xhigh − xmed
+
xmed − xlow ⎥ (6.7)
⎢ ⎥
⎢⎣ xmed xmed ⎥⎦

where y i , i = 1, 2,3 corresponds to y fr , ysw , y fw . The results are tabulated in table 6.2.

120
Variable name Lower Middle Higher Sensitivity

Sheet thickness t 0.76 0.86 0.96 0.83

Punch speed ps 9 10 11 0.21

Friction fp 0.2 0.25 0.3 0.61

Friction fd 0.1 0.15 0.2 0.89

Table 6.2. The sensitivity of initial selected process parameter.

121
From it, it is seen that the punch speed only has little effect on the maximum thinning of

the final cup comparing to the other three variables. Therefore punch speed is ignored.

6.6 Deterministic Design vs. Probabilistic Design

There are four variables considered in the design of experiments for the further

construction of RSM and optimization. They are sheet thickness: t , friction between

punch/blank: fp , friction between die/blank and binder/blank: fd , and the scale of the

fitted variable BHF: s , which is in the range of zero to one to enlarge or reduce the blank

holder force profile accordingly. In probabilistic design, the variables considered in the

DOE and RSM are classified into control variables and noise variables. Control variables

are those process parameters that can be easily changed and controlled at a certain level.

However, it is very hard or even impossible to set or control noise variables in the

ordinary manufacturing conditions. To illustrate, the adjustment a technician makes to the

office copier belongs to control factor while the humidity that fluctuates in the office

environment is noise factor. Ideally, processes should be adjusted to be insensitive to the

noise factors via the control factors, so that variation of noise variables does not cause

variation of product quality. Back to the copier example, a good designed copy machine

should handle paper properly during humid summer months as well as in the winter,

when conditions are dry and prone to static electricity.

122
In the study, the objective is to consistently make conical cup without defects. Therefore

the variables are classified as following:

I. Control variable: the scale of BHF: s

II. Noise variables: the sheet thickness: t , the friction between punch/blank: fp ,

the friction between die/blank and binder/blank: fd .

It should be emphasized that this classification is nowhere unchangeable. If some noise

variable is found to be the key to the success of the objective, a method may be figured

out to control it. For example, if the humidity is found to play an important role in the

right functionality of copy machine, a humidifier may be added to control the humidity in

the machine. Then the former noise variable changes to control variable. Of course,

sometime it is also impossible or too expensive to change it.

For the design of experiment (DOE), the Box-behken DOE plan is selected. This design

allows efficient estimation of the first and second order coefficients. Also because

Box-Behnken has fewer design points, it is less expensive to run than other design

methods such as central composite designs under the same number of factors [48]. With

four variables, there are total 27 runs. For each run, the maximum thinning y fr , sidewall

wrinkling ysw and flange wrinkling y fw are recorded. The full quadratic response models

are built after all the twenty-seven simulations are done. The default values for t , fp , fd

in the reference [44] are used and the performance responses plotted versus s,
123
shown in figure 6.3. From this figure, it is seen that when the BHF is increased, the

maximum thinning is going up while the sidewall wrinkling is the opposite. In the

reference [44], the criteria for the fracture failure and wrinkling are maximum thinning

being larger than 0.25, sidewall wrinkling 0.20 and flange wrinkling 0.05. Since all the

flange wrinkle magnitude within design range is much smaller than 0.05, its curve is not

shown in figure 6.3. Constrained by the failure criteria, the feasible range for s shrinks

to a very narrow band around 1 as illustrated in the shaded area, which is exactly the

optimum BHF generated by the PI control strategy in the reference [44]. After examining

the control strategy, it is found that it tends to maximize the drawing depth at the expense

of letting wrinkling especially sidewall wrinkling approaches its limit. By reducing the

BHF to just keep the wrinkling magnitude on the limit bound, the maximum thinning can

be achieved at a lower level so that drawing depth is increased. From the deterministic

point of view, no part will fail although their sidewall wrinkle magnitude reaches the

maximum allowed value. However, as discussed previously, the sheet properties and

process variations are unavoidable. Even a minor change of s , t , fp or fd will make the

wrinkling or fracture out of bound easily. Suppose s has a normal distribution with

mean equals unity and other parameters are constant, the probability for wrinkling failure

then will be about 50%. Adding the effect of variations of other variables, the defect rate

may be even higher. Therefore, the deterministic PI control strategy is not robust, and

sometime it is even very dangerous. The benefits of it, though, come from its ability to

predict a rough profile which grasps the metal flow characteristics evolution during

forming.

124
Wrinkling/thinning magnitude
D bottom

Scale of BHF: s

Figure 6.3. Maximum thinning and sidewall wrinkling (y-axis) at different s of BHF

(x-axis).

125
The probabilistic based optimization then should be able to take the process variation into

account and improve its robustness based on the predicted deterministic BHF profile.

Based on the current failure criteria, the feasible design space is so narrow that s equals

unity is the only choice. The probabilistic constraints will never get satisfied because no

change can be made to s and thus no solution available for the probabilistic

optimization. After literature checking and discussion with the author of the reference

[44], it is found that the sidewall wrinkling criterion is too strict. Therefore, it was

relaxed somehow to enable the solution of the probabilistic design optimization.

The deterministic optimization problem is formulated as:

ysw ( s ) y fr ( s )
Minimize : F = wsw + w fr
ssw s fr
Subjeect to : ysw ≤ csw (6.8)
y fr ≤ c fr
0.8 ≤ s ≤ 1.2

The probabilistic optimization problem is formulated as

wμsw wμ fr wσ sw wσ fr
Minimize : F = μ sw
2
+ μ 2fr + σ sw
2
+ σ 2fr
sμsw sμ fr sσ sw sσ fr
Subject to : μ sw + 3σ sw ≤ csw (6.9)
μ fr + 3σ fr ≤ c fr
s + 3σ s ≤ 1.2
0.8 ≤ s-3σ s

126
And the PI control optimization strategy can be described as

Minimize : F = y fr ( s)
Subject to : ysw ≤ csw (6.10)
0.8 ≤ s ≤ 1.2

In the figure 6.4 and 6.5, the sidewall wrinkling limit csw is relaxed from 0.21 to 0.23

gradually. The weights of fracture and sidewall wrinkling are chosen the same in the

deterministic optimization. The coefficient of variation for s is 1%, for the other three

is 5%. It is observed that PI control strategy and deterministic optimization always have

the same result, which is the intersection point of SW criterion line with the SW

magnitude curve. As explained earlier, PI control strategy tends to minimize the fracture

by letting the sidewall wrinkling magnitude approach limit. Therefore, as the criterion for

wrinkling is relaxed a bit, the blank holder force need be reduced to allow more

wrinkling.

For the deterministic design, if the weights for the two defects are the same, it is easy to

find out from the figure that the addition of those two curves has the lowest point at

s equals 0.8 and increases monotonously until it equals 1.2. The intersection point not

only satisfies both the fracture and wrinkling constraints, but also has the smallest

addition of their magnitude. However, concerning the process parameters variation,

neither of these two methods provides a reliable solution.

127
Wrinkling/thinning magnitude

Scale of BHF: s

Figure 6.4. The design solution of PI control, deterministic design and probabilistic design

for csw = 0.21 .

128
Wrinkling/thinning magnitude

Scale of BHF: s

Figure 6.5. The design solution of PI control, deterministic design and probabilistic design

for csw = 0.23 .

129
Csw=0.21 Csw=0.23

PI control Deter. Opt Prob. Opt PI control Deter. Opt Prob. Opt

S 0.925 0.925 0.9625 0.845 0.845 0.9125

Pr{no wrinkling} 45.78% 45.78% 91.04% 48.04% 48.04% 99.98%

Pr{no fracture} 100% 100% 99.86% 100% 100% 100%

(The coefficient of variation for s is 1%, other three are 5%)

Table 6.3. The deterministic and probabilistic design result with constraint reliability.

130
Csw=0.23
400

300

Frequency
200

100

0
0.217 0.224 0.231 0.238 0.245 0.252 0.259
Sidewall wrinkling magnitude

Cfr=0.25
350

300

250
Frequency

200

150

100

50

0
0.1870 0.1955 0.2040 0.2125 0.2210 0.2295 0.2380 0.2465
Maximum thinning magnitude

Figure 6.6. Histograms of sidewall wrinkling and maximum thinning for the

deterministic designs.

131
Csw=0.23
400

300

Frequency
200

100

0
0.198 0.204 0.210 0.216 0.222 0.228
Sidewall wrinkling magnitude

Cfr=0.25
300

250

200
Frequency

150

100

50

0
0.210 0.216 0.222 0.228 0.234 0.240 0.246
Maximum thinning magnitude

Figure 6.7. Histograms of sidewall wrinkling and maximum thinning for the probabilistic

designs.

132
On the contrary, the probabilistic design always pulls the optimum point away from the

deterministic constraint boundary. By this means, the probabilistic constraints can be

fulfilled with the larger fracture magnitude as tradeoff compared to the deterministic

design. In figure 6.4 where the sidewall wrinkling critical value is 0.21, it is also noted

that the probabilistic optimization gives no solution since the wrinkling reliability

constraint cannot be satisfied no matter what value s takes within its feasible region.

The best achievable probability of no wrinkling is 91.04% when s equals 0.9625. As the

sidewall wrinkling critical value is relaxed to 0.23, the probabilistic optimization has

solution and the probability of no wrinkling is 99.98% and the probability of no fracture

is 100%. The probabilistic design results are shown in table 6.3. The probabilities shown

here are calculated by the Monte Carlo Simulation and the histograms of the sidewall

wrinkling and fracture magnitude are compared side by side for the deterministic and

probabilistic optimization solutions, as drawn in figure 6.6 and 6.7. The sample size for

the Monte Carlo Simulation is 5000. It is easy to see that the distribution of thinning is

far away from the upper limit in the deterministic design. So it is very safe in terms of

avoiding fracture. On the other hand it turns out that half of the distribution of sidewall

wrinkling is outside of limit. It means half of the parts will be scrapped due to the

winkling problem. The probabilistic design, in contrast, pushes the distribution of

thinning a bit close to limit but gets rewarded by pulling most distribution of wrinkling

within limit. Therefore, the traditional deterministic design is not robust by forcing the

design solution on the constraint boundary.

133
Deterministic simulations were also conducted to evaluate the probabilistic design vs. the

deterministic design. The results are shown in figure 6.8. Figure 6.8(a)(b) shows the

thinning and wrinkling at the probabilistic design point. When the process is stable, we

could assume plus/minus 3 sigma. The most risky cases happen at these extremes. Figure

6.8(e)(f) shows the thinning and wrinkling at one of these most risky cases. The

deterministic design simulation result is shown in figure 6.8(c)(d). For the worse case of

scenario of probabilistic design, the wrinkling and thinning are still within the criteria

limits.

