Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407

j our nal homepage: www. el sevi er . com/ l ocat e/ j mat pr ot ec


Development of an analytical model for warm
deep drawing of aluminum alloys
Hong Seok Kim
b,
, Muammer Koc
a
, Jun Ni
b
a
NSF I/UCRC Center for Precision Forming (CPF), Department of Mechanical Engineering,
Virginia Commonwealth University (VCU), Richmond, VA 23284, USA
b
Department of Mechanical Engineering, University of Michigan, Ann Arbor, MI 48109-2125, USA
a r t i c l e i n f o
Article history:
Received 31 March 2007
Received in revised form 9 June 2007
Accepted 15 June 2007
Keywords:
Warm forming
Lightweight material
Analytical model
Deep drawing
Finite element analysis
a b s t r a c t
In this study, an analytical model was developed to investigate the effects of material, pro-
cess, and geometric parameters in the warm forming of aluminum alloys under simple
cylindrical deepdrawing conditions. The model was validatedwithbothexisting experimen-
tal ndings in the literature and FEAresults. The effects of the main process parameters (i.e.,
temperature, forming rate, blank holder pressure (BHP), and friction between a blank and a
tooling element) on formability were studied under a variety of warm forming conditions.
The developed model offers rapid, useful, and reasonably accurate results for the design of
warm forming process by predicting the deformation mechanism of the material and the
relationships between limiting drawing ratio (LDR) and process parameters in isothermal
and non-isothermal heating conditions. It was demonstrated that signicant formability
improvement could be achieved when a large temperature gradient was realized between
die and punch, while a slight decrease of LDR was observed when tooling elements and a
blank were heated up to same temperature levels.
2007 Elsevier B.V. All rights reserved.
1. Introduction
The realization of lightweight structures for transportation
vehicles (aerospace and automotive) is a prominent way of
improving fuel efciency and reducing emissions. Because
of their low density, comparable strength, and stiffness,
lightweight materials such as aluminum and magnesium
alloys offer great potential in replacing mild steel struc-
tures to reduce weight. Other important factors in selecting
lightweight materials for engineering applications, compared
to plastics and polymer matrix composites, include their ease
of recycling, thermal properties, anddimensional stability and
manufacturability (Avedesian and Baker, 1999). However, cost-
effective forming of lightweight sheet materials into desired
functional complex shapes is extremely difcult with the con-

Corresponding author. Tel.: +1 734 763 7119.


E-mail addresses: hongseok@umich.edu, redstein@hotmail.com (H.S. Kim).
ventional forming technologies (i.e., stamping) because of the
formability limitations of these materials at cold conditions.
To improve the formability of lightweight sheet materi-
als, warm forming processes have been widely investigated
since the 1940s as an alternative manufacturing process. It
was reported that with warmforming of lightweight materials
(i.e., 200350

C), up to 300%total elongation could be achieved


(Shehata et al., 1978; Li and Ghosh, 2003; Ayres, 1979). More-
over, many preliminary studies proved a signicant increase
in formability with 5XXX and 6XXX series of aluminum and
AZ31, AZ61 magnesium alloys by performing forming tests at
elevated temperatures for simple deep drawing and rectangu-
lar cup forming models (Naka and Yoshida, 1999; Doege and
Droder, 2001; Bolt et al., 2001; Moonet al., 2001). Bolt et al. (2001)
conducted warm forming experiments on various aluminum
0924-0136/$ see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.jmatprotec.2007.06.046
394 j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407
Nomenclature
Lb time increment
C
b
linear work hardening coefcient
F
c
critical failure load
F
p
drawing load
i nodal point on the blank
j time step
K strain hardening coefcient
n strain hardening exponent
P blank holder pressure
r
e
radius of inner ange
r
p
, r
d
radii of punch and die cavity
r
0
, r initial and current radii of a blank element
R
0
, R initial and current radii of outer ange

R normal anisotropy parameter


t
e
, t
b
thickness of a blank element before and after
bending
t
f
, t
ub
thickness of a blank element before and after
unbending
t
0
, t initial and current thickness of a blank element
u punch speed
v radial velocity of a blank element
W strain work of an elementary volume
Greek letters

r
,

,
z
plastic strain in radial, circumferential, and
thickness directions

0
material constant

r
.

.
z
plastic strainrate inradial, circumferential, and
thickness directions

equivalent plastic strain rate


z
b
displacement of the neutral axis from central
axis
j coefcient of friction
,
c
, ,
n
radii of central and neutral axes
,
p
, ,
d
radii of punch and die proles
o
e
, o
b
drawing stress before and after bending
o
f
, o
ub
drawing stress before and after unbending
o
r
, o

, o
z
radial, circumferential, and thickness stress
o
z
thickness stress at , =,
d
o
z
(+), o
z
() thickness stress in the tensile and compres-
sive regions
o equivalent plastic stress
o
b
.
b
equivalent stress and strain during bending
o
e
.
e
equivalent stress and strain at the beginning of
bending
contact angle of the blank with the punch pro-
le
alloys (1050, 5754-Oand6016-T4) between100 and250

Cusing
box shaped and conical rectangular dies. They showed that
increasing temperature increased formability (higher draw-
ing ability and higher cup heights before fracture). It was
also demonstrated that keeping the punch cool would help
increasing the formability. Doege and Droder (2001) conducted
very comprehensive experimental work including aluminum
and magnesium alloys at different temperatures and forming
speeds. It was observed that for AZ31B alloy maximum LDR
was achievable between 175 and 210

