You are on page 1of 26

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

net/publication/242348667

Simulation of the plug-assisted thermoforming of polypropylene using a


large strain thermally coupled constitutive model

Article  in  Journal of Materials Processing Technology · September 2013


DOI: 10.1016/j.jmatprotec.2013.02.001

CITATIONS READS

46 1,829

6 authors, including:

C. P. J. O’Connor P.eter Martin


Flow Science Queen's University Belfast
17 PUBLICATIONS   100 CITATIONS    29 PUBLICATIONS   303 CITATIONS   

SEE PROFILE SEE PROFILE

John Sweeney G.H. Menary


University of Bradford Queen's University Belfast
81 PUBLICATIONS   2,153 CITATIONS    69 PUBLICATIONS   824 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

INDICATE - A Novel Instrument for Characterization of Preforms and Bottles for Injection Stretch Blow Moulding View project

Cycling Mobile Augmented Spaces View project

All content following this page was uploaded by C. P. J. O’Connor on 25 April 2020.

The user has requested enhancement of the downloaded file.


Simulation of the Plug-assisted Thermoforming of Polypropylene Using a Large Strain

Thermally Coupled Constitutive Model

C.P.J. O’Connor1*, P.J. Martin1, J. Sweeney2, G. Menary1, P. Caton-Rose2, and P.E. Spencer2
1 School of Mechanical and Aerospace Engineering, Queen’s University Belfast, Ashby Building, Stranmillis Road, Belfast, BT9 5AH,
Northern Ireland, UK.
2 School of Engineering, Design and Technology/IRC in Polymer Science and Technology, University of Bradford, Bradford BD7 1DP, UK.

* Corresponding author: Email: ciaran.oconnor@qub.ac.uk

Abstract

Thermoforming is widely employed in industry for the manufacture of lightweight, thin-walled products from pre-extruded

plastic sheet and its largest application is in packaging. Over many years attempts have been made to simulate the process

and thereby exploit modern computational tools for process optimisation. However, progress in this area has been greatly

hampered by insufficient knowledge of the response of polymer materials under thermoforming conditions and an inability to

measure this and other processing phenomena accurately. In recent years some address has been made to these problems

through advances in measurement technologies, and in particular, the development of high speed, high strain, biaxial testing

machines that are designed to replicate the conditions in thermoforming processes. In this work the development of an

advanced finite element-based thermoforming process simulation is presented. At its heart is a sophisticated large strain

thermally coupled (LSTC) material model for polypropylene, which has been developed after several years of research and is

founded directly on biaxial test results at elevated temperatures. This material model has been demonstrated to provide an

excellent fit to the biaxial data and to offer a very stable computational platform for the process simulation. The performance

of the working simulation was validated through comparison with matching experimental test results, and this enabled

investigation of the sensitivity of the process output (in the form of part wall thickness distribution) to changes in a range of

other processing parameters. This work confirmed that the process is most sensitive to the parameters controlling

plug/sheet contact friction. Heat transfer parameters were also shown to be significant and the requirement for the model to

be fully thermo-mechanically coupled has been clearly established.

Keywords: Thermoforming, simulation, polypropylene, constitutive model, plug assist, thermal coupling

Page 1
Introduction

The manufacture of packaging is a large and very important industrial sector. Within packaging a large proportion of the

market is concerned with the production of thin-walled polymer containers that are manufactured using industrial

thermoforming processes. As the name implies, thermoforming involves the shaping of a heated polymer material through

the application of pressure in an appropriate thermal environment. Most commonly the feedstock is a heated pre-extruded

polymer sheet and air pressure is the principal means of forming. However, as the required depth of draw in the product

increases it becomes necessary to employ a two stage thermoforming process, where the heated sheet is initially pre-

stretched through contact with a moving mechanical plug, before forming is completed through the application of air

pressure. This process is known as plug-assisted thermoforming and it is very widely used in the manufacture of food

packaging products such as pots, tubs and trays. In the past such products were almost exclusively manufactured from

amorphous polymers such as polystyrene (PS) and polyvinylchloride (PVC), as these can be shaped with relative ease

above their glass transition temperatures and also exhibit forgivingly wide forming temperature windows. However, in recent

times the same market has become dominated by polypropylene (PP) which has gradually replaced its amorphous rivals due

to its superior mechanical properties and lower material costs. As a semi-crystalline material, PP is more difficult to

thermoform as it must be processed within a narrow temperature range just below its crystalline melting point. This, along

with its greater tendency to sag when heated, has significantly increased the challenge facing the thermoformed packaging

industry. At the same time the packaging market has become increasingly demanding, particularly with respect to product

aesthetics, performance and costs, and the industry has responded through improvements in design and manufacturing

technologies. This has included many advances in the speed and efficiency of production machines, but despite the

widespread use of simulation elsewhere, the thermoforming industry still largely remains wedded to traditional trial and error

methods for product and process development.

Over many years researchers have made significant advances in the simulation of most of the major polymer

processing techniques, for instance injection moulding, and this has increased the level of sophistication of materials

processing technology used throughout the industry. However, the equivalent body of work in thermoforming is

comparatively small, and to date its industry has failed to capitalise on many of the benefits of process simulation seen in

other areas. Nevertheless, advances in process simulation have been steadily reported, although the majority of this work

has been focussed upon simple vacuum forming processes such as the preliminary models described by Bourgin et al.

(1995), where a heated sheet was formed by air pressure alone, the material was treated as a membrane and there was no

interaction with a moving plug. Later vacuum forming studies published by Wang et al. (1999) of an acrylonitrile butadiene

styrene (ABS) square cup have reported good correspondence between experiment and simulation. In most cases

Page 2
simulations have employed finite element techniques for modelling the process components and in early work the

assumption of isothermal conditions was commonplace. As a result, these simulations generally employed membrane

elements to represent the sheet during thermoforming and tooling surfaces were modelled as rigid surfaces as in the study of

Karamanou et al. (2006). The performance of the membrane assumption compared to solid continuum elements was

presented in the study by Nam et al. (2000). It was proposed that under conditions of free inflation that the membrane

assumption corresponded well with the three-dimensional solids when the sheet width to thickness ratio is larger than 100.