6.7 Conclusion

In this chapter, a probabilistic design for the variable BHF in the cylindrical cup drawing

process is conducted and compared with the traditional deterministic design. The result

shows that the BHF predicted by the deterministic method is not robust under the

presence of process variations. However, given the same failure criteria, the probabilistic

design improved the yield (probability of good parts) to 99.98% from the 48.04%

obtained by the traditional deterministic design. In mass production environment, the

achieved process reliability is huge.

134
(a) thinning for probabilistic design

Cross-section cut of drawn cup at


25% drawing height to measure the
sidewall wrinkling of the Prob.
design

SW=0.212

(b) sidewall wrinkling for probabilistic design

Figure 6.8. Evaluation of probabilistic design vs. deterministic design. (to be continued)

135
Figure 6.8 Continued

(c) thinning for deterministic design

Cross-section cut of drawn cup at


25% drawing height to measure the
sidewall wrinkling of the
Deterministic design

SW=0.229

(d) sidewall wrinkling for deterministic design

136
Figure 6.8 Continued

(e) thinning for probabilistic design at one of the extreme cases where all noise variables
at 3sigma away from mean

Cross-section cut of drawn cup at 25%


drawing height to measure the sidewall
wrinkling of the Prob. Design at 3sigma
extreme case

SW=0.226

(f) sidewall wrinkling for probabilistic design at one of the extreme cases where all noise
variables at 3sigma away from mean

137
CHAPTER 7

7. SPATIALLY VARYING CONSTRAINTS AND

PROBABILISTIC DESIGN BY MULTI-OBJECTIVE

GENETIC ALGORITHM

In chapter 6, the drawing quality is improved by the introduction of temporal varying

blank holder force. However, many researches on the application of spatial varying

constraints have been done to successfully to improve the drawing quality. Traditionally,

these spatial varying constraints can be achieved through placing multiple beads at

various locations on the drawing die. It is especially useful in the drawing of complex

shaped parts. Besides drawing beads, other forms of spatial varying constraints are also

being studied. They are segmented elastic binder and very new discrete friction concept,

which will be focused in this chapter. The idea of segmented binder is that the blank

holder can be segmented and different pressing forces are then applied at each sub-binder

so that distributed restraining forces are generated on the sheet to regulate the metal flow

spatially. The discrete friction is a very new research field. The basic assumption is that

by introducing the micro-texture or tiny pockets on the die surface the fiction
138
condition can be altered at different locations. From former cylindrical cup drawing test

and study, we know that the part quality is more sensitive to the lubrication than the

blank holder force. So in this chapter, we select the discrete friction method as one

promising form of spatial varying constraints and apply it to Hishida parts to study its

effectiveness in improving deep drawing quality. Due to the part geometry, to investigate

the advantage of discrete fiction concept, fifteen design variables need to be determined.

Meantime, the design objective is to reduce both the wrinkling and fracture defects

simultaneously. Traditional DOE and RSM combined method cannot resolve this

problem because a CCD design for 15 factors needs 32799 simulation runs, which are

obviously prohibited. Instead, the optimization is realized through the integration of

Multi-Objective Genetic Algorithm (NSGA-II) and numerical simulation code to find the

optimal configuration of these local friction coefficients and other drawing process

parameter. The drawing quality by the uniform friction and the discrete friction are

compared. The results show that at the same draw depth and wrinkling level, around 12%

reduction of the maximum thinning has been obtained from discrete friction. If the

objective is to reduce both the wrinkling and fracture, an overall 33% improvement can

be obtained at the certain Pareto optimal setting from discrete friction comparing to

uniform friction. To incorporate process uncertainties and variations, the probabilistic

design search is implemented based on the deterministic design result from the

multi-objective genetic algorithm. The reliability analysis is carried out at each feasible

point from the last Pareto fronts. The design configuration with the highest reliability is

chosen as the optimal probabilistic design. For a complex case like this, the traditional

139
probabilistic design approach mentioned in chapter 5 and 6 is not applicable. However,

the proposed two phases design method illustrated in this chapter is able to effectively

find the optimal deterministic design pool and then search out the best probabilistic

design.

7.1 Introduction

Increase the forming window of the drawing process is an eternal topic in the sheet metal

forming industry. A typical stamping process is composed of three steps:

1. Binder holding blank: binder moves down to clamp the sheet;

2. Forming part: punch moves down to drag the sheet into the die cavity to take

shape of die and punch;

3. Releasing part: punch and binder move up and remove the formed part from the

die set.

Treating sheet metal forming process as a system, the forming window of the deep

drawing process is determined by intrinsic or constitutive material properties and

extrinsic factors in the manufacturing process, such as internal friction, forming speed,

tooling design, blank shape.

140
At a given sheet material and blank shape, tooling and process parameters, such as

forming speeds, temperatures etc., the metal flow into the die is determined by the

restraint imposed by the blank holder or the drawbead penetration. Previous studies have

shown that spatial distribution of these restraint forces could largely improve the

drawability of a given part.

The drawbeads are currently wide used in automobile industries to form complicated

auto-body panels. During the drawing, the blank holder force is usually uniformly

distributed along the binder. Its magnitude can be adjusted to roughly control the sheet

material flow into the die. The drawbeads, which are installed on the various spots of the

binder, could help more accurately control the material flow by exerting strong

restraining forces. One setup of such drawbeads for Hishida part drawing is shown in

figure 7.1.

Numerous researches have been done to model the bead, design the bead and study the

effect of bead on the part quality. Triantafyllidis et al. [49] considered the effects of the

drawbead using one-dimensional elasto-plastic shell element and compared numerical

results with experimental results. Cao and Boyce [50] analyzed the restraining force with

respect to the depth of drawbead. Naceur et al. [29] presents an optimization procedure of

drawbead restraining forces spatial distribution in order to improve the formability in

deep drawing process.

141
Figure 7.1. Example of one setup of drawbeads on the Hishida part drawing [51].

142
Siegert [51] first studied the segment-elastic binder and applied this concept in the

Hishida part drawing, as shown in figure 7.2(a). In this configuration, the binder has a

pyramidal support structure. On each pyramid a local blank holder force is introduced.

The blank holder forces are the pin forces of a multipoint cushion system. The pins,

which are adjustable by CNC, transfer the blank holder forces from the cushion plate to

the lower binder. Six pins are seen to support the binder, three on each long side of the

die. By spatially applying different forces along the binders, as illustrated in figure 7.2(b)

the limit drawing depth without any defects increased from 55mm to 63mm. Ayed [12]

applied the same concept on the optimization of the blank holder force distribution in the

stamping of a car front door panel. Although his work was focusing on the RSM method

to find the optimal force for the seven sub-binders, the result showed that the segmented

binder could reduce the inclination angle from 5 degree to 1 degree without wrinkling

and fracture defects. The segmented binder could alter the restraining force distribution

by changing the normal force applied on the blank to affect the flow pattern of the sheet

metal. We know that the restraining force is determined by both the normal force and the

friction coefficient. This reminds us that if we could control the friction condition

between die/punch and sheet, the drawability of sheet could also be improved without

using the complex segmented binder structure. Besides, the different friction can be

applied not only to binder but also to the die and punch. This actually expands the

forming window and gives us more control options. Therefore, in the following part we

will study the discrete friction concept and its application in Hishida part to form spatial

restraining force distribution.

143
(a) the segment-elastic binder

(b) the optimal pin forces for the segment-elastic binder

Figure 7.2. The segment-elastic binder used in the Hishida part drawing [51].

144
In sheet metal forming, the friction condition is affected by many parameters, such as

material properties of workpiece, tool and lubricant, sliding velocity and contact pressure,

and their effects are global, e.g. change the lubricant will change the friction everywhere

in the part interface.

However, the frictional conditions are also significantly dependent on the surface

topography of the tool [52] that can be locally tailored by applying micro texturing

techniques [53]. Excimer laser material processing is such a technique that offers the

opportunity to produce micro textures with high precision and flexibility on almost any

material – metal, ceramics and polymers [54]. The rectangle texture pocket formed by

Excimer lasers on the ceramics and its cross-section are shown in figure 7.3(a).

As illustrated in the figure 7.3(b), the friction zones in the textured stamping tooling can

be divided into textured area and non textured area. In the textured area, if the pressure is

higher enough to support the asperities of sheet blank, hydrodynamic lubrication situation

will be formed [53]. At the non-textured area, mixed lubrication regime happens

normally in the metal forming [52].

The micro texture tailored surface improves the friction condition by: a). forming

hydrodynamic pressure film to change real contact area; b) supplying lubricant by

working as small reservoirs. According to the experiments in [53], 14% of friction

coefficient reduction can be achieved by just varying the pocket length and depth. An

145
additional potential improvement is possible by changing the width and arrangement of

textures as well as the portion of the textured area. Besides adding the pocket texture to

the tooling surface, altering the surface roughness by micromachining is also another

effective measure to locally change the friction condition. Reported in [55], there is

maximum 54% friction reduction for die material M300 when the die surface roughness

Rz goes from 0.8 to 5.3 μm , see figure 7.4.

146
(a), rectangle texture pocket and the cross-section profile

(b), rectangle texture pocket and the cross-section profile

Figure 7.3. Micro-texture to alter the friction condition [53].

147
Figure 7.4. Topological effect at the roughness level [55].

148
7.2 Setup of Spatially Varying Constraints in Hishida Part Drawing

Based on the above discussion, friction condition on the stamping tooling interfaces can

be tailored by micro-texture process or micro machining process. A discrete friction

concept, which improves the deep drawing process by applying locally different friction

conditions, is proposed and demonstrated in this section. It is assumed that the friction

coefficients are design variables and the desired values can be realized at different zones

on the tool surfaces.

In the drawing of non-symmetrical geometry, the metal deformation changes with the

geometry and thus the required friction forces to control the metal flow are different.

Generally, the material at corner area is hard to flow in and prefers lower restraining

force while the straight edges have less compressive deformation caused thickening

tendency and thus need more retraining forces. The proposed discrete friction concept

tries to divide the deep drawing tooling into different friction zones and improves the

forming window by selecting different friction condition at the each friction zone. The

Hishida part, which has four corners with different radii and four tapered walls with

different slope angles on each side, is chosen, see figure 7.5 [39][47].

149
Figure 7.5. The Hishida Part.

150
Since the ideal friction condition depends on the drawing geometry, the segmentation of

these discrete zones follows the rule that each individual zone tends to have one unique

geometric feature and they are not overlapped. Based on this rule, the Hishida die/binder

and punch are divided into 10 zones and 4 zones, as denoted as μ1 , L , μ10 and

μ11 , L , μ14 respectively, see figure 7.6. Since metal flow is also largely controlled by the

blank holder force, BHF is added as another design variable. So for this new discrete
r
friction application, the design vector can be described as x = {μ1 , μ 2 , L , μ14 , BHF } . The

question following is how to find the optimal configuration of these friction conditions.

The next session will introduce the multi-objective genetic algorithm and the reason why

it should be used in this case.

151
Figure 7.6. Spatial distribution of discrete friction zones.

152
7.3 Multi-Objective Genetic Algorithm: NSGA-II

There are several reasons why multi-objective genetic algorithm should be used to find

the optimal configuration of the friction values in this discrete friction case.

