development of expanded thermoplastic … development of expanded thermoplastic polyurethane bead...

203
Development of Expanded Thermoplastic Polyurethane Bead Foams and Their Sintering Mechanism by Nemat Hossieny 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 Nemat Hossieny 2014

Upload: vodien

Post on 16-Apr-2018

225 views

Category:

Documents


5 download

TRANSCRIPT

Development of Expanded Thermoplastic Polyurethane Bead Foams and Their Sintering Mechanism

by

Nemat Hossieny

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 Nemat Hossieny 2014

ii

Development of Expanded Thermoplastic Polyurethane Bead

Foams and Their Sintering Mechanism

Nemat Hossieny

Degree of Doctor of Philosophy

Department of Mechanical and Industrial Engineering

University of Toronto

2014

Abstract

Polymer bead foaming technology represents a breakthrough in the production of low density

plastic foamed components that have a complex geometrical structure and has helped to expand

the market for plastic foams by broadening their applications. In this research, the unique

microstructure of thermoplastic polyurethane (TPU) consisting of phase-separated hard segment

(HS) domains dispersed in the soft segment (SS) matrix has been utilized to develop expanded

TPU (E-TPU) bead foam with microcellular morphologies and also to create inter-bead sintering

into three dimensional products using steam-chest molding machine. The phase-separation and

crystallization behavior of the HS chains in the TPU microstructure was systematically studied in

the presence of dissolved gases and also by changing the microstructure of TPU by melt-

processing and addition of nano-/micro-sized additives. It was observed that the presence of gas

improved the phase separation (i.e. crystallization) of HSs and increased the overall crystallinity

of the TPU. It was also shown that by utilizing the HS crystalline domains, the overall foaming

behavior of TPU (i.e. cell nucleation and expansion ratio) can be significantly improved.

Moreover, the HS crystalline domains can be effective for both sintering of the beads as well

strengthening the individual beads to improve the property of the moulded part. It was also

observed that unlike other polymer bead foaming technologies, the E-TPU bead foaming

iii

sintering does not require formation of double melting-peak. The original broad melting peak

existing in the TPU microstructure due to the wide size distribution of HS crystallites can be

effectively utilized for the purpose of sintering as well as maintenance of the overall dimensional

stability of the moulded part.

iv

Acknowledgments

I would like to express my sincere gratitude and appreciation to my supervisor, Professor Chul B.

Park for providing me with the continuous guidance, enthusiasm and encouragement to assist me

in conducting a successful research. His visions, insights and suggestions have an everlasting

influence on my personal and professional growth. I feel extremely honored and fortunate to

have such a supportive mentor.

I would like to thank my Ph.D. committee members, Professor Hani Naguib and Professor Glenn

D. Hibbard for their valuable comments and suggestions offered during the course of my Ph.D.

research. Also, I am grateful for Professor Anup Ghosh and Professor Lidan You for their

valuable feedback in my Ph.D. final oral examination.

I am grateful of the financial support and scholarships from Ontario Graduate Scholarship

(OGS), Consortium of Cellular and Micro-Cellular Plastics (CCMCP), and Natural Sciences and

Engineering Research Council of Canada (NSERC) funding for Network for Innovative Plastics

Materials and Manufacturing Processes (NIPMMP).

My special thanks goes to Kara Kim for her kind assistance. I would like to extend my

acknowledgment to my colleagues and other members of Microcellular Plastic Manufacturing

Laboratory. Their advice, assistance and friendship have contributed to the successful completion

of my research. Special thanks goes to Dr. Changwei Zhu, Dr. Saleh Amani, Hasan Mahmood,

Dr. Reza Barzegari, Dr. Reza Nofar, Dr. Amir Ameli, Alireza Tabatabaei, Mehdi Saniei, Vahid

Shaayegan, Lun Howe Mark, Weidan Ding, Davoud Jahani, Ali Rizvi, Mo Xu, Sai Wang,

Raymond Chu, Dr. Peter Jung, Dr. Anson Wong, Hui Wang, Anne Zhao as well as everyone else

who helped me in my Ph.D. studies. I am also grateful for the many undergraduate students who

have assisted me during the course of my research. Also, to the administrative staff in our

department: Konstantine, Brenda, Ceaser and Jho: thank you for your kind assistance on the

various administrative matters.

Finally, my special thanks go to my family members in India and Canada for their support,

encouragement and patience throughout the course of this Ph.D. research.

v

Table of Contents

Acknowledgments .......................................................................................................................... iv

Table of Contents ............................................................................................................................ v

List of Tables .................................................................................................................................. x

List of Figures ................................................................................................................................ xi

List of Symbols ............................................................................................................................ xix

Chapter 1 Introduction .................................................................................................................... 1

1.1 Thermoplastic Foams .......................................................................................................... 1

1.2 Classification of Thermoplastic Foams .............................................................................. 1

1.3 Bead Foam Technology ...................................................................................................... 2

1.4 Research Motivation ........................................................................................................... 2

1.5 Objective of Thesis ............................................................................................................. 3

1.6 Organization of Thesis ........................................................................................................ 4

1.7 References ........................................................................................................................... 5

Chapter 2 Literature Review ........................................................................................................... 7

2 Literature Review ....................................................................................................................... 7

2.1 Basic and General Principles of Foaming ........................................................................... 7

2.1.1 Polymeric foams and foaming process ................................................................... 7

2.1.2 Polymeric foams and foaming process ................................................................... 8

2.1.3 Supercritical CO2 (scCO2) foaming ...................................................................... 11

2.2 Extrusion Foaming Technology ........................................................................................ 17

2.3 Injection Foam Molding Technology ............................................................................... 19

2.3.1 Conventional foam injection molding and microcellular injection molding

technologies .......................................................................................................... 19

2.3.2 Low-pressure and high-pressure foam injection molding technologies ............... 20

2.4 Rotational Foam Molding Technology ............................................................................. 22

vi

2.5 Bead Foam Molding Technology ..................................................................................... 24

2.5.1 Bead fabrication .................................................................................................... 25

2.5.2 Bead bonding ........................................................................................................ 26

2.5.3 Bead foam materials ............................................................................................. 33

2.6 Thermoplastic polyurethane .............................................................................................. 40

2.7 References ......................................................................................................................... 42

Chapter 3 Phase Separation and Crystallization of TPU in the Presence of Dissolved Gas:-

Effects of Processing, Nano-/Micron-Sized Additives and Gas Types ................................... 57

3 Phase Separation and Crystallization of TPU in the Presence of Dissolved Gas .................... 57

3.1 Introduction ....................................................................................................................... 57

3.2 Experimental Procedure .................................................................................................... 59

3.2.1 Materials ............................................................................................................... 59

3.2.2 Sample preparation ............................................................................................... 59

3.2.3 Rheological analysis ............................................................................................. 60

3.2.4 Atomic force microscopy ...................................................................................... 61

3.2.5 Crystallization analysis of TPU at ambient pressure ............................................ 61

3.2.6 Crystallization analysis of TPU at high-pressure with dissolved gas ................... 62

3.2.7 Phase separation and crystallization analysis using X-ray diffraction .................. 64

3.3 Results and Discussions .................................................................................................... 65

3.3.1 Rheological behavior of TPU and TPU nano-/micro-composites ........................ 65

3.3.2 Atomic force microscopy ...................................................................................... 68

3.3.3 Crystallization analysis of TPU at ambient pressure ............................................ 70

3.3.4 Crystallization analysis of TPU in presence of high-pressure dissolved gas ........ 78

3.3.5 WAXS analysis ..................................................................................................... 89

3.3.6 SAXS analysis ...................................................................................................... 91

3.4 Conclusions ....................................................................................................................... 92

vii

3.5 References ......................................................................................................................... 93

Chapter 4 Foaming Behavior of TPU in Simulation Foaming Setup:- Effects of HS

Crystallites, Nano-/Micro-Sized Additives, Blowing Agent Types and Foaming Methods.... 97

4 Foaming Behavior of TPU in Simulation Foaming Setup ....................................................... 97

4.1 Introduction ....................................................................................................................... 97

4.2 Experimental Procedure .................................................................................................... 99

4.2.1 Materials ............................................................................................................... 99

4.2.2 Sample preparation ............................................................................................... 99

4.2.3 Butane sorption experiment ................................................................................ 100

4.2.4 Foaming setup and procedure ............................................................................. 100

4.2.5 Foam characterization ......................................................................................... 102

4.3 Results and Discussion ................................................................................................... 103

4.3.1 Sorption of butane in TPU .................................................................................. 103

4.3.2 Effect of HS crystallites on foaming of TPU with butane .................................. 103

4.3.3 Foaming of TPU and TPU nano-clay nanocomposites with CO2 and water ...... 111

4.4 Conclusions ..................................................................................................................... 114

4.5 References ....................................................................................................................... 115

Chapter 5 Modification of Steam-Chest Molding Technology .................................................. 118

5 Modification of Steam-Chest Molding Technology .............................................................. 118

5.1 Introduction ..................................................................................................................... 118

5.2 Theoretical Background .................................................................................................. 119

5.3 Modifications on Steam-Chest Molding Machine to Incorporate Hot Air ..................... 122

5.4 Experimentation .............................................................................................................. 123

5.4.1 Materials ............................................................................................................. 123

