Karafillis 1992

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Int. J. Mech. Sci. Vol.34, No. 2, pp. 113-131, 1992 0020-7403/92 $5.00+ .

00
Printed in Great Britain. ~ 1992PergamonPressplc

TOOLING DESIGN IN SHEET METAL FORMING USING


SPRINGBACK CALCULATIONS

A. P. KARAFILLISand M. C. BOYCE
Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, U.S.A.

(Received 13 March 1991; and in revised form 3 September 1991)

Abstract--A method to design tooling in sheet metal forming using springback calculations is
presented. The designed tooling produces a part which matches the desired shape, thereby
compensating for springback. To design the appropriate tooling, traction distributions on the sheet
in the fully loaded deformed state are computed using the finite element method. The calculated
tractions are then used to numerically reproduce springback of the desired part shape by elastic
unloading of the part in a reverse manner. The method is examined for materials covering a range of
steel strength and hardening, and is found to produce parts with negligible shape error. The success
of the tooling design algorithm leads to a proposal for an experimental method to design tooling
based on traction distribution measurements.

INTRODUCTION

In sheet metal forming operations, springback is a decisive parameter in designing the


appropriate tooling. Final part shape depends on the springback which occurs after the
removal of applied loads from the deformed sheet and results in the deviation of the product
from the applied tooling shape. This deviation depends on the shape of the part when it is
fully loaded and its contour matches that of the die, as well as on the interface loads and
material properties. Therefore, it is clear that fine-tuning of the tooling in order to eliminate
springback, prior to application in some high or low volume production, is a very difficult
task as it includes all the elasto-plastic and interface aspects of a real forming problem.
Springback can be considered as a dimensional change which happens during unloading
due to the occurrence of primarily elastic recovery of the part. In other terms, springback is
the additional deformation of the processed material from its fully loaded state due to the
application of reverse loeds which result in a net zero load on the part. Here, the
configuration in which the part shape matches the tooling shape during loading will be
referred to as the fully loaded state. The final unloaded configuration of the part includes
external deviations, i.e. shape errors and local thinning of the part and internal deviations,
i.e. residual elastic stresses and strains. Springback has been approached by several authors
with theoretical and/or experimental methods. Various theories on the prediction of
springback for simple cases of pure bending and/or stretching have been proposed [1-7].
Concepts of springback as a manufacturing problem have recently been discussed by
Tozawa [8] and Umehara [9]. Ayres [10] examined the problem of curling as a result of
springback in high-strength steels and proposed a method to eliminate it. Liu [11]
considered springback as a control problem in relation to the restraining force. While both
die and restraining force mechanism are relevant to the final product quality and to the
process success, the latter can be achieved by regulating the restraining force mechanism
(binder or drawhead) in order to avoid instabilities due to excessive in-plane tensile or
compressive conditions. This stability problem has been experimentally approached by
Hardt and Fenn [12] by providing estimations for the optimum binder force which
corresponds to the maximum cup height and by the construction of corresponding global
control trajectories to eliminate problems associated with in process variations. Sim and
Boyce [13] also demonstrated a computer modelling approach of the same process control
problem by modelling the forming operation using FEM and controlling the binder force
based on local failure information in order to avoid instability, thereby dramatically
increasing the cup height. However, as an initial step in the present work, the shape control
aspect of forming is treated separately from the binder constraint control in order to assess
springback perturbation to shape description independently from draw-in effects.
113
114 A.P. KARAFILLISand M. C. BOYCE

Assuming an appropriate binder condition, the work developed here is to design the die
shape which yields the correct part shape of some complex geometry. Relevant studies have
been demonstrated for various forming methods. Stelson and Gossard [14] have presented
a method where a simple three-point bending operation is controlled by in-process
measurements. Hardt et al. [15] have demonstrated the use of direct springback estimation
from in-process measurements for simple roll bending and further extended this to
automatic straightening [16] and twisting [17] by proposing methods for the in-process
estimation of either the elastic springback or the extrinsic constitutive relationships (i.e.
moment-curvature). In a recent work, Webb [18] and Webb and Hardt [19] have developed
a closed-loop control strategy to control the shape of the part produced based on a transfer
function between an incremental change in die shape and the resulting change in part shape,
estimated from a pair of successive forming cycles. In the same work, shapes are described in
the spatial frequency domain of the shape discrete Fourier transforms. Knowledge of the
elastic-plastic material behavior was not used and the interface, loads were not measured.
On the other hand, complete knowledge of the material behavior and interface normal
and friction stress distribution might allow for the construction of a closed-loop control
algorithm based on force measurements, instead of displacement and part-shape error
measurements. In the present work, a complex axisymmetric part shape is considered,
including a flat central area, a long flange and an intermediate double curvature area; see
Fig. 1. This shape has been selected in order to capture all the difficult features of a real
stamping case. The forming of this part from an initially flat metal sheet with tooling shape
matching the desired part shape and different material properties and friction conditions
has been simulated using the FEM. In each case, forming traction distributions and/or
produced part shapes were recorded and used in the developed closed-loop control
algorithms in order to design the correct die shape required to produce a desired part shape.

FINITE ELEMENT MODEL

The improvement of computational algorithms used in finite element analysis and the
increasing ease of access to fast and powerful computers has made the use of the finite
element method in the detailed analysis of complex sheet metal forming operations a
realistic and popular tool (e.g. [13, 20-22]). For the production of such complex shapes

(BrNDERLOAD) (BINDERLOAD)
I

(SUPPOR3]REACTION)

FIG. 1. Tooling in a sheet metal forming operation with a matching die.


