Modeling_and_Simulation_of_Stretch_Blow_Molding_of

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 46

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/229628335

Modeling and Simulation of Stretch Blow Molding of Polyethylene Terephthalate

Article in Polymer Engineering and Science · August 2004


DOI: 10.1002/pen.20142

CITATIONS READS
66 4,416

3 authors, including:

Francis Thibault Loong-Tak Lim


National Research Council Canada University of Guelph
152 PUBLICATIONS 1,165 CITATIONS 221 PUBLICATIONS 11,098 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Francis Thibault on 02 January 2018.

The user has requested enhancement of the downloaded file.


MODELING AND SIMULATION OF STRETCH BLOW MOLDING OF
POLYETHYLENE TEREPHTHALATE

X.-T. Pham*a , F. Thibault a, L-T. Lim b

a
, Industrial Materials Institute, NRC,

75, de Mortagne, Boucherville, Québec, Canada, J4B 6Y4

b
, Husky Injection Molding Systems

500 Queen Street South, Bolton, Ontario, Canada, L7E 5S5

Abstract
When polyethylene terephthalate (PET) is stretched, it exhibits strain-hardening properties, which are

temperature and strain-rate dependent. In this paper, two grades of PET are experimentally characterized

using biaxial tests. A visco-hyperelastic model is used to describe the stretching behavior for the polymer. A

biaxial characterization method is employed to determine the model parameters using a robust non-linear

curve-fitting program. This model can represent adequately well the stretching behavior of PET. Based on

this model, the membrane finite element formulation is developed to simulate the stretch blow molding

process. Two bottles of different designs, produced based on the single-stage injection blow molding process,

are used to validate the model. Good agreement on the bottle thickness profile is observed.

Keywords: PET, modeling, visco-hyperelastic, biaxial characterization, finite element, numerical simulation,

injection stretch blow molding.

*
Corresponding author, email: tan.pham@nrc.ca
INTRODUCTION
Due to its excellent properties, polyethylene terephthalate (PET) is used extensively in the food packaging

industry, particularly for the production of soft drink bottles using the injection stretch blow molding

process, i.e. injection molding of preform and blow molding of preform to form bottles. This can be done

in either a single-stage or a two-stage process. In the two-stage process, the preform which is obtained

from the injection stage and stored until required is heated up to the desired temperatures, then stretched

and blown with high pressure air to take the final shape in the mold. Alternatively, in the one-stage

process, the preform obtained directly from the injection phase is stretched and blown right away after a

short conditioning phase. A principle schematic diagram of these stages is shown in Fig.1.

PET undergoes strain-hardening when it is stretched. This provides a self-leveling effect on the stretching

preform, which is important for forming bottle of uniform wall thickness. It is also well known that this

behavior is temperature and strain-rate dependent. At any given strain, increasing the temperature reduces

the strain-hardening properties and vice versa. In contrast, at a fixed temperature, increasing the strain-

rate causes the polymer to strain-harden, above all at large strain values. Thus, to take advantage of the

strain-hardening properties of PET, the operating conditions during blow molding, as well as the bottle

and preform designs need to be considered in order to achieve bottles of desirable quality.

While the experimental method may be used to understand the forming of PET bottle, numerical

simulations offer an efficient and rapid tool to study the effects of temperature and strain-rate on the strain

hardening properties of the polymer during the stretch blow molding process. Moreover, by taking the

effect of contact and heat transfer into blow molding, numerical simulations provide a better

understanding on the deformation mechanism. As a result, the thickness of the final part can be better

predicted, which is important for process and preform/bottle optimization based on structural analysis.

In the literature, several attempts have been made for modeling such behaviors, either phenomenological

or physically based on the network concept. Phenomenological hyperelastic models such as Ogden and

2
Mooney-Rivlin represent reasonably well the strain hardening, but not the strain rate effect (1).

Christensen (2) proposed a viscoelastic model consisting of elastic and viscous components. The strain

rate tensor appeared in the viscous part of the formulation to consider the strain rate effect. However, this

model could not represent well the behaviors of PET either (1). Yang, et al. (3), in their study on rubber,

modified the Christensen model to propose a visco-hyperelastic model. In this model, the elastic part was

expanded to the Mooney-Rivlin-like model and the viscous part was extended with an expression of

invariants of the deformation tensor. This visco-hyperelastic model, called Christensen-Yang model in

this paper, represented quite well both the strain hardening and the strain rate effect of the rubber.

Schmidt, et al. (4) considered the flow of PET during molding as a viscoelastic fluid, whereas Wang, et

al. (5) employed a non-Newtonian creeping material model of the viscoplasticity. Billon, et al. (6,7) also

characterized amorphous PET above glass transition using a phenomenological viscoplastic model.

