You are on page 1of 142

EXPERIMENTAL AND NUMERICAL INVESTIGATIONS OF THE

BENDING CHARACTERISTICS OF LAMINATED STEEL

A Dissertation

Submitted to the Graduate School

of the University of Notre Dame


in Partial Fulfillment of the Requirements

for the Degree of

Doctor of Philosophy

by

Travis A. Eisenhour, B.S., M.S.

Edmundo Corona, Director

Graduate Program in Aerospace and Mechanical Engineering


Notre Dame, Indiana

July 2007
This document is in the public domain.
EXPERIMENTAL AND NUMERICAL INVESTIGATIONS OF THE

BENDING CHARACTERISTICS OF LAMINATED STEEL

Abstract
by

Travis A. Eisenhour

Laminated steel, as studied in this work, is a sheet metal product that consists

of two steel sheets with a relatively thin polymer layer between them. It has
been developed for applications requiring sound and vibration damping. The
objective of the study concerns bending operations, in particular how tooling and

sheet geometry, material properties, and process parameters affect the shape and
integrity of formed parts. Two cases are being considered: wipe bending and draw

bending. Investigation of each case has experimental and numerical components,


the latter using the finite element method.

Because of the particular construction of laminated steel, bending can involve


some issues that are not of concern when bending solid steel sheet. In wipe bending
an undesirable, permanent curl is present upon completion of the operation, while

in draw bending the shape of the part and the state of the polymer layer are of
concern. Therefore, it was important to establish how the geometry and material

state after bending depend on the process parameters. Experimentally, it was


necessary to design and build a set-up for each type of bending that had the
required flexibility to change tooling parameters. Numerically, the problem is

highly nonlinear because it involves large deformations in the steel sheets and
Travis A. Eisenhour

the polymer layer as well as nonlinear material models for both. In addition, for
draw bending the loading is complicated by factors such as localized deformations

under the draw beads and the sliding motion of the sheet with respect to the
tooling, which introduce loading/unloading cycles. Modeling of the mechanical
response of the polymer layer was particularly challenging, but appropriate models

have been implemented that give reasonable comparisons between experimental


measurements and numerical results. In addition, parametric studies have shown

how the curl and state of the polymer layer depend on various parameters that
are not easily varied in the experiments.

This work presents and explains the mechanics of bending laminated steel. It
is anticipated that the results can be used to improve processing of this material
during bending operations.
To Rex and Jo Ellen Eisenhour

ii
CONTENTS

FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x

ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . xi

CHAPTER 1: INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Sound Absorbing Laminated Steel . . . . . . . . . . . . . . . . . . 2
1.2 Bending of Laminated Steel . . . . . . . . . . . . . . . . . . . . . 5
1.3 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

CHAPTER 2: WIPING DIE BENDING . . . . . . . . . . . . . . . . . . . 13


2.1 Experimental Work . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Material Properties . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.3 Finite Element Analysis . . . . . . . . . . . . . . . . . . . . . . . 26
2.3.1 Finite Element Model . . . . . . . . . . . . . . . . . . . . 26
2.3.2 Simulation of Experiments . . . . . . . . . . . . . . . . . . 30
2.3.3 Parametric Study . . . . . . . . . . . . . . . . . . . . . . . 33
2.3.4 Variation of Thickness and Length . . . . . . . . . . . . . 39

CHAPTER 3: DRAW BENDING . . . . . . . . . . . . . . . . . . . . . . . 46


3.1 Experimental Work . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.1.1 Experimental Set-up and Procedure . . . . . . . . . . . . . 47
3.1.2 Experimental Results . . . . . . . . . . . . . . . . . . . . . 54
3.2 Finite Element Analysis . . . . . . . . . . . . . . . . . . . . . . . 67
3.2.1 Finite Element Model . . . . . . . . . . . . . . . . . . . . 67
3.2.2 Material Properties . . . . . . . . . . . . . . . . . . . . . . 68
3.2.3 Simulation of Experiments . . . . . . . . . . . . . . . . . . 81
3.2.4 Parametric Study . . . . . . . . . . . . . . . . . . . . . . . 96

CHAPTER 4: SUMMARY AND CONCLUSIONS . . . . . . . . . . . . . 101

iii
APPENDIX A: DETERMINATION OF THE STATIC COEFFICIENT OF
FRICTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

APPENDIX B: IMPLEMENTATION OF INTERFACE MODELS USING


ABAQUS’ UINTER USER SUBROUTINE . . . . . . . . . . . . . . . 112

APPENDIX C: IMPLEMENTATION SUMMARIES OF MONOTONIC


AND CYCLIC INTERFACE MODELS IN ABAQUS . . . . . . . . . . 118
C.1 Monotonic Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
C.2 Cyclic Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

iv
FIGURES

1.1 Frequency responses for laminated steel and solid steel showing the
damping effect of the polymer layer. . . . . . . . . . . . . . . . . . 3
1.2 Sequences of photos during bending in a wiping configuration. (a)
Solid steel test, (b) laminated steel sheet test, and (c) finite element
model predictions for laminated steel sheet. . . . . . . . . . . . . 6
1.3 Sketches that illustrate what causes the curl. (a) Each sheet bends
independently, causing a relative displacement between them. (b)
The shear stress in the polymer layer acts on the surfaces of the
sheets. (c) The shear stress induces a distributed moment that
results in each sheet bending. (d) The final shape with curling. . . 8
1.4 Sketches of different types of bending: (a) 3-point bending, (b) 4-
point bending, (c) wipe bending, (d) V-bending, and draw bending
without (e) and with beads (f). . . . . . . . . . . . . . . . . . . . 9

2.1 (a) Experimental wiping set-up mounted on a testing machine and


(b) schematic of the wiping apparatus. . . . . . . . . . . . . . . . 14
2.2 Definitions of wipe bending geometric parameters. . . . . . . . . . 16
2.3 Photograph of a typical specimen after wipe bending. . . . . . . . 17
2.4 Definition of bending angles. The deformation has been exagger-
ated for clarity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.5 Variation of punch and clamping forces during a typical experiment. 19
2.6 Overall bending angles measured experimentally for (a) Rd = 0.01 in.
and (b) Rd = 1/8 in. . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.7 Measured curl in specimens bent with die radii Rd = 0.01 and 1/8 in. 22
2.8 (a) Variation of overall bending angle with time for specimen bent
under identical conditions and (b) corresponding curl variation. . 23
2.9 (a) Variation of overall bending angle with time for specimens bent
with a die radius 1/8 in. and various strokes and (b) corresponding
curl variation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

v
2.10 Measured steel engineering stress-stain curve and points in piece-
wise linear fit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.11 Measured force-displacement curves from the lap-shear tests on the
polymer and points in piecewise linear fit. . . . . . . . . . . . . . 27
2.12 (a) Numerical model showing the specimen and contact surfaces,
(b) mesh near the bend region, and (c) detail of polymer model
using springs (the two steel layers are shown artificially separated
for clarity). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.13 Comparison of experimental results and numerical predictions for
Rd = 0.01 in. (a) overall bend angle and (b) curl. . . . . . . . . . 32
2.14 Residual shear stress distribution in polymer layer for three bound-
ary conditions (a) ‘Both’, (b) ‘Top’, and (c) ‘Bottom’. . . . . . . . 34
2.15 Comparison of experimental results and numerical predictions for
Rd = 1/8 in. (a) overall bend angle and (b) curl. . . . . . . . . . . 35
2.16 Parametric study results of varying steel sheet yield stress on the
(a) bend angle and (b) curl. . . . . . . . . . . . . . . . . . . . . . 36
2.17 Parametric study results of varying the thickness of the laminate
on the (a) bend angle and (b) curl. . . . . . . . . . . . . . . . . . 37
2.18 Parametric study results of varying the polymer layer stiffness on
the (a) bend angle and (b) curl. . . . . . . . . . . . . . . . . . . . 38
2.19 Photograph of two thicker (t = 0.084 in.) specimen after wipe
bending. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
2.20 Interface model for shear behavior of the polymer layer. . . . . . . 41
2.21 Comparison of UINTER models with and without damage for (a)
the bend angle and (b) the curl. (c) Plot of length of broken region
of specimen. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
2.22 Plot of the curl for specimen of different lengths. . . . . . . . . . . 45

3.1 Photographs of draw bending fixture, (a) front view and (b) oblique
view. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.2 Sectional view of fixture to illustrate internal features. . . . . . . . 49
3.3 Orthographic detail of binder block/die interface. . . . . . . . . . 50
3.4 Definitions of geometric parameters for wipe bending. . . . . . . . 52
3.5 Photograph of complete fixture, including punch guides. . . . . . 53
3.6 Photograph with punch guides in place. . . . . . . . . . . . . . . . 53

vi
3.7 Measured steel engineering stress-stain curves for two material sets
used in the draw bending study. . . . . . . . . . . . . . . . . . . . 56
3.8 Measured shear stress-displacement curves of the polymer layer in
lap-shear tests for the two material sets used in the draw bending
study. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.9 Photograph of a typical specimen after draw bending and definition
of the bend angles. . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.10 Detail of region around the draw bead that illustrates the separation
that occurs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.11 Plot of the punch and bead forces for an experiment with Rd = 6
mm and hb = 6 mm for Set 2. . . . . . . . . . . . . . . . . . . . . 59
3.12 Plots of the three bend angles for different die radii and bead heights
on each side. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.13 Plot of the thickness at different points on the bent specimen with
the largest thickness and spread in values. . . . . . . . . . . . . . 62
3.14 Location of the points where thickness measurements were made. 63
3.15 Plot of thickness extrema for all the draw tests. . . . . . . . . . . 64
3.16 Schematic illustrating where a mini-lap specimen was extracted
from a formed bending specimen. . . . . . . . . . . . . . . . . . . 65
3.17 Plot showing the polymer shear response after forming at location
a compared to straight and curved mini-lap specimens for Set 2. . 66
3.18 Plot showing the polymer shear response after forming at three
locations along a specimen for Rd = 6 mm, hb = 6 mm, and Set 2. 67
3.19 Illustration of the finite element model with the location of sym-
metry noted and the divisions between element size zones marked. 69
3.20 Measured steel engineering stress-strain curve and bilinear fit for
Set 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.21 Measured shear stress-displacement curves of the polymer layer in
lap-shear tests and its multi-linear fit. . . . . . . . . . . . . . . . . 72
3.22 Shear stress-displacement curves of the polymer layer in lap-shear
tests of 0.006 in. amplitude. Comparison of (a) cyclic response and
(b) load-reverse to failure response to the monotonic response. . . 73
3.23 Shear stress-displacement curves for lap-shear tests of 0.020 in. am-
plitude showing (a) comparison of cyclic and monotonic responses
plus a cyclic multi-linear fit and (b) the response for two loading
cycles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

vii
3.24 Shear stress-displacement curves of the polymer layer in lap-shear
tests of 0.015 in. amplitude, including the load-reverse to failure
and monotonic responses plus a cyclic, multi-linear fit. . . . . . . 75
3.25 Interface model for shear behavior of the polymer layer. . . . . . . 76
3.26 Sketch showing key points in the cyclic polymer behavior used in
modeling. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
3.27 The failure envelope in shear/normal stress space. . . . . . . . . . 80
3.28 Comparison of the analytical and experimental results of α1 , α2 ,
and α3 for hb = 6 mm for Set 2. . . . . . . . . . . . . . . . . . . . 83
3.29 Comparison of the analytical and experimental results of α1 , α2 ,
and α3 for hb = 8 mm for Set 2. . . . . . . . . . . . . . . . . . . . 85
3.30 Experimental results that demonstrate the affect of lubrication on
the punch force for Rd = 6 mm and hb = 6 mm for Set 1. . . . . . 86
3.31 Comparison of thickness at different points along specimen for the
standard model and experimental results for Rd = 3 mm, hb = 8
mm, and Set 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
3.32 History of (a) δ and (b) τ in the polymer for a point that experi-
enced the complete loading history possible. . . . . . . . . . . . . 90
3.33 History of (a) τ vs. δ and (b) τs and τd vs. δ in the polymer for a
point that experienced the complete loading history possible. . . . 91
3.34 History of (a) δn and (b) σn in the polymer for a point that expe-
rienced the complete loading history possible. . . . . . . . . . . . 92
3.35 History of σ11 vs ²p11
in steel for a point that experienced the com-
plete loading history possible. . . . . . . . . . . . . . . . . . . . . 93
3.36 History of (a) ²p11 and (b) σ11 in steel for a point that experienced
the complete loading history possible. . . . . . . . . . . . . . . . . 94
3.37 Final state of polymer for Rd = 6 mm and hb = 6 mm. . . . . . . 95
3.38 Parametric study results of varying steel sheet yield stress on (a)
α1 , (b) α2 , and (c) α3 . . . . . . . . . . . . . . . . . . . . . . . . . 98
3.39 Final shape for (a) σ̄o /σo = 0.90 and (b) σ̄o /σo = 2.70. . . . . . . 99
3.40 Parametric study results of varying steel sheet yield stress on (a)
punch and (b) bead forces. . . . . . . . . . . . . . . . . . . . . . . 99
3.41 Parametric study on the effect of varying hb on the angles for a
single die radius Rd = 6 mm. . . . . . . . . . . . . . . . . . . . . 100

A.1 Picture of experiment fixture for friction tests. . . . . . . . . . . . 109

viii
A.2 Picture of experiment fixture for friction tests. . . . . . . . . . . . 110
A.3 Coefficient of friction for test with clamping force of 300 lb. . . . . 111

ix
TABLES

3.1 FAILURE LENGTH FOR THE hb = 6 AND 8 mm CASES . . . 96

C.1 INFORMATION ON STATEV(i) . . . . . . . . . . . . . . . . . . 121


C.2 LIST OF PROPS(i) . . . . . . . . . . . . . . . . . . . . . . . . . 124

x
ACKNOWLEDGMENTS

I would first like to thank my advisor and mentor, Dr. Edmundo Corona. I

am very grateful for and appreciative of his guidance, enthusiasm, supportiveness,


and friendship. This work would be in shambles without his desire for detail and

constant support. It was nice to see firsthand how he balanced his family, work,
and other activities.

I am also extremely fortunate to have great parents, Rex and Jo Ellen Eisen-
hour. Without their love, guidance (slaps to the side of the head), example, and
teaching I would not be the person that I am today. They were the ones who

originally suggested engineering to me, and then they even paid for a lot of my
education. Which is especially nice because typically you have to pay people for

that kind of advice. I am grateful to my grandparents, Bud and Francie Haines,


for the love, humor, and visits.

I appreciate the time and help of my committee members Dr. David Kirkner,
Dr. James Mason, and Dr. Steven Schmid. For the wipe bending work, I would like
to thank Dr. Daniel Boss and Mr. Rich Williams both formerly of MSC Laminates

and Composites. For the draw bending work, I appreciate the contributions made
by Dr. Tom Stoughton, Dr. Siguang Xu, Dr. C.T. Wang, and Mr. Gary Telleck of

General Motors. I would also like to thank Shengjun Yin for pointing me in the
right direction when I was first learning how to use Abaqus.

The contributions and help that were provided by Leon Hluchota, Richard

xi
Strebinger, and Kevin Peters were instrumental in completing my experimental
work. I would like to thank Nancy Davis, Evelyn Addington, Judy Kenna, and

Nancy O’Connor for answering my (sometime random) questions and getting me


the necessary paperwork during my time here.
I also appreciate my fellow graduates and other friends for making my experi-

ence more enjoyable. (If I tried to list them I would probably forget someone and
never hear the end of it.)

The work was carried out with support from the Arthur J. Schmitt Foundation
and the Aerospace and Mechanical Engineering Department at Notre Dame. This

support is gratefully acknowledged.

xii
CHAPTER 1

INTRODUCTION

Laminated sheet metals come in many varieties and have a myriad of appli-

cations. They can range from simple sheet metals with decorative films used
in office partitions, ceiling panels, marine applications, and home appliances, to

aerospace-grade fiber-reinforced materials being used as fuselage skin in the new


Airbus A380 [1].
While the simplest laminates consist of two material layers, multi-layer lam-

inates are widely used as well. ARALL (Aramid fiber Reinforced ALuminum
Laminate) and GLARE (GLAss fiber REinforced laminate) are two examples of

laminates that have multiple layers and are used in the aerospace industry. They
both have fiber reinforced polymer layers between aluminum sheets and combine

good qualities of both metals and fiber composites, yielding favorable properties
such as high strength, lengthened fatigue life, good machinability, acoustic damp-
ing, and impact resistance [2]. Other laminates do not contain fiber reinforce-

ment. Rather, the metal sheets are separated by layers of homogeneous polymeric
materials. These laminates have qualities such as sound damping, reduction of

vibration, thermal insulation, and increased stiffness that make them desirable
for many applications. Laminates that consist of two metal sheets sandwiching

a thick polymer layer (normally 40-60% of total thickness) with low density are
designed mainly to achieve weight savings, when substituted for monolithic sheet

1
with comparable stiffness. Similar laminates, but with thin polymer layers (less
than 20% of the total thickness), are designed for sound-damping applications [3].

Laminates designed for weight saving have been found to save 3% of total vehicle
weight while maintaining similar torsional rigidity, bending stiffness, and stress
induced by seat loads management [4]. Laminates with dissimilar metal layers

can also be constructed to achieve properties that cannot be achieved by a single


kind of metal.

1.1 Sound Absorbing Laminated Steel

The laminates of interest in this work have steel sheets and belong to the
sound absorbing category above. Several patents have been filed in relation to

this type of product, for example [5–7]. Since sound and vibration are undesirable
in many applications, laminated steel has found uses in sheet metal products
that are in environments where sound or vibration must be minimized. The

sound and vibration absorbing characteristics of a laminated steel sample are


demonstrated in Fig. 1.1 [8], which compares the frequency responses of laminated

and solid steel sheets. The smoother response of the laminated sheet, without the
sharp resonance peaks apparent for the solid sheet, illustrates the relative damping

properties of the laminate.


The formulation of the polymer layer in laminated steel is of paramount im-
portance. Different polymer cores are available, such as acrylics and silicones,

that can be tuned to maximize damping under certain operating temperatures


and frequencies. As a part vibrates, resonant modes in the structure create local

regions of bending. These regions of bending cause shearing in the viscoelastic


polymer layer, which dampens the vibration. The shearing occurs due to the low

2
Figure 1.1. Frequency responses for laminated steel and solid steel
showing the damping effect of the polymer layer.

shear modulus of the polymer compared to that of the steel, which allows relative

displacement between the two sheets.


Laminated steel is produced by specialized manufacturers such as MSC Lam-

inates and Composites (tradename Quiet Steel), Roush (tradename Dynalam),


and others. In the case of MSC, the manufacturer receives steel coils of the
desired thickness. During the laminating process, steel sheet is simultaneously

drawn from two steel coils. The polymer is continuously applied to one side of the
sheets, which are then pressed together to create the laminate. Subsequently, the

laminated sheet is rolled into a coil and prepared for delivery to their customers.
The availability of laminated steel makes it possible to achieve weight and cost

savings in several applications. For example, since laminated steel sheet absorbs

3
sound and vibration it does not have to be further treated with damping materials
(solid steel sheet is usually treated with baked-on damping treatments and fibrous

sound absorption materials). This reduces the number of manufacturing steps


required to make a given part. Another advantage of laminated steel is that
it is easily recyclable because it can be put into a re-melt furnace without any

preparation [8].
Examples of current automotive applications of laminated steel include valve

covers, oil pans, brake shims, firewalls, and body panels. The 1999 Ford Explorer
used the laminate as a structural component, and by 2002 Ford expanded its use

to the Ford Ranger, Mercury Mountaineer, and Lincoln Navigator [9]. Since then,
the material has been used by other car manufacturers such as General Motors
and DaimlerChrysler.