C.
However, due to the complex nature of the warm forming
process including highly non-linear material behavior, con-
tinuously varying contact conditions, thermo-mechanically
coupled characteristics, and multi-faceted interactions
between material, process, and tooling factors, the design
of warm forming process greatly relies on the experience
of process engineers in addition to costly and lengthy
experimentations.
For cold forming conditions, comprehensive analytical
models have been developed by researchers considering
the relatively simple deep drawing cases. Chung and Swift
(1952a,b) analytically investigated and experimentally val-
idated the radial drawing process over a wide range of
operating conditions taking into account thickness changes,
bending and unbending effect, die prole friction, and tooling
geometry. Yamada (1961) adopted the incremental strain the-
ory based on a small strain formulation and proposed a nite
difference method to calculate the stresses and strains in the
ange region. Chang and Wang (1998) incorporated the effect
of friction, BHP, and radial thickness variation into their radial
drawing model, and developed separate radial drawing and
plastic bending analysis modules to systematically analyze
drawing and redrawing processes.
For warm forming conditions, however, very little ana-
lytical work has been carried out, and no full account has
been given due to the complex deformation conditions com-
pared to cold forming cases. Naka et al. (2000) developed
a simple analytical model for warm deep drawing using
temperature and strain rate dependent material properties
of an aluminum alloy without considering the thickness
change of the sheet and plastic bending effect. They reported
that the predicted drawing ability of cylindrical cups was
in good agreement with the corresponding experimental
results.
In recent decades, as an accurate and efcient analy-
sis and design tool, nite element analysis (FEA) techniques
have been increasingly used for the simulation of mate-
rial forming processes. The reliability of FEA for the
analysis of warm forming has been veried through var-
ious case studies in recent investigations. Takuda et al.
(2002) successfully predicted the improved drawing abil-
ity and the failure characteristics of aluminum alloy
using a 2D rigid-plastic and heat conduction nite ele-
ment method. Palaniswamy et al. (2004) compared the
experimental results on warm cup drawing of AZ31 mag-
nesium alloy with 2D and 3D FEA predictions. Similarly,
the authors of this paper previously presented the thermo-
mechanically coupled FEA results by comparing the predicted
part depth values and strain distribution with experiments
in various warm forming process conditions (Kim et al.,
2006).
In this study, in order to develop guidelines and extend
the fundamental understanding of warm forming process,
an attempt has been made to develop an analytical model
considering a simple circular cup part based on the previ-
ous investigations reported in the literature (Chung and Swift,
1952a; Yamada, 1961; Chang andWang, 1998; Naka et al., 2000).
All effects of tooling geometry, anisotropy, friction, and blank
j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407 395
holder pressure (BHP) were included in the model, and the
material parameters for hardening constants dependent on
temperature and strainrate were implemented to describe the
warm forming conditions. The accuracy and effectiveness of
the analytical model were veried through the comparisons
with the experimental results (Naka and Yoshida, 1999). A
nite element model was also developed for the same deep
drawing process conditions to compare with the analytical
model predictions. Thermo-mechanically coupled analyses
were performed both in isothermal and non-isothermal heat-
ing conditions and a static implicit solution procedure was
used to secure the reliability using a commercially available
FEA software called ABAQUS/Standard. In isothermal condi-
tion, stress and strain distribution, minimum wall thickness,
and LDR, under various warm forming processes, conditions
were evaluated using both analytical and FEA models. In
non-isothermal condition, the effect of temperature gradi-
ent between tooling elements on forming performance was
analyzed. Detailed failure characteristic and favorable tem-
perature condition for improved formability were explained
and suggested. In addition, the effect of forming speed (v),
friction (j), and BHP on formability was investigated. In the
next section, the derivation of the analytical model, assump-
tions, conditions, and solution procedure are explained in
detail considering deformation characteristics in ange, bend
(die radius), and cup wall regions in a typical axisymmetric
deep drawing case. In the third section, isothermal and non-
isothermal FEA models for the same conditions are described.
Comparison with the experimental ndings and discussion of
results are presented in Section 4.
2. Description of the analytical model for
an axisymmetric warm deep drawing case
Illustrated in Fig. 1 is the cylindrical deep drawing of a sheet
blank. A round blank with an initial radius of R
0
and thick-
ness of t
0
is clamped between a die and a blank holder. Then,
a at-faced punch moved down into the die cavity to form
an axisymmetric cup part. As shown in Fig. 1b, the defor-
mation process can be analyzed in ve distinct regions. The
regions from I to III successively undergo radial drawing (I)
(i.e., ange), plastic bending and unbending under tension
and frictional stresses (II), and side-wall stretching (III). The
tangential sites of a cup denoted by U and N in Fig. 1b are
the critical failure locations in isothermal (T
die
=T
punch
) and
non-isothermal (T
die
>T
punch
) process conditions, respectively
where the drawing stress reaches a maximum (Chung and
Swift, 1952b; Takuda et al., 2002). In the remaining regions,
the blank is subject to biaxial tension over the bottom of a cup
(V), and stretched on the punch prole radius combined with
a plastic bending (IV). Since the main tasks of this study are to
predict the LDR (i.e., the maximum ratio of blank diameter to
punchdiameter whichcanbe drawnintoacupwithout failure)
under warmforming conditions, and to establish the relation-
ship betweenthe LDRand process parameters, the derivations
are carried out from the edge of the ange to the beginning of
the punch prole region (i.e., regions IIII) in isothermal condi-
tions and to the end of the die prole region(i.e., regions III) in
non-isothermal conditions. Then, the maximum punch loads
obtained from the point U in isothermal cases and point N
Fig. 1 Deep drawing for a circular cup (a) initial stage and (b) intermediate stage.
396 j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407
in non-isothermal cases (Fig. 1b) are compared with the criti-
cal value from the instability criterion (Swift, 1952). When the
maximum load exceeds the critical value, the part is consid-
ered to fail.
Assuming that the material exhibits normal anisotropy (

R),
the stress and strain rate relations can be written using Hills
anisotropy yield criterion (Hill, 1950) as:


