Anda di halaman 1dari 105

EVALUATION AND STABILITY OF PET RESIN

MECHANICAL PROPERTIES

By
SUDHEER BANDLA

Bachelor of Technology in Mechanical Engineering

Acharya Nagarjuna University

Guntur, Andhra Pradesh, INDIA

2006

Submitted to the Faculty of the


Graduate College of the
Oklahoma State University
in partial fulfillment of
the requirements for
the Degree of
MASTER OF SCIENCE
July, 2010
EVALUATION AND STABILITY OF PET RESIN
MECHANICAL PROPERTIES

Thesis Approved:

Dr. Jay C. Hanan

Thesis Adviser

Dr. Raman P. Singh

Dr. Ranji Vaidyanathan

Dr. A. Gordon Emslie

Dean of the Graduate College

ii
ACKNOWLEDGMENTS

I take this opportunity to thank Dr. Jay C. Hanan, for being my advisor and

guiding me throughout my stay at OSU with innovative and thought provoking ideas. I

would also like to thank Dr. Raman P. Singh and Dr. Ranji Vaidyanathan for being on

my committee.

I would like to thank Niagara Bottling LLC., Mr. Rali Sanderson, and Mr. Bill

Hall for giving me opportunity to work on this project. I also would like to thank Mr.

Kaokam Sanouvong, Mr. Bradley Barbosa, Mr. Paul Wilson, Mr. Felix Ledesma, Mr.

Nick Hadad and other people working at their Dallas, California facilities for their timely

help regarding the test runs and allowing me to work there.

I would like to thank Dr. Hongbing Lu, for letting me use the Nanoindentor. I

would like to thank Dilek Cakiroglu for her assistance with Nanoindentation testing. I

would like to thank Dr. Raman P. Singh and his student, Arif Rahman for their assistance

with AFM scans. I would like to thank Dr. Justin Lang, from PerkinElmer for carrying

DSC analysis and Ms. Chris Grabon, from ThermoFisher Scientific for FTIR

spectroscopy data. I would like to thank Mr. Mike Lucas for his suggestions in the

design and fixture machining. I also would like to thank Mr. Alan Hall for his timely

support and providing access to the equipment at OSU Tulsa.

iii
I want to thank my mother, brother and other family members for their constant

support and encouragement for me towards pursuing higher education. I also want to

thank my colleagues Praful Bari, Dr. Hrishikesh Bale, Balaji Jayakumar, Massod

Allakarami, Nicholas B Phelps, Yanli Zhang, Advait Bhat, Reeaj Ahmed, Ranjan

Ganapthi Mahadevan, Chinmay Upponi, Rohit Vaidya for their assistance in my work. I

also would like to thank my friend Dhivakar Jeevan Kumar, and many others at and

outside OSU, last but not least my roommates for their support during my stay in

Stillwater.

iv
TABLE OF CONTENTS

CHAPTER Page

Abstract ...............................................................................................................................1
1. Introduction ....................................................................................................................3
1.1. Poly (ethylene Terephthalate) ...................................................................................... 3
1.2. Microstructure of PET ................................................................................................. 6
1.2. Relating Polymer Properties to the Microstructure...................................................... 9
1.3. Manufacturing Process............................................................................................... 11
1.3.1. Stretch-Blow Molding Process ............................................................................... 13
1.3.2. Blow Molding Process ............................................................................................ 14
2. Optimization of Polymer Resin Selection ..................................................................17
2.1. Motive ........................................................................................................................ 17
2.1.1. Concept of Nanocomposites ................................................................................... 18
2.2. Material Significance in Packaging Industry ............................................................. 23
2.3. Factors Effecting Resin Performance ........................................................................ 26
3. Materials and Methods ................................................................................................28
3.1. Materials .................................................................................................................... 28
3.2. Mechanical Testing Techniques ................................................................................ 29
3.2.1. Tensile Testing Method .......................................................................................... 29
3.2.1.1. Fixture Design...................................................................................................... 30
3.2.1.2. Testing Procedure ................................................................................................ 33
3.2.1.3. Strain Measurement Techniques .......................................................................... 36
3.2.2. Nanoindentation Technique .................................................................................... 38
3.2.2.1. Nanoindentation Method ..................................................................................... 40
3.3. Crystallinity Measurement Techniques ..................................................................... 43
3.3.1. Density Method....................................................................................................... 43
3.3.2. Differential Scanning Calorimeter (DSC)............................................................... 45
3.3.3. X-ray Diffraction .................................................................................................... 46
3.3.4. Spectroscopy Techniques........................................................................................ 48

v
3.3.5. Microscopy Techniques .......................................................................................... 51
4. Results ...........................................................................................................................56
4.1. Tensile Test Results ................................................................................................... 56
4.2. Nanoindentation Results ............................................................................................ 61
4.3. DSC Results ............................................................................................................... 63
4.4. X-ray Diffraction Results........................................................................................... 65
5. Discussion......................................................................................................................67
5.1. PET Resins................................................................................................................. 68
5.2. Stability of Mechanical Properties ............................................................................. 69
6. Conclusions ...................................................................................................................75
7. Future Work .................................................................................................................77
7.1. Impact of Stretch Ratio on Mechanical Properties .................................................... 77
References .........................................................................................................................80
Appendix I: Yield Strength Calculations....................................................................87
Appendix II: Process at Niagara .................................................................................89
Appendix III: Design and Apparatus..........................................................................92

vi
List of Tables Page

Table 1: Effect of nanoparticles on polymer properties[51] ............................................. 20


Table 2: Comparison of different non-metallic reinforcements[49, 59] ........................... 22
Table 3: Increase in mechanical properties of PET-Nanocomposites .............................. 26
Table 4: Yield Strength of resins through different approaches ....................................... 58
Table 5: Modulus (GPa) values for PET samples of different forms ............................... 63
Table 6: Crystallinity for different samples of B 0.82 PET resin ..................................... 64
Table 7: Crystallinity (relative to amorphous sample) corresponding to sample location 79

vii
List of Figures Page

Figure 1: Poly (ethylene Terephthalate) molecular structure [7] ........................................ 4


Figure 2: Flow chart of processes involved in PET polymerization[9]. ............................. 5
Figure 3: Trans and Gauche conformations for a PET dimer [25] ..................................... 8
Figure 4: Crystallization half-time curve of amorphous PET Vs Temperature, (b)
schematic of oriented PET[29] ........................................................................................... 9
Figure 5: (a) Areas of preform covered by different heating zones, (b) screenshot of
temperature setting display from control panel ................................................................ 12
Figure 6: Sections and their corresponding weights for a 12.5 gram bottle (source:
Niagara Bottling LLC.) ..................................................................................................... 15
Figure 7: General orientation map along the bottle length [34]........................................ 16
Figure 8: Types of Nano-particles .................................................................................... 19
Figure 9: Lateral and axial compression of thin cylinder ................................................. 24
Figure 10: Factors influencing PET properties ................................................................. 26
Figure 11: Preform and its cross-sectional view from a CAD model ............................... 30
Figure 12: (a) ASTM test sample specifications[68], (b) Cross-section view of the fixture
with sample ....................................................................................................................... 32
Figure 13: Preform mounted on materials tester using the designed fixture. ................... 33
Figure 14: Complete test setup with Laser Extensometer, (b) sample and fixture
assembly, and (c) Laser front panel. ................................................................................. 34
Figure 15: Schematic of Laser Extensometer parts, basic operation, and tensile sample
with reflective gauge marks [70, 71] ................................................................................ 37
Figure 16: Typical Load-Displacement curve from a Nanoindentor[73] ......................... 39
Figure 17: Location of biaxial stretched films samples collected along the length of a
bottle. ................................................................................................................................ 41
Figure 18: Triboindenter, important components [79] ..................................................... 42
Figure 19: Crystallinity and corresponding density, based on two different c values.[6]45
Figure 20: Representation of Meridional and Equatorial scans for fiber sample [90]...... 47
Figure 21: Finger print region of an FTIR spectrum, for PET resins with different
microstructures. ................................................................................................................. 50
Figure 22: Change in the peak ratio corresponding to crystallinity along the length of a
PET bottle [97].................................................................................................................. 51
Figure 23: Working of an Atomic Force Microscope [103] ............................................. 54
Figure 24: Example AFM scans of the surface of PET resin pellet showing molecular
chains (using Veeco AFM at Microscopy Laboratory, OSU) .......................................... 55
Figure 25: Load vs. Extension (crosshead) curve features related with the sample
(preform) state ................................................................................................................... 57
Figure 26: Cross-head and Laser Strains up to 0.25 mm/mm Engineering Strain ........... 58
Figure 27: Representation of approaches used for Yield Stress, (a) and (b) change in
linear region slope, (c) for 0.2% strain and (d) for 0.5% strain. ....................................... 59

viii
Figure 28: Young's Modulus of different resins, all resins are of 0.80 I.V. unless
otherwise marked. ............................................................................................................. 60
Figure 29: Yield Strength of different resins (data based on approach (b)) ..................... 60
Figure 30: PET resins ranked based on yield strength ...................................................... 61
Figure 31: (a) Preform sample typical failure, (b) brittle failure of the sample................ 61
Figure 32: Load-Displacement curves for PET film sample (resin F).............................. 62
Figure 33: Youngs Modulus of resins from Nanoindentation data of amorphous samples
........................................................................................................................................... 62
Figure 34: DSC curves for PET samples with different micro-structure.......................... 64
Figure 35: XRD scan of PET samples (B 0.82) (preform and pellet samples intensity
secondary axis) (using Bruker diffractometer using Cu-radiation.) ................................. 66
Figure 36: Plot showing resins with respect to Modulus and Yield Strength ................... 69
Figure 37: Youngs Modulus vs. Yield Strength plot showing the scattered individual
data along with error bars based on standard deviation .................................................... 70
Figure 38: Standard deviation for Youngs Modulus of resins ........................................ 71
Figure 39: Modulus vs. sampling number ........................................................................ 74
Figure 40: Locations used for diffraction scans ................................................................ 78
Figure 41: Diffraction patterns at each location compared with preform sample
(amorphous) ...................................................................................................................... 79
Figure 42: Complete Stress-Strain Curve ......................................................................... 87
Figure 43: Stress-Strain curve for Yield Strength............................................................. 88
Figure 44: Injection Molding Unit, Robotic Arm and Mould .......................................... 89
Figure 45: Bottle Mould inside a Blow Molder ................................................................ 90
Figure 46: Stages of Blow molding[97]............................................................................ 91
Figure 47: Polarized image of normal preform with defective ......................................... 91
Figure 48: Preform holding pin and bottom cup ............................................................... 93
Figure 49: Modulus of resins with Laser and cross-head strains ...................................... 94

ix
Abstract

Mechanical properties of materials differ by their chemistry and, processing

conditions. Polymers are no exception. Poly(ethylene Terephthalate) (PET), an

aromatic-polyester made from renewable sources (feedstock) is making an impact on the

environment due to the required volumes (13 Mega tons/annum). PETs properties make

it the markets choice for packaging. The amount of PET recycled is less than 30%

leaving virgin resin consumption to make up the majority of product. An emphasis on

environmental responsibility is one factor driving the plastics industry towards use of

renewable materials and light-weighting. Differences in micro-structure achieved

through processing changes can provide a range of properties. This highlights the

importance of testing samples processed under actual processing conditions when

considering improved designs. In the packaging industry, a superior product can be

obtained through usage of efficient part geometry and materials. Customer acceptance is

also an issue. The addition of nanoparticles is to obtain an ecologically-friendlier and

improved performance material. Even 1 wt% addition has been shown to significantly

improve stiffness.

The goal of the work is to examine the typical PET resins available (I.V. 0.74-

0.84 dL/g) and obtain baseline performance data on mechanical properties. To compare

the material properties, injection molded amorphous samples were collected

1
on industrial-scale equipment. Before this study, mechanical property data of samples

processed at quenching and flow rates typically used in industry is not available. A

convenient method for precise mechanical testing of these samples was developed. This

new method, including a custom fixture, accommodates the preform shape used as an

intermediate in packaging production. Along with this new tensile testing method,

nanoindentation was performed along the length of industrially produced films.

Nanoindentation is another potentially useful method for measuring properties from

material samples. The preform samples were verified to be amorphous through DSC and

X-ray diffraction analysis.

Outcomes of this work show a 13-27% performance range in modulus between

the resins. This is notable since it is within the typical property improvements reported

by 1 wt% nano-particle reinforcements. The results correlate the yield strength and

Youngs modulus with a ratio of 0.015. These results also highlight the industrys need

for forming a record of resins available for resin selection.

2
CHAPTER I

1. Introduction

A polymer is defined as a naturally occurring or synthetic compound consisting

of large molecules made up of a linked series of repeated simple monomers [1].

Polymers or plastics as designated by industry are considered wonder materials.

Thousands of types of polymers have been developed since the time the first man-made

polymer was synthesized in the early 18th century. Molecular chains when combined

together with the same or different groups gives rise to a wide range of polymers.

Advancement in characterization and production techniques, has led to the development

of special grades of the same polymer depending on the application. Today, the most

commonly used polymers are Polyethylene (PE), Polypropylene (PP), Poly (ethylene

Terephthalate) (PET), Polystyrene (PS), and Poly (vinyl chloride) (PVC) [2].

1.1. Poly (ethylene Terephthalate)

Poly (ethylene Terephthalate), an aromatic semi-crystalline polyester, holds a 5%

share (13 MTPA (Million tons per annum) out of the total 250 MTPA [3-5]) of the

worlds polymer industry volume. Poly (ethylene Terephthalate), also referred to as

PET, finds its largest volume application in fiber and packaging industries.

3
Its combination of chemical inertness and physical properties makes it one of the most

preferred plastic in the packaging industry. It has an interesting micro-structural

response, where longitudinal stretching of molecular chains forms strong fibers, and bi-

axial stretching forms strong films. Linear PET is naturally semi-crystalline. Heat

treatment and mechanical history can drive it to be amorphous or more crystalline.

Crystallinity can also be impacted by additives. The PET monomer consists of a phenyl

ring and two ester groups as shown in Figure 1. First synthesized in the early 1940s,

PET is mainly produced in two grades with a basic difference in intrinsic viscosity (I.V.).

An intrinsic viscosity of 0.64 dL/g I.V. is considered fiber grade and 0.80 dL/g I.V. is

considered bottle grade. Usage of PET in the bottling industry started only with the

development of resins suitable for injection molding in the early 70s. Now, resins with

higher molecular weight (I.V. 0.7-0.85 dL/g) are used in the packaging industry [6].

Figure 1: Poly (ethylene Terephthalate) molecular structure [7]

Poly (ethylene Terephthalate) is made through a step-by-step process, instead of a

direct polymerization, to minimize the content of acetaldehyde (AA). While harmless,

AA has a citric taste and diffuses through the walls of a bottle imparting unwanted taste

to the contents. A more detailed review of PET processing is provided by Long and

Ravindranath et al. in their works [8, 9]. A brief summary is provided here. The first

step involves the formation of a prepolymer by esterification of raw materials TPA

(Terephthalic Acid) and EG (Ethylene Glycol), (or) by trans-esterification of DMT

4
(Dimethyl Terephthalate) and EG (Ethylene Glycol). In the second step, the resin

formed is further polymerized through a melt-polycondensation process. PET produced

through melt-polycondensation has a higher molecular weight. A flow chart with the

basic steps involved is shown in Figure 2. Many catalysts are available for

accelerating the process at different stages; they can influence the final properties.

DP 1.5 - 4
DP
DP 100

20 - 30 DP 150

DP 1.5 - 4

DMT & EG

Figure 2: Flow chart of processes involved in PET polymerization[9].

