136
Rheology of Foaming Polymers and Its Influence on Microcellular Processing by Jing Wang A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Department of Mechanical and Industrial Engineering University of Toronto © Copyright by Jing Wang 2009

Rheology of Foaming Polymers and Its Influence on Microcellular Processing

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Rheology of Foaming Polymers and Its Influence on Microcellular Processing

by

Jing Wang

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy

Department of Mechanical and Industrial Engineering University of Toronto

© Copyright by Jing Wang 2009

ii

Rheology of Foaming Polymers and Its Influence on Microcellular

Processing

Jing Wang

Doctor of Philosophy

Department of Mechanical and Industrial Engineering

University of Toronto

2009

Abstract

The rheological properties of polymer melts and polymer/blowing agent (BA) solutions

are determined experimentally and the influences of material rheological properties and

crystallization on low-density foaming behaviour of polylactic acid (PLA) are investigated.

Understanding the rheological properties of foaming polymers allows the optimization of

polymer chemical structure and the development of technologies that produce desired cell

morphologies.

Although the technology for producing CO2-blown polystyrene (PS) foams is well

established, the rheological properties of a PS/CO2 solution, especially its extensional property,

are not well understood. In this study, these properties are determined with an in-house

developed, online technique, and the measured data are compared with those from commercial

rheometers. The online measurement system consists of a tandem foam extrusion system and a

die for measuring pressure drops. Shear viscosity is determined from the pressure drop over a

straight rectangular channel, while planar extensional viscosity from the pressure drop over a

thin hyperbolic channel, taking into account the pressure drop due to shearing. Measured

viscosities of the polystyrene without CO2 compare well with those from commercial

iii

rheometers. With the presence of dissolved CO2, both the shear and extensional viscosities of the

polystyrene are significantly reduced. The influence of CO2 on the two viscosities is found to be

similar to an increase of temperature.

Polylactic acid is the first mass-produced biodegradable polymer, and has potential to

replace petroleum-based polymers in foaming applications. In this study, the influences of

material rheological properties and crystallization on the low-density, microcellular extrusion

foaming behaviour of polylactic acids (PLAs) are investigated. Comparisons are made between

linear and branched PLAs and between amorphous and crystalline PLAs. The branched PLAs are

found to produce foams with higher expansion ratios and reduced open-cell content compared to

the linear PLA. The foaming behaviour of the linear PLA, then, is significantly improved by

adding a small amount of long-chain-branched PLA. The improved cell structure with branched

PLAs is attributed to their relatively high melt strength and strain to break. For the first time, it is

shown that crystallization, induced by cooling and macroscopic flow during processing,

increases melt strength, which aids the production of low-density foams.

iv

Acknowledgments

In a thesis project of this scale, there are certainly a multitude of people that have helped

to make it possible. I want to start by acknowledging my supervisors, Professors Chul Park and

David James, for guiding me into the world of science and technology. Professor Park taught me

how to approach complex engineering problems effectively by identifying the key issues and

focusing on them, and his industrial collaborations have always impressed me how useful

cellular polymers can be. I learnt from Professor James how to express my opinions in precise

scientific language, and that a serious researcher should be willing to cast doubt on what seems

established theories and practices. I wish my future career can combine both their styles and win

respects from practitioners as well as theoreticians.

I thank Professor Hani Naguib and Professor Markus Bussmann for agreeing to join my

thesis committee three years ago and having provided useful advice at each of the annual

progress review meetings. I thank Professor John Dealy for agreeing to be my external examiner.

His reputation in the rheology community has certainly been an encouragement for me to work

hard on this thesis.

During my graduate study at the University of Toronto, I received help from many

colleagues including all members of the microcellular plastics manufacturing laboratory. I thank

Dr. John Lee for collaborating with me on so many projects. He is one of the most dutiful

engineers I know, and I wish him all the best in your future career. I thank Nan Chen for always

being helpful. I hope he will enjoy receiving the same degree as I did, again. I thank Esther Lee

for tipping me off on the local life. With her tips the city of Toronto is never a boring place. I

thank Dr. Wenli Zhu for helping me with the experiments, and for all the dinners she cooked for

us: they are just delicious! I thank all other past and present MPML members, including Dr.

Guangming Li, Dr. Zhenjin Zhu, Dr. Gangjian Guo, Dr. Wenli Zhu, Dr. Gary Li, Dr. Qingping

Guo, Dr. Chunmin Wang, Dr. Wenge Zheng, Dr. Hongbo Li, Dr. Jin Wang, Xindong Zhu, Dr.

Donglai Xu, Lilac Wang, Hongtao Zhang, Jingjing Zhang, Qingfeng Wu, Mingyi Wang,

Professor Guoliang Tao, Dr. Changwei Zhu, Sunny Leung, Raymond Chu, Anson Wong, Dr.

Ryan Kim, Dr. Kevin Lee, Peter Jung, Dr. Yongrak Moon, Dr. Patrick Lee, Dr. Kyungmin Lee,

Dr. Jae Yoon , Richard Lee, Sue Chang, Johnny Park, Hojin Chi, Alex Lee, Dr. Takashi Kuboki,

Dr. Bhuwnesh Kumar, Dr. Mohammed Serry, Mohammad Hasan, Dr. Maridass

v

Balasubramanian, Professor Taher Azdast, Professor Wanrudee Kaewmesri, Florien Gunkel,

and Ivan Gutierrez, for both the collaborations on research projects and the technical discussions

that were always insightful and inspiring.

I thank Jeff Sansome, David Esdaile, Ryan Mendell, Mike Smith, and Tai Tran Do at the

MIE machine shop for their quality work and for always going out of the way to tell me better

mechanical designs.

Many staff members at the department of mechanical & industrial engineering have

helped me in one way or another. My special thanks go to Brenda Fung, Sheila Baker, Teresa

Lai, Joe Baptista, Lorna Wong, Geoffrey Chow, and Oscar Del Rio.

Finally, I thank my parents for their support throughout my student life. Working through

the years without getting a job is a big commitment, but they were always there to encourage me.

I will make every effort to make them proud of me in my future career.

vi

Table of Contents

Abstract ……………………………………….……….…...……………………….….……… ii

Acknowledgements ……………………………………………….……….…...……...…….. iv

List of Tables ……………………………………………………..………………………..….. ix

List of Figures ………………………………………………….………………...............…….. x

List of Symbols ………………………………………………………………….……….…..... xv

Chapter 1 Introduction

1.1 Thermoplastic Foams and Their Processing Technology ……...………….…….. 2

1.2 Rheology of Polymeric Fluids …...…………………………………………….... 4

1.3 Motivations for the Study ………………………………………………..…… 6

1.3.1 Challenges with Characterizing Rheological Properties ……….……… 6

1.3.2 Importance of Material Rheology Properties to Foam Processing …… 7

1.4 General Objectives …………………………..……...………………………….. 8

1.5 Overview of the Thesis ………………....…………………………………........ 9

Chapter 2 Background and Literature Review

2.1 Thermoplastic Foams and Their Processing Technology ……………………. 10

2.1.1 Categories of Foams ………………………………………………… 10

2.1.2 Microcellular Foam Processing – Formation of Polymer/BA Solution 11

2.1.3 Microcellular Foam Processing − Cell Nucleation and Growth ………. 17

2.2 Rheology of Polymer Melts and Polymer/BA Solutions …………...……...… 21

2.2.1 Shear Rheometry of Polymer/BA Solutions …………………………… 21

2.2.2 Extensional Rheometry Involving Shear-Free Flows ………………….. 24

2.2.3 Extensional Rheometry Involving Mixed Flows ……………………… 26

2.3 Relationship between Cell Growth and Rheological Properties ………….…… 29

2.4 Objectives of the Thesis ……………………………………………………….. 32

Chapter 3 Characterization of the Shear Properties

3.1 Experimental …………………………………………………………..........…. 33

3.1.1 The Hele-Shaw Channels ………………………………………………. 33

vii

3.1.2 The Processing System ………………………………………………… 35

3.2 Properties of the Polymer and the Polymer/Blowing Agent Solution ……...... 37

3.2.1 General Physical Properties ……............................................……….. 37

3.2.2 Rheological Properties Using Commercial Rheometers ……..………… 39

3.3 Shear Viscosity of Polystyrene Alone ….………….………………….……… 40

3.3.1 Calculating the Viscosity ……………………………………….……… 40

3.3.2 Viscosity Data ………………………………………………………… 42

3.4 Shear Viscosity of Polystyrene/CO2 Solution ………………………….…….. 46

3.4.1 Viscosity Data and Prediction by the Free Volume Theory ………… 46

3.4.2 Comparing Viscosity Reductions of Various BAs …………………… 48

Chapter 4 Characterization of the Extensional Properties

4.1 Introduction ………………………………….………………………........….... 52

4.2 Uniaxial Extensional Viscosity from EVF …………………..………………… 54

4.3 Calculating the Pressure Drop due to Extension ……………………………… 57

4.4 The Extensional Rate and Total Strain …………………………………..…….. 62

4.5 Comparing Extensional Viscosities ………………………………………….… 64

4.6 Extensional Flow Resistance of the Solution ………………………………… 71

Chapter 5 Influence of Rheological Properties on the Low-Density Microcellular

Foaming of Polylactic Acids

5.1 Introduction ……………………………….………………………...………… 75

5.2 Experimental ………………………………………………………………... 78

5.3 Properties of the Polymers …….….….….………………………………….. 78

5.3.1 General Physical Properties ………………………………………...… 78

5.3.2 Rheological Properties ………………………………………………. 82

5.4 Results and Discussions ……………………………………………………… 86

5.4.1 Processing Strategies …………………………….……..…………….. 86

5.4.2 Cell Densities …………………………………………....…………. 89

5.4.3 Expansion Ratios ………………………...………………..………….. 89

5.4.4 Cell Morphology from SEM ……………………………....………… 93

5.5 Influence of Processing Conditions on PLA Crystallization …………………. 96

viii

5.6 Controlling PLA Crystallization and Its Influence on Foaming ……………… 101

Chapter 6 Conclusions ……………………………………………………………...…… 105

Bibliography ………………………………………………………………………….……. 110

Appendix ……...………….……………………………………….………………………… 117

ix

List of Tables

Table 3-1 Relaxation times of PS685D, based on oscillatory shear data …………..…… 39

Table 3-2 Best fitting parameters for the master plot of PS685D ………………………. 44

Table 5-1 Properties of the three PLAs ………………………………………………...... 82

x

List of Figures

Figure 1-1 Illustrations of the main stages of microcellular processing: (a) formation of

polymer/blowing agent (gas) solution; reproduced from Park and Suh (1996); (b)

cell nucleation and growth ………………….………………………..….............. 3

Figure 1-2 Illustrations of simple flows for rheological characterization: (a) simple shearing;

(b) uniaxial extension; where V is the velocity ………………….......................... 5

Figure 2-1 Scanning electron microscopy (SEM) photos of: (a) open-cell structure; (b) close-

cell structure; reproduced from Park et al. (1998) …………………………… 12

Figure 2-2 Mixing elements on the plasticating screw ……………………………….......... 15

Figure 2-3 Schematic of heat exchanger containing static mixers used in foam extrusion

……………………………………………….………………………………… 16

Figure 2-4 Free energy of forming bubble from supersaturated polymer melt …………… 18

Figure 2-5 Surface force balance for a heterogeneously nucleated bubble, where θ is the

wetting angle and σ is the surface tension ……………………………………… 20

Figure 2-6 Schematic of an extruder setup for measuring PS/CO2 solution viscosity;

reproduced from Lee et al. (1999) …………………………………………...…. 23

Figure 2-7 Rotary clamp rheometers for measuring: (a) planar extensional viscosity; (b)

biaxial extensional viscosity; reproduced from Meissner (1987) …...…………. 25

Figure 2-8 Basic elements of an entry flow for flow from a large tube through an abrupt entry

into a small tube. The illustration applies to both axisymmetric contraction and

planar contraction; reproduced from Boger (1987) ……………………………27

Figure 2-9 Distribution of BA concentration in the surrounding of a cell (a gas bubble), as

described by the cell model; 𝑐 𝑟, 𝑡 is the BA concentration, 𝑐𝑅 𝑡 is the BA

concentration at cell surface, 𝑃𝑔 𝑡 is the cell pressure, 𝑘𝐻 is Henry‟s Law

xi

constant, 𝑅 𝑡 is the cell radius, and 𝑅𝑠 𝑡 is the outer radius of the melt envelope

[Amon and Denson (1984)] …………………………………………………… 30

Figure 3-1 The two test dies. The circles indicate the diaphragms of the pressure transducers.

The dimensions are: 𝐵0 = 30 𝑚𝑚 , 𝐵1 = 3 𝑚𝑚 , 𝐿0 = 20 𝑚𝑚 , 𝐿1 = 5 𝑚𝑚 ,

𝐿2 = 20 𝑚𝑚. The depth H (into the page) is 𝐻 = 0.94 𝑚𝑚 or 1.96 𝑚𝑚. and

–𝐻

2< 𝑧 <

𝐻

2 …………………………………………………………………. 34

Figure 3-2 Schematic of the tandem extrusion system …………………………………… 36

Figure 3-3 Density of polystyrene and polystyrene/CO2 solution; extracted from Li (2008)

………………………………………………………………………………… 38

Figure 3-4 (a) Estimates of the first normal stress difference in shearing for the PS at 172oC;

(b) Estimated exit pressure of flows in the Hele-Shaw channels at several

temperatures …………………………………….………………………….…. 41

Figure 3-5 (a) Shear viscosity data of the polystyrene from various rheometers; (b) master plot

of the shear viscosities from data in Figure 3-5(a) ……………………………… 43

Figure 3-6 Shear viscosity of polystyrene, determined from flow measurements in the Hele-

Shaw channels ………………………………………………………………… 45

Figure 3-7 (a) Shear viscosity of PS/CO2 solution, compared with that of PS only; (b) Master

plot of the PS/CO2 solution for a reference temperature of 172oC, compared to the

viscosity of PS alone at 172oC ………………………………………………… 47

Figure 3-8 Reduction of glass transition temperature for PS as a function of CO 2

concentration; reproduced from Wissinger and Paulaitis (1987) ….…………… 49

Figure 3-9 Viscosity reduction factors of the present PS with various blowing agents: (a) as a

function of molar concentration and (b) as a function of weight concentration.

Values for CFC-11 and CFC-12 are calculated from Han et al. (1983)

………………………………………………………………………………. 51

xii

Figure 4-1 (a) Schematic of the ARES-EVF (Extensional Viscosity Fixture); (b)

representative positions of the rotating cylinders, and corresponding Hencky

strains; Reproduced from the product note on EVF technology by TA Instruments

Inc ………………………………………………………….…………………. 55

Figure 4-2 Transient uniaxial extensional viscosity of PS determined with the EVF fixture;

the symbols “x1”, “x0.2”, and “x0.1” indicate that the original data were

multiplied by these factors to avoid overlapping ………………………………. 56

Figure 4-3 Ratio of pressure gradient neglecting the aspect ratio of a rectangular channel over

that considering the aspect ratio, 𝑘 =𝑑𝑝 /𝑑𝑥𝐻𝑆

𝑑𝑝 /𝑑𝑥, with values obtained from the

literature and from running an in-house code ………………………….............. 60

Figure 4-4 The total pressure drop and the pressure drop related to extension in the 0.94 mm

channel (see Equation 4-8 for definitions of ∆𝑃𝑐𝑜𝑛𝑣𝑒𝑟𝑔𝑒𝑛𝑡 and ∆𝑃𝑒) ……………. 61

Figure 4-5 Ratio of centreline velocity for a finite aspect ratio over that for an infinite aspect

ratio, used as a correction factor and determined from literature sources and from

an in-house numerical code ……………………………...................................... 63

Figure 4-6 Comparison of planar extensional viscosity 𝜂𝑃 𝜀 from two Hele-Shaw channels

with uniaxial extensional viscosity 𝜂𝐸 𝜀 predicted from EVF measurement

………………………………………………………………………………… 66

Figure 4-7 Trouton ratio of the PS calculated from the 1.96 mm Hele-Shaw channel and from

EVF measurement. Comparison of 𝜂𝑃 𝜀

𝜂 2∙𝛾 with

𝜂𝐸 𝜀

𝜂 3∙𝛾 …………………………. 68

Figure 4-8 Plot of 𝜏𝑦𝑦 /𝜏𝑥𝑥 in planar extensional flow, calculated using the upper-convected

Maxwell model ………………………………………………............................. 70

Figure 4-9 Planar extensional viscosity of PS and PS/CO2 solution from the 0.94 mm Hele-

Shaw channel ……………………………………………………….................... 72

xiii

Figure 4-10 Trouton ratios 𝜂𝑃 𝜀

𝜂 2∙𝛾 of the PS and the PS/CO2 solution calculated from the 0.94

mm Hele-Shaw channel data at several temperatures ……………...................... 74

Figure 5-1 Schematic of the tandem extrusion system for foam extrusion using a capillary

die; the system setup is similar to that in Naguib et al. (2002) ………………. 79

Figure 5-2 (a) Schematic of high molecular weight PLA molecule; (b) General structure of

the styrene-acrylic multi-functional oligomeric chain extenders; where R1 – R5 are

H, CH3, a higher alkyl group, or combinations of them; R6 is an alkyl group, and

x, y, and z are each between 1 and 20. Reproduced from Villalobos et al. (2006)

………………………………………………………….……………………… 80

Figure 5-3 Complex viscosities of the four PLAs at 180oC ………………………………... 83

Figure 5-4 Transient uniaxial extensional viscosities of: (a) the linear PLA at 140oC; (b) the

half-LCB PLA at 160oC; (c) the LCB PLA at 160

oC; (d) the LCB PLA with

lubricant at 160oC ………………………………………………………………. 85

Figure 5-5 Transient uniaxial extensional viscosities of: (a) the linear PLA at 140oC; (b)

blend of 10% LCB PLA and 90% linear PLA at 160oC; (c) the blend of 20% LCB

PLA and 80% linear PLA at 160oC ……………………………….……………. 87

Figure 5-6 Exit die pressure as a function of CO2 concentration by weight and die

temperature for the LCB PLA …………………….……………..……………. 88

Figure 5-7 Cell densities of the linear, the half-LCB, and the LCB (without lubricant) PLAs

from foam extrusion as a function of processing temperature ……………… 90

Figure 5-8 Expansion ratios of: (a) all four grades of PLAs; error bars are omitted for clarity;

(b) the linear PLA, the LCB PLA, and blends of the two …………………… 92

Figure 5-9 SEM images of the cellular structures; the temperatures correspond to the highest

expansion ratios at the given CO2 concentration ……………………………… 94

xiv

Figure 5-10 Cross sections of the extruded filament: (a) LCB PLA with 9% CO2 at 115oC; (b)

half-LCB PLA with 9% CO2 at 117oC; Notice the open-cell structure in the core

of the half-LCB PLA filament ………………………………………………… 95

Figure 5-11 Crystallinity of the foams as a function of CO2 content and temperature, for the

LCB PLA (without lubricant) and the half-LCB PLA ……………………… 97

Figure 5-12 Crystallization half time for the LCB PLA with and without shearing; the

crystallinity at these half times is approximately 15% ...............................…… 99

Figure 5-13 Ratio of the storage modulus of crystalline PLA to that of the amorphous PLA as

a function of crystallinity; the material is LCB PLA without lubricant ………. 100

Figure 5-14 Expansion ratios of foams produced after different time for crystallization. The

material is LCB PLA and crystallinity of the foam skin, determined on DSC, is

shown for several conditions ………………………………………………….. 102

Figure 5-15 SEM images of the cellular structures, the polymer is LCB PLA, and 9% CO2 is

used: (a) tcry ≈ 0 s, 112oC; (b) tcry ≈ 0 s, 120

oC; (c) tcry ≈ 90 s, 110

oC; (d) tcry ≈ 90 s,

120oC ………………………………………………………………………….. 103

xv

List of Symbols

𝛼𝑇 Shift factor for viscosity

𝛾 Shear rate, [𝑠−1]

𝛾 𝑎 Apparent shear rate, [𝑠−1]

∆𝐸𝐷 Activation energy, [𝐽]

∆𝐺 Gibbs free energy, [𝐽]

∆𝐺𝑕𝑜𝑚∗ Free energy to form critical nucleus during homogeneous nucleation, [𝐽]

∆𝐺𝑕𝑒𝑡∗ Free energy to form critical nucleus during heterogeneous nucleation, [𝐽]

∆𝐺𝑉 Free energy difference between the bubble phase and the polymer phase, [𝐽]

∆𝑃 Pressure drop, [𝑃𝑎]

∆𝑃𝑒𝑛 Entrance pressure drop, [𝑃𝑎]

∆𝑃𝑒𝑥 Exit pressure drop, [𝑃𝑎]

∆𝑃𝑒 Extra pressure drop, [𝑃𝑎] ; ∆𝑃𝑒 = ∆𝑃𝑒𝑛 + ∆𝑃𝑒𝑥

𝜀 Elongation (or extension, or stretch) rate, [𝑠−1]

𝜀𝐻 Hencky strain

𝜂 Shear viscosity, [𝑃𝑎 ∙ 𝑠]

𝜂0 Zero-shear-rate viscosity, [𝑃𝑎 ∙ 𝑠]

𝜂∞ Viscosity at infinite shear rate, [𝑃𝑎 ∙ 𝑠]

𝜂𝑃,𝑎𝑝𝑝 Apparent planar extensional viscosity, [𝑃𝑎 ∙ 𝑠]

𝜂∗ Complex viscosity, [𝑃𝑎 ∙ 𝑠]

xvi

𝜂𝐵+ Transient biaxial extensional viscosity, [𝑃𝑎 ∙ 𝑠]

𝜂𝐸+ Transient uniaxial extensional viscosity, [𝑃𝑎 ∙ 𝑠]

𝜂𝑃+ Transient planar extensional viscosity, [𝑃𝑎 ∙ 𝑠]

𝜆 Relaxation time, [𝑠]

𝜌 Density, [𝑔/𝑐𝑚3]

𝜍 Surface tension, [𝑁/𝑚]

𝜏21 or 𝜏𝑦𝑥 Shear stress, [𝑃𝑎]

𝜏𝑤 Wall shear stress, [𝑃𝑎]

𝜏11 − 𝜏22 Principal tensile stress difference, [𝑃𝑎]

𝜏22 − 𝜏33 Secondary tensile stress difference, [𝑃𝑎]

𝜒 Crystallinity [%]

𝛹1 First normal stress coefficient, [𝑃𝑎]

𝜔 Angular frequency, [𝑟𝑎𝑑/𝑠]

𝐴 Bubble surface area, [𝑚2]

𝑏 Extensional flow parameter, [0 < 𝑏 < 1]

𝐵 Width, [𝑚]; used in various contexts

𝐶 Weight concentration, [𝑔/𝑐𝑚3 or 𝑔/𝑔]

𝐶𝑚 Molar concentration, [𝑚𝑜𝑙/𝑐𝑚3 or 𝑚𝑜𝑙/𝑔]

𝐶𝑝 Heat capacity, [𝐽/𝑘𝑔 ∙ 𝐾]

𝐷 Diffusivity, or diffusion coefficient [𝑚2/𝑠]

xvii

𝑫 Rate-of-strain tensor [𝑠−1]

𝐷0 Maximum diffusion coefficient (at infinite temperature) [𝑚2/𝑠]

𝐷𝑒 Deborah number

𝑓 Fractional free volume

𝐹 Force, [𝑁]

𝑔 Gravitational constant, 9.8 𝑚/𝑠2

𝐺 Elastic modulus, [𝑃𝑎]

𝐺 ′ Storage modulus, [𝑃𝑎]

𝐺" Loss modulus, [𝑃𝑎]

𝐻 or 𝑕 Height, [𝑚]; used in various contexts

𝑘 Thermal conductivity, [𝑊/𝑚 ∙ 𝐾]

𝑘𝐵 Boltzmann constant, 1.38 × 10−23 𝐽/𝐾

𝑘𝐻 Henry‟s Law constant

𝑙𝐷 Striation thickness [𝑚]

𝐿 Length, [𝑚]; used in various contexts

𝑀𝑛 Number-averaged molecular weight, [𝑔/𝑚𝑜𝑙]

𝑀𝑤 Weight-averaged molecular weight, [𝑔/𝑚𝑜𝑙]

𝑚 Power Law parameter

𝑛 Non-Newtonian index in Power Law

𝑁𝐴 Avogadro number, 6.022 × 1023 𝑚𝑜𝑙−1

xviii

𝑁1 First normal stress difference, [𝑃𝑎]

𝑁𝑕𝑜𝑚 Cell nucleation rate during homogeneous nucleation [𝑐𝑒𝑙𝑙𝑠/𝑠]

𝑁𝑕𝑒𝑡 Cell nucleation rate during heterogeneous nucleation [𝑐𝑒𝑙𝑙𝑠/𝑠]

𝑃 Pressure, [𝑃𝑎]

𝑃𝑔 Gas pressure inside a cell, [𝑃𝑎]

𝑃𝑠 System pressure, [𝑃𝑎]

𝑄 Volumetric flow rate, [𝑚3/𝑠]

𝑟, 𝜃, and 𝑧 Cylindrical coordinates

𝑅 or 𝑟 Radius, [𝑚]

𝑟∗ Critical radius, [𝑚]

𝑅𝑔 Molar gas constant, 8.314 𝐽 ∙ 𝐾−1 ∙ 𝑚𝑜𝑙−1

𝑅𝑒 Reynolds number

𝑡1/2 Crystallization half time, [𝑠]

𝑡𝐷 Characteristic time for gas diffusion [𝑠]

𝑡𝑓𝑙𝑜𝑤 Characteristic time of the flow system [𝑠]

𝑇 Temperature, [𝐾]

𝑇𝑐 Crystallization temperature, [𝐾]

𝑇𝑔 Glass transition temperature, [𝐾]

𝑇𝑚 Melting temperature, [𝐾]

𝑇𝑟 Reference temperature, [𝐾]

xix

𝑇𝑟 Trouton ratio

𝑣 Velocity, [𝑚/𝑠]

𝑉𝑏 Initial volume of bubble, [𝑚3]

𝑉𝑡 Theoretical expansion ratio

𝑉𝑅𝐹 Viscosity reduction factor

𝑥, 𝑦, and 𝑧 Cartesian coordinates

1

Chapter 1 Introduction

In less than a century, polymers have evolved into a global industry that influences every

aspect of our lives. For the year 2000, nearly 200 million tons of synthetic polymeric materials,

or plastics, were produced worldwide to satisfy ever-growing market needs. This amount equals

to 2% of the wood and nearly 5% of the oil consumed by the world in that year. As a leading

producer, the United States produced $250 billion-dollar worth of plastics in the year 2000,

contributing about 4% of the gross domestic product [Carraher (2003)].