Experimental work by Collins et al. (2002) has presented clear evidence of significant changes in temperature during

thermoforming and has highlighted the need to include heat transfer in thermoforming process simulations. The

development of non-isothermal simulations have been reported by Laroche and Erchiqui (1999) who used a thermally

activated material model in the finite element analysis of a vacuum formed container and Christopherson et al. (2001) in a

thermally conducting computational fluid dynamic simulation of blister packaging. Use of more sophisticated element types

such as shell (Debbaut and Homerin, 1998) or continuum (solid) elements (O’Connor et al., 2007) have been reported in

thermoforming simulations to model the sheet material. The treatment of contact slip within thermoforming simulations has

also advanced substantially over the years. In early models it was common that no slip or absolute sticking conditions were

assumed, as is the case in the perfect plug slip assumption of Nam et al. (2000) and the perfect mould stick assumption of

Warby et al. (2003). Later models have included contact slip with the use of Coulomb friction functions and this has been

aided by efforts to experimentally measure friction either by using a moving sled technique (Collins et al., 2001) or a

rotational viscometer (Beilharz et al., 2007). Researchers such as McCool et al. (2006) have also recognised the need to

make frictional behaviour temperature dependent and have used non-isothermal friction models to improve their simulations.

Beilharz et al. (2007) proposed several experimentally based mathematical relations for frictional behaviour governed by

temperature, pressure and strain rate dependence. However, accurate measurement of friction during thermoforming is

recognised as an ensuing challenge for researchers and it remains as a serious impediment to the improvement of the

predictive accuracy of current simulations.

The most significant advance in the technology required for the simulation of thermoforming has come through the

development dedicated biaxial testing machines. For the first time these have been able to subject materials to testing

conditions precisely matching those experienced during thermoforming and have enabled the deformation response of

different polymer materials to be accurately mapped. Across the world a number of such machines are now being used by

research groups to create sophisticated and realistic models for different polymers used in thermoforming and blow moulding

applications. These include the dedicated biaxial testing machines at Oxford University (Buckley et al., 1996), the University

of Erlangen-Nuremberg (Capt et al., 2003) and Queen’s University Belfast (Martin et al., 2005). In early simulations

Newtonian or hyperelastic models were commonly employed (Nied et al., 1990), but with the combination of the need to

Page 3
introduce plug contact in many processes and the improved measurement and understanding of the materials, models are

now predominantly either viscoelastic (O’Connor et al., 2007) or viscoplastic (Wang et al., 1999). A number of very different

mathematical formulations have been proposed for different polymers as presented in the studies by O’Connor et al. (2007)

on PP, Buckley et al. (1996) on polyethylene terephthalate (PET), and Hummel and Nied (2004) on ABS. Researchers such

as Hosseini et al. (2006) have also recognised the need for these models to be thermo-mechanical and have introduced

mechanisms for thermal dependency of properties that is essential for modelling non-isothermal processes. In addition,

results from the biaxial testing of sheet polymers at high strain rates have suggested that internal or adiabatic heating during

stretching may raise the local sheet temperature during the process (Martin et al., 2005), and it has therefore been

suggested that material models need to be fully thermally coupled for accurate simulation of thermoforming (Makradi et al.,

2007). Despite the considerable research effort there is currently no approach to simulating thermoforming processes, which

has wide acceptance for its predictive capabilities and accuracy. Furthermore, thermoforming of PP presents particular

difficulties to the researcher because of the extreme temperature sensitivity of the material and the critical effect this has on

its mechanical behaviour.

The major aim of this work was the development of an effective and fully featured finite element simulation of the

plug-assisted thermoforming process for polypropylene. It was intended to base the simulation around a newly developed

constitutive model for polypropylene which was the result of several years of dedicated research involving extensive biaxial

testing of the polymer material. As well as providing accurate mapping of the deformation response of polypropylene, the

model was specifically configured to provide a fully coupled thermo-mechanical response and thereby ensure accurate

replication of all aspects of the thermoforming process. It was also planned that the simulation should permit appropriate

levels of heat transfer from both contacting and non-contacting surfaces, and slip across contacting surfaces. In the case of

the latter it was hoped that the development of the simulation would enable closer scrutiny of the importance of friction to the

overall process. The performance of the working simulation was to be validated through comparison with matching

experimental test results, and it was planned that this would enable investigation of the sensitivity of the process output (in

the form of part wall thickness distribution) to changes in a range of other processing parameters.

Product design specification and processing conditions

The simulation presented in this work was constructed specifically to replicate the manufacture of a typical industrially

thermoformed product. The cavity used to create the yoghurt container considered here is detailed in Fig. 1.

Page 4
Figure 1: Major dimensions (in mm) of the yoghurt pot cavity, provided in section view.
At the time of this study the receptacle was being manufactured from a single-layer polypropylene sheet using an

inline plug-assisted thermoforming process. The sheet material contained a 60%:40% blend of homopolymer (Basell Moplen

HP540J) and copolymer (Basell Moplen RP210G). The initial extruded sheet thickness was 1.23mm. The processing

conditions used in the manufacturing process were determined through consultation with an industrial collaborator. This

included process timings, forming speeds and the thermoforming temperature window. The latter was established using a

combination of laboratory tests, including Differential Scanning Calorimetry (DSC) and Dynamic Mechanical Thermal

Analysis (DMTA), and industrial practice. In industry direct sheet temperature measurement is not yet widely employed for

process monitoring and instead much of the industry still relies on long-established trial and error techniques during

production. In this case an acceptable processing window of 1555°C was determined, with a target temperature of 155°C.

The thermoforming process used to manufacture the part is illustrated in Fig. 2. This shows a representation of the

main elements of the production tool taken from the thermoforming simulation, which is described in detail later.

Page 5
Figure 2: Simulation of the plug-assisted thermoforming process.
In thermoforming a pre-extruded polymer sheet is heated, or re-heated, to its previously determined forming

temperature and it is then indexed into the open area of a two part split mould, the lower half of which is illustrated in 3-

dimesional section in Fig. 2(a) and the mesh structure in 2-dimensional form in Fig. 2(b). This shows a single cavity forming

tool, which in industry would only be a small part of a much larger multi-cavity production tool. In this case the complete

industrial tool contained 28 identical cavities. The two-part mould is then closed together and the sheet is clamped above

and below with a downholder built into the upper mould. In the first stage of the thermoforming operation a moving tool,

known as a plug, descends from the top of the two part split mould and pre-stretches the heated polymer sheet into the lower

mould cavity. Typically these plugs are made of temperature resistant polymer materials and can be actuated pneumatically,

hydraulically or with electrical servo motors to control their speed and displacement. For most packaging applications the

plugs are initially unheated but quickly rise to an equilibrium production temperature during repeated production cycles. The

production plug used in this case was made from an acetal homopolymer, polyoxymethylene (POM) which was chosen for its

high stiffness, low friction and low thermal conductivity. For this application the plug completes its downward stroke of 70mm

reaching to within 5.2mm from the centre of the base of the cavity. The plug stroke takes only 0.70 seconds to complete,

moving with an average speed of 100mm/s. At this point a forming air gauge pressure rising to 5bar is rapidly applied to the

plug side of the sheet, which stretches it and propels it downwards and outwards until contact is made with the cavity walls.