(1 +

R) o
=

r

R(o
r
o

) +(o
r
o
z
)
=

R(o

o
r
) +(o

o
z
)
=

z
(o
z
o

) +(o
z
o
r
)
(1)
o and

are dened by the following forms:
o =
1
_
1 +

R
_
(o
r
o
z
)
2
+(o

o
z
)
2
+

R(o
r
o

)
2
(2)

=
_
1 +

R
1 +2

R
_
(
r


R
z
)
2
+(



R
z
)
2
+

R(
r

)
2
(3)
The sheet material is assumed to follow a generalized
Swifts power-hardening law:
o = K(
0
+ )
n
(4)
In order to analyze forming performance at elevated tem-
peratures and different forming rates, the material constants
(i.e., K, n, and
0
) are made temperature and strain rate depen-
dent by tting the owstress curves measured at a wide range
of temperatures andstrainrates, andthe uniformstrainrate of
the material is assumed during forming. The equivalent strain
rate (

) of the material is estimated from the simple correla-


tion with the punch speed (u). If the plain strain condition
is assumed, the equivalent strain rate can be rewritten using
equation (3) as:

=
_
2(1 +

R)
_
1 +2

v
r

(5)
since

= (v,r). Here, v can be approximated as the following


form:
v
r
p
r
u (6)
By taking the average of radial velocities of elements at sev-
eral positions in the ange, the order of equivalent strain rate
can be nally calculated with respect to the punch speed.
2.1. Radial drawing
The stresses on an element of the ange are shown in Fig. 2. If
it is assumed that the blank holder force is distributed around
the edge of the ange that thickens most (Chung and Swift,
1952a), the thickness stress (o
z
) can be neglected. Then, the
radial equilibrium equation reduces to:
d(to
r
)
dr
=
t
r
(o

o
r
) (7)
Fig. 2 Stresses on an element in ange.
Examining the geometry of deformation of an element at
two different stages, the following strain compatibility equa-
tion can be derived:
d

dr
=
1
r
[1 exp(


r
)] (8)
where
r
=ln(dr/dR) and

=ln(r/R).
Since the strain rates in each direction are expressed as:

r
=
dv
dr
.

=
v
r
and
z
=

t
t
(9)
Combining it with Eq. (1) results in the following relations:
dv
dr
=
(1 +

R)o
r


Ro

(1 +

R)o



Ro
r
v
r
(10)

t =
(o
r
+o

)
(1 +

R)o



Ro
r
vt
r
(11)
The detailed computational scheme accounting for the
effect of BHP is developed using the nite difference method
as a slight extensionof the incremental straintheory proposed
by Yamada (1961). Hence, the basic equations (7), (8), (10), and
(11) are rewritten into nite difference forms as follows:
(to
r
)
i.j
= (to
r
)
i1.j
+
t
i1.j
r
i1.j
(o

o
r
)
i1.j
Lr
i.j
(12)
(

)
i.j
= (

)
i1.j
+
1
r
i1.j
[1 exp(


r
)
i1.j
]Lr
i.j
(13)
v
i.j
= v
i1.j
+
_
(1 +

R)o
r


Ro

(1 +

R)o



Ro
r
_
i1.j
v
i1.j
r
i1.j
Lr
i.j
(14)
j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407 397
t
i.j
= t
i.j1

_
(o
r
+o

)
(1 +

R)o



Ro
r
_
i.j1
v
i.j1
t
i.j1
r
i.j1
Lb (15)
where Lr
i.j
= r
i.j
r
i1.j
.
When the solution is obtained up to the points (i, j 1) and
(i 1, j), the radial stress (o
r
), thickness (t), and circumferential
strain(

) canbe calculatedat the point (i, j) using the foregoing


equations. Then, the strains
z
, and
r
, are determined based
on the current thickness value and under the assumption of
volume consistency of the material as follows:

z
= ln(t,t
0
) and
r
= (

+
z
) (16)
The strain rate at the current position (i, j) is approximated
from the strain difference between two time steps as:

i.j
=

i.j

i.j1
Lb
(17)
By substituting the strain rate in each direction into
equation (3), and combining it with the strain-hardening char-
acteristic in equation (4), the equivalent stress ( o) and strain
( ) values are obtained. Then, using equation (2), the circum-
ferential stress (o

) can be nally determined.


The calculation is performed from the edge of the ange
to the beginning of the die prole region with appropriate
boundary conditions. Since the friction force on the ange is
assumed to apply only around the rim, the radial stress at this
point is expressed as:
(o
r
)
0.j
=
jP
t
0.j
(r
2
0.j
r
2
e
)
r
0.j
(18)
where r
0,j
and t
0,j
denote the current radius and thickness of
the edge of the ange respectively. In addition, since the radial
drawing stress (o
r
) is usually small when compared to the cir-
cumferential stress (o

) at the edge of the ange, the initial


strains at the rim can be determined by assigning appropri-
ate initial radial velocity to the rim (v
0.j
) and using stress and
strain rate relationship in equation (1) as:
(

)
0.j
= ln
_
r
0.j1
+v
0.j
Lb
R
0
_
.
(
r
)
0.j
=

R(
z
)
0.j
=

R
1 +

R
(

)
0.j
(19)
2.2. Plastic bending
As shown in Fig. 3, when the element in the ange reaches
the die prole region, it starts to deform by bending along
the curved line of the die radius. Due to the tensile stress
at both ends before bending, the radius of the neutral axis
(,
n
) becomes smaller than that of the central axis (,
c
), and
the thickness decreases depending on the ratio of the original
thickness to the die prole radius. For simplicity, the deforma-
tion in the elastic regime is neglected in this study.
Based on the stress distribution shown in Fig. 3b, the equi-
librium equation for forces in the thickness direction can be
Fig. 3 Schematic of plastic bending (a) deformation in bending and tension and (b) stresses on an element in bending.
398 j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407
derived as
do
z
d,

o
r
o
z
,
= 0 (20)
In the die corner region, since we assumed that the circum-
ferential strain is negligible (Chang and Wang, 1998), Eqs. (2)
and (3) can be rewritten as
o =
_
1 +2