Hard disk drive covers is an application for the laminate that has been de-
veloped in the computer industry. The benefits of using laminated over solid

steel sheet include eliminating 12 steps in manufacturing, reducing overall prod-


uct costs, improving sound level by 2 dB, and saving space [10].

Since laminated steel is a relatively new material, its manufacturing and form-
ing characteristics have not been fully understood yet. Issues that have been
previously addressed for solid steel sheet have had to be revisited for laminated

steel. These include, for example, studies of wrinkling in stamping processes [11]
and studies of weldability [12]. Issues related to bending of laminated steel have

also been of interest and are of particular concern in the present work.

4
1.2 Bending of Laminated Steel

Many components manufactured using laminated steel to date have been rel-
atively flat, stamped shapes. For this case no significant problems are present.
On the other hand, bending has proven to be a more challenging operation. Key

issues include springback, curl, and delamination. Springback is the recovery of


elastic deformation and occurs with solid sheets as well. On the other hand, curl

occurs when the walls of a part away from the locally bent region, which ideally
should remain straight, become curved. This makes subsequent assembly of sheet

metal products more difficult and reduces the aesthetic appeal of the product.
To illustrate springback and the development of sidewall curl during wipe
bending operations, Figure 1.2 shows sequences of photographs taken during the

bending of (a) a solid steel strip and (b) a laminated steel strip subjected to 90 ◦
bends. In these photographs, sheet samples of 4 in. in length, 1 in. in width, and

0.04 in. in thickness are clamped between the bending die and the blank holder,
located on the right. A section 1.5 in. long protrudes from the die and is bent

90◦ as the punch moves up to a maximum stoke of 0.75 in. (third photograph in
both sequences). The punch subsequently retracts to its original position. Note
that while the solid sheet bends with a straight wall, the laminated sheet curls

substantially during the bending process, and that this curl remains after bending.
Springback is clearly visible for both materials after retracting the punch, but

the curl present in the laminated steel sample gives the appearance of a larger
springback. Sidewall curl has also been experimentally observed in three- and

four-point bend tests [13, 14] and in V-press bending [15, 16].
Qualitatively, the source of wall curl is well understood. It comes from the
relatively weak resistance to shear of the polymer. In the bend region, each sheet

5
(a)

(b)

(c)

Figure 1.2. Sequences of photos during bending in a wiping


configuration. (a) Solid steel test, (b) laminated steel sheet test, and (c)
finite element model predictions for laminated steel sheet.

basically bends independently of the other. Therefore, the sheets have different

bending radii and develop a relative displacement δ between them as shown in


Fig. 1.3 (a). For simplicity, the relative displacement is taken as zero at the left

end of the sheets in the figure. If the polymer stiffness is negligible, the offset
δ would also appear at the top end as shown. However, for a finite stiffness of

the polymer this displacement causes a shear stress τ to develop and act on the
steel sheets as illustrated in Fig. 1.3 (b). The shear stress induces a distributed
moment m (of magnitude τ 2t for unit width) in the same direction on both sheets,

causing them to bend as illustrated in Fig. 1.3 (c). Thus the laminate exhibits

6
curl, and the offset between the two sheets reduces to ε at the top edge as shown
in Fig. 1.3 (d). At one extreme, a very compliant polymer allows the sheets to

slide past each other with little resistance, the induced moment is small, little curl
occurs, and ε ≈ δ. On the other hand, a very stiff polymer will make the two
steel sheets tend to bend as one and therefore reduce the curl. Maximum curl will

occur for intermediate polymer layer compliance, as will be demonstrated in Sect.


2.3.3.

Industrial bending operations are usually conducted by wipe bending, V-


bending, or draw bending. In addition, three- and four-point bending are simple

arrangements used in laboratory settings. Figures 1.4 (a) and (b) show three- and
four-point bending, respectively. In wipe bending the blank is clamped between
the die and the blank holder with clamping force Fc . Then the punch moves and

“wipes” the blank around the die; see Fig. 1.4 (c) (also see Ch. 2 for more details
on wipe bending). In V-bending the punch pushes the blank into the die (both

are V-shaped) as shown in Fig. 1.4 (d). Draw bending, shown in Fig. 1.4 (e),
is very different from the other types of bending because the blank is in tension

during bending. The tension is generated by friction between the blank, the die,
and the blank holders due to the clamping force Fc . The friction can be regulated
by the clamping force applied to the holders. In some cases, draw beads are added

to the holders to increase the tension, as shown in Fig. 1.4 (f). (See Ch. 3 for
more details on draw bending.) Manufacturing textbooks, such as [17], discuss

the bending operations mentioned here in more detail.


Work on bending of laminated steel has been conducted by several researchers
in the past few years. Combined experimental and finite element work was done in

[14] on the bending characteristics of laminated steel under three-point bending.

7
B B B
δ
τ

A A A
δ
Rd inside outside
sheet sheet

(a) (b)

B B
B ε<δ

A A A
δ
inside outside Rd
sheet sheet

(c) (d)

Figure 1.3. Sketches that illustrate what causes the curl. (a) Each sheet
bends independently, causing a relative displacement between them. (b)
The shear stress in the polymer layer acts on the surfaces of the sheets.
(c) The shear stress induces a distributed moment that results in each
sheet bending. (d) The final shape with curling.

In particular, the photographs in Figs. 4 and 7 clearly show the relative motion

between the sheets, and the curl that develops during bending for laminates with
polymer layers of different stiffnesses. Further finite element simulations of the

same set of experiments were continued in [18]. Cantilever beam, three-point


bending, tensile, and shear tests on laminated steel were examined in [19], using

experimental, theoretical, and finite element work. Draw bending without beads

8
load load load

(a) (b)
Fc Fc
5 5

2 3 3 1
1 1

2
3 2
4
Fc

(c) (d) (e)


Fc Fc
5 5

1 Blank
3 1
2 Die

3 Punch
2
4 Blank Holder

5 Binder Block
(f)

Figure 1.4. Sketches of different types of bending: (a) 3-point bending,


(b) 4-point bending, (c) wipe bending, (d) V-bending, and draw bending
without (e) and with beads (f).

was studied in [20] to see the effect of the polymer layer’s thickness and properties

on springback. These works used solid elements for both the steel sheets and
polymer layer or shell elements with dissimilar material properties through the
thickness to accommodate the polymer layer. Alternatively, if the polymer layer is

thin, it can be modeled using discrete springs or interface models, thus simplifying

9
the modeling. In both cases, the thickness of the layer is neglected, but the
polymer influence on the mechanical coupling between the two steel sheets is

retained.
Springs are available in many commercially available finite element packages
and can be used to elastically connect the two steel sheets. Each spring connects

a node pair and applies a force on each node, which acts along the line between
the two nodes. The magnitude of the force is found from the relative displacement

between the nodes and the spring’s load-deflection characteristics. Springs were
first used in vibration analysis of laminates in [21, 22], and later for bending during

forming in [23].
Interface models are generally used to represent the interaction between two
bodies that are in contact. Models that treat the polymer layer as interfaces are

more flexible but require development and programming by the user. They can
be used for a variety of purposes. For example, in [24] an interface model was

used to implement a new tool/blank contact formulation for sheet metal bending
operations. In another example [25], which is somewhat closer to the present

study, an interface model was used to predict the initiation and propagation of
damage in a laminated fiber reinforced composite. In the current problem, an
interface model can be used to specify the mechanical interaction between the

two steel sheets including inelastic effects and damage. The last two items are
not usually included in spring elements found in the libraries of commercial finite

element codes.
Both the spring and interface methods are appropriate for the laminated steel
studied here because the polymer layer is very thin (0.001 in. for a 0.042 in. thick

laminate). Depending on the aspects to be modeled and the objectives of the

10
study, one may be more appropriate than the other.

1.3 Objectives

The objectives of the work conducted in this project relate to the behavior of

sound and vibration absorbing laminated steel in industrial bending operations.


The operations considered are wipe and draw bending. In the case of wiping,
careful experimental and numerical studies have been conducted to study the

development of curl with the objective of assessing its dependence on bending


parameters such as the die radius, the die-punch clearance, and the length of the

punch stroke. The objective of the numerical effort is to develop and use a finite
element model (using the commercial program ABAQUS) of the laminate that

can replicate the observations made in the experiments, which will be detailed in
Ch. 2. The determination of the material properties of the steel and the polymer
are ancillary experimental projects conducted in parallel to make the model as

realistic as possible. Comparison between the experimental and numerical results


reveals the fitness of the model to replicate the experimental results and hence

model the bending process. Once the predictive capabilities of the model are
established, it can also be used to conduct numerical studies of the response of

the laminate to changes in parameters that are difficult to vary experimentally,


such as the material properties of the steel sheets and the polymer layer. These
numerical experiments can help further understand the behavior of the laminates.

The objective and methods of work are similar for the part of the study con-
cerned with draw bending. As in the previous case one objective is to establish the

final geometry of the specimen and its dependence on parameters similar to the
ones mentioned above, plus the level of tension in the specimen. The tension in

11
the specimen is controlled with the use of draw beads, as will be discussed in Ch. 3.
The draw beads and the radius of the die introduce a complex bending-unbending

loading history as the material passes through these regions, and that can lead
to delamination. As a result, the state of the polymer layer after completing
the bending operation is also of interest in this case, and will be studied exper-

imentally and numerically. A finite element model of the draw bending process
is also developed with the objective of replicating the experimental observations

and conducting parametric studies of the bending characteristics of the laminate


for different material and geometric parameters.

The combination of experimental and numerical methods used in this study


give a reasonably clear picture of the behavior of laminated steel under two very
different bending operations. Although some of the characteristics of laminated

steel under bending without tension (wiping, V-bending, etc) had been discussed
in the literature prior to the current work, both the experiments and analysis

presented here have been carried out with a greater attention to detail, as was
briefly discussed above. For draw bending, the results presented here appear to

be the first where draw beads and their effect on the steel and polymer response
have been addressed. The results of the investigation are presented in the next
two chapters, which are followed by a summary of the main findings.

12
CHAPTER 2

WIPING DIE BENDING

The first problem addressed in this work is wiping die bending, which was
introduced in Figs. 1.2 and 1.4. This bending operation can be used in the man-

ufacture of box-like products. In these cases, starting from a flat sheet, four
segments are bent 90◦ with respect to the base to form the sides. Sidewall curl,

which arises when bending laminated steel, can significantly complicate assembly
and degrades the appearance of the box, as the sides are not straight. This chapter

addresses the behavior of laminated steel, including sidewall curl, when subjected
to wipe bending.

2.1 Experimental Work

In order to conduct the experiments, a wiping die set-up was designed and

constructed [26, 27]. The set-up is mounted in a 20-kip servo-hydraulic testing


machine as shown in Fig. 2.1 (a). A schematic showing the components of the

set-up is shown in Fig. 2.1 (b). The 90◦ bending die °


2 and the punch guide block

3 are attached to the fixed cross-head of the testing machine through a base plate
°
1 . The blank (specimen being tested) °
° 10 is a rectangular strip clamped between

the die and the blank holder °


5 using a screw clamping mechanism consisting

of parts °
4 and °
6 through °
8 . The clamping mechanism contains a load cell

13
6 to measure the clamping force. The punch °
° 9 is attached to the actuator of

the testing machine. The specimen is bent by moving the punch up. The control

system of the testing machine allows the punch stroke and speed (up to 3.5 in/sec)
to be prescribed. The set-up was designed to test specimens with widths of up to
2 in. and free lengths (the length protruding from the die) of up to 4 in.

2
3

10

4
5
6
7 9
8

1 Base Plate 6 Load Cell

2 Die 7 Clamping Reaction Plate

3 Punch Guide Block 8 Clamping Screw

4 Pivot 9 Punch

5 Blank Holder 10 Specimen

(a) (b)

Figure 2.1. (a) Experimental wiping set-up mounted on a testing


machine and (b) schematic of the wiping apparatus.

14
The main geometric parameters of the wipe bending set-up are defined in
Fig. 2.2. The radius of the die and the punch are Rd and Rp , respectively. The

specimen has thickness t and free length L. The clearance between the punch and
the die is given by t + g. Other parameters in the experiment not shown in the
figure include the punch stroke S, the punch speed V , and the clamping force Fc .

Three dies with radii of 1/8, 1/16, and 0.01 in. were used in this study. In the
interest of brevity, results for 1/8 and 0.01 in. only will be presented. The punch

radius was kept fixed at 1/8 in. All experiments were conducted using specimens
with thickness t = 0.042 in. in a quasi-static manner with V = 4 in/min and F c =

300 lb unless otherwise noted. The specimens used in the study had width of 1 in.
In most cases, the free length L was 1.5 in. except when the effect of specimen
length was considered (see Sect. 2.3.4). In all cases, a segment approximately

2.5 in. long resided between the die and the blank holder.
After cutting the specimens to the desired size, measuring them, choosing the

die radius, and fixing the tooling to a prescribed value of g, the following testing
procedure was conducted:

Step I: Mark the free length on the specimen.

Step II: Install the specimen between the die and the blank holder

and align the mark made above with the edge of the die.

Step III: Increase the clamping force Fc to the desired value using
the clamping screw.

Step IV: Move the punch up so it just touches the specimen.

Step V: Program the stroke length and punch speed information into
the testing machine controller.

Step VI: Start the data acquisition system (which records clamping

15
Die Punch
Guide
L
Rd
t
Blank Holder
Rp

t+g
Punch

Figure 2.2. Definitions of wipe bending geometric parameters.

force, punch force, and punch displacement).

Step VII: Start the program to move the punch to its full stroke at
constant velocity and then retract to its initial position.

Step VIII: Stop the data acquisition.

Step IX: Release the clamping force and remove the specimen.

After each experiment, the geometry of the specimen was similar to the one

shown in Fig. 2.3 and was measured using an optical comparator. Two parameters
were defined to characterize the geometry of the specimens after bending. The

first is the overall bending angle, α. This angle is defined in Fig. 2.4 as the angle
between the line BB and the tangent to the specimen at point A, right at the start

16
of the bend. When the specimen was released from the blank holder the section
to the left of A, which had been clamped, also developed a small amount of curl.

That is why the tangent at A is not shown as a horizontal line in the figure. The
second is the curl parameter C. It is defined as

αb − α t
C= . (2.1)
αb

Here αb represents the local bending angle measured between the tangent at point
C (just above the bend) and the tangent at A. The bending angle at the tip, αt ,

is measured between the tangent at point D and the line EE, which is parallel to
the tangent at A. A value C = 0 indicates no curl, and increasing values correlate
with increasing curl.

Figure 2.3. Photograph of a typical specimen after wipe bending.

Figure 2.5 shows the load exerted by the punch Fp and the variation of the

17
B

αt
E
D E

αb
C α
A

Figure 2.4. Definition of bending angles. The deformation has been


exaggerated for clarity.

clamping force ∆Fc as functions of the punch travel δp during a typical experiment
with Rd = 0.01 in. and g = 0 (experimental parameters similar to those in Fig.

1.2 (b)). Both Fp and ∆Fc rise rapidly as the specimen is bent over the die.
Once the specimen is bent, Fp drops rapidly and levels off at approximately 40 lb,

mostly due to friction at the punch/specimen and the punch/guide contacting


surfaces while ∆Fc remains essentially constant at a slightly lower load. As the
punch is retracted, Fp changes sign but ∆Fc remains at approximately the same

level because it is essentially a reaction holding the springback of the specimen.


Further retraction of the punch allows the specimen to springback, hence both

loads decrease in absolute value. Once the punch retracts to δp /L = 0.11 contact

18
is lost with the specimen, and Fp goes to zero. A small, residual ∆Fc remains.
This residual value is due to the fact that the part of the specimen under the blank

holder also develops a small amount of curl. Therefore, a small load is required
to keep that part flat. Experiments conducted with different values of die radius
or gap parameter have similar curves in character but different load values. For

example, if Rd = 1/8 in. the peak loads decrease approximately by 50%.

200
Fp , ∆Fc L = 1.5 in
Fp t = 0.042 in
(lb)
150 Rd = 0.01 in
g=0
S/L = 0.5
100 Fc = 300 lb

∆Fc
50

− 50
0 0.1 0.2 0.3 0.4 0.5
δ p /L

Figure 2.5. Variation of punch and clamping forces during a typical


experiment.

The first aspect of the problem to be examined is the dependence of the overall

bending angle α on the various parameters of the problem. Figure 2.6 shows

19
measured values of α for die radii Rd = 0.01 in. and 1/8 in. Results for Rd =
0.01 in., a rather sharp bend, are shown in Fig. 2.6 (a) as a function of the punch

stroke S (normalized by L) for three gap parameters g. Ideally, the closer α is to


90◦ the better, but all experiments show lower values because of springback, wall
curl, and in some cases the fact that g was not zero. It is clear that α increases

with increasing S, although most of the increase occurs in the range S/L < 0.5.
The maximum bending angles achieved are on the order of 85◦ . These angles are

smaller than those that would be obtained when solid steel specimens of similar
thickness are bent in the same set-up.

Similar patterns can be seen from the results in Fig. 2.6 (b) for the blunter
die with Rd = 1/8 in. In this case, however, the values of α are uniformly smaller
than in the previous case, with the maximum values being on the order of 80◦ . In

the main, longer strokes and smaller die-punch clearances result in larger bending
angles.

The sidewall curl parameter C, defined in Eq. 2.1, was calculated for the cases
with g = 0 and the two die radii above (in a separate series of tests for Rd =

0.01 in.). The results are shown in Fig. 2.7. The measurements indicate that the
curl is larger for the sharper die and that it decreases with increasing stroke length
(with the exception of one point in the series with Rd = 1/8 in.). Yet, the curl

cannot be eliminated by simply extending the punch travel.


In order to study the repeatability of the two bending parameters studied

above, a set of five specimens were consecutively bent under identical conditions:
Rd = 0.01 in. g = 0, and S/L = 0.5. The maximum variations in α and C were
±0.25◦ and ±0.01 from the mean, respectively.

Because the polymer layer is visco-elastic, it can be expected that both α and

20
90
α
(Deg)
85

80

L = 1.5 in
75 t = 0.042 in
Rd = 0.01 in
g/t
70 0
0.12
0.24
65
0 0.2 0.4 0.6 0.8 1.0
S/L
stroke/L
(a)

90
α
(Deg)

80

L = 1.5 in
70
t = 0.042 in
Rd = 1/8 in
g/t
60 0
0.12
0.22

50
0 0.2 0.4 stroke/L
0.6 0.8 1.0
(b) S/L

Figure 2.6. Overall bending angles measured experimentally for (a) Rd


= 0.01 in. and (b) Rd = 1/8 in.