These are typically manufactured from an anodised aluminium alloy, and on contact the heated polymer sheet cools rapidly.

The mould temperature is generally controlled by chilled water, which is pumped through internal cooling channels. Once

Page 6
sufficiently cooled the upper mould is retracted, the product is trimmed and then it is ejected from the lower mould and

collected by a stacking unit. The entire thermoforming cycle for this product takes less than two seconds to complete.

Product thickness distribution

In the majority of industrial thermoforming processes accurate control of the product wall thickness distribution is the principal

measure of part quality. Close control ensures that parts have adequate mechanical performance but it also enables

processors to reduce material costs by minimising the starting gauge necessary in the extruded sheet. In this study the wall

thickness distributions of industrially manufactured parts were measured using a magnetic induction thickness gauge. The

thickness profile was recorded through 29 selected locations from the base centre of each part through to its upper lip and an

average taken from a total of 112 containers randomly selected from the overall output from across all 28 cavities in the

production tool. In the later results presented these measured distributions were compared with those from the simulation in

order to judge its predictive performance. A plot of the measured wall thickness distribution for the production container is

shown in Fig. 3. Like most thermoformed products of any kind the distribution is uneven with significant variations in

thickness from one location to another. In this case the thinnest part is around the bottom corner of the pot (<0.2mm), which

contrasts sharply with the much greater thickness at the base centre (>0.6mm). In much of the sidewall the material

distribution is quite well balanced at values slightly above and below 0.3mm thickness. Previous research has clearly

demonstrated the important role that the plug plays in determining these relative thicknesses (Martin and Duncan, 2007) and

in particular it has highlighted the critical role of friction during contact between the plug and the sheet (Collins et al., 2002).

With lower friction and therefore more slip, movement of material from the base towards the corner and sidewall of the pot is

promoted. However, this change must not be excessive as otherwise the imbalance can precipitate quality issues in the final

product caused by an in appropriate thickness distribution.

Page 7
Figure 3: Wall thickness measurements for the product, taken as a profile from the centre of the base of the
container along a vertical profile terminating at the flat lip of the receptacle.
Development of the finite element simulation

It is clear from figures Fig. 1 and Fig. 2 that the product has rotational symmetry about its centreline. This therefore allowed

a simpler two dimensional analysis to be employed in the development of the process simulation. Subsequently an

axisymmetric analysis model was constructed using the commercial finite element analysis [FEA] pre-processor software

Altair HyperMesh. The plug, sheet and mould cavity are clearly detailed in the finite element analysis of the forming process

shown in Fig. 2, as is the axis of symmetry for the process and product. An implicit FEA model was created to simulate the

plug-assisted thermoforming of the product using the industry standard analysis code Simulia ABAQUS/Standard. ABAQUS

has been widely employed for analogous applications in the past and it is regarded as one of the most effective software

tools for constructing manufacturing process simulations. The finite element mesh of the polymer sheet was constructed of

341 SAX2T axisymmetric thermo-mechanically coupled shell elements. The SAX2T is a 3-noded thin or thick axisymmetric

shell element which offers linear temperature variation in the shell surface. The simulation used five integration points

through the element thickness for the calculation of thermal conduction across the sheet. The SAX2T is part of a group of

general-purpose elements that typically provide robust and accurate solutions in many loading conditions for thin and thick

shell problems. It can deform through its thickness as a consequence of in-plane deformations, which is an obvious

requirement in this application. The SAX2T is not prone to transverse shear locking, nor does it have any unconstrained

hourglass modes. SAX2T elements are programmed to calculate large (finite) membrane strains and their sizes, locations

and densities were arrived at by assessing the mesh sensitivity to achieve a progressive convergence of analysis results.

Page 8
The mesh was constructed iteratively, such that sufficient elements were available during the process simulation to take

account of the curvature of the cavity and other form tools, particularly in transitional areas of rapid changes of geometry.

The SAX2T is also a fully thermo-mechanically coupled element type so that mechanical straining of the element can

influence the thermal state and in turn internal heat generation can influence the mechanical behaviour.

The simulation was programmed to precisely replicate the measured processing conditions and therefore the input

thickness set for the polypropylene sheet was 1.23mm. The plug and cavity materials were modelled as isothermal rigid

bodies, incapable of deformation but capable of contact and thermal interaction with the sheet. The form tooling can

therefore act as either a heat sink or heat source, but does not increase or decrease in temperature during the analysis. This

assumption significantly reduces the computational complexity with very minor consequences in terms of the predictive

accuracy of the simulation. In the actual thermoforming process there is generally a thermal gradient between the sheet and

each of the tooling surfaces, but as previously explained the most important thermal interactions are normally between the

plug and the sheet. Clamping of the sheet material between the downholder and mould was simulated by constraining the

sheet around its clamped periphery. An accurate plug traverse was created in the model to displace the plug vertically

downwards into the sheet by 70mm, as discussed previously. Upon completion of this plug displacement a thermoforming

air gauge pressure of 5bar is achieved over a total pressurisation phase of duration 0.4s. This is performed using a fluid cell

in the ABAQUS software which introduces a mass flow rate of air into the plug-side cavity of a maximum value of 9e-03kg/s.

This ramping air-pressure forms the material into the shape of the cavity, where it cools and sets rapidly upon contact.

A fully coupled thermal-stress analysis was performed since the mechanical and thermal solutions in this situation

affect each other considerably and therefore the solution must be obtained simultaneously. The analysis provided for

thermal conduction across the heated sheet and into the cooler tool surfaces, and convection from the free surfaces of the

sheet to the surrounding air. In the case of conduction this included a simple gap conductance function:

q = k (A − B )

where  A and B are the temperatures of the respective surfaces in contact, q is the heat flux (W/m2) and k (W/m2K) is

the conductance coefficient. This function activates once the gap between the surfaces closed to within 0.01mm.

The thermal conduction properties of the materials were taken from literature and the values used for the gap

conductance and convection coefficients were estimated from experiments. In this case the convection coefficient chosen

was 25W/m2K and the gap conductance coefficient was 32W/m2K. Thermal radiation was not treated separately in the

analysis. However, its relatively minor influences were accounted for in the consideration and selection of the convection

coefficient chosen from the experiments. It should also be stated that at the time of writing it is not possible in

ABAQUS/Standard to perform a coupled temperature-displacement analysis that directly includes thermal radiation effects.