Unfortunately, the thermal dependence and even the strain-rate sensitivity were not accurately predicted

over the whole ranges of strain rate and temperature.

Buckley, et al. (8,9) proposed a visco-elasto-plastic model based on the network concept for the modeling

of amorphous PET. The energy associated with the deformation of the PET material consisted of two

parts: the molecular conformation change and the bond stretching. The first component was considered as

a spring element, whereas the second one was considered as a spring combined with a damping element

and a friction slipper connected in parallel as shown in Fig.2.

In the Buckley model, the energy of molecular conformation change is reversible and the principal

components of the stresses appearing in the first spring element are defined by the derivative of this part

of energy over the principal components of the Hencky deformation tensor. The deformation in the bond

stretching consists of the elastic and viscous parts. When the stress is still small, only the spring is

deformed. This is governed by a Hooke-like law, in which the Hencky deformation tensor is used here

instead and the stress is defined as in fluid mechanics. When the stress is over the threshold of the friction

3
slipper, the slippage starts in the slipper element and the viscous element is then activated. From this

moment, the viscoelastic deformation in the bond stretching is governed by a Maxwell-like law, in which

the deviatoric stress and deformation tensors are stated. Recently, Boyce, et al. (10,11) also used a model

that is very similar to the Buckley model for modeling PET. However, these models are complex and

required a well-defined characterization method to determine accurately their parameters.

BlowSim, a non-linear finite element software developed at IMI/NRC, allows to simulate the stretch blow

molding process from the heating, stretching and blowing of the preform to the cooling of the final part.

In this software, hyperelastic models such as Mooney-Rivlin and Ogden were employed for the

deformation of PET (12,13). In this paper, a Christensen-Yang-like model is studied for modeling the

PET behaviors. Two grades of PET fabricated by Voridian, CB12 and 9921W, are experimentally

characterized by the biaxial tests performed on the biaxial Bruckner tester at different constant strain rates

and different temperatures. A biaxial characterization method is developed to determine the parameters of

this model using a robust non-linear curve-fitting program based on the Levenberg-Marquardt algorithm.

The temperature dependence can be considered by using either the WLF function or the second order

polynomial. A non-linear finite element formulation of this phenomenological model is developed using

the membrane theory. This model is then implemented in BlowSim software for the stretch blow molding

process. Finally, the model is validated by the biaxial stretching test. The stress-strain curves obtained

from simulations are compared to experimental data. Furthermore two different types of bottles are used

to validate the model by comparing the final thickness of these bottles with numerical results.

BIAXIAL STRETCHER EXPERIMENTS

Experiment description

The Bruckner stretcher, shown in Fig. 3, is equipped with a special mechanism that allows the sample to

be clamped and then smoothly stretched in the sheet plane. The material is heated up to the desired

temperature before testing by hot air circulating in the conditioning chamber of the tester. The sample can

be stretched either simultaneously in the machine direction (MD) and the transverse direction (TD) or

4
sequentially, i.e. one after the other. The test can be performed at either constant velocity or constant

strain rate. The stretching forces versus time in both directions are recorded. Table 1 shows some

important characteristics of the Bruckner stretcher.

Stretching experiments

The stretching experiments of the PET materials were performed on the Bruckner stretcher with the

conditions indicated in Table 2. The dimensions of the samples are 85mm x 85mm x 1.5mm. The samples

were heated up to the target temperature and simultaneously stretched in MD and TD directions. To

ensure that the sample reached the target temperature, several tests with different heating times were

conducted. At 120 seconds further extension of heating time did not affect the stress-strain curve. This

duration was judged to be sufficient to equilibrate the sample to the test temperature. Three temperatures

close to the process window of PET blow molding were chosen for testing. At each temperature, the tests

were carried out at three constant strain rates to study the strain rate dependence of PET. A range from

10%/s to 200%/s was employed. To get average values, at least three runs were conducted for each test

condition.

Experimental results show that the mechanical behaviors of PET depend strongly on the temperature,

stretch ratio and strain rate. The stress curves can be better represented in 3D graphics in which the

Cauchy stress at each temperature is a surface function of stretch ratio and strain rate as shown in Fig 4.

The red triangle, green diamond and blue square symbols indicate the experimental Cauchy stress

obtained from the biaxial tests at 90oC, 105oC and 120oC, respectively. Each curve depicts the stress-

strain curve at a constant strain rate. The surface connecting the data points of the same symbol represents

the mechanical behaviors of CB12 at the corresponding temperature. Results show that the lower is the

temperature, the higher the Cauchy stress surface. At any given temperature, the strain hardening is

clearly observed at large deformation. The lower is the temperature, the more important this effect.