R
1 +

R
|o
r
o
z
| (21)
d =
1 +

R
_
1 +2

ln
_
,
,
n
_

(22)
since
d
r
= ln
_
,
,
n
_
(23)
Strain hardening is assumed to be linear (Chung and Swift,
1952a), hence the relation between equivalent stress ( o
b
) and
strain (
b
) during bending is dened as
o
b
= o
e
+C
b
(
b

e
) (24)
Re-writing Eq. (20) using Eqs. (21), (22), and (24) yields
do
z
d,
=

1
,
1 +

R
_
1 +2

R
_
o
e
+C
b
1 +

R
_
1 +2

ln(,,,
n
)

(25)
Here, the constant C
b
can be determined by taking the
instantaneous strain-hardening rate at the current strain level
as:
C
b
= Kn(
0
+
e
)
n1
(26)
Integrating Eq. (25) in the tensile and compressive regions
respectively yields the thickness stresses as follows:
o
z
(+) =
1 +

R
_
1 +2

R
o
e
ln
_
,
,
d
+t
b
_
+
(1 +

R)
2
2(1 +2

R)
C
b
_
_
ln
,
d
+t
b
,
n
_
2

_
ln
,
,
n
_
2
_
(27)
o
z
() = o
z

1 +

R
_
1 +2

R
o
e
ln
_
,
,
d
_

(1 +

R)
2
2(1 +2

R)
C
b
_
_
ln
,
d
,
n
_
2

_
ln
,
,
n
_
2
_
(28)
By equating Eqs. (27) and(28) at , =,
n
, the following relation
is obtained:
2 ln
_
,
n
_
,
d
(,
d
+t
b
)
__
1+

R
_
1 +2

R
o
e
+
(1 +

R)
2
2(1 +2

R)
C
b
ln
,
d
+t
b
,
d
_
= o
z
(29)
Substituting for o
r
from Eqs. (27) and (28) into Eq. (21), and
rearranging and integrating it yields the drawing stress after
bending (o
b
)
o
b
=
1
t
b
_
,
n
ln
_
,
d
(,
d
+t
b
)
,
2
n
_
_
(1 +

R) o
e
_
1 +2

R
+C
b
(1 +

R)
2
2(1 +2

R)
ln
_
,
d
+t
b
,
d
_
_
+o
z
(,
n
,
d
)
_
(30)
On the other hand, the increase in drawing stress during
bending can be also calculated using the strain energy as fol-
lows:
o
b
= o
e
+
_
W
,
c
t
b

(31)
W = V
_
o d (32)
where V=, d d, per unit width. By substituting Eqs. (21), (22),
and (32) into Eq. (31), the drawing stress after bending (o
b
) can
be obtained as
o
b
= o
e
+
1 +

R
,
c
t
b
_
1 +2

R
_
,
2
d
2
_
o
e
+
(1 +

R)
2
_
1 +2

R
C
b
_
1 ln
,
d
,
n
_
_
ln
_
,
d
,
n
_
+
(,
d
+t
b
)
2
2
_
o
e

(1 +

R)
2
_
1 +2

R
C
b
_
1 ln
,
d
+t
b
,
n
_
_
ln
_
,
d
+t
b
,
n
_
+
(2,
2
n
,
2
d
(,
d
+t
b
)
2
)o
e
4
+
(1 +

R)C
b
8
_
1 +2

R
t
b
(t
b
+2,
d
)
_
(33)
The stress and thickness values after bending are obtained
by equating Eqs. (30) and (33), and using geometric constraints
in Fig. 3 given by
t
b
=
,
n
,
n
+z
b
t
e
and ,
n
,
d
= z
b
+
t
b
2
(34)
2.3. Die friction and unbending
When a sheet metal is sliding over the die prole radius, there
will be an increase of drawing stress due to the frictional sheer
stress depending on the coefcient of friction (j) and con-
tact pressure. If the changes in thickness are neglected and
the contact angle of the blank over the die prole radius is
approximated as 90

, the increased drawing stress at the ver-


tical wall (o
f
) can be calculated based on the force equilibrium
as follows:
o
f
= o
b
r
e
r
d
e
(,2)j
(35)
Although the unbending process is more complicated
than the bending process due to the complex strain history
through the thickness, it can be treated in the same way with
j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407 399
the bending since the work done by unbending is not very
substantial. Finally, the drawing force (F
p
) in isothermal con-
ditions (T
die
=T
punch
) can be found at the punch corner region
taking into account the clearance between the die cavity and
the punch as follows:
F
p
= 2r
p
t
ub
_
o
ub
r
d
r
p
_
(36)
The above equation is also used for the calculation of draw-
ing force at the die corner region in non-isothermal cases
(T
die
>T
punch
) since the thickness change of the sheet in the
cup wall region (region III in Fig. 1b) is not included in this
study.
2.4. Instability criterion
In order to determine the LDR, as a measure of formability in
deep drawing process, the maximum drawing force obtained
in the previous section needs to be evaluated whether it
exceeds the critical load or not at the critical failure site. To
describe the stress states more accurately at the punch corner
regionwhere failure occurs inisothermal conditions, the equi-
libriumequations are derived based on the simplied analysis
of an axisymmetric shell (Wan et al., 2001) as shown in Fig. 4.
In non-isothermal cases, the same derivations can be used
neglecting the thickness change of the cup wall region.
The normal equilibrium equation is derived as:
p =
t
r,
p
(o

p
sin +o
r
r

) (37)
where ,

p
= ,
p
+(t,2), r

=r +(t/2) sin.
Re-writing Eq. (37) at the critical failure location(r
f
) denoted
by U in Fig. 1, the thickness stress (o
z
=p) is determined as
o
z
=
t
r
f
,
p
(o

p
sin +o
r
r

f
) (38)
Fig. 4 Stresses acting on a shell element.
where r
f
and r

f
are dened as:
r
f
= r
p
,
p
(1 sin ) and r

f
= r
f
+
t
2
sin (39)
Under the assumption of plane strain condition (

=0), the
equivalent stress and strain are rewritten from Eqs. (2) and (3)
as
o =
_
1 +2