Polymers have no competing materials in terms of weight, ease of processing, economy,

and versatility. However, polymers can be made lighter and more versatile by foaming them with

a blowing agent (BA), usually a gas under room temperature and pressure. The foams, also

called cellular polymers, first came into use during the 1940s, and have since enjoyed a

continuously growing market. As of 2005, the world consumption of foamed polymers was

approximately 23 billion pounds, with 1/3 of this amount occurring in the United States [The

Freedonia Group (2001)]. Growth is attributed to their light weight, excellent strength-to-weight

ratio, superior insulation properties, and energy absorbing capability.

During the past three decades, foaming technology has evolved from purely heuristic

efforts to fundamental approaches relying on science and engineering. Today, advanced

technologies such as the microcellular foaming technology [Baldwin et al. (1994a) and (1994b)]

enable engineers to control details of the cellular structures and to design foam-based materials

for advanced applications [Suh et al. (2000)].

Nearly all advanced foaming technologies subject the polymer/BA mixtures to a series of

well-defined kinematic events, such that the main stages of processing, dissolution of the

blowing agent, phase separation, and hardening of the cellular structure, are precisely controlled.

The development of these technologies requires accurate knowledge of the rheological (flow)

properties of the polymer/BA mixtures and how these mixtures react to processing conditions

and geometric constraints. Unfortunately, rheological properties of polymers and polymer/BA

mixtures are notoriously difficult to determine. Even if they can be determined, using this

information to analyze a foaming process and to predict cell morphology is not straightforward.

2

This thesis is concerned in general with these two challenges. The polymers of interest will be

polydisperse homopolymer, typically used in foam processing, and the BA will be supercritical

CO2, extensively investigated in recent years because of its environmental friendliness and its

tendency to induce high-cell-density, microcellular foams [Tomasko et al. (2003)]. In this thesis,

techniques for characterizing rheological properties will be presented, and the result will help

with understanding and predicting of cellular structures in processing.

1.1 Thermoplastic Foams and Their Processing Technology

Thermoplastic foams are cellular materials consisting of dispersed, usually spherical,

gaseous voids and a continuous thermoplastic matrix. They belong to a more general class of

foams with various matrices, including thermosetting foams (e.g., polyurethane foams),

naturally-occurring foams (e.g., wood and cancellous bones), food foams (e.g., steamed rice and

flour dough), and liquid foams (e.g., soap foam). „Foam‟ is different from „porous materials‟

because the former involves volume expansion and the latter does not.

Most thermoplastics can be foamed. Thermoplastic foams possess unique physical,

mechanical, and thermal properties, which are governed by the polymer matrix, characteristics of

the cellular structure (e.g., cell density and cell size), and the BA composition. They have found

widespread uses as insulating materials, light-weight structural components, cushioning

materials, filters, and many others. Moreover, thermoplastic foams which are recyclable or

biodegradable are expected to replace traditional non-recyclable thermosetting foams.

To generate thermoplastic foams, dissolved gas molecules have to be converted into

spherical bubbles through cell nucleation and growth, which typically take place when the gas

phase becomes supersaturated and the surrounding conditions change too abruptly to allow a

smooth and quasi-equilibrium phase separation through diffusion and vaporization [Lee et al.

(2007)]. Once nucleated, cells continue to grow as gas diffuses into it, and growth continues until

the cell stabilizes (usually when the melt becomes hardened by cooling) or ruptures (when the

melt is overstretched). Figure 1-1 illustrates the formation of polymer/gas (BA) solution and the

generation of cells during microcellular foaming, an advanced technology requiring complete

dissolution of the BA in the melt and inducing microcellular structures through a rapid pressure

3

(a)

(b)

Figure 1-1 Illustrations of the main stages of microcellular processing: (a) formation of

polymer/blowing agent (gas) solution, reproduced from Park and Suh (1996); (b) cell nucleation

and growth

4

drop. The key is to control the dynamics of cell nucleation and growth by controlling rheological

properties, processing conditions, and geometries of the flow channels.

1.2 Rheology of Polymeric Fluids

Rheology is the science that deals with deformation and flow [Bird et al. (1987)]. A

polymer is a large molecule composed of many repeating carbon-based chemical units, called

structural units. Different side groups may be attached to the carbon chain, giving rise to

different chemical properties. A polymer melt is both viscous and elastic. That is, polymers

exhibit both fluid-like, viscous behaviour, such that the stress is related to the strain rate, and

solid-like, elastic behaviour, such that the stress is related to the strain. The elastic component is

often described by the dimensionless Deborah number. This number may be interpreted as the

ratio of the elastic forces to viscous forces, and is defined as the ratio of a characteristic time of

the fluid, λ, to a characteristic time of the flow system, 𝑡𝑓𝑙𝑜𝑤 [Bird et al. (1987)]

𝐷𝑒 = 𝜆/𝑡𝑓𝑙𝑜𝑤 (1-1)

The characteristic time of the fluid is related to molecular motion, while the flow

characteristic time depends on macroscopic motion. At a low strain rate, corresponding to a long

characteristic flow time, the fluid has ample time to relax into its equilibrium state, and its

behaviour is mostly viscous. At high strain rates, however, the polymer chains are stretched by

the flow and do not have time to relax. Fluid behaviour is then more elastic in nature. The

viscoelasticity gives rise to highly non-Newtonian properties such as a shear-rate dependent

viscosity, a extension-thickening viscosity, and normal stresses in shearing [Macosko (1994)].

To characterize the flow behaviour, or rheology, of a polymeric fluid, the material is

generally subjected to two basic motions: shearing and elongation. As illustrated in Figure 1-2(a)

and Equation (1-2), during shearing the distance between local fluid elements on neighboring

streamlines grows linearly in time, and the rate-of-strain tensor D contains only off-diagonal

components

5

(a)

(b)

Figure 1-2 Illustrations of simple flows for rheological characterization: (a) simple shearing; (b)

uniaxial extension; where V is the velocity

6

𝑫 =

0 𝛾 𝑦𝑥 𝑡 0

𝛾 𝑦𝑥 𝑡 0 0

0 0 0

(1-2)

where 𝛾 𝑦𝑥 𝑡 is the time-dependent shear rate. Figure 1-2(b) and Equation (1-3) illustrates

uniaxial elongation, which is free of vorticity and local fluid elements move apart exponentially

with time, resulting in a much stronger deformation. The rate-of-strain tensor for elongational

flow contains only diagonal components

𝑫 = − 1 + 𝑏 𝜀 𝑡 0 0

0 − 1 − 𝑏 𝜀 𝑡 00 0 2𝜀 𝑡

(1-3)

where 𝜀 𝑡 is the time-dependent elongation rate and b is the type of flow. The major types are

uniaxial elongation (𝑏 = 0), planar elongation (𝑏 = 1), and biaxial elongation (𝑏 = 0), all

relevant to foam processing as will be discussed later in this thesis [Bird et al. (1987)].

1.3 Motivations for the Study

1.3.1 Challenges with Characterizing Rheological Properties

The fluids of interest in foam processing are polymer melts and polymer/BA solutions,

both exhibiting viscoelasticity which influences the fluids‟ processing behaviours. However, two

major challenges arise when characterizing the rheological properties of these fluids. One is the

need for high pressure during measurement of polymer/BA solutions, so that the BA stays

dissolved in the melt. The other is the difficulty of determining extensional properties. To

address the first challenge, customized rheometers, both pressure-driven and drag-driven, have

been developed because commercial ones cannot be operated under sufficiently high pressures.

As a result, shear characteristics of polymer/BA solutions are now relatively well established, at

least compared to extensional characteristics. Such studies have shown that even a few percent of

BA can reduce the shear viscosity of a polymer melt significantly, and that the shape of the

solution viscosity curve is similar to that of the polymer alone at a higher temperature. The

customized rheometers, however, have not been successful in determining other rheological

properties of the polymer/BA solutions.

7

The difficulty of determining extensional properties pertains to both the polymer and the

polymer/BA solutions. In contrast to a Newtonian fluid, in which the extensional viscosity in

uniaxial extension 𝜂𝐸 is three times the shear viscosity 𝜂 , polymeric liquids can exhibit

extensional viscosities that are orders of magnitude higher than the shear viscosity [Macosko

(1994)]. This extreme strain hardening contributes directly to unexpected behaviour in

processing, such as an excess pressure drop in convergent channel flow, die swell, and a high

strain to break, which aids the production of polymer thin films and low-density foams.

Characterizing the extensional viscosity 𝜂𝐸 requires a shear-free flow, as prescribed by

Equation (1-3). This is a difficult task, but uniaxial extensional viscosity can be routinely

measured nowadays because the associated shear-free flow can be generated with commercial

instrumentations. Planar and biaxial extensional viscosities, however, which are also relevant to

foam processing, are rarely determined because commercial instrumentation is not available. At

extensional rates of industrial relevance (usually above 1 s-1

), planar extensional viscosity has

been determined approximately from planar contraction flows, and biaxial extensional viscosity

from stagnation flows [Macosko (1994)]. These measurements, obtained with laboratory

instruments, also seem to be the only methods available for evaluating the extensional properties

of polymer/BA solutions. Ladin et al. (2001) and Xue and Tzoganakis (2003) determined the

apparent extensional viscosity of polymer/CO2 solutions by measuring the pressure drop in a

sudden planar contraction. They found that BA reduces the extensional viscosity, but their

methods lack detailed control of the extensional flow, and their data involved large errors.

Details of these studies will be discussed in Chapter 4.

1.3.2 Importance of Material Rheological Properties to Foam Processing

During foam processing, rheological properties of the polymer and the polymer/BA

mixtures determine the distributions of pressure and velocity in the processing system, and

therefore the generation and growth of cells. Specifically, cell nucleation occurs when the system

pressure drops below a critical pressure, called the solubility pressure; cell nucleation density and

initial cell growth rate, then, are determined by the degree of supersaturation of the blowing

agent, a close function of the pressure drop rate in a die or a mold; and finally, the velocity

8

history of a local flow element defines its current rheological state, which influences cell

nucleation and growth within this element.

Rheological properties influence cell growth. For example, the extensional properties of a

polymer influence the deformation at the surface of an expanding cell, which is essentially

biaxial stretching. Many studies (e.g., Münstedt & Stange (2006) and Spitael & Macosko (2004))

have found that foams produced with a long-chain-branched polypropylene show more uniform

cell size distributions and fewer cell openings (i.e., fewer ruptures of cell walls) than foams

produced with a linear polypropylene. The viscosity reduction effect of BA is also important. It

has beneficially allowed foams to be processed at very low temperatures such that the melt is

strong enough to sustain expansion of the cells. As the BA diffuses into the cells, the melt is

quickly hardened, causing cell growth to slow down and even to stop.

Despite the importance of material rheological properties to foam processing, previous

attempts to quantitatively relate these properties to cell growth behaviour have only been

limitedly successful, mainly because no constitutive model is capable of describing the

rheological properties accurately, and actual foaming involves complicated interactions between

momentum, energy, and mass transports.

1.4 General Objectives

Given the difficulty to determine rheological properties relevant to foam processing, and

the need to relate these properties to cell nucleation and growth, the general objectives of this

thesis are: first, to determine the rheological properties, especially the extensional properties of

polymer melts and polymer/BA solutions.; secondly, to analyze foam processing based on

knowledge of the rheological properties, especially to investigate how rheological properties

determine cell morphology. More detailed objectives will be presented in Chapter 2 after a

literature review.

9

1.5 Overview of the Thesis

Chapter 2 reviews relevant literature, leading to detailed objectives of this thesis.

Chapters 3 and 4 present a technique for determining the shear and extensional viscosities of a

polymer melt and its BA solution. The technique involves measuring pressure drops over well-

defined flow channels, and the analyses for extracting the rheological properties are presented in

detail. The measured viscosities from channels are compared to viscosity data from commercial

rheometers. In Chapter 5, then, the extrusion foaming behaviour of polylactic acid (PLA) with

different rheological properties is presented, and the optimal material compositions to produce

low-density PLA foams are investigated. Finally, Chapter 6 summarizes the contributions of this

thesis, and concludes by recommending future research directions.

10

Chapter 2 Background and Literature Review

Thermoplastic foam processing is a physical process driven by the diffusion of a gaseous

blowing agent into and out of the polymer melt under processing conditions. The diffusion

process is heavily influenced by the flows induced during processing. For example, shearing can

increase significantly the mixing efficiency of polymer and gas bubbles, and cell nucleation

density is closely related to the rate of pressure drop of the polymer/BA solution. From an

engineering point of view, it is important to model the cell nucleation and growth processes, and

thereby to predict cellular structures. Existing models have been derived from conservations of

momentum, energy, and mass at the single cell level, but they are only partially useful in actual

processing. A major limitation of these models is the lack of an accurate description of the

rheological properties of viscoelastic polymer/BA solutions, particularly their extensional

properties.

In this chapter, we start by reviewing the major scientific issues related to microcellular

foaming technology in Section 2.1. This is followed by a review of rheometric techniques for

polymer melts and polymer/BA solutions, especially techniques for determining extensional

properties, in Section 2.2. In Section 2.3, then, both the cell model and relevant experimental

foaming studies are reviewed to clarify the relationship between rheological properties and cell

growth. Finally, in Sections 2.4, detailed objectives of this thesis are presented.

2.1 Thermoplastic Foams and Their Processing Technology

2.1.1 Categories of Foams

Thermoplastic foams have been categorized by cell size, expansion ratio, and cell wall

integrity. The categories are independent of processing technology and the thermoplastics used,

and they reflect different applications and different physical properties of the foams. The major

categories by cell size are conventional (coarse) foams, with an average cell size above 100 μm,

fine-celled foams, between 20 and 100 μm, and microcellular foams, with an average cell size

below 20 μm [Klempner and Sendijarevic (2004)]. Studies have focused on microcellular foams

11

in recent years, because smaller cell size reduces convection and increases insulation of the

foam. Some mechanical properties, such as impact strength and fatigue life, are also improved

when cell size decreases, an effect attributed to the cells‟ ability to absorb micro-cracks.

Performance-to-weight ratios are also improved when cell size decreases [Suh et al. (2000)].

The major categories by expansion ratio, defined as the ratio of polymer density to foam

density, are high-density foams, with expansion ratios below 4, medium-density foams, between

4 and 10, and low-density foams, with expansion ratios above 10. High-density foams have been

used in structural applications where mechanical properties are important, while low-density

foams have been used in insulation and packaging applications, where energy absorption is

important [Throne (1996)].

Finally, the major categories by cell wall integrity are open-cell foams and closed-cell

foams. The former have openings in the cell walls such that adjacent cells interconnect with each

other (see Figure 2-1(a)), and the latter have complete cell walls such that adjacent cells are not

connected (Figure 2-1(b)). Open-cell foams have been used as sound insulation materials and

filters while closed-cell foams are suitable for packaging and cushioning applications [Lee et al.

(2007)].

2.1.2 Microcellular Foam Processing - Formation of Polymer/BA Solution

A. Solubility

As illustrated in Figure 1-1(a), microcellular processing begins with the formation of a

polymer/BA solution. Only a soluble amount of BA should be injected into the polymer melt,

because excess BA results in undesirable voids. These voids suppress cell nucleation because the

BA molecules preferentially diffuse to larger cells, resulting in hollow cavities in the final

product [Park and Suh (1996)]. The solubility of a BA also determines the plasticizing (viscosity

reduction) limit and the maximum expansion ratio of the foams. For example, carbon dioxide,

which has a solubility limit in polymers comparable to that of CFC blowing agents, has been a

good candidate for making low-density foams. Nitrogen, on the other hand, is much less soluble,

and has been used for making high-density foams with a high cell number density. The solubility

12

(a)

(b)

Figure 2-1 Scanning electron microscopy (SEM) photos of: (a) open-cell structure; (b) close-cell

structure; reproduced from Park et al. (1998).

13

limit usually depends on the polymer, the BA, the temperature, and the pressure. In general, only

loosely-packed phases in the polymer, usually amorphous phases, can dissolve BA, and densely-

packed phases, such as crystalline phases and solid particles, do not dissolve much BA [Tomasko

et al. (2003)].

B. Diffusivity

The diffusivity (or diffusion coefficient) of BA in a polymer is one of the key parameters

that determine the time needed to dissolve the BA and the kinetics of phase separation (i.e., cell

nucleation and growth). In general, BA diffusivity in a polymer follows an Arrhenius-type

temperature-dependence [Lee et al. (2007); Bird et al. (2002)]

𝐷 = 𝐷0 ∙ 𝑒𝑥𝑝 −∆𝐸𝐷/𝑅𝑔𝑇 (2-1)

where 𝐷0 is the maximum diffusion coefficient (at infinite temperature), ∆𝐸𝐷 is the activation

energy for diffusion, 𝑅𝑔 is the molar gas constant, and T is the absolute temperature. For

example, the typical diffusivity of CO2 and N2 in a thermoplastic at 200oC is 10

-6 cm

2/s, and that

at room temperature is 10-8

cm2/s [Bird et al. (2002)]. The much higher diffusivity at processing

temperatures facilitates formation of the polymer/BA solution in the extruder. The diffusivity

also depends on the type of blowing agent. In general, a BA with a lower molecular weight

exhibits higher diffusivity under the same temperature and pressure. This characteristic has

caused the cell nucleation and growth kinetics of CO2 and N2 to be much higher than those of

CFC blowing agents, making CO2 and N2 foaming more difficult to control.

C. Convective Diffusion and Convective Cooling

Convective diffusion and cooling are important characteristics of the microcellular

processing in this study, and they are briefly reviewed here. During microcellular processing, the

two-phase polymer/BA mixture evolves into a single-phase solution through gas diffusion under

elevated temperature and pressure. Before delivering the solution to the foaming element, usually

a die or a mold, the solution has to be cooled to achieve temperature uniformity and to increase

14

melt strength. Both mass transfer by molecular diffusion and heat transfer by conduction are very

slow processes when they occur in polymer melts under laminar flow conditions [Tadmor and

Klein (1970)]. Park and Suh (1996) demonstrated that convective flows can accelerate both

processes, thereby enabling microcellular processing to attain industrial efficiency. According to

mixing theory, convective flows bring fluid particles with a lower BA concentration or a lower

temperature into contact with particles with a higher BA concentration or a higher temperature,

thereby accelerating diffusion and cooling by inducing higher concentration and temperature

gradients [Tadmor and Klein (1970)].

Convective flows, i.e., flows driven by the bulk motion (observable movement) of fluids,

during polymer/BA mixing are induced by mixing sections on the plasticating screw, as

illustrated in Figure 2-2. A characteristic time 𝑡𝐷 to dissolve the BA completely [Tadmor and

Klein (1970)] is

𝑡𝐷 =𝑙𝐷

2

𝐷 (2-2)

where 𝑙𝐷 is the striation thickness, the average distance between two adjacent BA bubble

surfaces, and 𝐷 is the diffusion coefficient introduced in Equation (2-1). This time has been

estimated as 50 μm during typical extrusion processing [Park and Suh (1996)], and the time

needed to dissolve the BA in this case is only 20 s, giving rise to very high mixing efficiency.

Convective flows during cooling are induced by a second extruder with a cooling screw

and/or a heat exchanger containing static mixers, the latter being illustrated in Figure 2-3. Unlike

the mixing sections on the first plasticating screw, which induces mixing in the axial and circular

direction, the cooling screw and the static mixers induce mixing in the radial direction, cooling

the melt by keeping the extruder barrel at a low temperature. Because static mixers generate

pressure drops, they are usually used in systems with relatively low flow rates. In this thesis, the

processing systems are designed to induce both convective diffusion and convective cooling.

Details of the systems will be presented in Chapters 3.

15

Figure 2-2 Mixing elements on the plasticating screw

16

Figure 2-3 Schematic of heat exchanger containing static mixers used in foam extrusion

17

2.1.3 Microcellular Foam Processing - Cell Nucleation and Growth

A. Thermodynamics of Cell Nucleation

As illustrated in Figure 1-1(b), phase separation occurs when the BA becomes

supersaturated in the melt, usually as the result of a quick pressure drop. The excess amount of

BA diffuses out of the melt mainly through cell nucleation and growth. According to the

classical nucleation theory, an excess Gibbs free energy has to be exceeded in order to create the

bulk (bubble) phase and the bubble surface [Colton and Suh (1987)]

∆𝐺 = −𝑉𝑏 ∙ ∆𝐺𝑉 + 𝐴 ∙ 𝜍 (2-3)

where 𝑉𝑏 is the initial volume of the bubble, ∆𝐺𝑉 is the free energy difference between the

bubble phase and the polymer phase, A is the interfacial area (bubble surface area), and σ is the

surface tension. If the bubble is generated from a single homogeneous phase without impurity or

dirt, the process is called homogeneous nucleation. This is rarely the case, however, because

most polymers contain additives or impurities. If bubbles are formed at solid/liquid interface,

e.g., at the surface of foreign particles, the process is called heterogeneous nucleation.

For homogeneous nucleation into spherical bubbles, Equation (2-3) can be rewritten as

∆𝐺 = −4

3𝜋𝑟3 ∙ ∆𝑃 + 4𝜋𝑟2𝜍 (2-4)

where ∆𝑃 = 𝑃𝑠𝑜𝑙 − 𝑃𝑠 is the difference between the solubility pressure, i.e., the pressure required

to prevent phase separation, and the system pressure, i.e., the pressure sensed by a transducer,

and r is the initial bubble radius. The relationship between ∆𝐺 and r is plotted in Figure 2-4. It

suggests that in order for a bubble to grow larger, it has to exceed a critical radius 𝑟∗ =

𝑟 𝜕∆𝐺

𝜕𝑟=0

=2𝜍

∆𝑃. The free energy needed to form this critical nucleus is

∆𝐺𝑕𝑜𝑚∗ =

16𝜋𝜍3

3∆𝑃2 (2-5)

The nucleation rate, the number of bubbles formed per unit time, can also be calculated

following classical nucleation theory [Colton and Suh (1987)]

𝑁𝑕𝑜𝑚 = 𝐶𝑕𝑜𝑚 𝑓𝑕𝑜𝑚 ∙ 𝑒𝑥𝑝 −∆𝐺𝑕𝑜𝑚∗ /𝑘𝐵𝑇 (2-6)

18

Figure 2-4 Free energy of forming bubble from supersaturated polymer melt

19

where 𝐶𝑕𝑜𝑚 and 𝑓𝑕𝑜𝑚 relate to the kinetics of gas diffusion, 𝑘𝐵 is the Boltzmann‟s constant, and

T is the temperature. It is worth noting that if the rate of pressure drop is increased, such that ∆𝑃

becomes larger, the free energy to form bubbles will decrease, the cell nucleation rate will

increase, and the average cell size will decrease accordingly. This explains why pressure drop

rate is closely related to the cell morphology during microcellular processing.

The more common type of nucleation in a polymer melt is, however, heterogeneous

nucleation. Here the free energy barrier may be significantly reduced if the bubble is formed at a

solid/liquid interface. As illustrated in Figure 2-5, for a flat solid surface and a wetting angle of θ

for the liquid phase, the free energy barrier, Equation (2-3), is now a function of θ. The

nucleation rate in this case takes on a form similar to that of homogeneous nucleation [Ramesh et

al. (1994a) and (1994b)]

∆𝐺𝑕𝑒𝑡∗ = ∆𝐺𝑕𝑜𝑚

∗ ∙ 𝑓 𝜃 =16𝜋𝜍3

3∆𝑃2 ∙ 1

4 2 + 𝑐𝑜𝑠𝜃 1 − 𝑐𝑜𝑠𝜃 2 (2-7)

𝑁𝑕𝑒𝑡 = 𝐶𝑕𝑒𝑡𝑓𝑕𝑒𝑡 ∙ 𝑒𝑥𝑝 −∆𝐺𝑕𝑒𝑡∗ /𝑘𝐵𝑇 (2-8)

The free energy barrier can be further reduced if the solid surface has cavities. Many

types of solid particles have been used as nucleating agents during foam processing, talc being

the most common choice, and the search for the “ideal” nucleating agent is likely to continue

[Lee & Ramesh (2004); Spitael et al. (2004)].