5.4.2 Steam-chest molding setup and experimental design ......................................... 124

5.4.3 Surface quality characterization .......................................................................... 125

viii

5.4.4 Tensile property characterization ........................................................................ 126

5.4.5 Thermal property characterization ...................................................................... 127

5.5 Results and Discussion ................................................................................................... 127

5.5.1 Effect of hot air on the steaming time ................................................................. 127

5.5.2 Effect of hot air on the total processing temperature .......................................... 128

5.5.3 Effect of hot air flow rate on surface properties ................................................. 131

5.5.4 Effect of hot air temperature on surface properties ............................................ 135

5.5.5 Effect of hot air pressure on surface properties .................................................. 135

5.5.6 Thermal properties of molded EPP samples ....................................................... 137

5.5.7 Effect of hot air on tensile properties .................................................................. 139

5.6 Conclusions ..................................................................................................................... 141

5.7 References ....................................................................................................................... 142

Chapter 6 Processing of TPU Bead Foams In Lab-Scale Bead Foaming System and Sintering

Mechanism With Steam-Chest Molding Technology ............................................................ 145

6 Production and Sintering of E-TPU Beads ............................................................................ 145

6.1 Introduction ..................................................................................................................... 145

6.2 Materials and Experimental Procedure ........................................................................... 146

6.2.1 Materials ............................................................................................................. 146

6.2.2 Lab-scale bead foaming setup ............................................................................. 147

6.2.3 Expanded TPU (E-TPU) bead foaming procedure ............................................. 147

6.2.4 Thermal behavior of E-TPU beads ..................................................................... 148

6.2.5 Gel Permeation Chromatography (GPC) ............................................................ 148

6.2.6 Water up-take analysis ........................................................................................ 149

6.2.7 Foam characterization ......................................................................................... 149

6.2.8 Steam-chest molding of E-TPU beads ................................................................ 150

6.2.9 Mechanical property measurement ..................................................................... 150

ix

6.3 Results and Discussions .................................................................................................. 151

6.3.1 Foaming behavior of E-TPU beads ..................................................................... 151

6.3.2 Characterization of TPU ..................................................................................... 155

6.3.3 Thermal behavior of E-TPU bead foams ............................................................ 157

6.3.4 GPC analysis ....................................................................................................... 161

6.3.5 Sintering of E-TPU beads with steam-chest molding machine .......................... 162

6.4 Conclusions ..................................................................................................................... 176

6.5 References ....................................................................................................................... 177

Chapter 7 Conclusion and Future Recommendations ................................................................. 178

7 Conclusion and Future Recommendations............................................................................. 178

7.1 Summary of Major Contributions ................................................................................... 178

7.1.1 Effect of processing, nano-/micro-sized additives and dissolved gas on the

phase separation and crystallization behavior of TPU ........................................ 178

7.1.2 Effect of HS crystallites on the foaming behavior of TPU ................................. 179

7.1.3 Effect of HS crystallites on the foaming behavior of TPU ................................. 180

7.1.4 Lab-scale autoclave processing of E-TPU beads and sintering with steam-

chest molding machine ....................................................................................... 181

7.2 Summary of Major Contributions (Publications) ........................................................... 182

7.3 Recommendations for Future Research .......................................................................... 183

x

List of Tables

Table 3-1 Data of DSC measurements at ambient pressure (1bar) ............................................... 76

Table 3-2 Comparison of PR-TPU’s DSC measurements at ambient pressure (1 bar) and butane

pressure (55 bar) ........................................................................................................................... 85

Table 3-3Comparison of TPU-1GMS sample DSC measurements at ambient pressure (1 bar) and

butane pressure (55 bar) ................................................................................................................ 89

Table 5-1 Experimental parameters and design variables .......................................................... 125

Table 5-2 Experimental matrix ................................................................................................... 125

Table 5-3 Melting points and crystallinity of molded EPP samples at fix and moving mold

surface at different processing conditions of pure steam and steam with hot air ....................... 138

Table 6-1 Different E-TPU beads and conditions (steam pressure/time) used to produce molded

E-TPU samples ........................................................................................................................... 163

Table 6-2 Different conditions (steam pressure/time) used to produce molded E-TPU-90A

samples ........................................................................................................................................ 163

xi

List of Figures

Figure 2-1 Microcellular foaming process ...................................................................................... 9

Figure 2-2 Schematic pressure-temperature phase diagram for a pure component showing the

supercritical fluid (SCF) region .................................................................................................... 12

Figure 2-3 Methods for the production of expandable and expanded bead foams ....................... 25

Figure 2-4 Schematic of under-water pelletization as a following unit for foam extrusion ......... 26

Figure 2-5 Bead foam processing in a steam-chest moulding machine: 1: closing and filling the

mould, 2: steaming, 3: cooling, 4: ejection of moulded part ........................................................ 27

Figure 2-6 Steps for steaming bead foams: 1: purging, 2: cross-steam, 3: autoclave steaming ... 28

Figure 2-7 Concept of the crack filling method ............................................................................ 30

Figure 2-8 Concept of the pressure filling method ....................................................................... 30

Figure 2-9 A typical double-peak melting behavior of foamed beads .......................................... 32

Figure 2-10 SEM micrograph of a cross-section of an EPP bead made with autoclave foaming

setup .............................................................................................................................................. 36

Figure 2-11 Failure mechanism: a) inter-bead, b) intra-bead ....................................................... 39

Figure 3-1 Schematic of the saturation setup with butane ............................................................ 64

Figure 3-2 Complex shear viscosity plot of AR-TPU and PR-TPU ............................................. 65

Figure 3-3 Complex shear viscosity plot of AR-TPU, PR-TPU, TPU-1GMS, TPU-1NCl and

TPU-1NSi ..................................................................................................................................... 66

Figure 3-4 Time sweep rheological curves of AR-TPU and PR-TPU .......................................... 67

Figure 3-5 Time sweep rheological curves of AR-TPU, PR-TPU, TPU-1GMS, TPU-1NCl and

TPU-1NSi ..................................................................................................................................... 67

xii

Figure 3-6 AFM images: (a) AR-TPU, (b) PR-TPU; Scale: 5 μm side length in both micrographs

....................................................................................................................................................... 69

Figure 3-7 AFM image of PR-TPU after saturating at 160°C with butane at 55 bar pressure;

Scale: 5 μm side length ................................................................................................................. 69

Figure 3-8 AFM image of TPU-1GMS; Scale: 5 μm side length in the micrograph ................... 69

Figure 3-9 DSC curves of the AR-TPU and PR-TPU samples .................................................... 71

Figure 3-10 DSC cooling curves of PR-TPU and TPU-GMS samples ........................................ 72

Figure 3-11 DSC melting curves of PR-TPU and TPU-GMS samples: (a) regular plot, (b)

magnified plot for high temperatures ............................................................................................ 72

Figure 3-12 DSC curves of PR-TPU and TPU-NSi samples: (a) exotherms, (b) endotherms ..... 73

Figure 3-13 DSC curves of PR-TPU and TPU-NCl samples: (a) exotherms, (b) endotherms ..... 73

Figure 3-14 DSC endotherms of PR-TPU after annealing at various temperatures at ambient

pressure (1 bar) ............................................................................................................................. 75

Figure 3-15 Effect of annealing at high temperature producing low temperature peak (marked

with an arrow) after cooling .......................................................................................................... 75

Figure 3-16 DSC endotherm of AR-TPU and PR-TPU after annealing at 180°C for 60 min ...... 76

Figure 3-17 DSC endotherms of PR-TPU and TPU-GMS post annealing at 180°C .................... 77

Figure 3-18 DSC endotherms of PR-TPU after annealing at different saturation temperature and

time ............................................................................................................................................... 78

Figure 3-19 Non-isothermal melt crystallization behavior of TPU at different cooling rates: (a)

ambient pressure (1 bar), (b) CO2 pressure (45 bar) .................................................................... 79

Figure 3-20 Heat of crystallization of PR-TPU samples at different CO2 pressure and cooled

from the melt with different cooling rates .................................................................................... 80

xiii

Figure 3-21 Non-isothermal melt crystallization behavior of TPU in presence of different fillers

and in the presence of CO2 pressure (45 bar) ............................................................................... 80

Figure 3-22 DSC melting endotherms after annealing over a range of CO2 pressures at a fixed

saturation temperature and time for (a) PR-70A and (b) PR-90A ................................................ 81

Figure 3-23 DSC melting endotherms after annealing at 60 bar CO2 pressure for 30 min at a

range of saturation temperatures for (a) PR-70A and (b) PR-90A ............................................... 82

Figure 3-24 DSC melting endotherms after annealing over a range of saturation times at a fixed

saturation pressure and temperature for (a) PR-70A and (b) PR-90A .......................................... 82

Figure 3-25 .(a) Comparison of DSC endotherm of AR-TPU and PR-TPU after annealing at

atmospheric pressure (w/o butane) and 55 bar butane; (b) total heat of fusion of AR-TPU and

PR-TPU after saturation with butane. ........................................................................................... 84

Figure 3-26 Comparison of DSC endotherm of AR-TPU and PR-TPU after annealing at

atmospheric pressure (w/o butane) and 55 bar butane .................................................................. 85

Figure 3-27 Tg after annealing in ambient pressure (1 bar) and in the presence of butane (55 bar).