C will change with time after the bending operation [28]. The deformations of

several specimens were tracked over time in order to assess the changes in both α

21
0.10
C

0.08

0.06

0.04

0.02 L = 1.5 in Rd (in)


t = 0.042 in 0.01
g=0 1/8
0
0 0.2 0.4 0.6 0.8 1.0
stroke/L
S/L

Figure 2.7. Measured curl in specimens bent with die radii Rd = 0.01
and 1/8 in.

and C. Figure 2.8 (a) shows the variation of the overall bend angle over a period
of approximately 200 days for three of the specimens used in the repeatability

study mentioned above. The results show that α increased approximately 2.5◦
over this time period. Changes beyond this value are expected to be negligible

because the curves are relatively flat in the vicinity of 200 days. Similarly, the
curl of the specimens decreased by approximately 0.02 over the same period in

time, as shown in Fig. 2.8 (b). Again, the change in curl had become negligible
by the end of the measurement period.
Fig. 2.9 shows similar measurements for specimens bent using a die with Rd

= 1/8 in. and three strokes. The results show that the absolute changes in bend
angle are approximately the same for the three specimens. The curl change over

time is more pronounced for the shorter punch stroke, but it is not enough to

22
85
α
(Deg)
84

83

82 L = 1.5 in
t = 0.042 in
Rd = 0.01 in
81 g=0
S/L = 0.5

80 −3 −2 −1 0 1 2 3
10 10 10 10 (days)
log(t) 10 10 10
(a) t (days)

0.08
C

0.06

0.04
L = 1.5 in
t = 0.042 in
0.02 Rd = 0.01 in
g=0
S/L = 0.5

0 −3 −2 −1 0 1 2 3
10 10 10 10
log(t) (days) 10 10 10
(b) t (days)

Figure 2.8. (a) Variation of overall bending angle with time for specimen
bent under identical conditions and (b) corresponding curl variation.

decrease the curl to the level of the other two experiments.


Since the parameters α and C change with time, it is important to compare

23
α
(Deg) 84 S/L
0.33
0.5
82 1.0

80

78
L = 1.5 in
t = 0.042 in
76 Rd = 1/8 in
−3 −2 −1 0 1 2 3
10 10 10 10 10 10 10
(a) t (days)

0.08
C S/L
0.33
0.5
0.06 1.0

0.04

0.02 L = 1.5 in
t = 0.042 in
Rd = 1/8 in
0 −3 −2 −1 0 1 2 3
10 10 10 10 10 10 10
(b) t (days)

Figure 2.9. (a) Variation of overall bending angle with time for
specimens bent with a die radius 1/8 in. and various strokes and (b)
corresponding curl variation.

their measured values between different cases at a common time. The results

previously shown in Figs. 2.6 and 2.7 were measured immediately after the bending

24
operation was completed.

2.2 Material Properties

To better understand the experimental work and to get parameters needed for

the finite element analysis to be presented in Sect. 2.3, the material properties of
the steel sheets and polymer layer had to be determined experimentally. The ma-
terial properties of the steel were obtained from uniaxial tension tests on coupons

cut in the rolling direction of the sheet. The stress was calculated based on the
tensile load, measured with a calibrated load cell, and the strain was measured

directly on the specimen with an axial extensometer. The uniaxial engineering


stress-strain (σ-²) curve is shown in Fig. 2.10. Young’s modulus and the 0.2 %

offset yield stress are given in the figure.


Several lap-shear tests were conducted on appropriately machined laminated
steel specimens to better understand the mechanical behavior of the polymer layer.

The insert in Fig. 2.11 shows a sketch of the specimen. In the actual tests the
dimensions were W = 0.75 in. and b = 1 in. The relative motion between the

sheets was measured directly on the specimen via an axial extensometer. The
displacement rate in the test was 0.03 in/min, which is in the vicinity of shear

rates estimated to occur in the bending experiments. The three force-displacement


curves shown in Fig. 2.11 exhibited some scatter, even for specimens identically
machined from the same mother sheet and tested under identical conditions. The

range of scatter is shown by the dashed lines in the figure. The curves show a
hardening shear response at low strains, followed by a nearly linear region that

terminates in a relatively sharp peak as the polymer fails and the force drops
precipitously.

25
σ
(ksi)
40

30

Deep draw quality steel


20 E = 30.0 x 10 3 ksi
σo = 25.8 ksi

10 Test
Fit

0
0 0.05 0.1 0.15 0.2 0.25
ε

Figure 2.10. Measured steel engineering stress-stain curve and points in


piecewise linear fit.

2.3 Finite Element Analysis

The second part of the investigation consisted of the development and use of

a finite element model to simulate the experiments and to conduct parametric


studies of the problem.

2.3.1 Finite Element Model

In order to reduce computation time, the problem was reduced to two dimen-
sions by assuming a state of plane strain. This assumption is justified by the large

width-to-thickness ratio of the specimens tested. A few cases were simulated using
plane stress to test the sensitivity of the results to the choice of two-dimensional
model. The differences between the results produced by the two models were very

small.

26
350
P P
(lb) 300 Test
Fit
250

200 b

150 W

100
P
50

0
0 0.01 0.02 0.03 0.04 0.05
δ (in)

Figure 2.11. Measured force-displacement curves from the lap-shear


tests on the polymer and points in piecewise linear fit.

A plane strain model with four-noded, bilinear, reduced-integration, quadri-


lateral elements (Abaqus CPE4R [29]) for the steel sheets was considered. Figure

2.12 (a) shows the overall view of the model, with the specimen in the initial,
straight configuration. The die, punch, and blank holder were modeled as rigid

surfaces that made contact with the specimen. The contact conditions between
the specimen, the die, and the blank holder were modeled with Coulomb friction

and a coefficient of friction of 0.17, as measured in an ancillary study presented


in Appendix A. Figure 2.12 (b) shows a close up view of the bend region that re-
veals the mesh used. The sheet in contact with the blank holder had two elements

through the thickness while the one in contact with the die had four. The element
density in the longitudinal direction varied along the specimen. Away from the

bend, the element length was 0.075 in. to the far left and 1/16 in. to the far right.
A region 0.5 in. near the bend had a refined mesh in both sheets, where the length

27
of the elements was 0.01 in. Further refinement in the upper sheet at the bend
was required for cases with Rd = 0.01 in. as shown. The optimum element sizes

and the extent of the refined regions were determined by parametric studies of
the effect of the mesh on the predicted value of the overall bend angle α. The
studies demonstrated that the value of α calculated using the adopted mesh has

converged within ±0.5◦ .


The shear response of the polymer layer has been modeled using two methods:

nonlinearly elastic springs (discussed here) and an interface model (discussed in


Sect. 2.3.4). Chronologically, the elastic spring model was developed much earlier

than the interface model. At the time, it was used in the simulation of the exper-
imental results in Sect. 2.1. Interface models have been used in the simulation of
later experiments that will be discussed in Sect. 2.3.4 as well as in the simulation

of draw bending to be presented in Ch. 3.


The springs (Abaqus axial connector [29]) were attached to nodes in the steel

sheets as shown in Fig. 2.12 (c). It is important to attach the springs exactly as
shown to ensure that the most severely deformed springs are stretched. Otherwise

compression larger than the length of the springs may occur, which yields erro-
neous results. The length of the springs was equal to the spacing of the nodes in
the lower sheet in Fig. 2.12 (b). For clarity, the sheets are shown separated in Fig.

2.12 (c). In the model, however, the two steel sheets were constrained to remain
in contact at every point along the specimen (no separation was observed in the

wipe bending experiments presented until now). As it stands, the model cannot
account for the geometric changes as function of time shown in Fig. 2.8. The
numerical results will be compared only to experimental measurements conducted

within minutes after each experiment.

28
Die Specimen
A

Blank Holder Punch

(a)

(b)

a
(c)

Figure 2.12. (a) Numerical model showing the specimen and contact
surfaces, (b) mesh near the bend region, and (c) detail of polymer model
using springs (the two steel layers are shown artificially separated for
clarity).

The constitutive model for the steel sheets was J2 flow theory plasticity with
isotropic hardening. The material properties were obtained from the uniaxial

tension tests discussed in Sect. 2.2, using the piecewise linear fit shown in Fig.
2.10. A Poisson ratio of 0.3 was assumed. The load-displacement relation for

the springs modeling the polymer core was obtained from the piecewise fit of the

29
lap-shear test results shown in Fig. 2.11, as will be discussed next.
Because the polymer layer was modeled by discrete spring elements of different

initial lengths, their force-displacement response had to be related to the data in


Fig. 2.11. For each value of δ the force of the spring, f , is given by

a Wb
f =P , (2.2)
bW

where a is the initial length of the spring as shown in Fig. 2.12 (c), Wb is the width
of the bending specimens (unit width), W is the width of the lap specimen, and

b is the length of the overlap (see insert in Fig. 2.11). Analysis of the numerical
results indicated that the shear displacements in the polymer layer remained below
the value at the load peak throughout the bending process for the materials and

geometries considered in Sect. 2.1.

2.3.2 Simulation of Experiments

Figure 1.2 (c) shows the predicted specimen shapes that correspond to the

photographs in Fig. 1.2 (b). Qualitatively, the observed and predicted shapes
are in very close agreement. In the following, quantitative comparisons between

experiments and analysis will be carried out.


Figure 2.13 shows results of the analysis, compared with experimental mea-

surements for both the bending angle α and curl parameter C as a function of
punch stroke for bending with die radius Rd = 0.01 in. and g = 0. Two experi-
mental series are shown in Fig. 2.13 (a). In general, some spread of experimental

results as shown would result if the set-up was removed and the reinstalled. The
differences between the two series decrease for larger punch stokes. This indi-

cates that in actual production scenarios larger punch strokes may lead to more

30
consistent geometries after bending. The curl, shown in Fig. 2.13 (b), was only
measured for series 1. Three series of predictions are also shown in the figure.

Each series corresponds to one boundary condition implementation at the end A


of the specimen (shown in Fig. 2.12 (a)). In the series labeled ‘Both’, the horizon-
tal displacement of both sheets was restricted at that point. In the series labeled

‘Bottom’ only the lower sheet was restricted, while in the series ‘Top’ only the
upper sheet was restricted. In all three cases, the comparisons between analyt-

ical and experimental results are reasonable once S/L > 0.3 for both α and C.
The implementation of the boundary conditions makes the most difference for the

shortest stroke. Here, the curl predicted by the ‘Top’ series is much closer to the
experimental results, although the prediction for α is highest.
Because the curl is a direct result of the shear stress that develops in the

polymer, which is calculated from

f
τ= , (2.3)
Wb a

it is interesting to look at its distribution along the length of the specimen at the
conclusion of the bending simulation. Figure 2.14 shows these results (τ has been
normalized by τp , the shear stress corresponding to the peak force shown in Fig.

2.11) for the three boundary conditions considered and S/L = 0.5. The bend is
located at x/L = 0. The ‘Both’ and ‘Bottom’ cases have very similar distributions,

with negligible stress in the part of the specimen under the blank holder, relatively
drastic changes near the bend and then a smoothly decreasing shear stress as the
tip of the specimen is approached. The ‘Top’ case allows the sheets to slide under

the blank holder and therefore displays a slightly larger stress (in absolute value)
in that region. The peak stress near x/L = 0 is smaller than in the other two

31
90
α
(Deg)
85

80 L = 1.5 in
t = 0.042 in
Rd = 0.01 in
75 g=0
Exp. series 1
Exp. series 2
70 Both
Top Analysis
Bottom
65
0 0.2 0.4 0.6 0.8 1
(a) S/L
0.2
C L = 1.5 in
t = 0.042 in
0.16 Rd = 0.01 in
g=0
Experiment
0.12 Both
Top Analysis
Bottom
0.08

0.04

0
0 0.2 0.4 0.6 0.8 1
(b) S/L

Figure 2.13. Comparison of experimental results and numerical


predictions for Rd = 0.01 in. (a) overall bend angle and (b) curl.

cases, but the stress levels for x/L > 0 are about the same in all cases. One would

expect that the ‘Top’ and ‘Bottom’ cases would allow the sheets to slide relative

32
to each other under the blank holder. In the ‘Bottom’ case, however, the top sheet
tends to get locked at the bend due to the sharp die radius. Therefore, it gives

essentially the same results as the ‘Both’ case. Only in the ‘Top’ case, the lower
sheet is free to slide, and the results are different. Although the ‘Top’ case seems
to be more physically correct, numerical convergence was at times problematic,

so it was not pursued further.


Clearly, high residual shear stress in the polymer can be expected. Given the

viscoelastic characteristics of the polymer, it is not surprising that some of this


stress will be relaxed with time by changes in geometry with the results shown in

Figs. 2.8 and 2.9.


Figure 2.15 shows results similar to those in Fig. 2.13 but for Rd = 1/8 in.
As in the Rd = 0.01 in. case, α is reasonably predicted, although the predictions

are somewhat high. The predictions for C are also reasonable. Note that while
the ‘Bottom’ case yields more favorable predictions for the curl, the bend angle is

slightly better predicted by the ‘Both’ case. This contrasts with the fact that both
of these cases produced almost identical results for Rd = 0.01. A possible reason

for this difference is that the blunter die radius does not restrain the motion of
the upper sheet as much, which can then slide more freely over the die.

2.3.3 Parametric Study

The finite element model developed above yielded results that were reason-
ably representative of the experiments. In this section, the same model is used
to perform parametric studies of the dependence of the bend angle and curl on

variations of the yield stress of the steel, the thickness of the steel sheets, and
the stiffness of the polymer layer. The boundary condition ‘Both’ was used in all

33
1
τ / τp L = 1.5 in
0.8 Both
t = 0.042 in
0.6 Rd = 0.01 in
0.4 g=0
0.2 S/L = 0.5
0
−0.2
−1.5 −1 −0.5 0 0.5 1
(a) x/L
1
τ/ τp
0.8
0.6 Top
0.4
0.2
0
−0.2
−1.5 −1 −0.5 0 0.5 1
(b) x/L
1
τ / τp
0.8
0.6 Bottom
0.4
0.2
0
−0.2
−1.5 −1 −0.5 0 0.5 1
(c) x/L

Figure 2.14. Residual shear stress distribution in polymer layer for three
boundary conditions (a) ‘Both’, (b) ‘Top’, and (c) ‘Bottom’.

cases.
The dependences of the bend angle and curl on variations of the yield stress

of the steel sheets are shown in Fig. 2.16. The values of the geometric parameters
used are shown in the insert. In this study the shape of the steel stress-plastic

strain curve was kept constant, but the stress was raised or lowered at each plastic
strain value by the same amount to match the desired yield stress. The horizontal

34
Comparison of
90
α
(Deg) 85

80

75
L = 1.5 in
70 t = 0.042 in
R = 1/8 in
65 g=0
Experiment
60
Both
Bottom Analysis
55

50
0 0.2 0.4 0.6 0.8 1
(a) S/L

0.1
C
0.08

0.06

0.04

L = 1.5 in
0.02 t = 0.042 in Experiment
Rd = 1/8 in Both
Analysis
g=0 Bottom
0
0 0.2 0.4 0.6 0.8 1
(b) S/L

Figure 2.15. Comparison of experimental results and numerical


predictions for Rd = 1/8 in. (a) overall bend angle and (b) curl.

axes show the ratio of the yield stress (σ̄o ) to the yield stress of the curve in Fig.

2.10 (σo = 25.8 ksi). The response of the polymer layer was still dictated by the

35
curve in Fig. 2.11. The results indicate that increasing the yield stress causes a
decrease in the bend angle and an increase in the curl. Physically, the bending

operation is essentially deformation controlled. Therefore, the strains generated


during the bending phase are independent of the yield stress, but the elastic strain
is larger for materials with higher yield stress. Upon unloading (punch retraction),

the larger amount of recoverable strain leads to lager springback, which results
in smaller values for the local bending angles αb and αt , and yield smaller α and

larger C.

90 0.1
α L = 1.5 in
(Deg) C
t = 0.042 in
88 Rd = 0.01 in 0.08
g=0
S/L = 0.5
86 σo = 25.8 ksi 0.06

L = 1.5 in
84 0.04 t = 0.042 in
Rd = 0.01 in
g=0
82 0.02 S/L = 0.5
σo = 25.8 ksi

80 0
0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5
(a) σo / σo (b) σo / σo

Figure 2.16. Parametric study results of varying steel sheet yield stress
on the (a) bend angle and (b) curl.

Figure 2.17 shows the effect of varying the thickness of the laminate (t̄). Both

steel sheets had equal thickness. The results indicate that the bend angle increases
and the curl decreases as the thickness of the laminate increases. These trends are

36
the result of two competing effects. On one hand, a thicker sheet results in larger
relative displacements between the sheets at the bend zone. This has the effect of

raising the stress in the polymer layer and therefore increasing the moment on the
sheets, which tends to increase C. On the other hand, thicker sheets are stiffer
and present more resistance against the development of curl. The trends in the

figure represent the combined effects. Geometrically, an increase in C is generally


accompanied by a reduction in α. Raising t̄/t above one results in shear strains

in the polymer exceeding the failure point. Because the elastic spring polymer
layer model does not account for failure, results are not included for t̄/t > 1 in

Fig. 2.17. Cases with t̄/t > 1 will be discussed in Sect. 2.3.4, where the interface
model will be used in the predictions.

90 0.1
α
L = 1.5 in C
(Deg)
t = 0.042 in
88 Rd = 0.01 in 0.08
g=0
S/L = 0.5
86 0.06

84 0.04
L = 1.5 in
t = 0.042 in
82 0.02 Rd = 0.01 in
g=0
S/L = 0.5
80 0
0.5 0.625 0.75 0.875 1 0.5 0.625 0.75 0.875 1
(a) t/t (b) t/t

Figure 2.17. Parametric study results of varying the thickness of the


laminate on the (a) bend angle and (b) curl.

37
The third parametric study addressed the effect of the stiffness of the polymer
layer. For simplicity, the response of the polymer was set to be strictly linear with

slope K̄ as shown in the inserts in Fig. 2.18. The horizontal axes present the ratio
of the stiffness to the average slope of the ascending curve in Fig. 2.11 between 50
and 325 lb. The results indicate a high bending angle and zero curl for K̄ = 0.

As the stiffness increases, the bend angle decreases and the curl increases rapidly.
The maximum curl occurs at K̄/K = 1. Further increase in K̄ results in relatively

slow changes in bend angle and decreasing curl as shown. The physical reasons
for the observed variation in curl were explained in Sect. 1.2. Briefly, no curl is

expected when K̄ = 0 or ∞, so C must increase to a maximum and then decrease


with increasing K̄.

90 0.1
α L = 1.5 in
L = 1.5 in C
(Deg)
τ t = 0.042 in t = 0.042 in
88 Rd = 0.01 in 0.08 Rd = 0.01 in
K g=0
1 g=0
S/L = 0.5 S/L = 0.5
86 K = 23.2 ksi/in 0.06 K = 23.2 ksi/in
δ

84 0.04 τ
K
1
82 0.02

δ
80 0
0 2 4 6 8 10 0 2 4 6 8 10
(a) K/ K (b) K/ K

Figure 2.18. Parametric study results of varying the polymer layer


stiffness on the (a) bend angle and (b) curl.