During the process of polymerization, monomers, the basic unit of polymers, are

joined together to form long molecular chains. Arrangement of these molecular chains

inside the polymers could be crystalline (organized), amorphous (random), and semi-

crystalline. Arrangement of the polymer molecular chains plays a key role in the

properties of the polymer. In relating polymer molecular organization to macro scale

properties, a direct correlation is not always possible. Inclusion of processing history

assists in their correlation, having knowledge about the processing conditions helps as

polymers are sensitive to the way they are processed.[10]

5
Molecular weight plays a key role in determining the properties of the polymer;

molecular weight and I.V. are inter-related [11] (Equation 7). Usual melt-processing only

provides a lower molecular weight resin [I.V. of 0.60-0.65 dL/g]. This processing

typically produces amorphous pellets. Secondary polymerization processes like Solid

State Polymerization (SSP) increases the molecular weight. During SSP, the resin is

reheated and allowed to crystallize at a slower rate. This also facilitates the escape of AA

from the resin and prevents further formation, increasing the crystallinity and I.V.. A

balance is required, since resins with a high initial crystallinity require increased

processing temperature during injection molding--potentially leading to polymer damage

and increases in AA content. Molecular weight (I.V.) and crystallinity of the resin along

with many other factors affect the final mechanical properties. PET resin pellets

produced through the SSP process are of uneven crystallinity. In order to overcome the

problems associated with SSP, such as sticking of pellets which effects crystallization

and high production costs, industry is shifting towards new technologies for producing

uniformly crystallized PET resin [12].

PET is processed into several forms to suit their applications such as fibers for

textile industry and as films for packaging purposes. By changing the reagents and

catalysts used for PET synthesis, different grades can be produced. Manufacturing

process differs based on the final output, such as spinning for fibers, Injection Stretch

Blow Molding (ISBM) for bottles.

1.2. Microstructure of PET

Poly ethylene Terephthalate (PET) because of its industrial importance is under

constant focus for researchers to have a complete understanding of its structure and the

6
relation to its properties. A great deal of research carried out on PET was focused on

understanding the crystallization behavior, and relating it to properties. Jabarin et al.

studied the biaxial orientation, strain-induced crystallization, and property changes of

PET [13-16]. Mahendrasingam et al. studied the orientation of PET molecular chains,

strain-induced crystallization behavior, transient structure existence before strain-induced

crystallization, and temperature effects at industrial strain rates (10 -12 s-1) through X-ray

scattering techniques using synchrotron radiation [17-19]. Matthews et al. [20] discussed

tensile drawing conditions and their effect on properties. Dixon et al. [21] tried to inter-

relate the mechanical properties with molecular weight. Guzatto et al. [22] studied

mechanical properties under plane strain compression. Venkateswaran et al. [23] studied

the temperature profiles along the preform thickness and their effect on properties of

blown containers. Silberman et al. [24] studied the stretch effects on the mechanical

properties of a PET bottle.

PET, semi-crystalline polyester, contains a mixed micro-structure with crystalline

and amorphous phases. A fully crystalline micro-structure makes it brittle with an

increased stiffness, while the amorphous structure makes it ductile and more compliant.

As a semi-crystalline polymer, both amorphous and crystalline phases are present.

Therefore, determining the quantity of crystalline phase helps in determining the final

properties.

7
Trans
Gauche

Rotation of c-c bond


(CH2-CH2-)

Phenylene ring

Figure 3: Trans and Gauche conformations for a PET dimer [25]

The presence of a glycol linkage allows PET to have two different rotational

conformations trans (extended form) and gauche (relaxed form); with the rotation of the

C-C bond (-CH2-CH2-), the conformations change from one form to other. Initial studies

considered trans conformation for crystalline and gauche conformation for amorphous.

Later it was revealed that both exist in the amorphous region. Therefore, an increase in

crystallinity can be correlated with a decrease in the amount of PET with ethylene glycol

units in the gauche conformation [26], measured through spectroscopy techniques. The

amorphous phase is also divided as Free amorphous phase and constrained amorphous

phase based on their location with respect to the crystalline phase. Constrained

amorphous phase is closer to the crystal region. It is clear that different micro-structures

can be produced by varying the process parameters.[20, 26]

PET favors crystallinity because of its structural regularity and phenylene rings

(as shown in Figure 3) symmetry. The presence of the asymmetrical iso-phenylene ring,

acts to counter structural regularity. This further decreases the chain regularity and the

8
tendency of the polymer to crystallize. During the stretch blow process, PET tends to

crystallize under strain and form tiny crystallites, (10 - 200 ), making the containers

transparent. PET displays a relatively slow crystallization rate and this rate of

crystallization (inverse of crystallization half-time) increases in between the glass-

transition temperature (Tg) and melt temperature (Tm), shown in Figure 4. Crystallization

half-time is defined as the time taken for reaching half of the final crystalline quantity for

the specific temperature and process conditions. It is affected by I.V., resin chemistry

(catalyst, co-monomers), and moisture content. Physical ageing (enthalpic relaxation)

speed of PET decreases with the increase in crystallinity, as the amorphous content

trapped between the crystalline regions increases along with activation energy.[27-30]

10
Crystallization half-time (min)

(b)
120 180 240
Temperature (oC)

Figure 4: Crystallization half-time curve of amorphous PET Vs Temperature, (b)


schematic of oriented PET[29]

1.2. Relating Polymer Properties to the Microstructure

Material properties are influenced by their micro-structure. Therefore engineering

the micro-structure plays a crucial role in tailoring the properties of a material. The

9
molecular orientation and the presence of crystalline volume and its texture play the

primary roles in determining the mechanical properties of a polymer. Semi-crystallinity

in polymers is desirable as it combines the strength and flexibility of crystalline and

amorphous phases, such that they can be tough and bend without breaking. Semi-

crystalline polymers are differentiated by their percentage crystallinity. When a polymer

is cooled from its melting temperature, it starts nucleating and forming crystals trapping

the amorphous phase between the crystals. This restricts movement of the amorphous

phase. The properties of the polymer are affected by the type of nucleation which is

further impacted by the molecular weight as the mobility and homogeneity reflect back

on the quantity of crystalline phase. The average molecular weight (number average),

within 24,000 to 36,000 g/mol (I.V. from 0.75 dL/g to 1.00 dL/g) in the case of PET,

provides the best properties whereas higher molecular weights make them difficult to

process.[10, 31, 32]

Crystallinity is dependent on the I.V. and heat treatment of a resin. The final

crystallinity in a film product is also dependent on the stretch-ratio. In the case of PET,

melt history, strain, DEG (Diethylene Glycol) content, residual catalyst, stretch

temperature, stretching speed, and initial crystallinity affect the final crystallinity. The,

initial crystallinity is defined as the crystallinity level of process input (raw resin for

injection molding, preform for blow molding) and final crystallinity is defined as the

crystallinity level of the final product. For films, a decrease in crystallinity is from

disruptions in the spherulitic structure, an increase is from strain induced crystallinity.

Also with increases in initial crystallinity, the final crystallinity decreases [33]. Smith et

al. has found that the molecular orientation level is dependent on preform heating

10
temperatures, using a total IR reflectance spectroscopy technique to map the molecular

orientation [34]. They attribute this change to chain relaxation. Orientation of the

amorphous phase also plays a key role along with the crystallinity content in predicting

the final properties. Tensile modulus and yield strength of a material can be related to the

orientation of the amorphous phase. Tensile modulus doubles with increase of stretch

ratio (amorphous orientation increases with the stretch ratio, and modulus is strongly

dependent on amorphous molecular orientation [35, 36]) where as yield strength is

influenced by both crystallinity and orientation of amorphous content, with crystallinity

being more significant [35]. However, an increase in crystallinity results in the decrease

of the amorphous phase orientation (measured based on the material relaxation with

thermal treatment). With increase in crystallinity, the crystalline phase prevents this

relaxation[37].

1.3. Manufacturing Process

A typical manufacturing process for PET parts is described below. PET resin

pellets (I.V. 0.80) are dried at 165oC (329 oF) for about 6 hours, to eliminate the

moisture present in the resin. Dried PET resin is melted by passing through the screw

barrel at 293oC (560oF) in the injection molding unit. Heating elements are placed along

the length of the barrel to maintain the uniform melt temperatures and to prevent shearing

of polymer chains. Presence of shear causes the resin to degrade. Preforms are produced

on an injection molding machine (Husky, HyPET 500 model) running at a cycle time of 6

to 6.5 seconds. The cycle time varies based on the ease of processing of the resin or

external factors such as maintenance of the cooling water lines for the injection molds.

Decreasing the cycle time produces more parts, but also impacts the quality of the

11
preforms. For this equipment, below a cycle time of 6 seconds, the quality diminishes

with visible defects in the parts.

For a two stage processing, preforms produced are typically allowed to cool and

stabilize before blow molding. As part of the blow molding process, the preforms are

reheated above the recrystallization temperature (87oC (190oF)) for stretch blowing into

bottles. Preforms are divided into zones for the purpose of heating, as shown in Figure 5.

These zones are fixed by lamp configurations in the blow molding equipment. The

individual control of temperatures at different zones on a preform assists in effective

distribution of the material. Heated preforms are mechanically stretched in the mold

along the axial direction and then blown by pressurized cold air to stretch in the radial

direction, taking the final shape of the mold. This can be performed at up to 11.9

preforms per second. Both processes are performed at the fastest known rates in the

world.

Zone 1
Zone 2

Zone 3

Zone 4

Zone 5

10 mm (b)
(a)

Figure 5: (a) Areas of preform covered by different heating zones, (b) screenshot of
temperature setting display from control panel

12
In these processes, the molecular arrangement and microstructure of t h e PET

changes significantl y from stage to stage. During the injection molding process,

polymer molecules freeze with alignment in the direction of injection. Molecular

orientation can also change along the thickness of the part. In the case of a preform,

molecular orientation on the outer and inner face has been found to differ from that of

sub-skin layers. The low temperature of the mold faces makes the layers in contact cool

and preserves the polymer orientation; but, in the case of the sub-skin layers, shear flow

continues to orient the molecules. Middle layers represent the least orientation due to the

lower shear forces and high initial melt temperatures. In the case of slow crystallizing

polymers (Ex: PET, Natural rubber, polyphenylenelene sulfide), crystallization depends

on the cooling rate and the applied stress levels during injection. As mentioned by Ben

Daly et al., parts of a slow crystallizing polymer contains three different layers along the

thickness, an amorphous skin layer, a crystalline intermediate layer, and an amorphous

core layer. The crystalline intermediate layer is created because of the stress induced

crystallization produced from the high shear forces[38]. Development of the micro-

structures is purely dependent on the processing conditions (cooling time, mold

temperatures), faster cooling time result in completely amorphous samples.

1.3.1. Stretch-Blow Molding Process

During the Stretch-Blow molding process, the PET molecules are subjected to

strain by a mechanical stretching process. Next, during the blowing operation, the

pressure subjects the PET to biaxial strain. The molecular orientation of PET changes

under strain, causing the gauche conformations present in the amorphous region of the

sample to transform into trans conformations (as shown in Figure 3), thereby increasing

13
the crystalline content. This particular phenomena is called strain-induced

crystallization, it occurs at a very fast rate and only for a small time [17]. The rate of

crystal orientation is dependent on the extension rate, temperature, and draw ratio. The

temperature increases with the rate of orientation, as temperature increases molecular

mobility. The orientation level achieved for a draw ratio decreases with temperature and

with the decrease of draw rates [39]. The start of this strain induced crystallization is

dependent on the draw rate. At very small extension rates, there are no signs of

crystallinity development, but for low (intermediate) draw rates, crystallization starts

during the deformation process, and at very fast draw rates such as 10 to 12 s-1,

development does not start until the end of the deformation. Blundell et al. described this

phenomenon during extension as a local segments ability to provide chain mobility to

organize into a crystallographic register and at end of the process, onset of chain

relaxation would restore the freedom of local organization and allow crystal nucleation to

occur [17].

1.3.2. Blow Molding Process

In the blow molding process, during the stretching of PET, molecular chains also

tend to relax and try to counter crystallization. With an increase in crystallinity, the time

for chain relaxation increases due to higher effective viscosity (viscosity of the material

at the time of blowing), which permits a further increase in the orientation (crystallinity).

At higher draw ratios, the crystallization kinematics become more complex as the

entropic driving forces for chain relaxation and the rate of strain-induced crystallization

increases. Segmental mobility (movement of the crystal segments within the polymer

melt), increases with temperature while boosting both relaxation and crystallization.

14
Several researchers have come up with empirical orientation functions using lattice

parameters such as Hermans orientation function. The crystallization rate is sensitive

to the degree of molecular orientation and temperature. Change in the molecular

orientation can be tracked using techniques such as diffraction and IR spectroscopy.[17,

40-42]

12.5 gram

6.2 grams
Bell

70 mm
Upper
Panel 2.5 grams
0.2
127 mm grams

2 grams
Lower
178 mm Panel

1.8 grams
Base

Figure 6: Sections and their corresponding weights for a 12.5 gram bottle (source:
Niagara Bottling LLC.)

The bottle making process involves several interdependent parameters. The

design of the process will be based on specific grades of material (PET, 0.7-0.85 I.V.)

and with the final objective of better load bearing capacity with less resin, simplified

processing, and aesthetics. An optimum blowing temperature selection is critical for

achieving the best properties [43].

15
Crystallinity changes from location to location on the bottle, a molecular

orientation map for PET is shown in Figure 7. Properties are not only dependent on the

process but also on the quality of raw materials used. Therefore resin selection becomes

critical. Also, the required microstructure can be obtained by increasing the crystallinity

of oriented PET, which can be increased by heat setting method, where the final product

is reheated (to 100oC 110oC), thereby increasing its crystallinity. Heat setting increases

the apparent glass transition temperature of the resin. This increase in glass-transition is

because of the increase in thermal stability of the finished product with increase in

crystallinity [29]. The crystallization rate of an oriented PET is greater compared to the

random oriented polymer.

At neck, highly oriented


in Y (length) direction

Outside shoulder has


roughly equal amounts of Slightly preferred
chains in X (Hoop) and Y orientation to Y (length)
(length) direction direction

Orientation around
Slightly preferential X (Hoop) direction
orientation in X (Hoop) is higher for inside
direction, increasing surface then outside
towards the bottle base surface

Noticeable orientation switch


into Y (length) direction for
outside wall only

Figure 7: General orientation map along the bottle length [34]

16
CHAPTER - II

2. Optimization of Polymer Resin Selection

2.1. Motive

Polymers take up 70% of the packaging industry [44], with PET holding about

50% share in plastic bottling field only [45]. In the United States alone, 2.4 million tons

of PET is used yearly for packaging but, only 27% is recycled [46]. Production of highly

purified recycled PET was only made possible recently [8]. Packaging material demand

for industries related to food contact applications, such as the bottling industry, require

the use of virgin resin. With the need to use virgin resin and the markets demand to

become ecologically-friendlier, manufacturers have developed lighter weight bottle

designs.

Over the past decade, the bottling industry has moved from a 21 gram to 9.3 gram

package which is about a 60% reduction in the amount of polymer [47], and therefore

reduces its environmental impact. The industry is working towards further reductions in

bottle weight. Effective allocation of plastic at critical locations based on FEA (Finet

Element Analysis) simulations helps in further reducing the weight and at the same time

choosing a material with superior properties allowing further weight reduction.

17
To facilitate further reduction in weight strengthening existing materials is an option.

Materials used for the packaging industry are selected based on the properties such as

mechanical strength, chemical inertness (it should not react with the contents), low

permeability for maximum shelf life, and aesthetic aspects (optical clarity). Thus the

selection of strengthening techniques should consider these factors, such that these

properties would not degrade in the process.