....................................................................................................................................................... 85

Figure 3-28 Comparison of DSC melting endotherm of PR-TPU and TPU-1GMS after annealing

at ambient pressure (1bar) and in the presence of butane (55 bar) at 150°C for 60 min .............. 86

Figure 3-29 ΔHTot of PR-TPU and TPU-1GMS after annealing at ambient pressure (1bar) and in

the presence of butane (55 bar) for 60 min over a range of annealing temperature’s .................. 87

Figure 3-30 (a) Total heat of fusion (ΔHTot) of PR-TPU and TPU-1GMS over range of butane

pressure after annealing at 150°C for 60 min, (b) ΔHTm-low values of PR-TPU and TPU-1GMS

over range of butane pressure after annealing at 150°C for 60 min ............................................. 87

Figure 3-31 ΔHTm-high1 of PR-TPU and TPU-1GMS annealed under ambient pressure and butane

pressure of 55 bar over a range of annealing temperature’s for 60 min ....................................... 88

Figure 3-32 The Tm-high1 variations of PR-TPU and TPU-1GMS samples versus butane pressures

saturated at 165°C for 60 min ....................................................................................................... 88

xiv

Figure 3-33 Tg of PR-TPU and TPU-1GMS after annealing at ambient pressure and various

butane pressures ............................................................................................................................ 89

Figure 3-34 Comparison of XRD profiles of PR-TPU and TPU-1GMS ...................................... 90

Figure 3-35 Comparison of XRD profiles of TPU-1GMS annealed at ambient pressure (1bar) and

various butane pressures at a saturation temperature of 150°C .................................................... 91

Figure 3-36 SAXS profiles of TPU-1GMS samples after annealing at different pressure’s at

150°C ............................................................................................................................................ 92

Figure 4-1 Schematic of the simulation foaming setup with butane .......................................... 101

Figure 4-2 Schematic of the TPU and TPU nanocomposite foaming setup with water and CO2

..................................................................................................................................................... 102

Figure 4-3 The solubility of butane in AR-TPU and PR-TPU at 20.7 bar ................................. 103

Figure 4-4 Foam morphology of TPU prepared at 55 bar and 150ºC, 160ºC, and 165ºC: (a), (b),

and (c) AR-TPU; (d), (e), and (f) PR-TPU; Scale bars: 10 µm .................................................. 104

Figure 4-5 Foam morphology of TPU prepared at 103 bar and 150ºC, 160ºC and 165ºC: (a), (b),

and (c) AR-TPU; (d), (e), and (f) PR-TPU; Scale bars: 10 µm .................................................. 104

Figure 4-6 Characterization of AR-TPU and PR-TPU foams: (a) average cell size and (b) cell

densities ....................................................................................................................................... 106

Figure 4-7 Schematic of TPU/butane morphology displaying the possible broad HS length

distribution .................................................................................................................................. 107

Figure 4-8 Expansion ratios of AR-TPU and PR-TPU foams .................................................... 108

Figure 4-9 Foam morphology under 55 bar butane pressure at different saturation temperatures.

(a-d) PR-TPU; (e-h) TPU-05GMS; (i-l) TPU-1GMS ................................................................. 109

Figure 4-10 Cell densities of PR-TPU and TPU-GMS foams .................................................... 109

Figure 4-11 Expansion ratios of PR-TPU and TPU-GMS foams ............................................... 111

xv

Figure 4-12 Comparison of DSC melting endotherm of TPU-1NCl after annealing at ambient

pressure (1bar), in the presence of CO2 (55 bar) and in the presence of CO2 and water at 150°C

for 60 min .................................................................................................................................... 112

Figure 4-13 Foam morphology of PR-TPU prepared at 55 bar and 150°C: (a) CO2 and (b)

CO2+water .................................................................................................................................. 113

Figure 4-14 Foam morphology of TPU-1NCl prepared at 55 bar and 150°C: (a) CO2 and (b)

CO2+water .................................................................................................................................. 114

Figure 5-1 Double-peak melting behavior of EPP foamed beads ............................................... 120

Figure 5-2 A schematic of modified steam chest molding machine with hot air supply ............ 123

Figure 5-3 Rectangular area showing the location of line scans to characterize the surface

property on fixed mold and moving mold surface of molded EPP sample ................................ 126

Figure 5-4 Schematic of specimen preparation for tensile tests ................................................. 127

Figure 5-5 Effect of hot air and its flow rate on the total steaming time .................................... 128

Figure 5-6 Effect of hot air and its flow rate on the processing temperature during (a)1st

steaming cycle and (b) 2nd steaming cycle. (c) A schematic illustrating the locations where the

processing temperatures of T1 and T3 were measured. .............................................................. 130

Figure 5-7 Effect of hot air and its pressure on the processing temperature during (a) 1st

steaming and (b) 2nd steaming cycles ........................................................................................ 131

Figure 5-8 Comparison between actual line profile values measured over the scan length of EPP

parts molded with (a) pure steam and (b) steam mixed with hot air at 120 l/min ...................... 132

Figure 5-9 Effect of hot air and its flow rate on (a) Ra and (b) Rz surface roughness parameters

..................................................................................................................................................... 132

Figure 5-10 Effect of hot air and its flow rate on the waviness (Wa) values of molded EPP’s

surface ......................................................................................................................................... 133

xvi

Figure 5-11 Fixed mold surface micro-topography of EPP bead molded products using (a) pure

steam and (b) steam mixed with hot air with an air flow rate of 100 l/min ................................ 133

Figure 5-12 SEM micrographs of the cut surfaces of fixed mold surface of EPP samples

produced using steam and steam mixed with hot air at different flow rates (a) pure steam, (b) 80

l/min, and (c) 120 l/min .............................................................................................................. 134

Figure 5-13 Effect of hot air temperature on (a) Ra and (b) Rz surface roughness parameters . 135

Figure 5-14 Effect of hot air pressure on (a) Ra and (b) Rz surface roughness parameters ....... 136

Figure 5-15 Effect of hot air pressure on the waviness (Wa) values of molded EPP’s surface . 136

Figure 5-16 DSC thermographs of molded EPP samples (a) fixed mold surface and (b) moving

mold surface ................................................................................................................................ 138

Figure 5-17 Tensile strengths of molded EPP samples produced with pure steam and steam

mixed with hot air at different flow rates .................................................................................... 139

Figure 5-18 Tensile strengths of molded EPP samples produced with pure steam and steam

mixed with hot air at different temperatures ............................................................................... 140

Figure 5-19 Tensile strengths of molded EPP samples produced with pure steam and steam

mixed with hot air at different pressures ..................................................................................... 141

Figure 6-1 A schematic of autoclave bead foaming set-up ......................................................... 147

Figure 6-2 Steam-chest molding procedure ................................................................................ 151

Figure 6-3 Morphology of AR-TPU-90A beads at 55 bar CO2 pressure: (a), (b) without water;

(c), (d) with water ....................................................................................................................... 152

Figure 6-4 Morphology of AR-TPU-90A beads processed without water: (a), (b), (c) 55 bar CO2;

(d), (e), (f) 83 bar CO2 ................................................................................................................ 153

Figure 6-5 Morphology of AR-TPU-90A beads processed with water: (a), (b) 55 bar CO2; (c), (d)

83 bar CO2 ................................................................................................................................... 154

xvii

Figure 6-6 Morphology of AR-TPU-70A beads processed with CO2 pressure of 55 bar at 110°C:

(a) pressure-drop method (b) temperature-jump method ............................................................ 155

Figure 6-7 Morphology of AR-TPU-90A beads processed with CO2 pressure of 55 bar at 140°C:

(a) pressure-drop method (b) temperature-jump method ............................................................ 155

Figure 6-8 Expansion ratio of E-TPU beads produced with different methods: (a) AR-TPU-70A,

(b) AR-TPU-90A ........................................................................................................................ 156

Figure 6-9 Expansion ratio of different TPU foam beads processed with temperature-jump

method ......................................................................................................................................... 157

Figure 6-10 DSC melting curves of AR-TPU-90A after annealing at 150°C for 30 min with

different annealing conditions ..................................................................................................... 159

Figure 6-11 DSC melting curves of AR-TPU-90A bead foams processed with pressure-drop

method with water over a range of saturation temperature with 55 bar CO2 pressure .............. 160

Figure 6-12 DSC melting curves of AR-TPU-70A bead foams processed with different methods

..................................................................................................................................................... 161

Figure 6-13 Average molecular weight of the E-TPU beads processed with pressure-drop in the

presence of water: (a) AR-TPU-70A, (b) AR-TPU-90A ............................................................ 162

Figure 6-14 Actual E-TPU beads and their cellular morphologies: (a), (b) E-TPU-70A; (c), (d) E-

TPU-80A; (e), (f) E-TPU-90A .................................................................................................... 164

Figure 6-15 E-TPU-90A beads molded over range of steam pressure; (a) 1.5 bar, (b) 2 bar, (c)

2.2 bar, (d) 2.4 bar ....................................................................................................................... 165

Figure 6-16 Fractured E-TPU-90A bead foam molded part manufactured with 2.2 bar steam

pressure ....................................................................................................................................... 165

Figure 6-17 Water uptake percentage in E-TPU-90A beads over a range of temperature’s and

times ............................................................................................................................................ 167

xviii

Figure 6-18 E-TPU-90A beads soaked with water molded at 2 bar steam pressure ; (a) 50°C

water temperature, (b) 70°C water temperature .......................................................................... 167