Toolingdesignin sheetmetal forming 115

containing changes in curvature, the use of matched dies is generally necessary. The
matched dies result in a condition of double contact (contact on both sides of the blank)
during deformation, adding considerable complexity to the already numerically difficult
contact problem which exists in sheet forming simulations. For this reason, we have chosen
to model the problem with a detailed finite element mesh of the blank consisting of three
layers of axisymmetric continuum elements (as opposed to shell elements which cannot
handle through thickness compressive straining). The mesh thus provides six integration
points through the thickness of the blank for stress computation. This feature adequately
accounts for both the stretching and bending incurred by the blank. It is particularly
important to capture the bending effects in the simulation of forming relatively shallow
shapes containing curvature changes, since these constitute an important component of the
springback response of the sheet on unloading. Three rigid surfaces were used to simulate
the contact between the blank and the die, the binder and the punch. The contact algorithm
used, ABAQUS [23], computes the amount of initial overclosure between the surface of the
deforming body and the rigid surface, and the relative shear sliding between the two. Four-
node axisymmetric interface elements with an a priori defined tangent stiffness are used to
simulate contact having three nodes attached on the surface of the deformed body and one
node located in the rigid surface. The virtual work of the surface tractions is given by:

5w, = fs(pSh + z,67,)dS = 5w, o + 6IV,,, (1)

where p is the normal pressure, h is the distance from the surface in the normal direction, T,
is the shear contact stress and 7, is the relative shear with respect to the rigid surface. We
assume the rigid surface to be softened and therefore p = p(h), where p = 0 when
h > 0.015 mm. However, when h = 0, dp/dh ~ oo. For this case a Lagrange multiplier is
used to impose the constraint p = p, where/~ is an independent variable. The contribution
to surface virtual work due to surface pressure may then be written as:

6lPs, fs(~rh + 62(p -- }))as, (2)

where 62 is the Lagrange multiplier.


The friction was modelled by the standard Coulomb or, according to Dowson [24], the
Da Vinci or Amonton friction model. The friction coefficient was considered to be the same
for the material-punch and the material-binder interface and equal to 0.15 in all cases
unless otherwise mentioned. An upper limit in the developed shear stress was also
considered. The contribution to virtual work due to surface shear stresses with the
Lagrange multipliers formulation is:

fs (z'67" + ~,6~,)dS, (3)

where ~, are Lagrange multipliers introduced to enforce the constraints for sticking friction,
and ~, is the conjugate to ~, shear strain. Jacobians of Eqns (2) and (3) contribute to the
associated Newton algorithm of the implemented finite element code, ABAQUS [23] and
the discretized equilibrium equations are solved. Nodal forces of the surface of the deformed
body in the local coordinate system can be calculated using the stress distribution in the
interface elements:

F{ = ~s criknkNjdS' (4)

where F{ are the tractions at the jth node in the ith direction, Nj are the displacement
weighting functions of the element, and nk is a unit vector normal to the interface element
surface. Calculated nodal forces are then rotated to the global coordinate system. Boundary
conditions are imposed, fully constraining the die. The punch and the binder are con-
strained such that only vertical motion is permitted. Hence, the forming process is simulated
by imposing a prescribed downward displacement on the punch and the application of a
MS 34:2-C
116 A.P. KARAFILLISand M. C. BOYCE

vertical ring load of 4500 N on the binder, as in Fig. 1. An isotropic elastic-plastic


constitutive model with isotropic strain hardening was used to simulate the material
response. The elastic behavior was taken to be linear and the plastic response was modelled
using the von Mises yield criterion for isotropic materials. The yield surface was taken to
have evolved with plastic straining and an associated flow rule was assumed.

SPRINGBACK STUDIES
It is important to understand springback mechanics in order to surpass part shape
deviations due to springback by means of an algorithm which controls the tooling shape
and consequently the final part shape. For this purpose several stamping cases with the
tooling of Fig. 1 but with various forming parameters, i.e. various material properties and
friction conditions, are examined.
Two cases of material behavior are modelled, one with an initial yield stress of 250 M P a
(hereinafter referred to as low-yield) and the other with an initial yield stress of 400 MPa
(high-yield), see Fig. 2. Figure 3 depicts the deforming mesh at several stages during the
forming of high-yield material with the tooling of Fig. 1. The sequence of blank contact with
the punch and the matching die is shown. The sheet bending during loading is depicted,
where bending deformation clearly affects all regions of forming this shallow part. The
combined effects of bending and stretching can also be seen in Fig. 4 where the von Mises
equivalent tensile stress distribution near the top (punch-face) and bottom (die-face) surface
of the sheet in the fully loaded state is shown. The large difference between the von Mises
stress distribution near the punch-face and near the die-face is indicative of the bending
stresses and strains incurred by the sheet where the stress is always higher in the tensile side
of the bend due to the combination of a tensile bending stress with a tensile stretching stress.
Distinctive crossover of the punch-face von Mises stress with the die-face von Mises stress
occurs near all inflection points of the shape indicating the entire part shape is affected by
bending, i.e. there exist no long domains of simple stretching. We also note that the von
Mises stress greatly exceeds the yield stress in several locations, indicating a large amount of
strain hardening and plastic straining.
The corresponding traction distribution in the forming area for the fully loaded state is
depicted in Fig. 5. In this configuration the punch shape is offset from the die shape by the
sheet metal thickness. It is shown that regions of contact with the die cavity are alternating
with regions of contact with the punch, i.e. there exist no areas where the blank is
simultaneously in contact with both punch and die. We must also note the difference in the
traction distribution between the low-yield material and the high-yield material, where

800.0

?00.0

600.0 f -

i 600.0
400.0
300.0
200.0
Low-Yield Material
100.0
High-Yield Material

0.0 I t I t I a I i
0,0 0.2 0.4 0.6 0.8 t,O

T R U E STRAIN

FIG. 2. True.stress-true strain curves of the examinedmaterials.