One of the most accepted techniques to strengthen materials is by creating a

composite of the original material with a small quantity of reinforcements. The addition

of reinforcements has both positive and negative effects on the mechanical properties, for

example with improvement in strength of the material; there would often be a reduction

in the break elongation. The range of reinforcements available in the market helps in

developing different grades of materials.

2.1.1. Concept of Nanocomposites

With the raise of Nanotechnology, researchers started using materials ranging

from micro to nanometer scale as reinforcements. When at least one dimension of the

reinforcements used is in the order of nanometers, then that class of composites is called

as Nanocomposites. Further, they are classified as three different types based on the

geometry of the reinforcements, identified by number of nano-scale dimensions, such as

particulate (spherical), fibrous and layered as shown in Figure 8 [48].

18
d l
d
l t
l
Particle Fiber Platelet

Figure 8: Types of Nano-particles

The increase of properties in nanocomposites is attributed at least in the initial

stages to a nano effect, later researchers showed that nanocomposites can be

theoretically modeled using different continuum models with exceptions (Ex:- Tg) [49].

Research suggests the nanoparticles can act as nucleating agents for crystallization.

Increase in the strength of the polymer can be attributed to the increased crystalline

structure. Also due to the addition of the filler particles, the weight of the polymer is

reduced. This reduction is attributed to the degradation of the long chains to shorter

chains during the processing of nanocomposites [48, 50]. Nanoparticles have a higher

aspect ratio i.e. larger surface area for a given volume leading to higher surface energies.

Thus nanoparticles possess different properties compared to their bulk [51]. Polymer

chains are in the dimensional range of nanoparticles, molecular interaction at the nano

scale provides an added advantage for nanocomposites. The addition of nanoparticles

affects polymer processing and properties differently as listed in the Table 1.

19
Table 1: Effect of nanoparticles on polymer properties[51]

Improved Properties Disadvantages


Mechanical properties (tensile Viscosity (degradation) [52]
strength, stiffness, toughness) [52]
Fatigue sensitive (degrades
Gas barrier[53] mechanical properties) [56]
Flame-retardant Dispersion difficulties
Thermal expansion Optical issues[52]
Conductivity [54, 55] Sedimentation
Ablation resistance Color (when different carbon
containing Nanoparticles are used)
Chemical resistance

The significant advantage of nanoparticle additions will only be seen when the

nanoparticles are well dispersed in the polymer matrix. An uneven distribution or

agglomeration of the nanoparticles results in degradation of polymer properties.

Nanoparticles tend to form agglomerates because of their higher surface energy, making

them difficult to exfoliate and disperse in the matrix. Nanoparticles in a nanocomposite

act similar to short-fibers in short-fiber composites, at a smaller scale. Dispersed well

into the matrix, they can form a continuous network to facilitate load distribution. Also,

in the case of nano-platelets (like Clay platelets or Graphene), the stress concentration

levels are reduced due to the end effects of the nano-platelets [57].

The overall performance of polymer nanocomposites cannot be understood by

simple scaling rules that are applicable for traditional composites. Many factors can

affect the properties of a polymer nanocomposites (PNCs)[58]:

Types of the filler particle used and their surface treatments

20
Microstructure of the polymer nanocomposite (PNCs)

Process used for preparation of PNCs

Polymer chemistry, molecular weight, and crystallinity of polymer matrix

Since the first usage of PET in 1970 for blow molding, researchers have been

trying to understand it at the molecular level to better engineer its properties. Several

grades of PET which suit special purposes were developed, but they have not completely

addressed industrys evolving needs. PET even with its best properties, needs further

improvement in its strength and barrier properties, as the packaging industry is making

rapid progress. At present, industry is going with the aim of using the least possible

quantity of plastic (older products used to be thicker to compensate many of the defects

and more closely simulate glass); that way they can decrease the rate of raw material

(crude oil) consumption (as most of the resin used in packaging is primary in terms of the

recycling cycle) and protect the environment by reducing the plastic waste by quantity.

A variety of nanoparticles such as natural clays, silica, TiO2, Mica, Carbon black,

Carbon Nanotubes (CNT), Carbon Nanofibers (CNF), Graphene sheets, and cellulose

whiskers have been used as fillers for preparing polymer nanocomposites. Other than the

above mentioned, some additional particles have been tested for preparing

nanocomposites, such as: metals (Al, Fe, Au, and Ag), metal oxides (ZnO, CaCO3, and

Al2O3) and SiC. The selection of nanoparticles is based on the desired property

enhancement: electrical, mechanical, chemical, and barrier (for example Al nanoparticles

are added to improve electrical properties) [51]. Cost and process friendliness should

also be taken into consideration before selecting the nanoparticles. A few of the

21
important reinforcements available in market are compared for their physical and

mechanical properties in Table 2.

Table 2: Comparison of different non-metallic reinforcements[49, 59]

Property Graphene Clay Platelet SWCNTs Carbon Fiber Glass Fiber


Sheet (T-300) (S-Glass)
Physical Structure 1 nm x 100 nm 1 nm x 1000 nm 1 nm (D) 10 m 8-13 m
Tensile Modulus (GPa) 1000 170 1000 231 85
Tensile Strength (GPa) 10 20 1 20 50 3.65 4.6
Resistivity ( cm) 50 x 10-6 1010 - 1016 0 10 -4 18 x 10-4 -
Thermal Conductivity 3000 0.67 5000-8000 8.5 -
(W/m K)
CTE 1 x 10-6 8 16 x 10-6 - -0.6 (axial) -
Density (g/cm3) 2.0 2.5 3.0 - 1.76 2.49
D-spacing 0.34 1.85 - - -

Several nanoparticles like functionalized clay, metal oxides, and CNTs, have

been used for preparing PET-nanocomposites. Liu et al. [60] have used carbon-black as

an additive for making PET-nanocomposites. In their work, electrical properties of the

nanocomposite were studied. Sanchez et al. [53] in their work on PET-nanocomposites

used food-contact-compliant nanoclay, NanoterTM for improved barrier properties. They

also made a comparison with nanocomposites made from bio polymers such as PHB,

PHB/PCL blend. PETnanocomposites demonstrate superior performance compared to

others nanocomposites. Yang et al. [61] studied crystal growth and morphology in PET

nanocomposite thin films of 150-nm-thickness with -alumina (Al2O3) nano-particle

fillers of size 38 nm. Chang et al. [62] made PET nanocomposites using C12PPh-MMT

(organically modified montmorillonite, synthesized from Na+-montmorillonite (Na+-

MMT) and dodecyltriphenylphosphonium chloride (C12PPh-Cl)) organically modified

clay, and studied the thermo-mechanical properties and morphology of the

nanocomposite at 0-3% (wt%) addition of clay. They found about an 85% increase in the

22
tensile modulus at 3% addition with no change in elongation at break. Yao et al. [63]

prepared PET nanocomposites using Attapulgite (AT), a kind of natural, fibrous silicate

clay with exchangeable cations and reactive-OH groups on its surface by mixing the

clay with EG (Ethylene Glycol) during the polymerization process of PET. They found

an increase of about 45% in the tensile modulus and 10% increase in the strength of the

composite over the virgin PET. Brando et al. [64] studied PET and lamellar zirconium

phosphate nanocomposites. They used two different zirconium phosphates, ZrP (with

hydroxyl side chains) and ZrPP (with phenyl rings) and claimed an increase in the

modulus around 11% with ZrP, 30% with ZrPP in modulus. Anoop et al. [55] and Mun

et al. [54] has reported usage of CNTs with PET for preparation of PET-

Nanocomposites. In the work by Anoop et al. they found that CNTs addition till 3% had

no significant improvement in the thermal stability of the mechanical properties, but

showed improvement in the electrical properties. Mun et al. has prepared PET fibers

using MWCNTs with similar kind of observations. They attributed this reduction in

strength to the debonding in the interface of PET-MWCNT and void formation.

2.2. Material Significance in Packaging Industry

The objective for the packaging industry is to preserve the contents from damage

in harsh environments and convenient transportation. Requirements for a material for

packaging application include impact resistance, stiffness, gas barrier properties, creep

resistance, and transparency. Industry considers the top load capacity a bottle design

provides before clearing it for production. Top load is a significant factor for stable

transportation and storage of packaged inventory. Ideally, designs for optimum top load

would be done through FE simulations, but leading edge manufactures often do not have

23
the luxury of validated material properties making FEM results questionable. Material

properties such as Youngs modulus and yield strength are important in the process of

bottle design. As an example, the significance of material properties for package design,

such as a bottle, and their design attributes can be approximated from a thin walled

cylinder. Here the change in axial and lateral compression behavior can readily be

observed and the significant variables identified. The expressions for the maximum load

a cylinder can withstand before it buckles or collapses under axial compression or

maximum load before it gets crushed from side forces (during packaging, transportation)

is given in equation 1 and 2 [65, 66].

3 Er 2 t
Pa = 2 Pa E (1)
L

2 El
Pl 2 Pl E (2)
r

PL
Pa

Figure 9: Lateral and axial compression of thin cylinder

24
The main factors that determine the load bearing capability of PET products are:

Part geometry

Tensile Modulus

Final wall thickness

In the case of two cylinders of identical geometry, the key difference between

them is the material mechanical properties. PET is the best available material with

required properties for packaging application at high available production volumes.

Property limitations (permeability) are typically overcome through adding material

thickness. The addition of nanoparticles to PET through proper techniques can provide

the advantage of engineering the desired mechanical properties without the loss of

existing properties. This will compensate the material addition for additional

performance. Improved barrier properties also will help in increasing the shelf life of

food products. Uniform distribution of the nanoparticles is not expected to affect the

optical clarity of the final product. Mechanical properties of PET-Nanocomposites from

literature are shown in Table 3. It is evident that the addition of nanoparticles improves

the properties significantly. Also, PET virgin resins have a wide disparity in their

mechanical properties, with in the same grade of resins.

25
Table 3: Increase in mechanical properties of PET-Nanocomposites
Investigator Nanoparticle wt % Strength (MPa) Modulus (GPa)
0 46.00 2.21
Chang et al. MMT Clay 1 58.00 2.88
3 71.00 4.10
0 55.82 2.62
Yao et al. Attapulgite clay
2 61.35 3.80
0 42.10 1.17
Anoop et al. CNT
3 54.00 1.87
0 46.00 2.21
Mun et al. MWCNT
1.5 57.00 3.14
Zr-Phosphate 0 45.34 1.58
Brando et al.
(ZrPP) 5 58.80 2.22

2.3. Factors Effecting Resin Performance

Molecular weight plays a key role in determining the properties of the polymer.

Molecular weight, Intrinsic Viscosity (I.V.), and crystallinity of the resin are significant

factors along with the many others shown in Figure 10, in affecting the final properties of

packaging films such as bottles [6]. Catalysts used during the polymerization process

also influence crystallization behavior. This is because of the secondary molecular

structure evolved due to the usage of catalysts and because of catalyst residuals present

[67]. Selection of the catalysts becomes critical as their effects are not well documented.

Molecular Structure Processing History


Chemical Crystallinity level Melt temperatures
Properties Amorphous phase Cooling rate
orientation Blowing conditions

Macroscopic properties
Mechanical
Barrier
Physical

Figure 10: Factors influencing PET properties

26
In order to minimize their effect, a single catalyst is to be selected such that it

would be suitable for multiple stages during the polymerization process. Acetaldehyde

content is another critical factor which would affect resin selection and performance, as it

diffuses over time and can impart flavor to the contents of a package. The presence of

acid end groups in polymer chains leads to thermal and hydrolytic instability causing

thermal cleavage of the chains producing more AA and water which further increases

degradation. Another side product, diethylene glycol (DEG), reduces the melting point

and thermal stability of the final product. Reduced thermal stability increases the chances

of resin degradation during injection molding. Degradation of resin occur at all

processing stages. It reduces the average molecular weight which further affects the

mechanical properties. Side products produced during the polymerization process need

to be removed to prevent resin degradation. [9]

Properties of (PET) produced using DMT (Dimethyl Terephthalate) and EG

(Ethylene Glycol) are different from that produced using TPA (Terephthalic Acid) and

EG (Ethylene Glycol). The crystallization behavior of PET is affected by the presence of

side products from polymerization. Exposure to oxygen during resin processing also

leads to degradation of PET. [9]

27
CHAPTER - III

3. Materials and Methods

The importance of material selection, as discussed in the previous sections,

motivates the need to have a standard selection procedure. At industrial production

levels, manufacturers benefit from processing techniques that are different from others.

Therefore, the selection process should check for processing ability of the resin,

mechanical properties, and the resin chemistry (such as AA content). Lack of scientific

studies on samples from actual processing conditions, provoked us to carry out this study

using the samples processed on industrial scale machines, as they involve significant

temperature and rheological gradients not simulated in the typical laboratory. This would

allow us to understand the range in behavior of PET and also helps in forming a record of

the different PET resins available in the market.

3.1. Materials

Out of the vast pool of PET bottle grade (0.70 0.85 dL/g I.V.) resins available,

samples of 14 different resin were collected from industry relevant test runs on the

worlds fastest injection molding machine, Husky (HyPET 500) and stretch-blow molder.

During the test run, the processing parameters were modified to obtain a product with

optimum properties. Test runs were carried out on industrial scale machines. As

discussed previously, the microstructure of the PET resin changes during different

28
processing stages. Samples obtained were PET resin pellets - semi-crystalline form,

preforms amorphous form, and film samples from bottles semi-crystalline state.

Resin pellets are expected to have the highest percentage of crystallinity of all the

samples.

Preform samples used in the present work were processed at a cooling time of 1.4

sec, making them completely amorphous (confirmed by DSC measurements, and

diffraction scans). Thin film (125 m) samples were obtained from the stretch-blow

molded bottles processed at a stretch rate of 13 s-1 on a blow molder. In order to sort out

the resins based on their mechanical properties, different approaches were taken.

3.2. Mechanical Testing Techniques

Material properties such as Youngs modulus, tensile strength, hardness, and yield

strength can be obtained through tensile, compression, and other mechanical testing or

indentation techniques. Use of a particular technique is based on the samples attributes,

measurement resolution, and the type of properties needed. Techniques such as Micro

and Nanoindentation provide only local properties, whereas a tensile test on a universal

tensile tester provides the material bulk properties. In the present work, based on the

available samples, the bulk material properties were obtained through tensile testing and

the local properties were examined through Nanoindentation. The importance of the

local properties is discussed in the later section.

3.2.1. Tensile Testing Method

To obtain the bulk material properties, preform samples were tested on an Instron

5582 model universal tensile tester. Preforms produced through injection molding have a

29
tubular shape. However, their finished ends make them difficult to mount on standard

Instron grips. The shape of the preform is shown in Figure 11 using a cross-sectional

CAD model. Taking the advantage of being tubular, tensile tests were designed

according to the ASTM D638 standard. Based on section 6.2 of D638 Standard Test

Method for Tensile Properties of Plastics, the preform is considered a rigid pipe [68].

The ASTM standard details the procedure for testing rigid plastic pipes. Gripping the

preform was non-trivial primarily because of its closed design at one end.

10 mm

Figure 11: Preform and its cross-sectional view from a CAD model

3.2.1.1. Fixture Design

A fixture was developed to accommodate the preform-shape with minimal

alteration. Design of the fixture was based on material properties obtained through an

tensile test of a PET film sample taken along the radial direction. The top part of the

fixture was designed to take advantage of the perform shape and grip at the taper portion.

Holding the bottom of the preform was challenging. A method that worked was to drill

30
through the preform and insert a pin from inside; then a cup shape support screwed onto

the pin from the bottom and supported the preform from outside, as shown in Figure

12 (b). In order to avoid the preform slipping from the grips, a Devcon plastic welder

epoxy with a load capacity of 24 MPa (3500 Psi) was used between the pin and preform.