Increasing the strain rate also shifts the stress-strain curve upward. The same phenomena are also

observed for PET 9921W polymer.

5
MODELING

Visco-hyperelastic model

A Christensen-Yang-like model is adopted for modeling PET. The model consists of two parts: elastic

and viscous. The Mooney-Rivlin model is used for the elastic part and the viscous part has a form similar

to that of the Christensen-Yang model. If S, S e , Sv are the total, elastic and viscous second Piola-

Kirchhoff stress tensors, respectively, their relationship has the following form:

S  Se  Sv . (1)

When F and C ( T F)F denote the deformation gradient tensor and right Cauchy-Green deformation

tensor, the strain energy W  W (C) of the Mooney-Rivlin is defined as a function of C . For an isotropic

material it could be reduced to the following form (14):

W  W ( I1 , I 2 , I 3 ) , (2)

where I 1 , I 2 and I 3 are the three first invariants of C , and defined by

I1  tr(C) , (3)

I2 
1
2
 
(tr(C)) 2  tr (C 2 ) , (4)

I 3  det(C) . (5)
The stress of the elastic part is determined by

W I I I
Se  2  2{W1 1  W2 2  W3 3 } , (6)
C C C C
where

W
Wk  , with k  1,2,3. (7)
I k
Explicitly, the following form of the strain energy function is used:

6
W  M 1 ( I1 - 3)  M 2 ( I 2 - 3)  M 3 ( I1 - 3)( I 2 - 3) , (8)

where M 1 , M 2 , M 3 are the three first parameters of the model. On the other hand, the viscous part is

defined by

t  
t
S v   {M 4  M 5 I 2 ()} exp(  )E()d (9)
0
M6
 the strain rate tensor that can be
where M 4 , M 5 , M 6 are the other parameters of the model and E

obtained by

  1C
E   1 {( T F )F (T F)F } (10)
2 2
By assuming that the material is incompressible, after some transformations, we can get the constitutive

law as follows:

t  
t
S  2(W1  W2 I1 )δ  2W2C  pC1   {M 4  M 5 I 2 ()}exp(  )E()d (11)
0
M6

where C 1 , C and p are the Piola, right Cauchy-Green deformation tensors and the arbitrary hydrostatic

pressure, respectively.

For incompressible material, the Cauchy stress tensor σ can be calculated (15) by

σ  FS ( T F) . (12)
Finally, we can get the following relationship:

σ   pδ  2{M1  M 3 ( I 2  3)}c1  2{M 2  M 3 ( I1  3)}c


t  
t
(13)
 F(t ){ {M 4  M 5 I 2 ()} exp(  )E()d}(T F(t ))
0
M6

where c 1 , c and p are the Finger, left Cauchy-Green deformation tensors and the arbitrary hydrostatic

pressure, respectively.

7
Membrane formulation

In the stretch blow molding process, the thickness of bottles is relatively small in comparison to other

dimensions. It is reasonable to apply the membrane theory to study the deformation of PET during this

process. A membrane has a plane stress state (14). It means that the stress in the thickness of the

membrane vanishes. If we use the indices 3 for the thickness direction of the membrane, by using Eqs A3-

A5 of Appendix A, and from Eq 11, we have

t   23
t
1 1
23 2 0
S33  0  2(W1  W2 I1 )  2W223  p  {M  M I } exp(  ) d (14)
M 6 
4 5 2

where 3 is the stretch ratio in the thickness direction. By extracting p from Eq 14 and substituting p in

Eq 11, we get the constitutive law of the visco-hyperelastic model corresponding to the membrane theory

as follows:

t   23
t
S  [2(W1  W2 I1 )23 (t )  2W243 (t )  12 23 (t )  {M 4  M 5 I 2 }exp(  ) d]C1
0
M 6 
(15)
t 
t
 2(W1  W2 I1 )δ  2W2C   {M 4  M 5 I 2 }exp(  )Ed
0
M6

BIAXIAL CHARATERIZATION

Rheological analysis

To characterize the PET materials, the sample is simultaneously stretched with the same strain rate in

both MD and TD directions. The deformation of the sample can be represented by the deformation

gradient tensor as follows:

 0 0 

F 0  0 , (16)
 
 0 0 1 2 

8
where  is the stretch ratio in both MD and TD directions of the sheet plane. From Eq 16, based on the

continuum mechanics, the other deformation tensors are determined as pointed out in Appendix B. It is

noticed that the biaxial test is performed with a constant strain rate that is derived from the Hencky strain.

