You are on page 1of 10

6

th
Australasian Congress on Applied Mechanics, ACAM 6
12-15 December 2010, Perth, Australia

Bulge test of sheet metals using rubber as pressure carrying medium



Maziar Ramezani*, Zaidi Mohd Ripin

School of Mechanical Engineering, Universiti Sains Malaysia, 14300 Nibong Tebal, Penang, Malaysia

*Corresponding author. Email: mzramezani@gmail.com
Abstract: New dynamic bulge testing technique is developed in this paper which can be performed on
a split Hopkinson pressure bar (SHPB) system to evaluate the strain-rate dependent strength of
materials at different ranges of forming speed. The newly developed bulge forming technique uses a
rubber-pad as pressure carrying medium instead of hydraulic fluid. Natural rubber is used to bulge the
OFHC copper sheet. In the development of this method, analytical and finite element investigations
are performed followed by experiments at different strain rates. Hyper-elastic material model is used to
simulate the behavior of rubber. J ohnson-Cook material model is used for the copper sheet to
approximate the mechanical behavior of this material under high strain rate deformation. Theoretical
analysis is based on conventional hydraulic bulge test principle and SHPB relations. To verify the
efficacy of the developed technique, analytical and finite element approaches are compared with
experiments and show good correlation with each other. The results clearly show that as the strain-
rate increases, the strength of materials increases. From the study a robust method to determine the
material behavior under biaxial deformation conditions has been developed.
Keywords: Bulge test, Hyper-elastic, J ohnson-Cook, Split Hopkinson pressure bar (SHPB), Strain rate.
1 Introduction
The biaxial bulge test is a material test for sheet metals to evaluate formability and determine the flow
stress diagram. Due to the biaxial state of stress induced in this test, the maximum achievable strain
before fracture is much larger than uniaxial tensile test. In this test, the sheet metal clamped at its
edges is stretched against circular die using fluid/viscous as a pressure carrying medium. When the
medium in the lower chamber is pressurized, the sheet is bulged into the cavity of the upper die. The
clamping force between the lower and upper die has to be high enough to prevent sliding of the sheet
to the die cavity. Thus, the sheet will only be stretched and no draw-in will occur. When the
deformation of the material exceeds its formability limit, the sheet will fracture.
The basic quasi-static bulge test was developed in the late 1940s to investigate the plasticity and
strength of sheet metals. Broomheas and Grieve [1] studied the effect of strain rate on the strain to
fracture of sheets using bulge test. They used a drop hammer rig makes use of a falling weight to
impact a punch which in terns applies a pressure loading to the fluid above the sheet material. They
could determine the forming limit diagram of low carbon steel for strain rates of up to 70
1
s . Atkinson
[2] provided accurate explicit formulations of the thickness variation and the effective radius of
curvature near the pole of the bulge. He showed that the representative stress-strain relationship
determined for the biaxial stress state at the pole is significantly affected by his corrections.
Gutscher et al. [3] used the viscous pressure bulge (VPB) test for determination of flow stress under
biaxial state of stress. They performed FEM simulations and experiments in order to study the
interrelationship of the geometric and material variables such as dome wall thinning, dome radius,
dome height, strain hardening index, material strength coefficient, and anisotropy. From their study a
robust method to determine the flow stress under biaxial deformation conditions using a viscous
material as pressure medium had been developed.
Grolleau et al. [4] developed a dynamic bulge testing technique to perform biaxial tests on metals at
high strain rates. They used SHPB apparatus with viscoelastic nylon bars and water as pressure-
carrying medium to perform dynamic bulge experiments on aluminum sheets. They analyzed the
experimental system and measurement accuracy in details and found that bars made of low
impedance materials must be used to achieve satisfactory pressure measurement accuracy.
The objective of this study is to develop a dynamic bulge testing technique which can be performed on
a conventional SHPB system. This method can be used for the analysis of materials and structures,
which are dynamically loaded and subjected to different levels of strain rates. The new method is
based on conventional hydraulic bulge test principle where rubber is used as the pressure medium

instead of hydraulic fluid. In the development of this method, an analytical investigation is performed
followed by numerical simulations using finite element code ABAQUS/Explicit. To verify the efficacy of
the developed technique, analytical and finite element approaches are compared with experiments.
2 SHPB apparatus
Split Hopkinson pressure bar (SHPB) has become a commonly accepted test method for strain rates
in the range of medium and high strain rates (
1 4 2
10 10

s ) [5]. It consists of two long slender bars
that sandwich a short cylindrical specimen between them (see Figure 1). By striking the end of a bar, a
compressive stress wave is generated that immediately begins to travel towards the specimen. Upon
arrival at the specimen, the wave partially reflects back towards the impact end. The remainder of the
wave transmits through the specimen and into the second bar. It is shown that the reflected and
transmitted waves are proportional to the specimens strain rate and stress, respectively. Specimen
strain can be determined by integrating the strain rate. By monitoring the strains in the bars, specimen
stress-strain properties can be calculated.