B. Cell Growth and Stabilization

Once nucleated, cells continue to grow until they are either stabilized by cooling or

ruptured by overstretching. The growth process is very complicated because many variables

influence the polymer‟s rheological response to the deformation induced by expansion. With

amorphous polymers, as an example: first, when BA dissolves in the melt, it reduces both its

viscosity and its elasticity; secondly, as the BA bubbles grow, gas diffusion and expansion

induce cooling and polymer viscoelasticity increases due to the loss of BA; thirdly, the bubble

growth rate changes over time, again influencing the transient rheology of the polymer melt; and

finally, cell growth continues until the polymer reaches its 𝑇𝑔 or when the melt reaches its

20

Figure 2-5 Surface force balance for a heterogeneously nucleated bubble, where θ is the wetting

angle and σ is the surface tension

21

stretching limit, causing cell wall opening or cell coalescence.

Much effort has been expended to model cell growth. The models have evolved from a

single bubble surrounded by fluid with an infinite amount of BA available for its growth [Barlow

& Langlois (1962); Street (1968)] to the more recent “cell models” in which the melt is divided

into unit cells of equal and constant mass, each one consisting of a liquid envelope surrounding a

single bubble [Amon and Denson (1984), (1986)]. The cell model will be discussed in relation to

actual processing in Section 2.3.

2.2 Rheometry of Polymer Melts and Polymer/BA Solutions

2.2.1 Shear Rheometry of Polymer/BA Solutions

It is well known that the dissolution of a small amount of low-molecular-weight blowing

agent can reduce the viscosity of a polymer melt significantly [Ferry (1980)]. In this section we

review customized drag-driven and pressure-driven rheometers developed to measure

polymer/BA solution viscosity under sufficiently high pressures.

Drag-driven rheometers usually operate at low shear rates, because several types of flow

instability occur for viscoelastic fluids at even modest shear stresses and rates [Macosko (1994)].

Royer et al. (2002) developed a magnetically levitated sphere rheometer to study the viscosity of

a PDMS/CO2 solution. More recently, Wingert et al. (2009) determined the viscosity of a

PS/CO2 solution using a pressurized Couette viscometer. The maximum shear rate was 0.5 s-1

for

both rheometers. The most successful drag-driven rheometer has been the high pressure sliding

plate rheometer developed by Park and Dealy (2006). It shears a polymer film between two

parallel plates in a BA-pressurized chamber. The solution viscosities of HDPE/CO2 and

HDPE/N2 were determined over a wide ranges in pressure (0 to 50 MPa), in BA concentration (0%

to 25% by weight), and in shear rates (0.1 to 100 s-1

).

In general, a drag-driven rheometer requires a small amount of polymer, and, unlike a

pressure-driven rheometer, the stress and the strain rate are uniformly distributed over the sample.

The drag-driven devices have several limitations, however, which have prevented them from

widespread use. First, they require a long saturation time before measurement can start, because

22

the BA diffuses slowly into the melt in the pressure chamber, and reaching equilibrium can take

hours [Park and Dealy (2006)]. Secondly, the BA concentration has to be determined from a

separate solubility measurement, which takes time and is subject to error. Finally, many drag-

driven rheometers require the signals (torque, force, and displacement) to be transferred under

pressure through a dynamic seal, which may introduce serious errors. The last limitation has

been solved, at least partially, using magnetic signal transmissions or, in the case of sliding plate

rheometer, using local shear stress transducers.

Contrary to drag-driven rheometers, pressure-driven ones are ideal for high-shear-rate

measurements. According to a review by Tomasko et al. (2003), these rheometers fall into two

categories: in-house rheometers and customized commercial ones. Rheometers of the first

category usually consist of an extruder for generating a single-phase solution, a die at the

extruder exit for measuring pressure drops, and a regulator at the die exit for keeping the channel

pressure high (see Figure 2-6). Different types of extruders have been used to generate solutions.

These include single-screw extruders [Han and Ma (1983a) and (1983b); Lee et al. (1999) &

(2000); Royer et al. (2000) & (2001); Areerat et al. (2002)], tandem systems [Ladin et al.

(2001)], and twin-screw systems [Elkovitch et al. (1999), (2000), and (2001)]. The second

category includes customized capillary rheometers with high pressure seals [Gerhardt et al.

(1997) & (1998); Kwag et al. (1999) & (2001)] and a gear-pump driven online rheometer used

by Gendron et al. (1996), (1997), and (1998). The gear pump rheometer is useful in industry

because measurements can be performed at any time by sampling the melt stream, without

affecting the ongoing production.

Overall, pressure-driven techniques are convenient and reliable means of determining

viscosity. In particular, when a rectangular channel with a high aspect ratio is used, pressure

transducers can be flush mounted on the channel wall to eliminate pressure errors [Macosko

(1994)]. The BA concentration is also easily controlled by a gas pump. Pressure-driven

rheometers also suffer from several limitations: One limitation is a potentially-large temperature

gradient in the flow channels, caused by poor mixing or viscous heating, which may affect the

accuracy of data; another limitation is the shear history of the fluid preceding pressure

measurement; and finally, the variation of pressure along the flow channel makes measurement

at controlled pressure impossible. In spite of these limitations, pressure-driven rheometers

determine viscosity under conditions similar to those in processing, and are therefore chosen in

23

Figure 2-6 Schematic of an extruder setup for measuring PS/CO2 solution viscosity; reproduced

from Lee et al. (1999).

24

this thesis to characterize the viscous resistance.

2.2.2 Extensional Rheometry Involving Shear-Free Flows

Extensional properties of polymer melts and polymer solutions play a dominant role in

several processing techniques, including fiber spinning, film blowing, blow molding, and foam

processing. Despite their importance, few extensional properties are available compared to shear

properties, because controlling shear-free flows, like those prescribed by Equation (1-3), are very

difficult. When solid boundaries are used to control the flow, shear is introduced, which makes it

difficult to extract extensional properties [Macosko (1994)]. In this section, we briefly review

shear-free techniques. Techniques involving mixed flows will be reviewed in Section 2.2.3.

Perhaps the most versatile and accurate shear-free technique is achieved using a rotary

clamp device, such as those described in Meissner (1987) and Meissner and Hostettler (1994).

These devices can generate the major types of elongational flows, i.e., uniaxial, planar, and

biaxial flows, as well as combinations of the major types. Two possible configurations of the

rotary clamp devices, for measuring planar and biaxial extensional viscosities, are shown in

Figure 2-7. A constant strain rate is maintained by rotating the clamps at a constant speed, and

the type of elongation is determined by the clamps‟ geometrical arrangement. Extensional

viscosities are calculated from the stresses sensed at the clamps.

The rotary clamp devices, however, involve complicated designs, and are incapable of

generating extensional rates above 0.5 s-1

, except for uniaxial extension. Temperature control is

also difficult because a large sample is necessary. Recently, Sentmanat (2003) and (2004)

developed a compact shear-free fixture, called the Sentmanat Extensional Rheometer (SER) that

can be mounted on commercially available rotational rheometers. Two counter-rotating windup

drums replace the rotary clamps, and both constant and variable uniaxial extensional rates can be

induced by programming the rotation rate of the drums. The viscosity data from the SER was

found to agree well with that from the rotary clamp devices [Sentmanat et al. (2005)]. In this

study, a device similar to the SER will be used to determine the uniaxial extensional viscosity of

polymer melt without BA. Details of the device will be presented in Chapter 4.

25

(a)

(b)

Figure 2-7 Rotary clamp rheometers for measuring: (a) planar extensional viscosity; (b) biaxial

extensional viscosity; reproduced from Meissner (1987).

26

Instead of stretching a polymer sample, a shear-free flow can also be induced in a

squeeze-film flow. The sample ends are usually lubricated with a lower viscosity liquid to

eliminate shearing during compression, and the resultant devices are called lubricated squeezing

rheometers [Chatraei et al. (1981)]. Biaxial extensional viscosity has been determined by radial

flow, and planar extensional viscosity has been determined by keeping one dimension of the

sample constant. It was found that strain hardening during planar or biaxial extension is weaker

than that during uniaxial extension. Measurements could not be made under extensional rates

above 1 s-1

or Hencky strains above 1, however, because of the loss of lubricant [Macosko

(1994)].

2.2.3 Extensional Rheometry Involving Mixed Flows

Although shear-free rheometers allow measurement of the true extensional viscosity, they

may not be the most relevant devices for studying foam processing. In the first place, shear-free

devices cannot generally achieve industry-relevant extensional rates, usually of order 1 𝑠−1 ;

secondly, polymer/BA solutions cannot be measured on a shear-free rheometer because the BA

evaporates; and finally, during processing, both shearing and extension are present, and it is of

interest to study material rheological responses under mixed flow conditions. One technique to

overcome the limitations of shear-free devices is to determine extensional properties from

entrance pressure drops.

When a viscoelastic fluid flows internally from a large cross section to a smaller one, the

streamlines converge, producing an extensional flow. Because extensional stresses are produced,

an entrance pressure drop ∆𝑃𝑒𝑛 is necessary. The entrance pressure drop, though, is not uniquely

determined by the extensional flow component, because the shear component, caused by the

channel walls, also affects the velocity and stress distributions.

Flow through a sudden contraction is illustrated in Figure 2-8. The flow progresses from

being fully developed at some distance upstream from the contraction to being almost (e.g.,

98%) fully developed a distance 𝐿𝑒 downstream. Depending on the Reynolds number of the flow

and the characteristics of the fluid, a secondary-flow vortex may be present in the corner of the

upstream channel. In Figure 2-8, the Hencky strain is 𝐷𝑢/𝐷𝑑 2 for axisymmetric contraction

27

Figure 2-8 Basic elements of an entry flow for flow from a large tube through an abrupt entry

into a small tube. The illustration applies to both axisymmetric contraction and planar

contraction; reproduced from Boger (1987)

28

and 𝐷𝑢 /𝐷𝑑 for planar contraction. The extensional rate, then, is related to both the flow rate and

the channel cross section. A “better” design of the flow channel, a hyperbolic convergent

channel, has been proposed for extensional measurement [Kim et al. (1994)]. With this shape,

vortices are minimized or nonexistent, and a constant extensional rate is induced along the

centreline and in the region around it. Accounting for the shear stress, though, is not

straightforward, and previous studies simply assumed shear-free flow in the channel [Feigl et al.

(2003)]. For a planar hyperbolic channel, generating an adequate pressure drop due to extension

has been difficult, because a large aspect ratio is required to induce a two-dimensional flow. This

leads to low extensional rates, and hence weak strain hardening of the fluid, because the flow

rates of lab-scale experiments are usually limited [Kim et al. (1994)].

Two widely used analyses for estimating extensional viscosity from measurements

of ∆𝑃𝑒𝑛 are those by Cogswell (1972) and Binding (1988). Both analyses assume that ∆𝑃𝑒𝑛 can

be separated into shear and extensional components, that velocity distribution is determined by

shearing only, and that, in the case of a sudden contraction, the vortex size is determined by

minimizing pressure drop over the contraction. For a planar contraction, Cogswell calculated the

stresses and the vortex size by applying a simple force balance on an elemental wedge and

minimizing the sum of elemental pressure drops along the contraction. From his analysis, the

average extensional rate is [Macosko (1994)]

𝜀 =𝜏𝑤 ∙𝛾 𝑎

3∙ 𝜏11−𝜏22 (2-9)

where 𝜏𝑤 and 𝛾 𝑎 are the wall shear stress and the apparent shear rate in the downstream channel,

repectively. The apparent planar extensional viscosity is calculated from the normal stress 𝜏11 −

𝜏22 and the average extensional rate 𝜀

𝜂𝑃,𝑎𝑝𝑝 =𝜏11 −𝜏22

𝜀 =

1

2∙𝜀 𝑛 + 1 ∙ ∆𝑃𝑒𝑛 (2-10)

where n is the non-Newtonian index in the power law model.

Binding (1988) calculated the stresses and the vortex size for a sudden contraction by

applying energy balance and minimizing the energy consumption. Interestingly, Cogswell‟s and

29

Binding‟s approaches are related, and it has been shown that both analyses predict the same

values of apparent extensional viscosities [Tremblay (1989)]

2.3 Relationship between Cell Growth and Rheological Properties

Section 2.1.3 B briefly explains the relationship between cell growth and the rheological

properties of polymer/BA mixture. To be more specific, cell growth is driven by the pressure

difference between the cell and its surrounding, as a result of both BA diffusion into the cell

(diffusion-driven growth) and the decrease of system pressure (pressure-driven growth) [Amon

and Denson (1984)]. The dynamics of cell growth, then, is related to both this pressure difference

and the rheological properties of the polymer/BA mixture. A quantitative understanding of this

relationship not only allows cell growth to be controlled, thereby inducing desired cell

morphology, but also yields information about the type, rate, and strain of deformation relevant

to foaming, which is critical to the optimization of the chemical structure and composition of

foaming polymers. Information about cell growth comes from either theoretical modeling,

usually using the cell model [Amon and Denson (1984), (1986)], or from direct observation of

cell growth [Guo et al. (2006)]. In this section, we briefly review the results from these two

approaches and from previous studies aimed at relating material rheological properties to the

final cell morphology.

According to the cell model (Figure 2-9), each cell is surrounded by an envelope of melt,

of which the volume is inversely proportional to the cell density. The time-dependent momentum

balance near the cell surface is [Amon and Denson (1984)]

𝑃𝑔 𝑡 − 𝑃𝑠 𝑡 −2𝜍

𝑅 𝑡 + 2 ∙

𝜏𝑟𝑟 −𝜏𝜃𝜃 𝑡

𝑟

𝑅𝑠 𝑡

𝑅 𝑡 ∙ 𝑑𝑟 = 0 (2-11)

where 𝑃𝑔 𝑡 = 𝐶𝑅 𝑡 /𝑘𝐻 is the bubble pressure, with 𝐶𝑅 𝑡 being the BA concentration at cell

surface and 𝑘𝐻 the Henry‟s Law constant; 𝑃𝑠 𝑡 is the system pressure, the pressure measured by

a transducer; 𝜍 is the surface tension, usually negligible in calculation; 𝑅 𝑡 is the bubble

radius, 𝑅𝑠 𝑡 is the radius of the outer envelope, and 𝜏𝑟𝑟 − 𝜏𝜃𝜃 is the normal stress difference in

the melt.

30

Figure 2-9 Distribution of BA concentration in the surrounding of an expanding cell (a gas

bubble), as described by the cell model; 𝑐 𝑟, 𝑡 is the BA concentration, 𝑐𝑅 𝑡 is the BA

concentration at cell surface, 𝑃𝑔 𝑡 is the cell pressure, 𝑘𝐻 is Henry‟s Law constant, 𝑅 𝑡 is the

cell radius, and 𝑅𝑠 𝑡 is the outer radius of the melt envelope [Amon and Denson (1984)]

31

When a proper constitutive equation is chosen, Equation (2-11) allows the calculation of

cell growth rate and stress distribution from basic physical properties of the polymer/BA mixture,

such as the viscosity, the relaxation time, and the diffusivity of the BA. The calculated values

have been compared to experimentally observed cell growth rate, and the two were found to

agree reasonably [Leung et al. (2006)]. For typical processing conditions, the extensional rate

reaches a maximum following cell nucleation, usually of order 𝑂 10 𝑠−1 , and then slows down

monotonically due to strain hardening or cooling of the melt and decreased pressure difference at

the cell surface.

The Hencky strain at cell surface is obtained by integrating extensional rate over time

𝜀𝐻 = 𝜀 𝑅 𝑡 ∙ 𝑑𝑡𝑡

0= −𝑙𝑛

𝑅 𝑡

𝑅0 (2-12)

where R0 is the initial bubble radius, usually between 0.1 μm and 10 μm [Leung et al. (2006)].

The Hencky strain increases with the expansion ratio of the foams, and usually ranges between 2

and 6. Such strain can induce significant hardening in branched polymers and rupture of the melt

for both linear and branched polymers [McKinley and Hassager (1999)].

The above information about cell growth has been used in the literature to select or

synthesize polymers with optimized rheological properties for foaming. Most of these studies

focused on polypropylene, because of both the commercial value of PP foams and the

(undesirable) low cell density and low expansion ratio associated with conventional linear PP.

Park and Cheung (1997) and Naguib et al. (2002) used a long-chain-branching polypropylene

(LCB-PP), which exhibits significant strain hardening under extension, and demonstrated much

higher cell density and expansion ratio during foam extrusion compared to linear PP. Similar

results were obtained by Michaeli et al. (2004). All three studies attributed the better foaming

behaviour of branched PP to its higher melt strength and reduced cell coalescence during the

early stage of cell growth. Recently, Spitael & Macosko (2004) and Stange & Münstedt (2006)

characterized the uniaxial extensional viscosities of a series of linear PPs, LCB-PPs, and their

blends at conditions relevant to foaming, and attempted to relate rheological properties to cell

morphology. Besides showing that long chain branching suppresses cell coalescence, they found

that even a small amount of LCB-PP (e.g., 10% by weight) in the blend can improve the

expansion and reduce the cell opening of linear PP. Stange & Münstedt attributed the higher

32

volume expansion of LCB-PP and blends containing LCB-PP to their higher strains at rupture

and their more uniform deformation during extension compared to linear PP. The maximum

expansion ratio achieved by Stange & Münstedt, however, was only 3 times, and the associated

Hencky strain was relatively low. Spitael & Macosko were able to produce high-expansion-ratio

PP foams and induce high strain in the melt, but they did not find any direct correlation between

strain hardening and cell density or expansion ratio. The conclusions regarding linear PP and

LCB-PP still need to be confirmed for other polymers, because the interaction between the

polymer and the blowing agent can be very different. A detailed discussion of these issues is

found in Chapter 5.

2.4 Objectives of the Thesis

The objectives of this thesis are:

(1) To determine experimentally the rheological properties, especially extensional properties,

of a polymer melt and its BA solution. The rheological properties of the fluids will be

determined on various commercial rheometers, and the properties of polymer/BA

solution will be determined on an in-house rheometer by measuring pressure drop over

well-defined flow channels. Data from the two types of rheometers will be compared to

validate the in-house technique, and the influence of BA on the rheological properties

will be studied.

(2) To investigate the relationship between polymer rheological properties, especially

extensional property, and cell morphology, including cell size, cell density, volume

expansion, and cell opening. As mentioned in Section 1.5, a polymer other than

polypropylene will be used, and both linear and branched structures will be studied. The

results will be compared to those for polypropylene in the literature, and general

conclusions will be attempted.

33

Chapter 3 Characterization of the Shear Properties

This chapter describes a technique for determining the rheological properties of a

polymer melt and a polymer/blowing agent solution. The technique is based on measuring

pressure drops in a high-aspect-ratio straight rectangular channel for determining the shear

viscosity, and in a thin convergent rectangular channel with a hyperbolic shape for determining

extensional viscosity with the pressure drop due to shearing accounted for. For both the shear

and the extensional viscosities, channel values for the polymer alone will first be compared to

values from commercial rheometers. Then the technique will be utilized to determine the

corresponding rheological properties of a polymer/BA solution. This chapter describes

measurement of the shear viscosities, while determination of the extensional viscosity will be

presented in Chapter 4. For the shear viscosities, channel data of the polymer alone will be

compared to data from a commercial rotational rheometer and a commercial capillary rheometer.

The viscosity of the polymer/BA solution will be found from channel data, and will be compared

to previous measurements and predictions by the free volume theory.

3.1 EXPERIMENTAL

3.1.1 The Hele-Shaw Channels

The term „Hele-Shaw‟ is normally associated with slow flow of a Newtonian fluid

through a narrow gap between two parallel plates (a Hele-Shaw cell) [Batchelor (2000)]. In this

study, the term still refers to slow flow through a thin channel but the fluids will be viscoelastic.

Two dies were made with the same shape, which is illustrated in Figure 3-1, and different

thicknesses. The inlet section of each die is a three-dimensional wedge-shaped diffuser that

distributes tube flow to rectangular flow evenly. The subsequent Hele-Shaw section consists of a

high-aspect-ratio straight channel followed by a hyperbolic channel. The channel shape was

machined out of a metal insert to a specific depth, and the insert was then bonded to a flat plate

to create a channel with a rectangular cross section everywhere. The profile of the hyperbolic

channel, given by Equation (3-1), subjects a Newtonian fluid to a constant rate of extension near

the centreline [Kim et al. (1994)]

34

Figure 3-1 The two test dies. The circles indicate the diaphragms of the pressure transducers.

The dimensions are: 𝐵0 = 30 𝑚𝑚, 𝐵1 = 3 𝑚𝑚, 𝐿0 = 20 𝑚𝑚, 𝐿1 = 5 𝑚𝑚, 𝐿2 = 20 𝑚𝑚. The

depth H (into the page) is 𝐻 = 0.94 𝑚𝑚 or 1.96 𝑚𝑚. and –𝐻

2< 𝑧 <

𝐻

2

B0

L0

L2 L1

B1

Flow x

y

z

B(x)

ΔP shear ΔP convergent

35

𝐵 𝑥 = 1

𝐵0+

1

𝐵1−

1

𝐵0 ∙

𝑥

𝐿2

−1

, 0 ≤ 𝑥 ≤ 𝐿2 (3-1)

The dimensions in Figure 3-1 indicate that the aspect ratios of the straight sections were

32 and 15.3. The convergent section varied in width from 30 mm at the inlet to 3 mm at the

outlet, so that the contraction ratio was 10:1. The channel depths (or thicknesses) were 0.94 mm

and 1.96 mm. As indicated by the figure, two pressure transducers (PT462E-M10, Dynisco Inc.)

were flush-mounted along the centreline of the straight channel. A third transducer was installed

in the downstream reservoir immediately after the die exit. The pressure drop between the first

two transducers (∆𝑃𝑠𝑕𝑒𝑎𝑟 ) is caused by shear only, while the pressure drop between the second

and the third transducers (∆𝑃𝑐𝑜𝑛𝑣𝑒𝑟𝑔𝑒𝑛𝑡 ) is caused by both shear and extension. The three

transducers were calibrated at a representative processing temperature using a dead-weight tester

(M2800/3, Druck Inc.). Detailed design of the die (CAD drawing) can be found in the Appendix

of this thesis. While the Hele-Shaw channel was designed for rheological purposes, the channel

shape is not unrelated to the world of polymer processing; that is, the geometry is similar to the

shape of many injection molds and bears a close resemblance to annular and sheet extrusion dies.

3.1.2 The Processing System

Measurements with the Hele-Shaw channels were conducted on a tandem extrusion

system, shown in Figure 3-2, the type commonly used for foam processing [Lee et al. (2007)].

The system consists of a 0.75”-diameter plasticating extruder with a mixing screw, a 1.5"-

diameter extruder with a cooling screw, a syringe pump for injecting the blowing agent, a gear

pump, and a heat exchanger containing homogenizing static mixers. The first extruder plasticates

the polymer resin and disperses the blowing agent into the polymer melt. The mixing section on

the plasticating screw is already illustrated in Figure 2-2. The gear pump controls the flow rate,

independent of temperature and pressure. The second extruder provides further mixing and initial

cooling of the melt, and the heat exchanger, the design of which is already shown in Figure 2-3,

removes the remaining heat for testing at a particular temperature. The Hele-Shaw channel is

attached to the exit of the heat exchanger. Following the channel is a reservoir where the

pressure – the „back pressure‟ – is controlled by a valve. With the valve, a constant average

pressure can be maintained in the channel during flow measurements, so that the known effect of

36

Figure 3-2 Schematic of the tandem extrusion system [Ladin et al. (2001)]

firstextruder

second extruder

syringegaspump

gear pump

Hele-Shawchannel

pressure transducers

valve

heat exchanger

PS

CO2

ΔPshear ΔPconvergent

reservoir

37

pressure on the viscosity becomes negligible. The mass flow rate was found by collecting a

sample at the valve exit over a known time and weighing it. For all experiments, flow data were

not recorded until a steady state had been reached, typically in 10 to 20 minutes.

3.2 Properties of the Polymer and the Polymer/Blowing Agent Solution

3.2.1 General Physical Properties

The polymer was a general-purpose polystyrene from Dow Chemical Inc. (grade Styron

685D, 𝑇𝑔 = 100𝑜𝐶, 𝑀𝐹𝐼 = 1.5 𝑔/𝑚𝑖𝑛, 𝑀𝑛 = 120,000 𝑔/𝑚𝑜𝑙, 𝑀𝑊/𝑀𝑁 = 2.6, and no mineral

oil content), a polymer commonly used for foam processing. The blowing agent was CO2 of 99%

purity from BOC Inc. The concentration of CO2 in the PS was 5% by weight, a typical value in

industrial processes. Melt and solution densities depend on pressure, temperature as well as on

CO2 concentration, and values at two representative temperatures are presented in Figure 3-3.