Tooling design in sheet metal forming 117

10.0

0.0

-10.0
-80.0 -40.0 0.0 40.0 80.0
X [mm]
(a)

10.0

0.0
h~
-i0.0 , i , i , , , i , , , I , , i

-80.0 -40.0 0.0 40.0 80.0


X [mm]
(b)
10.0
r'-'n

0.0~ I '
i
-I0.0 I I

-80.0 -40.0 0.0 40.0 80.0

X [mm]
(¢)

10.0

0.0 ~ I '
>-
i
- 1 0 . 0 , , , i , , , i , , , I , ,

-80.0 -40.0 0.0 40.0 80.0

X [mm]

(d)
I0.0
r---n

0.0

-10.0
-80.0 -40.0 0.0 40.0 80l0
X [mm]

(e)

I0.0 , ,

g 0.0

-I0.0
-80.0 -40.0 0.0 40.0 80.0
X [mm]
(f)
FIG. 3. Deforming mesh of high-yield material during (a)-(d) loading and (e), (f) unloading.
118 A.P. KARAFILLIS and M. C. BOYCE

800.0 I I

Mises s t r e s s
700.0 P u n c h face (Hi gh-Y i e l d) ~ D m

Die face (High- Y i e l d)


P u n c h face ( L o w - Y i e l d )
800.0 Die face (Low-Yield)

600.0

400.0
m

300.0
/i',
200.0

, i , i
100.0
0.0
'
20.0
'
40.0 80.0 80.0
RADIUS Cram]

FIG. 4. Von Mises equivalent stress distribution in the fully loaded state.

2000.0

1000.0

0,0'

-1000.0

Horizontal direction
Low-Yield Materia l
-2000.0 P u n c h - f a c e side
D i e - f a c e side

-3000.0
0.0 20.0 40.0
RADIUS [ m m ]
(a)
8000.0
i Low Y- ield Material
4000.01 Die face side
P u n c h face side

2000.0 m m ~
a~
o 0.0

-2000.0

-4000.0
0.0 20.0 40.0
RADIUS [ram]

(b)

Figure 5(a,b)--caption on facing page


Tooling design in sheet metal forming 119

2000.0

1000.0

g oo

~0~ -1000.0

Horizontal direction
High-Yield Material
-2000.0 Punch-face side
Die-face side

-3000.0
0.0 20.0 40.0
RADIUS [ram]
(o)
6000.0

Vertical direction [
High-Yield Material
4000.0 Die face side . ,
P u n c h face side

Z 2000.0

0.0

-2000.0

-4000.0 i ' ,
0.0 20.0 40.0

RADIUS [ram]
(d)
F]G. 5. Traction distribution in the fully loaded state for (a), (b) the low-yield material and (c),(d) the
high-yield material.

tractions are higher in the case of high-yield material as a result of the higher stress regime
developed in this case.
During the unloading stages (see Fig. 3) springback occurs as the result of elastic
deformation effectively under tractions equal and opposite to those incurred at the fully
loaded state, where the traction distribution is depicted in Fig. 5. Analogously, the
springback or curvature recovery of an elastic-plastic beam in bending is simply the applied
moment divided by the elastic bending rigidity "EI". As in simple bending, the final
configuration of our formed part consists of a residual stress distribution, depicted in Fig. 6,
and a final shape which deviates from the desired one. The part shape error is depicted in
Fig. 7 as a function of the radial position, where all errors are calculated in the die-face side
of the material and refer to deviation in the vertical direction. Residual stresses are depicted
in Fig. 6 in terms of the equivalent von Mises stress. In the case of low-yield material the yon
Mises residual stress is everywhere less than the corresponding "forming" yon Mises stress
in the fully loaded condition and therefore secondary yielding during the unloading does
not occur. However, in the case of high-yield material the residual yon Mises stress exceeds
the forming stresses in a region of the part located at a distance of about 38 mm from the
120 A.P. KARAFILLISand M. C. BOYCE

800.0

700.0 Residual Mises stress


Low-Yield Material
i 600.0 P u n c h face
Die face

800.0

400.0

3~.0 f!

,=o[ AA. t
0.0 ~ ' ' i , , , I , I I i i i I
0.0 20.0 40.0 60.0 80.0

RADIUS [mini
(a)
800.0 , , , . ,

700.0 Residual Mises stress


a High-Yield Material
800.0 P u n c h face
Die face

I 800.0 A
fl
II
m 400.0

~ 300.0 A I',
200.0

11111.0

0.0 I I ' ' I , , , I , , ,

o.o zo.o 40.0 0o.o 8oo


RADIUS [mm]
(b)

FIG. 6. Residual yon Mises equivalent stress distribution for (a) the low-yield material and (b) the
high-yield material.

0.2
, , , i , , , , , I

0.0 - ~. -- I

i -0.4

--0.~
]
-O.e ' ~ / ¢ '

i
V

i i
....
~
w..
i i ,
r
ape error
Total, Low-Yield
Springback, Low-Yield -I
Total, High-Yield
Springback. High-Yield[
Thinning
i I
,
i
i

I
i i
1
I
[

0.0 20.0 40.0 60.0 80.0

RADIUS [ m m ]
FIG. 7. Shape error in the produced part for low-yield and high-yield material; thinning error is the
same for the two materials.
Tooling design in sheet metal forming 121

center, i.e. at the edge of the die cavity, and therefore additional plastic straining occurs
during the unloading at this point. As this region of reverse yielding is confined to a very
narrow area, it is not expected to significantly influence the assumed elastic nature of the
unloading process in our inverse springback calculations discussed later.
In Fig. 7 part error due to local thinning as a result of the in-plane stretching of the sheet
is also depicted. The difference in the thinning error between the low-yield material and the
high-yield material is negligible and therefore only the low-yield thinning error is re-
presented here. It can be seen that the amount of part error due to springback is much
higher than the amount of error due to thinning and dominates the total error. Therefore, it
appears as a consequence that fine-tuning of the tooling must be performed on the basis of
springback error elimination.
The influence of friction conditions on springback was also examined. The root mean
square (RMS) of part shape error with respect to the coefficient of Coulomb friction is
depicted in Fig. 8. It can be seen that springback is not sensitive to friction conditions for the
geometries considered and therefore slight discrepancies of the characterization of the
die-blank and binder-blank frictional interaction properties used in the FE modelling are
not strongly influencing the accuracy of the springback calculations. This is primarily a
result of the part geometry chosen where there is little relative motion of the material blank
with respect to the die and tooling as well as very little material draw-in from the binder. We
must note here that a standard binder force of 4500 N was used in all of our simulations.
However, the magnitude of the binder force is not expected to influence the springback
calculations, as long as it is strong enough to not allow any draw-in of the material.