Figure 6-19 Steam-chest molded E-TPU bead foams: (a) E-TPU-70A, (b) E-TPU-80A, (c) E-

TPU-90A ..................................................................................................................................... 168

Figure 6-20 Tensile property testing of E-TPU-70A molded sample: (a) loaded sample, (b)

fractured sample .......................................................................................................................... 168

Figure 6-21 Stress v/s strain curves of the samples: (a) E-TPU-70A, (b) E-TPU-80A .............. 169

Figure 6-22 Comparison of Young’s modulus and tensile strength of E-TPU, EPP and EPLA

molded samples: (a) Young’s modulus, (b) Tensile strength ..................................................... 170

Figure 6-23 SEM micrographs of the surfaces, the cut surfaces, and the fracture surfaces of

molded E-TPU-70A and E-TPU-80A samples ........................................................................... 172

Figure 6-24 DSC melting peak comparisons of neat-TPU, foamed E-TPU beads and molded E-

TPU beads: (a) E-TPU-70A, (b) E-TPU-80A, (c) E-TPU-90A .................................................. 175

xix

List of Symbols

CBA = Chemical blowing agents

PBA = Physical blowing agents

EPS = Expandable polystyrene

EPP = Expanded polypropylene

HFCs = Hydrofluorocarbons

PS = Polystyrene

PP = Polypropylene

Cs = Solubility of gas in the polymer (cm3/g or ggas/gpolymer)

H = Henry's law constant (cm3 [STP]/g-Pa)

ps = Saturation pressure

R = Gas constant (J/K)

Ho = Solubility coefficient constant (cm3 [STP]/g-Pa)

∆Hs = Molar heat of sorption (J)

D = Diffusivity

Do = Diffusivity coefficient constant (cm2/s)

W = Required work to generate a bubble

γpb = Surface tension

Ab = Surface area

Vb = Bubble of volume

∆ = Gibbs free energy in homogeneous nucleation

Co = Concentration of gas molecules in solution

fo = Frequency factor of gas molecules joining the nucleus

xx

k = Boltzman constant

r = Critical radius

σ = Surface tension

∆p = Pressure difference between the bubble and the melt.

N1 = Heterogeneous nucleation rate

∆ = Gibbs free energy in heterogeneous nucleation

= Surface energy of the polymer-bubble interface

∆P = Gas pressure used to diffuse the gas into the polymer

θ = Wetting angle of the polymer-additive gas interface.

dP/dt = Pressure drop rate

Psat = Saturation pressure

ρf = Foam density , g/cm3

ρ = Density of unfoamed sample, g/cm3

M = Mass of foam sample, g

V = Volume of foam sample, cm3

Φ = Volume expansion ratio,

No = Cell density

N = Number of bubbles in the micrograph

a = Area of the micrograph

M = Magnification factor of the micrograph

ΔHT = Experimental heat of fusion heat of fusion

1

Chapter 1 Introduction

1.1 Thermoplastic Foams

Thermoplastic foams consist of at least two phases: solid polymer matrix and a gaseous phase

that contributes to the formation of cells [1]. The manufactured polymer foam products possess

unique characteristics compared to their solid counterparts, such as higher specific tensile

strength, higher toughness, and superior thermal and sound insulation properties [2-6].

Additionally, polymer foamed parts are much lighter than their solid counterparts. Hence

thermoplastic foams keep stimulating manufacturers and users of foams to find new lucrative

application areas.

The main processing methods to produce thermoplastic foams are autoclave foaming [7-9],

extrusion foaming [10-16], injection foam molding [17-20], rotational molding [21-23], and

compression foam molding [24,25]. The two most popular methods are extrusion foaming and

injection foam molding due to their higher productivity. On the other hand, autoclave or the

batch foaming results in high quality foams.

1.2 Classification of Thermoplastic Foams

Generally thermoplastic foams are classified based on the cell size, the foam density and the cell

structure. Firstly, depending on cell size and cell density, thermoplastic foams are classified as

conventional foams, fine-celled foams, microcellular foams and nano-cellular foams [26]. The

foams are also classified based on the foam density as; high density foams (i.e. less than 4 times

expansion), medium density foams (i.e. between 4 and 10 times expansion), and low-density

foams (i.e. more than 10 times expansion). High density foams are usually used for construction

materials, furniture, and transportation products, whereas low-density foams are mainly used for

impact absorption, sound insulation, and packaging materials [27].To classify thermoplastic

foams based on the cell structure, they can be divided into the open-cell foams and the closed-

cell foams. The open-cell foams feature inter-connected cells. On the other hand, the closed-cell

foams have no openings in cell walls.

2

1.3 Bead Foam Technology

Foam extrusion and injection molding are the two predominated continuous processes in plastic

foam industry. In general, the process of foam extrusion allows production of two-dimensional

foam profiles of various densities and foam expansions. On the other hand, with the injection

foam molding, it is possible to fabricate foam and thin-wall foam components in complex, three-

dimensional shapes. Nevertheless, the volume expansion ratio for parts made from injection

foam molding is often limited to two to three-fold. In contrast to foam extrusion and injection

molding, the bead foaming technology is a manufacturing process which involves molding and

sintering of tiny foamed plastic beads into plastic foam components. This process can produce

three-dimensionally shaped foam products with ultra low densities. In this aspect, the bead

foaming technology is considered to be a highly promising alternative which possesses both the

foam expansion of extrusion foaming and the part geometry complexity of injection foam

molding.

The technology of bead foam molding, in general, comprises of two main steps: bead fabrication

and bead molding. There are two main approaches for manufacturing beads: batch autoclave

foaming and continuous extrusion foaming. The batch autoclave foaming approach is currently

being practiced in industry to fabricated foamed beads in batches, and bead foam products are

manufactured through a steam chest bead molding process with the foamed beads. A continuous

process which incorporates both the bead fabrication and molding processes has received great

attention from the plastic foam industry because it will introduce a cost-effective, continuous

foam process for ultra-low-density foam products with complex three-dimensional geometries.

In addition, such a cost-effective, continuous process will encourage the development of bead

foam with other polymeric materials tailored for particular applications.

1.4 Research Motivation

Although polymer bead foaming technology has provided a breakthrough in the production of

low-density foamed components with complex geometrical structure, there are only a few

polymer which have been successfully processed into expanded bead foams and their products.

One of the major issues is that every polymer beads may not fulfill the requirements of being

able to be welded into three dimensional parts using steam-chest molding machine. The sintering

technique used in expanded polypropylene (EPP) provided a promising solution for sintering

issues of polymer beads. In EPP, a double melting-peak is essential to have a balance between a

3

stable cellular structure and a proper inter-bead sintering. The low-temperature melting peak

formed during cooling as foaming occurs is used for bonding of the EPP beads. Whereas, the

high-temperature melting peak formed during the isothermal saturation step in a autoclave bead

foaming process are utilized to maintain the bead geometry even at the high temperature required

for good sintering.

The unique chemical structure of thermoplastic polyurethane (TPU) consisting of phase-

separated hard segment (HS) domains dispersed in the soft segment (SS) matrix can be

effectively utilized to develop expanded TPU (E-TPU) beads. Furthermore, it would be

necessary to investigate the desirable crystal melting structure required for a good sintering of

the E-TPU beads with steam-chest molding machine. The processing of E-TPU beads and its

three dimensional parts have a great potential to replace many important applications using

thermoset polyurethane, which are non-recyclable and are concern to the environment. The

knowledge would also help in utilizing other thermoplastic elastomeric materials for bead

foaming applications.

1.5 Objective of Thesis

The main objective of this thesis is to develop E-TPU bead foams with a desirable crystal

melting structure and foam morphology for molding with steam-chest molding machine. The

importance of achieving a desirable crystal melting peak is firstly to create a strong sintering

between the expanded beads in the molded E-TPU foam products by utilizing the crystals. The

crystals will also be beneficial to improve the foam morphology of the beads by increasing the

heterogeneous cell nucleation mechanism via the pressure variation around the existing crystals

or the crystals generated during the processing of the E-TPU bead foams. TPU are thermoplastic

elastomeric materials with a very unique crystallization behavior. It should also be noted that the

crystallization behavior of TPU is quite complicated and is significantly affected by the

processing conditions (i.e. melting and subsequent cooling from melt).

For this purpose, first of all, the crystallization behavior of TPU is extensively investigated by

varying the processing condition, by adding nano/-micron additives and in the presence of

dissolved gas at elevated pressures using regular DSC and HP-DSC.

4

Subsequently, TPU bead foams are processed in a simulation autoclave foaming chamber and in

a lab-scale bead foaming chamber. The effects of modifying the crystalline structure of TPU

during the foaming process and the parameters (i.e. saturation temperature, saturation pressure

and gas type) which affect this change are investigated in detail. Eventually, the effects of the

crystalline domains on the resultant E-TPU bead foam properties such as morphology and

thermal behavior are investigated.

Finally, the E-TPU bead foams are molded using a steam-chest molding machine and the

mechanism behind the bead-to-bead sintering for elastomeric bead foam materials is verified and

presented in detail. The tensile property of the molded E-TPU bead foam products is measured to

investigate the sintering behavior of the beads.