Page 9
The starting temperatures for the polymer sheet, plug and mould during the simulation were 155°C, 100°C and 60°C

respectively. Interfacial heat generation caused by contact was also accounted for with any heat generated from a frictional

source being added to the thermal balance of the polymer sheet.

For the treatment of contact friction, the simple Coulomb relationship was assumed for all contacting surfaces with

a single value of friction coefficient used to represent all conditions, whether static or dynamic. Initially relatively high values

for the coefficient of friction at the plug/sheet interface were applied (around  =0.80) as the plug tends to have a rather

sticky initial contact with the hot polymer sheet. A lower friction coefficient of  =0.40 was selected for the sheet/mould

interface, which reflects the much lower contact temperature at this interface. However, as is the case in practice, it was not

expected that this would result in any significant slip on the mould wall due to the much improved mechanical properties of

the sheet at these lower temperatures. In both cases the chosen values were consistent with practical thermoforming

experience and the published experimental measurements of Collins et al., 2001 from sled tests, and Beilharz et al., 2007

with a rotational viscometer. In the initial model the chosen coefficients of friction were independent of temperature as it was

deliberately planned to use the simulation to investigate the effects of friction in a decoupled manner.

The large strain thermally coupled [LSTC] constitutive model for p olypropylene

The key innovation in the development of the process simulation was the selection and implementation of a suitable

mathematical model to represent the deformation response of the polypropylene material at thermoforming conditions. A

review of literature performed for this study established that researchers are increasingly turning to viscoelastic models for

thermoforming applications, but there was little agreement on the best model to use for any given polymer material. For this

work the strategy chosen was to adapt an existing approach that had been developed by some of the authors over a number

of years for polycarbonate (PC) and polyethylene (PE) (Sweeney et al., 2007) and PP (Olley and Sweeney, 2011), all in cold

drawing experiments. This subsequently was demonstrated to provide excellent results in simulating the multiaxial large

strain response of PP at elevated temperature (Sweeney et al., 2009), but it had never been applied to a full industrial

process such as thermoforming before. In particular the existing model architecture needed to be enhanced to add

temperature dependence and to cope with the full non-linearity of the complete industrial operation. The modified version of

the model for this application has been entitled the Large Strain Thermally Coupled (LSTC) material model. A 1-dimensional

schematic of the LSTC model developed is shown in Fig. 4. The model structure consists of two circuits in parallel.

Page 10
Figure 4: A 1-dimensional schematic of the LSTC model for semi-crystalline polymers (ADJ = adjustable).
The top circuit (X-arm) contains a Neo-Hookean strain energy function W , which has the form of equation (1).

1
W = ( Nc kBT ) 12 + 22 + 32 − 3 (1)
2

where Nc is the number of cross-links per unit volume in this network, kB is the Boltzmann constant, T is the absolute

temperature in degrees Kelvin and i , ( i = 1, 2,3) are the principal stretch ratios. The Neo-Hookean spring represents

the change in entropy of the polymer network through the action of cross-linking behaviour in the material. There is no limit

to the chain extensibility in this X-arm of the circuit and the function makes use of Gaussian chain statistics. In series with

this is an Eyring process that provides a mathematical method to address the stress and rate dependency of the material

behaviour. In the Eyring model utilized in this paper a scalar plastic strain rate e is calculated using the expression detailed

in equation (2).

e = A exp (Vp ) sinh (Vs oct ) (2)

where A is a material constant which is often linked with initial induced strain at yielding,  oct is the octahedral shear

stress,  is the hydrostatic stress and Vp and Vs are related to the pressure and shear activation volumes of the material,

as follows:

Page 11
p s
Vp = ; Vs =
kBT kBT

where s and  p are the shear and pressure activation volumes respectively. The total strain in this X-arm is a

combination of the strain in the Eyring process and the strain in the neo-Hookean process, which can be represented by

e p
deformation gradient tensors F and F (the elastic and plastic components) respectively. These are combined

multiplicatively to provide an expression for the total deformation gradient F :

F = FeF p

Thus, 1, 2, 3 in equation (1) are the eigenvalues of the pure deformation V associated with the deformation gradient

F , i.e. the left pure stretch tensor V results when rigid body motion R has been removed from the deformation gradient

tensor F . The left stretch tensor V can be similarly decomposed into elastic and plastic components such that:

F = V e V pR

e p
where V is the pure elastic stretch tensor and V is the plastic component of the left stretch tensor.

p
The scalar strain rate e of equation (2) is related to the tensor plastic strain rate D by

1 p p
e= D :D
3

p
where D is the pure plastic rate of deformation tensor calculated as

Dp = V p V p−1

where V p is the rate of change of the plastic component of the left stretch tensor. The Levy-Mises flow rule is used to

specify how the scalar plastic strain increments defined by e are separated into the components of the plastic strain

increment tensor Δε pl . Details of the incremental analysis have been given by Sweeney et al. (2009).

The lower circuit (Y-arm) contains a construct that represents the change in entropy of the polymer network

through the action of cross-linking and slip-linking behaviours in the material. It was proposed by Ball et al. (1981) that in a

rubber network there will be two types of junction points: the permanent cross-links where the chains are joined by chemical

Page 12
bonds and entanglements which are temporary junction points caused by chains becoming entangled. They proposed that

these temporary junction points are replaced by slip-links, which allow the chains to slide over one another. The contribution

to the strain energy of this slip-linking behaviour is governed by a slipperiness factor  , which dictates the freedom of

motion of the chains over one another and holds a zero value in the case of no sliding, effectively making the junction a

cross-link. The Edwards-Vilgis modification of the Ball model introduces an additional parameter  , which is used as a
measure of the maximum extension ( max ) of the polymer network. max corresponds to a singularity in the stressing of

the network chains with  being its reciprocal (  = 1/ max ) (Edwards and Vilgis, 1988). An increase in the magnitude

of  represents a reduction in the maximum chain extension. The contribution to free energy is the summation of those

contributions from the slip-link and cross-linking behaviour. The Edwards-Vilgis strain energy function used in the model is

presented in equation (3)

  3  
  (1 −  ) i
2 2
1   
2
3
 ( Nc kBT )  i =1
2  1−  2  2
3
+ ln 

1 −  2
 i 

 

  
i =1
i
i =1

 
W =   (3)
     
  3  (1 +  ) 1 −  2  2
+ 1 (N k T )   ( ) i + ln 1+  2  + ln 1 −  2 3  2 
 
 2 s B  ( i )   i 
i =1   2
1 −   i  (1 − i )   
3

  2 2 i =1

     
 i =1   

where Ns is the number of slip-links per unit volume.