TOOLING DESIGN USING FORCE DESCRIPTOR

As mentioned above, springback is the recovery of the part during the unloading. The
unloading process is primarily elastic; however, secondary yielding is a possibility. In our
construction of an algorithm for tooling design we consider the unloading to be elastic.
Consider the fiat blank of Fig. 2(a) with initial position vectors noted by x and load it to a
desired shape described by the displacement field u*(x), see Fig. 2(d), which is incurred by
the traction field fl (x), see Fig. 9. When fl (x) is removed, the blank will spring back by an
amount of Usb~(fl, X) under the action of the traction field [ - fl(x)]. This leaves a
displacement field in the unloaded part of:

UU] I (X) = U*(X) -- llsb I (fl,x). (5)


However, it is desired that:
u,ll (x) = u* (x). (6)

0.6

0.4

r~ 0 C ~ ~"

0.2
O

Low-Yield Material
High-Yield Material

0.0 I I I I
0.0 O. 1 0.2 0.3 0.4 0.5
COULOMB FRICTION COEFFICIENT

FIG. 8. RMS of part shape error for various Coulomb friction coefficients.
122 A.P. KARAFILLISand M. C. BOYCE

10.0
r--n

t.==a
0.0

-10.0
-80.0 -40.0 0.0 40.0 80.0
X [mm]

FIG. 9. Traction vectors in the fully loaded part.

Therefore, since error U~b,(fl, x) is due to elastic unloading under the action of - fl (X), we
then consider elastic loading of the shape described by u*(x) with f~ (x). Assuming an elastic
response of the material, we obtain:
u12(x) = u*(x) + ue1(fl, x). (7)
However, it is assumed that residual stresses do not influence elastic behavior and therefore
they are not considered during the elastic loading. We now load the blank to a shape
described by the displacement field u u which will be incurred by the traction field f2(x).
When f2(x) is removed, the blank would spring back by an amount Usb~(f2, X) due to elastic
recovery from traction distribution [ - f2(x)]. The shape after unloading will be described
by:
Uul2(X) : BI2(X ) -- Usb2(f2, X). (8)

On the other hand:


ul~(x) = u*(x) + ucl(fx, x); (9)
SO~
Uulz(X) = U*(X) + Uel(fl, X) -- Usb2(f2, X). (10)

If fl (x) = f2 (x) and supposing reversibility of the elastic deformation occurred, we obtain:
u~,(f 1, x) = U~b~(f2, x ) (11)
and therefore
uu~2(x) = u*(x). (12)

Equation (12) shows that uu~2(x) coincides with the desired shape and motivates for
extraction of the forming tractions f(x) in some iterative way until we obtain fi(x) = fi + l(x).
This step is performed using the two alternative methods discussed below.
I terational traction extraction aloorit hm (I TEA )
This method is based on the algorithm outlined in Fig. 10(a). The forming tractions f~(x)
are extracted from simulation of the forming of the blank under the boundary conditions
described from the displacement field u~,(x). When fi(x) are calculated we replace fi-1(x)
with fi(x) and load elastically u*(x) with f~(x). The elastically deformed shape is described
from the displacement field u~,+l(x) which is considered as the next tooling shape. The
iterations are stopped when convergence has been obtained, i.e. the amount of part error
becomes very small and/or the difference in the tooling shape between two successive
iterations becomes negligible.
Extrapolation traction extraction algorithm (ETEA)
In the extrapolation traction extraction algorithm, prehistories of forming tractions f~(x)
are recorded as a function of punch displacement which is considered to be equal to the
central node displacement when loading to the desired shape u*(x). These histories are then
extrapolated to deeper punch displacements. Thus, a relation is provided between displace-
ment of the central node of the part and the nodal forces in the interface surfaces of the finite
element mesh. This relation is used to update the applied forces in the desired shape as its
deformation proceeds; see also Fig. 10(b). The difference between ETEA and ITEA is that
START >
START START

+ #
TOOLING SHAPE
MATCHINGDESIRED TOOLING SHAPE MATCHIN~ MAGNITUDE, TOOLING /
SHAPE DESIRED SHAPE | SHAPES° PART SHAPES./
+ DESIRED SHAPE /
+ LOADING AND UNLOADIN~
I b SIMULATION | 4,
TOOLING SHAPE LOADING SIMULATION + CALCULATE DFF'S
MATCHING ELASTICALLY RECORDING OF LOADING [
DEFORMED DESIRED TRACTION PREHISTORY
PART SHAPE ,-]
+ FILTER DATA 0
RECORDING OF 0
LOADING TRACTIONS SIMULATIONOF ELASTIC I m
+ LOADING OF DESIRED I
SHAPE WITH TRACTIONS t CALCULATE
SIMULATION OF CORRESPONDING TO THE PART F/~ROR
ELASTIC LOADING UNLOADING SIMULATION FULLY LOADED STATE
EXTRAPOLATION OF +
OF DESIRED SHAPE FULL LOADING TRACTION~
WITH RECORDED TO CURRENT TOOLING CALCULATE 1
I NEXT DIE SHAPE = TRANSFER FUNCTION
LOADING TRACTIONS SHAPE
PARTSHAPE DEFORMED SHAPE

GENERATE I E
NEW TOOLING DFr
÷ B.
t . NO NO CALCULATE IDFT
,I

TOOLING SHAPE
/
(=) (=) (=)
(a) (b) (c)
Webb, [18]
FIG. 10. Flow graphs of (a) ITEA, (b) ETEA and (c) TFA.
124 A.P. KARAFILLISand M. C. Bovc~

ETEA relies on force prediction from extrapolation of the traction histories of the first
forming case instead of the traction calculation from successive forming simulations in
ITEA. Therefore, although ETEA will not provide as accurate a description to successive
iterational traction distributions as ITEA, it only requires one elastic-plastic forming
simulation with tooling described from u*(x) as compared with the ITEA method which
requires one elastic-plastic forming simulation per iteration. However, convergence can be
confirmed when the difference in the tooling shape between two successive iterations
becomes negligible.