The data were determined from PVT measurements of a PS droplet in a CO2-pressurized

chamber [Li (2008)]. The plot shows the melt density varying little with temperature and

pressure, and the solution density by no more than 10%. Consequently, a constant density of 1.0

g/cc was assumed for both materials in the calculations which follow. According to solubility

measurements using a CO2-pressurized magnetic suspension balance [Li et al. (2004)], the

minimum pressure to dissolve CO2 in polystyrene is 1650 psi at 150oC and 1850 psi at 190

oC.

Pressures higher than these were maintained throughout the tandem system to prevent phase

separation. The zero-shear-rate viscosity of the polystyrene was measured before and after

processing. No difference was found, indicating that shear degradation in the tandem system was

negligible. Samples collected at the exit of the thinner (𝐻 = 0.94 𝑚𝑚) Hele-Shaw channel at a

low processing temperature (172oC) showed little melt fracture. Consequently, no slip at the

boundary was assumed to hold, at least approximately. Nucleating agents, often used in foaming,

were not used in this study.

38

0 1000 2000

0.88

0.90

0.92

0.94

0.96

0.98

1.00

1.02

De

nsity (

g/c

m3

)

Pressure (psi)

Melt and solution

density

150oC, PS

200oC, PS

150oC, PS/CO

2

200oC, PS/CO

2

Figure 3-3 Density of polystyrene and polystyrene/CO2 solution; Extracted from Li (2008)

39

3.2.2 Rheological Properties Using Commercial Rheometers

A. Shear Properties

Shear properties of the polystyrene were determined using several commercial

rheometers. Oscillatory shear and low-shear-rate data were obtained with an ARES rheometer

(TA Instruments Inc.) using 25-mm parallel disks and a 0.8-mm gap. For high shear rates,

viscosity values were determined using a twin-bore capillary rheometer (RH2000, Malvern Inc.).

For this device, melts were extruded through a 1-mm diameter, zero length capillary die and a 1-

mm diameter, 20-mm long capillary die at the same time. The pressure drop over the zero length

die is subtracted from that over the long die to account for entrance and exit pressure drops. The

remaining pressure drop over the long die is due to shearing alone, and allows the shear viscosity

to be calculated. The obtained viscosity data will be presented in Section 3.3, along with data

from the Hele-Shaw channels.

The linear viscoelastic properties, 𝐺 ′ and 𝐺", were measured at several temperatures, and

a single characteristic relaxation time was determined for each temperature from the crossover

point. This relationship between the moduli and this relaxation time may be derived from

Macosko (1994, pp. 124) by equating Equation (3.3.31) with (3.3.32) there and setting k=1. The

relaxation times, given in Table 3-1, were generally longer than 1 s, indicating a highly elastic

material.

Table 3-1 Relaxation times of PS685D, based on oscillatory shear data

Temperature (oC) 172 190 210

Relaxation time λ (s) 18.1 3.3 0.83

B. First Normal Stress in Shearing and Exit Pressure

In Chapter 4, it will be shown that the pressure drop in the hyperbolic channel in Figure

3-1 depends on the pressure at the exit, but the pressure there is different from that measured in

the downstream reservoir. The difference is the exit pressure (∆𝑃𝑒𝑥 ), and the only way of

estimating is to use the first normal stress difference N1 [see, e.g., Macosko (1994)]. However,

40

N1 could not be measured above about 1 s-1

, which was far below the channel shear rates,

because of edge failure. Instead N1 was estimated using, at low shear rate, the relation 𝑁1 𝜔 =

2 ∙ 𝐺 ′ 𝜔 , where 𝐺 ′ 𝜔 is the storage modulus, and, at high shear rate, Gleissle‟s correlation

[Macosko (1994)],

𝑁1 𝛾 ∗ = 2 ∙ 𝛾 ∗2 ∙ 𝑑𝜂

𝛾

𝛾 ∗

0 , where 𝛾 ∗ = 2.5 ∙ 𝛾 (3-2)

The two sets of data are presented in Figure 3-4(a), for one of the test temperatures. There

is a mismatch of 20%, which is similar to data for other melts [Macosko (1994)].

The exit pressure (∆𝑃𝑒𝑥 ) was then estimated from N1 by inverting the following relation

[Davies et al. (1973)]

𝑁1 = ∆𝑃𝑒𝑥 + 𝜏𝑤 ∙𝜕∆𝑃𝑒𝑥

𝜕𝜏𝜔=

𝜕 𝜏𝑤 ∙∆𝑃𝑒𝑥

𝜕𝜏𝑤 (3-3)

to yield

∆𝑃𝑒𝑥 𝛾 =1

𝜏𝑤 𝛾 ∙ 𝑁1 𝛾 ∙ 𝑑𝜏𝑤 𝛾

𝛾

0 (3-4)

where 𝜏𝑤 = 𝜂 ∙ 𝛾 𝑤 is the wall shear stress at the channel exit. Values of ∆𝑃𝑒𝑥 were determined

this way at several temperatures, and they are presented in Figure 3-4(b) as a function of the wall

shear rate 𝛾 𝑤 .

3.3 Shear Viscosity of Polystyrene Alone

3.3.1 Calculating the Viscosity

The pressure loss in the straight section of the Hele-Shaw channel is determined by the

fluid‟s viscosity, and thus it should be possible to calculate this property from pressure drop data.

In this section, we present the method to calculate the viscosity and then compare computed

values to those measured with commercial rheometers.

Our calculation method depends on channel dimensions and on measurements of pressure

drop and flow rate, and follows Laun (1983)

41

10-2

10-1

100

101

102

103

104

105

106

N1 = 2 * G'()

parallel plate, oscillatory measurement

using known shear viscosity at 172oC

capillary, Glessisle's correlation

First

no

rma

l str

ess d

iffe

ren

ce

N1 (

Pa

)

(Hz) or (s-1)

(a)

100

101

102

103

104

20

40

60

80

100

PS685D

164oC

172oC

190oC

Exit p

ressu

re d

rop

P

ex (

psi)

Wall shear rate at die exit w (s

-1)

(b)

Figure 3-4 (a) Estimates of the first normal stress difference in shearing for the PS at 172oC; (b)

Estimated exit pressure of flows in the Hele-Shaw channels at several temperatures

42

𝜏𝑤 = ∆𝑃

𝐿0∙

𝐻

1+𝐻/𝐵 (3-5)

𝛾 𝑤 = 6∙𝑄

𝐵0∙𝐻2∙

2+𝑏

3 , where 𝑏 =

𝑑 𝑙𝑜𝑔 6𝑄/𝐵0𝐻2

𝑑 𝑙𝑜𝑔 𝜏𝑤 (3-6)

𝜂 = 𝜏𝑤 /𝛾 𝑤 (3-7)

where ∆𝑃 is the pressure difference between the first two transducers in Figure 3-1, 𝜏𝑤 is the

wall shear stress, 𝛾 𝑤 is the corrected wall shear rate, η is the (shear) viscosity, and Q is the

volumetric flow rate. The aspect ratio of the straight channel is accounted for in Equation (3-5)

by the term 1 + 𝐻/𝐵, but was ignored in Equation (3-6) because an explicit expression for it is

not available [Laun (1983)]. The resultant error is negligible because these equations are only

used to calculate shear viscosity from the straight channel, which has a very high aspect ratio.

Equation (3-6) includes the Rabinowitch correction, which takes into account the wall shear rate

difference between a Newtonian fluid and a shear-thinning one [Macosko (1994)].

Viscosity values from the capillary rheometer were calculated from the pressure drop

difference between the long capillary and the zero-length capillary, as mentioned in Section 3.2.2

A. The Rabinowitch correction for tube flow was incorporated in the calculation.

3.3.2 Viscosity Data

Viscosity values from the 0.94-mm channel and those from the rheometers are presented

in Figure 3-5(a). There is a gap in the data at moderate shear rates, but the sets of data are not

inconsistent. The 210oC channel data compare well with data from the capillary rheometer.

The viscosity data were reduced to a master curve of 𝜂/𝛼𝑇 as a function of 𝛾 ∙ 𝛼𝑇 in

Figure 3-5(b), where 𝛼𝑇 is the temperature shift factor with 210oC as the reference temperature.

A shift factor is designated for each temperature such that the shifted curves agree with each

other. The master plot is well described by the Carreau-Yasuda model in Equation (3-8), using

the best-fitting parameters from Table 3-2 [Bird et al. (1987)]

𝜂−𝜂∞

𝜂0−𝜂∞= 1 + 𝜆𝛾 𝑎

𝑛−1

𝑎 (3-8)

43

10-2

10-1

100

101

102

103

104

101

102

103

104

105

106

Shear

vis

cosity (

Pa.s

)

Shear rate (s-1)

Hele-Shaw channel

H = 0.94 mm

164oC

172oC

190oC

210oC

rotational rheometer

164oC

172oC

210oC

capillary rheometer

210oC

(a)

10-2

10-1

100

101

102

103

104

105

101

102

103

104

Hele-Shaw 0.94mm

164oC

172oC

190oC

210oC

227oC

Hele-Shaw 1.96mm

164oC

172oC

190oC

210oC

227oC

Capillary Rheometer

210oC

Rotational Rheometer

210oC

/

T

T

(b)

Figure 3-5 (a) Shear viscosity data of the polystyrene from various rheometers; (b) master

plot of the shear viscosities from data in Figure 3-5(a)

fitted curve

using

Carreau-

Yasuda

model

44

The viscosity data were reduced to a master curve of 𝜂/𝛼𝑇 as a function of 𝛾 ∙ 𝛼𝑇 in

Figure 3-5(b), where 𝛼𝑇 is the temperature shift factor with 210oC as the reference temperature.

A shift factor is designated for each temperature such that the shifted curves agree with each

other. The master plot is well described by the Carreau-Yasuda model in Equation (3-8), using

the best-fitting parameters from Table 3-2 [Bird et al. (1987)]

𝜂−𝜂∞

𝜂0−𝜂∞= 1 + 𝜆𝛾 𝑎

𝑛−1

𝑎 (3-8)

Table 3-2 Best fitting parameters for the master plot of PS685D

𝜂0 (Pa.s) 𝜂∞ (Pa.s) 𝜆 (s) 𝑎 𝑛

2.5x104 0 0.566 0.5 0.158

The viscosities determined from the two Hele-Shaw channel flows are compared in

Figure 3-6. The included error bars overlap, indicating some reliability in the measurements and

in the non-simple calculation technique. The power-law index (n) at high shear rates is 0.16,

identifying a strongly shear-thinning melt. This characteristic means that, in the hyperbolic

channels, the velocity profiles are close to uniform in the core of the flows, creating large regions

of extensional flow with minimal shear, as will be shown in Chapter 4. High-shear-rate values of

the viscosity will be needed to determine the exit pressure, given by Equation (3-4) in Section

3.2.2 B, as well as to calculate the pressure drop due to shearing in the convergent channel, in

Chapter 4. The Carreau-Yasuda equation, with Table 3-2 parameters, will be used for these

calculations, along with the shift factor 𝑎𝑇 which describes temperature dependence of the

viscosity.

45

1 10 100 1000

100

1000

10000

Hele-Shaw

H = 0.94 mm

164oC

190oC

210oC

227oC

H = 1.96 mm

164oC

190oC

210oC

227oC

capillary rheometer

210oC

Vis

co

sity

(P

a.s

)

Shear rate (s-1)

Figure 3-6 Shear viscosity of polystyrene, determined from flow measurements in the Hele-

Shaw channels

46

3.4 Shear Viscosity of Polystyrene/CO2 Solution

3.4.1 Viscosity Data and Prediction by the Free Volume Theory

Having established techniques to determine the shear viscosity of the polystyrene, the

same techniques were applied to our PS/CO2 solution. For this fluid, the pressure in the system

was maintained well above the solubility pressure, so that the CO2 stayed in solution. In this

section and in Chapter 4, we will refer to CO2 as the „solvent‟ in spite of its low concentration,

because the gas has the same effect as a solvent of the polymer. That is, it reduces the viscosity

and elasticity of the melt.

The viscosity data of our PS/CO2 solution, determined at several temperatures and from

both Hele-Shaw channels, are presented in Figure 3-7(a), with PS values added for comparison

purposes. Note that the two materials are compared at the same viscosities, not at the same

temperatures, because the viscosities of the solution are so much lower than those of the

polystyrene alone that they could not be measured accurately at the same temperatures. The

viscosities of the solution at all temperatures were shifted to a master plot, following the same

procedures as described in Section 3.3.2 for PS alone. The results are shown in Figure 3-7(b),

where 172oC is the reference temperature, and which includes the previously-determined shear

viscosity of the PS alone at 172oC.

As to comparable solution viscosity data in the literature, Royer et al. (2000) determined

CO2 solution viscosity for several polystyrenes, including the present Styron 685D, and found

that the temperature dependence at a fixed CO2 concentration is well described by a modified

WLF equation in Equation (3-9)

𝑙𝑛 𝜂0,𝑇 𝐶

𝜂0,𝑇𝑔 𝐶

=−𝑐1∙ 𝑇−𝑇𝑔 𝐶

𝑐2+ 𝑇−𝑇𝑔 𝐶 (3-9)

where C is the CO2 weight concentration, 𝑐1 and 𝑐2 are fitting parameters independent of C, 𝜂0,𝑇𝑔

is the zero-shear-rate viscosity at the glass transition temperature 𝑇𝑔 , and T is the temperature.

Equation (3-9) was originally derived by Williams et al. (1955) from the free volume theory

without considering the effect of solvent concentration. The modified WLF equation suggests

that CO2 affects the solution viscosity mainly by suppressing the glass transition temperature of

the polystyrene.

47

100

101

102

103

104

PS + 5% CO2

H = 0.94 mm

130oC

140oC

150oC

172oC

H = 1.96 mm

130oC

140oC

PS alone

from both channels

172oC

190oC

210oC

Vis

co

sity

(P

a.s

)

Shear rate (s-1)

(a)

10-2

10-1

100

101

102

103

104

102

103

104

105

PS + 5% CO2, T

ref=172

oC

H = 0.94 mm

130oC

140oC

150oC

172oC

H = 1.96 mm

130oC

140oC

PS only at 172oC

o

r *

T

or / T

(b)

Figure 3-7 (a) Shear viscosity of PS/CO2 solution, compared with that of PS only; (b) Master

plot of the PS/CO2 solution for a reference temperature of 172oC, compared to the viscosity of

PS alone at 172oC

48

One way to compare our solution viscosity data to those by Royer et al. (2000) is to apply

Equation (3-9) to our data. For this, the glass transition temperature of PS/CO2 solution needs to

be determined. It was determined by Wissinger and Paulaitis (1987) as a function of CO2 weight

concentration and their results are reproduced in Figure 3-8. The relationship applies to most

commercial polystyrenes regardless of their molecular weight distribution. The ratio 𝑇𝑔/𝑇𝑔0is the

glass transition temperature of the solution over that of the PS alone, and it is 0.87 for a 5% CO2

concentration. The glass transition temperature of a PS + 5% CO2 solution, then, is 323 K, or

50oC, given that the 𝑇𝑔 of the PS alone is 373 K.

Knowing the glass transition temperatures of the two fluids, a BA concentration-

temperature equivalence can be established, as suggested by Equation (3-9). That is, the viscosity

reduction effect of 5% CO2 should be equivalent to a temperature increase of 50oC for the PS

alone, where 50oC is the difference of 𝑇𝑔 between the PS and the PS/CO2 solution. This is indeed

the case in Figure 3-7(a), where the solution viscosity at 140oC almost overlaps the PS viscosity

at 190oC, and the viscosities at other temperatures seem to follow a similar trend. The modified

WLF equation, therefore, accurately predicts the influence of CO2 on the viscosity of our

polystyrene.

3.4.2 Comparing Viscosity Reductions of Various BAs

Before moving on to extensional measurements in Chapter 4, it is worthwhile to compare

the viscosity reduction effect of CO2 with that of other blowing agents. Such information is of

interest to the polymeric foaming industry, because the present PS is a commonly used material,

and thus a comparison may be helpful for predicting the behaviour of a new agent. The

dependence of solution viscosity on BA concentration is derived in the present study from the

modified WLF equation in Equation (3-9):

𝑙𝑛 𝜂0,𝑇 𝐶

𝜂0,𝑇 0 = 𝑙𝑛

𝜂0,𝑇 𝐶

𝜂0,𝑇𝑔 𝐶

− 𝑙𝑛 𝜂0,𝑇𝑔

0

𝜂0,𝑇 0 =

−𝑐1 ∙𝑐2∙ 𝑇𝑔 0 −𝑇𝑔 𝐶

𝑐2+ 𝑇−𝑇𝑔 0 ∙ 𝑐2+ 𝑇−𝑇𝑔 𝐶 (3-10)

In Equation (3-10), if we define BA concentration such that the plots of 𝑇𝑔 versus this

concentration reduce to a single curve for different BAs, then the viscosity reduction of other

49

0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16

0.70

0.75

0.80

0.85

0.90

0.95

1.00

Tg /

Tg0

CO2 concentration by weight (wt%)

Figure 3-8 Reduction of glass transition temperature for PS as a function of CO2

concentration; reproduced from Wissinger and Paulaitis (1987)

50

BAs can be predicted. A proper definition of BA concentration in this sense is the molar

concentration 𝐶𝑚 = 𝐶/𝑀𝑠𝑜𝑙 , where 𝑀𝑠𝑜𝑙 is the molecular weight of the solvent. The molar

concentration naturally arises because the WLF equation is derived from the free volume theory

(e.g., Ferry (1980)), and the latter suggests that viscosity reduction of a polymer solution is

determined by the free volume contributed by the solvent (here free volume is defined as holes of

monomeric dimensions or smaller in the polymer resulting from packing irregularity). At the

same molar concentration then, the free volume contributions of different BAs to the same

polymer may be similar. In fact, Park and Dealy (2006) plotted BA concentration shift factors,

defined as the ratio of solution zero-shear-rate viscosity to that of the polymer alone, against BA

molar concentration in HDPE, and found that the plots for CO2 and N2 almost overlapped.

For our comparison between different BAs in the PS, a viscosity reduction factor is

defined based on 𝐶𝑚

𝑉𝑅𝐹 =𝜂 𝑇,𝐶𝑚 ,𝛾

𝜂 𝑇,𝛾 (3-11)

where 𝜂 𝑇, 𝐶𝑚 , 𝛾 is the solution viscosity at a fixed temperature and shear rate, and 𝜂 𝑇, 𝛾 is

the corresponding viscosity of the polymer only. Han and Ma (1983) used the same polystyrene

as in the present study and determined solution viscosities for the blowing agents CFC-11 and

CFC-12. Measurements showed that the VRF values based on Equation (3-11) from both studies

were nearly constant at high shear rates (i.e., at 𝛾 > 100 𝑠−1), and these high-shear-rate values

are presented in Figure 3-9(a). The plot shows that our VRF values with CO2 (𝑀𝑊 = 44 𝑔/𝑚𝑜𝑙,

𝜌𝑠𝑢𝑝𝑒𝑟𝑐𝑟𝑖𝑡𝑖𝑐𝑎𝑙 ≈ 1 𝑔/𝑐𝑐 ) are close to the values with CFC-11 ( 𝑀𝑊 = 137 𝑔/𝑚𝑜𝑙 ,

𝜌𝑠𝑢𝑝𝑒𝑟𝑐𝑟𝑖𝑡𝑖𝑐𝑎𝑙 ≈ 1.5 𝑔/𝑐𝑐 ) and CFC-12 ( 𝑀𝑊 = 121 𝑔/𝑚𝑜𝑙 , 𝜌𝑠𝑢𝑝𝑒𝑟 𝑐𝑟𝑖𝑡𝑖𝑐𝑎𝑙 ≈ 1.5 𝑔/𝑐𝑐 ).

Differences between the BAs grow at higher molar concentrations, but they are in general much

lower than those between VRFs based on the weight concentration C (Figure 3-9(b)), by a factor

of 𝑀𝐶𝐹𝐶 −𝑀𝐶𝑂 2

𝑀𝐶𝐹𝐶≈ 67%. It is thought, then, that the viscosity reduction plots of other blowing

agents in our PS, as a function of their respective molar concentrations, will follow the trend in

Figure 3-9(a).

51

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8

0.0

0.2

0.4

0.6

0.8

1.0

Vis

co

sity r

ed

uctio

n f

acto

r

solu

tion/

poly

mer

Cm (mol BA/kg PS)

For PS685D

power-law region

with CO2, at 172

oC

with CFC-11, 170oC

with CFC-12, 170oC

(a)

0 5 10 15 20

0.0

0.2

0.4

0.6

0.8

1.0

Vis

co

sity r

ed

uctio

n fa

cto

r

so

lutio

n/

po

lym

er

C (g BA/100g PS)

For PS685D

with CO2, at 190

oC

with CFC-11, 170oC

with CFC-12, 170oC

(b)

Figure 3-9 Viscosity reduction factors of the present PS with various blowing agents: (a) as a

function of molar concentration and (b) as a function of weight concentration. Values for

CFC-11 and CFC-12 are calculated from Han et al. (1983b)

52

Chapter 4 Characterization of the Extensional Properties

This chapter determines the extensional properties of PS melt and PS/CO2 solution from

pressure drop measurement over the thin hyperbolically-convergent rectangular channel

illustrated in Figure 3-1. Because of its shape, the channel induces approximately constant-

extensional-rate flow over a good fraction of its volume, which facilitates the calculation of

viscosity values. It should be mentioned that fluids behave differently under different types of

extensional flow. For example, the uniaxial extensional viscosity for a Newtonian fluid is three

times its shear viscosity, while the planar viscosity is four times, and the biaxial viscosity is six

times the shear viscosity. Whenever extensional viscosity is presented in this study, the type of

flow from which it is determined will be pointed out.

This chapter begins with a review of previous studies on the elasticity of polymer/BA

solutions in Section 4.1. This is followed by a presentation of the uniaxial extensional properties

of the PS alone determined on a shear-free rheometer in Section 4.2. In Sections 4.3, 4.4, and

4.5, a method is presented for extracting the planar extensional viscosity from pressure drop

measurements in the hyperbolic channels after the pressure drop due to shearing is accounted for.

Extensional viscosities from the channel are compared to those from the shear-free rheometer,

and error sources with the channel technique are discussed. Finally, in Section 4.6, planar

extensional viscosities of PS/CO2 solution are determined from the hyperbolic channels, and the

influence of the blowing agent on this property is clarified.

4.1 Introduction

In Chapter 3, it was shown that the shear viscosity of a polymer melt is significantly

reduced by a blowing agent even if the BA concentration is only a few percent by weight. The

dissolved gas reduces the viscosity by increasing the free volume of the melt and therefore chain

mobility [e.g., Utracki and Simha (2001) and Lee et al. (2007)]. This increase of free volume

likely reduces the material‟s elasticity as well because it reduces the relaxation time [Ferry

(1980)], but this has not been confirmed for polymer/BA solutions because measurement

techniques have not been generally available. That is, characterizing the elastic properties of a

53

polymer/gas solution is exceedingly difficult because the tests must be carried at 103 psi

pressures to prevent the gas from coming out of the solution. In contrast to shear properties, there

have been very few studies of the elastic properties of the solutions. One study was by Ouchi et

al. (2008) who measured the storage modulus of a LDPE/CO2 solution using a rotational

rheometer with a high pressure chamber. After taking magnetic signal errors into account, they

found that G’ decreased by about 30% when the LDPE was saturated with CO2 at 10 MPa.

Extensional flow resistance has been measured by Ladin et al. (2001) and Xue & Tzoganakis

(2003). They determined apparent extensional viscosity by measuring the pressure drop in a

sudden planar contraction. The contraction ratios in their flow channels were 10:1 and 18:1,

respectively. Ladin et al. found that the apparent extensional viscosity of a polybutylene

succinate (PBS) solution decreased by 40% to 60% for a CO2 concentration of 6% by weight,

while Xue and Tzoganakis reported a comparable decrease for a 4% CO2 concentration in

polypropylene. In both studies, the decrease of extensional viscosity was similar to the decrease

in shear viscosity at the same CO2 concentration. The aspect ratios of their rectangular channels,

however, were only 1 or 2, and extensional rates varied in unknown ways because of upstream

vortices. Moreover, their calculated values of the extensional viscosity had large errors because

the pressure drops due to extension were a small fraction of the total pressure drop. Both studies

followed Cogswell‟s approach in calculating the extensional viscosity [Cogswell (1972)], i.e.,

they assumed that the extensional viscosity is a function of strain rate only, ignoring the

dependence of this property on strain [Macosko (1994)].

Compared to the techniques of Ladin et al. and Xue & Tzoganakis, our technique is an

improvement because it is based on pressure drop measurements in a thin hyperbolically

convergent channel. With this geometry, extensional rates are nearly constant, at least in the core

of the flow, and so the flow field is more suitable and better defined than that in an abrupt

contraction. Also, our technique considers the dependence of extensional viscosity on strain as

well as on strain rate. Details of the technique will be presented after a presentation of the

uniaxial extensional viscosity of the PS alone, determined from a shear-free rheometer.