It means that


 k  constant , (17)

where k is the strain rate of the test. By considering principal components 1 and 3 of Eq 13, we can get

the following equation system:

σ11   p  f ( M i , I1 , I 2 , c11 and / or c111 , E11 ) (18)


1 
0   p  f ( M i , I1 , I 2 , c33 and / or c33 , E33 ) (19)
By using Eqs B1-B9 of Appendix B respectively in each component, Eqs 18 and 19 become

σ 11   p  g ' ( M i , , k ) (20)

0   p  g '' ( M i ,  , k ) (21)

Again, by substituting p from Eq 21 in Eq 20 we finally get the following equation of the biaxial test:

σ11  g (M i ,  , k ) . (22)
By developing Eq 22, the explicit form of the biaxial characterizing equation is obtained as follows:

9
σ11  6 M 36  (2 M 2  6 M 3 )4  (2 M 1  6 M 3 )2
1 1 1
 (2M 2  6M 3 )  (2M 1  6M 3 ) 4  6M 3 6
2
 
1
2kM 4 M 6 8 2kM 4 M 6  4 M 6 k
   
(1  4 M 6 k ) (1  4 M 6 k )
1
4
4
 2kM 5 M 6   2kM 5 M 6  M 6k
(23)
1
4kM 5 M 6 10 4kM 5 M 6  4 M 6 k
   
1  6M 6 k 1  6M 6 k
1 1
M 4 M 6 k 4 2 M 6 k M M k 2
 {   }  5 6 {8   M 6 k }
1  2M 6 k 1  6kM 6
1
2
 2 M 5 M 6 k{   2 M 6k
}

Parameter identification by non-linear regression

A non-linear curve-fitting program based on the Levenberg-Marquardt algorithm was developed in

MATLAB environment to determine the parameters of the visco-hyperelastic model by fitting the

experimental data obtained from the biaxial tests to Eq 23. When considering the strain rate k n as a

discrete variable, the characterizing equation for each temperature will have the following form:

σ11
Tm , k n
 g ( M i , , k n ) . (24)
First, the stress-strain curves at 100%/s and 200%/s of temperature 90oC were fitted altogether to Eq 24 to
o
determine the parameters M i90 C
which have to be common for all the curves in this temperature. Then

o
the fits were repeated for the stress-strain curves of temperatures 105oC and 120oC to get M i105 C
and

o
M i120 C . The parameters at a certain temperature between 90oC and 120oC can be obtained by

interpolation using either the WLF function or second order polynomials. Table 3 shows the coefficients

of PET model for CB12 using the second order polynomial.

10
Figures 5a and 5b depict the experimental data and the fitting curves of PET CB12 and PET 9921W,

respectively. Results show that the proposed model can represent very well the mechanical behaviors of

the PET materials. The strain hardening and the strain rate effect are intrinsically found in this model. In

these figures, the open symbols represent experimental data. The red triangle, green diamond and blue

square curves are the fitting curves at 90oC, 105oC and 120oC, respectively. It is observed that the model

fits better at large strain than at low strain. The stronger is the strain hardening effect, the more precisely

the model fitted the experimental data. Results show that the model fits very well the experimental data at

90oC, 105oC, but has some discrepancies at high temperatures (120oC) where the strain hardening is less

important.

To evaluate the precision of the temperature dependence model, the stress-strain curves obtained from the

interpolation are compared with the fitting curves. Figures 6a, 6b, 6c and 6d show the fitting and

interpolated curves of PET CB12 at 90oC and 105oC using the second order polynomials and WLF

functions, respectively. In these figures, the blue square and diamond curves are the fitting curves, and the

red triangles are the interpolated curves. Results show that the second order polynomials give a better

interpolation in the testing range, whereas the WLF functions are slightly less accurate. However, WLF

functions are smoother and give a better extrapolation due to their monotonic characteristics.

FINITE ELEMENT FORMULATION


The proposed model can be solved numerically by the finite element approximation, in which the

formulation is developed based on the virtual work principle and the incremental method.

Virtual work principle in Lagrangian variables

The equation of the virtual work principle described in Lagrangian variables has the following form (15-

17):

  E  {S} dV     u  {b} dV    u  {T} dA ,


V V A
(25)

11
where u , b ,  and T are the displacement vector, the unit volume force, the density and the unit

surface force respectively.

Incremental formulation

By applying the increment operator on Eq 25, we get directly the incremental formulation of the studied

problem (18-21)

{ ([ Bl1 ] [T Bnl ])[ D]([ Bl1 ]  [ Bnl ]){u n } [T G ][ Π][G ]{u n } }dV 
T

V
(26)
 [ Ψ]{T} dA   [ Ψ]{b} dV
T T

A V

where B l1 , B nl , Ψ , G are the transformation matrices, D the material stiffness matrix, Π the stress

stiffness matrix and u n the increment of the nodal displacement vector.