38
2.3.4 Variation of Thickness and Length

The results presented in Fig. 2.17 demonstrated that thickness plays a role

in the final geometry of the specimens. The reasons for this were pointed out in
the associated discussion. The trend presented indicated that as the thickness

increases, α gradually increases and C decreases. While generating these results,


care was taken to ensure that the relative displacement between the sheets re-

mained below δ = 0.026 in., where the load peaks, to avoid failure of the polymer
layer. Although the fit presented in Fig. 2.11 models the sudden drop in strength
at failure, one must remember that the springs are elastic and follow the same

curve upon loading and unloading. This behavior was quite unrealistic since the
failure of the polymer layer has to be irreversible.

A small series of experiments was conducted on specimens with thickness of


0.084 in. The specimens had two identical steel sheets of thickness 0.0414 in.

and a thin polymer layer of essentially the same thickness as the specimens used
previously. The stress-strain curve of the steel had a yield stress 38 ksi, which is
47% higher than that shown in Fig. 2.10. The polymer layer had a very similar

response to that in Fig. 2.11. Figure 2.19 shows a photograph of two specimens
after bending. It is obvious that the curl is negligible in both cases. Note that

one specimen showns delamination, indicating that failure of the polymer layer
occurred.

In order to include irreversible failure in the polymer layer model, a simple


interface model was developed. The shear behavior of the model is modeled using
a spring-dashpot system as shown in Fig. 2.20, where τ is the stress developed by

the relative motion of the sheets δ. The stress in the spring τs obeys the nonlinear
relation shown in Fig. 2.11. It was necessary to add some damping to the model

39
Figure 2.19. Photograph of two thicker (t = 0.084 in.) specimen after
wipe bending.

in order to obtain numerical convergence. As a result, a linear dashpot was added

to the model. The stress generated by the dashpot is given by

τd = cδ̇. (2.4)

Here, δ̇ is the rate of change of δ with respect to “time.” Note that here “time”

is just a parameter proportional to the prescribed loading and does not represent
“real time.” In all runs conducted, the time length to complete the analysis was

kept the same for consistency. The damping constant c was chosen to be the
smallest value that yielded converged solutions that closely agreed with those

40
obtained with the spring model. A value of 200 psi/(in/unit of time). The total
stress in the model is given by the sum of the two contributions

τ = τs + τd (2.5)

Figure 2.20. Interface model for shear behavior of the polymer layer.

The interface model must also address the interaction between the sheets in

the normal direction. To emulate the conditions used in the spring model that
prevent separation and penetration of the contacting surfaces, a simple linearly
elastic behavior given by

σn = k n δ n (2.6)

is adopted. Here, σn and δn are the stress and relative motion of the sheets in the

normal direction. The constant kn was chosen to have a large value 109 psi/in.
The polymer layer behavior in shear shown in Fig. 2.11 exhibits a sharp drop
in load if the relative displacement between the sheets exceeds a critical value.

This is the result of irreversible damage with the result that the load-carrying

41
capacity of the layer is almost completely eliminated. In the interface model,
damage is implemented by setting τs = 0 for all displacements δ after it exceeds

the critical value (0.026 in.) once. This model was implemented in Abaqus as a
user subroutine (UNITER [29]) as explained in Appendices B and C.
The results of this thickness study are shown in Fig. 2.21. Two sets of results

are shown for the range 0.5 ≤ t̄/t ≤ 2.25, which includes much thicker laminates
than Fig. 2.17. In one, the interface model for the polymer did not account for

damage, hence τs continues to increase indefinitely with increasing δ, yielding the


results shown by triangles. In this case the bend angle in Fig. 2.21 (a) showed little

dependence on thickness, and the curl in Fig. 2.21 (b) showed a small decrease,
but remained significant. As expected, these trends are continuations of those
presented in Fig. 2.17. The second interface model, which accounted for damage

of the polymer layer, yielded the results shown by circles. For specimens with
thickness up to thickness ratio of 1.75 the results are nearly coincidental with

those of the model without damage. For thicker specimens, however, the bend
angle suddenly jumps to the 88◦ level and the curl decreases to nearly zero.

The reasons for these abrupt changes in the trends can be explained with the
help of Fig. 2.21 (c), which shows the extent of damage in the specimen at the
conclusion of the bending process. For t̄/t ≤ 1 no damage takes place and the

results for bend angle and curl coincide with those obtained without accounting
for damage. Starting with t̄/t = 1.25 damage starts appearing for a short length

of the polymer layer near the bend. The damaged length continues to increase
gradually for thicker laminates. For t̄/t ≥ 1.875 the damage propagates along the
complete length of the polymer between the bend region and the tip. The total

failure of the polymer layer decouples the two sheets and allows them to bend

42
90
α L = 1.5 in Rd = 0.01 in
(Deg) 88 t = 0.042 in S/L = 0.5
86
84
Model with Damage
82 Model without Damage
80
0.5 0.75 1 1.25 1.5 1.75 2 2.25
(a) t/t
0.1
C 0.08
0.06
0.04
0.02
0
0.5 0.75 1 1.25 1.5 1.75 2 2.25
(b) t/t

Lb 1

0
0.5 0.75 1 1.25 1.5 1.75 2 2.25
(c) t/t

Figure 2.21. Comparison of UINTER models with and without damage


for (a) the bend angle and (b) the curl. (c) Plot of length of broken
region of specimen.

independently, thus eliminating the curl and increasing the bend angle. Since the
polymer has suffered damage, the damping and other properties of the laminate
may have been affected.

Another geometric parameter that has been held constant up until now is the
length of the specimen. Four-point bending experimental results presented in [13]

suggest that curl during wipe bending will increase with the free length of the

43
specimen. Figure 2.22 shows how varying the free length (L̄) affects the curl. In
the figure, L̄ has been normalized by L = 1.5 in. The values of all parameters

kept constant are given in the figure. The figure contains two experimental points
at L̄/L = 1 and 2, shown by circles, plus four numerical predictions with L̄/L
between 0.5 and 2 shown by squares. In all cases the section of the specimen

clamped between the die and the blank holder was constant and equal to 2.5 in.
The predictions show good agreement with the experimental points and indicate

a roughly linear increase in curl with specimen length, confirming the suggestions
inferred from four-point bending. The physical reason for the relationship between

curl and specimen length can be appreciated by first noting from Fig. 2.14 that
the shear stress in the polymer is relatively constant in the free length away from
the bend region. Since the shear stress induces a distributed moment and the

internal bending moment has to be zero at the tip of the specimen, the bending
moment in the sheets near the bend region has to increase with specimen length,

thus inducing more curl. The bend angle α has not been shown because it depends
on the length of the specimen even if the curl were to stay constant. Hence it is

an appropriate parameter only when comparing cases with equal length.

44
0.12
C
0.1

0.08

0.06

0.04 L = 1.5 in
t = 0.042 in
Rd = 0.01 in
0.02 Experiment g=0
Analysis S/L = 0.5
0
0.5 0.75 1 1.25 1.5 1.75 2
L/L

Figure 2.22. Plot of the curl for specimen of different lengths.

45
CHAPTER 3

DRAW BENDING

The second problem addressed in this work is that of draw bending. This
bending operation was introduced schematically in Figs. 1.4 (e) and (f). Here,

bending takes place in the presence of tension as explained in Sect. 1.2. Deep
drawing operations are commonly used in the automotive and other industries to

manufacture products such as car trunk wells, fuel tanks, kitchen sinks, etc. [17].
Many of these parts have complex, three-dimensional shapes, and their manufac-

turing operations can be affected by wall wrinkling and thinning instabilities that
are influenced by the level of tension in the blank. In addition, use of laminated
steel can influence the final shape of the part compared to solid sheet as discussed

in Ch. 2. Delamination is another issue that can arise when using this material.
The case to be studied here has a very simple geometry, and the objective of

the work is to make an initial assessment of the behavior of laminated steel in


deep drawing operations. As in the previous chapter, experimental results will be
presented first, followed by numerically generated results, comparisons between

experimental measurements and numerical predictions, and parametric studies.

46
3.1 Experimental Work

3.1.1 Experimental Set-up and Procedure

In order to preform the experimental part of the study, a draw bending fixture

that uses draw beads to control the tension in the blank, as shown in Fig. 1.4
(f), was designed and constructed. Figure 3.1 shows photographs of two different

views of the fixture. In addition, a sectional view showing some of the internal
features of the fixture is shown in Fig. 3.2.
The fixture is designed to bend an initially flat specimen of nominal dimensions

12 × 2 × 0.04 in. into a symmetric U-channel. It consists of a heavy steel base 13


in. long by 7.5 in. wide by 2.5 in. high that is mounted on the actuator rod of the

same 20-kip testing frame used in the wipe bending experiments presented in Ch.
2. Two die holder blocks are mounted symmetrically on the plate, each using four

1/2 in. bolts. The positions of the blocks on the plate are fixed using a set of two
dowel pins, which are press-fit in the plate, in each block, as shown in Fig. 3.2.
One bending die mounts on each die holder using four 3/8 in. bolts. Alignment

holes in the dies match dowel pins press-fit in the die holders to ensure proper
positioning. The dies are reversible, so each of the four corners can be used in

bending. Each corner has radius of 3, 6, 9, or 12 mm (0.118, 0.236, 0.354, 0.472


in). The die radii and other critical dimensions were chosen in consultation with

sheet metal forming experts at General Motors.


In order to generate tension during bending, the motion of the specimen is
restricted by draw beads whose location is shown in Fig. 3.2. A more detailed

schematic of the parts associated with the draw beads is shown in Fig. 3.3. The
draw bead fits in a binder block located above the die. The draw bead channel

in the die can accommodate beads of radius 5.3 mm (0.209 in.) and up to 8 mm

47
Figure 3.1. Photographs of draw bending fixture, (a) front view and (b)
oblique view.

48
Figure 3.2. Sectional view of fixture to illustrate internal features.

(0.315 in.) in height. Each bead fits in a slot in one binder block and is secured
to it using two screws. The smaller bead height used in this study was 6 mm.

Adding an insert in the slot allows increasing the height of the bead up to the 8
mm maximum, which was the largest value used. The specimen passes through a

small channel in the binder block so that the bead is the major restriction to the
motion of the specimen. Therefore, the bead geometry and the die radius control
the tension in the specimen during bending.

Horizontal motion of the binder blocks with respect to the die is prevented by
a pair of 5/8 in. dowel pins attached to each die. Each binder block has two holes

49
Load
Dowel Cell Dowel
Load Pin Pin
Cell
Bead
Binder Binder
Block Block

Specimen

Specimen

Die

Figure 3.3. Orthographic detail of binder block/die interface.

that engage the pins to maintain its horizontal position. The manner in which
the binder blocks are positioned is seen in the photograph in Fig. 3.1 (b) and in

the schematic in Fig. 3.2. The mechanism consists of a frame attached to each
die with two shoulder screws. At the top of each frame, a clamping screw presses

down on a plate that carries a load cell. It, in turn, pushes each binder block
and bead against the specimen, therefore locally bending it into the draw bead
channel in each die. The load cell indicates the vertical reaction force on the bead

and binder block. To ensure that the measured force values are not influenced the
contact between the binder block and the top surface of the die, a 0.002 in. shim

is placed at the top of the dies while clamping. The gap size is verified prior to
bending by sliding the shims between the binder block and the die.

The punch is directly attached to the load cell of the testing machine and
remains stationary during bending. The load cell provides a direct reading of the

50
punch force required to conduct the operation. The punch is symmetric and each
corner has a radius of 3 mm (0.118 in.). All surfaces in contact with the specimens

(punch, dies, and draw beads) have been machined from tool steel, heat treated
for hardness, and ground to smooth, mirror finishes.
The main geometric parameters of the draw bending set-up are illustrated in

Fig. 3.4. Because of symmetry, only half of the assembly is shown. Many of these
parameters were relevant in the wipe bending study. The radius of the die and

punch are Rd and Rp , respectively. The specimen has thickness t and undeformed
length Ls (not shown). The clearance between the punch and each die is given by

t + g. The bead has dimensions of height hb and radius Rb (not shown). Other
parameters not shown in the figure include the punch stroke S and the speed of
the stroke V . As previously mentioned, die radii of 3, 6, 9, 12 mm are available.

The punch radius is fixed at 3 mm. The speed V is low enough so the operation
is quasi-static.

Installation of the fixture in the testing frame must be conducted in a careful


manner to ensure proper alignment and spacing between the dies and the punch.

This is done by the addition of shims between the dies and the die holders and
adjusting the lateral position of the punch. In the work conducted here, the pa-
rameter t + g was set at 0.045 ± 0.001 in. Since the actuator rod of the testing

machine is free to rotate, punch guides must be used to ensure against misalign-
ment when the actuator moves. Figure 3.5 shows the complete configuration of

the fixture as used during the experiments, while Fig. 3.6 shows the detail design
of the punch guides. The most essential role of the punch guides is to ensure that
the fixture remains aligned with the punch as it first makes contact with the spec-

imen. Therefore, the guides extend above the top surface of the dies. In Fig. 3.6

51
Punch
t+g

Binder Block Rp
Bead
t
hb Rd

Die

Line of
Symmetry

Figure 3.4. Definitions of geometric parameters for wipe bending.

note that the guides have different sizes. The ones in the back always remain in

contact with the punch, while the ones in the front are shorter and can disengage
the punch to allow removal of the specimen after bending.
The specimens are cut with a shear from large sheets, so they have the di-

mensions indicated earlier. Prior to bending, the edges are filed to remove any
roughness introduced by cutting, and the specimens’ initial dimensions are mea-

sured using calipers and a micrometer. The bending procedure is as follows:


Step I: Lubricate the specimen with a heavy oil and carefully align

it on the dies.

52
Figure 3.5. Photograph of complete fixture, including punch guides.

Figure 3.6. Photograph with punch guides in place.

53
Step II: Place the binder blocks and frames over each die.

Step III: Move the actuator until the punch is just about to touch
the specimen.

Step IV: Program the stroke length and speed into controller of the
testing machine. In all experiments to be presented, the stroke

length was 2.5 in. while the speed was 1 in/min during forming
and 2 in/min during retraction.

Step V: Start the data acquisition system (recording binding forces,


punch force, and stroke).

Step VI: Tighten the clamping screws so the beads locally indent the

sheet to a predetermined depth. Verify the clearance between the


binder block and the top of the die.

Step VII: Start the program to move the punch to its full stroke at
constant velocity and then retract to its initial position.

Step VIII: Stop the data acquisition.

Step IX: Release clamping loads.

Step X: Dismount the frames, remove the binder blocks and the spec-

imen.

3.1.2 Experimental Results

Two different material sets were used in the draw bending study, which are

called “Set 1” and “Set 2”. The differences between the two can be seen in the
tension tests in Fig. 3.7 and lap-shear tests in Fig. 3.8. The steel of Set 1 has a
higher yield stress than that of 2, but the shapes of the curves after yielding are

54
very similar. The shear responses of the polymer layer are very different between
the two types of material. In Set 1 it is much stiffer and stronger than in Set

2. Both achieve stress maxima at similar values of relative displacement δ, but


the polymer layer in Set 1 breaks at a larger value. Another difference is the
mode of failure of the polymer layer. For Set 1, the type of failure observed in

the lap-shear tests was interface failure, where failure occurs at the steel/polymer
interface. Upon inspection, all the polymer remained attached to a single sheet,

while the other sheet was clean. On the other hand, for Set 2 an even coating of
polymer remained on both sheets, indicating a cohesive failure. These modes of

failure agree with observations in [30]; indicating that cohesive failure correlates
with low polymer layer shear strength, while interface failure correlates with high
shear strength. On a side note, the materials here are different from the material

used in the wipe bending study. In that case, the steel stress-strain behavior was
very similar to Set 1, but the polymer had the same lap-shear behavior as Set 2.

Figure 3.9 shows a photograph of a typical specimen after bending. The stroke
used ensures that a point in the specimen that starts in a region behind a draw

bead experiences a complete loading history, going around the draw bead and
over the die radius. The walls of the specimen exhibit some curl, but it must
be noted that even monolithic sheets exhibit curl when bent under tension [31].

Furthermore, it has also been shown that drawing sheet through beads results in
curvature of the specimen [32]. The question of interest here relates to the effect

of die radii and draw bead height on the shape of laminated steel specimens after
bending. The shape of the specimens will be quantified by the angles α1 , α2 , and
α3 in Fig. 3.9 in a manner similar to [33].

Another interesting observation in this experiment is that some separation

55
σ 40
(ksi)

30

20 Set 1
Set 2

10

0
0 0.02 0.04 0.06 0.08 0.1
ε

Figure 3.7. Measured steel engineering stress-stain curves for two


material sets used in the draw bending study.

between the sheets was observed in the material that was passing around the

draw beads upon completion of the experiment. This is more clearly shown in the
close up in Fig. 3.10, on the right side of the bead indentation. It appears that the

bending/reverse bending history under the bead is the reason for this separation
between the sheets. Separation has also been observed under four-point bending

in [34]. Note that the separation is not obvious beyond the bead region. The
strength of the polymer layer, however, may have suffered degradation. Therefore,
experimental and numerical investigations of the state of the polymer layer after

bending will be discussed in detail later, because it may influence the integrity
and acoustic properties of the laminate.

Figure 3.11 shows the load exerted by the punch Fp and the force each bead
exerts on the specimen, FC1 on Side 1 and FC2 on Side 2, as functions of the punch

stroke S for a typical experiment with Rd = 6 mm, hb = 6 mm, and material Set

56
700
τ
(psi) Set 1
600
Set 2
500

400

300

200

100

0
0 0.01 0.02 0.03 0.04 0.05
δ (in)

Figure 3.8. Measured shear stress-displacement curves of the polymer


layer in lap-shear tests for the two material sets used in the draw
bending study.

α3
α2

α1

Figure 3.9. Photograph of a typical specimen after draw bending and


definition of the bend angles.

2. Both FC ’s are first increased when S = 0 to the range of 1100 to 1200 lb by


tightening the clamping screw as described in the experimental procedure. After

57
Figure 3.10. Detail of region around the draw bead that illustrates the
separation that occurs.

the punch starts moving, the FC ’s vary slightly before settling to a fairly constant
value of 1050 lb. Fp increases rapidly initially before a small drop at S ≈ 0.4 in.

takes place. After that, Fp increases gradually by a relatively small amount until
the punch is retracted. When the punch first starts to retract, Fp and FC ’s drop

suddenly. While the punch is retracting, Fp is slightly negative due to friction


between the punch and the specimen. On the other hand, the FC ’s settle to a
value around half the value during forming. This residual force is due to the

specimen trying to springback. At approximately S = 0.4 in. the punch no longer


is in contact with the specimen. At that point Fp goes to zero and the FC ’s

decrease slightly.
Experiments conducted with different Rd and hb display similar trends for all

forces measured. The value of S where the (relatively) steady-state region starts,
end of region where Fp increases rapidly, depends on Rd . These values of S are
approximately 0.3, 0.4, 0.5, and 0.6 in. for Rd = 3, 6, 9, and 12 mm, respectively.

The magnitudes of the loads in the steady-state region, however, depend on the
particular values of Rd and hb . The value for Fp in this region depends on the

tension in the specimen, so decreasing Rd or increasing hb causes Fp to increase.

58
4000
F
(lb)

3000
Fp
FC1
2000 FC2

1000

0 0.5 1 1.5 2 2.5


S (in)

Figure 3.11. Plot of the punch and bead forces for an experiment with
Rd = 6 mm and hb = 6 mm for Set 2.