First, the traditional optimization techniques like sequential quadratic programming (SQP)

can only locate the local minimum. When the design space is high dimensional and the

response is not smooth or linear, it is more likely that the global optimal solution will be

missed. In this discrete friction scenario, the die is segmented to 10 friction zones.

Including the punch/sheet, die/binder/sheet friction coefficient and one blank holder force,

there are total fifteen design variables. Since the discrete friction zones on the die are

connecting to each other, the metal flow is affected jointly by the friction values. This

makes the system behavior highly nonlinear. Then traditional optimization method may

fail to find global minimum. MOGA, one of the exploratory optimization methods, is

very good at finding the global optimum.

Second, a good sheet metal forming process is judged by multiple criteria such as

wrinkling, fracture, and insufficient stretching. Traditional optimization methods usually

convert the multiple objectives to one single objective through a set of predetermined

weight coefficients. Most time the coefficients are determined based on expert

153
knowledge or trial-and-error and may not be right. However, the MOGA can compute

multiple objective functions simultaneously in one optimization run without converting

them into single objective by weighted linear combination. So no arbitrary weight

coefficients are needed.

Third, the multiple criteria in this sheet metal forming problem are conflicting with each

other. It is apparent that reduction of the tendency to fracture would increase the

probability of wrinkling. Therefore, there exists strong trade-off between these objectives.

The presence of multiple objectives then gives rise to a set of optimal solutions (known

as Pareto-optimal solutions), instead of a single optimal solution. Simply setting the

weights for them will likely to exclude the potential superior solution to the engineering

problem. However, the MOGA is capable of finding multiple Pareto-optimal solutions in

one single simulation run by the identification of the Pareto front.

Besides, unlike the numerical optimization techniques, the exploratory MOGA does not

require the calculation of the local gradient information of the objective to find the search

direction. In sheet metal forming design, getting this local gradient is very difficult or

even impossible.

154
7.3.1 Pareto optimal and Pareto front

*
For minimum problem, a vector of decision variables x is Pareto optimal if there does

not exist another x such that

*
f i ( x) ≤ f i ( x ) for all i = 1, L n

And
(7.1)
r *
f j ( x ) < f i ( x ) for at least one j , 1 ≤ j ≤ n.

n is the number of the objectives.

*
In words, this definition says that x is Pareto optimal if there exist no feasible vector of

decision variables x which would decrease some criterion without causing a

simultaneous increase in at least one other criterion. Unfortunately, this concept almost

always gives not a single solution, but rather a set of solutions called the Pareto optimal

*
set. The vectors x corresponding to the solutions included in the Pareto optimal set are

called non-dominated. The plot of the objective functions whose non-dominated vectors

are in the Pareto optimal set is called the Pareto front, as shown in figure 7.7, each design

objective corresponding to a coordinate axes. The multiple-objective genetic algorithm

can find the Pareto front through the following basic procedures. Initially, the algorithm

randomly selects multiple design points according to the chosen population size and

evaluates them against each of the objectives.


155
Figure 7.7. The Pareto optimal and Pareto front.

156
Then it finds the Pareto front to the current population by the so called domination

criterion. It is similar to the definition of Pareto solutions. A solution “a” is said to

dominate another “b” in the population if it is at least as good as b in every dimension

and better than b in at least one dimension (objective). The first set of Pareto front has

rank one, denoted by F1 . Then the next level of Pareto front F2 is found after the first set

of Pareto optimal is discounted from the population. This ranking process continues until

all the design points are ranked and denoted by F1 , F2 , L , Fk . After the ranking, the good

genes are selected, mutated and crossed-over to generate the next better population just

like the nature selection process by which a superior creature evolves whilst inferior

creatures fade out from their population as generations goes by.

7.3.2 Non-dominated Sorting Genetic Algorithm-II

Within the last two decades, a number of different EAs were suggested to solve

multi-objective optimization problem. Of them, Fonseca and Fleming’s MOGA [56],

Srinivas and Deb’s NSGA [57], Horn et al.’s NPGA [58], Zitzler and Thiele’s SPEA [59],

and Knowles and Corne’s PAES [60] enjoyed more attention. NSGA-II, the improved

version of NSGA, outperforms PAES and SPEA in terms of finding a diverse set of

solutions and in converging near the true Pareto optimal set [61]. Also, NSGA-II is more

computationally efficient by using elitism and crowded comparison operator that keeps

diversity without specifying any additional parameters.

157
The NSGA-II uses a fixed population size of N . In generation t , an offspring

population Qt of size N is created from parent population Pt and non-dominated

fronts F1 , F2 , L , Fk are identified in the combined population Pt ∪ Qt . The next

population Pt +1 is filled starting from solutions in F1 , then F2 , and so on as follows.

Let l be the index of a non-dominated front Fl that | F1 ∪ F2 ∪ L ∪ Fl |≤ N and

| F1 ∪ F2 ∪ L ∪ Fl ∪ Fl +1 |> N . First, all solutions in fronts F1 , F2 , L , Fl are copied to

Pt +1 , and then the least crowed ( N − | Pt +1 | ) solutions in Fl +1 are added to Pt +1 . This

approach makes sure that all non-dominated solutions ( F1 ) are included in the next

population if | F1 |≤ N , and otherwise the selection based on crowding distance will

promote diversity. The procedure is illustrated in figure 7.8.

The non-dominated sorting is described here. First, for each solution “ p ” two entities are

calculated: 1) dominated count n p , the number of solutions which dominate the solution

p , and 2) S p , a set of solutions that the solution p dominates. All solutions in the first

front F1 will have their domination count as zero. Now, for each solution p with

n p = 0 , we visit each member (q ) of its set S p and reduce its domination count by

one. In doing so, if any member q the domination count becomes zero, it is put in a

separate list Q . These members belong to the second non-dominated front F2 . Then the

above procedure is continued with each member of Q and the third front F3 is

identified. This process continues until all fronts are identified.


158
Figure 7.8. The procedure of NSGA-II [61].

159
The crowding distance is used to distribute the solutions more uniformly over the true

Pareto front to maintain a diverse population. Randomly select two solutions “a” and

“b”; if the solutions are in the same non-dominated front, the solution with a higher

crowding distance wins. Otherwise, the solution with the lowest rank is selected. Without

taking any preventive measures, the population tends to form relatively few clusters in

multi-objective genetic algorithm. The method used in NSGA-II is:

Step1. Rank the population and identify non-dominated fronts F1 , F2 , L , Fk . For each

front j = 1, L , k repeat steps 2 and 3.

Step2. For each objective function n , sort the solutions in F j in the ascending order.

Let

s =| F j | and x[i ,k ] represent the i th solution in the sorted list with respect to the

objective function n . Assign cd n ( x[1,n ] ) = ∞ and cd n ( x[ s ,n ] ) = ∞ , and for i = 2, L , s

z n ( x[ i +1,n ] ) − z n ( x[i −1,n ] )


assign cd n ( x[ i ,n ] ) =
z nmax − z nmin

Step3. To find the total crowding distance cd ( x) of a solution x , sum the solution

crowding distances with respect to each objective, i.e., cd ( x) = ∑ cd n ( x) .


n

160
7.4 Design Optimization Model

The simulation model for Hishida part forming process has been built up in the explicit

code PAM-STAMP2G, see figure 7.9. The Krupkowsky-law hardening equation:

σ y = K (ε 0 + ε p ) n is used for modeling the material behavior. The blank material

properties of the simulation model are listed in table 7.1.

7.4.1 Design Objectives

The objective functions needed in the NSGA-II to evaluate the quality of drawn Hishida

part are 1) fracture, 2) wrinkling. The building of these functions is based on the strains

on Forming Limit Diagram (FLD), which value can be obtained from the results file of

numerical simulation of Pamstamp 2G quickstamp full FEA.

An extensive literature review on fracture can be found in [62][63]. Generally, these

fracture criteria can be classified into the following three types: geometry based criteria,

e.g. FLD [64][65] and thinning at the part wall [66]; stress based, e.g. FLSD [67]; and

damage based criteria, e.g. Cockroft and Latham criterion [67]. In this study, thinning in

the part, a geometrical criterion, which has been traditionally used to estimate the

proximity to failure [66], is chosen as a criterion for determining the fracture. This is an

approximate method, because the limit of thinning is affected by strain paths.

161
Nevertheless, the thinning criterion is still useful and effective in most deep drawing

operations.

For wrinkling, Havranek [68] in his conical cup forming experiments found the

distribution of strain at the onset of sidewall formed a narrow band, which can be plotted

by a wrinkle-limit curve (WLC) in the negative section of the Forming-limit diagram

(FLD). According to his study on a range of material thickness, 0.25-0.99mm, this

wrinkling-limit curve is proved to be independent of sheet thickness. Hosford and

Caddell [69] have also shown that if the absolute value of principle strain ratio

β = ε min ε (where ε min and ε max are the major and the minor principal strains,
max

respectively) is greater than a critical value the wrinkling is supposed to happen, and the

larger the absolute value of β , the greater is the possibility of wrinkling to occur. Chen

[70] has successfully used this critical strain ratio value to predict sidewall wrinkling in

his studies on the formation of the tapered rectangular cups and stepped rectangular cups.

Sheng [39] used this wrinkling limit to predict the variable blank holder force for a

Hishida part. Therefore in this study, we will use his result and use 1 as critical wrinkling

ratio. This is equal to ε max = −ε min in the FLD diagram. The wrinkling objective

function can be defined as the proportion of strain points within the wrinkling zone of

forming limit diagram. The points below this curve represent the strong wrinkling

tendency.

162
(a) die with 10 different friction zones

(b) binder with 10 different friction zones

(c) punch with 4 different friction zones

Figure 7.9. The simulation model of discrete friction drawing of Hishida part.

163
Sheet material ECODAL
Initial sheet thickness 1mm

Flow stress (Mpa) σ y = 470(0.0013 + ε p ) 0.25

Table 7.1. Material properties used in the simulation

164
7.4.2 Design Variables

The design variables in this discrete friction case are: 1) ten discrete friction coefficients

μ1 , L , μ10 for die/binder, 2) four different discrete friction coefficients μ11 , L , μ14 for

punch, 3) one blank holder force. So there are fifteen design variables.

The multi-objective optimization for the discrete friction problem is formulated as,

Minimize:

F ( x) = {Obj f ( x), Objw ( x)}


Subject to:
400 KN ≤ BHF ≤ 700 KN ,
(7.2)
0.02 ≤ μ i ≤ 0.32, i = 1, L ,14
where:
x = {BHF , μi }, i = 1, L ,14

7.4.3 Integration of NSGA-II Optimization and Pam-Quickstamp Full FEA

In NSGA-II, at each generation, every design point is evaluated against the two

objectives: wrinkling and fracture. This calculation is based on the results output from the

finite element simulation of sheet metal drawing process. Suppose the population size is

thirty and evolves with fifty generations, NSGA-II requires approximately

165
30 × 51 = 1530 FEA simulations. If the formal Pam-Autostamp is used, given each

simulation needs 10 minutes, the whole optimization process requires at least ten days.