TOOLING DESIGN USING SHAPE DESCRIPTOR


Hardt and Webb 1-25] have demonstrated a simple form of tooling shape control which
changes the die shape locally in response to local shape error, regardless of any coupling
between adjacent points on the sheet metal part. This discrete control law led to poor
accuracy and very poor convergence due to its inability to capture the coupling between
loading and deformation at different positions. Another control strategy has been de-
veloped from the same authors 1-19] which considers the part shape error related to die
shape change via a shape-based transfer function as opposed to a position-based transfer
function. Part shape and die shape are described in the spatial frequency domain using
discrete Fourier transform of the shapes, thus capturing the dominating features of the
entire shape better than the piecewise shape description; a transfer function is developed for
each frequency individually. This control law, hereinafter referred to as transfer function
algorithm (TFA), is based on comparison of changes in the measured part shape error due
to die shape changes and requires two initial forming trials to be initiated; see Fig. 10(c). The
spatial frequency representation of the computed change in part shape error due to change
in the die shape between the two initial dies is used to construct the transfer function H(~a)
and design a new die using the algorithm. This algorithm has been demonstrated to
accurately design tooling and converge with few iterations. Convergence strongly depends
on the initial trials chosen. TFA is also simulated in the present analysis for comparison
purposes.

SIMULATION NOMENCLATURE
The discussed tooling shape algorithms have been used for the determination of the
tooling shape which yields part shape matching the tooling of Fig. 1. Simulations of the
forming process and of the artificial elastic loading used in the algorithms were conducted
for the various sets of matching die shape and material behavior considered; see Table 1.
Sets A-D consist of one initial or "trial" die set at 100% (first cycle) of the desired part
shape, followed by the simulation of the elastic loading of the desired shape under the
tractions incurred by the forming with the initial die in the fully loaded state. The elastically
deformed part after the elastic simulation is considered to be the designed tooling shape of
the second cycle of the algorithm. In the third cycle the desired shape is loaded elastically
with forces extracted either by ITEA or ETEA. Iterations are stopped after the third cycle as
convergence is obtained. However, even though some of the conducted simulations were
not required from the corresponding algorithm, they were conducted for the evaluation of
the resulting part error. For example, case C2 is not implemented in ETEA but it has been
conducted here in order to evaluate part shape error and observe the convergence of the
algorithm.
Sets E and F consist of two trial dies followed by a new die, designed using the TFA in
conjunction with the part shape errors computed from the two initial simulations. For
example, set E consists of the simulation of the forming using the low-yield material, where
the two trial dies were taken to be 95 and 100% of the desired part shape, and of the
simulation using the designed tooling which was designed using the transfer function
calculated in conjunction with the part shape errors computed from the simulations of the
two trials.
Tooling design in sheet metal forming 125

TABLE 1

Algorithm Set Simulation Material Case CPU time PI1 PI2

ITEA A Tooling 100% (lst cycle) Low-yield G1 9370.3


Elastic loading (2nd cycle) Low-yield AE 1 986.6
Design. Tooling (2nd cycle) Low-yield AI 9410.7 0.35 0.31
Elastic loading (3rd cycle) Low-yield AE2 986.6
Design. Tooling (3rd cycle)t Low-yield A2 9403.7
B Tooling 100% (lst cycle) High-yield G2 9409.3
Elastic loading (2nd cycle) High-yield BE 1 990.1
Design. Tooling (2nd cycle) High-yield B1 9390.4 0.38 0.33
Elastic loading (3rd cycle) High-yield BE2 990.1
Design. Tooling (3rd cycle)* High-yield B2 9408.1

ETEA C Tooling 100% (lst cycle) Low-yield G1 9370.3


Elastic loading (2nd cycle) Low-yield CE1 986.6
Design. Tooling (2nd cycle)* Low-yield C1 9410.7 0.62 0.47
Elastic loading (3rd cycle) Low-yield CE2 986.6
Design. Tooling (3rd cycle)* Low-yield C2 9405.1
D Tooling 100% (lst cycle) High-yield G2 9409.3
Elastic loading (2nd cycle) High-yield DE1 990.1
Design. Tooling (2nd cycle)t High-yield DI 9390.4 0.63 0.40
Elastic loading (3rd cycle) High-yield DE2 990.1
Design. Tooling (3rd cycle)* High-yield D2 9408.1

TFA E Tooling 95% (lst cycle) Low-yield G3 9443.3


Tooling 100% Low-yield G1 9370.3 0.40 0.24
Design. Tooling (2nd cycle)t Low-yield E1 9405.6
F Tooling 95% (lst cycle) High-yield G4 9382.1
Tooling 100% Low-yield G2 9409.3 0.40 0.38
Design. Tooling (2nd cycle)t Low-yield F1 9405.6

t Simulations not necessary for the algorithm but conducted here for the analysis of the part shape error. CPU
times of these simulations should not be included in the required CPU time. All CPU times were evaluated on a
SUN 4/60 machine.

RESULTS

Representative part shapes in the unloaded but clamped condition, yielded from toolings
of successive cycles when the force descriptor ITEA and the shape descriptor TFA methods
on the high-yield material were used, are depicted in Fig. 11. It can be seen that a coarse
correction in the part shape happens in the second cycle of the force descriptor algorithms,
see Fig. 1l(a), while in the third cycle these algorithms provide a fine correction in the part
shape. Iterations are stopped in the third cycle as both force descriptor algorithms converge
to a tooling shape. The final shape deviates from the desired one in regions where thinning
affects the shape of the measured surface (all considered surfaces describe the die-face side of
the deformed sheet), t In the case of TFA, see Fig. 1 l(b), the two first cycles are necessary
trials to initiate the algorithm and therefore they exhibit a large error. The third cycle
eliminates the part error by performing corrections on the die shape in the frequency
domain for frequencies which exhibited a part error greater than some threshold value
defined prior to initiation of the algorithm.
The error profiles of all cycles of the discussed algorithms appear in Fig. 12, where the
decrease of the part shape error using ITEA or ETEA can be seen, as compared with the
part shape error occurring when the G1 tooling is used. This comparison reveals the

tFor reasons of simplicity all calculations refer to unloaded but clamped parts. For the implementation of the
force descriptor algorithms, the hinder force in the elastic artificial reloading of the material was constant and
equal to the binder force incurred during the elasto-plastic forming of the material. However, possible changes of
the interface loads in the binder area due to unloading of the part were considered in the elastic reloading by the
implicit change of the interface loads as the forming loads were applied in the punch area.
126 A.P. KARAFILLISand M. C. BOYCE