For example, when Rd = 3 mm and hb = 8 mm, the value for Fp is 3800 lb at S =

2.5 in. for the same material (Set 2). The difference in forces for the two materials
used involves Fp . First, the values at beginning of the steady-state region are

higher for Set 2 than for Set 1, on the order of 10-15%, except for the Rd = 3
mm and hb = 6 mm case where both materials have similar values. Second, in the
steady-state region, Fp for Set 2 increases gradually while for Set 1 it is relatively

constant. The steady-state values of the FC ’s also vary slightly depending on the
test parameters and material, but remain in the 900 to 1100 lb range.

Immediately after bending each specimen, the three αi ’s defined in Fig. 3.9
were measured using an optical comparator. These angles are plotted in Fig. 3.12

with the horizontal axis being Rd and the vertical axis being αi . The results for
both sides are included to illustrate the symmetry of the specimen. By looking
at the values for different Rd ’s, some tends can be seen. Looking at α1 , its value

is relatively insensitive to the die radius or bead height. The values for Set 2 are

59
consistently larger than those for Set 1. This angle is primarily affected by the
radius of the punch, which does not vary in this study. Note that some experiments

were repeated to check for consistency, hence two symbols appear for some of the
cases. Looking at α2 , its value increases with increasing die radius. Again, its
values for Set 2 are larger than for Set 1. Now looking at α3 , it increases from Rd

= 3 to 6 mm. Then from 6 to 9 mm it is fairly constant but with some scatter.


This angle is difficult to measure for Rd = 12 mm since the length of the flat region

of the die between the bead region and the start of the die radius is essentially
non-existent. In summary, α1 is virtually insensitive in the range of geometric

parameters used. The values of α2 and α3 show some variation, but α2 displays a
more distinct trend as it monotonically increases with Rd .
In addition to the angles, the thickness along the length of the specimens was

measured as a first attempt to quantify the amount of separation between the


steel sheets as shown in Fig. 3.10 (Rd = 3 mm and hb = 8 mm). The thickness

variation supports separation. Results for the specimen with the largest thickness
measurement and largest spread in thickness are plotted in Fig. 3.13. (The results

for both sides are included to illustrate the symmetry of the specimen.) The
locations of the measured points are defined as A through K in Fig. 3.14 and are
plotted along the horizontal axis. The vertical axis is the ratio of the measured

thickness at each point, ti , to the nominal or original thickness, tnom , which had
a value of approximately 0.043 in. in the experiments. The results show that the

thickness varies along the specimen. Locations A and K had not undergone loading
and therefore retain the original thickness. At locations where ti /tnom > 1, such
as B, C, I, and J, separation has occurred. Note that locations B and J exhibit

the largest thickness increase, on the order of 20% of the nominal value. At

60
Side 1 Side 2
α1 90 α1 90
(Deg) 88 (Deg) 88

86 86
84 84
82 82
80 80
3 6 9 12 3 6 9 12
Rd (mm) Rd (mm)

α2 α2
(Deg) 76 (Deg) 76
74 74
72 72
70 70
68 68
61

3 6 9 12 3 6 9 12
Rd (mm) Rd (mm)

α3 100 α3 100
(Deg) (Deg)
98 98
96 96
94 94
92 92
90 90
3 6 9 12 3 6 9 12
Rd (mm) Rd (mm)
(a) hb = 6 mm hb = 6 mm
(b)
Set 1 { h b = 8 mm
Set 2 { h b = 8 mm

Figure 3.12: Plots of the three bend angles for different die radii and bead heights on each side.
places where ti /tnom < 1, stretching due to combined tension and bending caused
the steel sheets to become thinner. At the latter locations, separation could be

present; if so, it is smaller than the thinning. This cannot be determined with
the present measurements. The locations that exhibit thinning, D, E, G, and H,
correspond to the parts of the specimen that have gone over the die radius. The

region under the punch did not experience severe loading and retains the original
thickness as indicated by the measurement at point F.

ti
1.2 Rd = 3 mm
tnom h b = 8 mm

1.1

0.9

A B C D E F G H I J K
Position

Figure 3.13. Plot of the thickness at different points on the bent


specimen with the largest thickness and spread in values.

While only one example of thickness measurements was shown above, the
extremes of the thickness for all tests are plotted in Fig. 3.15. In all cases, the

thickness extrema occurred at the same locations, the maximum at position B

62
Figure 3.14. Location of the points where thickness measurements were
made.

or J in Fig. 3.14 and the minimum at location D or H. The vertical axis in the
figure represents the thickness ratio, while Rd is plotted in the horizontal axis.

One important item to notice is that both extremes tend towards the nominal
thickness as the die radius increases. Another important item is that the results

are insensitive to the values of hb considered. Also, Set 1 has consistently slightly
larger tmax /tnom and tmin /tnom than Set 2.

The next experimental step addresses the mechanical response of the polymer
layer, specifically its shear strength, after the bending operations. Evidence that
the polymer layer undergoes a severe loading history that may produce separation

between the steel sheets can be seen in Fig. 3.10, as mentioned previously. The
bending/unbending that occurs in the bead and die radius regions is the cause of

the separation. This raises the question: What is the shear response of the polymer
layer after the bending operation, especially in those regions of the specimen

that traversed the beads and/or the die radius? The question can be answered
by conducting lap-shear tests, similar to those described in Sect. 2.2 and at the
beginning of this section, on specimens cut from sheets that have been subjected to

the bending operation. These specimens will be called mini-lap specimens because

63
tmax , hb = 6 mm
tnom 1.2 Set 1 { hb = 8 mm
hb = 6 mm
tmin Set 2 {
h b = 8 mm
tnom
1.1

0.9

3 6 9 12
Rd (mm)

Figure 3.15. Plot of thickness extrema for all the draw tests.

they are smaller than those used previously and were extracted from formed sheets
as shown schematically in Fig. 3.16. The nominal dimensions of these were 2 in.

long and 0.5 in. wide.


Figure 3.17 shows the strength of the polymer after forming at location a,

indicated in the insert, for die radii of 3 and 6 mm and bead heights of 6 and
8 mm. Location a represents a portion of material that has passed through the
bead region and over the die radius, thus experiencing the full possible loading

history. Significant degradation in the shear strength of the formed material is


clearly seen when compared to the response of material that has not been formed

specified by straight 1 and straight 2 in the figure. For consistency these tests
were conducted on mini-lap specimens, as well, to show that the polymer layer

is not damaged while cutting the material to the appropriate size. These have
peak shear stresses of similar magnitude as the normal lap-shear specimens. After
forming, the bending specimens had a slight curvature along the width direction.

64
0.5 in 0.5 in

2 in

2 in 0.5 in

(a) (b) (c)

Figure 3.16. Schematic illustrating where a mini-lap specimen was


extracted from a formed bending specimen.

To check the effect of this, the depth of this curvature was measured with the aid of
a depth micrometer, and mini-lap specimens with the same curvature were made

by first rolling the material using a slip-roll forming machine and then cutting
to the appropriate configuration. As can be seen in Fig. 3.17, the responses of
the curved specimens are not identical to those from the straight ones, but they

are still significantly stronger than the formed material. Three curved specimens
were tested for consistency. It is interesting to notice that the δ where τmax occurs

for the curved specimens is between the corresponding values for those from the
formed material and the straight ones.
A more complete study of the strength of the polymer was conducted for a

specimen bent with Rd = 6 mm and hb = 6 mm. Here, additional mini-lap

65
τ 400
(psi) a Rd hb
(mm) (mm)
3 6
300 3 8
6 6
6 8
straight 1
200 straight 2
curved 1
curved 2
curved 3
100

0
0 0.01 0.02 0.03 0.04 0.05
δ (in)

Figure 3.17. Plot showing the polymer shear response after forming at
location a compared to straight and curved mini-lap specimens for Set 2.

specimens were cut at the locations shown in the insert of Fig. 3.18. Location

a is as before, while location b passed over the die radius but did not traverse
the bead region, hence it only underwent one bending/unbending cycle. Lastly,

location c was located at the plane of symmetry, under the punch. The figure
is, again, a plot of the shear response of the polymer. The response at location

a on Side 2 (also in Fig. 3.17) is higher than on Side 1, but still significantly
lower than the straight specimen ones. Obviously, some asymmetry was present
in the experiment. The polymer at location b also shows a significant decrease in

shear strength even though it underwent a “milder” loading history. As expected,


polymer behavior at location c is similar to that of the unformed material.

66
τ 400
Rd = 6 mm
(psi) a
h b = 6 mm
b
Pt. Side
300 c
a 1
a 2
b 2
200 c 2
straight 1
straight 2

100

0
0 0.01 0.02 0.03 0.04 0.05
δ (in)

Figure 3.18. Plot showing the polymer shear response after forming at
three locations along a specimen for Rd = 6 mm, hb = 6 mm, and Set 2.

3.2 Finite Element Analysis

The goal of the second part of the investigation of draw bending is to de-

velop and use a finite element model to simulate the experiments, investigate the
evolution of parameters that cannot be measured experimentally, and to conduct

parametric studies.

3.2.1 Finite Element Model

As was done in Ch. 2 for wipe bending, the problem was reduced to two-

dimensions by assuming a state of plane strain. This approximation is justified by


the large width-to-thickness ratio of the specimens tested. Due to symmetry about

the vertical, central plane of the bending fixture and therefore of the specimen,
further simplification was possible by modeling only one half of the set-up.
The model for the steel sheets consisted of plane strain, four-noded, bilinear,

67
reduced-integration, quadrilateral elements (Abaqus CPE4R [29]). Figure 3.19
shows a sketch of the model with the specimen in its initial, straight configuration.

The binder block and bead (modeled as one object), punch, and die were modeled
as rigid surfaces that contact the specimen. Contact conditions between these
objects and the specimen were frictionless in the tangential direction, and soft

contact was used for the normal direction. In soft contact, the two surfaces interact
through a nonlinear spring that activates when the surfaces are close to touching.

The pressure generated in this interaction, which tries to prevent interpenetration,


grows exponentially. Therefore, some interpenetration occurs, but it is very small

(many orders of magnitude smaller than the steel sheet thickness). The surface
with the symmetry boundary condition applied to it is noted by “Sym” in the
figure.

Each sheet had four elements through the 0.020 in. thickness. The element size
in the longitudinal direction varied along the specimen. Roman numerals mark

the edges of constant element size zones. Following is the longitudinal element size
in in each region: 0.0500 in. to the left of I, 0.0125 in. between I and II, 0.0250 in.

between II and III, and 0.0500 in. to the right of III. The optimum element sizes
and the extent of the regions were determined by parametric convergence studies
of the effect of the mesh density on the predicted values of the bend angles α1 , α2 ,

and α3 . The studies demonstrated that the values using the adopted mesh had
converged within ±0.7%.

3.2.2 Material Properties

To complete the model, the material responses of both the steel and poly-
mer components needed to be represented. The numerical work was restricted to

68
I Binder Block & Bead
Punch

Sym.

II

Die
III

Figure 3.19. Illustration of the finite element model with the location of
symmetry noted and the divisions between element size zones marked.

material Set 2. Although material models had been implemented in Ch. 2, the

nature of the loading at the material level in draw bending differs significantly
from that under wipe bending. Significant stress cycling occurs in draw bending

as material passes through the bead region and over the die radius. A preliminary
draw bending run was carried out using the material models in Ch. 2, isotropic

hardening plasticity for the steel and the interface model for the polymer. Looking
at the stress and strain histories in the steel and the relative displacement between
the two sheets, the thought that the material experiences cyclic loading was con-

firmed. This means the steel and polymer models should include the capacity to
model cyclic behavior as well as possible, and that suitable experiments should be

developed to obtain cyclic material properties.


The first set of desired properties was the cyclic hysteresis stress-strain curve

of the steel sheets. An attempt was made to obtain these properties through a
cyclic four-point bend test, but with the equipment available it was not possible

69
to obtain reliable data, especially in the strain range (on the order of 10%) needed
for this study. It must be noted that other researchers [35, 36] have been able to

develop test equipment and procedures for extracting the cyclic hysteresis stress-
strain curves of metallic sheets.
What was done in this study consisted of fitting the tensile, uniaxial, monotonic

stress-strain curve with a bilinear fit, as shown in Fig. 3.20, and use a linear
kinematic hardening plasticity model to represent the cyclic response of the steel,

including the Bauschinger effect. Note that, by the time the experiments were
conducted, the material had undergone some cyclic plastic deformation. First the

steel sheets were coiled at the sheet mill and then uncoiled, straightened and coiled
again at the laminating mill. Therefore, the stress-strain curve in Fig. 3.20 should
be at least somewhat representative of the cyclic hysteresis curve of the steel. The

values for the three parameters for the bilinear fit are the Young’s modulus E =
30 x 103 ksi, yield stress σo = 22.5 ksi, and modulus after yielding Ep = 124 ksi.

As in the case of the steel sheets, the cyclic behavior of the polymer had to
be included in the model. Therefore, an experimental program was conducted to

measure its response in lap-shear tests under cyclic loading. The objectives were
to first observe and measure the response and failure characteristics of the material
and second to gather sufficient data to allow the development of an interface model

with cyclic capability that captures the major response characteristics and can be
implemented in the finite element model.

Since the cyclic response of the polymer layer is somewhat complex, the famil-
iar monotonic part will be described first, followed by the cyclic response under
various conditions. In both cases, the experiments conducted are discussed first

followed by the corresponding fits for the model. The monotonic behavior for this

70
σ 40
(ksi)

30

20
Deep draw quality steel
E = 30.0 x 10 ksi
3

σo = 22.5 ksi
Ep= 124 ksi
10
Test
Fit

0
0 0.02 0.04 0.06 0.08 0.10
ε

Figure 3.20. Measured steel engineering stress-strain curve and bilinear


fit for Set 2.

material, Set 2, is shown in Fig. 3.21. To maintain as much simplicity as possible


in the cyclic model, the number of points in the fit was reduced to just four, as
shown in the figure. Note that upon failure, the shear stress drops to zero, where

it will remain regardless of the subsequent values of δ.


Whereas the monotonic lap shear tests were conducted by applying tension

to the specimens, the cyclic tests required tension and compression. Due to the
small thickness of the specimens, buckling became a concern. Therefore, 1/8 in.

thick aluminum tabs were glued to most of the length of the specimens to avoid
buckling. Two kinds of tests were conducted as part of the investigation of the
cyclic response of the polymer. The first consisted of loading to a predetermined

value of displacement δ and then reverse loading to failure. The objective was
to determine the effect of the initial loading on the unloading/reverse loading

response and on the strength. In the second, the specimens were cyclically loaded

71
τ 400
(psi)

300

200
Test
Fit

100

0
0 0.01 0.02 0.03 0.04 0.05
δ (in)

Figure 3.21. Measured shear stress-displacement curves of the polymer


layer in lap-shear tests and its multi-linear fit.

with a prescribed, symmetric displacement history of constant amplitude with the


shape of a triangular wave with respect to time. As in the bending experiments,

both kinds of tests were conducted at a slow, quasi-static rate, 0.03 in/min.
Results from cycling the polymer at 0.006 in. are shown in Fig. 3.22 (a). This

loading amplitude was the smallest one considered. The blue line is a segment of
the monotonic curve shown in Fig. 3.21 and is provided for comparison purposes.

The green curve is a segment of the load-reverse to failure of the polymer layer.
The red curve shows two cycles of the symmetric cyclic response plus subsequent
loading to δ = 0.01 in. It should be noted that the material did not fail where the

red curve ends. Two observations can be made from the data. First, the symmetric
cyclic response displays hysteresis, but with an amplitude of under 10 psi, which

compared with the stress range of the monotonic response, is negligible. Second,
loading to higher values of δ after cycling, the response follows the monotonic curve

72
very closely. In Fig. 3.22 (b) the load-reverse to failure behavior (absolute value of
δ and τ ) is compared to the monotonic curve. It can be seen that reverse-loading

from this small displacement value does not affect the shape of the response of the
polymer since it is identical to the monotonic one. It only affects the magnitude at
which the polymer fails. In conclusion, it can assumed that after cycling at small

values of displacement, the monotonic stress-displacement curve remains effective,


but with a modest decrease in stress at failure.

60
Monotonic Monotonic
τ Load−Reverse to Failure τ Load−Reverse to Failure, Mag.
(psi) 40 Cycle at 0.006 in. (2 cycles +) (psi) 300

20
200
0

−20
100

−40

−60 0
−0.01 0 0.01 0 0.01 0.02
δ (in) δ (in)
(a) (b)

Figure 3.22. Shear stress-displacement curves of the polymer layer in


lap-shear tests of 0.006 in. amplitude. Comparison of (a) cyclic response
and (b) load-reverse to failure response to the monotonic response.

Similar experiments were also conducted for a cycle amplitude of 0.020 in. The

results are shown in Fig. 3.23 (a). Again, the monotonic (blue), load-reverse to
failure (green), and the cyclic (red) results are plotted. For this cycle amplitude,
the behavior is significantly different from before. Prior to the first load reversal,

73
all curves coincide, as expected. First unloading occurs along a steeper curve and
the stress quickly reaches a value near zero. This is followed by a relatively long

plateau that extends until δ ≈ −0.01. At that point the stress starts rising (in
absolute value). This trend continues until failure occurs when τ reaches 250 psi
(green line). If the loading is reversed the second time at δ = −0.02 unloading

occurs along a steeper curve once again, which is followed by the stress plateau
and then a rising curve that is close to the first unloading curve. The response

under further cycling follows a curve that, in the main, traces the first and second
unloading curves as shown in Fig. 3.23 (b). Behavior after unloading was modeled

using a “simple” multi-linear fit, shown as a black, dashed line in Fig. 3.23 (a),
which will be discussed in detail below.

300 Monotonic 300 1st cycle


τ Load−Reverse to Failure τ 2nd cycle
(psi) Cycle at 0.020 in. (1 cycle) (psi)
200 Fit 200

100 100

0 0

−100 −100

−200 −200

−300 −300
−0.02 −0.01 0 0.01 0.02 −0.02 −0.01 0 0.01 0.02
δ (in) δ (in)
(a) (b)

Figure 3.23. Shear stress-displacement curves for lap-shear tests of


0.020 in. amplitude showing (a) comparison of cyclic and monotonic
responses plus a cyclic multi-linear fit and (b) the response for two
loading cycles.

74
A third set of tests was conducted for a cycling amplitude of 0.015 in. The
results are shown in Fig. 3.24, which shows the monotonic (blue), load-reverse to

failure (green), and the fit (black, dashed). The load-reverse to failure response is
qualitatively similar to the one shown in Fig. 3.23 (a). Hence a similar fit captured
this behavior as well.

300 Monotonic
τ Load−Reverse to Failure
(psi) Fit
200

100

−100

−200

−300
−0.02 −0.01 0 0.01 0.02
δ (in)

Figure 3.24. Shear stress-displacement curves of the polymer layer in


lap-shear tests of 0.015 in. amplitude, including the load-reverse to
failure and monotonic responses plus a cyclic, multi-linear fit.