To alleviate the computation cost and find the solution quickly, the Pam-Quickstamp Full

Process code is used instead for its quickness. Pam-Quickstamp is a simplified approach

where some components of the press tool are deduced from the initial part and where the

real kinematics is not completely defined. Since this is a multi-objective trade-off

optimization problem, the Pam-Quickstamp, although with a little bit lower fidelity

comparing to Pam-Autostamp, can still yield a good global result. The optimization flow

chart is plotted in figure 7.10. In this NSGA-II optimization method, population size is

set as fifty and generations are set as eighty. Crossover probability is 0.9, crossover

distribution index is 20 and mutation distribution index is 100.

7.4.4 Hishida Part Drawing Process Heuristics

From the previous study, especially the segmented binder research for the Hishida part,

we found that the spatial distribution of the restraining forces was not random. In fact,

they go after some rules or heuristics which are followed by process engineers. We call

them process heuristics. One heuristics example is given here: in the corner area, in order

to pull the sheet more easily into the die, the friction between the punch/blank should be

larger than the friction between die/blank. We could transform these heuristics to

following explicit constraints bounds.


166
μ1 < μ14
μ4 < μ11
μ6 < μ12 (7.3)
μ9 < μ13

Therefore after multi-objective genetic algorithm searches out all the possible

combinations of these process parameters, we could use these heuristics rules to screen

out the designs that make sense to engineers. In addition, new rules can be added in or

existing heuristics can be modified. By this way, the sheet metal forming knowledge is

brought back to the design rather than using the genetic algorithm solely to find the

optimum mathematically.

7.5 Optimization Result and Analysis

After 4050 simulations, the Pareto front and all the other offspring objective evaluations

are shown in the figure 7.11. The blue solid dots represent the Pareto front and the light

blue dots are all the points evaluated by the genetic algorithm. The brown solid dots are

the design points that comply with the heuristics rule we mentioned above. However, in

the following analysis, this heuristics rule is ignored. We simply try to find any

opportunity to improve the part quality even if the friction settings seem abnormal. The

lowest level of thinning and wrinkling can approach to 0.07 and 0.01. Since the two

objectives are confronting to each other, it is impossible to achieve low values for both

thinning and wrinkling.

167
Start

FEA Modeling & Initial

Update
model
Pam-stamp FEA
NSGA-II
Mutation
Crossover
Objective Evaluation Selection

Terminate?

Drawing
End Process
Heuristics

Figure 7.10. The optimization flow chart.

168
0.23

Evaluation points in NSGA


0.21
Pareto front

Best uniform design


0.19
Comparable to uniform
design
0.17 Best discrete friction
design
Heuristics
0.15
Thinning

0.13

0.11

0.09

0.07

0.05
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45
Wrinkling

Figure 7.11. The Pareto front of discrete friction NSGA-II optimization.

169
To see the improvement of discrete friction in drawing of complex shape part, the

optimization of the uniform friction drawing is conducted according to the method in the

previous work [47] and the results are compared. When the weighs of wrinkling and

thinning objectives are equal, the optimum configuration for the uniform friction is that

blank holder force equals 605KN, die/binder friction equals 0.09 and punch friction

equals 0.12. The resulting maximum thinning and wrinkling are 0.098112 and 0.173657.

However, for the discrete friction case, in the Pareto front of figure 7.11 there is one

comparable point that has wrinkling equal 0.172064 but the maximum thinning is only

0.086309. This means that if the wrinkling is controlled at the same level, by applying

discrete friction the maximum thinning is improved by 12%. Furthermore, if the thinning

can be allowed to increase a little bit (the upper limit is 0.12), there exists a great

potential for the reduction of wrinkling. This best design point is shown in the figure 7.11

where the thinning is 0.10818 and the wrinkling dropped to 0.075453. Therefore, the

total objective (assume equal weights) is changed from 0.271769 of uniform friction to

0.183633 of the discrete friction case. The percentage of improvement is around 33% and

it is huge. The friction coefficient values and strain distributions of the uniform friction

and best discrete friction designs are plotted in figure 7.12 and 7.13 for comparison

purpose. The other benefit of applying discrete friction conditions is the more uniform

strain distributions within the part. Figure 7.14 illustrates the strain distribution, at the

corner walls A and B and bottom, the difference between the maximum strain and

minimum strain drop respectively from 0.178 to 0.156, from 0.176 to 0.136 and from

0.0214 to 0.013. The deformation on the part wall becomes more uniform.

170
(a) uniform friction design

(b) the strain distribution

Figure 7.12. Uniform friction design and strain distribution after drawing.

171
(a) discrete friction design

(b) the strain distribution


Figure 7.13. Discrete friction design and strain distribution after drawing.

172
0.25

Frist strain at the out


0.20

surface
0.15
Uniform
0.10 friction

0.05 Discrete
0.00
0.00 20.00 40.00 60.00 80.00
Curlinear distance (mm)

a) Strain distribution at the corner A

0.25
Frist strain at the out

0.20 Uniform
friction
surface

0.15

0.10
0.05 Discrete
fi i
0.00
0.00 20.00 40.00 60.00 80.00
Curlinear distance (mm)

b) Strain distribution at the corner wall B


First strain at the out

0.03
0.03 Uniform
0.02 friction
surface

0.02
0.01
Discrete
0.01 friction
0.00
150.00 250.00 350.00 450.00
Curvilinear distance (mm)

c) Strain distribution at the bottom

Figure 7.14. Strain distribution at different locations of the part after drawing.

173
7.6 The Probabilistic Design Search

To incorporate process uncertainties and variations, the probabilistic design search is

implemented based on the deterministic design result from the multi-objective genetic

algorithm. The reliability analysis is carried out at each feasible point from the Pareto

front. The criterion of wrinkling for Hishida part is 0.20, and the criterion of thinning is

0.12. Within the 50 points of the Pareto front, 19 of them have both wrinkling and

thinning smaller than the limits. They are denoted as feasible points.

For those 19 feasible points, reliability analysis is conducted at each point. The

aforementioned mean value first order method (MVFO) is used here to access the

reliability. The results are shown in figure 7.16, 7.17, 7.18, where the coefficient of

variation for blank holder force is kept at 0.05, and for die/punch friction it is changed

from 0.1 to 0.2 and 0.4. The purpose of three different levels of variations is to see when

the uncertainties of friction are enlarging, which point is the most robust design point.

All the results are summarized in the figure 7.15. The solid diamonds represent the

non-feasible deterministic points in the Pareto front. The solid square stands for the

feasible deterministic points in the Pareto front. The hollow triangles correspond to the

most reliable points (stochastic) when the BHF COV=0.05 and all other friction

COV=0.1. By referring to the figure 7.16, those 11 points all have zero percent of

wrinkling and thinning defect. When the uncertainties of friction is increased to

174
COV=0.2, the most reliable points are identified by the hollow square. In this case, only

3 points have zero percent of wrinkling and thinning defect. If we further increase the

friction COV to 0.4, we find there is no point which could have zero percent defect. The

best point has reliability for wrinkling equal 0.89 and for thinning 0.92.

The interesting observations from figure 7.15 lie in four aspects: (1), all the reliable

points are away from the deterministic constraints. (2), the higher the uncertainties, the

farther the robust point should be away from the deterministic constraints. (3), the most

robust design is in the middle of the robust points under lower uncertainties. (4), the

deterministic optimum is very different with the probabilistic optimum, which again

signifies the probabilistic design.

175
Reliability analysis of the Pareto front points

0.18
Non-feasible feasible good at cov=0.1
0.16 good at cov=0.2 best at cov=0.4

0.14

0.12
Best probabilistic design
0.1

0.08
Best deterministic design

0.06

0.04

0.02

0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45

Figure 7.15. Reliability analysis of the Pareto front points.

176
Reliability at COV=0.1
1.050 Prob(no thinning) Prob(no wrinkling)
1.000

0.950

0.900

0.850

0.800

0.750

177
0.700

0.650

0.600

0.550

0.500

0.450

0.400
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19
Prob(no thinning) 1 0.8 1 0.88 1 0.960 1.000 1.000 1.000 1.000 1.000 1.000 0.779 1.000 1.000 1.000 1.000 1.000 1.000
Prob(no wrinkling) 0.52 1 0.96 1 1 1.000 0.994 0.494 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000

Figure 7.16. Reliability of feasible points in the last Pareto front when BHF COV=0.05, friction COV=0.1
Reliability at COV=0.2
1.050 Prob(no thinning) Prob(no wrinkling)
1.000

0.950

0.900

0.850

0.800

0.750

178
0.700

0.650

0.600

0.550

0.500

0.450

0.400
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19
Prob(no thinning) 1 0.72 1 0.727 0.971 0.880 1.000 1.000 1.000 0.996 0.971 1.000 0.709 0.998 1.000 0.960 0.960 0.968 1.000
Prob(no wrinkling) 0.48 1 0.88 1 1 1.000 0.920 0.466 0.960 0.965 0.994 1.000 1.000 0.998 1.000 1.000 1.000 1.000 1.000

Figure 7.17. Reliability of feasible points in the last Pareto front when BHF COV=0.05, friction COV=0.2
Reliability at COV=0.4
1.050 Prob(no thinning) Prob(no wrinkling)
1.000

0.950

0.900

0.850

0.800

0.750

179
0.700

0.650

0.600

0.550

0.500

0.450

0.400
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19
Prob(no thinning) 0.993 0.6 0.945 0.576 0.798 0.679 0.937 1.000 0.912 0.919 0.839 0.918 0.618 0.868 0.921 0.786 0.796 0.824 0.898
Prob(no wrinkling) 0.467 1 0.708 0.987 0.98 0.994 0.799 0.420 0.790 0.811 0.893 0.888 0.902 0.929 0.810 0.905 0.984 0.976 0.894

Figure 7.18. Reliability of feasible points in the last Pareto front when BHF COV=0.05, friction COV=0.4
7.7 Conclusion

Spatial varying constraints can successfully improve the drawing quality. Traditional

method is to use drawbeads or segmented binder to generate strong restraining forces on

the sheet to regulate its flow. In this chapter, we investigated discrete friction concept and

its application in the Hishida part drawing. By this method, more flexible spatial

constraints can be applied. Due to the Hishida part geometry, fifteen design variables

need to be determined. Meantime, the design objective is to reduce both the wrinkling

and fracture defects simultaneously. Traditional DOE and RSM combined method cannot

resolve this problem. Instead, the optimization is realized through the integration of

Multi-Objective Genetic Algorithm (NSGA-II) and numerical simulation code to find the

optimal configuration of these local friction coefficients and other drawing process

parameter. The results show that at the same draw depth and wrinkling level, around 12%

reduction of the maximum thinning has been obtained from discrete friction. If the

objective is to reduce both the wrinkling and fracture, an overall 33% improvement can

be obtained from discrete friction comparing to uniform friction. To incorporate process

uncertainties and variations, the probabilistic design search is implemented based on the

deterministic design result from the multi-objective genetic algorithm. The reliability

analysis is carried out at each feasible point from the Pareto front. The design

configuration with the highest reliability is chosen as the optimal probabilistic design.