The Neo-Hookean spring and Eyring process detailed in the 1-dimensional schematic of Fig. 4 are depicted with an

arrow crossing through them with the abbreviation “ADJ.” beside the arrowhead. This is to indicate that the behaviour of the

Neo-Hookean spring and Eyring process are now adjustable. In the model presented here, the spring and the Eyring

process have temperature dependent material parameters. These effects are accounted for in the model by introducing

temperature dependence into the cross-linking density parameter ( Nc ) in the X-arm. Additionally in the X-arm of the

constitutive model, the pre-exponentiation factor ( A ) in the Eyring process expression, detailed in equation (2), is made a

function of temperature and this changes the flow behaviour of the material. In effect the pre-exponentiation factor ( A ) is

adjusted as the deformation evolves. If there is an increase in the material temperature this will activate an increased flow

from the Eyring process. Conversely a drop in temperature will restrict the activated flow processes of the Eyring dashpot.

Larger values of pre-exponentiation factor will precipitate yielding at lower values of stress and strain, which is observable

Page 13
with many polymers. The material model is provided with a range of values for the pre-exponentiation factor ( A ) and the

number of crosslinks per unit volume ( Nc ) for a range of temperatures covering the thermoforming window of the polymer.

A kinematic approach to accommodate the potential for internal heat generation in the material caused by

mechanical deformation is included in the model. A significant amount of energy is dissipated during the stretching process.

However, a significant proportion of this energy can lead to adiabatic heating of the material itself. In the operation of the

pl
model a scalar volumetric heat generation per unit volume, r , caused by the mechanical working of the material is

calculated using a backward Euler scheme to integrate an incremental plastic strain expression (4)

1
r pl =   Δε pl  : σtX + σtX+t  (4)
2t

where the stress in the X-arm only at the previous configuration, σ t , and the current configuration, σt +t , are summed
X X

and the double dot product of the resultant stress tensor and a tensor of true plastic strain increments, Δε pl , in the Eyring
process is formed. The product of the two represents the total energy dissipated in the Eyring process alone over the

increment. This is divided by the time increment between the previous and current configuration, t . The relevant stress

is assumed to be the average of σt and σt +t , hence the divisor of 2. The parameter  represents the fraction of the

energy dissipated by the Eyring process, which is returned to the material as heat. It can hold values between 0 and 1, but

in the current model its value was set to 0.30 after a range of values were checked for consistency with experimental

pl
observation of temperature rises during biaxial testing. The heat generation per unit volume, which is calculated as r in

equation (4), is returned to the heat balance equations.

Material testing, characterisation and fitting of model parameters

In order to determine the values of the material parameters for the LSTC model, samples of polypropylene sheet were

extensively tested using the custom-built flexible biaxial stretching machine (FBSM) at Queen’s University. Details of the

features and operation of the test machine are described in Martin et al. (2005). The experiments were conducted over a

wide range of conditions representing the thermoforming process. This included temperatures from 145-159°C, strains from

0.00 to 3.16, and nominal strain rates from 0.5 to 31.5/s. In addition, data was obtained for different modes of deformation

including constant width, equi-biaxial and sequential in order for it to be fully representative of the different stages of

thermoforming processes. An example of the model constants obtained by fitting to biaxial test data for the polypropylene

blend described earlier is shown in Table 1. The data given was obtained at a test temperature of 155°C, which matched the

Page 14
chosen forming temperature for the product. The data presented was tested at a nominal strain rate of enom =1.0/s and a

stretch ratio of  =3.0. The temperature dependency of the material response was provided by piecewise fitting of the

model constants from the available biaxial test data as presented in Table 2.

Table 1: LSTC model material parameters for the polypropylene blend at 155°C, enom =1.0/s, stretch ratio of  =3.0.

Table 2: Temperature dependent materials parameters for the polypropylene blend at given temperatures, enom
=1.0/s, stretch ratio of  =3.0.

In general the fit obtained is quite good for the equi-biaxial results presented in Fig. 5. Fig. 5 compares true stress-strain

data for the PP blend in equi-biaxial stretch conditions at three different test temperatures. The model predictions for each of

the three conditions are displayed in Fig. 5. The simulation results presented in Fig. 5 are a deliverable of using the LSTC

model in conjunction with the data set presented in tables 1 and 2. The initial stiffness response is predicted well by the

LSTC model in the curves detailed in Fig. 5. The yielding behaviour is similarly accurately simulated by the mathematical

model as is the strain hardening evident in the PP blend at strains greater than 60%. However the LSTC construct found it

difficult to replicate the significant post-yield strain softening response of this polypropylene blend which is very prevalent at

lower test temperatures in the 10-50% strain region. This strain softening is attributed by the authors to a reduction in the

strength of the material caused by internal heating of the material, which is itself a deliverable of the considerable mechanical

working of the polypropylene as it is progressively stretched and deformed during tests. This is an area of the model

performance that is currently being addressed by the authors.

Page 15
7.00
Exp. T=149°C Sim. T=149°C
6.00 Exp. T=155°C Sim. T=155°C
True stress, N/mm2

5.00 Exp. T=159°C Sim. T=159°C

4.00

3.00

2.00

1.00

0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
True strain

Fig. 5: Model [Sim.] vs. experimental [Exp.] data (equi-biaxial tests on the PP blend at 155°C, enom =1.0/s, stretch
ratio of  =3.0.
Simulation results

The performance of the process simulation employing the LSTC model for polypropylene in modelling the thermoforming of

the small yoghurt pot detailed in figures Fig. 1 and Fig. 2 is demonstrated in figures Fig. 6 to Fig. 12. In all of these figures

the output from the process is assessed in terms of the final wall thickness distribution in the formed product. All the data

presented in figures Fig. 6 to Fig. 12 are drawn from fully thermo-mechanically coupled simulations. The evolution of the part

wall thickness (STH) with time is shown pictorially in stages in Fig. 6. The progressive changes in sheet material shell

thickness (STH) during the plugging stage and the subsequent blowing stage respectively, at six different junctures in the

process, is illustrated in Fig. 6. The thermoforming process is illustrated from the start to the end of the plug’s displacement

into the clamped sheet, which covers a total period of 0.70s, and then the final application of air pressure, which takes a

further 0.40s. In the images shown, reductions in the local sheet thickness are indicated by increases in the relative

brightness of bands in the walls of the partially formed parts. The PP material is seen to effectively wrap around the plug

shape as it progressively deforms and presses into the sheet during its traverse into the lower recesses of the cavity up until

time t=0.70s. Once the pressure is ramped during the pressure cycle at t>0.70s the sheet is seen to billow and peel away

from the plug and ultimately comes into contact with the surrounding cavity, whereupon it cools and sets forming the final PP

product shape.