Based on the observations from the cyclic tests presented above, an interface

model for shear behavior of the polymer will be developed next. As in Sect. 2.3.4,
it will consist of a nonlinear spring in parallel with a linear dashpot, as shown
in Fig. 3.25. The behavior of the nonlinear spring will be based on the lap-shear

75
test data while the dashpot will be used once more to improve convergence during
equilibrium iterations in the draw bending model. The total shear stress in the

model, τ , is given by
τ = τs + cδ̇. (3.1)

where τs is the stress due to the spring, δ is the relative displacement between

the steel sheets, δ̇ is the rate of change of δ with respect to “time,” and c is the
damping constant. Recall that here “time” is just a parameter proportional to
the prescribed loading and does not represent “real time.” In all runs conducted,

the time length to complete the analysis was kept the same for consistency. The
damping constant c was chosen to be the smallest value that yielded converged

solutions. The interface model was implemented in Abaqus as a user subroutine


(UNITER [29]) as explained in Appendices B and C.

Figure 3.25. Interface model for shear behavior of the polymer layer.

Figure 3.26 shows how the monotonic and cyclic behaviors are combined to
define the response of the nonlinear spring. Starting at τs = 0 and δ = 0, the

monotonic behavior is used upon loading. Point 1 is where the slope in the

76
monotonic behavior increases. If unloading occurs between Pts. 0 and 1, the
monotonic behavior is still used. Recall that small amplitude cycles, such as

those shown in Fig. 3.22, have no effect on the shape of the initial, monotonic
part of the response.
After passing Pt. 1 though, continued loading can result in failure as shown in

Fig. 3.21 at a value τs = 429 psi, unless unloading occurs prior to failure. Unload-
ing will cause the model to use the cyclic behavior. In Fig. 3.26, unloading occurs

at Pt. 2, where δ = δ2 . The slope of line 2-3 was found by fitting the unloading
part of the 0.020 in. cycle amplitude results in Fig. 3.23. If unloading while still in

2-3 stops, and the material reloads past Pt. 2, the behavior returns to monotonic.
But if unloading continues beyond Pt. 3, τs = 0 until Pt. 4. The location of Pt. 4,
where δ = δ4 , depends on the value of δ2 . As can be seen by comparing the re-

sponses in Figs. 3.23 and 3.24, as δ2 increases, so does δ4 . Therefore, the value of
δ4 for an arbitrary δ2 was found using a linear interpolation/extrapolation of the

values for δ4 for initial unloading at 0.015 and 0.020 in. If the loading direction
changes in 3-4 (going towards 3 now), it is possible for the model to return to 2-3

(and therefore also return to the monotonic fit).


If, on the other hand, loading in 3-4 (going towards 4) passes Pt. 4, 4-9 is
entered. The slope in this region was found from the experiments with cycle size

of 0.020 in. Point 9 is where the magnitude of τs is 250 psi and corresponds to
the location of failure during cycling. This value of τs was the smaller of the two

failure shear stresses from the load-reverse to failure experiments shown in Figs.
3.23 and 3.24. The slope of line 9-10 is the same as the slope from the monotonic
fit between the stress peak and zero shear stress, see Fig. 3.21. If the loading

direction changes while in 9-10, a line with the same slope as in 2-3 is used until

77
τs = 0 (this line is not shown in the figure). Once τs reaches a value of zero after
failure, its value will remain zero from then on.

τs

12

0 3, 11
δ
6
10 8 4 13

9 7

Figure 3.26. Sketch showing key points in the cyclic polymer behavior
used in modeling.

If the direction of loading changes while in 4-9, a second unloading occurs.


This happens at Pt. 5 in Fig. 3.26, and 5-6 is entered. The response of the model

after the second unloading is shown in dashed lines. The slope in this region is

78
the same as in 2-3 after the initial unloading. If the loading switches direction
and passes Pt. 5, the model returns to 4-9. If unloading continues past Pt. 6,

the model enters 6-11. The model can leave this region by the loading direction
changing and passing Pt. 6, thus entering 6-7. Two exits of 6-7 exist: by failing
(at Pt. 7) or by returning to 6-11. The stress at Pt. 7, the failure stress, has

magnitude of 250 psi. The behavior in and after 7-8 is the same as in and after
9-10. The other way out of 6-11 causes 11-12 to be entered. Point 12 is at the

failure stress like Pts. 7 and 9. The behavior in and after 12-13 is the same as
in and after 9-10. Once the response enters 6-11, all further cycling occurs along

the lines defined by 7-6-11-12, unless failure occurs at 7 or 12. Figures 3.23 and
3.24 show, in dashed line, the fit of the model as it applies to these two cycling
amplitudes.

The model behaves the same if the relative displacement at Pt. 1 is in the
“negative” direction. The only difference is that the sign of δ is switched, and

therefore the sign of τs is also switched, from what is shown in Fig. 3.26.
The interface model must also address the interaction between the sheets in

the normal direction. In order to do this, a spring-damper system like the one in
Fig. 3.25 was used. The behavior of the polymer in the normal direction is given
by

σn = kn δn + cn δ̇n . (3.2)

Here, σn and δn are the stress and relative displacement of the sheets in the normal
direction. The constant kn has different values depending on whether the steel

sheets are separated or overlapping. When overlapping kn was chosen to have a


large value of 109 psi/in. On the other hand, when separating, the value of kn

was selected to be 29.1 x 103 psi/in. as found experimentally in a pull test. The

79
damping constant cn was chosen to be the smallest value that yielded converged
solutions. Here, δ̇n is the rate of change of δn with respect to “time.” It is possible

for the polymer to fail by separating too much. This happens if separation is
occurring and kn δn is greater than or equal to the separation failure stress, 291
psi.

When failure in the polymer occurs either by separation or by shear, both σn


and τs are set to zero, that is failure in one direction causes failure in the other.

Note that the failure stress in each mode is independent of the stress in the other.
In other words the failure envelope in shear/normal stress space is as shown in

Fig. 3.27.

σn
σno

τso τso
τs

Figure 3.27. The failure envelope in shear/normal stress space.

The model is a relatively simple first attempt to represent the cyclic behavior

of the polymer. It has neglected some aspects of the response that were observed

80
in the experiments. For example, it uses only straight lines to represent nonlinear
curves and it sets τs to zero in 3-4 and 6-11, when in the actual material it has

a small non-zero value. Looking at Fig. 3.23 (a), however, it can be seen that
the differences between the straight line fits and the actual response are relatively
small. The slopes of 6-7 and 11-12 are constant, but the material loses some

stiffness with cycling as shown in Fig. 3.23 (b). In spite of this, the draw bending
model only causes the polymer layer to cycle at the most two times, so the level

of sophistication of the model is appropriate to represent its pertinent behavior.


In addition, the experimental data were obtained in a limited number of tests,

so some aspects of the response may still need to be determined, in particular


whether the shear stress affects the stiffness and strength in the normal direction
and vice versa. As a result, the model could still be improved, but its performance

in its current form was evaluated by comparing the results it produced in the draw
bending simulations to the experimental observations.

3.2.3 Simulation of Experiments

After developing the model, simulation of the experiments was conducted to


assess its performance. Several aspects related to the draw bending operation will

be addressed in what follows. Among these are the final shape of the specimen,
the stress and deformation histories at different points during forming, and the

state of the polymer layer after the operation is completed. Recall that numerical
simulations were only conducted for material Set 2.
In order to generate the numerical results, the steps followed in the experiment

were replicated. First the draw bead was brought in contact with the sheet and
pushed into the draw bead channel in the die by the appropriate amount. This

81
was followed by the punch stroke and retraction, and finally moving the tools
(bead/binder block and die) away from the specimen in order to release it. In

order to complete a run, however, it was necessary to adjust the two damping
parameters c and cn in the interface model. It could take as few as one or as many
as 38 attempts to achieve convergence. What made the process more problematic

was that the values of c and cn that yielded convergence depended on the geometric
(Rd and hb ) and material (σo ) parameters of the problem. The total time to

produce a converged run could vary between one hour and a few days.
The first aspect to be addressed is the final shape of the bending specimen

after forming. Figure 3.28 shows results of the analysis for hb = 6 mm, compared
to experimental measurements of α1 , α2 , and α3 for different values of Rd . Two
experimental values are shown, one from each side of the specimen. No numerical

results for Rd = 3 mm are shown, because the model did not converge in this
case. In Fig. 3.28 (a), the close agreement between experimental and analytical

results for α1 can be seen. Both set of results indicate that α1 does not depend
on Rd . The values for α2 , shown in Fig. 3.28 (b), are close to the same range, but

the results from analysis are higher than those from experiment by 5.5% or less.
Another difference is that as Rd increases, the value of α2 decreases slightly for
the analytical results but increases for the experimental measurements. Figure

3.28 (c) shows that the same trend for the variation of α3 as in the experiments
is predicted by the model, and that the values are relatively close. Overall, the

experimental and analytical values for α1 are in good agreement, for α2 have some
differences present, and for α3 are again in fairly good agreement.
The numerical results for α1 , α2 , and α3 were also compared to experimental

results for the other bead height (hb = 8 mm) used in the experiments. The results

82
α1 90
Experiment
(Deg) h b = 6 mm
88 Analysis

86
84
82
80
3 6 9 12
Rd (mm)
(a)
α2 80
(Deg)
78
76
74
72
70
3 6 9 12
(b) Rd (mm)

α 3 100
(Deg)
98
96
94
92
90
3 6 9 12
(c) Rd (mm)

Figure 3.28. Comparison of the analytical and experimental results of


α1 , α2 , and α3 for hb = 6 mm for Set 2.

are shown in Fig. 3.29. Here, numerical results were obtained for all four values

of Rd . The values of α1 do not depend on Rd yet again, as shown in Fig. 3.29 (a),
and the two sets of results are in close agreement. Figure 3.29 (b) shows that the

analytical values of α2 vary slightly but are relatively constant. The analytical

83
value at Rd = 3 mm is 5% high, but the agreement is better for other values of Rd .
From Fig. 3.29 (c) it can be seen that the analytical values of α3 increase slightly

over the range of Rd looked at here. The analytical values at each Rd are higher
than the experimental values. Overall, the experimental and analytical values for
α1 are in good agreement; for α2 are in fairly good agreement; and for α3 have

some differences, which are in the range of 4% or less.


A second set of parameters from the results of the analysis that can be com-

pared to the experimental results are the punch and bead forces. One can expect
large differences between the experiment and numerical results for the punch force

because the model neglected the effect of friction. Attempts to include friction
in the model failed to produce converged results. Figure 3.30 illustrates the ef-
fect that friction can have on the punch force. It compares two experimental

results: (1) the specimen was lubricated with heavy oil (standard case) and (2)
the specimen was bent without lubrication. The punch force in the second case is

approximately twice of the first at S = 2.5 in. As expected, the numerical results
yield punch forces that are well below the experimental measurements. For the

case with Rd = 6 mm and hb = 6 mm, the predicted punch force is one-third of


the measured one.
The force exerted by the beads on the specimen was also calculated. The

numerical results underpredicted this by a factor of one half. Comparison of


numerical results on solid sheet with and without friction indicate that this force

should not be significantly affected by friction. The most likely explanation is that
the value may be affected by contact between the specimen and the binder block
away from the beads, which due to tolerances in the construction of the set up,

may not be accurately represented by the numerical model. Between the top of the

84
α1 90
Experiment
(Deg) h b = 8 mm
88 Analysis

86
84
82
80
3 6 9 12
Rd (mm)
(a)
α2 80
(Deg)
78
76
74
72
70
3 6 9 12
(b) Rd (mm)

α 3 100
(Deg)
98
96
94
92
90
3 6 9 12
(c) Rd (mm)

Figure 3.29. Comparison of the analytical and experimental results of


α1 , α2 , and α3 for hb = 8 mm for Set 2.

die and the bottom of the binder block there was space for the specimen thickness

plus clearance. The nominal value of this clearance, also the standard value in
the numerical analysis, was 0.006 in. Varying its value was studied numerically.

Decreasing this value increases the bead force because additional contact occurs

85
between the specimen and the binder block. Since the laminate in the experiments
was slightly thicker than in the numerical analysis, 0.043 vs. 0.040 in., the clearance

was smaller in the experiments, which may have contributed to this discrepancy.
The punch force is not significantly affected by changing the clearance. It should
be noted that the numerical models with smaller clearance values had convergence

problems while the punch was moving down, so other results from these models
cannot be studied.

6000
Fp
(lb)
5000
No Lubrication
4000 Heavy Oil (std.)
Rd = 6 mm
3000 h b = 6 mm

2000

1000

0 0.5 1 1.5 2 2.5


S (in)

Figure 3.30. Experimental results that demonstrate the affect of


lubrication on the punch force for Rd = 6 mm and hb = 6 mm for Set 1.

Another aspect of the problem where it is interesting to compare the numerical

results to the experimental ones is the thickness at various locations along the
bending specimen after forming. A comparison is shown in Fig. 3.31 for Rd = 3

86
mm, hb = 8 mm, and Set 2. Recall from Fig. 3.14 that location F was at the plane
of symmetry, so the two experimental values included are the measurements on

each side of the specimen. The thickness ratio for the analysis at locations A and
F have a value of 1.0, just like in the experiment. At the rest of the locations, the
predicted thickness ratios are closer to 1.0 than in the experiment, but they follow a

similar trend. In particular, the tendency of the steel sheet to separate at location
B was predicted reasonably well. The thinning of the specimen at locations D

and E was underpredicted. Looking more in detail at the numerical results, it was
found that the thinning of the steel sheets themselves was underpredicted. This

is likely a result of the lower tension in the numerical model due to the friction
being neglected.

ti
1.2 Experiment
tnom Analysis
Rd = 3 mm
h b = 8 mm

1.1

0.9

A B C D E F
Position

Figure 3.31. Comparison of thickness at different points along specimen


for the standard model and experimental results for Rd = 3 mm, hb = 8
mm, and Set 2.

87
Having a numerical simulation allowed the study of items that are of interest
but could not be addressed experimentally. Here several such items are addressed

for the case with Rd = 6 mm and hb = 6 mm. The first such item is the relative
displacement between the steel sheets, δ, as well as the total shear stress in the
polymer, τ , during the forming process. The loading history experienced by each

point in the specimen depends on its starting location, whether under the punch,
between the punch and the die, in front or behind the bead, etc. Figure 3.32

shows δ and τ histories for a point in the polymer that experienced the complete
loading history possible. That means this point started behind the bead, went

through the entire bead region, and passed around the die radius during forming.
The horizontal axis in the figure represents progress through the bending process.
The Roman numerals represent the end of a loading step and the beginning of the

next. The points labeled with letters indicate key situations during the loading
history. Between i and ii the bead moved into the draw bead channel. The point

being considered had just entered the bead region at ii, so the material here was
bent, causing the nonzero δ and τ values. Between ii and iii the punch moved down

causing the point to traverse the bead region and go over the die radius. At A the
point was about a quarter of the way through the bead region, where the material
was straightened. Here δ and τ achieve extrema. At B the point reached the tip

of the bead (half way through the bead region), where δ = 0 and τ ≈ 0. (Recall
that the damping term will contribute to τ as long as a relative velocity between

the steel sheets is present.) At C the material straightened a second time, about
three-quarters through the bead region. Here, the relative displacement between
the sheets was now in the opposite direction, and, correspondingly, the shear stress

had changed sign. C is also an instance where δ and τ achieve extrema. The point

88
exited the bead region at D, which coincides with a change in the direction in
which δ was moving. By E the point just started to pass over the die radius. It

marks the start of a significant change in δ into negative values thus completing
one full cycle in the polymer stress-deflection response. Shortly after, the polymer
failed due to excessive shear stress at a location close to the end of the die radius.

This situation is labeled as f on the horizontal axis. After failure the laminate
straightened, and the value of δ did not change significantly for the rest of the

process. Therefore, τ remained close to zero. At iii the punch had traveled its
full stroke. From iii to iv the punch retracted. Between iv and v the bead/binder

block moved away from the specimen, and from v to vi the die moved away from
the specimen. In summary, the polymer at the point described underwent a one
and one-quarter cycle in its stress-deflection response prior to failing near the end

of the die radius.


From the history of δ at this one point, the histories at most other points can

be seen. For any point that passed A, the value of δ here was very close to the
value shown in Fig. 3.32. The same is true for the other key situations indicated

in the figure. For points that started further behind the bead region, the histories
of δ and τ are very similar to that in Fig. 3.32, but they are shifted to the right
and are truncated upon reaching iii. For points that started in front or in the

bead region, their δ histories from ii on start at positions midway through what is
shown in the figure, depending on their starting locations. For instance, for points

that started between the bead region and the die radius, ii corresponds to a point
in the curve between D and E. In addition, these points did not undergo as much
cycling and therefore have a somewhat different τ history.

The history of the polymer at the same point considered in detail above is

89
δ 0.02 C
(in) E
0.01
B
0
−0.01 D
−0.02 A
i ii f iii iv v vi
(a)
τ 200
C
(psi) 100 B E
0
−100 D
−200
−300 A
i ii f iii iv v vi
(b)

Figure 3.32. History of (a) δ and (b) τ in the polymer for a point that
experienced the complete loading history possible.

shown in a τ vs. δ plot in Fig. 3.33 (a). This figure illustrates the amount of

cycling this point experienced more clearly. Recall that the values for the damping
constant c were chosen to be as small as possible and still yield convergence. The

individual components τs and τd are shown in Fig. 3.33 (b) plotted vs. δ. The
case shown, Rd = 6 mm and hb = 6 mm, required one of the largest values for c.
Therefore, other cases have similar or smaller values of τd . In general, the value of

τd is less than 20% of the peak τs value and is most significant when τs = 0. Note
that when both τs and τd are non-zero, the value of τs is dominant. Therefore,

although including damping was necessary to obtain converged solutions, it should


not have affected the results obtained significantly.

A second item, also related to the polymer layer, that could not be addressed
by the experiments but could be by the finite element analysis is the loading

90
τ 200 τ 200
(psi) Rd = 6 mm (psi) τs
h b = 6 mm τd
100 100

0 0

−100 −100

−200 −200

−300 −300
−0.02 −0.01 0 0.01 0.02 −0.02 −0.01 0 0.01 0.02
δ (in) δ (in)
(a) (b)

Figure 3.33. History of (a) τ vs. δ and (b) τs and τd vs. δ in the polymer
for a point that experienced the complete loading history possible.

history in the normal direction. Figure 3.34 shows the histories of δn and σn

for same point considered above. In this figure, positive δn indicates separation
between the two sheets. Recall that the behavior in the normal direction was

prescribed to be very stiff when the surfaces overlap but relatively compliant
when separating. This is why δn is larger in the positive direction than in the

negative direction, but σn is much larger in the negative direction. Again, the
horizontal axis represents the steps in the process. The same key situations are
indicated in Fig. 3.34 that were shown in Fig. 3.32. Several interesting items can

be seen in the figure: first, some separation between the steel sheets occurs as the
point of interest traversed the bead region; second, the large negative spikes in

σn occur as the sheet contacted different parts of the tooling; third, once failure
occurs σn goes to zero and a relatively large separation develops all through the

end of the punch stroke; fourth, the separation reduces considerably upon release
of the specimen.
Another item that could only be studied numerically is the stress-strain history

91
δn 8
(x10 −4 in) 6
A C E
4
B D
2
0
i ii f iii iv v vi
(a)
σn
0
(psi)
−1000

−2000

−3000
i ii f iii iv v vi
(b)

Figure 3.34. History of (a) δn and (b) σn in the polymer for a point that
experienced the complete loading history possible.

in the steel sheets. Figure 3.35 shows the normal stress/plastic strain curve in the

longitudinal direction of the specimen at a point in the steel located by the point
in the polymer that was discussed above. This point was on the top of the top

sheet. It should be noted that the histories at the four points at the top and the
bottom of the each sheet that start aligned vertically have similar characteristics

but with variations in magnitudes. The subscript “11” indicates that these are the
components in the direction aligned with the length of the specimen. Note that
the coordinates rotate with the material. Also, the σ11 -²p11 loops are biased towards

positive strain due to the tension that develops in the specimen. The histories of
²p11 and σ11 are also shown individually in Fig. 3.36 (a) and (b) as functions of

the progress through the bending process. The key situations A through E defined
previously are also shown for reference. Looking at both figures, it is clear that the

92
material underwent three stress cycles. Hence, it is imperative that a kinematic
hardening plasticity model be used to model the constitutive behavior of the steel

in this problem.