For a complex case like this, the traditional probabilistic design approach mentioned in

chapter 5 and 6 is not applicable. However, the proposed two phases design method

180
illustrated in this chapter is able to effectively find the optimal deterministic design pool

and then search out the best probabilistic design.

181
CHAPTER 8

8. CONCLUSIONS AND FUTURE WORK

The exclusion of inherent process variations in the current deterministic design methods

for sheet metal forming can lead to very unreliable result that may cause high scrap rate,

frequent rework, machine shut down and thus huge loss of profit. Extensive research has

been done in exploring the deterministic effect of each factor on the part, but the impact

of their variations on the fluctuation of output quality for the stamping process is seldom

addressed and quantified. The inclusion of uncertainty in the design and optimization

cycle should lead to better understanding of the impact of uncertainty associated with

system input on the system output. This understanding can then be applied for managing

such uncertainties. To date, there is very few report on incorporating the uncertainties and

variations in the design of stamping process.

In this research we propose three different probabilistic design approaches. It uses sheet

metal forming finite element method (FEM) simulation as the fundamental tool. When

the system meta-model is not complex, the design of experiments (DOE) technique and

response surface method (RSM) are integrated with FEM to build an explicit function to

connect process input to process performance output. With the quantification


182
of uncertainties of input variables, by Monte Carlo Simulation (MCS) or other reliability

analysis methods, the probability that the product conforms to its specification would

therefore be assessed. Through the right formulation of probabilistic optimization the

robust optimal design configuration could be found. In the case where the number of

process inputs is too many so that the design of experiment cannot handle and the system

meta-model cannot be approximated, the alternate approach that integrates FEM,

multi-objective genetic algorithm and reliability assessment is illustrated. By the

proposed probabilistic design approaches, deeper understanding about the relationship

between process variables uncertainties with the part qualities is achieved. Ultimately,

the process output variability could be reduced, defect rates could be minimized and

product quality would be improved. The strategies, their advantages and disadvantages

are listed in table 8.1.

183
Strategies Advantages Disadvantages
DOE + FEM + z Quality index is a simple z Cannot handle the problem
RSM + MCS + metric to represent the design where the number of input
Quality Index to quality and can be easily variables is over 10. The DOE
lump understood. runs are too many.
multi-objective z Once RSM is build, MCS can z The MCS is evaluated at each
be used to accept all forms of design point; therefore great
input variables distributions computation effort is needed.
z Reliability is analyzed at the z Since no optimization routine
uniformed distributed design is used, the search for the
points and no optimization optimal is not efficient.
routine is required. Therefore z The design formulation only
no issue of local minima or concerns the defect rate. It
maxima exists. does not care about whether
the quality metrics is highly
consistent within the spec.
DOE + FEM + z The weighted sum of μ 2 + σ 2 z Cannot handle the problem
RSM + POE for design formulation not only where the number of input
σ + Weighted requires the quality metrics variables is too many. The
Sum of μ 2 + σ 2 should be within the spec. but DOE runs are too many.
design also the variation of them z Must assume that the quality
formulation should be minimized. This is metrics (the responses) follow
reflecting the Taguchi loss normal distribution, which
function idea. may be totally incorrect in
z The propagation of error will some cases.
derive an explicit form of z The multi-objectives are
standard deviation of quality lumped by the weighted sum
metrics and incorporate them of mean square and variance,
into the weighted sum of which are not straightforward
as the formulation in the first
μ 2 + σ 2 . Then difficult
strategy.
probabilistic design problem is z The mean square and variance
transformed to traditional may be at different scale,
optimization problem. which rises difficulty to
z No MCS is needed, which assign scale factors and
saves a lot computation effort. weights.
z The method is very efficient z The optimization routine may
since optimization routine is find the local minima or
used to search the optimal maxima instead of global
design. minima or maxima.

Table 8.1. The summary of prob. & determ. design strategies. (to be continued)

184
Table 8.1 Continued.

FEM + z Can handle very complex z The computation cost is still


Multi-objective problem where the number very huge. Each function
Genetic of design variables is very evaluation is a finite element
Algorithm + large. simulation since no meta-model
Reliability z The genetic algorithm is an or surrogate model is used.
Analysis at explorative method. The z The reliability analysis is
Pareto front design space is more conducted after the
thoroughly searched than the deterministic design by
second strategy. multi-objective genetic
z The multi-objective genetic algorithm.
algorithm does not need the z The selection of optimal design
arbitrary setting of weights still requires experience. This is
for different objectives. harder than the first strategy
Instead, it gives the Pareto where a simple quality index is
front, from which the adopted.
designer can select the right
optimal design based on the
ranking of the importance of
objectives.
z No DOE and RSM are
needed.
z No gradient is needed.

185
The future work includes:

1. Gathering more quantitative information of the process control and noise parameters,

especially their distribution statistics.

2. Investigate the more efficient probabilistic design algorithm when the process

parameters do not follow normal distributions.

3. Integrate the reliability analysis directly to the multi-objective genetic algorithm loop.

4. Investigate other system approximation methods besides response surface model,

such as the suitability of Kriging model in the sheet metal forming behavior

approximation.

186
LIST OF REFERENCES
[1]. Nee, A.Y.C., PC-based Computer Aids in Sheet-metal Working, J.Mech. Work.
Technol. 19 (1989) 11-21

[2]. Majeske, K.D., and Hammett, P.C., Identifying Sources of Variation in Sheet Metal
Stamping, International Journal of Flexible Manufacturing Systems, 2002

[3]. Kobayashi, S., Oh, S., and Altan, T., Metal Forming and the Finite-element Method,
Oxford University Press, New York, 1989.

[4]. Zeinkiewicz, O, The Finite Element Method, 3rd Edition, McGraw-Hill, New York,
1977.

[5]. Wifi, S.A., An Incremental Complete Solution of the Stretch-forming and


Deep-drawing of a Circular Blank Using a Hemispherical Punch, Int. J. Mech. Sci., 1976,
Vol.18, p. 23

[6]. Wang, N.M., and Budiansky, B., Analysis of Sheet Metal Stamping by a
Finite-element Method, General Motors Research Publication GMR-2423, 1978

[7]. Onate, E,. and Zienkiewicz, O.C., A Viscous Shell Formulation for the Analysis of
Thin Sheet Metal Forming, Int. J. Mech. Sci., 1983, Vol. 25, p.305

[8]. Toh, C.H., and Kobayashi, S., Finite-element Process Modeling of Sheet Metal
Forming of General Shapes, Grundlagen der Umformtechnik I Symposium, Stuttgart,
1983, p.39

[9]. Toh, C.H., and Kobayashi, S., Deformation Analysis and Blank Design in Square
Cup Drawing, Int. J. Machine Tools Des. Res, 1985, Vol. 25, No. 1, p.15

[10]. Pam-stamp 2G version, user manual, 2005

[11]. Isight 8.0 version, reference manual, 2004

187
[12]. Ayed, L.B., et., Optimization of the blankholder force distribution with application
to the stamping of a car front door panel, Proceedings of Numisheet, 2005, pp. 849-854

[13]. Gelin, J.C., and Labergere, C., Numerical design and optimal control for sheet
metal forming and tube hydroforming process, 7th Int. Conf. in Numercial Mehods in
Forming Processes, NUMIFORM 2001, Toyahashi, Japan, 18-20

[14]. Kim, S.H., and Huh, H., Design of the Bead Forces and Die Shape in Sheet Metal
Forming Processes using a Rigid-plastic Finite Element Method and Response Surface
Methodology, Trans. KSTP (in Korean), Vol.9, No.3, pp.284-292, 2000

[15]. Kim, S.H., Huh, H. and Tezuka, A., Optimum Design of Draw-bead Force in Sheet
Metal Stamping using Rigid-plastic FEM and Response Surface Methodology,
Proceedings of KSTP Spring Conference (in Korean), pp.143-146, 1999

[16]. Tezuka, A., Kim, S.H., and Huh, H., Process Parameter Design in Sheet Stamping
Processes with Rigid-plastic Finite Element Analysis, Trans of JSCES, Paper
No.20000011, 2000

[17]. Lepadatu, D., Hambli, R., Kobi, A., and Barreau, A., Optimization of Springback in
Bending Processes using FEM Simulation and Response Surface Method, Int. J. Adv.
Manuf. Technol., 2005, 27:40-47

[18]. Wang, L., and Lee, T.C., Controlled Strain Path Forming Process with Space Variant
Blank Holder Force using RSM Method, Journal of Materials Processing technology, 167,
2005, 447-455

[19]. Forsberg, J. and Nilsson, L., On Polynomial Response Surfaces and Kriging for Use
in Structural Optimization of Crashworthiness, Struct. Multidisc. Optim., 2005, 29:
232-243

[20]. Huang, Y., Lo, Z.Y., and Du, R., Minimization of the Thickness Variation in
Multi-step Sheet Metal Stamping, Journal of Materials Processing Technology, 177, 2006,
84-86

[21]. Yamazaki, K., Itoh, R., Han, J., Watanabe, M., and Nishiyama, S., Optimum Design
of Aluminum Beverage Can Ends Using Structural Optimization Techniques,
Proceedings of Numisheet, 2005, pp.719-724

[22]. Lin, J.C., and Tai, C.C., The Application of Neural Networks in the Prediction of
Spring-back in an L-shaped Bend, Int. J. Adv. Manuf. Technol., 1999, 15: 163-170

[23]. Ji, M.X., and Shivpuri, R., Reduction of Random Seams in Hot Rolling Through
FEA Based Sensitivity Analysis”, Materials Science and Engineering: A, Vol. 425, Issue:
1-2, pp. 156-166, June 15, 2006.
188
[24]. Hambli, R., Prediction of Burr Height Formation in Blanking Processes using
Neural Network, Int. J. Mech. Sci. 44, 2002, 2089-2102

[25]. Hasofer, A.M. and Lind, N.C., Exact and Invariant Second Moment Code Format,
Journal of Engineering Mechanics, ASCE, Vol.100, No. EM1, February, pp.111-121

[26]. Chen, X., Hasselman, T.K. and Neil, D.J., Reliability Based Structural Design
Optimization for Practical Applications, 38th AIAA/ASME/ASCE/AHS/ASC, Structures,
Structural Dynamics and Materials Conference, Kissimmee, FL, pp.2724-2732, Paper No.
AIAA-97-1403.