Figure 1: Typical compressive split Hopkinson pressure bar apparatus.


The current SHPB technique which is established by Kolsky [6] is based on one-dimensional wave
propagation analysis in pressure bars. The engineering stress, strain rate and strain, defined on the
specimen length, are obtained from:
)] ( ) ( ) ( [
. 2
.
) (
0
t t t
A
A E
t
T R I s
c c c o + + = (1)
)] ( ) ( ) ( [
) (
) (
0
0
t t t
L
C
t d
t d
R I T
s
c c c
c
+ = (2)
}
+ =
t
R I T s
dt t t t
L
C
t
0
0
0
)] ( ) ( ) ( [ ) ( c c c c (3)
where

E
C =
0
is the elastic stress wave speed in pressure bars, E and are Youngs modulus
and the density of pressure bars, respectively,
0
L and
0
A (
4
2
0
0
d
A
t
= ; where
0
d is the diameter of
the specimen) are the original length and cross-section area of the specimen and A is the cross-
section area of pressure bars, ) (t
I
c ; ) (t
R
c and ) (t
T
c are the recorded incident, reflected and
transmitted strain pulses with the time being shifted from the strain gage locations to the interfaces
between pressure bars and specimen according to the elastic wave speed in pressure bars.
Figure 2 shows a schematic of the dynamic bulge testing set-up. We considered a movable bulge
cell with natural rubber as pressure carrying medium which can be used to perform dynamic bulge
tests in a conventional SHPB system. When bulging with rubber, sealing problems and the possibility
of leakage of the high-pressure liquid employed in hydraulic bulging are eliminated. The need for the

filling and removal of fluid or the cleaning of the bulged specimen after forming is eliminated. The
insertion of the rubber is quick and convenient and the rubber can be re-used [7].

Figure 2: Schematic of the dynamic bulge testing set-up.
3 Theoretical analysis
3.1 Membrane theory
For a thin spherical shell expanded uniformly by internal pressure, the membrane stress is given very
closely by the approximation
d
d
t
pR
2
= o (4)
where p is the bulge pressure and
d
t ,
d
R are the thickness and radius at the top of the dome,
respectively. The equivalent strain can be calculated using the sheet thickness:
|
|
.
|

\
|
=
0
ln
t
t
d
c (5)
The radius at the top of the dome can be calculated by
d
d c d c
c
d
h
h R h R
d
R
2
2
2
2
2
+
|
|
.
|

\
|
+ |
.
|

\
|
= (6)
where
c
R is the radius of the fillet of the cavity,
c
d is the diameter of the cavity and
d
h is the dome
height.
Chakrabarty and Alexander [8] developed analytical methods to describe the deformation in the
hydraulic bulge test. They assumed that the shape of the bulge is spherical. With this assumption, the
thickness at the top of the dome can be calculated by the following equation:
n
c d
d
d h
t t

|
|
.
|

\
|
+
=
2
2
0
) / 2 ( 1
1
(7)
where n , is the strain hardening coefficient.
3.2 Dynamic bulge theory
The input bar of SHPB is used to measure the bulging pressure. The rubber pressure in the bulge cell
is determined directly from the incident strain pulse ) (t
I
c and the reflected strain pulse ) (t
R
c at the
input bar/rubber interface:
)] ( ) ( [ ) ( t t E t p
R I
c c + = (8)
The corresponding input bar/rubber interface velocity is:
)] ( ) ( [ ) (
0
t t C t u
R I in
c c + = (9)
The velocity of the bulge cell ) (t u
out
is determined from the transmitted wave ) (t
T
c

at the output
bar/die interface.
) ( ) (
0
t C t u
T out
c = (10)
The effective bulge velocity is the difference of interface velocities,