40
σ11
(ksi)

20

−20

−40
−0.04 −0.02 0 0.02 0.04 0.06
ε 11
p

Figure 3.35. History of σ11 vs ²p11 in steel for a point that experienced
the complete loading history possible.

One last important aspect to be investigated from the finite element results

relates to the predicted failure of the polymer layer as function of the die radius
and the bead height. Recall that the final state of the polymer was addressed

experimentally using the mini-lap specimens cut from formed bending specimens
in Figs 3.17 and 3.18 and related discussion. For the Rd = 6 mm and hb = 6 mm
case, the predicted final state of the polymer with respect to failure is shown in

Fig. 3.37. The two regions where the polymer reached failure are indicated. The
insert shows the original locations of the ends of the broken regions. The first

region (the longer one) went from the end of the die radius at Pt. 1 to a point

93
ε 11
p
0.06 C D
0.04
0.02
B E
0
−0.02 A
−0.04
i ii iii iv v vi
(a)
σ11 40
(ksi) C
20
E
0

−20
A
B D
−40
i ii iii iv v vi
(b)

Figure 3.36. History of (a) ²p11 and (b) σ11 in steel for a point that
experienced the complete loading history possible.

that started close to the end of the bead region, Pt. 2. The polymer in this region

experienced at least some cycling while passing through the bead region, and then
failed close to the end of the die radius. The polymer in the second broken region,
between Pts. 3 and 4, did not experience cycling. It was loaded beyond failure

by the bending and straightening that occurred at the start of the process in this
region combined with its proximity to the punch radius. On the other hand, the

polymer between Pts. 2 and 3 did not break because it neither experienced cycling
nor was it loaded beyond failure. (Recall that, in the model, the failure stress was
largest during first loading but decreased after the first reversal.) The state of

the polymer shown in Fig. 3.37 agrees with the results from the mini-lap tests
(refer back to Fig. 3.18). The polymer at locations a and b in these tests showed

degradation in shear strength, while the numerical results indicate that all or a

94
significant portion of the polymer at these locations has failed. Location a for the
mini-laps was just below Pt. 1 here, while b had Pt. 3 close to the middle of it.

1 2 3 4 Broken region

2
3
Broken region
4

Figure 3.37. Final state of polymer for Rd = 6 mm and hb = 6 mm.

As a way to estimate the extent of failure in the polymer layer, a quantity


Lf was defined by the sum of the lengths of the broken regions, measured in the

undeformed configuration. For the case with Rd = 6 mm and hb = 6 mm Lf has


a value of 1.69 in. The values of Lf for all cases considered thus far are shown in

Table 3.1. This table also indicates the start of the first broken region. For four
of the cases the broken region started at the end of the die radius (Pt. 1 in Fig.

3.37), but for three of the hb = 8 mm cases this region started midway through
the bead region, just to the right of the apex of the bead. Consequently, Lf is
considerably larger in these cases.

95
TABLE 3.1

FAILURE LENGTH FOR THE hb = 6 AND 8 mm CASES

Rd hb Lf Start of broken region

(mm) (mm) (in)

3 6 — ——

6 6 1.69 end of die radius

9 6 1.68 end of die radius

12 6 1.60 end of die radius

3 8 2.39 midway through bead region

6 8 2.46 midway through bead region

9 8 2.59 midway through bead region

12 8 1.60 end of die radius

3.2.4 Parametric Study

The finite element model developed yielded results that were, in several ways,
reasonably representative of what was observed in the experiments. In this section,

the same model was used to preform parametric studies on how the yield stress of
the steel and the bead height affect the bending process, primarily the final shape

of the specimen.
The effect of the yield stress on α1 , α2 , and α3 for Rd = 6 mm and hb = 6 mm

is shown in Fig. 3.38. The horizontal axes contain the ratio of the current yield

96
stress σ̄o to the yield stress of the steel in material Set 2, σo = 22.5 ksi, while α1 ,
α2 , and α3 are plotted in the vertical axes. Figure 3.38 (a) shows that α1 decreases

less than 2◦ over the range of σ̄o /σo shown. In Fig. 3.38 (b) the dependence of α2
on the yield stress is evident. By going from σ̄o /σo = 0.90 to 2.70, the value of
α2 decreases by more than 20◦ . Figure 3.38 (c) shows that α3 varies slightly but

does not depend on the yield stress significantly.


The effect of varying the yield stress on the final shape of the specimen is shown

in Fig. 3.39 for the two extreme values of σ̄o /σo . The wall of the U-channel (portion
between the die radius and the punch radius) is the main difference between the

two cases. For σ̄o /σo = 0.90 this wall is much straighter than for σ̄o /σo = 2.70.
Other items addressed in the steel sheet yield stress study were the punch and
bead forces. The value of Fp for σ̄o /σo = 2.70 is more than twice its value for

σ̄o /σo = 0.90, see Fig. 3.40 (a). Figure 3.40 (b) shows FC for the two extreme
yield stresses that were studied. FC nearly triples as σ̄o /σo went from 0.90 to

2.70. Recall that the model underestimates Fp and FC , so actual values would be
higher. The characteristics of the broken region of the polymer layer in all cases

are similar to those obtained for all cases with hb = 6 mm. For σ̄o /σo > 1.50, the
two regions shown in Fig. 3.37 coalesced.
Overall, increasing the yield stress decreases α1 slightly, decreases α2 signifi-

cantly, and has minimal effect on α3 . It causes the part of the specimen between
the die radius and punch radius to be more curved. Also, increasing the yield

stress increases the values of Fp and FC .


A parametric study of the effect of the bead height was conducted next. Figure
3.41 shows how varying hb affected α1 , α2 , and α3 for constant Rd = 6 mm and σo

= 22.5 ksi. Values of hb = 0, 3, 6, and 8 mm were considered. The case with hb =

97
α1 90
(Deg) σo = 22.5 ksi
88 Rd = 6 mm
86 h b = 6 mm

84

82

80
1 1.5 2 2.5
(a) σo / σo
α2 80
(Deg) 75
70
65
60
55
50
1 1.5 2 2.5
(b) σo / σo
α 3 100
(Deg)
98

96

94

92

90
1 1.5 2 2.5
(c) σo / σo

Figure 3.38. Parametric study results of varying steel sheet yield stress
on (a) α1 , (b) α2 , and (c) α3 .

0 yielded a significantly different shape of the specimen, hence the values of the

three angles are smaller than in the other three cases. The differences between
the values for hb = 3, 6, and 8 mm are not very significant. The broken region of
the polymer layer in the case of hb = 3 mm was very similar to that shown in Fig.

3.37. For hb = 0, the broken region is limited to a small region next to the punch
radius.

98
(a) (b)

Figure 3.39. Final shape for (a) σ̄o /σo = 0.90 and (b) σ̄o /σo = 2.70.

Fp FC 1200
(lb) (lb)
3000 σo / σ o = 0.90 σo = 22.5 ksi
σo / σ o = 2.70 Rd = 6 mm
1000
h b = 6 mm

2000 800

600
1000
400

0 200

0
0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2 2.5
S (in) S (in)
(a) (b)

Figure 3.40. Parametric study results of varying steel sheet yield stress
on (a) punch and (b) bead forces.

99
α 100
(Deg)

90

80

70 α1
σo = 22.5 ksi
α2
α3
Rd = 6 mm

60
0 3 6 8
h b (mm)

Figure 3.41. Parametric study on the effect of varying hb on the angles


for a single die radius Rd = 6 mm.

100
CHAPTER 4

SUMMARY AND CONCLUSIONS

The objective of this dissertation was to study the forming characteristics of

laminated steel under two bending operations: wipe bending and draw bending.
In both cases, the items of interest were the loads and deformations applied dur-

ing the bending process and the state of the specimen after bending. The work
discussed here consisted of experimental and numerical investigations for both
bending operations. The type of laminated steel considered here has been devel-

oped as a means of sound and vibration attenuation in sheet metal products. It


consists of two layers of steel separated by a thin layer of a viscoelastic polymer.

The application of this research is of interest in industry since the material is


relatively new (it has been in development for around 30 years) and some of its

forming characteristics remain to be determined. Furthermore, successful use of


laminated steel promises to reduce weight, cost, and manufacturing steps in the
production of objects where it can substitute more traditional materials.

The first part of this work addressed the bending characteristics of laminated
steel in wipe bending. A dedicated experimental set-up was developed for this

study. The specimens in the experiments consisted of relatively wide strips that
underwent a 90◦ bend. The results show that the specimens curl during bending,

and that the curl remains after the bending operation ends, as shown in Fig.
1.2 (b). As a consequence of the curl, the specimens also appear to have larger

101
springback angle (or smaller bending angle) than specimens of similar dimensions
made of solid steel. These factors can result in manufacturing complications when

laminated steel is used. Increasing the stroke length of the punch somewhat
decreases the curl, but not enough to eliminate the problem. The polymer layer
between the two sheets of steel is viscoelastic and, as a result, the curl decreases

and the bend angle increases somewhat with time after bending.
The numerical predictions for wipe bending were based on a plane strain finite

element model of the problem that included contact between the specimen and the
tools. The shear response of the polymer layer was modeled by a series of properly

calibrated nonlinear spring elements. No separation between the steel sheets was
allowed. These sheets were treated as isotropically hardening elastic-plastic solids.
The material properties for the polymer layer and the steel sheets were obtained

from ancillary material tests. The numerical predictions of the bend angle and
curl were in reasonable agreement with the experimental measurements. In some

instances the results depend somewhat on the boundary conditions at the end of
the specimen under the blank holder. A parametric study of the dependence of the

curl and bend angle on variations of the yield stress of the steel, the thickness of the
steel sheets, and the stiffness of the polymer layer show that the curl decreases and
the bend angle increases as the yield stress decreases and the thickness increases.

The dependence on the polymer layer stiffness is more complicated. The curl is
zero when the stiffness is zero since the sheets bend independently, but increases

rapidly with stiffness until it reaches a maximum. For even higher stiffnesses the
curl decreases at a relatively slow rate. The bend angle initially decreases rapidly
with stiffness but does not change much after the curl achieves its maximum.

One particular aspect of the modeling that affected both bending operations

102
concerns how to represent the thin polymer layer that separated the steel sheets.
Originally, it had been modeled by a set of discrete springs, which adds significant

burden to building the model and was unable to include damage of the polymer.
More recent efforts were directed to the development of an interface model that
was implemented in the finite element analysis. It provided not only a more

convenient way to model the polymer, but also a way to incorporate a more
realistic polymer behavior, including effects such as damage and the effect of

cycling. The first interface model used a spring-dashpot system for the shear
behavior of the polymer. The nonlinear spring had the same behavior as the spring

model with the addition of damage by setting the stress in the spring to zero after
the peak in Fig. 2.11 is passed. The damping was necessary in order to obtain
numerical convergence. This interface model was used for two wiping bending

studies. One was on the effect of increasing the thickness to values beyond those
considered earlier. The shear deformation of the polymer layer becomes larger

as the thickness increases and causes failure near the bend region. The length
of the region where the polymer fails increases rapidly with thickness until the

whole length of the specimen protruding from the die has failed for thicknesses
larger than or equal to 1.875 times the standard thickness of the typical specimens
tested. This results in a sudden increase in the value of α and a sudden decrease

in the value of C. These predictions were confirmed by experiments conducted on


specimens with twice the standard thickness. The other study was on the effect

that the length of the specimen had on the curl. It turns out that as the length
increases, so does the value of C.
The main problem that arises when forming laminated steel using wipe bending

is curl. It is inherent to this process and must be addressed. Suggestions have

103
been made to raise the temperature of the workpiece and tooling in order to reduce
the curl. This should work because it decreases the stiffness of the polymer layer,

and, therefore, the curl would also be reduced as indicated by the parametric
study. Potential problems with this idea concern whether it would be practical
and economical to implement. Another possible solution is to develop polymers

that can be broken during bending but then heal naturally, by heat treatment,
or by a different technique. Again, whether this is practical or even feasible is

questionable and beyond the scope of this investigation.


The second problem addressed in this work was draw bending. The mechanics

of draw bending are considerably more complicated than those of wipe bending
because bending occurs in the presence of tension and material points undergo a
much more complex loading history as they are drawn through the beads and the

die radii (see Fig. 3.3). Instead of a single bending event, draw bending causes
the sheet to experience several bending/unbending/reverse-bending events. A

dedicated experimental set-up was developed in order to study draw bending as


well. The specimens were relatively wide strips that were formed into a symmet-

ric U-channel. This experimental work was conducted on two sets of material
with slightly different steel stress-strain responses and significantly different shear
stress-displacement responses of the polymer. Experimental observations indicate

that the geometry of the tooling affects the final geometry of the specimens and
the integrity of the polymer layer, as some separation between the steel sheets

was observed. The final geometry of the specimen was quantified using three
bend angles. The values of the angles depend on the radius of the die and, by a
lesser amount, the height of the bead. The thickness at different points along the

length of the specimen was also addressed. The thickness in undeformed portions

104
of the specimen remain equal to the original, or nominal, thickness. Some por-
tions were thicker than the nominal value, indicating that separation had occurred

there. Other portions were thinner than the nominal, and this means thinning
occurred in the steel sheets due to the tension that developed during bending.
The extreme values of the thickness tend towards the nominal thickness as the die

radius increases and are insensitive to the bead height. The last item addressed
experimentally was the state of the polymer layer after forming. By conducting

lap-shear tests similar to those used to characterize the polymer response, it was
found that the polymer shear response located in a segment between the die ra-

dius and the punch radius at the end of the bending process degraded significantly.
On the other hand, polymer from an undeformed (mildly deformed at the most)
region had a shear response very similar to that from the intact material.

The numerical predictions for draw bending were based on a plane strain fi-
nite element model, including contact between the specimen and the tools. The

shear response of the polymer layer was modeled using a second, more detailed
interface model that included cyclic effects and failure. In addition to addressing

the shear response, this interface model also defined the normal behavior: very
stiff when the surfaces overlap to keep interpenetration minimal, relatively com-
pliant when separating, and polymer failure if the amount of separation exceeded

a critical value. If failure occurred in shear or in the normal direction, the poly-
mer completely failed. The steel sheets were treated as kinematically hardening

elastic-plastic solids. The material properties for the polymer layer and the steel
sheets were obtained from ancillary material tests. The numerical predictions of
the three bend angles were in reasonable agreement with the experimental mea-

surements. The predictions resulted in reasonable agreement on which locations

105
where thicker and which where thinner than the nominal thickness, but the pre-
dicted thicknesses were in general closer to the nominal value. Where along the

length of the specimen the polymer had failed in the numerical analysis was noted
to check the final state of the polymer layer. This agreed well with the experi-
mental results. Parametric studies on how the yield stress of the steel and the

bead height affect the bending process, primarily the final shape of the specimen,
were conducted. Increasing the yield stress affects one of the angles significantly

and has much smaller effects on the other two angles. It causes the part of the
specimen between the die radius and punch radius to be more curved. Varying

the bead height has very little effect on the three bend angles or the final state of
the polymer, except when there is no bead.
Polymer degradation is a big problem when using draw bending to form lami-

nated steel, so it would be ideal if the degradation could be minimized. A possible


technique to do this is use friction instead of beads to restrain the motion of the

material and therefore reduce cycling of the polymer layer. This could raise other
problems, for instance, increase loads on the tools. Another possible solution

would be to develop polymer layers with higher strength without affecting the
damping characteristics of the laminate.
This work sheds some light on the forming characteristics of laminated steel

under two bending operations. The results can be used to optimize the tooling and
process parameters used, which under the right economic and technical conditions,

can lead to an expanded use of laminated steel. Achieving the full potential of
laminated steel as a sound absorbing material that can be easily integrated into
sheet metal structural parts can lead to lightweighting of structures and reduce

the number of parts and manufacturing steps required to build a product, all with

106
a material that can be easily recycled.

107
APPENDIX A

DETERMINATION OF THE STATIC COEFFICIENT OF FRICTION

Relatively simple pull tests were conducted in order to find the static coefficient
of friction, µ, for use in the wipe bending finite element model. Most of the

important parts of the experimental set-up are shown in Fig. A.1. The die and
clamping fixture were from the wiping die set-up and consisted of the following

from Fig. 2.1: die °


2 , pivot °
4 , blank holder °
5 , load cell °
6 , clamping reaction

plate °
7 , and clamping screw °
8 . The die was mounted on the vertical part of a

reaction frame, as shown in the figure. A laminated steel strip was clamped as
in the wipe bending experiments and then pulled downwards by contracting the
hydraulic cylinder. A load cell measured the pull force, P , while a displacement

transducer measured the displacement of the strip, d. The strip had a width of
1.0 in. just like the wipe bending specimens.

Figure A.2 shows a closer, side view of the friction test set-up. The support
fixture was necessary to prevent the clamping fixture from rotating relative to the
die and was mounted on the reaction frame, just below the die. This ensured

that the clamping force remained nearly constant during the test and was applied
perpendicularly to the surface of the specimen.

After the experimental set-up was in place, the following testing procedure was
conducted:

108
Figure A.1. Picture of experiment fixture for friction tests.

Step I: Install the specimen between the die and the blank holder
and align it vertically.

Step II: Start the data acquisition system (which records the clamp-
ing force, hydraulic cylinder force, and displacement of the spec-

imen).

Step III: Increase the clamping force Fc to the desired value using
the clamping screw.

Step IV: Place the arm of the displacement transducer on the top of

the specimen.

Step V: Apply hydraulic pressure to the hydraulic cylinder, so it pulls

109
Figure A.2. Picture of experiment fixture for friction tests.

on the specimen.

Step VI: Stop loading after the specimen starts slipping in the clamp-
ing fixture.

Step VII: Stop the data acquisition system.

Step VIII: Release the clamping force and remove the specimen.

From the experimental data, the value of µ was calculated for each data point

using
P
µ= , (A.1)
2Fc

where the factor of 2 in the equation accounts for the contact on both sides of

110
the specimen. Figure A.3 shows the values of µ for the test with Fc = 300 lb.
Recall that this was the standard value for the force during wipe bending. The

horizontal axis is the displacement of the specimen, d. All three symbols represent
data. The square and triangle indicate the start of slip, where µ = 0.162 and 0.176,
respectively. Experiments conducted with Fc = 100 and 400 lb yielded values of µ

also in the 0.16-0.18 range. Therefore, µ = 0.17 was taken as the static coefficient
of friction and used in the numerical analysis.