R(o
r
o

) =
_
1 +2

R
1 +

R
(o
r
o
z
) (40)
d =
1 +

R
_
1 +2

R
d
z
(41)
Substituting Eq. (38) into Eq. (40) and using Eq. (4) gives the
drawing stress
o
r
=
1
_
1 +2

R
1 +

R +(t,

p
sin ,r
f
,
p
)
1 +(t,

p
sin ,r
f
,
p
) +(tr

f
,r
f
,
p
)
K(
0
+ )
n
(42)
According to the instability criterion by Swift (1952), the
critical failure condition can be determined by the following
relation:
o

=
_
1 +2

R
1 +

R
o (43)
Then, by inserting Eq. (4) into (43), the equivalent plastic
strain at failure (
c
) is obtained:

c
= n
1 +

R
_
1 +2

0
(44)
The drawing force in the cup wall can be dened as
F = 2r
p
to
r
(45)
By substituting Eqs. (42) and (44) into the above equation
and approximating =90

, the critical punch force can be


determined as
F
c
=
2r
p
t
1 +

R
_
1 +

R
_
1 +2

R
_
n+1
1 +

R +(t,

p
,r
f
,
p
)
1 +(t,

p
,r
f
,
p
) +(tr

f
,r
f
,
p
)
Kn
n
(46)
where t = t
0
exp(n +(
_
1 +2

R,(1 +

R))
0
) based on Eqs. (16)
and (41).
Finally, by comparing the maximum force in Eq. (36) with
the critical loading condition in Eq. (46), the LDR can be deter-
mined as a measure of deep drawing ability. The effect of
friction, BHP, temperature, andstrainrate onLDRcanbe inves-
tigated by using different coefcients of friction, BHP values,
and material models obtained at a wide range of temperatures
and strain rates. The detailed calculation scheme is illustrated
in Fig. 5.
400 j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407
Fig. 5 Calculation scheme of the analytical model.
Fig. 6 Axisymmetric FE model for deep drawing of a circular cup.
j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407 401
3. Finite element model
FEA was carried out for the deep drawing process using an
implicit FEA package, ABAQUS/Standard, to investigate the
effects of process parameters, such as temperature, form-
ing speed, friction, and BHP, on forming performance and to
compare the predicted results with analytical and experimen-
tal investigations. As illustrated in Fig. 6, the FE model was
developed with a 2D continuum element (CAX4RT) based on
the axisymmetric geometry and loading condition assuming
isotropic material behavior. To further simplify the analysis,
heating and cooling devices for the tooling elements (i.e.,
die, blank holder, and punch) were not designed, and tool-
ing distortion due to temperature changes was ignored by
using rigid body constraints. Uniform temperature distribu-
tion was directly assigned on the tooling surfaces to describe
warmforming condition. The sheet was modeledwithve ele-
ments inthe thickness directiontoinclude non-linear bending
and frictional shear effects, and the element size was initially
0.3mm in the radial direction.
For the process parameters, the Coulomb friction coef-
cients in a range of 0.10.5 were prescribed between a blank
and tools, and uniform BHP of 110MPa was applied on the
top surface of a blank holder plate. The contact heat trans-
fer coefcient for non-isothermal simulations was assumed to
be uniform (1400W/m
2
K) regardless of the temperature and
pressure at the interface based on the study of Takuda et al.
(2002).
4. Results and discussion
The reliability of the developedanalytical model was validated
through the comparison with the experimental ndings avail-
able in the literature and FEA results. Using the analytical and
FEAmodels, the warmforming behavior of analuminumalloy,
Al5083, was investigated over a wide range of warm forming
process conditions.
First, in isothermal conditions, where blank and tooling
elements are heated up to the same temperature levels,
stress and strain distributions, minimum thickness values at
failure, and interactions between LDR and various process
parameters (i.e., temperature, BHP, punch speed, and friction)
were examined both analytically and numerically. Then, in
non-isothermal condition, where blank temperature changes
temporally and spatially depend on tooling temperatures, the
inuence of a temperature gradient between tooling elements
on warm forming behavior was evaluated. The comparison of
deformation mechanisms is also made between isothermal
and non-isothermal conditions to explain the differences in
forming performance.
The tooling dimension and material properties were deter-
mined based on the experimental set up by Naka and Yoshida
(1999). The ow stress of Al 5083 was measured in the ten-
sile tests at a wide range of temperatures (20, 80, 100,150,
200, and 250

C) and strain rates (5.5610


5
, 5.5610
4
,
5.5610
3
, 5.5610
2
, and 5.2810
1
) as illustrated in Fig. 7,
and used in the analytical and FEA models by tting the
curves up to the maximum tensile strength. The stressstrain
relationships at other temperatures and strain rates were
interpolated based on the given experimental measurements
in Fig. 7. A blank thickness was initially 1mm, and the
radius of at-bottomed punch and die cavity was 18 and
20mm, respectively. Both die and punch had a prole radius
of 4mm.
4.1. Isothermal conditions
Fig. 8 shows the stress and strain distributions of the ange
region obtained from the analytical and FEA models at a part
depth of 10mm. In most of the ange region, the predicted
andcalculatedvalues of stress andstrainare very close to each
other (i.e., <5%). However, some deviations in stress values can
be foundinthe inner ange regionof Fig. 8a. The reasonseems
that the plastic bending over the die prole region affects part
of the inner ange region in the FEA model, while it is not
reected in the analytical model.
As shown in Fig. 8a, the circumferential compressive stress
(o