This final design has been refined over a period of trials with improvements to the initial

fixture models (Appendix 3). The main hurdle here was the slipping of the pin due to

adhesive failure during the test. Devcon plastic welder worked effectively out of other

commercially available adhesives including Super glue, 2-Ton Epoxy from TAP

Plastics and Loctite epoxy of 17 MPa (2500 Psi) which failed to withstand the

maximum load during the tensile test. When the epoxy was applied over an insufficient

area, the preform would slip out of the pin during the test process, not allowing to test till

failure. A bottom cup was introduced to the fixture and the length of the pin was

increased to increase the gripping area. The preform samples to be tested were machined

carefully to avoid damage, such as indents and groove formation from lathe chuck, for

accommodating the pin which is later glued to the sample and allowed to cure for around

20 to 24 hrs.

31
Connect to
Top Platen

Sample with
Grip for
uniform cross-
holding
section (Preform)
Samples

10 mm In to Bottom
Grips

Figure 12: (a) ASTM test sample specifications[68], (b) Cross-section view of the
fixture with sample

The final design presented in Figure 12 (b) shows all the components in their

relative sizes with respect to the preform. With an improved success rate compared with

all the earlier designs, the rate limiting step is the curing time of the adhesive. To

compensate this, multiple pins were fabricated to carry out the testing discussed here.

Two parts of the top fixture are made of stainless steel (SS 304), the bottom pins are

made from cold worked steel.

32
To load cell

Fixture

Preform with
reflective tabs
25 mm

Bottom cup

Instron grips

Figure 13: Preform mounted on materials tester using the designed fixture.

3.2.1.2. Testing Procedure

Preform samples were collected from a single injection shot of the 144 cavity

mold during each resin test run at optimized and stable processing conditions, typically

24 hours after setting up injection. A random set from the 144 samples were collected for

testing and machined on a lathe using its tailstock with an 8.33 mm (0.328) diameter

drill bit. The preforms were held gently on the lathe headstock with the open end

towards the tailstock to avoid impressions. Devcon plastic welder comes as epoxy and

hardener; has a working time of 4 min and a handling time of 15 min. The samples were

set for curing for a period of more than 20 hrs, to avoid bond failure during the test

process. Properly cured preform samples were tensile tested using the fixture at a

33
displacement rate of 5 mm/min. The strain rate was selected as drawing of PET is

independent of rate at and below 5 mm/min [35, 69] and according to the standard ASTM

D638. Preforms were tested till failure. Failure of the preform samples occurs near the

neck portion. Holding pins used with these samples are ready for reuse after removing

from tested parts and cleaning it with acetone to remove adhesive residue.

c
b

Figure 14: Complete test setup with Laser Extensometer, (b) sample and fixture
assembly, and (c) Laser front panel.

As the test sample cross-section changes along its length, care is taken in

developing the test to have a uniform cross-section gauge length. The cross-section area

for the uniform cylindrical section was calculated as 121.3 3 mm2. The cross-section

dimensions were measured by bisecting the preform along the length and also across the

length, to have actual dimensions compared to the drawing dimensions. These were

verified with the random samples used during the test.

34
A Laser extensometer model LE-05, manufactured by Electronic Instrument

Research was used for strain measuring. In the present work, for preform testing, settings

with a full scale of 5 mm, target distance at 380 mm and averaging of 64 scans for a data

point were used. It has a measurement range of 8 to 127 mm with a maximum resolution

of 0.001 mm. The laser extensometer is equipped with standard RS-232 port and an

analog port for transferring the data. The Laser extensometer gives a DC output of 10

volts and does not have an inbuilt system for data recording.

In order to have a simplified, faster data recording system, a Keithley data logger

Model 2701 with plug-in switching and control module 7706 was used. The 2701 data

logger has a buffer memory of 450,000 readings and can accommodate two modules.

Module 7706 is an all-I/O module with a capacity of reading 20 analog channels. The

data logger can be interfaced with the computer through an RS-232 port or Ethernet

cable, making it more flexible. It can be controlled through an add-in program for Excel

called ExceLINX or from the front panel or through commands given over the IE

browser diagnostic page. Data from the test was recorded using the data logger, where

load data was collected from the Instron auxiliary output and extension from the Laser

extensometer analog output. The data logger can record up to a maximum of 50 data

points per second per channel. Data in this work was recorded at every 0.1 sec (10 per

second per channel) interval to have adequate data points, decided based on the total time

usually the experiment takes (20 min).

35
3.2.1.3. Strain Measurement Techniques

The load extension curve obtained from the crosshead includes the machine,

fixture, and sample compliance. In order to have true material properties, we need a

stress-strain curve free of compliance. Machine and the fixture compliance can be

corrected by using strain gauges or other external strain measurement techniques. To

compensate the machine and fixture compliance, contact extensometers was tried

initially; but the sample showed brittle failure during loading originating at the sharp

edges of the contact extensometer placed on to the preform. Thus, the contact

extensometer acted as stress concentrator lowering the apparent yield stress.

A non-contact Laser based strain measurement technique was used to avoid

sample damage. In the laser extensometer, a laser scan line is focused onto two targets.

These targets were two self-reflecting glass bead tapes mounted on the sample gage

section to measure strain. A schematic view of the basic components present in the laser

extensometer, sample and target position are shown in Figure 15. A top to bottom line

scan was generated using visible light laser from a diode transmitter through a two-facet

rotating mirror. As the laser scans along the sample, the amount of energy reflected back

will be more at the reflective tape than from the rest of the sample. The reflected beam is

analyzed using position sensitive detectors or edge detectors and a video amplifier along

with an inbuilt algorithm for calculating the two angles marked in Figure 15, one between

the reflectors, second from one end to the first reflector and the extension from these

angles. [70, 71]

In order to improve the accuracy of the extension measurements, smaller

measurement ranges can be selected by changing the settings from the Laser

36
extensometer front panel. Along with the target distance, the distance between the

sample and Laser can be changed for higher resolution measurements. The signal-to-

noise ratio is another critical factor in electronic devices, which can be improved in this

case by increasing the averaging rate for collecting a single reading. The initial distance

between the two reflecting tapes can be offset to zero, which allows direct recording of

the change in length. The reference positions for making the measurements can be

selected as the top ends of the two reflecting gauge marks for single deformation zone

samples or between four marking for multiple zone measurements.

+ 40 Deg

Edge detector & Specimen


Video amplifier/PSD Laser diode
module
A

B Gauge
marks

Photo detector

Receiver Lens Rotating mirror

- 40 Deg

Figure 15: Schematic of Laser Extensometer parts, basic operation, and tensile
sample with reflective gauge marks [70, 71]

37
3.2.2. Nanoindentation Technique

Hardness measurements through indentation is one of the earliest techniques

(Mohs hardness scale, 1822 [72]) for finding material properties. Hardness is defined as

the resistance provided by the material against plastic deformation. An indenter of a

definite shape is used to make an indent on the material under investigation, based on the

depth of penetration for a certain load, its hardness is calculated. Based on the length

scale at which the indent is made, the technique can be classified into Macro-, Micro-,

and Nano-hardness testing. Techniques through which material properties are obtained

by making an indent at the nano-scale is called Nanoindentation. In conventional

indentation, the area of the indent is measured directly from the residual indent. During

Nanoindentation, the depth of the indent is continuously monitored a tip area function is

used to find the area of the indent. Thus, nanoindentation is also a form of depth-sensing

indentation (DSI) or instrumented indentation testing (IIT). [72]

An instrument capable of producing nano scale displacements and applying loads

on the order of 1000 mN maximum is called a Nanoindentor. A nanoindentor provides a

load-displacement curve for every indent, material properties such as hardness can be

determined using the residual displacement and youngs modulus from the slope of the

unloading curve. A typical load-displacement curve from a nanoindentor is shown in

Figure 16. In order to prevent overestimation of the material hardness, the elastic

contribution needs to be removed from the total displacement (hfinal). Plastic depth is

defined as the depth in contact with the sample under load. Modulus (E) of the material

is calculated by knowing Poissons ratio () and using the following equations 3, 4. Here

38
dP/dh is the slope of the unloading curve, Er is the reduced modulus of the indenter and

sample system, Eo and o are indenter properties.

1
dP 2 2
= DEr (3)
dh

1 1 2 1 o2
= + (4)
Er E Eo

Indentation Curve

Creep
Load

Slope =
1/compliance
Loading

Unloading

hfinal hplastic
Depth

Figure 16: Typical Load-Displacement curve from a Nanoindentor[73]

The above described process is valid only within the assumption that the material

is elastically isotropic and elastic deformation is the only mode during unloading. This

is not the case for soft and viscoelastic materials like polymers, leading to uncertainty in

the measurements. It is hard to measure the modulus of soft polymers (E < 1 GPa) using

nanoindentation [74]. Cook et al. [75] discussed the issues related with the indentation of

polymers and proposed a viscoelastic-plastic indentation procedure. In the VEP model,

the material is subjected to indentation with extended loading rates for calculating the

39
material properties over time. The indentation depth is more than the normal in this

process. VanLandingham et al. [74] found that the modulus values provided through the

nanoindentation technique provides higher values than that of the ones from tensile

testing. The viscoelastic behavior during the unloading process causes these relatively

high modulus values.

Different indenter types are available for carrying out nanoindentation on a

material. Selection of the indenter is based on the type of material under investigation

and maximum load used for testing. Material properties can be obtained using different

indenters based on the concept of geometrical similarity [72]. For polymer materials, a

spherical indenter is preferred because its blunt shape would not damage the polymer.

Kumar et al. [76] have studied the usage of a spherical indenter for linear viscoelastic

materials. They compared the viscoelastic response of PMMA from nanoindentation and

FE simulations and found them to correlate with experimental values. Lu et al. [77] has

proposed a method for calculating the relaxation modulus of time-dependent materials

using nanoindentation. Ion et al. [78] has studied micron-scale indentations on

amorphous and drawn PET surfaces. They found a decrease in the modulus from

undrawn to drawn and mentioned that this will hold if the indentations were carried in the

direction of drawing.

3.2.2.1. Nanoindentation Method

The choice of nanoindentation is based on an allowance to test small samples.

PET samples collected from the test runs were spherical or cylinder beads of few

millimeters in radius (resin pellets), amorphous thick films (preforms), and thin (125 m)

films (walls of blown bottles). Each of these samples represents the PET resin at

40
different stages of industrial processing. These samples will be helpful in understanding

the changes to the mechanical properties from stage to stage. They also represent

different micro-structures, resin pellets with high crystallinity content, preform being

amorphous, and bottles semi-crystalline. An amorphous block sample was cut from the

preform near the neck portion for nanoindentation. Film samples of 125 m thickness

and about 3 x 4 mm2 were collected along the length of the bottle as shown in the Figure

17.

1
2 3 4

Figure 17: Location of biaxial stretched films samples collected along the length of a
bottle.

Nanoindentation was carried on a TriboindenterTM 900 instrument from Hysitron

Inc., with a displacement resolution of 0.0002 nm and a minimum load of 1 nN. The

maximum load that can be applied is 30 mN. A spherical indenter made of diamond,

with a 5.02 m radius and 60o angle was used for testing in the Quasi-static mode.

Samples were attached to a metal disk using adhesive tape to prevent air trapping

between the sample and disk, and mounted on to a magnetic base. The instrument is

controlled using the Triboscan software. The transducer along with the focusing

microscope was mounted on a granite base to isolate it from vibrations, shown in Figure

18. The entire setup was enclosed in a chamber isolating it from the outside

41
environment. PET samples were tested under load control indentation to a maximum

load of 1 mN at a loading and unloading rate of 0.02 mN/s.

Transducer Microscope

Positioning stage

Figure 18: Triboindenter, important components [79]

42
3.3. Crystallinity Measurement Techniques

Degree of crystallinity is considered as the single most important characteristic

of a polymer in determining its mechanical properties [80]. As a semi-crystalline

polymer, PETs properties are governed by crystallinity levels. Polymer crystallization is

largely classified into three different groups: (a) crystallization during polymerization, (b)

crystallization induced by orientation (also called strain-induced crystallization), and (c)

crystallization under quiescent conditions [81, 82]. A detailed classification is made by

Ratta [81], as a part of his dissertation. Out of all the crystallization types, strain-induced

crystallization is considered to be important. Because of this importance, several

methods were developed for measuring the crystallinity in polymers. Crystallization

behavior, crystallinity content characterization was studied using X-ray diffraction,

infrared spectroscopy, and density as early as the 1960s [83, 84]. Some of the

techniques such as AFM or SEM help in visualizing surface details (crystallinity),

where as other techniques can provide bulk information, listed below:

Density Method

Differential Scanning Calorimeter (DSC, MDSC)

Spectroscopy (Infrared, Raman)

Microscopy (AFM, TEM, SEM)

X-ray Diffraction (SAXS, WAXS)

3.3.1. Density Method

A change in crystallinity leads to a different structure affecting the density of

polymers. By knowing how the density changes with processing, we can control the

43
properties. In order to measure density a standard density gradient column (CCl4 -

heptane mixture at 23oC) calibrated with samples of known density would typically be

used. In this method, standard densities for 100% crystalline and amorphous PET

samples are used to find out the crystallinity of the test sample using equation 5 [6].

c sample a
X c = (5)
sample
c a

Here, c is the density of a 100% crystalline sample and a, density of an

amorphous sample. While using the density method, samples should be void-free. Micro

voids are formed due to improper processing conditions such as higher stretch ratios than

required and stretching at lower temperatures. Although it has high accuracy, the density

method is not preferable because of the handling difficulties associated with chemicals. [6]

Another way to map the density is based on local/micro hardness values which

are dependent on the chain packing density. Variations in chain packing density can be

attributed to changes in chain orientation, isomer conformations, or physical aging. A

chart showing density and its corresponding crystallinity values with two different

reference c values is shown in Figure 19. The most accepted crystalline density is

1455 kg/m3.

44
Figure 19: Crystallinity and corresponding density, based on two different c
values.[6]

3.3.2. Differential Scanning Calorimeter (DSC)

Differential Scanning Calorimetry is one of the most used techniques for thermal

analysis of polymers. It involves the measurement of heat flow into or from a sample

during isothermal heating or cooling processes. The DSC is an instrument which can

monitor even slight changes in the heat flow values during the heating or cooling process.

The crystal domains present in a polymer tend to absorb more heat during the melting

process; this will be reflected in the output curve. A DSC can be used to measure the

glass-transition temperature of the curve and crystallinity content. Crystalline content of

a sample can be calculated using equation 6, by knowing the heat of fusion and heat of

cold crystallization [80, 85].

45
H fusion H cold _ crystallization
XC = (6)
H co

Where Hco is the heat of fusion for 100% crystalline polymer and is taken as

140.1 J/g [86]. To have a more accurate measurement, a new method was developed

called Modulated temperature DSC (MDSC). This method employs a linear heat rate

with a sinusoidal temperature oscillation [6]. Vilanova et al. [87] used DSC to study

crystallization kinetics of low molecular weight PET under isothermal crystallization

conditions. Bashir et al. [6] has compared the DSC results between the standard and

modulated methods on PET samples and found that MDSC is suitable for measurement

on film samples.

Sample requirements for DSC measurements are very small (10-30 mg), therefore

the samples used for testing should be selected such that they represent the sample or

section of interest. A DSC instrument is calibrated for its temperature and heat flow

using an Indium reference, as it has the least uncertainty for the transition temperature

[88]. Crystallinity of the samples under testing were measured using a Perkin Elmer

DSC 8000 setup, using a method where the sample were heated from 50oC to 300oC at a

heating rate of 20 o/min, held at 300oC for 2 minutes and cooled back to 50oC at a rate of

200 o/min. The samples are reheated back to the same temperatures for obtaining the

second heat curve, useful for identifying glass transition temperature (Tg).