54

4.2 Uniaxial Extensional Viscosity from EVF

Measurements of uniaxial extensional viscosity were made using an Extensional

Viscosity Fixture (EVF, TA Instrument Inc.) attached to the ARES rheometer. This device is

similar to the fixture developed by Sentmanat et al. (2005). As shown in Figure 4-1, the EVF

design is based on the original Meissner concept [Meissner and Hostettler (1994)] of elongating

a sample within a confined space by a pair of rotary clamps. Instead of the rotary clamps, the

EVF uses two cylinders to wind up the sample: one cylinder rotates, while the other measures the

force. In order to wind up the sample equally on both sides, the rotating cylinder moves on a

circular orbit around the force measuring cylinder while rotating around its own axis at the same

time. Time-dependent extensional stress and extensional viscosity can be calculated from the

force 𝐹 𝑡 sensed by the static cylinder

𝜏𝑥𝑥 𝑡 − 𝜏𝑦𝑦 𝑡 =𝐹 𝑡

𝐴 𝑡 =

𝐹 𝑡

𝐴0∙ 𝑒𝜀 0𝑡 (4-1)

where 𝐴0 is the initial cross-sectional area of the sample, 𝐴 𝑡 is the instantaneous cross-

sectional area, and 𝜀 0 is the constant extensional rate [Sentmanat et al. (2005)]. The transient

uniaxial extensional viscosity, then, is calculated from the normal stress difference and the

extensional rate

𝜂𝐸+ 𝑡 =

𝜏𝑥𝑥 𝑡 −𝜏𝑦𝑦 𝑡

𝜀 0 (4-2)

To characterize our PS685D on the EVF, rectangular samples of 18 mm × 10 mm × 0.8

mm in dimensions were prepared on a compression molding machine. For these dimensions,

sagging of the sample during measurement is negligible if the zero-shear-rate viscosity of the

material is above 104 Pa.s [Sentmanat et al. (2005)]. The maximum strain rate achievable with

the EVF is 5 s-1

and the maximum Hencky strain achievable is usually 3.5.

Values of uniaxial extensional viscosity were obtained for the PS at temperatures of

172oC, 190

oC and 210

oC, and at extensional rates of 0.1, 0.5, 1.0, and 3.0s

-1. These temperatures

and extensional rates are relevant to flows in the Hele-Shaw channels. The EVF data are

presented in Figure 4-2, showing significant strain hardening because of a high molecular weight

component in the PS [Münstedt (1980)], and necking at the two lowest strain rates.

55

(a)

(b)

Figure 4-1 (a) Schematic of the ARES-EVF (Extensional Viscosity Fixture); (b) representative

positions of the rotating cylinders, and corresponding Hencky strains; reproduced from the

product note on EVF technology by TA Instruments Inc.

56

10-1

100

101

103

104

105

106

0.5 s-1

1.0 s-1

=3.0 s-1

0.1 s-1

0.5 s-1

1.0 s-1

=3.0 s-1

210oC, x0.1

190oC, x0.2

0.1 s-1

1.0 s-1

0.5 s-1

Tra

nsie

nt

unia

xia

l exte

nsio

nal vis

cosity

E

+ (

Pa.s

)

Time (s)

=3.0 s-1

172oC, x1

Figure 4-2 Transient uniaxial extensional viscosity of PS determined with the EVF fixture;

the symbols “x1”, “x0.2”, and “x0.1” indicate that the original data were multiplied by these

factors to avoid overlapping

57

4.3 Calculating the Pressure Drop due to Extension

The channels for determining extensional properties are the hyperbolically convergent

channels in Figure 3-1. Using a thin hyperbolic channel for extensional flows of melts may be

novel but such channels have been utilized in microfluidics, where the cross section can only be

rectangular and the aspect ratio can be large [Whitesides and Strook (2001)]. In a channel

similar to that in Figure 3-1, Oliveira et al. (2007) studied flow of a Newtonian fluid, both

experimentally and numerically, at Reynolds numbers up to 20, in order to evaluate the channel‟s

usefulness as an extensional rheometer. They reported that the velocity gradient in the y

direction was negligible when the aspect ratio (B/H) was high, and that the extensional rate along

the centreline of the channel was approximately constant. However, they questioned whether

extensional stresses could be extracted from the measurement of pressure drop because the shear

stresses are so large. All the same, we anticipated that extensional stresses in our melt would be

large enough to be deduced, enabling a determination of the fluid‟s extensional viscosity.

Detailed procedures are presented below.

When viscoelastic fluid flows through the converging Hele-Shaw channel, part of the

pressure drop between the second and the third transducers, designated ∆𝑃𝑐𝑜𝑛𝑣𝑒𝑟𝑔𝑒𝑛𝑡 in Figure 3-

1, depends on the elastic normal stresses. The precise dependence is found by integrating the

momentum equation in the convergent section.

For steady isothermal flow of an incompressible fluid, the inertialess momentum

equations in the x and z directions (Figure 3-1) are

𝜕𝑃

𝜕𝑥+

𝜕𝜏𝑥𝑥

𝜕𝑥+

𝜕𝜏𝑦𝑥

𝜕𝑦+

𝜕𝜏𝑧𝑥

𝜕𝑧= 0 (4-3)

𝜕𝑃

𝜕𝑧+

𝜕𝜏𝑥𝑧

𝜕𝑥+

𝜕𝜏𝑦𝑧

𝜕𝑦+

𝜕𝜏𝑧𝑧

𝜕𝑧= 0 (4-4)

[Bird et al. (1987)]. The convection terms are omitted because Reynolds numbers in the channel

are of order 10-4

, and the equation in the y direction is not needed because our analysis focuses

on the centre plane, 𝑦 = 0, in the convergent section. When Equation (4-3) is integrated along

the centreline from the upstream location (0, 0, 0) to the downstream location (L2, 0, 0), the result

is

58

𝑃 0,0,0 − 𝑃 𝐿2, 0,0 = 𝜏𝑥𝑥 𝐿2, 0,0 − 𝜏𝑥𝑥 0,0,0 + 𝜕𝜏𝑦𝑥

𝜕𝑦+

𝜕𝜏𝑧𝑥

𝜕𝑧 𝑑𝑥

𝐿2

0 (4-5)

The upstream pressure transducer is located at the wall at 𝑥 = −𝐿1. The pressure which is

sensed there, 𝑃𝑤 −𝐿1 = 𝑃 −𝐿1, 0,𝐻

2 , is related to 𝑃 0,0,0 by first integrating Equation (4-4)

in the z direction at 𝑥 = −𝐿1 from 𝑧 = 𝐻/2 to 𝑧 = 0, and then by integrating Equation (4-3)

along the centreline from 𝑥 = −𝐿1 to 𝑥 = 0. The result is

𝑃 0,0,0 = 𝑃 −𝐿1, 0,0 − 𝜕𝜏𝑧𝑥

𝜕𝑧 𝑑𝑥 =

0

−𝐿1 𝑃𝑤 −𝐿1 + 𝜏𝑧𝑧 −𝐿1, 0,

𝐻

2 −

𝐿1

𝐿0∙ ∆𝑃𝑠𝑕𝑒𝑎𝑟

(4-6)

where the term 𝜏𝑧𝑧 −𝐿1, 0,𝐻

2 is the normal stress at the wall due to shearing, which can be

determined from the N1 measurement in Figure 3-4(a). In deriving Equation (4-6), the terms 𝜕𝜏𝑥𝑥

𝜕𝑥

and 𝜕𝜏𝑦𝑥

𝜕𝑦 in Equation (4-3) and

𝜕𝜏𝑥𝑧

𝜕𝑥 and

𝜕𝜏𝑦𝑧

𝜕𝑦 in Equation (4-4) can be neglected because the x-

derivatives are zero along the straight channel between 𝑥 = −𝐿1 and 𝑥 = 0, and the y-derivatives

vanish along the centreline due to symmetry. The downstream centreline pressure 𝑃 𝐿2, 0,0 in

Equation (4-5) is related to the pressure sensed in the reservoir (𝑃𝑟𝑒𝑠𝑣 = 𝑃𝑤 𝑥 > 𝐿2 ) and to the

exit pressure ∆𝑃𝑒𝑥𝑖𝑡 by

𝑃 𝐿2, 0,0 = 𝑃𝑟𝑒𝑠𝑣 + ∆𝑃𝑒𝑥𝑖𝑡 (4-7)

The exit pressure can be determined from N1, as described in Section 3.2.2 B.

Combining Equations (4-5) to (4-7) and omitting the normal stress term 𝜏𝑧𝑧 in Equation

(4-6) because its value is no more than a few percent of 𝑃𝑤 −𝐿1 , we obtain a relationship

between the measured pressure drop over the convergent channel ∆𝑃𝑐𝑜𝑛𝑣𝑒𝑟𝑔𝑒𝑛𝑡 = 𝑃𝑤 −𝐿1 −

𝑃𝑟𝑒𝑠𝑣 and the pressure drop due to extension ∆𝑃𝑒 = 𝜏𝑥𝑥 𝐿2, 0,0 − 𝜏𝑥𝑥 0,0,0 in the convergent

channel. The latter is the increase in the extensional stress along the channel, to the maximum at

the exit centreline, where 𝑦 = 𝑧 = 0 and 𝑥 = 𝐿2. The relationship between ∆𝑃𝑒 and ∆𝑃𝑐𝑜𝑛𝑣𝑒𝑟𝑔𝑒𝑛𝑡

takes into account the pressure drop due to shear over the length L1 by proportionality:

∆𝑃𝑒 = ∆𝑃𝑐𝑜𝑛𝑣𝑒𝑟𝑔𝑒𝑛𝑡 −𝐿1

𝐿0∙ ∆𝑃𝑠𝑕𝑒𝑎𝑟 −

𝜕𝜏𝑦𝑥

𝜕𝑦+

𝜕𝜏𝑧𝑥

𝜕𝑧 𝑑𝑥

𝐿2

0− ∆𝑃𝑒𝑥𝑖𝑡 (4-8)

59

In order to evaluate the third term on the right side of Equation (4-8), which is the

pressure drop due to shearing in the convergent channel (∆𝑃𝑠), the following assumptions are

made regarding the velocity distribution and constitutive properties: (1) the flow is fully-

developed everywhere in the converging channel because of the low Reynolds numbers; (2) the

velocity distribution depends on shear stresses only, and not on extensional stresses; (3) the shear

viscosity is described by the power law (𝜂 = 𝑚 ∙ 𝛾 𝑛−1), which Figure 3-5 and 3-6 show is

acceptable at the high channel shear rates (except at the highest temperature, 227oC).

The pressure drop due to shearing in the convergent channel is calculated from the

following equation, which takes into account shearing on both the curved and flat walls.

∆𝑃𝑠 = 𝜕𝜏𝑦𝑥

𝜕𝑦+

𝜕𝜏𝑧𝑥

𝜕𝑧 𝑑𝑥

𝐿2

0=

21+𝑛 ∙𝑚

𝐻

𝐿2

0∙

2+1

𝑛 ∙𝑄

𝐵 𝑥 ∙𝐻2

𝑛

∙1

𝑘∙ 𝑑𝑥 =

21+𝑛 ∙𝑚

𝐻∙

2+1

𝑛 ∙𝑄

𝐻2

𝑛

𝐿21

𝐵1−

1

𝐵0

∙ 𝑡𝑛

1−0.536∙𝐻∙𝑡

1/𝐵1

1/𝐵0∙ 𝑑𝑡 (4-9)

Equation (4-9) incorporates the correction factor 𝑘 =𝑑𝑝 /𝑑𝑥𝐻𝑆

𝑑𝑝 /𝑑𝑥 , where the subscript “HS”

denotes Hel-Shaw flow, i.e., it indicates that the shearing caused by the two curved walls

contributes nothing to the pressure gradient. The denominator in 𝑘 is the pressure gradient for

shearing on all four walls, accounting for the finite aspect ratio of the channel. The value of k

approaches 1 for a wide rectangular channel (𝐻/𝐵 ≪ 1). Values of k were found for a power-law

fluid from numerical studies in the literature [Syrjälä (1995) and Kostic (1993)] and from an in-

house finite element program with a Newton-type algorithm for viscosity iteration. The values

are plotted in Figure 4-3, showing how k depends on the power-law index n and on H/B. For our

polystyrene, n is 0.158, and for our hyperbolic channels, 𝐻/𝐵 = 0.03 to 0.68. The plot reveals

that the relevant dependence on n is minimal and that the data are nearly linear over the relevant

H/B range. Thus the straight line 𝑘 = 1 − 0.536 ∙ 𝐻/𝐵, which closely fits the data, is used in the

denominator of the integral of Equation (4-9).

Values of ∆𝑃𝑒 were calculated from Equation (4-8), based on (a) the measurements

of ∆𝑃𝑐𝑜𝑛𝑣𝑒𝑟𝑔𝑒𝑛𝑡 , (b) the measured pressure drop between the first two transducers, (c) Equation

(4-9) for the third-term integral, and (d) Equation (3-4) for the exit pressure. The values of the

two pressure drops obtained for the 0.94 mm channel are presented in Figure 4-4, where the flow

60

0.0 0.2 0.4 0.6 0.8 1.0

0.4

0.5

0.6

0.7

0.8

0.9

1.0

H/B range for

0.94 mm insert

Pre

ssu

re g

rad

ien

t ra

tio

k

H/B

power law index n=1.0

n=0.5

n=0.2

n=0.1

H/B range for 1.96 mm insert

Figure 4-3 Ratio of pressure gradient neglecting the aspect ratio of a rectangular channel over

that considering the aspect ratio, 𝑘 =𝑑𝑝 /𝑑𝑥𝐻𝑆

𝑑𝑝 /𝑑𝑥, with values obtained from the literature and from

running an in-house code.

61

100

101

102

103

To

tal p

ressu

re d

rop

an

d p

ressu

re d

rop

rela

ted

to

exte

nsio

n (

psi)

Flow rate (cc/min)

PS

0.94 mm Hele-Shaw

channel

Pconvergent

164oC

172oC

190oC

Pe

164oC

172oC

190oC

Figure 4-4 The total pressure drop and the pressure drop related to extension in the 0.94 mm

channel (see Equation 4-8 for definitions of ∆𝑃𝑐𝑜𝑛𝑣𝑒𝑟𝑔𝑒 𝑛𝑡 and ∆𝑃𝑒)

62

rate is in units of cc/min, the units of measurement. The figure shows the pressure drop due to

extension is a significant fraction of the total pressure drop, and that ∆𝑃𝑒 increases more rapidly

than ∆𝑃𝑐𝑜𝑛𝑣𝑒𝑟𝑔𝑒𝑛𝑡 with flow rate. This latter behaviour is expected because Equation (4-9) shows

that the pressure drop due to shearing is proportional to 𝑄𝑛 , while the pressure drop due to

extension is proportional to the product of 𝑄 and the extensional viscosity, as will be seen

shortly. The large values of ∆𝑃𝑒 in Figure 4-4, attributed to the extreme shear thinning and the

high elasticity of the fluid, are contrary to the speculation of Oliveira et al. (2007) that a thin

planar hyperbolic channel may not be useful for determining the extensional viscosity.

4.4 The Extensional Rate and Total Strain

In order to relate the EVF data to the extensional measurements in the convergent

channel, it is necessary to know the extensional rate and strain in the channel. If shearing on the

curved walls is neglected, the extensional rate along the centreline is constant and found by

differentiating Equation 3-1, i.e.,

𝜀 𝐻𝑆 = 𝑑𝑣𝐻𝑆

𝑑𝑥 =

2𝑛+1 ∙𝑄

𝑛+1 ∙𝐻∙

1

𝐵1−

1

𝐵0

𝐿2 (4-10)

where the subscript “HS” continues to indicate an infinitely-wide channel or Hele-Shaw type

flow, where 𝑣𝐻𝑆 is the centreline velocity, and n is the power-law index. The Hencky strain 𝜀𝐻 in

this case is 𝜀𝐻,𝐻𝑆 = 𝑙𝑛 𝐵0

𝐵1 .

The actual centreline extensional rate is higher than that in Equation (4-10) because

shearing on the curved walls increases the centreline velocity. The higher extensional rates were

found through the velocity ratio, 𝑣/𝑣𝐻𝑆 , where the numerator accounts for the finite aspect ratio

and the denominator pertains to an infinite aspect ratio. This ratio of velocities is plotted in

Figure 4-5, where values were again found from literature sources and our in-house finite-

element code, like the values in Figure 4-3. Figure 4-5 indicates that, for 𝑛 = 0.16, the parameter

in the Carreau-Yasuda equation for our polystyrene melt, the actual centreline velocities were 12

to 18% higher than those determined from Equation (4-10). Data for 𝑛 = 0.5 and 1.0 are also

included in the plot to illustrate the influence of this parameter.

63

0.0 0.2 0.4 0.6 0.8 1.0

0.8

0.9

1.0

1.1

1.2

1.3

1.4

1.5

Ce

nte

rlin

e v

elo

city r

atio

v /

vH

S

H/B

power-law n=1.0

n=0.5

n=0.2

n=0.1

H/B range for

0.94 mm insert

H/B range for 1.96 mm insert

Figure 4-5 Ratio of centreline velocity for a finite aspect ratio over that for an infinite aspect

ratio, used as a correction factor and determined from literature sources and from an in-house

numerical code

64

The curves in Figure 4-5 which apply to our case – namely, 𝑛 = 0.16 and 𝐻/𝐵 =

0.03 𝑡𝑜 0.68 − were used to calculate the actual centreline extensional rate, 𝜀 =𝑑𝑣

𝑑𝑥. Thus the

actual Hencky strain 𝜀𝐻 is

𝜀𝐻 = 𝜀 𝑡

0∙ 𝑑𝑡 =

𝜀

𝑣

𝐿2

0∙ 𝑑𝑥 = 𝑙𝑛

𝑣 𝑥=𝐿2

𝑣 𝑥=0 (4-11)

Following the adjustment in the centreline velocity, the values of Hencky strain were

12% to 18% above those based on simple Hele-Shaw flow, i.e., above 𝑙𝑛 𝐵0

𝐵1 = 2.3. These

corrections for extensional rate and strain were applied to all flow rates and temperatures.

4.5 Comparing Extensional Viscosities

The calculations of ∆𝑃𝑒 , 𝜀𝐻, and 𝜀 in the two previous sections enable us to compute the

extensional viscosity based on the channel data, and then compare channel results to EVF data.

In principle, a comparison can be made only for flows with similar histories – namely, for flows

having the same constant extensional rates and similar strains − because extensional viscosity

depends on both variables. In the EVF, all of the fluid is subjected to a constant extensional rate;

in the channel, the core of the flow is subjected to an extensional rate which is almost constant.

Some extensional rates are common to both flows, and thus extensional rates can be matched up.

As for the strain, the fixed channel strain lies within the range of EVF measurements. Therefore,

the EVF data at the known channel strain and various extensional rates can be related to the

channel results at the same strain and similar extensional rates.

Relating the results will be in terms of the extensional viscosity, and so the primary

purpose of this section is to show how extensional viscosity is determined from channel

measurements. The two extensional viscosities, though, are different. That is, the EVF provides

measurements of the uniaxial extensional viscosity 𝜂𝐸 , while the channel data provides estimates

of the planar extensional viscosity 𝜂𝑃 [Dealy (1984)]. The two viscosities have different values

at the same extensional rate, but the results should indicate whether channel measurements can

be used to determine the extensional viscosity of a fluid which cannot be tested in an EVF − a

fluid, say, which is evaporative or a fluid requiring a high pressure, such as a foamed plastic.

65

Calculation of the planar extensional viscosity from channel measurements is much less

straightforward than the calculation of the shear viscosity. To start with, the term „planar‟ must

be explained because, for truly planar flow, shearing on the curved walls must be negligible and

H/B must be much greater than unity. In the present case, shearing on the curved walls is small

compared to that on flat walls because W/B is much above unity everywhere. As for H/B, it is of

order 1/10, and not 𝑂 10 . However, because of the extreme shear thinning (𝑛 = 0.16), the

velocity in the z direction is close to uniform over a good fraction of the z-direction depth, and so

the flow in the central region is close to planar. Hence the relevant extensional viscosity is

planar, but the value obtained will be approximate because the flow is not truly planar.

In Section 4.3, it was shown that the pressure drop due to extension ∆𝑃𝑒 in the Hele-Shaw

channel equals the difference in 𝜏𝑥𝑥 between a downstream location and an upstream location

(∆𝑃𝑒 = 𝜏𝑥𝑥 𝐿2, 0,0 − 𝜏𝑥𝑥 0,0,0 ). Extensional viscosity is based on the first normal stress

difference 𝜏𝑥𝑥 − 𝜏𝑦𝑦 ; since only 𝜏𝑥𝑥 is available from ∆𝑃𝑒 , 𝜏𝑦𝑦 is neglected for the moment.

Using the centreline extensional rate at the channel exit, as determined in Section 4.4, the

channel extensional viscosity is calculated from the following Equation (4-12)

𝜂𝑃 =𝜏𝑥𝑥 −𝜏𝑦𝑦

𝜀 ≈

∆𝑃𝑒

𝜀 =

𝜏𝑥𝑥 𝐿2 ,0,0

𝜀 (4-12)

where 𝜏𝑥𝑥 0,0,0 is considered zero because the channel is straight before that location.

Values of the extensional viscosity 𝜂𝑃 𝜀 , 𝜀𝐻 determined this way from Equation (4-12)

are compared with values of 𝜂𝐸 𝜀 , 𝜀𝐻 from EVF measurements in Figure 4-6. Here 𝜀 and 𝜀𝐻 are

the channel centreline extensional rate and strain, as determined previously in Section 4.4. The

data from the two channels should be close, having been determined in the same way. The plot

shows reasonable agreement between the two channels at the two lowest temperatures, when the

pressure differences were highest and therefore the measurements were most accurate. At all

three temperatures, 𝜂𝑃 is higher than 𝜂𝐸 . The only prior comparison of these two viscosities

appears to be in the paper by Laun and Schuch (1989). They determined 𝜂𝑃 properly by

stretching an inflated hollow cylinder of an LDPE melt. For a maximum strain rate of 0.08 s-1

,

they found that values of the normalized planar extensional viscosity (𝜂𝑃 divided by the

Newtonian limit of 4) were lower than values of the normalized uniaxial extensional viscosity

(𝜂𝐸). Because the strain rates in our experiments were much higher, elastic effects were much

66

0.1 1 10

105

106

Po

r

E (

Pa

.s)

Extensional rate (s-1)

Hele-Shaw channel

H = 0.94 mm

164oC

172oC

190oC

H = 1.96 mm

164oC

172oC

190oC

predicted from EVF

164oC

172oC

190oC

Figure 4-6 Comparison of PS planar extensional viscosity 𝜂𝑃 𝜀 , 𝜀𝐻 from two Hele-Shaw

channels with uniaxial extensional viscosity 𝜂𝐸 𝜀 , 𝜀𝐻 from EVF measurement; Here 𝜀 and 𝜀𝐻

are the channel centerline extensional rate and strain, as determined in Section 4.4

67

stronger, and therefore it is expected that the planar extensional viscosity of our polystyrene melt

should be less than the uniaxial extensional viscosity, not more as shown in Figure 4-7.

A relationship between the two properties was worked out by Jones et al. (1987) for a

generalized Newtonian fluid. More specifically, they argued that planar values 𝜂𝑃 𝜀 should

equal uniaxial values calculated as 𝜂𝐸 2∙𝜀

3 ∙

4

3. We found that the Jones et al. relationship brings

our two sets of data closer together, but the applicability of this near-Newtonian model to our

highly elastic polystyrene melt is questionable and thus values are not presented.

In Figure 4-6, the curves are extension thinning. Because the fluid is shear thinning,

Trouton ratios are of interest and thus these ratios are presented in Figure 4-7. The values are

calculated as 𝜂𝑃 𝜀

𝜂 2∙𝛾 or

𝜂𝐸 𝜀

𝜂 3∙𝛾 , based on the 𝜂 𝛾 data in Figure 3-5. The ratios show the expected

behaviour of increasing with extensional rate, and the values are all much higher than 4 or 3, the

planar and axisymmetric values for Newtonian fluids, respectively. In fact, since the ratios are of

order 10 and higher, the graph demonstrates the importance of extensional effects in a

converging flow.

The only prior attempt to compare channel and rheometrical values of extensional

viscosity appears to be that by Gotsis and Oriozola (1998). They defined apparent extensional

viscosity as the measured pressure drop due to extension divided by the average extensional rate

over the entry region. In their conical channel, the extensional rate increased as r3, where r is the

distance from the apex, and r varied by a factor of 6 or 12. Hence the extensional rate was far

from constant and an average value does not seem appropriate. They also argued that these

apparent values from the conical channel should equal rheometrical values of uniaxial

extensional viscosity averaged over the strain. The appropriate value of the extensional viscosity,

though, should be that at the exit of the channel, as calculated here, and not an average value.

Still, they found agreement between channel and rheometrical values to within ±30%. Our

discrepancies are larger although our approach appears more valid.