Material stiffness matrix

For the membrane formulation, the material stiffness matrix is defined by the following relationship:

S X   DXX DXY DXS  E X 


    
 SY    DYX DYY DYS   EY  , (27)
 S   D DSS   
 S   SX DSY  ES 
where

S X  S XX , (28a)

SY  SYY , (28b)

S S  S XY  SYX , (28c)

E X  E XX  12 (C XX  1) , (29a)

EY  EYY  12 (CYY  1) , (29b)

E S  2 E XY  C XY . (29c)

Because S and E are symmetric, so D is symmetric and determined as follows:

12
S X S S
D XX   XX  2 XX , (30a)
E X E XX C XX
S X S XX S
D XY    2 XX , (30b)
EY EYY CYY
S X S XX S
D XS    XX , (30c)
E S 2E XY C XY
SY SYY S
DYY    2 YY , (30d)
EY EYY CYY
SY SYY S
DYS    YY , (30e)
E S 2E XY C XY
S S S XY S
DSS    XY . (30f)
E S 2E XY C XY

VALIDATIONS
To validate the material model implemented in BlowSim software, a simple test of biaxial stretching was

numerically studied and compared with the experimental measurements. In addition, we validated the

proposed model using two bottles with different preform designs. This validation allowed us to study the

influence of material model on the container thickness prediction.

Numerical biaxial stretching of the square sample

By using BlowSim, several numerical biaxial stretching tests were preformed to verify whether the

implemented model predicts a stress response adequately close to experimental data. A mesh of a square

of 73mmx73mm as shown in Fig. 7 was used to simulate the biaxial tests performed on the biaxial

stretcher. This square was stretched on all four sides with constant strain rates that were considered as

boundary conditions for numerical simulations. Both the visco-hyperelastic and Ogden models were used

to perform numerical simulations with the conditions indicated in Table 4.

The stresses obtained from numerical simulations are compared to experimental stress-strain curves as

shown in Fig. 8. As shown, numerical results using the visco-hyperelastic model agree well with

13
experimental data. Both the strain rate effect and the strain hardening of PET are accurately predicted. On

the contrary, the Ogden model fits the experimental data less accurately, simply because the strain rate

effect is not taken into account in this model.

Description of injection stretch blow molding of PET bottles

A 22 g wide mouth spice jar (case 1) and a 32 g narrow neck shampoo bottle (case 2) were used to

validate the model as shown in Fig. 9. The bottles were produced in a single-stage process using Husky

IndexSB stretch blow molding machine (Husky Injection Molding Systems Ltd., ON, Canada). During

plastic injection, the dual-face turret clamped up to the stationary platen, forming the mold in which the

molten PET was injected as pointed out in Fig. 10. When the in-mold cooling phase was completed, the

index turret moved to the open position and flipped 180o. This transposed the molded preforms to the

take-off side of the turret, where the preforms continued to be cooled (inner side of preforms was still in

contact with the mold core). Meanwhile, the next shot of preforms was being made on the clamp up side

of the turret. At the end of the on-core cooling phase, with the forward opening stroke of neck ring, the

preforms on the take-off side were ejected to a robot, which transferred and positioned the preforms to the

mandrel located on an index table. The table rotated 60o at every one and half of the injection cycle time

interval, conveying the preforms through two conditioning stations and finally to the blow mold.

To optimize the bottle wall thickness distribution, an elliptical oven was used at the conditioning station

to heat up the neck support ledge area of the preform. The optimal temperature profiles for the two

preforms, as recorded by the Prism DS infrared camera (3.6 to 5 m wavelength range; FLIR Systems,

Inc., Portland, OR) were summarized in Fig. 11. At the blow station, the preforms were first stretched in

the axial direction by means of a stainless steel stretch rod. In order to prevent the stretched preform from

touching the stretch rod, a pre-blow could be imposed during the stretch rod movement. When the stretch

rod reached the maximum axial position (i.e., high blow position), the preforms were then inflated in the

hoop direction with high blow air to form the final shape. The blown bottles were then indexed to the

14
discharge station where the bottles were stripped off from the mandrel and discharged from the machine.

Table 5 summarizes the key process conditions for the blow molding process.

Numerical simulation of stretch blow molding of PET bottles

The stretch blow molding of PET bottles was numerically simulated by using BlowSim software. It is

noticed that the temperature profiles obtained from experiments as shown in Fig. 11 are the preform

surface temperature. It is supposed that the preform temperature through the thickness is approximately

close to the experimental surface temperature at the end of the conditioning phase. Figures 12a and 12b

summarize the operating conditions as the temperature profile, the rod movement and the blow air

pressure that were used for each case of validation respectively.