3.3.3. X-ray Diffraction

X-ray diffraction is a technique capable of providing sensitive information

regarding the changes at a molecular level. The earliest crystallography studies of PET

were reported back in the 1950s by Johnson et al. [83]. Crystalline PET is made up of a

46
triclinic unit cell [40]. Diffraction scans of a semi-crystalline PET sample have three

reflections at 17o, 22.5o, and 25.5o angles for Cu radiation corresponding to (100), (010),

(110) crystal planes [83], represented in terms of d-spacing 5.2 , 3.9 , and 3.4 . A

large hump created due to the scattering of X-rays by the amorphous content present in

the sample is overlapped by the crystalline peaks and the shape is independent of

degree of crystallinity, yet the relative area under the hump to the diffraction peaks is

proportional to crystallinity [83]. PET crystal orientation is also characterized in terms

of the angle between the c-axis of the unit cell and the drawing direction using the

normal to the lattice plane (105) with the corresponding reflection at 43o (2.1 ) [40]. In

the case of film samples, the in-plane structure can be inferred through an equatorial

scan about an axis parallel to the surface normal [89]. An equatorial scan is obtained

by aligning the fiber axis or film longitudinal direction with the diffractometer axis and

scanning on the surface parallel to the axis, which is different from that of a Meridional

scan taken over surface perpendicular to axis (shown in Figure 20).

M
2
= 90o
Fiber
M
A

E
E 2
= 0o

M : Meridional scan ( = 90o)


E : Equatorial scan ( = 0o)
A : Azimuthal scan
: Azimuthal angle (Chi)

Figure 20: Representation of Meridional and Equatorial scans for fiber sample [90]

47
In a diffraction experiment, the strength of the diffracted beam is important to

distinguish between the peaks and noise. Based on the sample properties (geometry,

absorption constants) and the type of data of interest, different kinds of studies can be

conducted such as transmission, reflectance, small angle, and wide angle X-ray

diffraction.

Researchers like Mahendrasingam et al., Jabarin et al., and Schmidt et al. have

used X-ray diffraction to study the molecular orientation and also the effect of factors

such as temperature, and strain rate on the crystallization of PET. Diffraction can be

carried out in-situ, which allows having a clear visualization of the chain packing during

the strain-induced crystallization process. Using the high energy X-rays available at

synchrotron sources, complex mechanisms can be explored. Through diffraction, the

change in the molecular orientation can be tracked by measuring the orientation order

parameter < P2 (cos) > , where is defined as the angle between the molecular axis

and specimen axis. The orientation distribution for polymers and liquid crystals can be

obtained from the azimuthal profile of an arbitrary reflection [41]. Therefore, X-ray

diffraction is a promising technique for understanding the molecular dynamics that occur

during the transformation of PET from phase to phase.

3.3.4. Spectroscopy Techniques

Spectroscopy is defined as the study of the interaction of radiation with matter.

The nature of these interactions between the materials and radiation change from material

to material. Spectroscopy covers a broad range, and is classified based on the type of

radiation used such as UV, visible light, Infrared, and Raman. In the field of polymers,

spectroscopy can be used for measuring the changes in the orientation and molecular

48
conformation of polymers, which help in estimating their mechanical properties [91].

Infrared radiation is sensitive to the rotation and vibration of molecules. A molecule

under vibrational motion is considered Infrared active if the dipole moment changes or

Raman active if the polarizability changes. Therefore, a molecule under a certain

vibrational mode can be active for either IR or Raman spectroscopy. Each material will

have a unique spectrum, called the finger print region. Tobin et al. [92] indentified most

of the peaks wave numbers (cm-1) present in an Infrared spectrum with the individual

vibrations of different bonds present in PET. Daniels et al. [93] listed them with respect

to their wavelengths. Spectroscopic measurements directly would not provide the

orientation function as provided by diffraction. Ward [91] in his work has listed the

orientation functions for different spectroscopic techniques and the manner through

which they are measured.

Infrared spectroscopy can be classified as two different techniques based on way

the measurements are made; they are absorption spectroscopy and reflectance

spectroscopy. Miyake et al. [94] studied the effect of crystallization on infrared

spectroscopy. Fourier transform IR spectroscopy helps to identify and quantify the

spectral changes associated with trans and gauche rotational isomers of ethylene glycol

residue present in PET. These conformations are identified with their unique band

intensities and normalized with the phenyl ring vibration intensity band (1410 cm-1),

because it would not get affected by the structural changes. Jabarin et al. [ 26]

developed a method through which they calculate the modified structural factor for

trans (1340 cm -1 ) and gauche (1370cm -1 ) rotational isomers. Based on the intensity

changes in the spectrum after normalizing, percentage of crystallinity is determined.

49
Andanson et al. [25] has used attenuated total reflection FTIR spectroscopy for in-situ

degradation studies of PET during methanolysis. Smith et al. [34] used an ATR-FTIR

spectroscopy technique for mapping the biaxial orientation in PET bottles.

FTIR spectrum in the finger print region is shown in Figure 21 for different

samples. The plot shows absence of peaks at 970 cm-1 and 1470 cm-1 wavelength in the

spectrum for amorphous block sample. The peak representing phenyl ring is at 1408 cm-1.

723.19
Welman-block
1713.28

Welman-pellet-edge 2

1242.60
Welman-block
Welman-pellet-edge

1096.65
Welman-pellet-side
Welman-pellet-edge 2
Welman-film

Welman-pellet-edge

1116.77
Welman-pellet-side

1017.37
Welman-film

872.14
1338.72
1408.27

970.53
1039.67

846.97
1174.82

792.70

631.59
1504.89

1454.71
1577.87
1615.42

1800 1600 1400 1200 1000 800


cm-1

Figure 21: Finger print region of an FTIR spectrum, for PET resins with different
microstructures.

Raman spectroscopy was used by Melveger et al. [95] to track the crystallinity

changes in PET fibers drawn at different draw ratios. Ellis et al. [96] studied PET

orientation and crystallization processes using FT Raman spectroscopy. The crystallinity

content of PET from a Raman scan can be estimated by the ratio of bands at 1095 cm-1

and 1120 cm-1.

50
Figure 22: Change in the peak ratio corresponding to crystallinity along the length
of a PET bottle [97].

3.3.5. Microscopy Techniques

The structure of polymers can be visualized through different imaging techniques

such as TEM, SEM, and Scanning Probe Microscopy. SEM and TEM have their own

limitations for obtaining the structural details. Scanning probe microscopy, a relatively

new branch of microscopy provides new and powerful tools that can image at the nano

level. One such tool belonging to the scanning probe microscopy family is the Atomic

Force Microscope (AFM). AFM evolved in the process of developing cantilever

deflection monitoring mechanisms for STM. In AFM, surface imaging is obtained using

attractive and repulsive forces between a sharp tip and the sample surface. AFM is

capable of generating atomically resolved images by displaying deflection of a cantilever

as a function of its position. This makes it useful for imaging non-conducting materials

such as polymers up to the nano level. [98]

51
In the AFM, vertical deflection of the sample is measured as the tip moves along

the sample surface by measuring the change in position of a reflected laser beam that falls

on the position sensitive detectors (PSDs). Selection of a cantilever tip is critical to

record the desired tiny changes of the specimen surface. Mechanical properties play an

important role in tip selection. The concept behind the selection of cantilever is based on

the physics of atomic interactions. As explained by Dror [99], a typical atom vibrates at a

frequency () of 1013 radians and has mass, m = 1025 kg; therefore a spring constant

equivalent for an atom is given by k = 2m = 10 N/m. This provides the maximum

spring constant value for an AFM cantilever to avoid any damage to the sample.

Choosing materials like Si, SiO2, and Si3N4 will be advantageous as cantilevers having as

low as a 1 N/m spring constant and several tens of kHz in resonance frequency.

Cantilevers with these mechanical properties are good enough to permit rapid raster

scanning. AFM Imaging of a hard sample is based on surface topology whereas soft

sample imaging is dominated by viscoelastic properties [98].

In-plane mechanical properties of a material can be measured using AFM by

small lateral oscillation, which are used for material characterization based on shear force

interactions. These can be carried out using higher excitation frequencies near to the

torsional resonances (TR) of the cantilever [100]. Advances in AFM, such as a heating

stage helped in carrying out polymer evolution studies in-situ up to 250oC. AFM can be

operated in three different modes, they are: contact mode, non-contact mode, and tapping

mode. Early work on polymer imaging was carried out using contact mode. Later it

became evident from works of Gould et al. that contact mode operation of AFM for

polymer imaging can modify the surface or even damage the surface of softer samples

52
[101, 102]. Tapping mode operation avoids the effect of lateral forces that exists in

contact mode. In tapping mode, tip and sample interactions are based on several factors

like cantilever vibration amplitude, cantilever stiffness, sample stiffness, tip shape,

viscoelasticity and adhesion. Based on the amount of force that is applied in tapping

mode operations, they are classified as hard and light tapping. Imaging samples with

different materials, results in change of contrast of the final image based on density and

stiffness, this is more evident in the hard tapping mode. When high resolution imaging is

of importance, light tapping is best but it is difficult to achieve because of the capillary or

adhesive forces interfere with the cantilever vibration amplitude.[102]

Gould et al. used AFM for nano level imaging of PET. They came to

differentiate crystalline and amorphous phases based on the local roughness values.

Large variation in the values of mean roughness (Ra) and maximum height (Rm) indicate

crystalline and amorphous regions. In tapping mode, imaging of PET does not provide

enough phase contrast to distinguish crystalline and amorphous regions. Cupere et al.

studied surface crystallization of PET films cast at 85oC. They imaged formation of

spherulites on the top layer of PET while the rest of the sample remained amorphous.

Obtaining productive information using AFM for characterizing polymers needs

thorough knowledge of the physical and chemical reactions (if any) involved between the

tip and surface of the sample along with various key factors such as the signal-to-noise

ratio, imaging force, and scan speed.

In the present work, PET sample (resin pellet and film) surfaces were scanned on

an AFM (Asylum Researchs MFP-3DTM AFM, Veeco Multimode SPM (AFM) and

AC240TS scan tip (Si) were used). Surface topography of PET resin pellet, at 500 nm

53
scan size and scan rate 0.999 Hz is shown in Figure 24. In order to measure the surface

roughness, unpolished samples were used. AFM scans with scan size from 500 nm to 10

m were tried to have defect (presence of foreign particles and regions with relatively

high surface roughness causes artifacts) free images. The scanning time will be based on

the scan parameters (size, rate, resolution). Selecting the scan parameters is critical and

can help in minimizing the artifacts. Crystallinity measured through this technique only

represents the surface and not the bulk. Also, data from polished surfaces does not

represent real surface data. Considering these issues, AFM is not considered further for

measuring crystallinity in this work.

mirror A B Sensor output , F


C D
cantilever
Position sensitive
photodetector

= 10 o- 15o Error =
Actual signal +
sample
set point
Courtesy: Asylum Research

Piezoelectric Z
Y Feedback loop
Scanner

Figure 23: Working of an Atomic Force Microscope [103]

54
Figure 24: Example AFM scans of the surface of PET resin pellet showing
molecular chains (using Veeco AFM at Microscopy Laboratory, OSU)

55
CHAPTER IV

4. Results

4.1. Tensile Test Results

Material properties such as Youngs modulus and yield strength were obtained

from the tensile test of preform samples. Multiple samples (7 samples) were tested from

each of the 14 different kinds of resins collected. A usual load vs. extension curve

obtained from the experiment is shown in Figure 25. Modulus is obtained from the slope

of the Engineering stress-strain curve in the elastic region. Yield stress values can be

obtained by three different ways:

Stress value at the intersection of a line drawn parallel to the elastic region

from 0.2% strain (point c, from Figure 27) with the Engineering stress-strain

curve.

Stress from the point at which the linear region slope changes (points a and

b, from Figure 27).

Stress from the intersection of a vertical line drawn at 0.5% strain (point d,

from Figure 27) with the Engineering stress-strain curve.

56
Yielding
Hardening

Plastic deformation

Elastic zone Failure

Figure 25: Load vs. Extension (crosshead) curve features related with the sample
(preform) state

In the present work, yield stress values are collected using three different

approaches, based on the laser strain values. A plot showing yield stress values through

different approaches are shown Figure 27. Yield strength obtained through 0.5% offset

(d) approach was not considered due to the relatively low values. Yield strength values

obtained from different approaches follow a similar trend, are shown in Table 4. During

the tensile test, the preform sample was stretched up to 400% of its initial length.

Different features of the loadextension curve of a preform can be related to the sample

state as shown in Figure 25. The difference between the strain values measured through

the cross-head and Laser extensometer are shown in detail in Figure 26.

57
Table 4: Yield Strength of resins through different approaches
Yield Yield
Yield
Resin Standard Strength Standard Strength Standard
Strength
Name Deviation (MPa) Deviation (MPa) Deviation
(MPa) (a)
(b) (0.2 %)
B 0.76 38.1 0.5 30.7 0.3 29.5 0.5
B 0.82 39.5 0.6 31.2 0.3 29.4 0.5
K 39.7 0.6 31.4 0.4 29.7 0.6
H 39.2 0.5 31.6 0.2 29.2 0.2
J 39.1 0.5 31.7 0.3 29.7 0.7
E 0.77 40.1 0.2 32.4 0.7 30.3 0.7
B 41.0 0.5 33.0 1.4 31.0 0.1
E 0.84 41.3 0.9 33.1 0.7 30.6 0.6
I 41.2 0.6 33.2 0.2 31.3 1.1
F 42.1 0.6 33.5 0.4 31.1 0.9
G 41.4 0.7 33.7 0.2 31.4 0.5
A 41.2 0.6 33.7 0.6 31.5 0.5
D 43.6 0.6 34.2 0.4 31.8 1.1
C 42.9 0.3 34.3 0.1 31.6 0.1

40
Engineering Stress (MPa)

35
30
25
20
15 Laser Strain
10 Cross-head Strain
5
0
0.1 0 0.150.05 0.2 0.25
Engineering Strain
Figure 26: Cross-head and Laser Strains up to 0.25 mm/mm Engineering Strain

58
50
45 (a)

Engineering Stress (MPa)


40
35 (b)
30 (c)
25
20
15
(d)
10
5
0
0 0.01 0.02 0.03 0.04 0.05
Engineering Strain

Figure 27: Representation of approaches used for Yield Stress, (a) and (b) change in
linear region slope, (c) for 0.2% strain and (d) for 0.5% strain.

Based on the Youngs modulus values, the resins are ranked and plotted in their

respective order. This ranking is based on the average modulus values. A plot showing

the absolute and relative modulus values with their relative ranking for all the 14 resins

are shown in Figure 28. Error analysis is based on the standard deviation of the values

for the each sample set. Relative modulus of each resin is calculated with respect to the

maximum modulus (resin C). A typical preform failure is shown in Figure 31, along with

an abnormal failure due to brittle fracture.

59
2700
110

Young's Modulus (MPa)


2600

Relative Modulus (%)


2500 105
2400 100
2300 95
2200 90
2100 85
2000 80
1900 75
B B J K I F E H G E D A B C
0.82 0.76 0.77 0.84
Resins

Figure 28: Young's Modulus of different resins, all resins are of 0.80 I.V. unless
otherwise marked.

Similarly, yield strength values are also plotted for different resins based on their

modulus rankings as shown in Figure 29. Also, the resins are ranked based on their yield

strength values (Figure 30). From the modulus plots, it was clear that there is about 13%

increase in modulus from left to right.