) is greatest at the edge of the ange while the radial


stress (o
r
) is a minimum. As the blank material approaches
to the inner ange region, the radial stress increases and
the circumferential stress becomes less compressive. Since
the existence of a high circumferential compression in the
outer ange region causes wrinkling, the proper level of BHP
needs to be applied to prevent wrinkling during forming.
However, in warm forming, this compressive stress tends to
decrease due to the decreased ow stress level of the mate-
rial at elevated temperatures. Hence, it can be implied that
relatively smaller BHP values can be used in warm forming
processes to eliminate wrinkling when compared to room
temperature forming cases. In terms of the strain distribu-
tion (Fig. 8b), the overall ranges of the predicted values are not
very severe since the failure by localized thinning occurred
around the punch corner. The blank becomes thicker as the
blank radius increases due to the increasing circumferential
compression.
The minimum thickness values around the punch corner
are analytically calculated as one of the formability measure-
ments, and compared with the simulation results at various
warm forming process conditions. As shown in Fig. 9, in
general, the analytical results overestimates the minimum
thickness values since thickness decreases over the punch
prole radius and in the clearance between the die and the
punch are ignored to simplify the analysis. However, the
magnitude of errors is within 10% in all process conditions.
In addition, it should be noted that the minimum thick-
ness values are not very sensitive to the temperature and
forming speed levels in both analytical and FEA models. Max-
imum 8% of thickness variation is observed as temperature
changes from 20 to 250

C. Hence, it can be concluded that the


formability improvement cannot be expected by increasing
temperature of all tooling elements to the same level.
Fig. 10 shows the LDR values obtained from the analyti-
cal and FEA models at two different punch speed conditions.
To determine the LDR in simulations, blank radii were pro-
gressively increased from 32 to 45mm by 0.5mm until the
part failed. Then, the largest blank radius which can be drawn
into a cup without failure was used for the calculation of LDR
(i.e., the maximum ratio of blank diameter to punch diam-
402 j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407
Fig. 7 Stressstrain relationship of Al5083 at six different temperatures and strain rates (Naka and Yoshida, 1999).
eter which can be drawn into a cup without failure). In this
study, the failure is considered only in terms of necking. How-
ever, it should be noted that wrinkling is also one critical
failure in an actual forming process. The compressive stress
in the ange increases with increasing blank diameter and
its effect becomes prominent at room temperature due to
the higher material strength level compared to warm form-
ing cases. When wrinkling occurs, material ow into the die
cavity is restrained and the increased stretching effect in the
punch corner region leads to localized necking. At elevated
j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407 403
Fig. 8 Stress and strain distributions in ange at T=250

C. (a) Stress distribution and (b) strain distribution (BHP=1MPa;


=0.1; v = 10mm/min; DR=2.11).
temperatures, onthe other hand, frictioneffect becomes dom-
inant in restraining the blank movement of the ange because
wrinkling tendency is reduced due to the decreased material
strength.
As expected from the minimum thickness values in Fig. 9,
a slight decrease of LDR (10%) is observed in the temperature
range of 20250

Cunder isothermal condition (Fig. 10). Since a


similar trend has been obtained in the experimental study by
Sugamata et al. (1987), it can be conrmed that the isothermal
heating conditions are not favorable to increase the formabil-
ity in circular cup part forming cases. The prediction errors
between two models are 510%.
Figs. 11 and 12 illustrate the effects of BHP and friction. A
monotonic decrease of the LDR with increasing BHPs and fric-
tion coefcients is observed in both room and warm temper-
ature conditions. As BHP increases from 1 to 10MPa, about a
7%and a 12%decrease of the LDR values are observed in room
and warm temperature conditions, respectively. In the case
of the friction variation, the absolute decrease of the LDR are
26% at T=20

C and 32% at T=250

C with the varying friction


Fig. 9 Effect of temperature on cup wall thickness at different temperatures (BHP=1MPa; =0.1; v = 10mm/min; DR=2.11).
404 j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407
Fig. 10 Effect of temperature on LDR under different punch
speeds (BHP=1MPa; =0.1).
coefcient of 0.10.5. Therefore, it is recommended that the
friction and BHP should be kept as low as possible to achieve
increased formability. However, as mentioned earlier, a com-
promise is requiredwhendetermining these factors toprevent
other types of failure, such as wrinkling and surface defects.
Even though the analytical and FEA models developed in
this section for the analysis of warm forming in isothermal
conditions have not been directly compared with the exper-
imental results, their reliability can be reasonably deduced
from the previous investigation by the authors (Kim et al.,
2006, 2004), and the accurate validation results of the non-
isothermal FEA in the next section performed with the same
tooling geometry andmaterial model withthe isothermal FEA.
In addition, it is proved that the analytical model can provide
an efcient analysis tool for warmforming compared to costly
and lengthy experimental trials and FEA. Only less than 30s
is required to nish one calculation in the analytical model,
while it takes about 30min for one simulation in the case of
the FE model.
Fig. 11 Effect of BHP on LDR at different temperatures
(v = 10mm/min; =0.1).
Fig. 12 Effect of friction on LDR at different temperatures
(v = 10mm/min; BHP=1MPa).
4.2. Non-isothermal conditions
In order to apply the analytical model to the non-isothermal
forming cases, the blank temperature contacting with the
ange and punch prole radius was assumed to be the same
with the die and punch temperature. For the temperature of
the sheet between these two regions, the value in the middle
range of die and punch temperature is assigned to realize the
gradual temperature change of the blank. In the case of FEA,
the blank was initially set 25

C and allowed to heat up by heat


transfer from the tooling elements. The BHP was 1MPa, and a
temperature dependent friction coefcient between 0.05 and
0.25 for a temperature of 20 and 300