Page 16
Figure 6: Evolution of the wall thickness (STH) distribution in millimetres during the plugging step (t=0.00 to 0.70s)
and blowing stage (t=0.70 to 1.10s), illustrated with a 180 slice of container.
The equivalent wall thickness distribution from the centre of the base to the clamped lip of each of the partially formed parts

is plotted in Fig. 7 for the plugging step and in Fig. 8 for the blowing step. The final wall thickness distribution of the fully

formed part from experiments is included for comparison. It is clear from these illustrations that the majority of the thinning in

the material initially takes place in the unsupported sidewall. They also show that the material progressively wraps around

the plug as the stroke continues. The material that comes into physical contact with the plug around the base of the part is

observed to largely maintain its thickness throughout the plug displacement.

Page 17
1.2 Experimental
t=0.23s
t=0.47s
1.0
t=0.70s
Wall thickness, mm

0.8

0.6

0.4

0.2

0.0
0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 100.0
Distance from the centre of the cup base, mm

Figure 7: Evolution of the wall thickness distribution during the plugging step.

1.2

Experimental
1.0
t=0.83s
Wall thickness, mm

t=0.89s
0.8
t=1.10s

0.6

0.4

0.2

0.0
0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 100.0
Distance from the centre of the cup base, mm

Figure 8: Evolution of the wall thickness distribution during the blowing step.
This material is clearly restricted from thinning by the high frictional forces preventing slip of the material in this area. This

precipitates the thinning of the material progressively up the sidewall in the direction of the lip of the container. Overall the

unsupported material in the gap between the plug face and the clamp area at the mould interface has performed the

overwhelming majority of the deformation. At the end of the plugging stroke (0.70s in Fig. 7) some material elements in the

Page 18
sidewall have lost nearly 75% of their initial thickness. As time progresses further, the pressure ramps and progressively

starts to peel the material off the plug. The sheet noticeably billows at 0.83s in Fig. 6, and a significant alteration to the

deformation occurs. At this point the sidewall material starts to reach the outer cavity surface and due to a combination of

rapid cooling and frictional contact the thinning of this material ceases as it becomes locked to the surface. The forming

pressure then continues to stretch the remaining free material around the bottom of the part until there is contact of the sheet

across the entire internal surface of the cavity (1.10s in Fig. 6 and Fig. 8). The switching of the major deformation of the

material from the sidewall to the base of the part is very apparent from 0.70s onwards in Fig. 6. Beyond this point the

pressure effectively acts like a physical plug and forces the material down into the lower reaches of the cavity, which has a

major effect on the thickness distribution of the material. This can be verified by observing the very substantial difference in

the wall thickness distribution from 0.83 to 0.89s in Fig. 8. It is clear from this figure that the forming of much of the upper

sidewall of the part is almost complete after the plug step. The blowing step is therefore largely concentrated on defining the

final wall thicknesses in the part base and lower sidewall. It is evident that this step is responsible for taking the relatively

thick regions of the sheet that were initially largely captive through contact with the base of the plug and redistributing this

material towards the furthest extremities of the lower part of the cavity.

The overall performance of the simulation may be seen by comparing the final predicted part wall thickness

distribution at 1.10s in Fig. 8 with the measured distribution obtained from experiments. The correlation between the two is

generally very good, particularly in the base region of the part from 0-25mm and around the top lip (90-100mm), where the

predictions are within the scatter of the physical experimental measurements as indicated by the error bars. The largest

discrepancies are in the sidewall where the simulation underpredicts the thickness in the lower sidewall (25-50mm) and

balances this with a modest overprediction of the thickness in the upper sidewall (55-95mm). However, the average sidewall

thickness is almost the same. It is thought that these differences in the prediction of the sidewall thickness are not due to

any fundamental weakness in the LSTC material model, but instead are largely attributable to the adoption of a single, fixed

coefficient of friction for the plug/sheet contact in the model. This appears to permit too much sliding across the face of the

plug in the early stages of plugging (Fig. 7) which is unresponsive to thermal interactions in that area caused by plug-sheet

contact. As previously introduced, experimental studies on plug friction have already confirmed that the measured

coefficients of friction are very temperature sensitive and have concluded that friction models for thermoforming need to be

temperature dependent. Despite this weakness, the predictive performance of the LSTC model-based thermoforming

simulation shown here is much improved over results previously published for the same product when using material models

based on standard linear viscoelasticity theories (O’Connor et al., 2007). Additionally, throughout these simulation

Page 19
processes the LSTC model proved to be very robust computationally and, providing close control was maintained over the

solution time step, it proved to be considerably more stable than analyses utilising standard linear viscoelastic models.

Processing parameters study

As part of this investigation, the sensitivity of the simulation to changes in a number of the key processing parameters was

investigated as the research evolved. The availability for the first time of a proven material model within the simulation

provided a unique opportunity for the effects of other parameters to be examined. These results are shown in Fig. 9 through

to Fig. 12. The influence that contact friction between the plug and sheet has on the part wall thickness distribution is

provided in Fig. 9. Here the coefficient of friction between the plug and the polymer sheet has been varied between  =0.40

and  =3.20. In these examples the fully coupled model was employed along with the contact conditions and process

variables outlined earlier. The results show that the lower the coefficient of friction used, the greater the underestimation of

the wall thickness in the base region of the part. The best results are clearly obtained for a friction coefficient of  =0.8.

This value of friction coefficient is consistent with published data for PP in contact in the work of Beilharz et al. (2007). These

results confirm the previously reported observations of the importance of friction to the process and confirm the need for the

simulation to include realistic models for contact friction.