1.6 Organization of Thesis

This thesis is organized into 7 chapters:

Chapter 1 presents an introduction to thermoplastic foams and their classification, brief

introduction on bead foaming method is described and the motivation and objectives of the thesis

is systematically described.

Chapter 2 presents a detailed literature review and the theoretical background of the thesis topics.

The various foaming technologies are discussed and special emphasis is given to bead foaming

technology, the variety of polymeric materials commercially processed using bead foaming

technology and emerging bead foam materials. A thorough review on the crystallization behavior

of TPU is also presented.

Chapter 3 extensively present’s the effects of melt-processing, the addition of nano/-micron

additives and the presence of dissolved gas on the crystallization behavior of TPU studied using

regular DSC and HP-DSC.

Chapter 4 demonstrated the effect of crystals on the foaming behavior of TPU with different

physical blowing agents. The results from chapter 3 are correlated to the foams processed in

Chapter 4.

5

Chapter 5 shows the modifications performed to the existing steam-chest molding machine by

addition of hot air to the steam supply. This was done to reduce the sensitivity of the temperature

to the pressure variation inside the mold of the steam-chest molding machine. The effects of the

flow rate, the pressure and the temperature of the hot air on the surface roughness, thermal

properties, and mechanical properties of the molded products were studied.

Chapter 6 demonstrates the manufacturing of E-TPU bead foams with different foaming

techniques and the effect of foaming on the crystalline domains in the TPU microstructure.

The sintering of the E-TPU beads was achieved with steam-chest molding machine and the

mechanism behind the sintering was investigated. To verify the effectiveness of the sintering

between the E-TPU beads, the tensile property was measured and reported in this chapter.

Chapter 7 provides a summary of major contribution and conclusion remarks as well as the

recommendations for the future research.

1.7 References

[1] D. Klempner, and V. Sendijarevic, Handbook of Polymeric Foams and Foam Technology,

2nd Edition, Hanser Publishers (2004)

[2] D. F.Baldwin, and N. P. Suh, SPE ANTEC Tech. Papers, 38, 1503 (1992)

[3] D. I. Collias, D. G. Baird, and R. J. M. Borggreve, Polymer, 35, 3978 (1994)

[4] D. I. Collias, and D. G. Baird, Polym. Eng. Sci., 35, 1167 (1995)

[5] K. A. Seeler, and V. Kumar, Journal of Reinforced Plastics and Composites, 12, 359

(1993)

[6] L. M. Matuana, C. B. Park, and J. J. Balantinecz, Cellular Polymers, 17, 1 (1998)

[7] L. Glicksman, Notes from MIT Summer Program 4.10S, Cambridge, MA (1992)

[8] J. Reignier, J. Tatiboue¨t, and R. Gendron, Polymer, 47, 5012 (2006)

[9] M. Shimbo, D. F. Baldwin, and N. P. Suh, Polym. Eng. Sci., 35, 1387 (1995).

[10] J. H. Schut, Plastics Technology, July (2001)

[11] D. I. Collias, and D. G. Baird, R. J. M. Borggreve, Polymer, 25 3978 (1994)

6

[12] D. I. Collias, and D. G. Baird, Polym. Eng. Sci., 35, 1167 (1995)

[13] E. P. Giannelis, Adv. Mater., 8, 29 (1996)

[14] M. Okamoto, P. H. Nam, P. Maiti, T. Kotaka, N. Hasegawa, and A. Usuki, Nanoletters, 1,

295 (2001)

[15] X. Han, C. Zeng, L. J. Lee, K. W. Kurt, D. L. Tomasko, SPE ANTEC Tech. Papers, 48,

Paper #354 (2002)

[16] M. Kwak, M. Lee, and B. K. Lee, SPE ANTEC Tech. Papers, 48, Paper #381 (2002)

[17] C.A. Villamizar, C. D. Han, Polym. Eng. Sci.,18, 699 (1978)

[18] D. Maldas, B. V. Kokta, and C. Daneault, J. Vinyl. Technol., 11, 2 (1989)

[19] N. E. Zafeiropoluos, C. A. Baillie, and F. L. Matthews, Adv. Compos. Lett., 9, 291 (2000)

[20] G. Cantero, A. Arbelaiz, R. Llano-Ponte, and I. Mondragon, Comp. Sci. Techno. 63 1247

(2003)

[21] A. Arbelaiz, B. Fernandez, G. Cantero, R. Llano-Ponte, A. Valea, and I. Mondragon,

Compos. Part A, 36, 1637 (2005)

[22] P. Balasuriya, L. Ye, Y. Mai, and J. Wu, J. Appl. Polym. Sci., 83, 2505 (2002)

[23] B. V. Kokta, D. Maldas, C. Daneault, and P. Beland, Poly. Plast. Technol. Eng., 29, 87

(1990)

[24] B. N. Kokta, D. Maldas, C. Daneault, and P. Beland, J. Vinyl. Technol., 12, 146 (1990)

[25] K. L. Pickering, A. Abdalla, C. Ji, A. G. McDonald, and R.A. Franich, Composites: Part A,

34, 915 (2003)

[26] K. C. Frisch, J. H. Saunders, Plastics Foams, Marcel Dekker Inc., New York (1972)

[27] J. L. Throne, Thermoplastic Foams, Sherwood Publishers, Ohio (1996)

7

Chapter 2 Literature Review

2 Literature Review

2.1 Basic and General Principles of Foaming

2.1.1 Polymeric foams and foaming process

Polymeric foams [1, 2] are lightweight structures with a gas phase dispersed in the form of

bubbles. They have been widely used in various applications such as cushioning, insulation,

packaging and absorbency. Foams with interconnected pore structures are being studied recently

for their applications in tissue engineering as scaffolds for cell attachment and growth.

Various polymers have been used for foam applications, e.g., polyurethane (PU), polystyrene

(PS), polyolefin (polyethylene (PE) and polypropylene (PP)), poly(vinyl chloride) (PVC),

polycarbonate (PC), just name a few. In US market PU occupies the largest market share (53%)

in terms of the amount consumed, while PS is the second (26%).

Polymeric foams can be classified depending on their composition, cell morphology and physical

properties into two categories, rigid or flexible foams. Rigid foams are used in applications such

as building insulation, appliances, transportation, packaging, furniture, flotation and cushion, and

food and drink containers, whereas flexible foams are used as furniture, transportation, bedding,

carpet underlay, textile, gaskets, sports applications, shock and sound attenuation, and shoes.

Based on the size of the foam cells, polymer foams are classified as macrocellular (>100µm),

microcellular (1-100 µm), ultra-microcellular (0.1-1 µm) and nano-cellular (0.1-100nm).

Polymer foams can also be defined as either closed cell or open cell foams. A closed cell has the

foam cells isolated from each other by complete cell walls. Whereas, in open cell foams, cell

walls are broken and the structure consists of ribs and struts. Generally, closed cell foams have

lower permeability, leading to better insulation properties. Open cell foams, on the other hand,

provide better absorptive capability.

8

The foaming process consists of a system composing polymer (or monomer), blowing agents,

nucleating agent, and other necessary additives (fire retardants, surfactant, catalyst etc). Blowing

agent plays a vital role in the foam cell morphology.

Typically there are two types of blowing agents: physical blowing agents and chemical blowing

agents. Chemical blowing agents produce gases by chemical reactions or thermal decomposition

which are trapped within the polymer matrix to form foams. Physical blowing agents consists of

volatile chemicals such as chloroflurocarbons (CFCs), hydrocarbons/alcohols, and inert gases

(CO2, N2, argon). Current concerns with the ozone layer depletion has gradually reduced the use

of CFCs. Inert gases especially CO2 has become a favorable choice due to its environmentally

benign and supercritical fluid working properties.

2.1.2 Polymeric foams and foaming process

Plastic foams with cell sizes smaller than 10 µm and cell densities larger than 109 cells/cm

3 are

defined as microcellular foams [3, 4]. Nam Suh [5] was the first who proposed an idea of

introducing small bubbles in solid polymers. The rationale is that if the cell size is smaller than

the critical flaws, which already exist in the bulk polymer matrix and is generally introduced in

sufficient numbers, then the material density could be reduced while maintaining the essential

mechanical properties. Microcellular foams compared to conventional polymeric foams offer

higher impact strength, increased toughness and longer fatigue life [6, 7, 3, 8, 9, 10].

Extensive research has been carried out in this area during the past several decades. A wide

range of polymers such as PS [3, 5], PC [11], and PMMA [12, 13] have been successfully

synthesized into microcellular parts.

2.1.2.1 Microcellular foams

Microcellular foams can be produced by a batching, semi-continuous, and continuous process.

Each process mentioned has three basic steps: mixing/saturation, cell nucleation and cell growth

as shown in Figure 2.1 [14].

9

Figure 2-1 Microcellular foaming process

The batch foaming process [11, 15] of polymer materials is carried out by placing the polymer

samples in a pressurized autoclave and saturating it with the blowing agent at certain saturation

temperature and saturation pressure. If the temperature at which the polymer is saturated is

higher than the glass transition temperature, Tg, of the polymer matrix, sudden release of pressure

would result in super- saturation and cell nucleation and growth. Cell nucleation is usually fixed

by cooling the materials below its Tg. However when the saturation temperature is lower than Tg,

the cell is not able to nucleate and grow after the release of pressure even if the gas is in the

super saturation state. This is because of the glassy nature (high rigidity) of the polymer matrix.