C was used based on the


experimental results by Naka et al. (2000).
In Fig. 13, the LDR values predicted from the non-
isothermal FEA and analytical models in this study are
compared with the experimental and analytical results by
Naka and Yoshida (1999) and Naka et al. (2000) at two different
temperatures and various forming rates. The LDR values pre-
dicted from the non-isothermal FEA model matches best with
the experimental measurements. The maximum prediction
error is less than 3% when the punch speed is 500mm/min.
The analytical model developed in this study also shows
accurate prediction results. The error is within 6%. In both
room and warm temperature condition, the LDR decreases
with increasing punch speed. However, the variation is more
sensitive in warm forming condition due to the increased
strain rate sensitivity. The remarkable formability improve-
ment can be achieved especially when larger temperature
gradient between die and punch (T
die
>T
punch
) is realized
together with the slow punch speed. In the rectangular cup
part forming analyses (Kim et al., 2006), similar results were
obtained. Hence, it can be concluded that the favorable heat-
ing mode for warmforming is to introduce a large temperature
gradient between the die and punch (i.e., cooled punch, heater
die, and blank holder).
This result can be also explained based on the analytical
model developed in this paper. The critical failure load (F
c
)
and the maximum drawing force (F
p
) in Eqs. (36) and (46),
j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407 405
Fig. 13 Comparison of LDR at various temperatures and
punch speeds (a) T
die
=25

C, T
punch
=25

C and (b)
T
die
=180

C, T
punch
=25

C.
whichdetermine the forming performance, canbe regardedas
functions of material parameters and tooling geometry. When
the forming temperature increases under the given tooling
geometry, the values of strength coefcient (K) and hardening
exponent (n) of Aluminum alloys generally decrease (Li and
Fig. 14 Material characteristic of Al5083 alloy at various
temperatures (strain rate=5.5610
3
s
1
).
Fig. 15 Effect of temperature on maximum drawing load
(BHP=1MPa; =0.1).
Ghosh, 2003). The Al5083 alloy used in the study also shows
the similar trend (Fig. 14). To see the combined effect of these
two factors, the maximum load at the failure site is calculated
withtemperature inFig. 15 using the Kandn values inFig. 14. It
is foundthat the maximumloadis signicantly reducedat ele-
vated temperature. Onthe other hand, whenthe punchis kept
cold at the room temperature level, the critical failure loads
are the same in isothermal and non-isothermal forming con-
ditions since there is no change in the material properties of
the sheet contacting with the punch corner region. Hence, the
formability can be improved in the heated die, blank holder,
and the cooled punch conditions by reducing the maximum
load at the failure site and keeping the critical failure load to
the higher level.
The deformation characteristic in non-isothermal condi-
tion is quite different from that of the isothermal forming
case. For an in-depth evaluation, the comparison of thick-
ness distributions is made between these two cases in Fig. 16.
Fig. 16 Comparison of thickness distribution between
isothermal and non-isothermal forming cases
(v = 10mm/min; BHP=MPa; =0.1).
406 j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407
Fig. 17 Deformed shaped and failure at isothermal and
non-isothermal conditions (a) T
die
=180

C, T
punch
=180

C
and (b) T
die
=180

C, T
punch
=25

C.
In isothermal condition, the thickness strains are mainly
concentrated around the critical punch corner region as
mentioned earlier. However, those from the non-isothermal
simulation are not very signicant at the same part depth
(10mm). In addition, it is noted that the edge of the ange
moved more into the die cavity in non-isothermal condi-
tion. Hence, it can be seen that relatively higher ductility of
the ange at elevated temperatures and increased material
strength around punch corner at a lower temperature helps
to increase formability by delaying the localized thinning. As
showninFig. 17, innon-isothermal condition, the failure even-
tually occurred around the die corner region as different with
the isothermal forming case, and the deeper part depth value
of 15mm can be achieved.
In summary, the prediction capability of the FEA and ana-
lytical models can be validated in this section through the
comparison with the experimental results reported in the lit-
erature (Naka and Yoshida, 1999). The detailed deformation
characteristic of the material and favorable heating condition
in warm forming can be successfully analyzed and suggested
with reasonably small prediction error with the experiments.
However, since the current analytical model does not fully
account for the complex strainhistory of the material andheat
transfer fromthe tooling elements tothe blank, further studies
on these issues are required to determine the optimal tem-
perature condition in each tooling region, hence, to maximize
the formability. Some discrepancy observed between experi-
ments, and FEA and analytical models can be also attributed
to the inaccurate material property denition and the incom-
plete assumption of process parameters. The stress and stain
curves obtained from the uniaxial tension tests have a limi-
tation in predicting the biaxial deformation characteristics of
the material. In addition, the friction coefcient and contact
heat transfer coefcient have not been fully evaluated as a
function of various process parameters (i.e., temperature and
contact pressure). Tooling distortion by temperature change
is also ignored. Therefore, for more reliable results, studies on
the accurate material modeling and many unknown contact
properties are necessary.
5. Conclusions and future work
For the purpose of providing guidelines and further extend-
ing basic understanding of warm forming process, analytical
and numerical models were developed in this study as rapid
and cost effective prediction tools. The dependence of warm
forming performance on temperature, punch speed, BHP, and
friction, identied as main factors inuencing the formability
signicantly, was investigated under various warm forming
process conditions. In summary, the followings can be con-
cluded:
(1) Ananalytical model was developed to evaluate deep draw-
ing process at elevated temperatures and under different
BHP andfrictionconditions using a temperature andstrain
rate dependent material model. The results of calculations
were shown to be in good agreement with the corre-
sponding FEA predictions and experimental results. The
required calculationtime to nishone calculationwas less
than 30s; hence, its cost effectiveness could be veried.
However, since the constant temperature and uniform
strain rate conditions were assumed for the analysis of
each distinct region (i.e., ange, die prole, and punch
prole), further developments integrating non-isothermal
effects in the same deformation regions are required for a
wide range of application to industrial cases.
(2) A thermo-mechanically coupled FEA model was
developed using an implicit software package called
ABAQUS/Standard. The prediction error of the model
was found to be less than 3% based on the comparison
of LDR with experimental measurements. The slight
deviations of predictions were mostly due to incomplete
material modeling and inaccurate assumption of contact
conditions between tooling and blank. For more reliable
results, accurate stress-strain relationship under various
j ournal of materi als processi ng technology 1 9 7 ( 2 0 0 8 ) 393407 407
loading conditions and complete experimental data on
the anisotropic behavior and yield locus of the material
are required. In addition, contact factors such as friction
and heat transfer coefcients need to be evaluated as a
function of temperature and contact pressure.
(3) LDR values were not very sensitive to forming tempera-
tures in isothermal condition, while a remarkable increase
of formability was observed when the punch was kept at
the roomtemperature level. Hence, it is concluded that the
formability of aluminum alloys can be enhanced by intro-
ducing appropriate a temperature gradient on the work
piece. Additional studies to nd out the optimal heating
condition of tooling elements are necessary to maximize
formability.
(4) Detailed deformation characteristics were compared
between isothermal and non-isothermal conditions. In
the former case, the critical failure location, where limit
strain developed, was the punch corner region. However,
inthe latter case, relatively uniformstraining andthinning
was observed at the same part depth. It seems that the
increased temperature of ange region delays the onset of
localized thinning and shifts the failure site to the die cor-
ner region due to the improved ductility of ange material
and the increased ow stress of punch corner region.
(5) BHP and friction showed signicant effects on forma-
bility. Lower BHP and friction coefcient were preferred
to achieve increased formability due to the decreased
restrain force of the material. However, these factors need
to be carefully determined in practical application to pre-
vent other types of failure such as wrinkling and surface
defects.
r e f e r e nce s
Avedesian, M.M., Baker, H., 1999. Magnesium and Magnesium
alloys, ASM Specialty Handbook. ASM International, Materials
Park, OH.
Ayres, R.A., 1979. Alloying aluminum with magnesium for
ductility at warm temperatures (25250