6.0 l , , , [ , , , i , , , I

P a r t shape(High Yield)
(ITEA)
G1
lql ......
B2
Desired shape .....
0.0

-5.0

-10.0 I I

0.0 20.0 40.0 60.0 80.0


RADIUS [ram]
(a)
5.0 ' ' I ' ' , I ' ' I

P a r t shape(High Yield)
G1
G3 ......
F1
Desired shape .....
0.0

f'~ . / '
o

-5.0

-10.0 i , , , , , , , , , ,
0.0 20.0 40.0 60.0 80.0

RADIUS [ram]
(b)
FIG. |I. Part shapes after unloading for the high-yie|d material using (a) ITEA and (b) TFA.

effectiveness of both force descriptor algorithms, as the part shape error is strongly
eliminated in the second cycle of the algorithms. Note that both ITEA and ETEA yield the
same tooling and, consequently, the same part shape error in their second cycle; see also
Table 1. The RMS value of the part shape error with respect to the cycle number is depicted
in Fig. 13. RMS values are normalized by the corresponding value of the part error due to
thinning, RMSth, which is equal to 0.05 mm. The second tooling of the force descriptor
algorithms reduces the normalized RMS value of the part shape error distribution to 1.8 or
28.3% of its initial value for the low-yield material, and to 2.8 or 28.4% of the initial value,
for the high-yield material. The third cycle yields part shape error distribution for ITEA and
for ETEA with RMS slightly less than in the second cycle, i.e. the greater reduction in the
part error occurs in the second cycle. However the RMS of the part shape error in the third
cycle is very close to the RMS of the part shape error due to thinning of the sheet, indicating
springback error has been completely compensated. In Fig. 13 the width W = [(maximum
error) - (minimum error)] of the part shape error, normalized by the value of the error with
Wth due to the thinning of the sheet, is also depicted. We can see that the width of the shape
error is also decreased considerably in the second iteration of the force descriptor
algorithms. However, the third cycle of ETEA tends to increase the value of the error width
Tooling design in sheet metal forming 127

0.2 0.2

,l

0.0 0.0

i -0.2 -0.2

-0.4 . -0.4
w .l Part s h a p e error

-o.8
A1 . . . .
A 2 .--
. ..-.Z
.
-0.6 "v' N/ cz -. .T. . -
DI__..

-0.B
02 .......
B1 . . . .

-0.8
0.0
' ' '
20.0
' ' t ,
D2
i ,
.

,
. . . . .

+ i ,
.

, ,
o.o 2o.o 40.0 eo.o 8o.o 40.0 60.0 00.0
RADIUS [ m m ] RADIUS [ram]

(a) (b)
0.2

0.0

-0,2

L==e
-0.4

-0.6
~VPart
"~ ! l (rFA)
shape
"
error

GI--

-0.8

-I.0 \l,V,
0.0
,

20.0
RADIUS
+ I

40.0
G4--
E2 . . . .

[ram]
80,0
, , ,

80.0

(o)
FIG. 12. Error profiles using (a) ITEA, (b) ETEA and (c) TFA.

as it slightly overforms the part. Although the results of ETEA appear to be somewhat
inferior as compared with ITEA we must keep in mind that computational effort in ETEA is
considerably lower than in ITEA; see also values of CPU time spent on each algorithm
tabulated in Table 1.
A comparison of the force descriptor algorithms with TFA shows that the reduction of
the part shape error in those algorithms is of the same order as that found with TFA.
However, the final error distribution differs due to the thinning of the sheet as TFA acts in
the displacement field and cannot distinguish between springback error and thinning error.
A discussion on the performance of each algorithm in terms of path shape error and
computational time can be found in the Appendix.
Webb [18] has shown that evaluation of the part shape error in the frequency domain for
successive forming cycles when the TFA method is used, reveals useful information on error
peaks in discrete frequencies; TFA modifies the die shape in order to eliminate these peaks.
In the present analysis, shape errors of parts yielded from successive cycles of all discussed
algorithms were evaluated in the frequency domain using the discrete Fourier transform, see
Fig. 14. It can be seen that peak errors occur for some discrete frequency numbers in cases
G1, G2, G3 and G4. TFA compensates for these individual peaks (for example frequency
number 2) reducing the error in the corresponding frequencies to values which are a small
128 A . P . KARAFILLISand M. C. BOYCE

15'0I ' Low-'Yield Material


| RMS/RMS~:
,t ITEA o
10.0 ~ . ETEA . . . . .
TFA _-
W/W~:
ITEA - - Q ' -
ETEA ----4.--
TFA .=,=0.--

0.0 I I
Z 3
ITERATION CYCLE
(a)
15.0 i i

High-Yield Materi~l

RMS/RMS~:
ITEA o
10.0 • 11 ETEA --'~'--
\ ~ TFA :
\ \ w/w,~:
o~ "~bM
\~ ErEdk"-'--~---

~ 5.0

0.0 I I
Z 3

ITERATION CYCLE

(b)
FIG. 13. Normalized RMS and width of part shape error in the discussed algorithms for (a) the low-
yield material and (b) the high-yield material.