) ( ) ( ) ( t
out
u t
in
u t u = A (11)
The effective bulge displacement can be determined by integration of (11) as a function of the
measured strain histories:
( ) t t c t t c t c d d C t u
t
T R
t
I
) ( )] ( ) ( [ ) (
0
} }
+ + = A (12)
3.3 Mechanical behavior of sheet metal
The J ohnson-Cook [9] material model is used to analyze the mechanical behavior of OFHC copper.
This model is particularly suited to model high strain rate deformation of metals in adiabatic transient
dynamic analysis. It expresses the equivalent von-Mises flow stress as a function of the equivalent
plastic strain, strain rate, and temperature. In quasi-static conditions, metals work harden along the
well-known relationship which is known as parabolic hardening rule:
n
kc o o + =
0
(13)
where
0
o is the yield stress of the metal, n is work hardening exponent and k is the exponential
factor. Dynamic events often involve increases in temperature due to adiabatic heating and so the
thermal softening must be included in the constitutive model. The effect of temperature on the flow
stress can be described by following relation:
(
(

|
|
.
|

\
|

=
m
r m
r
r
T T
T T
1 o o (14)
where
m
T is the melting point,
r
T is a reference temperature at which reference stress,
r
o is
measured and m is material dependant constant. The strain rate effect can be simply expressed with
following relationship, which is very often observed at strain rates that are not too high.
c o ln (15)
Based on the above relations, J ohnson and Cook [9] presented the following equation for strength
model, where the von-Mises flow stress is given as:
] ) ( 1 )][ ln( 1 ][ ) ( [
*
0
m n
T C B A + + =
c
c
c o

(16)
where m n C B A , , , , are material constants which are experimentally determined; c is the effective
plastic strain rate and
0
c is the reference strain rate which can for convenience be made equal to 1
(
1
0
1

= s c ). The expression in the first set of brackets gives the stress as a function of strain for
1
0
=
|
|
.
|

\
|
c
c

and 0
*
= T . The expressions in the second and third sets of brackets represent the effects
of strain rate and temperature. The homologous temperature
*
T , is the ratio of current temperature T
to the melting temperature
m
T .
r m
r
T T
T T
T

=
*
(17)
where
r
T is the reference temperature at which
0
o is measured. The J ohnson-Cook constants for
OFHC copper are listed in Table 1.



Table 1: Material constants for OFHC copper sheet.
) (
3
m
kg

) (MPa A ) (MPa B C
n m
8960 90 292 0.025 0.31 1.09

3.4 Hyper-elasticity
The uniaxial compression stress-strain diagrams of natural rubber measured by SHPB apparatus at
different strain rates are illustrated in Figure 3. As can be seen from Figure 3, as the strain rate
increases, the strength of the material increases. Rubber-like materials have very little compressibility
compared to their shear flexibility and have nonlinear stressstrain characteristics for relatively large
deformations. Under such conditions, they are generally assumed as nearly incompressible. To model
these hyper-elastic materials through FEM, a constitutive law based on total strain energy density W
has to be adopted [10]. ABAQUS allows the automatically evaluation of hyper-elastic material behavior
by creating response curves using selected strain energy potentials. Among several approaches,
Ogden [11] theory is used based on the polynomial development of total strain energy. The Ogden
material model has previously been used with success to predict the behavior of rubber materials at
high strain rates (see e.g., Ramezani et al. [10]). The form of the Ogden strain energy potential is:
ij
ij
W
c
o
c
c
= (18)
i el
i
N
i i
i
N
i
J
D
W
i i i
2
1
3 2 1
2
1
) 1 (
1
) 3 (
2
+ + + =

= =
o o o

o

(19)
where W is the strain energy per unit of reference volume;

i
are the deviatoric principal stretches
which can be defined by
i i
J
3
1
= ;
i


are the principal stretches;
el
J is the elastic volume ratio;
and
i
,
i
o and
i
D

are temperature-dependent Ogden constants. Compressibility can be defined by
specifying nonzero values for
i
D , by setting the Poisson's ratio to a value less than 0.5, or by
providing test data that characterize the compressibility. We assumed a fully incompressible behavior
for natural rubber with 4997 . 0 = v and
i
D

equal to zero and so the second expression in (19) can be
eliminated. To determine the strain energy density W, ABAQUS uses a least-squares fitting algorithm
to evaluate the Ogden constants automatically from experimental data.

Figure 3: Stress-strain behavior of natural rubber under compression at different strain rates.

4 Experiments and finite element simulations
The bulge cell is composed of a thick-walled rubber chamber and a die (Figure 2). For experiments,
the round OFHC copper sheet specimen of thickness 1mm is clamped between the chamber and the
die. The input bar is inserted into the chamber which is filled with natural rubber to transmit the
pressure from the input bar to the sheet surface. The outer diameter of the SHPB pressure bars match
the inner diameter of the container and die. The pressure bars are made of silver steel and have
1500mm length and 12mm diameter. Several collars support the pressure bars, allowing it to slide
freely and to remove any bending waves due to an impact. When the striker bar impacts the input bar
at a defined velocity, a compressive stress wave is generated propagating towards the input
bar/rubber interface. This stress wave is transmitted through the rubber and ultimately causes the
bulging of the sheet specimen while both the bulge cell and the output bar are accelerated [12].
Throughout each experiment, the incident, reflected and transmitted waves are measured at the
center of the pressure bars by strain gages attached to the bars. The strain gage signals are recorded
using a HYLDE FE-H359-TA amplifier and a PICOSCOPE 3206, 200MHz digital oscilloscope. The
experimental set-up is shown in Figure 4.