In calculating the values in Figure 4-6 and 4-7, two assumptions were made: (a) 𝜏𝑦𝑦 was

ignored in the core of the channel; and (b) the velocity distribution in the channel depends on

shearing alone. Because the stress 𝜏𝑦𝑦 cannot be found experimentally, a simple way of

68

0.1 1 10

0

20

40

60

80

100

Tro

uto

n r

atio

Extensional rate (s-1)

Hele-Shaw channel

H = 1.96 mm

164oC

172oC

190oC

from EVF

164oC

172oC

190oC

Figure 4-7 Trouton ratio of the PS calculated from the 1.96 mm Hele-Shaw channel and from

EVF measurement. Comparison of 𝜂𝑃 𝜀

𝜂 2∙𝛾 with

𝜂𝐸 𝜀

𝜂 3∙𝛾

69

estimating it is to apply the upper-convected Maxwell model, which has the fewest material

constants for a highly elastic fluid like our PS melt

𝝉 + 𝜆∇𝝉

= 2𝜂0𝑫 (4-13)

where 𝝉 is the stress tensor, 𝑫 is the deformation tensor, and λ and η0 the relaxation time and the

constant shear viscosity, respectively. Values of 𝜏𝑦𝑦 /𝜏𝑥𝑥 for planar extensional flow were

calculated as

𝜏𝑦𝑦

𝜏𝑥𝑥 = −

𝜏𝑦𝑦

𝜏𝑥𝑥=

1−2𝜆𝜀

1+2𝜆𝜀 ∙

1−𝑒−𝑡 1+2𝜆𝜀 /𝜆

1−𝑒−𝑡 1−2𝜆𝜀 /𝜆 (4-14)

A plot of 𝜏𝑦𝑦 /𝜏𝑥𝑥 versus 𝜆𝜀 produces a single curve independent of λ and η0, as shown in

Figure 4-8. The plot clearly shows that 𝜏𝑦𝑦 becomes negligible when the „strength‟ of the

extensional flow, indicated by 𝜆𝜀 , is of order unity and higher. Hence it is not unreasonable to

neglect this stress in 𝜏𝑥𝑥 − 𝜏𝑦𝑦 for determining the extensional viscosity.

The error associated with assuming wholly extensional flow in the core is more difficult

to evaluate. As noted previously, the viscosities from the two channels in Figure 4-6 are expected

to be closer at the same temperature and extensional rate. The disagreement is thought to result

from the interaction between shearing and extension in the channel flows. In the thickness (z)

direction, the flow changes from pure shearing at the walls to approximately pure extension

along the centreline. In between the flow is mixed, and there is little guidance in the literature

regarding the stress field in such flows. Fuller and Leal (1980 and 1981) measured birefringence

of dilute and moderately-concentrated polymer solutions under various flow conditions in a four-

roll mill. They found that the birefringence, which is proportional to the viscoelastic stress

according to the stress-optical relation [Macosko (1994)], is almost uniquely determined by the

extensional flow component. Assuming this holds for melts as well, the total stress in a mixed

flow is not simply the addition of shear and extensional stresses calculated from the

corresponding viscosities. In the same way, the total pressure drop in a channel is not likely the

sum of the pressure drops due to shearing and extension, which has been assumed in calculating

the channel extensional viscosities.

70

0.01 0.1 1 10 100

0.01

0.1

1

yy/

xx

For H=2.4 and any value

of and 0

Figure 4-8 Plot of 𝜏𝑦𝑦 /𝜏𝑥𝑥 in planar extensional flow, calculated using the upper-convected

Maxwell model

71

Despite the above considerations, our analysis of the planar hyperbolic convergent flow

has been as rigorous as possible. The sources of error are clearly identified in our case. With the

melt alone presented, the road is now clear for determining the extensional viscosity of a foamed

plastic, a material whose extensional properties cannot be found using an EVF.

4.6 Extensional Flow Resistance of the Solution

Since the results in Section 4.5 with PS show that flow in a hyperbolic channel can

provide a useful estimate of extensional viscosity, the technique is now applied to our PS/CO2

solution, to learn the effect of CO2 on this property. The planar extensional viscosity of the PS +

5% CO2 solution was calculated from Hele-Shaw channel measurements at several temperatures

(130oC, 140

oC, and 150

oC), using Equation (4-12), and the values are presented in Figure 4-9.

For comparison, the extensional viscosities of the PS are given, but at three different

temperatures (164oC, 172

oC, and 190

oC). Again, comparisons could not be made at the same

temperatures because either the melt or solution failed to generate reliable extensional stresses.

Figure 4-9 shows that the solution produced the similar extensional viscosity values at a much

lower temperature; hence extensional viscosity values at the same temperature would be much

lower for the solution than for the PS.

The previous studies of the apparent planar extensional viscosity, of a PBS/CO2 solution

by Ladin et al. (2001) and of a PP/CO2 solution by Xue & Tzoganakis (2003), indicate that

viscosity reductions by a 5% CO2 concentration were approximately 30% to 50% for PBS at

140oC, and 50% to 70% for PP at 230

oC. The reductions were similar to those for shear viscosity

under the same conditions. Although the extensional viscosity reduction in our case cannot be

found because the temperature ranges for the melt and solution are different, Figure 4-9 suggests

that the reduction for our PS/5% CO2 solution is higher than 80% because the extensional

viscosity curve for the melt alone at 172oC almost overlaps the solution curve at 130

oC, while the

solution curve at 172oC is some distance below the solution curve at 150

oC. In fact the ratio of

the two viscosities at the same temperature appears to be about 10:1. This reduction in the

extensional viscosity is higher than the 50% to 60% reduction of shear viscosity calculated from

Figure 3-7. The larger difference in the extensional viscosity may be expected, though, because

72

100

101

105

106

PS only

0.94 mm Hele-Shaw

164oC

172oC

190oC

PS+5% CO2

0.94 mm Hele-Shaw

130oC

140oC

150oC

P (

Pa.s

)

Extensional rate (s-1)

Figure 4-9 Planar extensional viscosity of PS and PS/CO2 solution from the 0.94 mm Hele-Shaw

channel

73

the CO2 increases the free volume of PS, an effect equivalent to an increase of temperature

[Utracki and Simha (2001); Ferry (1980)]. Because the melt is more Newtonian like at an

increased temperature, the PS/CO2 solution should also be more Newtonian like than the PS

alone, resulting in the more significant reduction of the extensional viscosity of the solution. This

analogy between the influences of temperature and CO2 concentration on the rheological

properties of the PS is illustrated in Figure 3-7(a) and Figure 4-9 where we notice that, for both

the shear and the extensional viscosities, the PS values at 140oC overlap the solution values at

190oC.

Trouton ratios are compared in Figure 4-10. The extensional viscosities are from Figure

4-9, and the shear viscosities of PS/CO2 at corresponding shear rates were found from the data in

Figure 3-7 using time-temperature superposition. As in the previous figure, the solution produced

the same Trouton ratios at much lower temperatures, suggesting that, at the same temperature,

the solution is a less elastic fluid than the PS alone. All values in Figure 4-10 are higher than 4,

the value for Newtonian fluids, and increase with strain rate for all temperatures. The PS values

in Figure 4-10 are for the 0.94 mm channel and similar PS data were found for the 1.96 mm

channel in Figure 4-7. Both graphs show ratios well above 4 and not tending to that limit at low

extensional rates, in contrast to the EVF data in Figure 4-7. Since flow in the upstream straight

channel contained significant shearing, the abnormally high Trouton ratios may have been

caused by pre-shearing, an effect known to increase extensional stresses in a converging channel

[James et al. (1987); Yao and McKinley (2008)].

74

1 100

20

40

60

80

100

120

140

160 from 0.94 mm

Hele-Shaw channel

PS only

164oC

172oC

190oC

PS + 5% CO2

130oC

140oC

150oC

Tro

uto

n r

atio

Extensional rate (s-1)

Figure 4-10 Trouton ratios 𝜂𝑃 𝜀

𝜂 2∙𝛾 of the PS and the PS/CO2 solution calculated from the 0.94 mm

Hele-Shaw channel data at several temperatures.

75

Chapter 5 Influence of Rheological Properties on the Low-Density

Microcellular Foaming of Polylactic Acid

This chapter investigates the relationship between rheological properties of a polymer,

especially extensional property, and cell morphology from foam processing, characterized by

cell density, open-cell content, cell size uniformity, and expansion ratio. Because most previous

investigations of such relationship focused on polypropylene, a different polymer, polylactic acid

(PLA), is used in this study. This biodegradable polymer has the potential to replace traditional

non-biodegradable polymers in foaming and other applications. Establishing a relationship

between rheological properties and cell morphology is important for PLA, because the foams

may be used in different applications requiring different cell density, cell opening, and expansion

ratios.

Section 5.1 reviews previous efforts to develop foaming technology for PLA,

summarizing challenges and proposing solutions. Sections 5.2 and 5.3 present details of the

foaming experiment and physical properties of the PLAs used, especially their rheological

properties. Section 5.4 discusses the results from foaming experiments, in particular the

influence of material rheological properties and processing conditions on cell morphology. In

Section 5.5, flow-induced crystallization of semi-crystalline PLA is studied. And finally, Section

5.6 investigates experimentally the influence of crystallization on cell morphology from

extrusion foam processing.

5.1 Introduction

Polylactic acid (PLA), or polylactide, is a thermoplastic, aliphatic polyester derived from

renewable resources such as corn starch and sugarcanes, which can be fermented to produce

lactic acid monomer. The monomer is converted to lactide, from which high molecular weight

PLA is synthesized through ring-opening polymerization [Drumright et al. (2000)]. Due to its

biodegradability and biocompatibility, PLA has been used in biomedical applications such as

sutures [Frazza and Schmitt (1971)], tissue engineering scaffolds [Langer and Vacanti (1993);

Mikos et al. (1993)], and drug delivery devices [Johansen et al. (2000)]. During the past decade,

76

this once exclusively biomedical material has found numerous low-cost applications such as food

containers, grocery bags, and controlled release matrices for fertilizers and pesticides, because of

mass production by companies like NatureWorks LLC [Sawyer (2003)]. The foaming industry

has also expressed interest in this material, because PLA foams can potentially replace

polystyrene (PS) and polyurethane (PU) foams, which are widely used, but are either non-

biodegradable (PS) or non-recyclable (PU) [Lee et al. (2007)]. In particular, PLA foaming with

supercritical-CO2 is considered a 100% “green” technology because, unlike organic solvents,

CO2 can be removed completely after foaming [Mooney et al. (1996)].

Early efforts to produce PLA foams involved batch foaming in a chamber using various

blowing agents and solvents [Mooney et al. (1996); Maquet et al. (2000); Di et al. (2005); Ema

et al. (2006); Wang et al. (2007)]. This process is of little commercial value, however, because it

can only produce high-density foams, and productivity is extremely low. Recently, efforts have

been made to produce PLA foams from extrusion, especially low-density foams using CO2 as the

blowing agent, because the foams may be used as packaging and insulation materials. Lee et al.

(2008) and Reignier et al. (2007) studied comprehensively the CO2-foaming behaviour of

commercially available linear PLAs during extrusion. They were both able to achieve high

expansion ratios (over 20 times) using a capillary die and a high CO2 content (e.g., 9% by

weight), but the foams showed high open-cell content, poor mechanical properties, and shrank to

60% to 80% of their initial volumes after left in the atmosphere for 48 hours. Pilla et al. (2008)

studied the influence of epoxy-functionalized linear chain extender (CE), which connects chain

ends and hence increases viscosity, on the foaming behaviour of a linear PLA. They found that

expansion ratio increased with CE content, but the maximum expansion ratio was only 4 times.

The open-cell content, on the other hand, appeared unrelated to the CE content. The mechanical

and insulation properties of the low-density PLA foams produced in these studies are in general

inferior to foams of polystyrene and polyolefins, because of inferior cell morphologies.

In summary, development of low-density CO2-foaming process for PLA has suffered

several challenges. First, the low melt strength of commercially available PLAs, resulting from a

linear molecule and a relatively low molecular weight, gives rise to significant cell coalescence

and cell wall opening. Secondly, the faster permeation of CO2 through PLA, compared to

conventional polymers [Bao et al. (2006)], and the high open-cell content of the low-density

foams, cause the foams to shrink after processing [Reiginer et al. (2007)]. Thirdly, only limited

77

data are available of the solubility and diffusivity of CO2 in PLA and the rheological properties

of PLA/CO2 solution, making it difficult to optimize foaming process for better control of cell

nucleation and growth. And finally, the temperature window for producing low-density PLA

foams is very narrow, requiring high tolerance for process control.

The above challenges may be solved by optimizing the molecular structure of PLA and

thereby its physical properties, especially its rheological properties. As already discussed in

Section 2.3 using the cell model, when two neighbouring cells grow during foaming, the melt

between them is subjected to approximate biaxial stretching, and the cell wall will rupture if the

melt has reached its strain to break. If rupture occurs during the early stage of foaming, i.e., when

the melt is barely cooled, adjacent cells will merge into one cell, and the cell number density will

be reduced in the final foams. If rupture occurs when the melt is partially cooled, the ruptured

cell wall will maintain and cell opening is resulted. Increasing the melt viscosity and the strain to

break will prevent the cells from growing too fast and increase the maximum strain the melt can

endure. It should be mentioned that high cell number density may increase the chance of cell

opening, because the average cell wall thickness will be reduced compared to that of lower cell

number density.

The melt strength and strain to break of linear PLA may be increased by introducing long

chain branches into the molecules and also by controlling processing conditions. Increasing the

melt strength also suppresses permeation of gas through the polymer, and widens the temperature

window for producing low-density foams. The present study therefore investigates the low-

density, CO2 extrusion foaming behaviour of both linear and branched PLAs with different

molecular weight, molecular weight distribution, and, in the case of branched PLAs, branching

topology. The production of low-density foams involves higher extensional rates and higher

Hencky strains compared to the production of high-density foams. As a result, cell morphologies

are expected to be more sensitive to the rheological properties of the melt. Details of the study

are presented below.

78

5.2 Experimental

The same tandem system described in Section 3.1.2 was used. As shown in Figure 5-1, a

capillary die replaced the thin rectangular die in Figure 3-2 at the exit of the heat exchanger. Two

capillary dies were used, the flow channel of each begins with a reservoir of 10 mm in diameter

and 50 mm in length, followed by a small capillary of 1 mm in diameter and either 6 mm or 10

mm in length. The small capillaries provide enough resistances for the reservoir pressure to stay

above the solubility pressure during processing. The reservoir and the capillary are connected by

a cone with a total angle of 120o, which suppresses the formation of vortices at the entry. The

CO2 flow rate was controlled by the syringe gas pump in Figure 5-1, and the first extruder, the

gear pump, and the second extruder were all synchronized such that pressure was above the

solubility pressure everywhere in the system. Mass flow rate was determined by collecting a

sample at the die exit and weighing it, and foamed samples were collected by hand. For all

experiments, flow data and samples were not collected until a steady state had been reached,

typically within 20 minutes.

5.3 Properties of the Polymers

5.3.1 General Physical Properties

Four grades of polylactic acids were supplied by NatureWorks LLC. They will be

referred to as the linear PLA, the half-long chain branched (LCB) PLA, the LCB PLA, and the

LCB PLA with lubricant. According to the manufacturer, the linear PLA is a linear and low

crystallinity polymer (Ingeo 2002D), while the other three grades are branched polymers with

higher crystallinity prepared from melt blending of a linear polymer (Ingeo 8051D) with an

epoxy-based multi-functional oligomeric chain extender (Joncryl ADR-4368C, BASF Inc.). The

structures of the linear PLA and the chain extender are shown in Figure 5-2 (a) and (b),

respectively. According to literature [Villalobos et al. (2006); Bikiaris and Karayannidis (1995)],

the main reaction between the two is esterification of carboxyl end groups on PLA with epoxy

group on the chain extender, resulting in star-branched polymer. The functionality of the chain

extender, the number of epoxy groups per molecule, has a broad distribution, which maximizes

the elasticity of the branched polymer, and the average epoxy content is 285 g/mol. The weight

79

Figure 5-1 Schematic of the tandem extrusion system for foam extrusion using a capillary die;

the system setup is similar to that in Naguib et al. (2002)

reservoir

80

(a)

(b)

Figure 5-2 (a) Schematic of high molecular weight PLA molecule; (b) General structure of the

styrene-acrylic multi-functional oligomeric chain extenders; where R1 – R5 are H, CH3, a higher

alkyl group, or combinations of them; R6 is an alkyl group, and x, y, and z are each between 1

and 20; reproduced from Villalobos et al. (2006)

81

percentage of chain extender is 0.7% for preparing the two LCB PLAs and 0.35% for preparing

the half-LCB PLA. Assuming the chain extender has fully reacted with the carboxyl end groups,

the two LCB PLAs should have approximately two times the branching points as the half-LCB

PLA. Lubricant was added to one of the LCB PLAs to improve its processability and to delay the

onset of melt fracture during extrusion [Kulikov et al. (2007)]. For all three branched PLAs,

0.4%wt of talc was added as nucleating agent. It is well known that PLA has two stereoisomers,

PLLA and PDLA, and copolymer or blend of the two at different molar fractions show different

phase behaviour, including melting point and crystallization kinetics [Bastioli (2005)]. All four

PLAs have a PDLA molar content of 4.2%, but the crystallization kinetics of the linear PLA is

much slower than the branched ones, as will be presented below. This is mainly because of the

lack of talc, which facilitates heterogeneous nucleation, in the linear PLA.

Molecular weight distribution of all four PLAs are determined from size-exclusion

chromatography (SEC) by the manufacturer and shown in Table 5-1. The table indicates that

LCB PLAs have slightly more high-molecular-weight component than the half-LCB PLA, and

all three branched PLAs have significantly more high-molecular-weight component and wider

molecular weight distribution than the linear PLA. Also shown in Table 5-1 are the glass

transition temperature 𝑇𝑔 , the melting temperature 𝑇𝑚 , the isothermal crystallinity 𝜒𝑖𝑠𝑜 , and the

crystallinity from a cooling rate of -1 oC/min, 𝜒−1 , for the four PLAs, all determined on a

differential scanning calorimeter (DSC Q2000, TA Instruments). The melting temperature 𝑇𝑚

corresponds to the tip of the melting peak, while the isothermal and the -1 oC/min crystallinity

are determined by first increasing the sample temperature to well above 𝑇𝑚 , then either cooling it

to 100oC and letting it fully crystallize or cooling it to below 𝑇𝑔 at a rate of -1

oC/min, and finally

re-heating the sample to determine crystallinity. The data in Table 5-1 confirms that the linear

PLA has significantly lower crystallinity and slower crystallization kinetics compared to the

branched PLAs. The isothermal crystallinity is the maximum crystallinity attainable under static

conditions.

82

Table 5-1 Properties of the four PLAs

𝑀𝑛

(g/mol)

𝑀𝑊/𝑀𝑛 𝑇𝑔

(OC)

𝑇𝑚

(OC)

𝜒𝑖𝑠𝑜

(%)

𝜒−1

(%)

Linear PLA 132,000 1.4 57 N/A 7 0.04

Half-LCB PLA 215,000 2.5 57 151 29.1 30

LCB PLA 232,000 2.7 57 148 27.4 23.9

LCB PLA + lub. 232,000 2.7 57 148 27.2 27.7

For all characterization and processing experiments, the PLAs were dried at 60oC for 8

hours prior to use. This drying stage is necessary to prevent hydrolysis (degradation) in the

molten state. Solubility of CO2 in PLA, required for pressure control during extrusion, was

determined by Li et al. (2006). The solubility pressure for 9% CO2 in PLA at 180oC, for

example, is 19 MPa, or 2800 psi. The solid density of all three PLAs is 1.25 𝑔/𝑐𝑚3.

5.3.2 Rheological Properties

A. Shear Properties

Oscillatory shear data were determined on an ARES rheometer (TA Instruments Inc.)

using 25-mm parallel disks setup and a 0.8-mm gap. Complex viscosities at 180oC are presented

in Figure 5-3, where the viscosities of the branched PLAs are higher than that of the linear PLA

because of their higher molecular weights. The branching structure influences the complex

viscosity. In fact, in Figure 5-3 the viscosity of the half-LCB PLA is more similar to the linear

PLA than it is to the LCB PLAs, although all three branched PLAs have similar molecular

weight distributions.

The linear viscoelastic properties, 𝐺 ′ and 𝐺" , were determined at 180𝑜𝐶 , an arbitrary

processing temperature, for all three PLAs. The relaxation time determined from the crossover

point is 0.02 s for the linear PLA, 0.04 s for the half-LCB PLA, and 0.2 s for the two LCB PLAs.

83

0.1 1 10 100

103

104

Co

mp

lex v

isco

sity

* (P

a.s

)

Frequency (Hz)

180oC

LCB PLA

LCB PLA + Lub.

half-LCB PLA

linear PLA

Figure 5-3 Complex viscosities of the four PLAs at 180oC

84

This indicates that the linear viscoelastic properties of the half-LCB PLA are also more similar to

the linear PLA than they are to the LCB PLAs.

B. Extensional Properties

Measurements of the uniaxial extensional viscosities were made using the Extensional

Viscosity Fixture attached to the ARES rheometer, as described in Section 4.2. To prevent

sagging of the sample because of too low zero-shear viscosity, the linear PLA was characterized

at 140oC. The three branched PLAs, on the other hand, were characterized at 160

oC to avoid

crystallization during measurement. Values of the uniaxial extensional viscosities are presented

in Figure 5-4 at several extensional rates. The linear PLA exhibits little strain hardening, and its

extensional viscosity approaches a steady-state value for the two higher extensional rates.

Necking of the sample, indicated by a decaying viscosity, is found at the two lower extensional

rates. The Hencky strain where necking occurs, the strain to break, is 0.2 for the lowest

extensional rate, and it increases with the extensional rate. This increase is attributed to less

relaxation of the chains during extension, and, according to molecular theory [McKinley and

Hassager (1999)], the strain to break approaches a maximum in the limit of rapid extension when

the chains do not relax at all. The increase of strain to break with extensional rate is beneficial to

foaming because cell growth normally induces high extensional rates in the melt [Guo et al.

(2006)].

The uniaxial extensional viscosities of the three branched PLAs exhibit significant strain

hardening at all extensional rates, and contrary to the shear data in Section 5.3.2 A, the

extensional viscosities of the half-LCB PLA are more similar to those of the LCB PLAs than

they are to the linear PLA, indicating, possibly, that the side chains affect the extensional

properties more significantly than the shear properties [Auhl et al. (2004)]. Because of the

Hencky strain limit with the EVF, strain to break cannot be determined for the branched PLAs.

Nevertheless, experiments by other authors (e.g. Auhl et al. (2004)) and predictions from the

molecular theory suggest that it increases with the length and density of side chains and also with

a widening of molecular weight distribution of the polymer. A higher strain to break for the

branched PLAs will be beneficial to the processing of low-density foams.

85

(a) (b)

0.01 0.1 110

3

104

105

106

LCB PLA

160oC

0.1s-1 0.5s

-1

1.0s-1 3.0s

-1

E

+(P

a.s

)

H

0.01 0.1 110

3

104

105

106

LCB PLA + lub.

160oC

0.1s-1 0.5s

-1

1.0s-1 3.0s

-1

E

+(P

a.s

)

H

(c) (d)

Figure 5-4 Transient uniaxial extensional viscosities of: (a) the linear PLA at 140oC; (b) the

half-LCB PLA at 160oC; (c) the LCB PLA at 160

oC; (d) the LCB PLA with lubricant at 160

oC

0.01 0.1 110

3

104

105

106

linear PLA

140oC

0.1s-1 0.5s

-1

1.0s-1 3.0s

-1

E

+ (

Pa

.s)

H

0.01 0.1 110

3

104

105

106

E

+ (

Pa

.s)

H

half-LCB PLA

160oC

0.1s-1 0.5s

-1

1.0s-1 3.0s

-1

86

In order to investigate the influence of branching topology on foaming, two blends of the

LCB PLA (without lubricant) and the linear PLA, at weight ratios of 10% - 90% and 20% - 80%,

were prepared using a twin-screw compounder. The uniaxial extensional viscosities of these

blends were determined at 140oC, and they are compared to the extensional viscosities of the

linear PLA in Figure 5-5. Clearly, as the amount of long-chain-branched component increases,

the blends become more strain hardening and more viscous. Foaming experiments using these

blends will be presented shortly.

5.4 Results and Discussions

5.4.1 Processing Strategies

For each PLA or blend, foaming experiments were conducted in the following way:

beginning with the highest processing temperature (usually 170oC in the die), 5% CO2 by weight

was injected and a flow rate was chosen such that, at steady state, the capillary die pressure was

above the solubility pressure. After foams were collected at the die exit, the die temperature and

the heat exchanger temperature were lowered together, and foamed sample was not collected at

the new temperature until the system has reached a steady state for at least 5 minutes. The

temperature was then further lowered until the die pressure was too high (e.g., above 4500 psi),

when the CO2 content was increased to 7% by weight. As a result, experiment could be

continued at lower temperatures because of increased plasticizing effect of CO2, and when the

die pressure was too high again, the CO2 content was increased to 9% by weight, and the

temperature was further decreased until the experiment had to stop.

The capillary die pressures, measured at the exit of the heat exchanger, are shown for the

LCB PLA in Figure 5-6. Die pressures for the other PLAs and blends follow a similar trend. The

die pressures suggest that the plasticizing effect of each additional 2% CO2 by weight is

approximately equivalent to a temperature increase of 10oC. The flow rate ranged between 11

and 14 g/min, depending on the polymer and the processing conditions. Except for the two LCB

PLAs, which were extruded from the 6-mm die, all the other PLAs and blends were extruded

from the 10-mm die.