fraction of the initial error, see Fig. 14(c). However, TFA is not activated for all frequency
numbers and therefore there are some frequencies (for example frequency number 4) where
error magnitude in the final cycle of the algorithm is not decreased as compared with the
base cases G 1 for the low-yield material and G2 for the high-yield material. To eliminate the
error in these frequencies other "trial" toolings should be used in order to excite activation
of the algorithm in all discrete frequencies of the die shape description. Therefore, initial
toolings must be chosen carefully in order to excite the algorithm in as many frequencies as
possible. In the present analysis, TFA was excited in two frequencies for both low-yield and
high-yield material. However, the final part error spectrum is confined to very small error
values for all frequencies and therefore the algorithm was considered to converge very well.
It must also be noted that TFA treats the part shape error as a dimensional disturbance in
the frequency domain and therefore allows for elimination of both springback and thinning
error. Therefore, it seems that TFA is a very effective approach to the die-tuning problem
regarding shape error elimination in stamping operations.
In Fig. 14 (a) and (b) the discrete Fourier transform of the part shape error encountered in
all cycles of the force descriptor algorithms is depicted. It can be seen that force descriptor
"[81] u!~mop £ouonbo~j oql
m .zo u!emop [e!]eds u~.zsol.~uD oq) u! ~oql!o 'os~o s!ql u! poSeu~u.z oq ~ouu~o 1uomo~nsuom
:to~t:to od~qs l:t~d oq~ se qldop s~! u! soi~ueqo )dnzq~ jo ooaj oq ol odttqs ~.~d oql oz!nboa
spoqlatu aold.uosop odeqs '~sualuoo u I "papnpu! oae suopoos leO!laOAUOqA~U0AO'SO!a~otuOZ$
laed oIq.tssod lle aoj pasta oq utto ,~aql se lsnqoa LaoA oq o~ tuoos OSle spoq~om aold!aos
-op oozo~I "suo!luaop!suoo adeqs uo paseq Su!oq aoqlo oql pue suo!leaop!suoo uo!peal
uo poscq Su!oq ouo aql 'u.tetuop ,~auonboaj oql u! £[ael!tu!s l~e spoqlotu qloq se poqlotu
aold.uosop odeqs oql ol £aeluamoldmo~ sg poaop!suoo oq ueo poqlom zo~d!aosop ooaoj
aq~ 'oat.logds.zod .Ioqaou~ u.l 'oaoJoaOq.L "pOleAl.:t01es'em V~I,L o~zoq~ so!ouonb~j ou.z~s ~ql ut
suoddeq p o q l o m ao~d!.zosop oozoj aql zoj aoaao odgqs laed oql jo uoponpoa o)niosqe .~o~eoafl
o q I "so.touonboaj o)o.zos!p l[e .zoj SOnlt~A aoaao llems £aOA O~ pOU~IUO0St. pu~ aoaao ~u.tuu!ql
aql ol spuodso~uoo qo!qm mna~oods zozao ~u!u!emoa [eug e ~u.zplo!,( so!ouonboaj i[~ ut
,([[t~uo!l.zodoad lo~ smq).uO~le ~o~d.uosap oo.zoj 'JOAO,~OH 'so!ouonboaj ~som ut, op£o puooos
oql u~ql ao~z~tosso I ql.z,~ od~zqs Lzed splo!£ o[o£o pa!ql o q i "o[o/~o puooos .z.zoq, ,~q s~!ouonboaj
lie ~oj u ! e m o p £~uonbo.zj oql u.z opnz.m~gm ~zo~.za odeqs z~ed oq) oonpoa suaqLuo$1¢
"Va£ (o) pu~ Va~La (q) 'VELLI(t~)~oJ u!emop ~ouonbo.U ,~q:l u.t ~o~aood~qs la~d "t'l "oI~I
(o)
zt:~tzn.fz~qxolq:~flo~zi.~i
OC 0 ~- 01 I
. . . . . 0"0
~3
- - ~,o III
....... I~D
.... t3
.... CO
.zoa.z~ od~qs ~tJ~cl
li~lJilliilliiliillllllJilJi
O'OZ
(q)
Zi3flRfZN k3NEl~'dd
0C OZ 0I OC 02 0I
0"0 o'o
0'9 0"9
.... gG
0'01 o'oi
' = I(I "--" Ifl
.... ~O ttJ .... gV
.... IO
- IO O'g~ o'gl
(rata)
Jo.z,ze edeqs "~.,md ,zo~aa ad~t.lS la~d
, i . . . . . i J I I • I * * A , , j : . . . . . . .
O'OZ lJlJlitil,lJ*iJ,lJ~l|i*J,J,iJ o'0~
6EI Su.ttmoj legato ~aaqs u.t u~!sap Su.qoo£
130 A.P. KARAF1LLISand M. C. BOYCE

The success of the force descriptor algorithm utilizing finite element analysis also
provides the motivation for an experimental method which would measure interface
tractions. The interface traction distribution could be extracted by performing a real
forming experiment instead of simulating one by using FEA. An elastic FEA could then be
conducted applying the experimentally obtained traction distributions to the desired part
shape. The elastic simulation is simple, fast and accurate. The deformed part shape could be
used as the next tooling shape and the algorithm continued depending on required
accuracy, where one iteration may be all that is required. The experiments for traction
extraction are suggested as an alternative to the forming simulation due to the still
remaining inherent assumptions and difficulties associated with numerical analysis of
forming, i.e. contact algorithm and friction law accuracy, material anisotropy effects, etc.
We recognize that traction distribution measurement in metal forming is a very difficult
task. However, progress in this field has been made recently by Fayfield et al. [26], and such
measurements can be performed. Such a technique would provide a valuable and systematic
method for tooling design.

CONCLUSIONS

Springback studies for the sheet metal forming process using FEA were conducted for
two materials covering a range of steel strength, one low-yield and one high-yield. The
results were utilized for tool design regarding elimination of the part shape error. The
analyses conducted can be outlined in the following conclusions:
(1) Springback is a result of release of the traction distribution in thc fully loaded part. It
is higher for the high-yield material, insensitive to friction for the cases considered
here, and results in a final configuration including a part shape deviation and a
residual stress distribution.
(2) Information on traction distribution incurred in the forming process can be used to
design the die shape which yields the desired part shape. In this procedurc inverse
elastic simulation of the material unloading is implemented. We caUed this method
the force descriptor method as opposed to the already established shape descriptor
methods.
(3) Force dcscriptor algorithms converge very fast and with rcmarkable accuracy to the
desired shape. Comparison of force descriptor algorithms with known shape descrip-
tion algorithms reveals the effectiveness of the force descriptor method.
(4) Extraction of forming tractions provides useful knowledge about springback in sheet
metal forming process. The traction distribution in the fully loaded state determines
the amount of the springback and therefore complete understanding of springback for
individual part geometries requires knowledge of the associated tractions. Informa-
tion on traction distribution can be uscd as a powerful tool to modify tooling shape in
order to eliminate part shape error due to the springback of the part.
(5) Given accuratc constitutional laws for matcrials and interfaces, the finite element
analysis approach to tooling design as given in this paper can be used to replace
current experimental trial and error die design procedures. Indced, with only
adequate material description, our traction based approach can be used as a tool to
obtain an excellent preliminary die and thus eliminate several hardware steps in
existing procedures.