Relative Strength (% w.r.t. C)


35
34 100
Yield Strength (MPa)

33 96
32 92
31
88
30
29 84

28 80
B B J K I F E H G E D A B C
0.82 0.76 0.77 0.84
Resins

Figure 29: Yield Strength of different resins (data based on approach (b))

60
36 104

Relative Strength (% w.r.t. C)


35
Yield Strength (MPa) 34 100
33 96
32
92
31
30 88
29
84
28
27 80
B B K H J E B E I F G A D C
0.76 0.82 0.77 0.84
Resins

Figure 30: PET resins ranked based on yield strength

10 mm
(a)
(b) 10 mm

Figure 31: (a) Preform sample typical failure, (b) brittle failure of the sample

4.2. Nanoindentation Results

The load-displacement curves provide a reduced modulus (Er) and hardness

values. Reduced modulus (Er) is based on the deformation of a sample and indenter and

is obtained from the slope of the unloading curve and the tip area function. The Youngs

modulus of PET is calculated using equation 4, properties of the indenter used were

modulus - 1060 GPa and Poissons ratio - 0.17 and for PET, 0.4 [104] (also from the

DSM technical guide on polymers). A minimum of three similar load-displacement

61
curves (similar in slope) were obtained from each sample, for calculating the error.

Moduli for all the three curves are calculated using the reduced modulus (Er) obtained

through equations 3. An average is considered for final modulus, standard deviation is

used for representing the error.

1200

Reading 1
1000
Reading 2
800 Reading 3
N)
Load (

600

400

200

0
0 50 100 150 200 250 300 350 400
Depth (nm)

Figure 32: Load-Displacement curves for PET film sample (resin F)

6.00
120.00

Relative % Modulus w.r.t Resin D


5.00
100.00
Young's Modulus (GPa)

4.00
80.00

3.00
60.00

2.00 40.00

1.00 20.00

0.00 0.00
B 0.82 B I F A C D
Resins

Figure 33: Youngs Modulus of resins from Nanoindentation data of amorphous


samples

62
Additionally, nanoindentation was performed on film samples collected along the

length of bottle as shown in Figure 17, to measure the properties along the length of the

plastic container (bottle) obtained from the test run. Modulus values obtained from the

film samples are compared with that of amorphous sample values are listed in Table 5.

Table 5: Modulus (GPa) values for PET samples of different forms

Amorphous Film Film Film


Resin Name Pellet
Sample (Location 2) (Location 3) (Location 4)

Resin F 6.80 3.72 5.76 6.32 5.60


Resin B 4.62 3.49 3.81 4.01 3.38
Resin B 0.82 3.79 3.31 3.51 4.18 3.40

4.3. DSC Results

Crystallinity of the PET samples under testing was measured using a Perkin

Elmer DSC 8000 setup. A sample DSC curve obtained for an amorphous sample

collected from the preform is shown in Figure 34. A preform sample, a resin pellet, and

film samples from blown bottles were used to measure and confirm their crystallinity

level. It was found that the samples collected from the preform had < 3% crystallinity,

the film samples had about 30% crystallinity, and the resin pellet had the highest

crystallinity (42%), the heat values and corresponding crystallinity for different samples

are tabulated in Table 6.

63
Table 6: Crystallinity for different samples of B 0.82 PET resin
Cold
Melting First Second
Sample - Crystallization Heat of Fusion %
Point Heat Tg Heat Tg o
type o Heat (J/g) @ (J/g) @ 246 C Crstallinity
Tm ( C) (o C) o
( C) o
142 C
Pellet 239.35 81.94 - - 59.6 42
Film 246.41 80.19 - - 44.1 31
Film 247.33 80.14 - - 41.8 29
Film 246.21 80.46 - - 59.6 31
Amorphous 247.14 81.1 80.61 -30.5 35.1 3.3
Amorphous 247.36 81.2 81.9 -30.3 34.3 4
Amorphous 246.87 80.8 80.7 -30.3 34.7 3.1

Tg

Figure 34: DSC curves for PET samples with different micro-structure

64
4.4. X-ray Diffraction Results

Diffraction scans for PET resins samples are obtained using a Cu radiation source,

shown in Figure 35. PET resins sample provides crystalline peaks with amorphous

background, whereas the film samples display a sharp peak indicating strong influence of

texture in the sample. Using Bruker EVATm software, resin pellet sample crystallinity is

quantified as 30%, less compared with that of DSC results. This is due to lack of proper

approach for estimating air scatter contribution to the amorphous halo. The data for

preform sample confirms their amorphous nature (collected from injection molding) used

for tensile testing. Data collected represents material up to a thickness of 0.2 mm on the

sample. This eliminates the chances of having any crystallization on the outer layers,

which come in contact will mold walls at chilled temperatures.

65
Figure 35: XRD scan of PET samples (B 0.82) (preform and pellet samples intensity
secondary axis) (using Bruker diffractometer using Cu-radiation.)

66
CHAPTER V

5. Discussion

Poly (ethylene Terephthalate) with its interesting semi-crystalline microstructure

and wide range of applications is still under explored [8]. Beginning with the

polymerization process itself, there are different variables such as the type, quantity of

metal catalysts used, stabilizers, colorants, and the way they influence the synthesis and

later processing that remains ambiguous. The number of side reactions during the

polymerization process and the presence of these side products such as diethylene glycol

(DEG), dioxanes and short chain oligomers affect resin performance. PET degrades in

the melt stage (above the melting point) due to thermal and thermo-oxidative degradation

processes. These will reduce the viscosity (I.V.) of the resin and also increase

acetaldehyde content. Acetaldehyde provides unwanted flavor to the contents of a bottle,

this makes it critical to keep it low (< 1 ppm). Thermal degradation reactions are

influenced by the metal catalysts used during the polymerization of PET.[8, 105]

In the process of bottle making, PET resin will be heated to the melting

temperature and cooled and reheated to above its glass-transition before it takes the final

shape. Also, the presence of humidity influences the rate of degradation of PET and at

the same time it influences the rate of crystallization at higher temperatures [106]. Much

67
of these changes happen to the resin during the injection molding process itself since the

resin must be melted to inject. PET is also sensitive to photo-degradation. These factors

make the resin properties change after processing.

The mechanical properties of PET have been studied by several researchers as

mentioned earlier, using laboratory prepared samples. Relevance of this data to industrial

products was not well established, as their investigations do not involve all the factors

shown in Figure 10. After availability, resins are generally selected based on their

processing ability and handling properties. Studying the change in properties of PET,

processed at industry relevant conditions is important for understanding its relevant

mechanical behavior. However, today the package engineer is forced to accommodate

the differences between the resins and the possible degradation from resin processing as a

factor of safety. This results in producing over-engineered (heavy) products. In the

process of reducing the polymer usage in products (light-weighting), testing of industrial

samples provides realistic error estimates to base dimensional changes (thickness). With

an increased understanding of the mechanical properties and potential validation with FE

simulations, there is potential for significantly (>25% with tight material tolerances)

reducing the polymer used without compromising performance.

5.1. PET Resins

PET resins used in this work are selected based on their availability. This

selected group of resins covers PET resins with different polymerization processes,

demographic regions, and intrinsic viscosity (I.V.) values (0.74 0.85 I.V.). Resins with

a high and low intrinsic viscosity are also tested to make a better understanding on their

mechanical behavior. All the resins are of 0.80 dL/g I.V. unless specified.

68
5.2. Stability of Mechanical Properties

Youngs modulus of PET was found to range from 1950 to 2600 MPa, over the

group of resins tested. This is in the range of reported values in literature (0.89 GPa 3.5

GPa)[20, 48]. Mechanical properties of the resins are plotted against their Youngs

modulus and yield strength as shown in Figure 36. This plot elucidates the relationship

between modulus and yield of PET, with a ratio of 67.5 (y/E - 0.014-0.015). Brown

[107] in his work has mentioned that there is a strong correlation between yield strength

and modulus of glassy polymers, and gave a range for their ratio 0.015-0.039. Also, this

plot depicts clearly an increase in performance of resins from left to right. Overall, the

average modulus of the resins ranges across 13%. The maximum and minimum give a

range of 27%.

2650

2550

2450
Young's Modulus (MPa)

C
B A
2350 D
E 0.84
G
H
E 0.77
2250 F
I
K
2150 J y = 67.55x + 59.14
B 0.76 R = 0.69
B 0.82
2050

1950
29 30 31 32 33 34 35 36
Yield Strength (MPa)

Figure 36: Plot showing resins with respect to Modulus and Yield Strength

69
2650.00

2550.00 A
E 0.84
2450.00 E 0.77
Young's Modulus (MPa)

K
2350.00 G
D
F
2250.00
C
I
2150.00
J
H
2050.00 B 0.76
B
1950.00 B 0.82
29.00 30.00 31.00 32.00 33.00 34.00 35.00 36.00
Yield Strength (MPa)

Figure 37: Youngs Modulus vs. Yield Strength plot showing the scattered individual data along with error bars based on
standard deviation

70
A detailed plot with individual sample data is shown in Figure 37, with standard

deviation error bars. It can be inferred that a few resins have more scatter in the

properties. The standard deviation error for different resins are plotted in Figure 38;

provides information on the stability in performance of the resins. It is evident from the

mechanical property results that the deviations from the mean were not consistent and

changes with resin. Resins with lower deviation signify better stability than other resins.

This requires a factor of safety to incorporate the material property range with a single

design.

250

200
Standard Deviation

150

100

50

0
F J E C D G B H E B B I K A
0.77 0.76 0.84 0.82
Resins

Figure 38: Standard deviation for Youngs Modulus of resins

Modulus values from the nanoindentation (shown in Figure 33) show a parallel

trend in comparison with the results from tensile test. Nanoindentation results of film

samples (shown in Table 5) illustrate that the modulus change from location to location

along the length of bottle, the highest value being on the walls of the bottle. In the blow

molding process, the stretch ratio changes from location to location on the bottle, this will

71
result in non-uniformly stretched bottles. In the case of film samples (from blown

product) used in this work, the stretch ratio changes from 3.5 to 4.5 along the length.

Stretching of the preform after heating it above the glass-transition temperature

crystallizes PET through a strain-induced crystallization process. With the change in

crystalline quantity and also wall thickness along the length of a bottle, properties

changes from location to location. This can be observed with nanoindentation.

The modulus values obtained through nanoindentation appear higher than that

from the macroscopic testing procedures. The higher modulus is attributed to the

viscoelastic behavior during the quasi-static mode of testing. PET as explained by Ion et

al. [78] in their work, falls in to the category of materials with complex elastic-plastic

deformation. Plastic recovery during the unloading curve affects the measured

properties. The rate of loading and hold time and indentation have a large effect on the

final properties measured in the case of viscoelastic materials [108].

Furthermore, the modulus curves suggest resin mechanical properties decrease

with I.V. (molecular weight). Molecular weight can be related to the intrinsic viscosity

(I.V.) by using the Mark-Houwink equation [11], shown in equation 7, where K and a

were given as 3.72 x 10-4, 0.73 for Mn and 4.68 x 10-4, 0.68 for Mw. A decrease in

properties sometimes observed at higher I.V.s can be attributed to the degradation of

resins during processing. This can be described by the following process. With an

increase in I.V., chain length increases causing chains to shear during the processing, this

leads to chains scission and thermal degradation.

[ ] = KM a (7)

72
Comparing these results with the mechanical properties of PET-Nanocomposites

from literature, as shown in Table 3, the increase in modulus for 1% clay addition is in

the range of property improvements simply from judicious resin selection.

PET samples used here are collected from industrial process; change in sample

dimensions resulted from the processing conditions can be accounted up to 3% variation

in the material properties. Load cell drift is accounted as 2N over the test time, which is

considered negligible compared with the maximum load. Processed resins exhibit

different colors based on the resin chemistry for example, resin B yellow tint, A

green tint, and D blue tint. During the testing, few resin samples exhibit cloudy

(haziness) behavior as they stretch (resins F, D). This behavior can be related to the

increase in crystallinity during the test, termed as the strain-induced crystallization.

Other factors have also been identified. During processing, there is a 1%

variation in preform weights, which leads to changes in the final wall thickness of the

bottle. Also, in measuring the stiffness, the yield strength was observed. Yield strengths

ranged over 11% for the resins tested. Such a range could be exaggerated above Tg and

may lead to differing percent crystallinity in films, assuming the yield strength is

proportional to the molecular alignment factors involved in stress induced crystallinity.

Further investigation of yield at increased temperature is needed to confirm this

parameters impact on film processing.

It is understood that variations will exist in the same basic type of plastic from

different manufacturers and even in different batches from the same manufacturer. For

reduction of manufacturing tolerances and accounting for equipment or manufacturing

variables, checking materials upon receipt is important. Since, in manufacturing the

73
checks must proceed quickly or even inline, automated spot measurement of mechanical

performance will become necessary to support further efficiency improvements and light-

weighting.

2900 A
2800 E 0.84
2700
B 0.80
Modulus (MPa)

2600
2500
2400
2300
2200
2100
2000
1900
1 2 3 4 5 6 7
No of sample

Figure 39: Modulus vs. sampling number

In the above plot, modulus for samples of three different resins are plotted against

their corresponding sample numbers, error bars indicate the change in deviation for the

average modulus with increase in sampling size. Resins E 0.84 and B 0.80 are having

modulus within the deviation range even for a small size, higher sampling size is useful

in case of Resin A to minimize error from sampling size. This confirms the reliability

of present sample size for making an appropriate comparison between the resins.

74
CHAPTER VI

6. Conclusions

The present work testing of PET resins provides the following conclusions:

Mechanical testing of PET resins showed a performance difference based on

modulus from a minimum 13% to about 27% between different manufacturers when

processed under similar conditions on one machine.

Mechanical performance of PET resins can significantly differ with the change in

resin process chemistry.

Preform samples collected from the latest injection molding process have an

amorphous microstructure with less than 3% crystallinity.

Yield strength to Youngs modulus ratio for PET resins (in amorphous form) is

determined as 0.015 across 11 resin suppliers around the world.

PET resin pellets have higher crystallinity compared with the processed resin

samples (preform, film). Material properties are sensitive to intrinsic viscosity (I.V.).

Clearly material selection provides further room for reduction in the amount of

resin used in packaging industry (light-weighting) without compromising performance.

This justifies the need for manufacturers to develop and update a resin mechanical

performance database.

75
Material properties and crystalline quantity of PET films change with change in

their stretch ratios. Understanding the way stretch ratio affects material properties, help

in developing better designs.

Testing of amorphous samples (preforms) is an effective approach for studying

effect of industrial processing on material performance.

76
CHAPTER - VII

7. Future Work

FEM validation is also necessary for adequate industrial usage of these results. A

better understanding of the product geometry variables and optimization of the geometry

design is a next direction for this work.

With increased environmental awareness, industry started using rPET (recycled

PET) in clear packaging applications to a maximum of 25% [109]. Studies on rPET

suggest its low usage towards the difficulties in producing product with fewer impurities.

Understanding the effect of increased rPET content on the process and final mechanical

properties is required to make progress in increasing the usage of recycled PET.

Finally, reducing the polymer usage is only possible to certain extent without

trailing on important packaging attributes such as permeability (barrier properties).

Addition of nanoparticles in the parent material (PET) can be considered to compensate

the trailing on important properties. Choosing nanoparticles optimally can improve

mechanical and barrier properties.

7.1. Impact of Stretch Ratio on Mechanical Properties

Nanoindentation of PET film samples collected from the stretch-blow molded

bottles show that the material properties change along the bottle length, therefore

77
developing an entire map of crystallinity and correlating it with the local properties will

help in understanding the effects of stretch ratio. Strain-induced crystallization is the

dominant mechanism that influences the crystallinity content of the final product. The

mechanism of strain-induced crystallization is strongly based on factors such as stretch

rate, stretch ratio and material temperature. It is necessary to include all the parameters

for developing a useful model.

Crystallinity measurements were made on specific locations along the bottle

length using X-ray diffraction. Crystallinity determined through this approach needs to

be supported by other techniques. To study the micro-structural changes during the

stretching process, X-ray diffraction is a promising tool with feasibility to make in-situ

observations. This approach data will help in developing next generation green bottles

by reducing the plastic usage.