µ 0.25

0.2

0.15

0.1

0.05

0
0 2 4 6 8 10
d (x10 −3 in)

Figure A.3. Coefficient of friction for test with clamping force of 300 lb.

111
APPENDIX B

IMPLEMENTATION OF INTERFACE MODELS USING ABAQUS’ UINTER


USER SUBROUTINE

The finite element program Abaqus allows the user to write “user subroutines”
to model aspects of problems that are not included in the standard package. In
this way, the user can implement his/her own element, material model, etc. In this

work, a user subroutine, UINTER, was used to implement the interface models
developed to introduce the effect of the mechanical behavior of the polymer layer

on the bending simulations of laminated steel.


According to the Abaqus manual [29], the user subroutine UINTER:

• is called at the slave nodes of a contact pair with a user-defined constitutive


model defining the interaction between the surfaces;

• is used to define the mechanical interactions between the surfaces;

• is used when a more complex surface interaction behavior than is available


otherwise is needed;

• must provide the entire definition of the mechanical interaction between the

contacting surfaces (hence, no additional surface behaviors can be specified


in conjunction with this capability); and

• is used to update solution-dependent state variables.

112
UINTER is implemented in the input file as the behavior of a contact pair
by modifying the *surface interaction command line. The contact pair here

consists of the bottom of the top steel sheet and the top of the bottom steel sheet
in the laminate, the two surfaces in contact with the polymer layer.The subroutine
was written in Fortran and is outlined below.

SUBROUTINE UINTER(STRESS,DDSDDR,FLUX,DDFDDT,DDSDDT,DDFDDR,
1 STATEV,SED,SFD,SPD,SVD,SCD,PNEWDT,RDISP,DRDISP,
2 TEMP,DTEMP,PREDEF,DPRED,TIME,DTIME,CINAME,SLNAME,MSNAME,
3 PROPS,COORDS,ALOCALDIR,DROT,AREA,CHRLNGTH,NODE,NDIR,NSTATV,
4 NPRED,NPROPS,MCRD,KSTEP,KINC,KIT,LINPER,LOPENCLOSE,LSTATE,
5 LSDI,LPRINT)
C
INCLUDE ’ABA_PARAM.INC’
C
CHARACTER*80 CINAME,SLNAME,MSNAME
DIMENSION STRESS(NDIR),DDSDDR(NDIR,NDIR),FLUX(2),DDFDDT(2,2),
1 DDSDDT(NDIR,2),DDFDDR(2,NDIR),STATEV(NSTATV),
2 RDISP(NDIR),DRDISP(NDIR),TEMP(2),DTEMP(2),PREDEF(2,NPRED),
3 DPRED(2,NPRED),TIME(2),PROPS(NPROPS),COORDS(MCRD),
4 ALOCALDIR(3,3),DROT(2,2)

user coding to define STRESS, DDSDDR, FLUX, DDFDDT,


DDSDDT, DDFDDR,
and, optionally, STATEV, SED, SFD, SPD, SVD, SCD, PNEWDT,
LOPENCLOSE, LSTATE, LSDI

RETURN

END

A brief description of each variable in the argument list is provided below.

The following variables are passed into UINTER:

• (RDISP) relative position between the two surfaces,

• (DRDISP) increments in relative position between the two sheets,

113
• (TEMP) temperature at contact pair,

• (DTEMP) increment of temperature,

• (PREDEF) predefined field variables,

• (DPRED) increments in predefined field variables,

• (TIME(1)) value of step time at the end of the increment,

• (TIME(2)) value of total time at the end of the increment,

• (DTIME) current increment in time,

• (CINAME) user-specified surface interaction name,

• (SLNAME) slave surface name,

• (MSNAME) master surface name,

• (PROPS) property values written in the input file

• (COORDS) current coordinates of this point,

• (ALOCALDIR) array of the direction cosines of the local surface coordinate


system (defined by the geometry of the master surface),

• (DROT) rotation increment matrix

• (AREA) surface area associated with the contact point,

• (CHRLNGTH) characteristic contact surface face dimension,

• (NODE) slave node number involved with this contact point,

• (NDIR) number of force components at this point,

114
• (NSTATV) number of solution-dependent state variables,

• (NPRED) number of predefined field variables,

• (NPROPS) number of property values read in from the input file,

• (MCRD) number of coordinate directions at the contact point,

• (KSTEP) step number,

• (KINC) increment number,

• (KIT) iteration number,

• (LINPER) linear perturbation flag, and

• (LPRINT) flag indicating if a detailed contact printout is requested.

These are the variables that need to be defined or have value of zero:

• (STRESS) stress between the slave node and the master surface at the end
of the increment;

• (DDSDDR) interface stiffness matrix, DDSDDR(I,J) defines the change in


the Ith stress component at the end of the time increment caused by an
infinitesimal perturbation of the Jth component of the relative displacement

increment array;

• (FLUX) heat flux flowing into the slave and master surfaces at the end of

the increment;

• (DDFDDT) the negative of the variation of the flux at the two surfaces with
respect to their respective temperatures, for a fixed relative displacement;

115
• (DDSDDT) variation of the stress with respect to the temperatures of the
two surfaces for a fixed relative displacement; and

• (DDFDDR) variation of the flux with respect to the relative displacement


between the two surfaces.

The variables that can be updated if active are:

• (STATEV) an array of the solution-dependent state variables,

• (SED) the elastic energy density at the end of the increment,

• (SFD) the incremental frictional dissipation,

• (SPD) the incremental dissipation due to plasticity effects in the interfacial

constitutive behavior,

• (SVD) the incremental dissipation due to viscous effects in the interfacial


constitutive behavior,

• (SCD) the incremental dissipation due to creep effects in the interfacial


constitutive behavior,

• (PNEWDT) ratio of suggested new time increment to the time increment


currently being used,

• (LOPENCLOSE) integer flag that is used to track the contact status (whether

open or closed),

• (LSTATE) integer flag where a simple open/close status in not appropriate

or enough to describe the state, and

116
• (LSDI) flag to trigger the current iteration to be treated as a severe discon-
tinuity iteration.

117
APPENDIX C

IMPLEMENTATION SUMMARIES OF MONOTONIC AND CYCLIC

INTERFACE MODELS IN ABAQUS

C.1 Monotonic Model

The interface model used to implement the monotonic, piecewise linear fit

shown in Fig. 2.11 does not require the use of all of the variables listed in Appendix
B because it does not involve heat transfer and simply involves a constitutive law
that governs stress/displacement behavior. Since the bending analysis conducted

is two-dimensional, the relative position (RDISP), the increment in relative posi-


tion (DRDISP), and stress (STRESS) are arrays each with two components. The

first component is the value in the direction normal to the interaction surface, and
the second component is the value in the direction tangent to the surface. Another

variable used is PROPS. It has two components that are read in from the input
file. The first property is the modulus in the normal direction, En , while the sec-
ond property is the damping coefficient, c. The points in the fit shown in Fig. 2.11

were defined in the subroutine itself. The increment in time is DTIME. There is
only one state variable, so STATEV has one component. For STATEV(1) = 0, the

polymer at that point is not broken, whereas STATEV(1) = 1 means the polymer
is broken. The remaining variable used is DDSDDR, which is a 2x2 array. DDS-

DDR(I,J) is the partial derivative of the equation for the Ith stress component
with respect to the Jth component of the relative displacement increment.

118
The subroutine UINTER is called for each pair of contact points, which are
assigned by Abaqus, along the surfaces. Following is an outline of the procedures

used in the implementation of the interface model:

1: The tangential relative position, RDISP(2), is used to find the currently


active linear segment of the piecewise linear fit in Fig. 2.11.

2: Whether or not the polymer has previously failed is checked. If no failure has
occurred, the parameters of the current linear segment are used. If failure

has occurred, the tangential stress component is zero.

3: The tangential relative velocity, δ̇ is calculated for later use.

DRDISP(2)
δ̇ = (C.1)
DTIME

4: The increment in stress in the normal direction, dσn , is determined simply


by multiplying the modulus in the normal direction and the increment in
relative position in the normal direction.

dσn = En DRDISP(1) (C.2)

5: The stress in the normal direction at the end of the increment is updated
by adding the calculated increment to its previous value.

6: The stress in the tangential direction at the end of the increment is found
from

τ = τs + τd (C.3)

119
where τs is the contribution of the spring, given by

τs = (Eti abs(RDISP(2)) + bi ) sign(RDISP(2)). (C.4)

Here Eti and bi are the slope and y-intercept of the current linear segment.
The stress due to the damper is τd , given by

τd = cδ̇. (C.5)

7: The interface model’s stiffness matrix at the end of the increment, DDSDDR,
is found from  
 En 0
DDSDDR =  (C.6)


c
0 Eti + DTIME

8: The values for the desired parameters are printed to the output file and

control is returned to Abaqus.

C.2 Cyclic Model

The implementation of the cyclic model shares many aspects with the mono-
tonic one. Specifically, the shear stress-displacement relation of the spring is taken
to be piecewise linear and therefore can be calculated using Eq. C.4, with the stiff-

ness matrix given by Eq. C.6. The cyclic characteristics of the model, however,
significantly increase the complexity of the model, and make it necessary to keep

track of 18 state variables listed in Table C.1. The property array now contains
14 elements, listed with the values used in this investigation in Table C.2. In the

normal direction, the value of En depends on whether the sheets are separating or
overlapping and can go to zero if failure occurs. In addition a damping term is also

120
used in the normal direction, which requires the specification of another damping
coefficient. In summary, the stresses and stiffness matrix are calculated using Eqs.

C.1-C.6, but the logic to choose the values of Eti , bi and En is much more compli-
cated because it is influenced not only by the current value of displacement, but
also by its previous history.

TABLE C.1

INFORMATION ON STATEV(i)

i Value Details

1 0 Not broken polymer layer

1 Broken polymer layer

2 1 Polymer in first part of the monotonic region, 0-1 in Fig. 3.26

2 Polymer in second part of the monotonic region, Pt. 1 to peak in


the monotonic response (see Fig. 3.21)

3 Polymer is breaking, going from peak to zero load

4 Polymer in failed region

11 Polymer in region 2-3

12 Polymer in region 3-4

13 Polymer in region 4-9

19 Polymer in region 9-10

21 Polymer in region 5-6

Continued on Next Page. . .

121
TABLE C.1 – Continued

i Value Details

22 Polymer in region 6-11

23 Polymer in region 11-12

24 Polymer in region 6-7

28 Polymer in region 12-13

29 Polymer in region 7-8

3 Spring stress in the previous increment

4 0 Using monotonic behavior

1 Using first reversal behavior

2 Using second reversal behavior

5 Increment in δ, DRDISP(2), for the previous increment

6 Value of shear stress at first unloading (at Pt. 2)

7 Value of δ at 1st unloading, δ2

8 Value of δ where shear stress reaches zero during unloading, δ3

9 Value of δ where shear stress leaves 0 during reverse loading, δ4

10 Value of shear stress at second unloading (at Pt. 5)

11 Value of δ at second unloading, δ5

12 Value of δ where shear stress reaches zero during reverse-


unloading, δ6

Continued on Next Page. . .

122
TABLE C.1 – Continued

i Value Details

13 Value of δ at failure in 11-12, δ12

14 Value of δ at the end of breaking region when fail in 11-12, δ13

15 Value of δ at failure in 4-9, δ9

16 Value of δ at the end of breaking region when fail in 4-9, δ10

17 Value of δ at failure in 6-7, δ7

18 Value of δ at the end of breaking region when fail in 6-7, δ8

123
TABLE C.2

LIST OF PROPS(i)

i Value Description

1 109 psi/in Modulus in the normal direction for overclosure

2 Varied Damping coefficient in tangential direction

3 4,210 psi Modulus in tangent direction for 0-1 in Fig. 3.26

4 21,900 psi Modulus in tangent direction for 1-2 in Fig. 3.26

5 0.00795 in. Value of δ at Pt. 1 in Fig. 3.26

6 -1.024 x 106 psi Modulus in tangent direction for breaking region, slope
between failure stress and τs = 0

7 0.02601 in. Value of δ where failure occurs in monotonic loading

8 80,000 psi Modulus in tangent direction during unloading, slope in


2-3 in Fig. 3.26

9 35,000 psi Modulus in tangent direction during reverse loading,


slope in 4-9 in Fig. 3.26

10 250 psi Failure shear stress after cycling

11 -1.024 x 106 psi Modulus in tangential direction for breaking region after
cycling (same value as PROPS(6) for this work)

12 Varied Damping coefficient in normal direction

13 291 psi Failure stress in normal direction during separation

14 29,100 psi/in. Modulus in the normal direction during separation

124
BIBLIOGRAPHY

1. J. Hinrichsen. The material down-selection process for the a3xx. In C. Ver-


meeren, editor, Around Glare: A New Aircraft Material in Context. Kluwer
Academic Publishers, 2002.

2. S. Krishnakumar. Fiber metal laminates-the synthesis of metals and compos-


ites. Materials and Manufacturing Processes, 9(2):295–354, 1994.

3. J.-K. Kim and T.-X. Yu. Forming and failure behaviour of coated, laminated
and sandwiched sheet metals: a review. Journal of Materials Processing Tech-
nology, 63(1-3):33–42, 1997.

4. C. J. Osborn and J. M. Kennedy. Integrating advanced composite materials


into the chassis of a sports car. American Society of Mechanical Engineers,
Design Engineering Division (Publication) DE, 117:915–932, 2004.

5. D. Caldwell and T. Gardeski. High-temperature damping composite. U.S.


Patent 4,223,073, filed October 30, 1978, issued September 16, 1980.

6. A. Eustice. Adhesive interlayer suitable for constrained layer vibration damp-


ing. U.S. Patent 3,658,635, filed April 2, 1968, issued April 25, 1972.

7. W. Hart and W. Korez. High heat, sound damping metal-polymer laminate.


U.S. Patent 4,599,261, filed February 6, 1984, and issued July 8, 1986.

8. Private communication with Dr. Daniel Boss, formerly at MSC Laminates


and Composites, Inc.

9. E. Garsten. Quiet steel mutes vehicle noise. The Detriot News, March 30,
2004. www.detnews.com.

10. J. Jancsurak. Quiet steel drives manufacturing advancements. Appliance


Design, October 26, 2001. www.ammagazine.com.

11. H. Cheng, J. Cao, H. Yao, S. Liu, and B. Kinsey. Wrinkling behavior of


laminated steel sheets. Journal of Materials Processing Technology, 151(1-3
SPEC ISS):133–140, 2004.

125
12. S. Ramasamy. Welding laminated sheet steel. Advanced Materials and Pro-
cesses, 157(3):43–45, 2000.
13. J. Kim and P. Thomson. Forming behaviour of sheet steel laminate. Journal
of Materials Processing Technology, 22(1):45–64, 1990.
14. A. Makinouchi, H. Ogawa, and K. Hashimoto. Finite element simulation of
bending process of steel-plastic laminate sheets. Computational methods for
predicting material processing defects, pages 221–230, 1987.
15. K. Ito, T. Sagawa, and M. Terada. Simple analysis of double-bent in v-type
bending of steel/resin/steel laminates vibration damping sheet. Journal of the
Japan Society for Technology of Plasticity, pages 1490–1496, 1989.
16. M. Yoshida. Press formability of vibration-damping sheets. Journal of the
Japan Society for Technology of Plasticity, 26:394–399, 1985.
17. S. Kalpakjian and S. Schmid. Manufacturing Engineering and Technology.
Perarson/Prentice Hall, 5th edition, 2006.
18. Y.-M. Huang and D.-K. Leu. Finite-element simulation of the bending pro-
cess of steel/polymer/steel laminate sheets. Journal of Materials Processing
Technology, 52(2-4):319–337, 1995.
19. X. Xiao, C. Hsiung, and Z. Zhao. Flexural response of laminated steel. Amer-
ican Society of Mechanical Engineers, Manufacturing Engineering Division,
MED, pages 1011–1019, 2005.
20. M. Weiss, B. Rolfe, M. Dingle, and P. Hodgson. The influence of interlayer
thickness and properties on springback of sps- (steel/polymer/steel) laminates.
Steel Grips (Suppl. Metal Forming 2004), 2:445–449, 2004.
21. Y. Lu and J. Killian. A finite element modeling approximation for damp-
ing material used in constrained damped structures. Journal of Sound and
Vibration, pages 352–354, 1984.
22. L. A. Mignery. Proceedings of the 1995 ASME design engineering technical
conference. part c. American Society of Mechanical Engineers, Design Engi-
neering Division (Publication) DE, 84(3 Pt C):23–33, 1995.
23. H. Yao, C. Chen, S. Liu, K. Li, C. Du, and L. Zhang. Laminated steel
forming modeling techniques and experimental verification. SAE Technical
Paper Series, (2003-01-0689), 2003.
24. B. Gearing, H. Moon, and L. Anand. A plasticity model for interface friction:
application to sheet metal forming. International Journal of Plasticity, 17(2):
237–271, 2001.

126
25. Y. Zhang, P. Zhu, and X. Lai. Finite element analysis of low-velocity impact
damage in composite laminated plates. Materials & Design, 27(6):513–519,
2006.

26. E. Corona and T. Eisenhour. Wiping die bending of laminated steel. Inter-
national Journal of Mechanical Sciences, 49(3):392–403, 2007.

27. E. Corona, T. Eisenhour, S. Yin, and J. Mason. Wall curl in bending of


laminated steel. NUMIFORM, (712):964–969, 2004.

28. M. Kleiner and V. Hellinger. New possibilities for improved bending of vibra-
tion damping laminated sheets. CIRP Annals - Manufacturing Technology,
48(1):217–220, 1999.

29. Abaqus/Standard user’s manual, Version 6.6. Hibbit, Karlsson, & Sorenson,
Inc., 2006.

30. C. Wang, X. Sun, and Y. Bai. Failure mechanism of laminated damping steel
sheet during tensile-shearing. Journal of Materials Science and Technology,
18(1):80 – 82, 2002.

31. F. Pourboghrat and E. Chu. Prediction of spring-back and side-wall curl in 2-d
draw bending. Journal of Materials Processing Technology, 50(1-4):361–374,
1995.

32. B. Chun, H. Kim, and J. Lee. Modeling the bauschinger effect for sheet metals.
ii. applications. International Journal of Plasticity, 18(5-6):597–616, 2002.

33. W. Carden, L. Geng, D. Matlock, and R. Wagoner. Measurement of spring-


back. International Journal of Mechanical Sciences, 44(1):79–101, 2002.

34. J. Kim and P. Thomson. Separation behaviour of sheet steel laminate during
forming. Journal of Materials Processing Technology, 22(2):147–161, 1990.

35. F. Yoshida, M. Urabe, and V. Toropov. Identification of material parameters


in constitutive model for sheet metals from cyclic bending tests. International
Journal of Mechanical Sciences, 40(2-3):237 – 249, 1998.

36. K. Zhao and J. Lee. Generation of cyclic stress-strain curves for sheet metals.
Transactions of the ASME. Journal of Engineering Materials and Technology,
123(4):391 – 397, 2001.

127

You might also like