Figures 13 and 14 show the preform deformations in the two validation cases during stretching and

blowing into the mold. Four consecutive steps of deformation are presented. From the initial state,

preforms are simultaneously stretched and/or pre-blown. Then they are blown with high-pressure air to

take the final shape in the mold.

Figures 15 and 16 show the final thickness distributions of the spice jar and round shampoo bottle,

respectively. The thickness is measured at nine different locations from the bottom up to the neck for the

spice jar and at fourteen different positions for the shampoo bottle. For each position, several samples are

picked out randomly for each case to evaluate the thickness variation in the hoop direction. Both

numerical results predicted by the visco-hyperelastic and Ogden models are compared with experimental

data in terms of maximum, average and minimum thickness values. It can be observed from these figures

that the thickness predictions using the visco-hyperelastic material model agree quite well with the

experimental thickness measurements. Almost all thickness predictions are adequately close to

experimental results. Furthermore, the visco-hyperelastic model predicted thickness profiles much more

similar to the actual data as compared to the Ogden model. The strong thickness variation observed with

15
the Ogden model along the bottle can be explained by the inability of this model to represent the strain-

hardening effect at high strain rates.

CONCLUSION
A visco-hyperelasic model was developed and implemented in BlowSim software for the stretch blow

molding process. The biaxial characterization method was used to determine the parameters of the studied

model. The second order polynomial was used for temperature dependence. The validation with two

materials PET CB12 and PET 9921W was performed with the biaxial stretching test and the stretch blow

molding of two bottles by using both the visco-hyperelastic and Ogden models. The investigation

demonstrated that:

 The biaxial stretching test is a very good way to characterize PET materials used in stretch blow

molding process, because the biaxial deformation is the most important mode of deformation in

this process.

 The non-linear curve-fitting program based on the Levenberg-Marquardt algorithm is robust to fit

the experimental stress-strain curves of PET to the proposed model.

 Experimental results show that the strain-hardening properties of PET are strain-rate and

temperature dependent. The visco-hyperelastic model represents well these phenomena under

large strain rates and various temperatures.

 The proposed model passed well the biaxial stretching test. Very good agreements were found

between the experimental curves and numerical results.

 Numerical simulations of stretch blow molding process using the visco-hyperelastic model

predict much better the stress and the deformation of PET than using the Ogden model. However,

the thickness prediction could be improved when the temperature profile in the preform thickness

and the induced crystallization will be taken into account in future works. In Figs. 15 and 16, the

16
numerical results are slightly higher than the thickness of PET bottles. This can be explained by

the increased specific density of PET in final parts due to strain-induced crystallization.

The thickness distribution in the final product is one of the most important criteria in the production of

bottles. The preform design, the temperature profile, the heat transfer, the harmony between stretching

and blowing, the cooling phase etc. need to be optimized in order to get a good bottle wall thickness in the

final products. Numerical simulations offer an efficient tool to better understand the behavior of the

process before finalizing tooling for production.

ACKNOWLEDGEMENTS
We would like to gratefully thank Voridian Company for supplying PET materials for this project, to

Husky Injection Molding Systems Ltd., Canada for experimental data using in the validation. The works

of M-A. Rainville, P. Courtois, A. Malo and B. Lantcot are also gratefully acknowledged.

17
REFERENCES
1. X-T. Pham, Modelling and characterization of PET, Internal Report, Industrial Materials

Institute/NRC, (2002).

2. R. M. Christensen, Theory of viscoelasticity, an introduction, Academic Press, (1982).

3. L. M. Yang, V.P.W Shim, and C.T. Lim, International Journal of Impact Engineering, 24, 545-560

(2000).

4. F. M. Schmidt, J-F. Agassant, M. Bellet, and L. Desoutter, J. Non-Newtonian Fluid Mech., 64, 19-42

(1996).

5. S. Wang, A. Makinouchi, M. Okamoto, T. Kotaka, T. Tosa, K. Kidokoro, and T. Nakatawa, Simulation

of Materials Processing: Theory, Methods and Applications, Huetink and Baaijens (eds), 441-446 (1998).

6. E. Gorlier, J-F. Agassant, J. -M. Haudin, and N. Billon, Plastics, Rubber and Composites, 30(2), 48-55

(2001).

7. N. Billon, E. Gorlier, and J. M. Haudin, The 4th Int. ESAFORM Conf. On Material Forming, Liège

(23-25 April 2001), 391-394.