[27]. Pegada, V., Chun, Y., Santhanam, S., “An Algorithm for Determining the Optimal
Blank Shape for the Deep Drawing of Aluminum Cups,” Journal of Materials Processing
Technology, 125-126 (2002) 743-750

[28]. Moshksar, M.M., and Zamanian, A., “Optimization of the Tool Geometry in the
Deep Drawing of Aluminum,” Journal of Materials Processing Technology, 72 (1997)
363-370

[29]. Naceur, H., Guo, Y.Q., “Optimization of drawbead restraining forces and drawbead
design in sheet metal forming process,” International Journal of Mechanical Sciences, 43
(2001) 2407-2434

[30]. Gantar, G., Kuzman, K., “Sensitivity and Stability Evaluation of the Deep Drawing
Process,” Journal of Materials Processing Technology, 125-126 (2002) 302-308

[31]. Karthik, V., Comstock, R.J., Wagoner, R.H., “Variability of Sheet Formability and
Formability Testing,” Journal of Materials Processing Technology, 121 (2002) 350-362

[32]. Cao, J., Kinsey, B.L., “Next Generation Stamping Dies –Controllability and
Flexibility,” Robotics and Computer Integrated Manufacturing, 17 (2001) 49-56

[33]. Sahai, A., Cao, J., Xia, C.Z., “Sequential Optimization and Reliability Assessment
Method for Metal Forming Processes,” NUMIFORM 2004

[34]. Haldar, A., Mahadevan, S., Reliability Assessment Using Stochastic Finite Element
Analysis, John wiley & Sons, INC., 2000

[35]. Irfan, K., McMahonand, Xianyi, C. M., “Reliability-based Structural Optimization


Using the Response Surface method and Monte Carlo Simulation,” 8th International
Machine Design and Production Conference, Sep.9-11, 1998, Ankara Turkey

[36]. Sirkirk, J.F., “Process Variable Effects on Sheet Metal Quality,” Journal of Applied
Metalworking, American Society for Metals, (1986) pp.262-269
189
[37]. Jaisingh, A., Narasimhan, K., “Sensitivity Analysis of a Deep Drawing Process for
Miniaturized Products,” Journal of Materials Processing Technology, 147 (2004) 321-327

[38]. Lee, B.H., Kenum, Y.T., Wagoner, R.H., “Modeling of the Friction Caused by
Lubrication and Surface Roughness in Sheet Metal Forming,” Journal of Materials
Processing Technology, 130-131 (2002) 60-63

[39]. Sheng, Z.Q., Shivpuri, R., “Adaptive PI Control Strategy for Prediction of Variable
Blank Holder Force,” ESAFORM 2004, Trondheim, Norway, April 28-30, 2004

[40]. Hirose, Y., Hishida, Y., Furubayashi, T., Oshima, M., and Ujihara, S., 1990, “Part II:
Applications of BHF-Controlled Forming Techniques,” Proc. 4th Symposium of the
Japanese Society for the Technology of Plasticity.

[41]. Hardt, D.E., and Fenn, R.C., 1993, “Real-Time Control of Sheet Stability During
Forming,” ASME Journal of Engineering for Industry, 115(3), pp. 299-308.

[42]. Sim, H.B., and Boyce, M.C., 1992, “Finite Element Analysis of Real-time Stability
Control in Sheet Forming Processes,” ASME Journal of Engineering Materials and
Technology, 114(2), pp. 180-188.

[43]. Cao, J., and Boyce, M.C., 1994, “Design and Control of Forming Parameters using
Finite Element Analysis,” Computational Material Modeling, PVP-Vol. 294, pp. 265-285.

[44]. Sheng, Z.Q., Jirathearanat, S., and Altan, T., 2004, “Adaptive FEM Simulation for
Prediction of Variable Blank Holder Force in Conical Cup Drawing,” International
Journal of Machine Tools and Manufacture, 44, pp. 487-494.

[45]. Buranathiti, T., Cao, J., Xia, Z.C., and Chen, W., 2005, “Probabilistic Design in A
Sheet Metal Stamping Process under Failure Analysis,” NUMISHEET, L.M. Smith et al.,
eds., pp. 867-872.

[46]. Li, Y.Q., Cui, Z.S., Ruan, X.Y., and Zhang, D.J., 2005, “Application of Six Sigma
Robust Optimization in Sheet Metal Forming,” NUMISHEET, L.M. Smith et al., eds., pp.
819-824.

[47]. Zhang, W.F., Sheng, Z.Q., and Shivpuri, R., 2005, “Probabilistic Design of
Aluminum Sheet Drawing for Reduced Risk of Wrinkling and Fracture,” NUMISHEET,
L.M. Smith et al., eds., pp. 247-252.

[48]. Montgomery, D.C., and Myers, R.H., 1995, Response Surface Methodology:
Process and Product Optimization Using Designed Experiments, John Wiley & Sons.

190
[49]. Triantafyllidis, N., Maker, B., Samanta, S.K., An Analysis of Drawbeads In Sheet
Metal Forming: Part I-Problem Formulation, J. Engr. Mater. Technol. 108, 321-327
(1986)

[50]. Cao, J. and Boyce, M.C., Drawbeads Penetration as A Control Element of Material
Flow, Sheet Metal and Stamping Symposium, SAE 930517, Detroit, 1993, pp.145-153

[51]. Siegert, K., Ziegler, M., Wagner, S., Closed loop control of the friction force. Deep
drawing process, Journal of Materials Processing Technology 71 (1997) pp.126-133

[52]. Schey, J.A., Friction in Sheet Metal Working, SAE 970712

[53]. Neudecker, T., Popp, U., Schraml, T., Engel, U., Geiger, M., Towards optimized
lubrication by micro texturing of tool surfaces, Advanced Technology of Plasticity, Vol. I,
Proceedings of the 6th ICTP, Sept. 19-24, 1999, Nurmberg, Germany

[54]. Tonshoff, H.K., Hesse, D., Mommsen, J., Micromachining Using Excimer Lasers,
Annals of the CIRP 42 (1993) 1, p.247-251.

[55]. Wagner, S., Tribology in Drawing Car Body Parts, SAE, 1999-01-3228

[56]. Fonseca, C.M., and Fleming, P.J., Genetic Algorithms for Multiobjective
Optimization: Formulation, Discussion and Generalization, In Proceedings of The 5th
International Conference On Genetic Algorithms, pp.416, V423, 1993.

[57]. Srinivas, N., and Deb, K., Multiobjective Optimization Using Nondominated
Sorting in Generic Algorithm, Evolutionary Computation, Vol.2, No.3, pp.221, V248,
Fall 1994

[58]. Horn, J., Nafpliotis, N., and Goldberg, D.E., A Niched Pareto Genetic Algorithm for
Multiobjective Optimization, In Proceedings of the First IEEE Conference on
Evolutionary Computation, Vol.1, pp.82-87, 1994

[59]. Zitzler, E., and Thiele, L., Multiobjective Evolutionary Algorithms: A Comparative
Case Study and the Strength Pareto Approach, IEEE Transactions on Evolutionary
Computation, Vol.3, No.4, pp.257, V271, 1999

[60]. Knowles, J., and Corne, D., The Pareto Archived Evolution Strategy: A New
Baseline Algorithm For Pareto Multiobjective Optimization, In Proceedings of 1999
Congress on Evolutionary Computation, Vol.1, pp.105, 1999

[61]. K.Deb, A.Pratap, S.Agarwal, and T.Metarivan, A Fast and Elitist Multiobjective
Genetic Algorithm: NSGA-II, IEEE Transactions on Evolution Computation, Vol. 6, No.2,
2002, pp.182-197.
191
[62]. Sheng, Z.Q., Yang, J.B., Jirathearanat, S., Altan, T., Drawing of Conical
Cups-prevention of Wrinkling and Fracture by Controlling Blank Holder Force, ERC
Report/ERC/NSM-01-R-47-A, 2001, Engineering Research Center for Net Shape
Manufacturing, the Ohio State University.

[63]. Keeler, S.P., and Backofen, W.A., 1964, Plastic Instability and Fracture in Sheet
Stretched Over Rigid Punches, ASM Transactions Quarterly, Vol.56, pp.25-48.

[64]. Goodwin, G..M., 1968, Application of Strain Analysis to Sheet Metal Forming
Problems in the Press Shop, SAE paper, No.680093.

[65]. Cockcroft, M.G.., Latham, D.J., 1968, J. Inst. Metals, 33-39.

[66]. Shulkin, L.B., Mendelsohn, D.A., Kinzel, G..L., Altan, T., 1997, Blank Holder
Pressure (BHP) Control with Flexible Blank Holder in Sheet Metal Forming,
ERC/NSM-S-97-14, Engineering Research Center for Net Shape Manufacturing, the
Ohio State University.

[67]. Arrieux, R., Determination and Use of the Forming Limit Stress Diagrams in Sheet
Metal Forming, Journal of Materials Processing and Technology 53 (1995) 47-56.

[68]. Havranek, J., Wrinkling Limit of Tapered Pressings, J. Aust. Inst. Met. 20(2) (1975)
114-119.

[69]. Hosford, W.F., Caddell, R.M., Metal Forming: Mechanics and Metallurgy, 2nd edn,
1993.

[70]. Chen, F.K., Liao, Y.C., An Analysis of Draw-Wall Wrinkling in A Stamping Die
Design, Int. J. Adv. Manuf. Technology (2002) 19:253-259.

192
APPENDIX A

A.1 Objective:

1. To check whether the neural network can have the same level of prediction accuracy

as regression when the same set of data is given?

2. What is the optimal structure for the neural network for the fixed data size?

3. When the data are collected by different experimental design methods, which one

will give the least prediction error for the neural network modeling?

A.2 Approach:

First, the prediction accuracy of regression or neural network can not be simply taken for

granted as good or bad. In some cases, one method may be better than the other, while at

some time, it is opposite. So in order to have a comprehensive investigation of their

performance, we design an experiment with three test functions, four settings of data size

and three experimental design methods.

The three test functions are chosen so that their surface complexes are increased one by

one. They are: (1) Cubic function (2) Bivariate normal function (3) Cowboy hat

function, as shown in figure A.1.


193
(a) Cubic function

(b) Bivariate normal function

(C) Cowboy hat function

Figure A.1. Shape of three test functions

194
The four settings of data size are: (1) 9 data points, (2) 16 data points, (3) 25 data points,

(4) 36 data points. The three experimental design methods are; (1) D-optimal design, (2)

Bayesian D-optimal design, (3) Latin hypercube design.

For the same setting of data size, different data is collected according to the three

experimental design methods. Here, we notice that D-optimal can give several design

matrixes by the order of the assumed model when the number of runs is the same. For

example, when the number of runs is 25, the possible models are: (1) full first order, (2)

full second order, (3) full third order, (4) full forth order, (5) full fifth order. Since we

don’t know whether the prediction error is higher or lower when the assumed model

orders change, we use the table A.1 to illustrate all the model orders with different run

numbers.

195
9 runs 16 runs 25 runs 36 runs
full 2nd model
full 3rd model full 3rd model full 3rd model full 3rd model
4th full Bayesian full 4th model full 4th model full 4th model
Latin hypercube 5th full Bayesian full 5th model full 5th model
Latin hypercube 6th full Bayesian full6th model
Latin hypercube full 7th model
8th full Bayesian
Latin hypercube

Table A.1. The model orders with different run numbers.