C). Metall. Trans. 10A,


849854.
Bolt, P.J., Lamboo, N.A.P.M., Rozier, P.J.C.M., 2001. Feasibility of
warm drawing of aluminium products. J. Mater. Process.
Technol. 115, 118121.
Chang, D.F., Wang, J.E., 1998. Analysis of drawredraw processes.
Int. J. Mech. Sci. 40, 793804.
Chung, S.Y., Swift, H.W., 1952a. Cup-drawing from a at blank.
Part II: Analytical investigation. In: Proceedings of the
Institution of Mechanical Engineers, vol. 165, pp. 211223.
Chung, S.Y., Swift, H.W., 1952b. Cup-drawing from a at blank.
Part I: Experimental investigation. In: Proceedings of the
Institution of Mechanical Engineers, vol. 165, pp. 199211.
Doege, E., Droder, K., 2001. Sheet metal forming of magnesium
wrought alloys-formability and process technology. J. Mater.
Process. Technol. 115, 1419.
Hill, R., 1950. The Mathematical Theory of Plasticity. Clarendon
Press, Oxford, UK.
Kim, H.S., Koc, M., Ni, J., 2004. Determination of appropriate
temperature distribution for warm forming of aluminum
alloys. Trans. NAMRI SME, 573580.
Kim, H.S., Koc, M., Ni, J., Ghosh, A., 2006. Finite Element modeling
and analysis of warm forming of aluminum alloysvalidation
through comparisons with experiments and determination of
a failure criterion. ASME J. Manuf. Sci. Eng. 128, 613621.
Li, D., Ghosh, A., 2003. Tensile deformation behavior of aluminum
alloys at warm forming temperatures. Mater. Sci. Eng. A 352,
279286.
Moon, Y.H., Kang, Y.K., Park, J.W., Gong, S.R., 2001. Tool
temperature control to increase the deep drawability of
aluminum 1050 sheet. Int. J. Mach. Tools Manuf. 41, 12831294.
Naka, T., Yoshida, F., 1999. Deep drawability of type 5083
aluminiummagnesium alloy sheet under various conditions
of temperature and forming speed. J. Mater. Process. Technol.
8990, 1923.
Naka, T., Hino, R., Yoshida, F., 2000. Deep drawability of 5083
AlMg alloy sheet at elevated temperature and its prediction.
Key Eng. Mater. 177180, 485490.
Palaniswamy, H., Ngaile, G., Altan, T., 2004. Finite element
simulation of magnesium alloy sheet forming at elevated
temperatures. J. Mater. Process. Technol. 146, 5260.
Shehata, F., Painter, M.J., Pearce, R., 1978. Warm forming of
aluminum/magnesium alloy sheet. J. Mech. Work. Technol. 2,
279290.
Sugamata, M., Kaneko, J., Usagawa, H., Suzuki, M., 1987. Effect of
forming temperature on deep drawability of aluminum alloy
sheets. Adv. Technol. Plast., 12751281.
Swift, H.W., 1952. Plastic instability under plane stress. J. Mech.
Phys. Solids 1, 118.
Takuda, H., Mori, K., Masuda, I., Abe, Y., Matsuo, M., 2002. Finite
element simulation of warm deep drawing of aluminum alloy
sheet when accounting for heat conduction. J. Mater. Process.
Technol. 120, 412418.
Wan, M., Yang, Y.Y., Li, S.B., 2001. Determination of fracture
criteria during the deep drawing of conical cups. J. Mater.
Process. Technol. 114, 109113.
Yamada, Y., Studies on formability of sheet metals, Report of the
Institute of Industrial Science, University of Tokyo 11, 1961.

You might also like