87

(a)

0.01 0.1 110

3

104

105

106

E

+ (

Pa

.s)

H

140oC

0.1s-1

0.5s-1

1.0s-1

3.0s-1

0.01 0.1 110

3

104

105

106

E

+ (

Pa

.s)

H

140oC

0.1s-1

0.5s-1

1.0s-1

3.0s-1

(b) (c)

Figure 5-5 Transient uniaxial extensional viscosities of: (a) the linear PLA at 140oC; (b) blend of

10% LCB PLA and 90% linear PLA at 160oC; (c) blend of 20% LCB PLA and 80% linear PLA

at 160oC

0.01 0.1 110

3

104

105

106

linear PLA

140oC

0.1s-1 0.5s

-1

1.0s-1 3.0s

-1

E

+ (

Pa

.s)

H

88

120 130 140 150 160 170

2000

2500

3000

3500

4000

4500

5000

5500

Die

Pre

ssure

(psi)

Die temperature (oC)

LCB PLA

5% CO2

7% CO2

9% CO2

Figure 5-6 Exit die pressure as a function of CO2 concentration by weight and die temperature

for the long-chain-branched (LCB) PLA

89

5.4.2 Cell Densities

The foamed samples were dipped in liquid nitrogen and snapped to reveal the cells on the

cross section. The cells were then observed on a scanning electron microscope (SEM), and the

cell density 𝑁𝑐𝑒𝑙𝑙 was calculated using the following equation

𝑁𝑐𝑒𝑙𝑙 = 𝑛

𝐴

3/2

∙𝜌𝑃

𝜌𝑓 (5-1)

where n is the number of cells on an SEM image, A is the area of the image, and 𝜌𝑃 and 𝜌𝑓 are

the densities of the unfoamed and foamed PLA, respectively. The cell densities are shown for the

linear, the half-LCB, and the LCB (without lubricant) PLAs in Figure 5-7. At the same

temperature and CO2 content, cell densities for the LCB PLA appeared higher than those for the

half-LCB PLA, which are in turn higher than those for the linear PLA. The higher cell densities

for the branched PLAs may result from the presence of nucleating agent, less cell coalescence as

the melt strength increases with the number of branching points, or even the nucleating effect of

crystal precursors in the branched PLAs extruded at lower temperatures.

For any given PLA, the cell density increases with the CO2 content, but is not very

sensitive to the change of temperature. According to classical nucleation theory (see Equations

(2-5) to (2-8) in Chapter 2), the nucleation rates for both homogeneous and heterogeneous

nucleation are determined by the oversaturation of BA (∆𝑃 in Equation (2-5)), which, in this

case, is a function of pressure drop rate in the die. The pressure drop rate is calculated as the die

pressure divided by the average residence time of melt in the capillary. It varies mildly

between 500 𝑀𝑃𝑎/𝑠 and 1 𝐺𝑃𝑎/𝑠 as a function of temperature, and nucleation rates are

therefore expected to be similar over this range [Xu et al. (2003)]. In contrast to temperature,

higher CO2 content increases the chance of forming nuclei larger than the critical radius, giving

rise to higher nucleation rates, and therefore higher cell density.

5.4.3 Expansion Ratios

Expansion ratio is defined as the ratio of foam density to the density of unfoamed

polymer. In general, the expansion ratios of a thermoplastic polymer at higher processing

90

110 120 130 140 150 160 170 18010

5

106

107

108

109

Ce

ll d

en

sity (

ce

lls/c

m3)

Die temperature (oC)

LCB PLA

9% CO2

7% CO2

5% CO2

Half-LCB PLA

9% CO2

7% CO2

5% CO2

Linear PLA

7% CO2

5% CO2

Figure 5-7 Cell densities of the linear, the half-LCB, and the LCB (without lubricant) PLAs

from foam extrusion as a function of processing temperature

91

temperatures are low because low melt strength gives rise to significant cell coalescence,

creating channels for the BA to escape to the environment [Naguib et al. (2002)]. The escaping

process is accelerated by high diffusivity of the BA at these temperatures. The expansion ratio

increases as the temperature is lowered, and it reaches a maximum when the melt strength and

the BA diffusivity are well balanced. Below the optimal temperature, expansion ratio decreases

again because the melt is too stiff to be stretched. The expansion ratio usually increases with BA

concentration, although a too high concentration causes undissolved gas pockets in the foams. In

this section, we present expansion ratios of the PLA foams obtained from bulk density

measurement. The cell morphology of these foams will be presented in the next section.

The expansion ratios of all four grades of PLAs are shown in Figure 5-8(a), where they

increase with CO2 concentration and decrease with processing temperature without exception.

For the linear and the half-branched PLAs, expansion ratios are very low (< 2 times) except with

9% CO2 and at the lowest temperature. The dotted line for the linear PLA indicates a range over

which the foams are reticulated due to cell rupture, as will be shown in Section 5.4.4. In contrast,

expansion ratios of the two LCB PLAs are much higher, and ultra low-density foams with

expansion ratios over 40 were produced at temperatures below 117oC using 9% CO2. For very

high expansion ratio foams, it is of interest to calculate the theoretical expansion ratio, and

thereby to learn the foaming efficiency of the CO2. This theoretical ratio assumes no escape of

the BA to the environment, and is calculated as [Naguib et al. (2002)]

𝑉𝑡 = 1 + 𝑤𝑡%𝐵𝐴 ∙𝜈𝐵𝐴

𝜈𝑃= 1 + 𝑤𝑡%𝐵𝐴 ∙

𝑅𝑔 ∙𝑇/ 𝑃∙𝑀𝑊 ,𝐵𝐴

𝜈𝑃 (5-2)

where 𝑤𝑡%𝐵𝐴 is the weight percentage of the BA, 𝜈𝑃 is the specific volume of the polymer, 𝜈𝐵𝐴

is the specific volume of the BA calculated from the ideal gas law, and 𝑀𝑤 ,𝐵𝐴 is the molar mass

of the BA. From this equation, the maximum expansion ratio is 65 times for 9% CO2 at 115oC.

Considering shrinkage of the BA due to cooling while the cells were still growing, the foams

produced at 115oC have almost reached the full potential of the blowing agent!

Expansion ratios of the PLA blends, for 9% CO2 only, are compared to those of the base

PLAs in Figure 5-8(b). The expansion ratios increase with the amount of LCB PLA added.

Interestingly, expansion ratios of the blend with 10% fully-branched PLA already show

significant improvement over those of the linear PLA. This is consistent with findings in the

92

110 120 130 140 150 160 1700

5

10

15

20

25

30

35

40

45LCB PLA

5% CO2

7% CO2

9% CO2

LCB PLA + lubricant

9% CO2

half-LCB PLA

9% CO2

Linear PLA

9% CO2

Expansio

n r

atio

Die temperature (oC)

(a)

115 120 125 130 135 140

0

5

10

15

20

25

30

35

40

45

50

Expansio

n R

atio

Temperature (oC)

9% CO2

LCB PLA

20% LCB PLA

+ 80% linear PLA

10% LCB PLA

+ 90% linear PLA

Linear PLA

(b)

Figure 5-8 Expansion ratios of: (a) all four grades of PLAs; error bars are omitted for clarity; (b)

the linear PLA, the LCB PLA, and blends of the two

93

literature [Spitael & Macosko (2004); Stange & Mϋnstedt (2006)] that expansion ratios of linear

PP may be improved by adding a small amount of long-chain-branched PP.

5.4.4 Cell Morphologies from SEM

Figure 5-9 presents cell morphologies of the PLAs and the blends, taken from the centre

of the extruded filament and observed on a scanning electron microscope (SEM). The

temperatures correspond to the highest expansion ratios at the given CO2 concentration. For the

linear and the half-LCB PLAs, the foams with 5% CO2 show isolated cells (Figure 5-9 (a) and

(d)) resulting from cell coalescence at high temperature. Cells for the LCB PLA with 5% CO2

are better grown (Figure 5-9 (g)), but the cell walls appear thick because of a low expansion

ratio. The foams are better expanded at 7% CO2 (Figure 5-9 (b), (e), and (h)), and then at 9%

CO2, foams with the linear PLA show complete open-cell structure even at the lowest

temperature (Figure 5-9 (c)). Foams with the half-LCB PLA also show open-cell structure

(Figure 5-9 (i)), although the cell shapes are better maintained compared to linear PLA foamed at

the same temperature, and foams with the two LCB PLAs at 9% CO2 show closed cells (Figure

5-9 (i) and (j)). Foams with the 10% blend also show open-cell structure (Figure 5-9 (k)) similar

to the half-LCB PLA, but foams with the 20% blend (Figure 5-9 (l)) show closed cells similar to

the LCB PLAs, possibly because of the lower expansion ratio of the 20% blend in Figure 5-9 (l)

and the lower processing temperature compared to the half-LCB PLA in Figure 5-9 (f).

Figure 5-10 presents SEM images of the extruded filaments using 9% CO2 and at the

lowest temperature, for the LCB PLA (Figure 5-10 (a)) and the half-LCB PLA (Figure 5-10 (b)).

The cells on the outer edge show lower cell density and higher closed-cell content compared to

those in the core. The result occurs because melt on the outer circle suffers from shearing in the

die and a consequent loss of BA to the environment.

From the results in Figure 5-8 and 5-9, it can be concluded that cell morphology and the

expansion ratios of foams are closely related. If an open-cell structure is induced, BA will

quickly diffuse out from the melt, and the foams will not expand. If a closed-cell structure is

induced, gas will diffuse into individual cells and blow them up. Loss of BA to the environment

is reduced compared to open-cell structure because of resistance from the cell walls. For the

94

(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

(j) (k) (l)

Figure 5-9 SEM images of the cellular structures; the temperatures correspond to the highest

expansion ratios at the given CO2 concentration: (a) linear PLA, 5% CO2, 140oC; (b) linear PLA,

7% CO2, 130oC; (c) linear PLA, 9% CO2, 110

oC; (d) Half-LCB PLA, 5% CO2, 140

oC; (e) Half-

LCB PLA, 7% CO2, 130oC; (f) Half-LCB PLA, 9% CO2, 116

oC; (g) LCB PLA, 5% CO2, 140

oC;

(h) LCB PLA, 7% CO2, 130oC; (i) LCB PLA, 9% CO2, 115

oC; (j) LCB PLA + lubricant, 9%

CO2, 110oC; (k) 10% LCB + 90% linear PLA, 9% CO2, 117

oC; (l) 20% LCB PLA + 80% linear

PLA, 9% CO2, 114oC

95

(a)

(b)

Figure 5-10 Cross sections of the extruded filament: (a) LCB PLA with 9% CO2 at 115oC; (b)

half-LCB PLA with 9% CO2 at 116oC; Notice the open-cell structure in the core of the half-LCB

PLA filament

96

linear and the half-LCB PLAs, open-cell structures are induced in the core of the filament, and

volume expansion is not caused by expansion of the cells in the core, but by expansion of the

outer part of the filament. The foams in this case have poor mechanical properties because of

reduced connectivity among cells in the core. For the closed-cell foams produced with the LCB

PLAs, then, foam expansion is driven by the growth of individual cells, and the expansion ratio

is determined by the gas pressure inside the cells and the melt strength.

Cell morphology is also closely related to the rheological properties of the polymers. As

discussed in Section 2.3 and 5.1, low-density foaming involves significant extension of the melt.

If the melt has adequate viscosity and strain to break, cells will grow at a reasonable rate, and

significant strain can be induced without rupturing the cell wall, resulting in closed-cell low-

density foams. If the melt viscosity is low, cell wall may rupture due to overstretching, and open

cell is resulted. Molecular branching increase the extensional viscosity and the strain to break,

and this is why open-cell content is significantly reduced as the number of branching points is

increased in the melt (e.g., for the LCB PLAs).

5.5 Influence of Processing Conditions on PLA Crystallization

In Section 5.5, we see that material rheological properties, in particular the melt viscosity

and the strain to break, have a strong influence on the foaming behaviours of PLA. Since the

PLAs used in this study, especially the LCB PLAs, are semi-crystalline, crystallization may

affect cell growth. The crystallinity of foams with the LCB PLA and the half-LCB PLA was

determined on DSC, and the results are presented in Figure 5-11. Crystallinity of the foamed

polymer generally increases with CO2 content and decreases with temperature. Interestingly, the

increase of crystallinity matches the increase of expansion ratios presented in Figure 5-8(a). In

the literature, it is well-known that PLA has very slow crystallization kinetics compared to

polyolefins [Takada et al. (2004)], and that extruded PLA usually show little crystallinity. The

crystallinity developed in PLA foams must have been induced by the unique processing

conditions associated with foam processing. This section therefore investigates the influence of

processing conditions on the crystallization of PLA.

Crystallization of PLA occurs when it is cooled from a molten state to below its melting

97

110 120 130 140 150 160 1700

5

10

15

20

Cry

sta

llin

ity

(%

)

Die temperature (oC)

LCB PLA

5% CO2

7% CO2

9% CO2

Half-LCB PLA

9% CO2

Figure 5-11 Crystallinity of the foams as a function of CO2 content and temperature, for the

LCB PLA (without lubricant) and the half-LCB PLA

98

temperature. During foam extrusion, PLA is cooled to below its melting temperature in the heat

exchanger (see Figure 5-1), and it stays in the die isothermally, for about 30 seconds, before it is

exposed to the atmosphere and cooled to below its glass transition temperature, when

crystallization stops. Possible mechanisms that accelerate PLA crystallization during foam

processing include the presence of CO2, shearing in the flow channels, and extensional flows

induced by cell growth. According to literature (e.g., Yu et al. (2008)), the presence of CO2

lowers the crystallization temperature of PLA, but whether it accelerates crystallization depends

on the temperature. Shearing and extension are well-known to increase the kinetics of

crystallization, sometimes by several orders of magnitude, because the polymer chains are

oriented by the flow [Schultz (2001)]. To investigate this effect, a series of experiments were

conducted by shearing the LCB PLA prior to crystallization: A ring-shaped sample was placed

between a cone and a plate on the ARES rheometer. It was heated to above 𝑇𝑚 and given 10

minutes to melt. Shearing was then imposed at a shear rate of 50 s-1

for 20 seconds, following

which the sample was cooled to a temperature below 𝑇𝑚 to crystallize isothermally. The storage

modulus was measured using very low strain, and the modulus was found to increase as the melt

crystallizes. The time needed to attain 50% of the steady state crystallinity, the crystallization

half time 𝑡1/2 , is shown in Figure 5-12, in comparison with 𝑡1/2 of PLAs not subjected to

shearing. Clearly, shearing accelerated crystallization significantly, and the associated 𝑡1/2 is

comparable to the residence time in the die after cooling by the heat exchanger. The mechanism

of crystallization is therefore clarified.

From the same experiment, 𝐺 ′ was determined as a function of crystallinity and the

results are plotted in Figure 5-13. The modulus increased by one order of magnitude for a 15%

crystallinity, the value found in the ultra low-density foams of LCB PLA. It is therefore expected

that crystallization increased the melt strength during cell growth. In the literature, crystallization

during foam extrusion is often assumed to suppress the expansion ratio, because it stiffens the

melt [Naguib et al. (2002)]. If the crystallinity is properly controlled, however, crystal structures

may increase the melt strength favorably and suppress diffusion of BA through the melt

[Hedenqvist and Gedde (1996)]. Foams with crystal structures also show better mechanical

properties and dimensional stability compared to amorphous foams [Klempner & Sendijarevic

(2004)]. Crystallization-induced gelation is also superior to that induced by cross-linking agents,

which are sometimes used in foam extrusion of low-melt-strength polymers, but introduces

99

90 100 110 120 130

0

5

10

15

Cry

sta

llization h

alf t

ime t

1/2 (

min

)

Isothermal crystallization temperature (oC)

LCB PLA

without shearing

after shearing at 160oC

shear rate = 50 s-1

tshear

= 20 s

Figure 5-12 Crystallization half time for the LCB PLA with and without shearing; the

crystallinity at these half times is approximately 15%

100

0 5 10 15 20 25 30 35

1

10

100

G' c

rysta

llin

e/G

' am

orp

ho

us

Crystallinity (%)

Figure 5-13 Ratio of the storage modulus of crystalline PLA to that of amorphous PLA as a

function of crystallinity; the material is LCB PLA without lubricant

101

permanent linkage between the chains and causes non-recyclability. In comparison,

crystallization is controlled by processing conditions and geometry of the die channel, and does

not impose any additional costs on the manufacturing process. Given all these potential

advantages, an experimental study was conducted to control crystallization during foam

processing and thereby to compare the resultant cell morphology. The results are presented in the

next section.

5.6 Controlling PLA Crystallization and Its Influence on Foaming

Crystallization of PLA during foam processing was controlled by varying the length of

the reservoir in the capillary die, as is already shown in Section 5.2, and thereby controlling the

isothermal residence time prior to cell nucleation and growth. Two capillary dies were designed.

Both dies have a capillary of 1 mm in diameter, but the capillary lengths are 6 mm and 10 mm,

respectively. The 6 mm die has a long reservoir corresponding to a residence time of

approximately 90 seconds prior to cell nucleation, while the 10 mm die has negligible reservoir,

corresponding to a residence time of approximately 0 seconds. Foaming experiments were

conducted using LCB PLA, 9% CO2, and following the same strategy as already described in

Section 5.2.

Expansion ratios of the foams as a function of processing temperature are presented for

the two dies in Figure 5-14. It is clearly seen that the die with a longer residence time induced

higher crystallinity, and high-expansion ratio foams could be produced over a much wider

temperature window compared to the short die. Below 115oC, the expansion starts to decrease

because of too high melt strength of the polymer. Foams produced from the short die, however,

show much lower crystallinity at the same temperature, and the expansion ratio keeps increasing

until the experiment had to be stopped because of too high die pressure, indicating that melt

strength is not high enough to induce the maximum expansion ratio.

SEM images of the cell structure are presented in Figure 5-15. Again, the influence of

crystallization is clearly seen. Foams with the highest crystallinity, those produced from the long

die at 112oC (Figure 5-15 (a)), show close-cell structure, while those produced from the long die

at a higher temperature (Figure 5-15 (b)) or those from the short die (Figure 5-15 (c) and (d))

102

110 115 120 125 130 1350

5

10

15

20

25

30

35

40

Exp

an

sio

n r

atio

Die temperature (oC)

LCB PLA

9% CO2

crystallization time = 0

crystallization time

= 1.5 minutes

Figure 5-14 Expansion ratios of foams produced after different time for crystallization. The

material is LCB PLA and crystallinity of the foam skin, determined on DSC, is shown for several

conditions

χ =3%

χ =18%

χ =15%

103

(a) (b)

(c) (d)

Figure 5-15 SEM images of the cellular structures, the polymer is LCB PLA, and 9% CO2 is

used: (a) time for crystallization ≈ 0 s, 112oC; (b) time for crystallization ≈ 0 s, 120

oC; (c) time

for crystallization ≈ 90 s, 110oC; (d) time for crystallization ≈ 90 s, 120

oC

104

show more open cells. For the short die, open cell content increases with the expansion ratio,

which is evidence that crystallization almost did not occur in this case. Foams with higher

crystallinity also have better surface finish and better mechanical properties.

105

Chapter 6 Conclusions

As stated in Chapter 1 of this thesis, the production of microcellular foams involves

subjecting a polymer/BA mixture to a series of well-defined kinematic motions, such that the

main stages of processing are precisely controlled. In spite of the rapid progress in developing

microcellular technology for commercial polymers during the past two decades, many challenges

still remain. The work presented in this thesis contributes to understanding the rheological

properties, especially the extensional properties, of polymer melts and polymer/BA solutions and

to understanding the relationship between material rheological properties and cell morphology

from processing. The polymers and the experimental techniques have been chosen so that the

physical processes of interest become dominant and can be investigated quantitatively. Among

the major accomplishments of this thesis are:

1. Design and construction of a rheological die for characterizing the shear and planar

extensional properties of polymer melts and polymer/BA solutions from pressure drop

measurement. The flow channel consists of a thin straight rectangular channel with a high

aspect ratio, for determining shear properties, followed by a thin hyperbolically

convergent rectangular channel, for determining extensional properties once shearing is

accounted for. This die design has several advantages over previous designs (e.g., Wang

and Park (2006)), which have contributed to the reliability of measured data: first,

pressure transducers can be flush-mounted on the channel wall to eliminate errors;

secondly, the flow channel can be easily machined to a high precision and its dimensions

allow the flow to be treated as a motion 2D; and finally, temperature can be well

controlled throughout the die.

2. Determination of the shear properties of a polymer melt and a polymer/CO2 solution

utilizing the rheological die, and comparison of data with those in the literature and those

determined on commercial rheometers. The viscosity of the melt was determined over

shear rates between 10-2

s-1

and 104 s

-1, and the data agree excellently with data using the

rheometers. The viscosity of the solution was determined over two decades of shear rate,

and the influence of CO2 on melt viscosity was found to follow a WLF-type equation.

106

3. Identified necessary conditions that enable fluid extensional properties to be determined

from a hyperbolically convergent channel. In previous studies using such channels, the

authors found that shearing is dominant and extensional information could not be

extracted. The results in this thesis demonstrate that, if the fluid is very elastic and very

shear-thinning, an approximately extensional flow is induced over a good fraction of the

hyperbolic channel, and planar extensional behaviour of the fluid can be characterized.

4. Determination of the planar extensional viscosity of a polymer melt and a polymer/CO2

solution from the rheological die, comparison of data with uniaxial extensional viscosity

determined on a shear-free rheometer, and investigation of the difference between the

two sets of data. Planar extensional viscosity was determined from pressure drop

measurements in the convergent channel after deducting the calculated pressure drop due

to shearing. Unlike previous entry flow analyses, which considered the extensional

viscosity as a function of extensional rate alone, the analysis in this thesis made a more

realistic assumption that extensional viscosity depends on both extensional rate and

strain. Extensional viscosity of the melt alone determined from the channel and that from

the shear-free rheometer compare reasonably, with possible error sources clearly

identified in the thesis. A comparison of the shear and extensional viscosities of the

polymer/CO2 solution indicates that the influence of CO2 on fluid rheological properties

is similar to an increase of temperature, because both increase the free volume of the

melt.

5. Characterized the thermal and rheological properties of biodegradable polymers with

different molecular structures, and investigated the influence of molecular branching on

the polymers‟ low-density, microcellular extrusion foaming behaviour. These polymers

have the potential to replace traditional non-biodegradable polymers in foaming

applications. With this polymer, the number of branches is well controlled, and it was

found that rheological properties, especially extensional properties, provide the most

sensitive probe to branching topology. Foaming experiments with these biodegradable

polymers indicate that long chain branching increases the cell density of foams either by

increasing the nucleation density or by suppressing cell coalescence during the early

stage of foaming. It was also found that long chain branching also increases melt

107

viscosity and strain to break, allowing the production of low-density foams with closed

cells.

6. Characterized the crystallization behaviour of both linear and branched biodegradable

polymers, and investigated the influence of CO2- and flow-induced crystallization on cell

morphology from low-density, microcellular extrusion foam processing. The influence of

crystallization on cell morphology has not been studied in the literature because, for most

commercial polymers, the crystallization kinetics is either too fast, such as polyolefins, or

too slow, such as polycarbonate, making it very difficult to control the degree of

crystallinity during processing. The biodegradable polymers used in this thesis, polylactic

acids, were synthesized to have mild, controllable crystallization kinetics. The kinetics

was characterized under well-controlled thermal and flow conditions relevant to

processing, and the information was used to design foaming experiments that produced

foams with controlled crystallinity. It was found that flow-induced crystallization

increases the melt strength and suppresses the permeation of CO2 through the melt,

thereby significantly reducing open-cell content in the foams. Because part of the

crystallinity was induced by cell growth, the melt did not become too stiff before they

reach their extensional limit, resulting in the production of high-expansion-ratio foams

with closed cells and good mechanical properties. For the first time, it was shown that

foams produced from crystallized melt also show better surface finish and is free from

post-processing volume shrinkage.

Many questions remain, of course, and in continuing this project in the future, the

following topics may be interesting from both theoretical and practical points of view:

1. Determining the actual velocity distribution over the hyperbolically convergent

rectangular channel for a highly-elastic fluid. In this thesis, it is assumed that velocity

distribution depends on the shear properties alone, and that no secondary flow is present.

These assumptions are reasonable considering that the channel is so thin and stress

gradient in secondary flow directions is small compared to the stress gradient generated

in the primary flow direction. Nevertheless, acceleration of flow by the convergent

profile and development of excessive extensional stress along a streamline may affect the

velocity distribution, possibly making the velocity profile in the channel thickness

108

direction “blunter” than the calculated profile in this thesis. The velocity distribution, at

least over the mid-plane of the channel, may be determined by particle imaging

velocimetry (PIV) technique.

2. Calculating stress and velocity distribution over the convergent channel using proper

constitutive models for viscoelastic fluids. From the literature (e.g., Larson (1988)), we

know that no constitutive equation is capable of describing all types of extensional

properties for a polymer melt using only one set of parameters. Nevertheless, the free

parameters may be determined using shear and uniaxial extensional viscosities

determined on commercial rheometers and then the stress distribution in the thickness

direction may be calculated for the center plane of the hyperbolic channel. Such

calculations will yield information about the stress-strain relationship for a viscoelastic

fluid in a mixed flow and allowing theory and experimental results to be compared.