An increase in temperature above the Tg can cause foaming. Cell structure is again fixed by

cooling. The latter method allows an independent manipulation of saturation and foaming

condition, leading to higher process flexibility. However, diffusion of the gas is inevitable while

transferring the gas-saturated material to the high temperature environment, leading to thick skin

region.

Kumar et al. [16] developed the semi-continuous foaming process. It was used to produce

polymer sheets in a solid state. In this method, a gas channeling material (gas permeable

materials) is rolled by interleaving them between layered polymer sheets. Subsequently, the roll

is saturated with the blowing agent at room temperature. Finally, the pressure is released and the

saturated polymer sheets are separated from the channeling material. The bubble nucleation and

growth is induced by pulling the sheets through a heating station.

10

The continuous extrusion foaming process is a attractive method because of its mass production

features of foamed polymers leading to high productivity, easy control, and flexible product

shaping [4, 17]. Extrusion foaming can be carried out on either a single-screw extruder or twin-

screw extruder. During the extrusion, it is better to reduce the temperature profile from the

hopper to the die. A homogeneous single-phase solution is achieved by mixing the blowing agent

into the barrel with the polymer. Cell nucleation is induced by either a rapid, large pressure drop

or a sudden temperature increase through the die. Cells will expand until the extrudate

temperature is below the glass transition temperature of the polymer. The foam shape and

expansion is controlled by a shaping die. The two distinctive characteristics of extrusion foaming

compared to a batch foaming process is that instead of saturated amount of gas a metered amount

of gas is mixed with the polymer. Secondly, the driving force for bubble nucleation is controlled

by the flow instead of the saturation pressure.

2.1.2.2 Microcellular foam properties

Many polymers have been synthesized as microcellular foams. However, very limited

development has taken to understand their mechanical properties.

A brief status on the previous research on the mechanical properties of microcellular foams can

be summarized as follows. In case of most polymers, microcellular foams exhibited superior

impact strength, toughness and fatigue life compared to solid polymers. The extent of

improvement differs among different polymers. Further, different research groups have reported

different results for the same polymer-gas system. To conclude, a direct comparison of the

mechanical properties between microcellular foams and macrocellular foams with the same

density is very limited. The review to follow is focused on the impact, tensile, and compressive

properties.

The microcellular foams prepared from PVC [18-20] and PC [21] showed an improvement in

their impact strength. A void fraction of 80% increased the impact strength of PVC foams by

four times compared to solid PVC [18]. Barlow et al. work on impact strength of PVC reported

the strength to be a strong function of both the cell density and cell size [21]. There are some

controversial results as well such as from Kumar et al. [22] reporting lowered impact strength by

introducing microcellular structure in PVC compared to that of neat PVC. The reason is yet not

clear.

11

Tensile strength and modulus of microcellular foams were studied for PC, ABS, PET and PVC

[18, 23, 24]. Though not much improvement in these two properties is seen in microcellular

foams over their bulk counterparts a marginal increase in the relative tensile strength is noticed.

It was noticed that a linear relationship exists between the tensile strength and the foam density

for the polymer systems examined. Waldman [23] reported a 400% increase in toughness of PS

foams compared to solid samples. Additionally, the tensile toughness peaked at a relative foam

density of 0.75.

Arora [25] carried out a systematic study of the compressive behavior of microcellular PS foam.

An anisotropic model was proposed to describe the effect of cell size and cell shape on the

compressive strength. It was reported that the compressive strength of PS foams increases as the

size of the cell increases. The development of a stable neck in the polymer while subjected to a

uniaxial tension correlated the phenomenon of heterogeneous, progressive buckling of the

microcellular structures. From an energy balance consideration, a model was established

describing the densification process of microcellular foams under compression.

The fatigue life characterizing the behavior of materials under repeating external forces were

studied in case of foams. PC foams with a relative density of 0.9 (10% of the weight of

reduction) showed the same fatigue life as that of the PC solid. Furthermore, the PC foams

exhibited a fatigue life one order of magnitude higher than that of solid with an increase of

relative density to 0.97[15].

2.1.3 Supercritical CO2 (scCO2) foaming

Carbon dioxide is a clean and versatile solvent for the synthesis and processing of a wide range

of materials. Supercritical (scCO2) as a processing fluid has made noticeable developments in the

past decade and have been extensively used in a variety of applications such as polymerization,

polymer fractionation and extraction, impregnation, polymer foaming and blending, surface

modification, coating and microlithography [26, 27]. A supercritical fluid (SCF) as seen in

Figure 2.2 [26] may be defined as a substance for which both temperature and pressure are both

above the critical values. Under supercritical conditions the SCF exhibits gas like diffusivity and

liquid like density with zero surface tension. The high solvation power and fast diffusion are

especially beneficial to polymer processing and there is a great deal of research in using scCO2 in

polymer processing and foaming technology. Additionally, the critical point of CO2 is relatively

12

low, 31° C and a pressure of 73.8 bar. Furthermore, CO2 is abundantly available at low cost; they

are not-toxic, non-flammable, and environmentally benign. All these advantages make ScCO2 a

promising blowing agent for polymeric foaming production.

Figure 2-2 Schematic pressure-temperature phase diagram for a pure component showing

the supercritical fluid (SCF) region

2.1.3.1 Formation of polymer/foaming agent homogeneous solutions

Formation of gas/polymer solution is one of the fundamental steps of the gas foaming process.

Solubility and diffusivity are the two important factors that describe the gas absorption behavior

into polymers. Solubility denotes the maximum concentration of the gas in the polymer which

can be described by Henry’s law as,

C=H.P (Eq 2.1)

where, P is the pressure, C is solubility of gas in the polymer and H is the Henry constant, which

is dependent on the temperature. While diffusivity denotes how fast the gas can enter or disperse

out of the polymer. The diffusivity can be described by Arrhenius relationship as,

RT

EDD aexp0 (Eq 2.2)

Where D is the diffusivity, D0 is the diffusion constant, Ea is the activation energy for diffusion

of a gas in a polymer, R is the gas constant, and T is the absolute temperature. An ideal foaming

13

condition is a condition with higher gas solubility in a polymer assisting in greater cell

nucleation and growth. A higher diffusivity is sought in this step because of a shorten saturation

time and better productivity. However, this may not assist in cell growth, discussed latter.

Both the solubility and diffusivity are highly dependent on the pressure and temperature. A

lower temperature generally results in a higher solubility, a highly desirable situation. However,

a decreased processing temperature decreases the diffusivity of gas in polymer reducing the

productivity. In order to improve the productivity, a higher gas pressure is usually used thereby

increasing the solubility. Wissinger et al [29, 30] and Zhang [31] reported that in a PS-CO2

system there is a linear relationship between the solubility and the saturation pressure (Henry’s

law). Similar results were noticed in the PP-CO2 system. Handa et al [32] investigated the

solubility of CO2 in PMMA over a wide range of temperature (0-1670C) and pressure up to 61

atm. They reported that the linear relationship between the solubility and pressure only exists at

high temperature regions. However, at lower temperature, the solubility was convex towards the

pressure.

Gas solubility being affected by various other factors has been reported in recent research

studies. Effect of nanoclay on the kinetics of CO2 gas in PMMA was studied by A.Manninen et

al. [33]. It was reported that diffusivity increased with a higher nanoclay concentration while the

solubility remained unchanged by the presence of nanoclay. Handa et al. [13, 32] reported that

the diffusivity of highly pressurized CO2 in PMMA at a lower temperature may be higher than

that at a higher temperature because of the shifting of the glass transition temperature (Tg). The

change in crystallinity of semi-crystalline polymer was also found to change the solubility of gas

in a polymer [34].

Polymers with electron donor groups such as ether, fluro, and carbonyl groups, usually exhibits a

higher solubility of CO2. Kazarian et al [35] have shown that CO2 can participate in Lewis acid-

base type interactions with polymers containing electron-donating groups such as carbonyls. In

this case, CO2 is considered as Lewis acid and the polymer with those functional groups as the

Lewis base.

14

2.1.3.2 Cell nucleation

Formation of a gas/polymer solution is followed by a rapid drop in pressure and/or a increase in

temperature. According to the Henry’s law the solubility decreases during this process. The

resulting over saturation induces large number of cell nucleation’s because the gas tends to

escape out of the polymer matrix. The morphology of the final foam product is determined by

the cell formation in a polymer and hence cell nucleation is of great importance in the foaming

process.

Classical nucleation [36] theory is commonly adopted to explain the nucleation process. The

theory classifies the cell nucleation into two different types: homogeneous nucleation and

heterogeneous nucleation. Homogeneous nucleation occurs in a pure gas/polymer solution. There

are no additional impurities added to the solution. The rate of homogeneous nucleation is

expressed as,

)/exp( *

hom00hom kTGCfN (Eq 2.3)

where, is the frequency factor for homogeneous nucleation a function of both the surface

tension and the mass of the gas molecule, is the concentration of gas molecules, is the

free energy required for the homogeneous nucleation to form a nucleus with critical size, is the

Boltzmann’s constant, T is the temperature in Kelvin. The critical nucleation energy is expressed

as,

3

2

*

hom)(3

16bp

PG (Eq 2.4)

and the corresponding critical bubble size is,

Pr

2* (Eq 2.5)

15

Here, is the liquid-gas surface tension, and ΔP is the pressure difference between that of the

inside of critical nuclei and the surrounding liquid. Assuming the polymer is fully saturated at by

the blowing agent and the partial molar volume of blowing agent in the polymer is zero, ΔP can

be taken as the saturation pressure.