196
1.5 1.5

1 1

0.5 0.5

0 Series1 0 Ser i es1


-1.5 -1 -0.5 0 0.5 1 1.5 -1.5 -1 -0.5 0 0.5 1 1.5

-0.5 -0.5

-1 -1

-1.5 -1.5

Run=9, 2nd full model D-optimal Run=9, 3rd full model D-optimal
1.5
1
1
0.8
0.5
0.6
Series1
0 Series1
0.4
-1.5 -1 -0.5 0 0.5 1 1.5
-0.5
0.2
-1
0
-1.5
0 0.5 1 1.5

Run=9, 4th full model Bayesian D-optimal Latin Hypercube design

Figure A.2. The data 2-D scatter plots for 9 points.

197
1.5
1.5
1
1
0.5
0.5
0
-1.5 -1 -0.5 0 0.5 1 1.5 Series1 0 Series1
-0.5 -2 -1 0 1 2
-0.5
-1
-1
-1.5
-1.5

Run=16, 3rd full model D-optimal Run=16, 4th full model D-optimal
1.5 1.2

1 1

0.8
0.5

0.6 Series1
0 Series1
-1.5 -1 -0.5 0 0.5 1 1.5
0.4
-0.5
0.2
-1
0
-1.5 0 0.2 0.4 0.6 0.8 1 1.2

Run=16, 5th full model Bayesian D-optimal Latin Hypercube design

Figure A.3. The data 2D scatter plot for 16 points.

198
1.5 1.5

1 1

0.5 0.5

0 Series1 0 Series1
-1.5 -1 -0.5 0 0.5 1 1.5 -1.5 -1 -0.5 0 0.5 1 1.5
-0.5 -0.5

-1 -1

-1.5 -1.5

Run=25, 3rd full model D-optimal Run=25, 4th full model D-optimal
1.5
1.5
1 1
0.5 0.5

0 0 Series1
-1.5 -1 -0.5 0 0.5 1 1.5 Series1
-2 -1 0 1 2
-0.5 -0.5

-1 -1

-1.5 -1.5

Run=25, 5th full model D-optimal Run=25, 6th full model Bayesian D-optimal

1.2

0.8

0.6 Series1
0.4

0.2

0
0 0.5 1 1.5

Latin Hypercube design

Figure A.4. The data 2D scatter plot for 25 points.

199
1.5
1.5

1 1

0.5 0.5
Series1
0 0 Series1
-1.5 -1 -0.5 0 0.5 1 1.5 -1.5 -1 -0.5 0 0.5 1 1.5
-0.5 -0.5

-1 -1

-1.5 -1.5

Run=36, 3rd full model D-optimal Run=36, 4th full model D-optimal
1.5 1.5

1 1

0.5 0.5

0 Series1 0 Series1
-1.5 -1 -0.5 0 0.5 1 1.5 -1.5 -1 -0.5 0 0.5 1 1.5
-0.5 -0.5

-1 -1

-1.5 -1.5

Run=36, 5th full model D-optimal Run=36, 6th full model D-optimal

1.5 1.5

1 1

0.5 0.5

0 Series1 0 Series1
-1.5 -1 -0.5 0 0.5 1 1.5 -1.5 -1 -0.5 0 0.5 1 1.5
-0.5 -0.5

-1 -1

-1.5 -1.5

Run=36, 7th full model D-optimal Run=36, 8th full model Bayesian D-optimal

1.2

0.8

0.6 Series1

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2

Latin Hypercube design


Figure A.5. The data 2D scatter plot for 36 points.

200
For each of these data sets, we will fit to the respective regression model. Then we

randomly select 1000 data points from the design space and calculate the squared

difference between the predicted value and the true value by our known test functions.

The overall prediction error is measured as follows:

)

1000
( yi − yi ) 2
RMSE = i =1

1000

However, to calculate RMSE for the neural network, things become complicated. The

reason is that first, the network can have different number of hidden nodes as you want. It

is not fixed at all. The second, there are a lot of training methods. Different ones may

have different effects. The third, the training convergence error can be set manually. To

see what effect can these factors have, for each data set we actually use 16 different

combination of number of hidden nodes, training method and training error to build the

neural network prediction model. The table A.2 represents the RMSE resulted from

different treatment for test function #3 at run 25 and 3rd full model D-optimal design. We

have 16 results for each design matrix.

201
Number
of Training Training
RMSE1 RMSE2 RMSE3 RMSE4 RMSE5 RMSEave RMSEvar
hidden method error
nodes
3 1 0.01 0.3608 0.35363 0.44372 0.55582 0.49739 0.43397 0.007607
3 1 0.001 0.39378 0.40901 0.39584 0.3677 0.40945 0.39954 0.000288
3 2 0.01 0.54614 0.6134 0.62496 0.60945 0.60945 0.61077 0.00097
3 2 0.001 0.51948 0.43873 0.43873 0.43873 0.43874 0.43873 0.001304
6 1 0.01 0.37859 0.32452 0.32269 0.48226 0.49886 0.39512 0.007166
6 1 0.001 0.68158 0.4759 0.38801 0.47085 0.45272 0.46649 0.012247
6 2 0.01 0.30436 0.23813 0.25355 0.33544 0.31371 0.29054 0.001712
6 2 0.001 0.27586 0.34779 0.35557 0.45858 0.39668 0.36668 0.004519
10 1 0.01 0.37595 0.6785 0.29018 0.31661 0.35506 0.34921 0.02478
10 1 0.001 0.50662 0.99966 0.34177 0.50057 0.26386 0.44965 0.081993
10 2 0.01 0.37809 0.311 0.37789 0.4026 0.26432 0.35566 0.003286
10 2 0.001 0.35574 0.40952 0.40331 0.53573 0.47556 0.42946 0.004931
15 1 0.01 0.41408 0.37616 0.50912 0.58991 0.61954 0.50437 0.01129
15 1 0.001 0.44542 0.77606 0.57462 0.33361 0.36945 0.46316 0.032371
15 2 0.01 0.31992 0.37212 0.22116 0.5473 0.59756 0.41311 0.024807
15 2 0.001 0.44199 0.41923 0.37942 0.32656 0.49367 0.41355 0.003996

Table A.2. RMSE of neural network from different training methods fro test function #3

at run 25 and 3rd full model D-optimal design.

202
In the table A.2, there are five RMSE values. Each of them is from the same neural

network model. The tricky thing about the neural network is that every time the same

network is trained by the same data set, the nodes weights are not the same. Thus the

prediction errors are varied. Sometime, the same network will even not converge. To

overcome this disadvantage, I make thirty same structured networks and train them by

the same data. After the training, I choose the five networks that have the least training

errors. Then I calculate the RMSE for each of those five networks and average them to

give the RMSEave for each network treatment.

For run equals to nine, and test function one, we run the three-way ANOVA, and found

that the number of nodes and training methods do matter in prediction error. The other

terms like the training error and other interaction terms are not significant. The similar

results apply also to other run number and test function combinations.

203
(a). Effect of network methods for test function 1

(b). Effect network methods for test function 2

(c). Effect of network methods for test function 3

Figure A.6. Effect of neural network methods for three test functions at run=9 and 2nd full
model D-optimal design.

204
Main Effects Plot (data means) for 3th_ave_1
nodes numer train method
0.5

0.4

0.3

Mean of 3th_ave_1
0.2

0.1
3 6 10 15 1 2
train error
0.5

0.4

0.3

0.2

0.1
0.001 0.010

(a). Effect of network methods for test function 1


Main Effects Plot (data means) for 3th_ave_2
nodes numer train method

0.4

0.3
Mean of 3th_ave_2

0.2

0.1
3 6 10 15 1 2
train error

0.4

0.3

0.2

0.1
0.001 0.010

(b). Effect of network methods for test function 2


Main Effects Plot (data means) for 3th_ave_3
nodes numer train method

0.7

0.6
Mean of 3th_ave_3

0.5

3 6 10 15 1 2
train error

0.7

0.6

0.5

0.001 0.010

(c). Effect of network methods for test function 3

Figure A.7. Effect of neural network methods for three test functions at run=9 and 3nd

full model D-optimal design.

205
The plots for the other order D-optimal or Latin hypercube designs are not shown here.

But we found that there is a principal in choosing the structure of network and some of

the training parameters. Basically, when the data size is small, we’d better select small

number of hidden nodes. When the data size is larger, we’d better select large number of

hidden nodes. The multiplication of input nodes number N input and number of hidden

nodes N hidden in addition of N hidden is recommended smaller than the data size N data .

Also when the number of hidden nodes is larger, the training error is better not set to be

too small. We can roughly regard the number of weights coefficient as the number of the

unknown parameters and the data size as the number of degree of freedom in regression

analysis. So if the number of hidden nodes is too many when our data size is relative

small, the network weights cannot be well estimated. For each data size settings, we

recommend corresponding neural network structures and training parameters. They are

tabulated in table A.3. So for each design matrix, we have 16 network results. From

these 16 results, we select the one by the criteria above. Then by this way, we can

compare the accuracy of regression and neural networks given the same data set.

The figure A.8 compares all the regression and network RMSE for different run numbers

(9, 16, 25, 36) with the three test functions. We observed that when the data size is small,

say 9 in this case, the regression outperforms the network a lot. However, when the data

collected are more and more, the network tends to have the similar performance as the

regression.

206
9 runs 16 runs 25 runs 36 runs
# of hidden nodes 3 3 6 10
Training method LM LM LM LM
Training error 0.001 0.001 0.001 0.01

Table A.3. Recommended neural network structure and training parameters.

207
Comparison of regression and neural net error

1
0.9
0.8
0.7
0.6
RMSE

neural net
0.5
regression
0.4
0.3
0.2
0.1
0
16

16

16

25

25

25

25

36

36

36

36

36
9

Figure A.8. Comparison of regression and neural network prediction error.

208
Comparison of regression and ANN, regre RMSE
(test function: cubic function) net RMSE
0.16
0.14
0.12
0.1

RMSE
0.08
0.06
0.04
0.02
0
9 9 9 9 16 16 16 16 25 25 25 25 25 36 36 36 36 36 36 36
Sample number

(a). Comparison of regression and ANN for cubic test function.


Comparison of regression and ANN, regre RMSE
(test function2: bivariate normal) net RMSE

0.35
0.3
0.25
RMSE

0.2
0.15
0.1
0.05
0
9 9 9 9 16 16 16 16 25 25 25 25 25 36 36 36 36 36 36 36

Sample number

(b). Comparison of regression and ANN for bivariate normal function.


Comparison of regression and ANN, regre RMSE
(test function3: cowboy hat function) net RMSE

0.8

0.6
RMSE

0.4

0.2

0
9 9 9 9 16 16 16 16 25 25 25 25 25 36 36 36 36 36 36 36
sample number

(c). Comparison of regression and ANN for cowboy test function.

Figure A.9. Individual plot of regression and ANN prediction error for three test
functions.