Acknowledgements--This research has been supported by the National Science Foundation, Manufacturing
systems program, under Grant No. 8714686-DMC. Computations were conducted on a SUN 4/60 purchased with
MIT Bradley Foundation funds.

REFERENCES
1. F. J. GARDINER, The spring back of metals. Trans. ASME 79, 1 (1958).
2. J. INGVARSON,Residual stresses. Effect on buckling. Proc. 3rd Int. Spec. Conf. on Cold Formed Steel Struct.,
p. 85 (1975).
3. W. JOHNSON and T. X. Yu, Springback after the biaxiai elastic-plastic pure bending of a rectangular plate. Int.
J. Mech. Sci. 23, 619 (1981).
Tooling design in sheet metal forming 131

4. W. JOHNSON and T. X. Yu, On the range of applicability of results for the springback of an elastic/perfectly
plastic rectangular plate after subjecting it to biaxial pure bending. Int. J. Mech. Sci. 23, 631 (1981).
5. S. C. TANG, Analysis of springback in sheet forming operation. Adv. Tech. Plastic. 1, 193 (1987).
6. X. Ju, J. C. GERDEEN and E. G. BERAK, Buckling of imperfect panels under temperature variation. ASME
Press. Vessels Piping Conf. 149, 7 (1988).
7. W. Y. D. YUEN, Springback in the stretch-bending of sheet metal with non-uniform variation. J. Mater. Proc.
Tech. 22, 1 (1990).
8. Y. TOZAWA,Forming technology for raising the accuracy of sheet-formed products. J. Mater. Proc. Tech. 22,
343 (1990).
9. Y. UMEHARA,Technologies for the more precise press-forming of automobile parts. J. Mater. Proc. Tech. 22,
239 (1990).
10. R. A. AVRES, SHAPESET: a process to reduce sidewall curl springhack in high-strength steels rails. J. appl.
Metalworking 3, 127 (1984).
11. Y. C. L~u, The effect of restraining force on shape deviations in flanged channels. Trans. ASME, J. Engng
Mater. Technol. 110, 389 (1988).
12. D.E. HARDT and R. C. FENN, Real-time sheet forming stability control. Proc. 16th Biennial IDDRG Congress,
Borlange, Sweden (1990).
13. H. B. SIM and M. C. BOYCE,Finite element analyses of real-time stability control in sheet forming processes.
Trans. ASME, J. Engng Mater. Teehnol (in press).
14. K. A. STELSON and D. C. GOSSARD, An adaptive pressbrake control using an elastic-plastic material model.
Trans. ASME, J. Engng Ind. 104, 289 (1982).
15. D. E. HAROr, M. A. ROaERTS, and K. A. STELSON, Material adaptive control of sheet metal roll bending.
J. Dyn. Syst. Meas. Control 104 (1982).
16. D. E. HAROT and M. HALE, Closed-loop control of a roll-straightening process. Ann. CIRP 33 (1984).
17. D. E. HARDT, T. JENNE, M. DOMROESE and R. FARRA, Real-time control of twist deformation processes. Ann.
CIRP (1987).
18. R. D. WEan, Spatial frequency based closed-loop control in sheet metal forming. Ph.D. thesis, M.I.T. (1987).
19. R.D. WEan and D. E. HAROT, A transfer function description of sheet metal forming for process control. Trans.
ASME, J. Engng Ind. 113, 44 (1991).
20. S. KIM and K. A. STELSON, Finite element modelling for the analysis of a material property identification
algorithm for pressbrake bending. Trans. ASME, J. Engng Ind. 110, 218 (1988).
21. N. REaELLO, J. C. NAGTEGAALand H. D. HlaalTT, Finite element analysis of sheet forming processes. Int. J.
Numer. Meth. Engng 30, 1739 (1990).
22. S. KOaAYASHI,S. OH and T. ALTAN, Metal Formin0 and the Finite Element Methods. Oxford University Press,
New York (1989).
23. ABAQUS, User's Manual, Theory Manual, version 4.7. Hibbitt, Karlsson and Sorensen, Providence, Rhode
Island (1988).
24. D. DowsoN, History of Tribology. Longman, London (1979).
25. D. E. HARDT and R. D. WEan, Sheet metal die forming using closed-loop shape control. Ann. CIRP (1982).
26. T. FAYF1ELD, V. SCOTT, W. P. RoamNS, S. RAMALINGHAM,B. KLAMECKI and M. Ll, Piezoelectric thin film
stress sensors for metal forming operations. Proc. 1989 IEEE Ultrasonics Symposium, Montreal, October
(1989).

APPENDIX
The effectiveness of each algorithm regarding error elimination and CPU time spent, was measured with two
performance indices (PI 1) and (PI2). (PI.0 is equal to the ratio of the total reduction of the RMS value of the initial
part error distribution in parts produced from the 100% tooling with respect to the total required CPU time used
in each algorithm. Reduction of RMS of shape error is normalized with respect to the RMS of the shape error of
the part produced from tooling set at 100% of the desired shape, while CPU time was also normalized with respect
to the CPU time spent on the simulation of forming with the 100% tooling. Therefore
(NRRMS)
(PIl) (NRCPUT) ' (A1)

where NRRMS is the normalized reduction of RMS of part error and NRCPUT is the normalized required CPU
time.
Similarly
(NRW)
(PI2) (A2)
(NRCPUT)'
where (NRW) is the normalized reduction of width of part error.
Values of performance indices are tabulated in Table 1. It can be seen that, when CPU time is considered, ETEA
seems to be the most effective algorithm as it provides the maximum value of decrease of part shape error per CPU
second. However the final shape error when ETEA is used is higher than in the other two algorithms.

MS 34:2-D

You might also like