8. C. P. Buckley and D. C. Jones, Polymer, 36(17), 3301-3312 (1995).

9. C. P. Buckley, D. C. Jones, and D. P. Jones, Polymer, 37(12), 2403-2414 (1996).

10. M. C. Boyce, D. M. Park, and A. S. Argon, Mechanics of materials, 7, 15-33 (1988).

11. M. C. Boyce, S. Socrate, and P. G. Llana, Polymer, 41, 2183-2201 (2000).

12. R. DiRaddo, D. Laroche, R. Aubert and D.M. Gao, SPE ANTEC Tech. Papers, 850-856 (1997).

18
13. L. Martin, D. Stracovsky, D. Laroche, A. Bardetti, R. Ben-Yedder and R. DiRaddo, SPE ANTEC

Tech. Papers, 982-987 (1999)

14. A. J. M. Spencer, Continuum mechanics, Longman Inc., New York, (1980).

15. G. A. Holzapfel, Nonlinear solid mechanics: a continuum approach for engineering, John Wiley &

Sons, (2000).

16. A. E. Green & W. Zerna, Theoretical elasticity, Dover Publications Inc., (1968).

17. A. E. Green & J. E. Adkins, Large Elastic Deformations, Oxford Uni. Press, (1970).

18. K-J. Bathe, Finite Element Procedures, Prentice Hall, (1995).

19. O. C. Zenkiewicz, Robert L. Taylor and R. L. Taylor, Finite Element Method: Volume 1, The Basis,

Butterworth-Heinemann, (2000).

20. O. C. Zenkiewicz and R. L. Taylor, Finite Element Method: Volume 2, Solids Mechanics,

Butterworth-Heinemann, (2000).

21. J. N. Reddy, Introduction to the finite element, McGraw-Hill, (1993).

19
APPENDIX A: MEMBRANE DEFORMATION STATE
According to the membrane theory (14), if we use the indices 3 for the thickness direction of the

membrane and denote X , Y and x, y the coordinates in the plane, respectively in the reference and

deformed configurations, the deformation gradient tensor has the following form:

 FxX FxY 0
 
F   FyX FyY 0 , (A-1)
 0  3 
 0

where  3 is the stretch ratio in the thickness direction. If the material is incompressible, so det F  1, the

value of  3 can be determined by

1
3  . (A-2)
FxX FyY  FxY FyX
From Eq A-1, we can determine the right Cauchy-Green and Piola deformation tensors as follows:

 FxX2  FyX2 FxX FxY  FyX FyY 0


 
C ( T F)F   FxX FxY  FyX FyY FxY2  FyY2 0 , (A-3)
 0 0 23 

C 1  (F 1 )(T F 1 ) 
 ( FyY2  FxY2 )23  ( FyY FyX  FxY FxX )23 0 
 , (A-4)
   ( FyX FyY  FxX FxY )23 ( FyX  FxX ) 3
2 2 2
0 
 0 0 1 23 

The strain rate tensor can be obtained by

  1C
E   1 {( T F )F (T F)F } . (A-5)
2 2

20
APPENDIX B: DEFORMATION IN THE EQUI-BIAXIAL TEST
The deformation gradient tensor in the equi-biaxial test is determined by

 0 0 

F 0  0 , (B-1)
 
 0 0 1 2 
where  is the stretch ratio in each direction of the sheet plane. From Eq B-1, we can get

1  0 0
F  0 1  0 ,
1
(B-2)
 
 0 0 2 

2 0 0 
1  
c  F.( T F)   0 2
0 , (B-3)
0 0 1 4 

1 2 0 0
T 1 1  
c ( F ).( F )   0 1 2 0. (B-4)
 0 0 4 

Because F is symmetric in this case, we have
C  c 1 . (B-5)

4 0 0 
 
C2   0 4
0 , (B-6)
0 0 1 8 

Then the two first invariants of C can be obtained as follows:
I1  22  1 / 4 , (B-7)

2
I 2  4  . (B-8)
2
Finally, the strain rate tensor has the following form:

21
  
  0 0 
   0 
E 0  (B-9)
 2 
0 0  5
  

22
List of tables
Table 1. Characteristics of the Bruckner stretcher.

Table 2. Stretching experiments.

Table 3. Parameter coefficients of PET model for CB12 using the second order polynomial.

Table 4. Numerical biaxial stretching tests.

Table 5. Process information for blow molding.

23
List of figures
Figure 1. Principle schematics of the injection stretch blow molding process.

Figure 2. Schematics of the Buckley model.

Figure 3. Bruckner biaxial stretcher.

Figure 4. Stress-strain behaviors of PET CB12.

Figure 5. Experimental data and fitting curves of: a) PET CB12, b) PET 9921W.

Figure 6. CB12 interpolation curves.

Figure 7. Mesh of the square sample.

Figure 8. Validation of the visco-hyperelastic model with biaxial stretching tests.