3. Predicting pressure drop over complicated die channels and injection mold cavities using

rheological properties determined from commercial or in-house techniques. A full-scale

simulation using constitutive equations involving multiple relaxation times is time

consuming and subject to error. The technique used in this thesis for calculating pressure

drops allows for simple, albeit approximate, calculation to be carried out. This involves

assuming that the velocity distribution depends on shearing alone, and that the pressure

drop due to shearing and extension can be determined independently by integrating the

stress along a streamline. The usefulness and reliability of this technique for calculating

general flow channels needs to be examined.

4. Studying the development of crystal structure during foam processing. For a crystalline

polymer such as PLA, it is useful to control the crystal morphology, including crystal

size, shape, and orientation, so that they are favorable for developing desired cell

morphology and for further processing. One example of further processing concerns the

production of expanded-PLA foams. This involves cutting the extruded PLA foam

filament into small beads, and then joining these beads in a mold under elevated

temperature and pressure, usually using hot steam, to produce a foamed product of any

shape. The crystal structure on the surface of the foam beads is critical to the bonding

109

strength between beads, and hence the mechanical properties of the entire foamed

product.

5. Finally, once the foaming behaviour of the PLA alone has been clarified, it is important

to investigate the behaviour of blends of PLA with other polymers. These blends

typically show phase behaviour and rheological properties between those of the two

components, and their foaming behaviour will depend on factors such as the domain

morphology, the interface compatibility, and solubility of blowing agent in the two

components. Foams produced from these blends may have different physical properties

than foams produced from either of the components, and may be used in a variety of

applications.

110

Bibliography

1. Amon, M., C. D. Denson, “Study of the dynamics of foam growth: analysis of the growth of

closely spaced spherical bubbles,” Polym. Eng. Sci. 24, 1026-1034 (1984)

2. Amon, M., C. D. Denson, “Study of the dynamics of foam growth: simplified analysis and

experimental results for bulk density in structural foam molding,” Polym. Eng. Sci. 26, 255-

267 (1986)

3. Areerat, S., T. Nagata and M. Ohshima, “Measurement and prediction of LDPE/CO2 solution

viscosity,” Polym. Eng. Sci. 42, 2234-2245 (2002)

4. Auhl, D., J. Stange, H. Münstedt, et al., “Long-chain branched polypropylenes by electron

beam irradiation and their rheological properties,” Macromolecules 37, 9465-9472 (2004)

5. Baldwin, D. F., D. Tate, C. B. Park, S. W. Cha and N. P. Suh, “Microcellular plastics

processing technology (1),” Journal of Japan Society of Polymer Processing (SEIKEI-

KAKOU) 6, 187-194 (1994a)

6. Baldwin, D. F., D. Tate, C. B. Park, S. W. Cha and N. P. Suh, “Microcellular plastics

processing technology (2),” Journal of Japan Society of Polymer Processing (SEIKEI-

KAKOU) 6, 245-256 (1994b)

7. Bao, L., J. R. Dorgan, D. Knauss, S. Hait, N. S. Oliveira and I. M. Maruccho, “Gas

permeation properties of poly(lactic acid) revisited,” J. Membrane Sci. 285, 166-172 (2006)

8. Barlow, E. J. and W. E. Langlois, “Diffusion of gas from a liquid into an expanding bubble,”

IBM J., 329-337 (1962)

9. Bastioli, C. Eds., Handbook of biodegradable polymers, Rapra Technology Ltd., Shawbury,

UK (2005)

10. Batchelor, G. K., An introduction to fluid dynamics, Cambridge University Press, Cambridge

(2000)

11. Bikiaris, D. N. and G. P. Karayannidis, “Chain extension of polyesters PET and PBT with

N,N‟-Bis(glycidyl ester) Pyromellitimides. I,” J. Polym. Sci.: Part A: Polym. Chem. 33,

1705-1714 (1995)

12. Binding, D. M., “An approximate analysis for contraction and converging flows,” J. Non-

Newtonian Fluid Mech. 27, 173-189 (1988)

13. Bird, R. B., R. C. Armstrong and O. Hassager, Dynamics of Polymeric Liquids, Vol. 1 –

Fluid Mechanics, John Wiley & Son, New York, (1987)

14. Boger, D. V., “Viscoelastic flows through contractions,” Ann. Rev. Fluid Mech. 19, 157-182

(1987)

15. Carraher, C. E., Polymer Chemistry, Sixth ed., Marcel Dekker, New York (2003)

16. Chatraei, S. M., C. W. Macosko and H. H. Winter, “Lubricated squeezing flow: a new biaxial

extensional rheometer,” J. Rheol. 25, 433-443 (1981)

17. Cogswell, F. N., “Converging flow of polymer melts in extrusion dies,” Polym. Eng. Sci. 12,

64-73 (1972)

111

18. Colton, J. S. and N. P. Suh, “The nucleation of microcellular thermoplastic foam with

additives: part I: theoretical considerations,” Polym. Eng. Sci. 27, 485-492 (1987)

19. Davies, J. M., J. F. Hutton, and K. Walters, “Theory for normal stresses in slits and

capillaries,” J. Phys. D: Appl. Phys. 6, 2259-2266 (1973)

20. Dealy, J. M., “Official nomenclature for material functions describing the response of a

viscoelastic fluid to various shearing and extensional deformations,” J. Rheol., 28, 282-295

(1984)

21. Di, Y., S. Iannace, E. D. Maio and L. Nicolais, “Poly(lactic acid)/organoclay

nanocomposites: thermal, rheological properties and foam processing,” J. Polym. Sci. Part B:

Polym. Phys. 43, 689-698 (2005)

22. Drumright, R. E., P. R. Gruber and D. E. Henton, “Polylactic acid technology,” Adv. Mater.

12, 1841-1846 (2000)

23. Elkovitch, M. D., D. L. Tomasko and L. J. Lee, “Supercritical carbon dioxide assisted

blending of polystyrene and poly(methyl methacrylate),” Polym. Eng. Sci. 39, 2075-2084

(1999)

24. Elkovitch, M. D., D. L. Tomasko and L. J. Lee, “Effect of supercritical carbon dioxide on

morphology development during polymer blending,” Polym. Eng. Sci. 40, 1850-1861 (2000)

25. Elkovitch, M. D., D. L. Tomasko and L. J. Lee, “Effect of supercritical carbon dioxide on

PMMA/rubber and polystyrene/rubber blending: viscosity ratio and phase inversion,” Polym.

Eng. Sci. 41, 2108-2125 (2001)

26. Ema, Y., M. Ikeya and M. Okamoto, “Foam processing and cellular structure of polylactide-

based nanocomposites,” Polymer 47, 5350-5359 (2006)

27. Feigl, K., F. X. Tanner, B. J. Edwards and J. R. Collier, “A numerical study of the

measurement of elongational viscosity of polymeric fluids in a semihyperbolically

converging die,” J. Non-Newtonian Fluid Mech. 115, 191-215 (2003)

28. Ferry, J. D., Viscoelastic Properties of Polymers, John Wiley & Sons, New York (1980)

29. Frazza, E. J. and E. E. Schmitt, “A new absorbable suture,” J. Biomed. Mater. Res. 5, 43-58

(1971)

30. Fuller, G. G. and L. G. Leal, “Flow birefringence of dilute polymer solutions in two-

dimensional flows,” Rheol. Acta 19, 580-600 (1980)

31. Fuller, G. G. and L. G. Leal, “Flow birefringence of concentrated polymer solutions in two-

dimensional flows,” J. Polym. Sci.: Polym. Phys. Edition 19, 557-587 (1981)

32. Gendron, R., L. E. Daigneault and L. M. Caron, “Rheological behavior of

polystyrene/blowing agent mixtures,” SPE ANTEC Tech Papers, 1118-1122 (1996)

33. Gendron, R. and L. E. Daigneault, “Rheological behavior of mixtures of various polymer

melts with CO2,” SPE ANTEC Tech Papers, 1096-1100 (1997)

34. Gendron, R. and A. Correa, “The use of on-line rheometry to characterize polymer melts

containing physical blowing agents,” Cell. Polym. 17, 93-113 (1998)

35. Gerhardt, L. J., C. W. Manke and E. Gulari, “Rheology of polydimethysiloxane swollen with

supercritical carbon dioxide,” J. Polym. Sci., Part B: Polym. Phys. 35, 523-534 (1997)

112

36. Gerhardt, L. J., A. Garg, C. W. Manke and E. Gulari, “Concentration-dependent viscoelastic

scaling models for polydimethysiloxane melts with dissolved carbon dioxide,” J. Polym. Sci.:

Part B: Polym. Phys. 37, 1911-1918 (1998)

37. Gotsis, A. D. and A. Odriozola, “The relevance of entry flow measurements for the

estimation of extensional viscosity of polymer melts,” Rheol. Acta 37, 430-437 (1998)

38. Guo, Q, J. Wang, C. B. Park and M. Oshima, “A microcellular foaming simulation system

with a high-pressure drop rate,” Ind. & Eng. Chem. Res. 45, 6153-6161 (2006)

39. Han, C. D. and C. Y. Ma, “Rheological properties of mixtures of molten polymer and

fluorocarbon blowing agent. I. Mixtures of low-density polyethylene and fluorocarbon

blowing agent,” J. Appl. Polym. Sci. 28, 831-850 (1983a)

40. Han, C. D. and C. Y. Ma, “Rheological properties of mixtures of molten polymer and

fluorocarbon blowing agent. II. Mixtures of polystyrene and fluorocarbon blowing agent,” J.

Appl. Polym. Sci. 28, 851-860 (1983b)

41. Hedenqvist, M. and U. W. Gedde, “Diffusion of small-molecule penetrants in semicrystalline

polymers,” Prog. Polym. Sci. 21, 299-333 (1996)

42. James, D. F., B. D. McLean and J. H. Saringer, “Presheared extensional flow of dilute

polymer solutions,” J. Rheol. 31, 453-481 (1987)

43. Johansen, P., Y. Men, H. P. Merkle and B. Gander, “Revisiting PLA/PGLA microspheres: an

analysis of their potential in parenteral vaccination,” Eur. J. Pharm. Biopharm. 50, 129-146

(2000)

44. Jones, D. M., K. Walters and P. R. Williams, “On the extensional viscosity of mobile

polymer solutions,” Rheol. Acta 26, 20-30 (1987)

45. Kim, H. C., A. Pendse and J. R. Collier, “Polymer melt lubricated elongational flow,” J.

Rheol. 38, 831-845 (1994)

46. Klempner, D. and V. Sendijarevic, editors, Handbook of Polymeric Foams and Foam

Technology, Hancer, Munich (2004)

47. Kostic M., “Influence of viscosity function simplification on non-Newtonian velocity and

shear-rate profiles in rectangular ducts,” Int. Comm. Heat Mass Transfer 20, 515-525 (1993)

48. Kulikov, O., K. Hornung and M. Wagner, “Silanols cured by borates as lubricants in

extrusion of LLDPE. Impact of elasticity of the lubricant on sliding friction,” Rheol. Acta 46,

741-754 (2007)

49. Kwag, C., C. W. Manke and E. Gulari, “Rheology of molten polystyrene with dissolved

supercritical and near-critical gases,” J. Polym. Sci.: Part B: Polym. Phys. 37, 2771-2781

(1999)

50. Kwag, C., C. W. Manke and E. Gulari, “Effects of dissolved gas on viscoelastic scaling and

glass transition temperature of polystyrene melts,” Ind. Eng. Chem. Res. 40, 3048-3052

(2001)

51. Ladin, D., C. B. Park, S. S. Park, H. E. Naguib and S. W. Cha, “Study of shear and

extensional viscosities of biodegradable PBS/CO2 solutions,” J. Cellular Plastics 37, 109-148

(2001)

52. Langer, R. and J. P. Vacanti, “Tissue engineering,” Science 260, 920-926 (1993)

113

53. Larson, R. G., Constitutive equations for polymer melts and solutions, Butterworth

Publishers (1988)

54. Laun, H. M., “Polymer melt rheology with a slit die,” Rheol. Acta 22, 171-185 (1983)

55. Laun, H. M. and H. Schuch, “Transient elongational viscosities and drawability of polymer

melts,” J. Rheol. 33, 119-175 (1989)

56. Lee, M., C. B. Park and C. Tzoganakis, “Measurements and modeling of PS/supercritical

CO2 solution viscosities,” Polym. Eng. Sci. 39, 99-109 (1999)

57. Lee, M., C. Tzoganakis and C. B. Park, “Effects of supercritical CO2 on the viscosity and

morphology of polymer blends,” Adv. Polym. Technol. 19, 300-311 (2000)

58. Lee, P. C., J. Wang and C. B. Park, “Extruded open-cell foams using two semicrystalline

polymers with different crystallization temperatures,” Ind. Eng. Chem. Res. 45, 175-181

(2006)

59. Lee, S.-T. and N. S. Ramesh, Polymeric Foams – Mechanisms and Materials, CRC Press,

Florida (2004)

60. Lee, S.-T., C. B. Park and N. S. Ramesh, Polymeric Foams – Science and Technology, CRC

Press, Florida (2007)

61. Lee, S. T., L. Kareko and J. Jun, “Study of thermoplastic PLA foam extrusion,” J. Cellular

Plastics 44, 293-305 (2008)

62. Leung, S. N., C. B. Park, D. Xu, H. Li and R. G. Fenton, “Computer simulation of bubble-

growth phenomena in foaming,“ Ind. Eng. Chem. Res. 45, 7823-7831 (2006)

63. Li, Y. G., “Development of a novel apparatus for the PVT visualization and measurement of

polymer/gas solutions,” Ph.D. Thesis, University of Toronto, Canada (2008)

64. Li, G., J. Wang, C. B. Park, P. Moulinie and R. Simha, ”Comparison of the SS-based and SL-

based estimation of gas solubility,” SPE ANTEC Tech. Papers, Paper #421 (2004)

65. Li, G., H. Li, L. S. Turng, S. Gong and C. Zhang, “Measurement of gas solubility and

diffusivity in polylactide,” Fluid Phase Equili. 246, 158-166 (2006)

66. Macosko, C. W., Rheology – Principles, Measurements, and Applications, VCH Publishers,

New York (1994)

67. Maquet, V., S. Blacher, R. Pirard, J. – P. Pirard and R. Jerome, “Characterization of

Polylactide Foams by Image Analysis and Impedance Spectroscopy,” Langmuir 16, 10463-

10470 (2000)

68. McKinley, G. H. and O. Hassager, “The Considère condition and rapid stretching of linear

and branched polymer melts,” J. Rheol. 43, 1195-1212 (1999)

69. Meissner, J., “Polymer melt elongation – methods, results, and recent developments,” Polym.

Eng. Sci. 27, 537-546 (1987)

70. Meissner, J. and J. Hostettler, “A new elongational rheometer for polymer melts and other

highly viscoelastic liquids,” Rheol. Acta 33, 1-21 (1994)

71. Michaeli, W., J. Lorenz, R. van Hoorn and H. Schumacher, “Assessment of the foamability

of polymers on the basis of their biaxial stress/strain behaviour,” Proceedings of the Blowing

Agent and Foaming Processes (2004)

114

72. Mikos, A. G., G. Sarakinos, S. M. Leite, J. P. Vacanti and R. Langer, “Laminated three-

dimensional biodegradable foams for use in tissue engineering,” Biomaterials 14, 323-330

(1993)

73. Mooney, D. J., D. F. Baldwin, N. P. Suh, J. P. Vacanti and R. Langer, “Novel approach to

fabricate porous sponges of poly(D,L-lactic-co-glycolic acid) without the use of organic

solvents,” Biomaterials 17, 1417-1422 (1996)

74. Münstedt, H, “Dependence of the elongational behavior of polystyrene melts on molecular

weight and molecular weight distribution,” J. Rheol. 24, 847-867 (1980)

75. Naguib, H. E., C. B. Park, U. Panzer and N. Reichelt, “Strategies for achieving ultra low-

density PP foams,” Polym. Eng. Sci. 42, 1481-1492 (2002)

76. Oliveira, M. S. N., M. A. Alves, F. T. Pinho and G. H. McKinley, “Viscous flow through

microfabricated hyperbolic contractions,” Exp. Fluids 43, 437-451 (2007)

77. Ouchi, S., Y. Masubuchi and H. Shikuma, “The effect of CO2 pressure on viscoelasticity of

LDPE,” Inter. Polym. Processing 5, 173-177 (2008)

78. Papanastasiou, A. C., L. E. Scriven and C. W. Macosko, “Bubble growth and collapse in

viscoelastic liquids analyzed,” J. Non-Newt. Fluid Mech. 14, 53-75 (1984)

79. Park, C. B. and N. P. Suh, “Filamentary extrusion of microcellular polymers using a rapid

decompressive element,” Polym. Eng. Sci. 36, 34-48 (1996)

80. Park, C. B. and L. K. Cheung, “A study of cell nucleation in the extrusion of polypropylene

foams,” Polym. Eng. Sci. 37, 1-10 (1997)

81. Park, C. B., A. H. Behravesh and R. D. Venter, “Low-density, microcellular foam processing

in extrusion using CO2,” Polym. Eng. Sci. 38, 1812-1823 (1998)

82. Park, C. B., “Continuous production of high-density and low-density microcellular plastics in

extrusion,” in Foam Extrusion, edited by S. T. Lee, pp. 263-305, Technomic Publishing

Company, Lancaster, Pennsylvania (2002)

83. Park, H. E. and J. M. Dealy, “Effects of pressure and supercritical fluids on the viscosity of

polyethylene,” Macromolecules 39, 5438-5452 (2006)

84. Pilla, S., S. G. Kim, G. K. Auer, S. Gong and C. Park, “Microcellular extrusion of polylactide

with chain-extender,” SPE ANTEC Tech. Papers, 1908-1912 (2008)

85. Ramesh, N. S., D. H. Rasmussen and G. A. Campbell, “The heterogeneous nucleation of

microcellular foams assisted by the survival of microvoids in polymers containing low glass

transition particles. part I: mathematical modeling and numerical simulation,” Polym. Eng.

Sci. 34, 1685-1697 (1994a)

86. Ramesh, N. S., D. H. Rasmussen and G. A. Campbell, “The heterogeneous nucleation of

microcellular foams assisted by the survival of microvoids in polymers containing low glass

transition particles. part II: experimental results and discussion,” Polym. Eng. Sci. 34, 1698-

1706 (1994b)

87. Reignier, J., R. Gendron and M. F. Champagne, “Extrusion foaming of poly(lactic acid)

blown with CO2: toward 100% green materials,” Cellular Polymers 26, 83-115 (2007)

115

88. Royer, J. R., Y. J. Gay, J. M. Desimone and S. A. Khan, “High-pressure rheology of

polystyrene melts plasticized with CO2: experimental measurement and predictive scaling

relationships,” J. Polym. Sci.: Part B: Polym. Phys. 38, 3168-3180 (2000)

89. Royer, J. R., J. M. Desimone and S. A. Khan, “High-pressure rheology and viscoelastic

scaling predictions of polymer melts containing liquid and supercritical carbon dioxide,” J.

Polym. Sci.: Part B: Polym. Phys. 39, 3055-3066 (2001)

90. Royer, J. R., Y. J. Gay, M. Adam, J. M. Desimone and S. A. Khan, “Polymer melt rheology

with high-pressure CO2 using a novel magnetically levitated sphere rheometer,” Polymer 43,

2375-2383 (2002)

91. Sawyer, D. J., “Bioprocessing – no longer a field of dreams,” Macromol. Symp. 201, 271-

281 (2003)

92. Schultz, J. M., Polymer crystallization: the development of crystalline order in thermoplastic

polymers, Oxford University Press, Washington D. C. (2001)

93. Sentmanat, M. L., “Dual windup extensional rheometer,” U.S. Patent No. 6,578,413 (2003)

94. Sentmanat, M. L., “Miniature universal testing platform: from extensional melt rheology to

solid-state deformation behavior,” Rheol. Acta 43, 657-669 (2004)

95. Sentmanat, M., B. N. Wang and G. H. McKinley, “Measuring the transient extensional

rheology of polyethylene melts using the SER universal testing platform,” J. Rheol. 49, 585-

606 (2005)

96. Spitael, P. and C. W. Macosko, “Strain hardening in polypropylenes and its role in extrusion

foaming,” Polym. Eng. Sci. 44, 2090-2100 (2004)

97. Spitael, P., C. W. Macosko and R. B. McClurg, “Block copolymer micelles for nucleation of

microcellular thermoplastic foams,” Macromolecules 37, 6874-6882 (2004)

98. Stange, J. and H. Münstedt, “Rheological properties and foaming behavior of polypropylenes

with different molecular structures,” J. Rheol. 50, 907-923 (2006)

99. Street, J. R., “The rheology of phase growth in elastic liquids,” Trans. Soc. Rheol. 12, 103-

131 (1968)

100. Suh, K. W., C. P. Park, M. J. Maurer, M. H. Tusim, R. D. Genova, R. Broos and D. P.

Sophiea, “Lightweight cellular plastics,” Adv. Mater. 12, 1779-1789 (2000)

101. Syrjälä, S., “Finite-element analysis of fully developed laminar flow of power-law non-

Newtonian fluid in a rectangular duct,” Int. Comm. Heat Mass Transfer 22, 549-558 (1995)

102. Tadmor, Z. and I. Klein, Engineering Principles of Plasticating Extrusion, Van Nostrand

Reinhold, New York (1970)

103. Takada, M., S. Hasegawa and M. Ohshima, “Crystallization kinetics of Poly(L-lactide) in

contact with pressurized CO2,” Polym. Eng. Sci. 44, 186-196 (2004)

104. Taki, K., K. Tabata, S. Kihara and M. Ohshima, “Bubble coalescence in foaming process

of polymers,” Polym. Eng. Sci. 46, 680-690 (2006)

105. The Freedonia Group, Foamed Plastics, Industry Study #1436 (2001)

106. Throne, J. L., Thermoplastic Foams, Sherwood Publishers, Ohio (1996)

116

107. Tomasko, D. L., H. Li, D. Liu, X. Han, M. J. Wingert, L. J. Lee and K. W. Koelling, “A

review of CO2 applications in the processing of polymers,” Ind. Eng. Chem. Res. 42, 6431-

6456 (2003)

108. Tremblay, B., “Estimation of the elongational viscosity of polyethylene blends at high

deformation rates,” J. Non-Newt. Fluid Mech. 33, 137-164 (1989)

109. Utracki, L. A. and R. Simha, “Free volume and viscosity of polymer-compressed gas

mixtures during extrusion foaming,” J. Polym. Sci.: Part B: Polym. Phys. 39, 342-362 (2001)

110. Villalobos, M., A. Awojulu, T. Greeley, G. Turco and G. Deeter, “Oligomeric chain

extenders for economic reprocessing and recycling of condensation plastics,” Energy 31,

3227-3234 (2006)

111. Wang, X., V. Kumar and W. Li, “Low density sub-critical CO2 blown solid-state PLA

foams”, Cellular Polymers 26, 11-35 (2007)

112. Wang, J. and C. B. Park, “Pressure profile in annular die using PP/CO2 solution

viscosity,” SPE ANTEC Tech. Papers, 896-902 (2006)

113. Whitesides, G. M. and A. D. Stroock, “Flexible methods for microfluidics,” Phys. Today

54, 42-48 (2001)

114. William, M. L., R. F. Landel and J. D. Ferry, “The temperature dependence of relaxation

mechanisms in amorphous polymers and other glass-forming liquids,“ J. Amer. Chem.

Soc.77, 3701-3707 (1955)

115. Wingert, M. J., S. Shukla, K. W. Koelling, D. L. Tomasko and L. J. Lee, “Shear viscosity

of CO2-plasticized polystyrene under high static pressure,” Ind. & Eng. Chem. Res., to

appear

116. Wissinger, R. G. and M. E., Paulaitis, “Swelling and sorption in polymer-CO2 mixtures at

elevated pressures,” J. Polym. Sci.: Part B: Polym. Phys. 25, 2497-2510 (1987)

117. Xu, X., C. B. Park, D. Xu and R. Pop-Iliev, “Effects of die geometry on cell nucleation of

PS foams blown with CO2,” Polym. Eng. Sci. 43, 1378-1390 (2003)

118. Xue A. and C. Tzoganakis, “Rheological properties of polystyrene/supercritical CO2

solution from an extrusion slit die,” J. Polym. Eng. 23, 1-22 (2003)

119. Yao, M. and G. H. McKinley, “Shear history effects on extensional flow of non-

Newtonian fluids in filament stretching rheometers,” XVth International Congress on

Rheology, Paper #CF53, Monterey, California (2008)

120. Yu, L., H. Liu, K. Dean and L. Chen, “Cold crystallization and postmelting

crystallization of PLA plasticized by compressed carbon dioxide,” J. Polym. Sci.: Part B.:

Polym. Phys. 46, 2630-2636 (2008)

121. Zhu, Z., D. Xu, C. B. Park and R. G. Fenton, “Finite element analysis of cell coarsening

in plastic foaming,” J. Cellular Plastics 41, 475-486 (2005)

117

Appendix

CAD Drawing of the Rheological Die