In the presence of nucleating agents, heterogeneous nucleation takes place in the polymer matrix.

It occurs at the interface between the polymer/gas solution and the nucleants. The heterogeneous

nucleation rate is given by [37]:

)/exp( *

11 kTGCfN hethet (Eq 2.6)

Where, is the frequency factor, is the concentration of the heterogeneous nucleation sites,

which can be related to the particle concentration. The term is given by,

)()(3

16 3

2

* fP

G bphet (Eq 2.7)

where, is the surface energy of the polymer, ΔP is gas saturation pressure, and is wetting

angle geometric factor.

The homogeneous and heterogeneous nucleation’s are not different from each other. The mixed

model describes the nucleation by,

hetNNN '

hom (Eq 2.8)

where, N is the combined nucleation rate of both homogeneous and heterogeneous nucleation’s,

is the modified homogeneous nucleation rate, and is the heterogeneous nucleation

rate. Modified homogeneous nucleation rate can by given by,

16

kT

GCfN hom'

00

'

hom exp (Eq 2.9)

where is the concentration of gas molecules in solution after heterogeneous nucleation has

occurred.

The free energy required for heterogeneous nucleation is generally much lower that required for

homogeneous nucleation. Therefore, additives such as talc, nano-clay or nanotubes can decrease

the energy required to create bubbles and therefore promote the cell nucleation. However there

are certain criteria to be fulfilled for being an ideal nucleant [38]. Three of the most important

criterion are: first, highest nucleation efficiency can only be achieved when the nucleation on the

nucleant surface is energetically favored and is relative to homogeneous and heterogeneous

nucleation; secondly, ideal nucleants have uniform size and surface properties; thirdly, ideal

nucleants are easily dispersible.

2.1.3.3 Cell growth and stabilization

The process of cell growth involves mass, momentum and heat transfer of the fluid. The models

describing the cell growth evolve from a basic model [39] used to describe the cell growth from

a single bubble that is surrounded by an infinite sea of fluid with an infinite amount of available

gas.

Cells come too close to each other as they grow. A solid wall of polymer separates the gaseous

phase. The increased pressure inside the bubbles stretches the cell walls to become thinner. Ones

the pressure inside a cell is high enough it ruptures the cell wall and two adjacent bubble

becomes a single large bubble. This transformation is referred to as cell coalescence [40]. Cell

coalescence adversely affects the cell sizes and hence should be avoided. Decreasing the

flexibility of the polymer by cooling down the polymer is common way to prevent cell

coalescence. A drop in temperature below the glass transition temperature (Tg) or the

crystallization temperature (Tc) fixes the foam morphology.

17

2.2 Extrusion Foaming Technology

Extrusion foaming possesses an important feature that the polymer foams made are

manufactured in a continuous process contrary to batch foaming and also has a higher

productivity. Both CBA and PBA can be used for extrusion foaming depends on the material and

the desired product properties. PBA-based processing is not limited by decomposition

temperatures and can therefore be processed below critical temperatures. In addition, it induces

less cost and produces better cell morphology.

Continuous extrusion foaming with a PBA involves a few basic steps: firstly, there is a uniform

formation of a polymer /gas solution, secondly there is cell nucleation followed by cell growth

and timely solidification of the polymer melt. A rapid-pressure-drop nucleation die [41] is where

cell nucleation occurs. Setting the polymer/gas solution to a thermodynamic instability can

generate large number of bubble nuclei inside the polymer melt. Thermodynamic instability

itself is induced by reducing the solubility of gas in the solution and by creating a rapid pressure

drop that results in the nucleation of numerous microcells. Cell nucleation directly influences the

number of cells. Directly influencing the number of cells generated in the polymer makes cell

nucleation a critical step. Cell after nucleation continue to grow even while exiting the mold and

it only stops when the dissolved gas is consumed or when the part is cooled to become stiff. Cell

coalescence and cell collapse are very critical issues in cell growth. Park and Behravesh [42] has

developed effective methods to prevent cell coalescence and gas escape during cell growth. Cell

coalescence can be suppressed by cooling the polymer/gas solution homogeneously, which

increases the melt strength. Whereas, gas escape can be controlled by cooling the surface of the

extrudate to form a solid skin layer, thereby, blocking the gas from escaping from the polymer.

When the polymer melt is extruded out of the die and its temperature decreases, it will solidify

through classification or crystallization. Timely solidification is important, for a delayed

solidification may result in gas loss, whereas solidification that is too fast will not produce a

desired volume expansion ratio ( or density reduction) [43].

The geometries of the filamentary dies, i.e. the die diameter and the dies length induce different

die pressures and different pressure drop rates, and consequently, different final foam structures.

Xu et al. [44] designed three interchangeable groups of 9 dies with the same pressure or the same

18

pressure drop rate. They assumed that the polymer melt was described by a “ power law model”

and generated the theoretical equations to calculate die pressure and pressure drop rate [45].

Naguib et al. carefully analyzed experimental results of extrusion foaming at various processing

conditions. They concluded that the final volume expansion ratio of extruded PP foams blown

with n-butane was governed either by the loss of the blowing agent through the foam skin or the

crystallization of polymer matrix [46].

The diffusivity of blowing agents at elevated temperatures is very high. Therefore, gas can easily

escape from the extruded foam because of its higher diffusivity at elevated temperatures. In

addition, as the cell expansion increases, the thickness of the cell wall decreases and the resulting

rate of gas diffusion between cells increases. Consequently, the rate of gas escape from the foam

to the environment increases. Gas escape through the thin cell walls decreases the amount of gas

that is available for the growth of cells, resulting in lowered expansion. Moreover, if the cells do

not solidify quickly enough, they tend to shrink due to loss of gas through the foam skin, causing

overall foam contraction. This indicates that the gas loss phenomena are a dominant factor that

constrains the volume expansion when the melt temperature is high.

Another critical factor that affects the maximum expansion ratio in plastic foam processing is the

crystallization behavior of semi-crystalline materials. Semi-crystalline polymer melt, such as PP,

solidifies at the moment of crystallization during cooling. Therefore, the foam structure solidifies

at the crystallization temperature during the foaming process. If the crystallization occurs in the

primitive stage of foaming, i.e., before the dissolved blowing agent fully diffuses out of the

plastic matrix and into the nucleated cells, then the foam cannot fully expand. Therefore, in order

to achieve the maximum volume expansion ratio, the crystallization (or solidification) should not

occur before all of the dissolved gas diffuses out into the nucleated cells. Upon exiting the die,

the temperature of melt decreases due to external cooling outside the die and the cooling effect

which is attributed to isentropic expansion of the blowing gases Thus, the processing temperature

at the die determines the time for the solidifying of the polymer melt. Therefore, in order to

provide adequate time for the gas to diffuse into the polymer matrix, the processing temperature

should be sufficiently high. It should be noted that if the processing temperature is too close to

the crystallization temperature, the polymer melt would solidify too quickly before the foam has

expanded fully.

19

This indicates that there is an optimum processing temperature for achieving maximum

expansion. If the melt temperature is too high, then the maximum volume expansion ratio is

governed by gas loss and it will increase as the processing temperature decreases. If the melt

temperature is too high, then the maximum volume expansion ratio is governed by gas loss and it

will increase as the processing temperature decreases. If the melt temperature is too low, then the

volume expansion ratio is governed by the solidification (i.e., the crystallization ) behavior and it

will increase as the temperature increases.

2.3 Injection Foam Molding Technology

2.3.1 Conventional foam injection molding and microcellular injection molding technologies

Foam Injection Molding (FIM) technology, one of many conventional injection molding

processes, is also like other thermoplastic foam manufacturing technologies. Polymer is melted,

mixed with a gas blowing agent, and injected into a mold through a shut off nozzle. The large

pressure differentiation between the melting chamber and the mold would induce a significant

pressure drop in the polymer since the mold is not pressurized. The injected material foams

during the pressure drop and expands in volume to fill the mold.

The FIM technology produces a number of advantages compared to other methods. It reduces the

material cost, the parts weight, the molding cycle time, the residual stress, the viscosity and the

processing temperature. Other advantages include the elimination of surface sink marks on the

parts, enhanced dimensional stability, high stiffness-to-weight ratio, and minimized fiber-type

fillers damages.[47]

In the 1980s, Dr. Suh and his students at the Massachusetts Institute of Technology developed a

microcellular plastic to reduce material usage and increase material stiffness by crack arrestors

formed by tiny bubble. Contrary to other researches, the cell diameter for this plastic is around 5

to 50 μm and the cell density is higher than 106[48]. The majority of cells also must be closed

cells with less amount of weight reduction. The team focused on the microcellular structures

development then moved on to continuous polymer manufacturing processes. Other researchers

also participated in developing the microcellular injection molding process and the

manufacturing equipment. Trexel Inc. is the company responsible in cooperating with the MIT

20

research team to develop the MuCellⓇ system that injects PBA such as N2 and CO2. FIM is now

also generally known as “microcellular injection molding”.