1.2

1.0
Experimental
Wall thickness, mm

0.8 Friction coefficient, μ=3.2


Friction coefficient, μ=1.6
0.6 Friction coefficient, μ=0.8
Friction coefficient, μ=0.4
0.4

0.2

0.0
0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 100.0
Distance from the centre of the cup base, mm

Figure 9: The influence of friction coefficient of the plug sheet interface upon the wall thickness distribution.
The influence that the speed of the plug has upon the thickness distribution is illustrated in Fig. 10, which shows

the wall thickness distribution obtained for speeds of 0.01, 0.10, and 1.00m/s. For these results the fully coupled model and

Page 20
a friction coefficient of  =0.8 were employed, all other model variables remaining unchanged. It is apparent that the overall

effect of speed is very small for small changes in plugging rate, which confirms the findings from previously published

experimental observations of the thermoforming process (McCool and Martin, 2010). However if the speed is increased by

larger magnitudes, i.e. factors of 10 as in Fig. 10 more pronounced changes in the behaviour of the material can be

observed. In this instance a significant change in plug speed effects the thickness distribution at the plug-sheet interface

and near to the entrance of the cavity at the lip of the container. A considerable increase in plug speed is shown to cause

the sheet to maintain significantly more material in the base of the pot and precipitate increased thinning near the lip of the

receptacle. Outside of these two areas, a change in plug speed has relatively minor effects on the distribution of the material

in most of the remainder of the product.

1.2

1.0
Wall thickness, mm

Experimental
0.8 Plug speed=1.00m/s
Plug speed=0.10m/s
0.6 Plug speed=0.01m/s

0.4

0.2

0.0
0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 100.0
Distance from the centre of the cup base, mm

Figure 10: The influence of average plug speed upon the wall thickness distribution.
The effect of thermal convection from the sheet, in the form of the convective heat transfer coefficient, or film

coefficient ( h ), chosen for the simulation is shown in Fig. 11. In this case the results for the chosen initial value of 25W/m2K

for the baseline simulation are compared to those with convection film coefficients of 5W/m2K and 100W/m2K. This lower

value of h =5W/m2K would be representative of a still air condition whereas the higher value would be appropriate for a very

energetic convective airflow. For thermoforming the actual situation is more complex as the sheet is also moving rapidly

(through stretching), so it is expected that the higher values for h would be more representative of actual convection

conditions. Changing the film coefficient is seen to have a sizeable effect on the overall wall thickness distribution in the part,

Page 21
although its influence is again most pronounced in the base and lip regions. At the lower convection values the thickness in

the base is slightly reduced and further away from the experimental measurements, thus confirming that the higher value is

more appropriate for the actual thermoforming process. The higher value convection film coefficient influences the material

to sustain the thickness of the material underneath the plug face. Since the LSTC material model is thermo-mechanically

coupled then changes caused by the more rapid cooling of the material are to be anticipated and indeed this is consistent

with thermoforming practice.

1.2

1.0 Experimental
Convection film coefficient, h=100W/(m^2K)
Wall thickness, mm

Convection film coefficient, h=25W/(m^2K)


0.8
Convection film coefficient, h=5W/(m^2K)

0.6

0.4

0.2

0.0
0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 100.0
Distance from the centre of the cup base, mm

Figure 11: The effect of convective heat transfer film coefficient upon the wall thickness distribution.
The final parameter investigated was the rate of application of the forming air pressure, which effectively controls

the straining rate experienced by the sheet material during the blowing stage. In this case this was adjusted by changing the

mass flow rate of air into the upper plug-side fluid cell. Varying the mass flow of air into the upper plug-side cell controls the

length of time required to reach the full forming gauge pressure of 5bar. A higher mass flow rate into the fluid cell will see the

forming pressure achieved sooner, whilst a lower mass flow rate will increase the length of time required to reach the full

5bar. A higher air mass flow rate leads to an increased rate of straining of the material and has the effect of maintaining a

thicker distribution of material in the base of the PP blend container. PP as a material is rate dependent and the fact that the

wall thickness distribution changes when part of the thermoforming process is slowed or accelerated is consistent. This

result is clearly demonstrated in Fig.12 where it is seen that the lowering of the mass flow rate into the fluid cell causes the

receptacle base to thin as compared to the two higher rate simulations presented.

Page 22
1.2

Experimental
1.0
Air mass flow rate=3.60e-02kg/s
Wall thickness, mm

Air mass flow rate=9.00e-03kg/s


0.8
Air mass flow rate=2.25e-03kg/s
0.6

0.4

0.2

0.0
0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0 80.0 90.0 100.0
Distance from the centre of the cup base, mm

Figure 12: The influence of the pressure cycle air mass flow rate upon the wall thickness distribution.
A comparison of the results from the parameters study shows that all of the key processing parameters

investigated have clearly discernible effects on the wall thickness distribution. In most cases the greatest effects are seen

around the base and lip regions of the formed part, and it is clear that these could not be ignored without significantly

compromising the overall performance of the simulation. Of the parameters investigated, variations in the coefficient of

friction at the sheet/plug interface have clearly the largest capacity to change the performance of the simulation. The

establishment of firm temperature-dependent friction relationships for use in thermoforming and the development of reliable

test methods therefore remain as clear priorities for future research. The importance of predicting the thickness distribution

in the sidewall of the container is underlined as this is a direct deliverable of the plug-sheet interaction and the frictional

conditions therein. Improved modelling approaches must be developed to predict this interaction with greater accuracy.

However, this work has clearly demonstrated the significant effects that the various heat transfer parameters can have on the

simulation and have confirmed the need for a fully thermo-mechanically coupled approach to the process model.

Conclusions

A fully featured process simulation has been successfully constructed for the plug-assisted thermoforming process and it has

been shown to provide good correlation with experimental measurements. It contains a sophisticated large strain thermally

coupled (LSTC) material model for polypropylene that has been carefully constructed using the results of extensive biaxial

stretching tests at thermoforming conditions. This material model has been demonstrated to provide a superior fit to biaxial

data and to offer a very stable computational platform for the process simulation. It has enabled a range of other processing

Page 23
parameters to be investigated, which has confirmed that the process is most sensitive to the parameters controlling

plug/sheet contact friction. Heat transfer parameters have also been shown to be significant and the requirement for the

model to be fully thermo-mechanically coupled has been established.

References

Ball, R.C., Doi, M., Edwards, S.F. and Warner, M., 1981. Elasticity of entangled networks. Polymer, 22, 9.

Beilharz, F., Kouba, K., Blokland, H. and Eyerer, P., 2007. Experimental investigation of friction, thermal conductivity and
heat transfer and their mathematical description for the thermoforming process simulation. In: Proceedings of Polymer
Processing Society Regional Meeting, Gothenburg, Sweden.

Bourgin, P., Cormeau, I. and Saint-Matin, T., 1995. A first step towards the modelling of the thermoforming of plastic sheets.
Journal of Materials Processing Technology, 54, 1-11.

Buckley, C.P., Jones, D.C. and Jones, D.P., 1996. Hot-drawing of poly(ethylene terephthalate) under biaxial stress:
application of a three-dimensional glass--rubber constitutive model. Polymer, 37(12), 2403-2414.