Figure 4: Dynamic bulge cell set-up mounted on SHPB.
In order to simulate dynamic bulge forming, a finite element model is built in commercial software
ABAQUS. An explicit nonlinear approach with negligible temperature effects is assumed for
simulations. By taking advantage of axisymmetry, it is possible to simulate the process as 2-D
axisymmetric model. The die and rubber container are made of mild steel and are modeled as linear
elastic bodies with the Youngs modulus GPa E 206 = , Poissons ratio 3 . 0 = v , and mass density
3
m
kg
7800 = . Natural rubber is used to bulge an OFHC copper blank with diameter 60mm and
thickness of 1mm. One of the major requirements for computer simulations is the incorporation of
material properties through realistic models. Rubber is modeled as a hyper-elastic material, as
explained in Section 3.4. The J ohnson-Cook material model is used to simulate the behavior of metal
at high strain rates. The experiments were performed using two different striker bar velocities; i.e.,
s m V
st
/ 11 =

and s m V
st
/ 14 = . An empirical relationship was found between the striker bar velocity
and natural rubber strain rate at SHPB system, which is

0
2L
V
st
~ c (20)
where
st
V is the striker bar velocity and
0
L is the length of rubber. The initial length of the natural
rubber in the dynamic bulge test is 30mm and according to (20), the approximate strain rate of natural
rubber during the test will be
1
180

s and
1
230

s .
Since the model is developed by taking advantage of axisymmetry, the component nodes at the
symmetry edges are restrained in the appropriate directions. The end of output bar is also constrained
in order to model the momentum trap. The Coulomb friction model with the coefficient of friction of
0.25 is used to model the interface between rubber and sheet [7]. All other contact surfaces are
modeled as friction-free interfaces. The interactions between all components are modeled using an
automatic surface to surface contact algorithm. All geometric entities are modeled using CAX4R
elements. CAX4R is a 4-node bilinear axisymmetric quadrilateral, reduced integration, hourglass
control element. The pressure bars mesh comprises only one element row in the radial direction. The
pressure bars are made of silver steel and are modeled as linear elastic with the Youngs modulus of
GPa E 214 = and the mass density of
3
/ 7830 m kg = . The Poisson ratios are set to a non-
physical value of 0 = v . Thus, uniaxial waves in the computational model are not altered when
traveling along the bar axis [10].
The initial velocity is applied to the striker bar to impact the input bar. The simulation begins with the
input bar in contact with natural rubber. The copper blank is then introduced between the die and the
flexible rubber. After the striker bar hits the input bar, the impact stress pulse travels throughout the
input bar and upon arrival to input bar/rubber interface, a part of that reflects back to the input bar and
the rest of the stress pulse travels toward the rubber and makes the rubber to move down and bulge
the sheet. The axisymmetric finite element model of the high strain-rate rubber-pad forming at the end
of the process with s m V
st
/ 14 = is shown in Figure 5.

Figure 5: Axisymmetric finite element model of the bulge cell at the end of the process.
5 Results and discussions
Figure 6 shows the strain signals at the middle of pressure bars for experiments performed at striker
bar velocity of 14m/s. The incident strain pulse rises to its plateau strain level of about
5
10 194

during a time interval of about 19s. The time profile of the incident wave is of rectangular shape and
its amplitude remains quite constant during the simulation. The pulse shape of reflected wave is
triangular and reaches the maximum strain of
5
10 95

which happens at 15s. The reflected strain
signal becomes compressive after time duration of 34s. The transmitted strain pulse arrives at the

die/output bar interface after 39s and reaches the maximum amplitude of
5
10 120

at the end of
the simulation.

Figure 6: Representation of the incident, reflected and transmitted strain signals at striker velocity of
14m/s.
Figure 7 shows the impact pressure at input bar/rubber interface with impact velocity of 14m/s
evaluated according to (8). As can be seen from the figure, the pressure-time history shows an initial
peak at about 17s and then the pressure level increases monotonically until it reaches its maximum
level at 96s. The maximum pressure is about 142 and 173MPa respectively for the experiments
carried out at the striker velocities of 11 and 14m/s. subsequently, the pressure amplitude decreases
because of the end of the incident pulse. To compare the results of experiments and finite element
simulations, the pressure-time history at the input bar/rubber interface is monitored directly throughout
the simulations. The comparison of the curves demonstrates the good correlation between
experimental and simulation results. In general the simulation tends to predict slightly lower peak
pressure than the experimental and theoretical analysis. The maximum errors at the peak pressures
are 4.1% and 4.3% respectively at striker velocities of 11 and 14m/s.