1
2 3 5

Figure 40: Locations used for diffraction scans

78
Figure 41: Diffraction patterns at each location compared with preform sample
(amorphous)

Table 7: Crystallinity (relative to amorphous sample) corresponding to sample


location

Sample
Crystallinity
Location

1 0%
2 19%
3 53%
4 31%
5 20%
6 0%

79
References

[1] C. Fellbaum, "WordNet: An Electronic Lexical Database," MIT Press, 1998.


[2] "Chemistry Daily: The Chemistry Encyclopedia." vol. 2010.
[3] J. Maddox, "NAFTA PET resin & container industry," in Packaging Conference,
Lasvegas, 2010.
[4] "Polymer scenario in 2009." vol. 2010, 2010.
[5] J. G. Rodwan, Jr., "U.S. and International Bottled Water Developments and
Statistics for 2008," International Bottled Water AssociationApril 2009 2008.
[6] Z. Bashir, I. Al-Aloush, I. Al-Raqibah, and M. Ibrahim, "Evaluation of three
methods for the measurement of crystallinity of pet resins, preforms, and bottles,"
Polymer Engineering and Science, vol. 40, pp. 2442-2455, 2000.
[7] "Polyethylene Terephthalate." vol. 2010: Wikipedia, 2006.
[8] T. E. Long, Modern Polyesters: Chemistry and Technology of Polyesters and
Copolyesters: John Wiley and Sons, Ltd, 2003.
[9] K. Ravindranath and R. A. Mashelkar, "Polyethylene terephthalate--I. Chemistry,
thermodynamics and transport properties," Chemical Engineering Science, vol.
41, pp. 2197-2214, 1986.
[10] H. E. H. Meijer and L. E. Govaert, "Mechanical performance of polymer systems:
The relation between structure and properties," Prog. Polym. Sci., vol. 30, 2005.
[11] N. B. Sanches, M. L. Dias, and E. B. A. V. Pacheco, "Comparative techniques for
molecular weight evaluation of poly (ethylene terephthalate) (PET)," Polymer
Testing, vol. 24, pp. 688-693, 2005.
[12] J. M. Therese, E. M. Paul, S. L. Dalmacio, W. C. Lee, and W. L. Cates, "Thermal
crystallization of polyester pellets in liquid." vol. 7674877, U. S. Patent, Ed. USA:
Eastman Chemical Company, 2010.
[13] S. A. Jabarin, "Strain-Induced Crystallization of Poly(Ethylene Terephthalate),"
Polymer Engineering and Science, vol. 32, p. 9, Septermber 1992 1992.
[14] P. Chandran and S. Jabarin, "Biaxial orientation of poly (ethylene terephthalate).
Part II: The strain-hardening parameter," Advances in Polymer Technology, vol.
12, pp. 133-151, 1993.
[15] P. Chandran and S. Jabarin, "Biaxial orientation of poly(ethylene terephthalate).
Part I: Nature of the stress - strain curves," Advances in Polymer Technology, vol.
12, pp. 119-132, 1993.
[16] P. Chandran and S. Jabarin, "Biaxial orientation of poly(ethylene terephthalate).
Part III: Comparative structure and property changes resulting from simultaneous
and sequential orientation," Advances in Polymer Technology, vol. 12, pp. 153-
165, 1993.
[17] D. J. Blundell, D. H. MacKerron, W. Fuller, A. Mahendrasingam, C. Martin, R. J.
Oldman, R. J. Rule, and C. Riekel, "Characterization of strain-induced
crystallization of poly(ethylene terephthalate) at fast draw rates using synchrotron
radiation," Polymer, vol. 37, pp. 3303-3311, 1996.
[18] A. Mahendrasingam, C. Martin, W. Fuller, D. J. Blundell, R. J. Oldman, J. L.
Harvie, D. H. MacKerron, C. Riekel, and P. Engstrom, "Effect of draw ratio and

80
temperature on the strain-induced crystallization of poly (ethylene terephthalate)
at fast draw rates," Polymer, vol. 40, pp. 5553-5565, 1999.
[19] A. Mahendrasingam, C. Martin, W. Fuller, D.J.Blundell, R. J. Oldman, D. H.
MacKerron, J. L. Harvie, and C. Riekel, "Observation of a transient structure
prior to strain-induced crystallization in poly(ethylene terephthalate)," Polymer,
vol. 41, pp. 1217-1221, 2000.
[20] R. Matthews, A. Ajji, K. C. Cole, and M. M. Dumoulin, "Roll and Tensile
Drawing of PET: Effect of Process Conditions on Structure and Properties," in
ANTEC' 98, Atlanta, Georgia, 1998, pp. 1614-1618.
[21] E. R. Dixon and J. B. Jackson, "The Inter-Relation of Some Mechanical
Properties with Molecular Weight and Crystallinity in Poly(ethylene
terephthalate)," Journal of Materials Science, vol. 3, pp. 464-470, 1968.
[22] R. Guzatto, M. Becker da Roza, E. L. Gasparotto Denardin, and D. Samios,
"Dynamical, morphological and mechanical properties of poly(ethylene
terephthalate) deformed by plane strain compression," Polymer Testing, vol. 28,
pp. 24-29, 2009.
[23] G. Venkateswaran, M. R. Cameron, and S. A. Jabarin, "Effects of temperature
profiles through preform thickness on the properties of reheat-blown PET
containers," Advances in Polymer Technology, vol. 17, pp. 237-249, 1998.
[24] A. Silberman, M. Omer, A. Ophir, and S. Kenig, "The Effect of Stretch and Heat
Transfer on the Thermo-Mechanical Properties of PET Bottles," in ANTEC' 98,
Atlanta, Georgia, 1998, pp. 803-809.
[25] J. M. Andanson and S. G. Kazarian, "In situ ATR-FTIR Spectroscopy of
Poly(ethyelene terephthalate) subjected to High-Temperature methanol,"
Macromol. Symp., vol. 265, pp. 195-204, 2008.
[26] E. A. Lofgren and S. A. Jabarin, "Polarized internal reflectance spectroscopic
studies of oriented poly(ethylene terephthalate)," Journal of Applied Polymer
Science, vol. 51, pp. 1251-1267, 1994.
[27] R. J. Caldicott, "Fundamentals of PET stretch Blow Molded Containers," in
ANTEC' 98, Atlanta, Georgia, 1998, pp. 810 - 815.
[28] S. Tang and Z. Xin, "Structural effects of ionomers on the morphology,
isothermal crystallization kinetics and melting behaviors of PET/ionomers,"
Polymer, vol. 50, pp. 1054-1061, 2009.
[29] N. C. Lee, Practical Gudie to Blow Moulding. Shropshire: Rapra Technology
Limited, 2006.
[30] A. Aref-Azar, F. Arnoux, F. Biddlestone, and J. N. Hay, "Physical ageing in
amorphous and crystalline polymers. Part 2. Polyethylene terephthalate,"
Thermochimica Acta, vol. 273, pp. 217-229, 1996.
[31] L. S. Lever, "Crystallinity." vol. 2008.
[32] T. Belha, "Mechanical porperties of semicrystalline polymers." vol. 2010.
[33] S. Venkataraman, E. A. Lofgren, and S. A. Jabarin, "Orientation and Structure
Development of Highly Crystalline and Clear Poly(ethylene Terepthalate)," in
ANTEC' 98, Atlanta, Georgia, 1998, pp. 1629 -1635.
[34] M. R. Smith, S. J. Cooper, D. J. Winter, and N. Everall, "Detailed mapping of
biaxial orientation in polyethylene terephthalate bottles using polarised attenuated
total reflection FTIR spectroscopy," Polymer, vol. 47, pp. 5691-5700, 2006.

81
[35] J. C. Viana, N. M. Alves, and J. F. Mano, "Morphology and mechanical
properties of injection molded poly(ethylene terephthalate)," Polymer
Engineering and Science, vol. 44, pp. 2174-2184, 2004.
[36] A. K. Oultache, X. Kong, C. Pellerin, J. Brisson, M. Pzolet, and R. E.
Prud'homme, "Orientation and relaxation of orientation of amorphous
poly(ethylene terephthalate)," Polymer, vol. 42, pp. 9051-9058, 2001.
[37] S. K. Sharma and A. Misra, "The effect of stretching conditions on properties of
amorphous polyethylene terephthalate film," Journal of Applied Polymer Science,
vol. 34, pp. 2231-2247, 1987.
[38] H. B. Daly, B. Sanschagrin, K. T. Nguyen, and K. C. Cole, "Effect of polymer
properties on the structure of injection-molded parts," Polymer Engineering &
Science, vol. 39, pp. 1736-1751, 1999.
[39] D. J. Blundell, A. Mahendrasingam, C. Martin, D. H. Mackerron, W. Fuller, R. J.
Oldman, J. L. Harvie, and C. Riekel, "Orientation prior to crystallisation during
drawing of poly(ethylene terephthalate)," Polymer, vol. 41, pp. 7793-7802, 2000.
[40] F. Chaari, M. Chaouche, and J. Doucet, "Crystallization of poly(ethylene
terephthalate) under tensile strain: crystalline development versus mechanical
behaviour," Polymer, vol. 44, pp. 473-479, 2003.
[41] R. Lovell and G. R. Mitchell, "Molecular orientation distribution derived from an
arbitrary reflection," Acta Cryst., vol. A 37, pp. 135-137, 1981.
[42] D. J. Blundell, R. J. Oldman, W. Fuller, A. Mahendrasingam, C. Martin, D. H.
Mackerron, J. L. Harvie, and C. Riekel, "Orientation and crystallisation
mechanisms during fast drawing of poly(ethylene terephthalate)," Polymer
Bulletin, vol. 42, pp. 357-363, 1999.
[43] "Polyethylene Terephthalate (PET) for food packaging applications,"
International Life Science Institute2000.
[44] "Golbal Packaging Consolidation: Lead, Follow, or Get Out of the Way," in
Packaging Conference, Las Vegas, 2010.
[45] "Recycling Symbols (U. S.)." vol. 2010.
[46] D. Sabourin, "PET: A Sustainable Package," in Packaging Conference, Las
Vegas, 2010.
[47] M. Stones, "Nestle launches lightweight PET bottle," AP-Food Technology, 2009.
[48] L. V. Todorov and J. C. Viana, "Characterization of PET Nanocomposites
Produced by Different Melt-Based Producation Methods," Journal of Applied
Polymer Science, vol. 106, p. 11, 2007.
[49] E. T. Thostenson, C. Li, and T.-W. Chou, "Nanocomposites in context,"
Composites Science and Technology, vol. 65, 2005.
[50] X. Xu, Y. Ding, Z. Qian, F. Wang, B. Wen, H. Zhou, S. Zhang, and M. Yang,
"Degradation of poly(ethylene terephthalate)/clay nanocomposites during melt
extrusion: Effect of clay catalysis and chain extension," Polymer Degradation and
Stability, vol. 94, pp. 113-123, 2009.
[51] F. Hussain, M. Hojjati, M. Okamoto, and R. E. Gorga, "Review article: Polymer-
Matrix Nanocomposites, Processing, Manufacturing and Application: An
Overview," Journal of Composite Materials, vol. 40, 2006.

82
[52] L. V. Todorov and J. C. Viana, "Characterization of PET nanocomposites
produced by different melt-based production methods," Journal of Applied
Polymer Science, vol. 106, pp. 1659-1669, 2007.
[53] M. D. Sanchez-Garcia, E. Gimenez, and J. M. Lagaron, "Novel PET
Nanocomposites of Interest in Food Packaging Applications and Comparative
Barrier Performance with Biopolyester Nanocomposites," J. of Plastic Film &
Sheeting, vol. 23, 2007.
[54] S. J. Mun, Y. M. Jung, J.-C. Kim, and J.-H. Chang, "Poly(ethylene terephthalate)
nanocomposite fibers with functionalized multiwalled carbon nanotubes via in-
situ polymerization," Journal of Applied Polymer Science, vol. 109, pp. 638-646,
2008.
[55] A. K. Anoop, U. S. Agarwal, and R. Joseph, "Carbon nanotubes-reinforced PET
nanocomposite by melt-compounding," Journal of Applied Polymer Science, vol.
104, pp. 3090-3095, 2007.
[56] R. D. Averett, M. L. Realff, and K. I. Jacob, "The effects of fatigue and residual
strain on the mechanical behavior of poly(ethylene terephthalate) unreinforced
and nanocomposite fibers," Composites Part A: Applied Science and
Manufacturing, vol. 40, pp. 709-723, 2009.
[57] J. H. Song, H. Huh, and H. T. Hahn, "Stress Evaluation of Nanocomposites with
Nanoplatelets," in 14th Int. Conf. on Composite Materials, 2003.
[58] J. H. Koo, Polymer Nanocomposites: McGraw-Hill, 2006.
[59] L. Pilato and M. J. Michno, Advanced Composite Materials: Springer, 1994.
[60] Y. Liu, Zhizhongsu, X. Li, W. Guo, Q. Li, and C. Wu, "Effect of Dispersion of
Carbon Black on Electrical and Thermal Properties of Polyethylene
Terephthalate/Carbon Black Composites," J. of Macromolecular Sci., vol. 48,
2009.
[61] H. Yang, P. Bhimaraj, L. Yang, R. W. Siegel, and L. S. Schadler, "Crystal growth
in alumina poly(ethylene terephthalate) nanocomposite films," J. of Polym Sci.
Part B: Polymer Phys, vol. 45, pp. 747-757, 2007.
[62] J.-H. Chang, S. J. Kim, Y. L. Joo, and S. Im, "Poly(ethylene terephthalate)
nanocomposites by in situ interlayer polymerization: the thermo-mechanical
properties and morphology of the hybrid fibers," Polymer, vol. 45, pp. 919-926,
2004.
[63] X. Yao, X. Tian, X. Zhang, K. Zheng, J. Zheng, H. Zhang, L. Chen, Y. Li, and P.
Cui, "Poly(ethylene terephthalate)/attapulgite nanocomposites: Preparation,
structure, and properties," Journal of Applied Polymer Science, vol. 110, pp. 140-
146, 2008.
[64] L. S. Brando, L. C. Mendes, M. E. Medeiros, L. Sirelli, and M. L. Dias,
"Thermal and mechanical properties of poly(ethylene terephthalate)/lamellar
zirconium phosphate nanocomposites," Journal of Applied Polymer Science, vol.
102, pp. 3868-3876, 2006.
[65] H. Kobayashi and M. Daimaruya, "Dynamics and quasi-static lateral compression
tests of ceramics tubes," Journal De Physique IV, vol. 4, pp. 275-280, 1994.
[66] E. P. Popov, Engineering Mechanics of Solids, second ed.: Prentice-Hall of India,
2005.