209
To have clearer view of the comparison, we split the RMSE by the test functions. The

figure A.9 (a) is the RMSE for the test function one. Since the full 3rd full polynomial

model includes the cubic function, the regression method actually gives no prediction

error. This tells us when the system response is simple the regression should be the better

choice.

The RMSE for the test function two is plotted in the figure A.9(b). We see that when the

response is starting to deviate from polynomial form, the performance of neural network

comes close to the regression. There are several peaks in the regression RMSE curve.

These points are corresponding to the Latin Hypercube designs. It is obvious that this

design method is not suitable for regression modeling and the regression method gives

higher requirement of the data than the neural network does. The test function three is

cowboy hat function, which is quite nonlinear. We know the neural network is linear

combination of basic activation functions, most of which are nonlinear too. So the

regression is supposed to work worse than network does in this case.

However, from the figure A.9(c), we see that regression still gives lower RMSE generally

than the network. This phenomenon is even apparent when the data size is small. The

explanation we tend to give is that using D-optimal design or Bayesian design, most

points gather at the corner of the design space. There is no much information given at the

center of the space. Since the network doesn’t know the underlying polynomial form as

the regression method, it is just being trained to remember the response at the

210
design space boundary. This is why it gives high prediction error over the whole space.

But when the data size is increased the network is doing better and better.

When the data are collected by different experimental design methods, which one will

give the least prediction error for the neural network method? We discussed the answer

by three different levels of response complexity. When the response is highly nonlinear

like the cowboy hat function, from the main effects plots, we found that given the small

size of data, the prediction is very bad no matter what experimental designs methods are

used. However, D-optimal design gives relative small prediction error (In the x-axis of

figure A.12, 9 means Bayesian D-optimal design, 10 means Latin hypercube design). It

also verifies our conclusion aforementioned: when the data is not big enough, we’d better

use less hidden nodes in the neural network otherwise the more the unnecessary hidden

nodes, the more spurious responses it will catch. When the number of runs is increased

from nine to thirty-six, the prediction error decreased from 0.55 to 0.15. We see in all

these cases, the Bayesian D-optimal design almost gives the lowest prediction error

within all the design methods for the function approximation by neural networks. My

explanation is that when the response becomes highly nonlinear, the traditional

D-optimal design still focuses on the boundary and some locations which may carry

low-order terms’ information. Bayesian D-optimal, on the other hand, contains lots of

high order terms and thus brings more information about the system’s behavior.

Therefore, it gives neural network lower prediction error. The Latin hypercube method

only gives good prediction when the number of experiments is big.

211
Main Effects Plot (data means) for ave Main Effects Plot (data means) for ave
model number of nodes model number of nodes
0.5 0.3

0.4
0.2
0.3

0.2
0.1
Mean of ave

Mean of ave
0.1
2 3 9 10 3 6 10 15 3 4 9 10 3 6 10 15
training method training method
0.5 0.3

0.4
0.2
0.3

0.2
0.1
0.1
1 2 1 2

Run=9, test function 1 Run=16, test function 1

Main Effects Plot (data means) for ave Main Effects Plot (data means) for ave
model number of nodes model number of nodes

0.150 0.10

0.125 0.08
0.100
0.06
0.075
Mean of ave

Mean of ave
0.04
0.050
3 4 5 9 10 3 6 10 15 3 4 5 6 7 9 10 3 6 10 15
training method training method

0.150 0.10

0.125 0.08
0.100
0.06
0.075
0.04
0.050
1 2 1 2

Run=25, test function 1 Run=36, test function 1

Figure A.10. Effect of DOE and neural network methods on prediction error for test
function 1.

212
Main Effects Plot (data means) for ave Main Effects Plot (data means) for ave_1
model number of nodes model number of nodes
0.5 0.30

0.4 0.25

0.3 0.20

0.2 0.15

Mean of ave_1
Mean of ave

0.10
0.1
2 3 9 10 3 6 10 15 3 4 9 10 3 6 10 15
training method training method
0.5 0.30

0.4 0.25

0.3 0.20

0.15
0.2
0.10
0.1
1 2 1 2

Run=9, test function 2 Run=16, test function 2

Main Effects Plot (data means) for ave_1 Main Effects Plot (data means) for ave_1
model number of nodes model number of nodes
0.18
0.12
0.16

0.14
0.10
0.12

Mean of ave_1
Mean of ave_1

0.10
0.08
3 4 5 9 10 3 6 10 15 3 4 5 6 7 9 10 3 6 10 15
training method training method
0.18

0.16 0.12

0.14
0.10
0.12

0.10
0.08
1 2 1 2

Run=25, test function 2 Run=36, test function 2

Figure A.11. Effect of DOE and neural network methods on prediction error for test
function 2.

213
Main Effects Plot (data means) for ave_2 Main Effects Plot (data means) for ave_2
model number of nodes model number of nodes
0.48
0.80
0.46
0.75
0.70 0.44

0.65 0.42
Mean of ave_2

Mean of ave_2
0.60 0.40

2 3 9 10 3 6 10 15 3 4 9 10 3 6 10 15
training method training method
0.48
0.80
0.46
0.75
0.70 0.44

0.65 0.42

0.60 0.40

1 2 1 2

Run=9, test function 3 Run=16, test function 3

Main Effects Plot (data means) for ave_2 Main Effects Plot (data means) for ave3
model number of nodes models number of nodes
0.40
0.40
0.35
0.35 0.30

0.30 0.25
Mean of ave_2

Mean of ave3
0.20
0.25
3 4 5 9 10 3 6 10 15 3 4 5 6 7 9 10 3 6 10 15
training method training method
0.40
0.40
0.35
0.35 0.30

0.30 0.25

0.20
0.25
1 2 1 2

Run=25, test function 3 Run=36, test function 3

Figure A.12. Effect of DOE and neural network methods on prediction error for test

function 3.

214
A.3 Conclusion:

In this research, I try to answer the three questions:

1) To check whether the neural network can have the same level of prediction accuracy

as regression when the same set of data is given?

Through the comparative study, I found that the network at most times gives worse

prediction accuracy than the regression method. Especially when the system real

response is quite linear, the disadvantage of network is more apparent. But when the

system’s behavior is more and more nonlinear, the network tends to have the similar

accuracy as the regression. Since in this project, only three test function is used, we can

only assume that when the response is even more complex, the network may outperform

the regression.

2) What is the optimal structure for the neural network when the data size is increasing?

In this project, I found that the number of the hidden nodes is the most important

parameter we should take care in the building of neural network. Basically, when the data

size is small, we’d better select small number of hidden nodes. When the data size is

larger, we’d better select large number of hidden nodes. The multiplication of input

nodes number N input and number of hidden nodes N hidden in addition of N hidden is

215
recommended smaller than the data size N data . Also when the number of hidden nodes is

larger, the training error is better not set to be too small. We can roughly regard the

number of weights coefficient as the number of the unknown parameters and the data size

as the number of degree of freedom in regression analysis. So if the number of hidden

nodes is too many when our data size is relatively small, the network weights cannot be

well estimated. In function approximation, the LM training method gives the fastest

training speed and best result.

3) When the data are collected by different experimental design methods, which one will

give the least prediction error for the neural network method?

Through the study, we found that for building the system response model by the neural

network technique, the Bayesian D-optimal design gives the better prediction accuracy

than the other experimental design methods such as Latin Hypercube design or D-optimal

designs. In the future work, the EIMSE will be compared with Bayesian D-optimal.

Reference:
[1]. Choueiki, M., Training Data Development With The D-optimality Criterion, IEEE
Transaction On Neural Networks, vol. 10, No. 1, 1999

[2]. DuMouchel, W., A Simple Bayesian Modification of D-optimal Designs to Reduce


Dependence on An Assumed Model, Technometrics, Vol. 36, NO. 1, 1994

[3]. Johnson, M, Some Guidelines for Constructing Exact D-optimal Designs on Convex
Design Spaces, Technometrics, Vol. 25, NO. 3, 1983
216
[4]. DeVeaux, R., Prediction Intervals for Neural Networks via Nonlinear Regression,
Technometrics, Vol.40, NO. 4, 1998

[5] Cohn, D., Neural Network Exploration Using Optimal Experimental Design,
Advances in Neural Information Processing Systems 6, Morgan Kaufmann, 1994

[6]. Poland, J., Different Criteria for Active Learning in Neural Networks: A Comparative
Study, University of Tubingen, Germany, 2003

217
APPENDIX B

In this section we will briefly show the validation result of using aluminum RSM to

predict the thinning of steel material we referenced in table 5.1. The aluminum RSM is

from chapter 5.

The high and low levels of the aluminum material properties used in DOE and the

variation of the referenced steel material properties are compared and listed in table B.1.

From the table, we could see that their K values are not overlapped at all. Their n values

are partially overlapped though.

We first keep the blank holder force and lub1 at the normal level. Then we change the

value of n and K from low to high in the Pam-stamp simulation. There are 4

combinations for n and K. After the simulation, the thinning value is extracted. Meantime,

we use the known RSM of aluminum material to predict the thinning of steel material at

this n and K combination. The results are shown in the first 4 data rows in table B.2. We

see that the predictions are quite close. The error percentages are all below 2%.

Then we keep both the n and K values at low level, change the values of blank holder

force and lub1, and check whether the RSM will still give the good prediction. From last

218
4 rows in table B.2, we could see that the predictions are also very close to the simulation

values. Figure B.1 plots all these data in one figure so that we can easily see the

difference.

From this testing, we see that the RSM build from the aluminum can be used to

extrapolate the thinning behavior of the steel material as long as their n and k values are

not that far away from the range used in the aluminum DOE. I think the fact is due to the

monotonous characteristic of the thinning or wrinkling behavior with respect to material

properties n and K in this sheet metal drawing process. So even the steel has different

n/K value, their trends are the same. This conclusion is very helpful because it could save

us a lot of effort to re-run another 57 time-consuming FEM simulations in order to get

RSM for steel material.

219
Steel Aluminum
n+ 0.256 0.2625
n- 0.184 0.2375
k+ 596 493
k- 496 446

Table B.1. The range of n and K for aluminum and steel materials.

220
thinning
n k bhf lub1 thinning simu error perc
pred
1 1 0 0 0.075025 0.074228614 0.010615
1 -1 0 0 0.08073 0.080705725 0.000301
-1 1 0 0 0.062263 0.062307029 -0.00071
-1 -1 0 0 0.067563 0.06878414 -0.01807
-1 -1 1 1 0.095322 0.096296131 -0.01022
-1 -1 1 -1 0.056419 0.054879634 0.027285
-1 -1 -1 1 0.083661 0.086482383 -0.03372
-1 -1 -1 -1 0.052839 0.052050062 0.014931

Table B.2. The comparison of predicted values vs. the simulation values.

221
0.12
0.1
0.08
sim ulation
0.06
prediction
0.04
0.02
0
1 2 3 4 5 6 7 8

Figure B.1. The comparison of prediction and simulation.

222

Vous aimerez peut-être aussi