Figure 9. Preforms and bottles used in the validation: a) spice jar, b) round shampoo bottle.

Figure 10. Top view of Husky IndexSB (Stretch Blow) single-stage machine.

Figure 11. Optimal temperature profile of preform before stretch blow molding: a) case 1, b) case 2.

Figure 12. Numerical stretch blow molding of PET bottles: a) case 1 b) case 2.

Figure 13. Stretch blow molding of the spice jar.

Figure 14. Stretch blow molding of the 15 oz round bottle.

Figure 15. Final thickness comparison of the spice jar.

Figure 16. Final thickness comparison of the round shampoo bottle.

24
(Pham, Table 1.)

Dimensions of sample (mm x mm) From 85x85 to 100x100


Sample thickness (mm) Up to 3
Testing temperature (oC) Up to 250
Strain rates (%/sec.) Up to 300
Stretch ratio (%) Up to 800
Sequences Sequential or Simultaneous

25
(Pham, Table 2.)
Type Biaxial
Sequence Simultaneous
Temperature (oC) 90 105 120
Heating time (sec.) 120 120 120
Strain rate (%/sec.) 10,50,100,200 10,50,100,200 10,50,100,200
Max. stretch ratio [3.0, 3.5] [3.5, 4.0] [5.0, 6.0]

26
(Pham, Table 3.)

Coefficients
Parameters A B C
M1 -7.9525E-05 -2.4418E-02 4.4030E+00
M2 -3.7289E-05 1.5616E-02 -1.3572E+00
M3 2.1043E-06 -6.1208E-04 4.3280E-02
M4 2.3517E-02 -5.8313E+00 3.6358E+02
M5 -8.7632E-05 2.0763E-02 -1.2313E+00
M6 0.0000E+00 0.0000E+00 1.8000E-03

27
(Pham, Table 4.)
Voridian CB12
PET and Ogden models
Temperature (oC) Strain rates (%/s)
90 100 and 200
105 100 and 200
120 100 and 200

28
(Pham, Table 5.)
Case 1 Case 2
Description Spice jar Round bottle
PET Resin Voridian CB12 Voridian 9921 W
Cycle time (s) 9 12
Stretch rod speed (mm/s) 750 1000
Pre-blow pressure (MPa) 0.0 0.26
High-blow pressure (MPa) 4.0 3.5
Pre-blow time (s) 0.0 0.02
High-blow time (s) 2.0 1.2
Stretch rod diameter (mm) 16.5 12

29
(Pham, Fig. 1)

30
(Pham, Fig. 2)

31
(Pham, Fig. 3)

32
(Pham, Fig. 4)

33
(Pham, Fig. 5)

a) b)

34
(Pham, Fig. 6)

a) 2nd order polynomial at 90oC b) 2nd order polynomial at 105oC

c) WLF function at 90oC d) WLF function at 105oC

35
(Pham, Fig. 7)

36
(Pham, Fig. 8)

37
(Pham, Fig. 9)

Preforms

Bottles

a) case 1 b) case 2

38
(Pham, Fig. 10)

39
(Pham, Fig. 11)

140
120

Temperature (C)
100
80
60
40
20
0
0 10 20 30 40 50 60 70
Axial Position (mm)

a)

140

120

100
Temperature (C)
80

60

40

20

0
0 20 40 60 80 100 120
Axial Position (mm)

b)

40
(Pham, Fig. 12)

250 4.5
4

Stretch rod displacement


200 3.5

Pressure (MPa)
3
150
2.5

(mm)
2
100
1.5

50 1
Stretching Rod
Pressure 0.5
0 0
0 1 2 3
Time (sec)

a)

b)

41
(Pham, Fig. 13)

Initial Complete stretch Blow 1 Final

0s 0.31s 0.33s, 0.345MPa 2.8s, 4.0MPa

42
(Pham, Fig. 14)

Initial Complete stretch Blow 1 Final

0s 0.29s, 0.26MPa 0.30s, 0.35MPa 2.8s, 3.5MPa

43
(Pham, Fig. 15)

1.0
VHE Max
VHE Average
0.9
VHE Min
Expimental
Ogden Max
0.8 Ogden Average
Ogden Min

0.7
Thickness (mm)

0.6

0.5

0.4

0.3

0.2

0.1

0.0
0 20 40 60 80 100 120
Position from the bottom (mm)

44
(Pham, Fig. 16)

1.2
VHE Max
VHE Average
VHE Min
Experimental
1
Ogden Max
Ogden Average
Ogden Min

0.8
Thickness (mm)

0.6

0.4

0.2

0
0 20 40 60 80 100 120 140 160
Position from the bottom (mm)

45

View publication stats

You might also like