83
[67] S. A. Jabarin, "Crystallization Kinetics of Polyethylene Terephthalate. I.
Isothermal Crystallization from the melt," Journal of Applied Polymer Science,
vol. 34, pp. 85-96, 1987.
[68] "ASTM D638 - 08 Standard Test Method for Tensile Properties of Plastics," in
Annual Book of ASTM Standards. vol. 8.01: ASTM, 2009.
[69] R. B. Dupaix and M. C. Boyce, "Finite strain behavior of poly(ethylene
terephthalate) (PET) and poly(ethylene terephthalate)-glycol (PETG)," Polymer,
vol. 46, pp. 4827-4838, 2005.
[70] J. Noh, J. La, W. Youm, C. Kim, and K. Park, "A New Optical Laser
Extensometer using a Position Sensitive Detector (PSD)," in International
Conference on Advanced Inteligent Mechatronics, 2003.
[71] U. T. Systems, "Optical Extensometer - Features, Performance and Applications
of Laser Extensometer." vol. 2010.
[72] A. C. Fischer-Cripps, Nanoindentation, second ed.: Springer, 2004.
[73] M. F. Doerner and W. D. Nix, "A method for interpreting the data from depth-
sensing indentation instruments," J. Mater. Res, vol. 1, pp. 601-609, 1986.
[74] M. R. VanLandingham, J. S. Villarrubia, W. F. Guthrie, and G. F. Meyers,
"Nanoindentation of Polymers: An Overview," Macromolecular Symposia, vol.
167, pp. 15-44, 2001.
[75] R. F. Cook and M. L. Oyen, "Nanoindentation behavior and mechanical
properties measurement of polymeric materials," International Journal of
Materials Research, vol. 98, pp. 370-378, 2007.
[76] M. V. R. Kumar and R. Narasimhan, "Analysis of spherical indentation of linear
viscoelastic materials," Current Science, vol. 87, pp. 1088-1095, 2004.
[77] G. Huang and H. Lu, "Measurement of Young's relaxation modulus using
nanoindentation," Mech Time-Depend Mater, vol. 10, pp. 229-243, 2006.
[78] R. H. Ion, H. M. Pollock, and C. Roques-Carmes, "Micron-scale indentation of
amorphous and drawn PET surface," J. of Materaials Science, vol. 25, pp. 1444-
1454, 1990.
[79] K. J. V. Vliet, "Experimental Aspects of Nanoindentation," MIT OCW, 2003.
[80] Y. Kong and J. N. Hay, "The measurement of the crystallinity of polymers by
DSC," Polymer, vol. 43, pp. 3873-3878, 2002.
[81] V. Ratta, "Crystallization, Morphology, Thermal Stability and Adhesive
Properties of Novel High Performance Semicrystalline Polyimides," in Chemical
Engineering. vol. PhD: Virginia Tech, 1999.
[82] L. C. Sawyer, D. T. Grubb, and G. F. Meyers, " Introduction to Polymer
Morphology " in Polymer Microscopy, Third ed New York: Springer 2008, pp. 1-
25.
[83] J. Johnson, "X-ray Diffraction studies of the crystallinity in Polyethylene
Terephthalate," Journal of Applied Polymer Science, vol. II, pp. 205-209, 1959.
[84] C. J. Heffelfinger and P. G. Schmidt, "Structure and Properties of Oriented
Poly(ethylene Terephthalate) Films," Journal of Applied Polymer Science, vol. 9,
pp. 2661-2680, 1965.
[85] W. J. Sichina, "DSC as Problem Solving Tool: Measurement of Percent
Crystallinity of Thermoplastics," Perkin Elmer Inc.2000.

84
[86] A. Mehta, U. Gaur, and B. Wunderlich, "Equilibrium melting parameters of
Poly(ethylene Terephthalate)," J. of Polym Sci: Polymer Physics Edition, vol. 16,
pp. 289-296, 1978.
[87] P. C. Vilanova, S. M. Ribas, and G. M. Guzman, "Isothermal crystallization of
poly(ethylene terephthalate) of low molecular weight by different scanning
calorimetry: 1. Crystallization kinetics," Polymer, vol. 26, pp. 423-428, 1985.
[88] E. Gmelin and S. M. Sarge, "Calibration of differential scanning calorimeters,"
Pure and Appl. Chem., vol. 67, pp. 1789-1800, 1995.
[89] E. Lifshin, X-ray Characterization of materials: Wiley-VCH, 1999.
[90] C. Weder, B. H. Glomm, P. Neuenschwander, and U. W. Suter, "Ordering of
liquid crystalline solutions of rigid-rod aramids using mechanical shearing and
electric-field poling," Macromolecular Chemistry and Physics, vol. 196, pp.
1113-1127, 1995.
[91] I. M. Ward, "The Measurement of Molecular Orientation in Polymers by
Spectroscopic Techniques," J. of Polym Sci. Polymer Symposium, vol. 58, pp. 1-
21, 1977.
[92] M. C. Tobin, "The Infrared Spectra of Polymers. II. The Infrared Spectra of
Polyethylene Terephthalate," J. Phys. Chem, vol. 61, pp. 1392-1400, 1957.
[93] W. W. Daniels and R. E. Kitson, "Infrared Spectroscopy of Polyethylene
Terephthalate," J. of Polym Sci, vol. 33, pp. 161-170, 1958.
[94] A. Miyake, "The Infrared Spectrum of Polyethylene Terephthalate. I. The effect
of Crystallization," J. of Polymer Science, vol. 38, pp. 479-495, 1959.
[95] A. J. Melveger, "Laser-raman study of crystallinity changes in poly(ethylene
terephthalate)," Journal of Polymer Science Part A-2: Polymer Physics, vol. 10,
pp. 317-322, 1972.
[96] G. Ellis, F. Romn, C. Marco, M. A. Gmez, and J. G. Fatou, "FT Raman study
of orientation and crystallization processes in poly(ethylene terephthalate),"
Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy, vol. 51,
pp. 2139-2145, 1995.
[97] "A Study of Polymer Crystallinity," in Raman Spectroscopy, P. E. L. a. A.
Science, Ed., 2007.
[98] M. Radmacher, R. W. Tillmann, M. Fritz, and H. E. Gaub, "From Molecules to
Cells: Imaging Soft Samples with the Atomic Force Microscope," Science, vol.
257, 1992.
[99] D. Sarid, "Scanning Force Microscopy with applications to Electric, Magnetic and
Atomic Forces," Oxford series in Optical and Imaging Sciences, 1991.
[100] A. Yurtsever, A. M. Gigler, C. Dietz, and R. W. Stark, "Frequency modulated
torsional resonance mode atomic force microscopy on polymers," APPLIED
PHYSICS LETTERS, vol. 92, 2008.
[101] S. A. C. Gould, D. A. Schiraldi, and M. L. Occelli, "Analysis of Poly(ethylene
terephthalate) (PET) Films by Atomic Force Microscopy," 1997.
[102] D. A. Ivanov and S. N. Magonov, "Atomic Force Microscopy Studies of
Semicrystalline Polymers at Variable Temperature," 2003.
[103] C. Ortiz, "AFM Imaging," in Nanomechanical of Materials and Biomaterials. vol.
2008: MIT: OCW, 2008.

85
[104] R. Feng and R. J. Farris, "Linear thermoelastic characterization of anisotropic
poly(ethylene terephthalate) films," Journal of Applied Polymer Science, vol. 86,
pp. 2937-2947, 2002.
[105] J. P. Jog, "Crystallization of Polyethylene terephthalate," J. M. S - Rev.
Macromol. Chem. Phys., vol. C 35, pp. 531-553, 1995.
[106] S. A. Jabarin, "Crystallization kinetics of poly(ethylene terephthalate). III. Effect
of moisture on the crystallization behavior of PET from the glassy state," Journal
of Applied Polymer Science, vol. 34, pp. 103-108, 1987.
[107] N. Brown, "The relationship between yield point and modulus for glassy
polymers," Materials Science and Engineering, vol. 8, pp. 69-73, 1971.
[108] B. D. Beake and G. J. Leggett, "Nanoindentation and nanoscratch testing of
uniaxially and biaxially drawn poly(ethylene terephthalate) film," Polymer, vol.
43, pp. 319-327, 2002.
[109] "Nestle launches its first rPET bottled water product," in Food Production Daily,
2009.

86
Appendix I: Yield Strength Calculations

The Stress-Strain plotted using Laser extensometer strain data is adjusted to

obtain the final yield strength values for each resin. The curves are adjusted as the sample

started to deform outside the gauge length and proceeds in to the gauge length, the stress

value decreases and increases later in the uniform thickness, stress increase and reaches

the yield point. Process of adjustment is shown for DAK resin sample data.

Stress Vs Strain
50

40

30
Stress

20

10

0
0 0.1 0.2 0.3 0.4 0.5
Strain

Figure 42: Complete Stress-Strain Curve

87
Stress Vs Strain
45

40

Stress
35

30

25
0.01 0.015 0.02
Strain

Yield strength values are taken from the following obtained after removing the

unwanted data.

Stress Vs Strain
40
35
30
25
Stress

20
15
10
5
0
0 0.01 0.02 0.03 0.04 0.05
Strain

Figure 43: Stress-Strain curve for Yield Strength

88
Appendix II: Process at Niagara

PET samples used in this work are collected after test running each resin to their

optimum processing conditions at Niagara Bottling LLC, Dallas facility. They are

producing a 10.3 gram bottle at this facility. PET resins from different resin

manufacturers were used in the process. The production process starts with drying the

resin to 165oC for about 6 hrs, to remove moisture. Later, dried resin passes through the

screw barrel where it is melted by heating to a temperature of 290oC. The molten resin is

injected into the preform moulds through a hydraulic system. Two HyPET injection

molding machines (Manufactured by Husky Co) capable of producing 144 preforms at a

cycle time of 6-6.5 sec are used.

Robotic arm

Mould

Figure 44: Injection Molding Unit, Robotic Arm and Mould

89
Preforms produced are either stored for later use or taken directly into the blow

molder through conveyor system. Three blow molders, manufactured by Sidel Corp. are

used for blowing the preforms. Each blow molder contains 24 moulds and produce

43,000/hr at their maximum speed. Preforms taken into the blow molder are heated above

the glass-transition temperature of PET to 90oC. Preforms are divided into five zones for

heating purpose. They pass at high speed through ceramic lined chamber where 20

heating filaments are placed to provide uniform heating. The temperatures of the zones

are set based on the required amount of material at each point along the length of the

bottle.

Figure 45: Bottle Mould inside a Blow Molder


Bottles produced are inspected for any blowing and heating defects through a high

speed imaging system, which measures the dimensions, color and centering of the bottle

base. Preforms produced are inspected under magnifier and polarized light for any

injection molding defects such as micro air bubbles, cold shot (local crystallization,

leading to residual stress can be seen through polarized light), foreign particles, and

90
injection lines. Some are symptoms for mold cooling and other machine issues; other can

be corrected by changing the molding parameters. Bottles blown are tested for their part

weight and top load. This data will help in changing the blowing parameters.

Niagara also has in house cap production units; they use J50 grade HDPE resin.

The bottles produced are conveyed to the filling unit through closed conveying system by

using pressurized air. After filling and labeling process, they are stacked on to pallets and

wrapped before shipped to the customer.

Figure 46: Stages of Blow molding[97]

Figure 47: Polarized image of normal preform with defective

91
Appendix III: Design and Apparatus

In the process of fixture development, initial approach was to take advantage of

the threaded portion of the preform to grip the top portion on to the Instron. With this

approach, for the simplification of machining process, a four-piece fixture idea is put

front and developed as a CAD model. At this point, verification of design highlighted the

drawbacks such as machining difficulties, not enough contact area to hold the preform,

elevating the necessity to redesign. In the new design, same four-piece idea is

implemented by extending the fixture contact till the uniform thickness region of the

preform i.e. the tapered region provided the required extra support for holding the

preform. Further simplified design is obtained through a two piece design as shown in

Figure 12 (b).

Initial experiments were carried out by holding one side of the preform using the

fixture and the bottom was held between round-jaw faces for grips. The preform was

distorted by the gripping force. Later, test trails were carried out using sand/salt & metal

insert inside the preform to resist the force also turned out unsuccessful. At this point, the

concept of holding the preform using metal pin parallel to the ASTM standard D638 as

mentioned previously. Preform samples are tested using the evolved fixture design.

Lower success rate during initial testing raised the question of repeatability to proceed

with this design, for developing a standard testing procedure. To solve this issue, the

bottom pin design is changed and started application of glue between the pin and

92
preform. Bottom holding pin is designed keeping the load bearing requirements and the

size constraints with the preform sample. Further to reduce the failures that result due to

the reduced clearance between pin outer diameter and diameter of the drill performed to

accommodate the pin, outside mount is introduced as shown in Figure 48. This is

intended to increase the area load being distributed.

10 mm

Figure 48: Preform holding pin and bottom cup

Modulus of resins calculated using the strain measured through cross-head

displacement does not provide accurate comparison between the resins. Comparison of

modulus data based on Laser and cross-head strain is shown in Figure 49. This highlights

the importance of having accurate strain measurements. Error in these measurements

occurs due to the machine and sample compliance. Presence of a number of moving

parts and deflection of the test frames at higher loads gives raise to compliance.

93
2700 950
Laser Strain

Modulus (MPa) - Cross head stain


900
Modulus (MPa) - Laser strain 2500 Cross-head Stain
850
2300 800
750
2100
700
1900 650
600
1700
550
1500 500
B B J K I F E H G E D A B C
0.82 0.76 0.77 0.84 0.80
Resins

Figure 49: Modulus of resins with Laser and cross-head strains

94
VITA

Sudheer Bandla

Candidate for the Degree of

Master of Science

Thesis: EVALUATION AND STABILITY OF PET RESIN MECHANICAL


PROPERTIES

Major Field: Mechanical and Aerospace Engineering

Biographical:

Personal Data: Born to Prasada Rao and Sasikala Bandla in Guntur, India, on
August 22, 1985.

Education: Received Bachelor of Engineering degree in Mechanical


Engineering from Acharya Nagarjuna University, Guntur, India in May
2006.

Completed the requirements for the Master of Science degree with a


major in Mechanical and Aerospace Engineering at Oklahoma State
University, Stillwater, Oklahoma in July, 2010.

Experience: Graduate Research Assistant in Mechanical and Aerospace


Engineering Department, Oklahoma State University, Stillwater,
Oklahoma May, 2008 July 2010.

Maintenance Engineer at Iron Ore Beneficiation plant, ESSAR


Steels Ltd., Bailadilla, Chhattisgarh, India (September 2006 December
2007)

Professional Memberships: American Society of Mechanical Engineers


(ASME), and Materials Research Society (MRS)
Name: Sudheer Bandla Date of Degree: July, 2010

Institution: Oklahoma State University Location: Stillwater, Oklahoma

Title of Study: EVALUATION AND STABILITY OF PET RESIN MECHANICAL


PROPERTIES

Pages in Study: 94 Candidate for the Degree of Master of Science

Major Field: Mechanical and Aerospace Engineering

An emphasis in environmental responsibility is one factor driving the plastics


industry towards use of renewable materials and light-weighting. Poly (ethylene
Terephthalate) (PET) made from renewable sources (feedstock) is a concern due to the
volume required (13 MTPA). PETs properties make it the markets choice for
packaging. Its properties are robust but sensitive to processing and resin chemistry.
Differences in micro-structure achieved through processing changes provide a range of
properties. This highlights the importance of testing samples with actual processing
conditions when considering improved designs. In the packaging industry, a superior
product can be obtained through usage of efficient part geometry and materials.
Customer acceptance is also an issue. In the process of going ecologically-friendlier and
maintaining performance, some researchers are trying to improve material properties by
the addition of nano-particles. Even 1 wt% addition has been shown to significantly
improve stiffness. The present work aims at examining the typical resins available (I.V.
0.74-0.84 dL/g) to obtain a baseline of performance based on their mechanical properties.
To have a comparison of material properties, amorphous samples were produced on the
latest generation of industry injection molding equipment. Prior studies have not
published testing of samples processed at quenching and flow rates typically used in
industry.

A convenient method for precise mechanical testing of the preform samples was
developed. This new method, including a custom fixture, accommodates the typical
preform shape used as an intermediate in packaging production. The test samples were
also analyzed for crystallinity and microstructural differences. Outcomes of this work
show 13-27% performance range between the resins. This is notable since it is within the
typical property improvements reported by 1 wt% nano-particle reinforcements. These
results also highlight the industrys need for forming a record of resins available for resin
selection.

ADVISERS APPROVAL: Dr. Jay C. Hanan

Anda mungkin juga menyukai