Capt, L., Rettenberger, S., Münstedt, H. and Kamal, M.R., 2003. Simultaneous biaxial deformation behavior of isotactic
polypropylene films. Polymer Engineering & Science, 43(7), 1428-1441.

Christopherson, R., Debbaut, B. and Rubin, Y., 2001. Simulation of pharmaceutical blister pack thermoforming using a non-
isothermal integral model. Journal of Plastic Film and Sheeting, 17(3), 239-251.

Collins, P., Martin, P.J., Harkin-Jones, E.M.A., 2001. Experimental investigation of slip in plug-assisted thermoforming. In:
Proceedings of SPE ANTEC Conference, Dallas, TX, USA.

Collins, P., Harkin-Jones, E.M.A. and Martin, P.J., 2002. The Role of Tool/Sheet Contact in Plug-assisted Thermoforming.
International Polymer Processing, 17(7), 361-369.

Debbaut, B. and Homerin, O., 1998. 3D numerical simulation of thermoforming: prediction of thickness and extensions. In
Proceedings of SPE ANTEC Conference, Atlanta, GA, USA.

Edwards, S.F. and Vilgis, T.A., 1988. The tube model theory of rubber elasticity. Reports on Progress in Physics, 51, 55.

Hosseini, H., Vasilivuch, B.B and Mehrabani-Zeinabad, A., 2006. Rheological modeling of plug-assist thermoforming. Journal
of Applied Polymer Science, 101, 5.

Hummel, S.R. and Nied, H.F., 2004. A procedure for measuring biaxial viscoelastic behaviour of thermoplastics.
Experimental Mechanics, 44(4), 381-386.

Karamanou, M., Warby, M.K. and Whiteman, J.R. , 2006. Computational modelling of thermoforming processes in the case
of finite viscoelastic materials. Computer Methods in Applied Mechanics and Engineering, 195, 5220-5238.

Laroche, D. and Erchiqui, F., 1999. Experimental and theoretical study of the thermoforming of industrial polymers. Journal
of Plastic Film and Sheeting, 15(4), 7.

Martin, P.J., Tan, C.W., Tshai, K.Y., McCool, R., Menary, G., Armstrong, C.G. and Harkin-Jones, E.M.A., 2005. Biaxial
characterisation of materials for thermoforming and blow moulding. Plastics, Rubber and Composites, 34, 276-282.

Martin, P.J. and Duncan, P., 2007. The role of plug design in determining wall thickness distribution in thermoforming.
Polymer Engineering & Science, 47(6), 804-813.

Makradi, A. Belouettar, S., Ahzi, S. and Puissant, S., 2007. Thermoforming process of amorphous polymer sheets: modeling
and finite element simulations. Journal of Applied Polymer Science, 106, 7.

McCool, R., Martin, P.J. and Harkin-Jones, E., 2006. Process modelling for control of product wall thickness in
thermoforming. Plastics, Rubber and Composites, 35, 340-347.

Page 24
McCool, R. and Martin, P.J., 2010. The role of process parameters in determining wall thickness distribution in plug-assisted
thermoforming. Polymer Engineering and Science, 50(10), 1923-1934.

Nam, G.J., Ahn, K.H. and Lee, J.W., 2000. Three-dimensional simulation of thermoforming process and its comparison with
experiments. Polymer Engineering & Science, 40(10), 2232-2240.

Nied, H.F., Taylor, C.A. and Delorenzi, H.G., 1990. Three-dimensional finite element simulation of thermoforming. Polymer
Engineering & Science, 30(20), 1314-1322.

O’Connor, C.P.J., Martin, P.J., Menary, G.H., McConville, E., 2007. Modelling the thermoforming of polypropylene using
Abaqus. In: Proceedings of Abaqus User’s Conference, Daventry, UK.

Sweeney, J., Spares, R., Caton-Rose, P. and Coates, P.D., 2007. A unified model of necking and shearbanding in
amorphous and semicrystalline polymers, Journal of Applied Polymer Science, 106, 1095 – 1105.

Olley, P. and Sweeney, J., 2011. A multi-process Eyring model for large strain plastic deformation. Journal of Applied
Polymer Science, 119, 2246-2260.

Sweeney, J., Spares, R. and Woodhead, M., 2009. A constitutive model for large multiaxial deformations of solid
polypropylene at high temperature. Polymer Engineering and Science, 49(10), 1902-1908.

Wang, S., Makinouchi, A., Tosa, T., Kidokoro, K., Okamoto, M., Kotaka, T. And Nakagawa, T., 1999. Numerical simulation of
Acrylonitrile-butadiene-Styrene material’s vacuum forming process. Journal of Materials Processing Technology, 91, 219-
225.

Warby, M.K., Whiteman, J.R., Jiang, W.G., Warwick, P. and Wright, T., 2003. Finite element simulation of thermoforming
processes for polymer sheets. Mathematics and Computers in Simulation, 61(3-6), 209-218.

Figure captions

Figure 1: Major dimensions (in mm) of the yoghurt pot cavity, provided in section view.
Figure 2: Simulation of the plug-assisted thermoforming process.
Figure 3: Wall thickness measurements for the product, taken as a profile from the centre of the base of the
container along a vertical profile terminating at the flat lip of the receptacle.
Figure 4: A 1-dimensional schematic of the LSTC model for semi-crystalline polymers (ADJ = adjustable).

Fig. 5: Model [Sim.] vs. experimental [Exp.] data (equi-biaxial tests on the PP blend at 155C, enom =1.0/s, stretch
ratio of  =3.0.
Figure 6: Evolution of the wall thickness (STH) distribution in millimetres during the plugging step (t=0.00 to 0.70s)
and blowing stage (t=0.70 to 1.10s), illustrated with a 180 slice of container.
Figure 7: Evolution of the wall thickness distribution during the plugging step.
Figure 8: Evolution of the wall thickness distribution during the blowing step.
Figure 9: The influence of friction coefficient of the plug sheet interface upon the wall thickness distribution
Figure 10: The influence of average plug speed upon the wall thickness distribution.
Figure 11: The effect of convective heat transfer film coefficient upon the wall thickness distribution.
Figure 12: The influence of the pressure cycle air mass flow rate upon the wall thickness distribution.
Table captions

Table 1: LSTC model material parameters for the polypropylene blend at 155C, enom =1.0/s, stretch ratio of  =3.0.
Table 2: Temperature dependent materials parameters for the polypropylene blend at given temperatures, enom
=1.0/s, stretch ratio of  =3.0.

Page 25

View publication stats

You might also like