Figure 7: Pressure-time history at input bar/rubber interface.
The strain history of the sheet during dynamic bulging process calculated using (5) is shown in Figure
8. According to Figure 8, the sheet deformation starts at about 40s after the input bar impacts the
rubber and reaches the maximum strain of 0.26 and 0.38 at striker velocities of 11 and 14m/s.
Combining Figures 7 and 8, we arrive at Figure 9 which shows the pressurestrain curve of OFHC
copper sheet at two different impact velocities during high strain-rate bulge test. As depicted in Figure

9, as the strain rate (impact velocity) increases, the formability of the material increases. In general, an
increase in the strain rate also increases the flow stress. However, when the strain rate is increased
significantly, only a small fraction of the work required to deform the material is stored; the remainder
is converted into heat. This results in adiabatic conditions, which lead to work softening [10].

Figure 8: Strain history of copper sheet during bulge forming.


Figure 9: Pressurestrain diagram of OFHC copper at different impact velocities.
6 Conclusions
In this paper, dynamic bulge test was investigated experimentally and numerically. The main
conclusions of this research are summarized below:
- The incident strain pulse is compressive and rectangular shape and it remains quite constant
during the simulation. The pulse shape of reflected wave is triangular in the tensile range.
- The pressure-time history at input bar/rubber interface shows an initial peak at about 17s and
then the pressure level increases monotonically until it reaches its maximum level of about
142 and 173MPa respectively for the striker velocities of 11 and 14m/s.
- Comparison of pressure-time curves obtained by experiments and finite element simulations
show good correlation between experimental and simulation results.
- As the strain rate increases, the strength of OFHC copper sheet increases.


References
1. P. Broomhead and R.J . Grieve, 1982, Effect of strain rate on the strain to fracture of a sheet steel
under biaxial tensile stress conditions, J Eng Mater Technol, 104(2), pp. 102106.
2. M. Atkinson, 1996, Accurate determination of biaxial stress-strain relationships from hydraulic bulging
tests of sheet metals, Int J Mech Sci, 39(7), pp. 761769.
3. G. Gutscher, H.C. Wu, G. Ngaile and T. Altan, 2004, Determination of flow stress for sheet metal
forming using the viscous pressure bulge (VPB) test, J Mater Process Technol, 146, pp. 17.
4. V. Grolleau, G. Gary and D. Mohr, 2008, Biaxial Testing of Sheet Materials at High Strain Rates Using
Viscoelastic Bars, Exp Mech, 48, pp. 293306.
5. M. Ramezani, Z.M. Ripin and R. Ahmad, 2010, Plastic bulging of sheet metals at high strain rates, Int
J Adv Manuf Technol, 48(9-12), pp. 847858.
6. H. Kolsky, 1949, An Investigation of the Mechanical Properties of Materials at Very High Rates of
Strain, Proc Roy Phys Soc, B 62, pp. 676700.
7. M. Ramezani, Z.M. Ripin and R. Ahmad, 2009, Computer aided modelling of friction in rubber-pad
forming process, J Mater Process Technol, 209, pp. 49254934.
8. J . Chakrabarty and J .M. Alexander, 1970, Hydrostatic bulging of circular diaphragms J Strain Anal
5(3), pp.155161.
9. G.R. J ohnson and W.H. Cook, 1983, A constitutive model and data for metals subjected to large
strains, high strain rates and high temperatures, Proceedings of the seventh international symposium
on ballistic, The Hague, The Netherlands, pp. 541547.
10. M. Ramezani, Z.M. Ripin, R. Ahmad, H.M. Akil and M. Damghani, 2010, High strain rate bulge forming
of sheet metals using a solid bulging medium, Proc IMechE B: J Eng Manuf, 224(2), pp. 257270.
11. R.W. Ogden, 1972, Large deformation isotropic elasticity on the correlation of theory and experiment
for incompressible rubberlike solids, Proc R Soc Lond, A 326, pp. 56584.
12. M. Ramezani and Z.M. Ripin, 2010, Combined experimental and numerical analysis of bulge test at
high strain rates using split Hopkinson pressure bar apparatus, J Mater Process Technol, 210, pp.
10611069.

You might also like