Microcellular injection molding differs from commercial FIM by the fact that it fills the mold

without foaming. It injects a full shot of polymer material into the mold to account for the

volume shrinkage from cooling. In contrast, low pressure FIM often achieves a high void

fraction. The microcellular process utilizes only PBA to create foam structure instead of the PBA

or CBA that FIM uses. The process can create thin-walled products due to the gas present in the

polymer matrix. The injection molding process provides the same dimensional stability because

volume expansion counter acts shrinkage and warpage in cooling, The packing and holding

phases is no longer necessary thus cycle time is reduced. Disadvantages are shared between FIM

and microcellular injection molding. The mold parts have poor surface finish, and there are

limited applications for a nontransparent part. The process also requires a strictly balanced

runner system, complicated design and technology, and requires a significant amount of

investment. [48].

2.3.2 Low-pressure and high-pressure foam injection molding technologies

2.3.2.1 Low-pressure FIM

FIM is classified as low pressure by two major factors, the relative cavity pressure between 0.5

to 10 MPa, and smaller injection shot size from 65 to 80% of the full shot volume. Low pressure

is achieved by the small shot size and can also lower tonnage in the molding machine. Expansion

in this method fills most of the cavity volume in the mold and is therefore very suitable for thick-

walled products. Low pressure FIM also reduces residual stress, lowers cost, and increases

dimension stability. Products that perform simple functions can be made effectively using this

process.

Flow swirl marks on the product surface and non-uniform cellular morphology are two defects

that have to be overcome. Swirl marks are created when the cells are nucleated and squeezed

onto the mold surface. The nucleated cells have to travel all the way to fill up the mold. Cell

growth can be excessive during this travel and cause significant cell coalescence. The low

injection pressure limits the minimum thickness of parts made. It is because when the thickness

21

is small, the polymer cools faster and would encounter a higher resistance during the flow and it

is hard to overcome when the injection pressure is low. Generally parts with thickness of less

than 0.25’’ is not considered in this FIM process.

2.3.2.2 High-pressure FIM

High pressure FIM is classified by the fast completion of the mold cavity filling, the

disconnection of the core foaming and solid skin layer formation, and the core foaming due to

cooling polymer shrinkage. In high pressure FIM, the mold cavity is filled completely with

polymer quickly. When the polymer shrinks during the cooling stage, volume expansion again

fills the free volume in the mold. Uniform cell structures are created with this process since

foaming occurs after the cavity is completely filled without much cell movement. Swirl mark

effect is reduced due to the fast filling and increases the surface quality. The volume expansion

from foaming is minimum with this process which requires more cost in material.

2.3.2.3 Investigation of foaming behaviors in foam injection molding using mold pressure profile

Different experiments are done to explore the foaming mechanism of the FIM process. However,

it is difficult to control the mechanism as various factors contribute to the outcome. One

researcher, Lee, experimented FIM with mold pressure profiles [47, 49]. Pressure is measured in

three different locations, corresponding Location A, Location B, Location C, within the plaque-

shaped mold with multiple fan gates. A comparison is made between the results of Lee’s FIM

experiment and foam extrusion with the assumption that when the system pressure is

significantly lower than the solubility pressure, cell nucleation occurs. From foam extrusion, the

gas-added polymer part leaves the mold in steady state and the cell grows also in steady state.

From FIM, however, the gas-added polymer flow in the mold experiences different pressures at

different parts of the mold. The degree of injection of the polymer/gas mixture varies and affects

cell nucleation. While the gas pressure drop also changes according to the filling of the mold and

the pressure of the flow front.

Most experiment parameters in the two processes were maintained the same to extract a fair

comparison. Only the pressure drop and pressure drop rates are different in the two technologies.

The polymer materials used in the two processes are polypropylene (PP), thermoplastic poly-

olefin (TPO), and high density polyethylene (HDPE). N2 acts as the PBA for FIM as well as

22

foam extrusion. Pressure drop and pressure drop rate conditions for the foam extrusion process

are derived from cell densities measured from products of different trails of varying processing

parameter. As for the FIM, cell density values are measured at the 3 locations where pressures

were measured. Pressure drop values give an assumption that the larger the pressure drop would

determine the final cell density.. Measured cell density values from FIM closely correspond to

the estimated values from foam extrusion. Eventually, the foam extrusion data and the mold

cavity pressure profile can be used to estimate the cell density values and foam structure in the

FIM process. In conclusion, specifically desired cell density values from the FIM process can be

achieved by varying certain processing parameters.

2.4 Rotational Foam Molding Technology

Rotational foam molding is an example of chemical foaming processes. The technique is evolved

from the conventional rotational molding process. Conventional rotational molding process is

widely used in the plastic industry to manufacture storage tanks, furniture, playground

equipment, toys, and components for aircrafts and automobiles [50-52]. The process is believed

to be developed in the late 1930s to early 1940s along with the development of highly plasticized

liquid polyvinyl chloride, a thermoplastic alternative to latex rubber. At that period of time,

rotational molding was mainly used to produce toys such as squeezable toy dolls and beach balls.

During the World War II, the process was utilized to produce items such as syringe bulbs,

squeezable bottles, bladders and air-filled cushions. During these early years, plastic parts were

rotational molded inside a hollow metal mold over an open flame. With the introduction of

rotational –molding-graded polyethylene powders and hot air ovens in the 1950s, the process

advanced rapidly and more and more types of hollow plastic products could be manufactured

from the process [53].

Rotational molding required low equipment and mold cost and has relatively low waste [53, 54].

It produces hollow parts that are low in residual stressed [53]. It is also capable of manufacturing

parts of complex geometries, sizes, and variable thicknesses and layers [50]. Because of these

advantages, rotational molding has been expanding at a rate of 10 to 15% per annum over the last

few decades [55]. A possible drawback to utilizing rotational molding process would be the

material requirements of the process. Materials suitable for rotational molding are relatively

expensive due to the need of special additives and fine powder sizes. Low-density polyethylene

23

(LLDPE)and high-density polyethylene (HDPE) are widely used in rotational molding. They

represent about 85%of all polymers consumed in the rotational molding industry. The reasons for

that are their modest melting temperatures and their ability to sinter together under low-shear

conditions [53]. Although polyethylene materials are easy to rotational mold, their relatively low

stiffness and strength make them less favorable for engineering applications [54]. Other

polymeric materials used in liquid polymers [53]. Polypropylenes-based materials are selected

for the experiments in this thesis project because of their better mechanical properties over

polyethylene. Until the last decade, very few studies have been conducted for developing plastic

foams from this technique. Due to the nature of the rotational molding process, physical blowing

agents such as high-pressure gases are not applicable to create the gaseous phase in the material

matrix [56, 57]. Needham attempted to produce plastic foams from a rotational molding process

by introducing a chemical blowing agent (CBA), mainly sodium bicarbonate into the material

mixture [58]. The Uniroyal Chemical Co. Inc., in 1996, successfully synthesized LDPE foams

from rotational molding using cologne OT as the CBA. The company reported that the foam

density decreased and the wall thickness increased as a result of an increased concentration of

Celogen OT used during the foaming process. They also suggested that the cooling step in the

foaming cycle determined the quality for the resulting foam [50]. Liu et al. and Pop-Iliev et al.

investigated the processing of polyethylene and polypropylene foams and the effect of the

blowing agent and processing temperature on the resulting cell microstructure [50, 56, 59]. Fine-

celled foams of three-fold to six-fold expansion could be made from the rotational foam molding

technique [50, 56, 59].

In general, the process of rotational foam molding can be summarized into four main steps [53,

60]:

i) Charging the Mold with Materials: A predetermined amount of polymer powders and CBA

particles mixture will be charged into the mold. The mold is then set to rotate uni-axially and is

heated simultaneously inside an oven at the desired temperature.

ii) Polymer Powder Sintering: The polymer powders begin to melt and sinter together. Due to the

temperature gradient along the radial direction of the mold, powders on the mold surface will

sinter first. As the heating process continues, all the powders in the mold will eventually sinter

together forming a continuous polymer matrix.

24

iii) Decomposition of CBA and Foaming: As the temperature of the polymer melt inside the

mold rises to certain point, the CBA particles in the melt will start to decompose and liberate

gases creating bubbles or pores inside the polymer matrix.

iv) Cooling: While the mold is still in rotation, it is being cooled by air or water jets upon the

completion of the heating cycle. The polymer melt inside the mold begins to cool and solidify

starting from the mold surface towards the center of the mold. At the end of the cooling cycle,

the solidified polymer part will be released from the mold.

As suggested by Pop-Iliev et al., the sintering temperature of the polymer should be lower than

the decomposition temperature of CBA, the foaming temperature, and the coalescent

temperature. The reason for that is to allow the polymer to sinter into a continuous phase for

improved cell quality. It is also important that the molten polymer has to flow and wet to the

blowing agents well in order to eliminate undesired air bubbles encapsulation, which impairs the

cell morphology. The zero-shear viscosity is also a key parameter in the sintering of the polymer

matrix and the resulting cell morphology [56, 59, 60].

2.5 Bead Foam Molding Technology

Foam extrusion and injection molding are the two predominat continuous processes in plastic

foam industry. In general, the process of foam extrusion allows production of two-dimensional

foam profiles of various densities and foam expansions. On the other hand, with the injection

foam molding, it is possible to fabricate foam and thin-wall foam components in complex, three-

dimensional shapes. Nevertheless, the volume expansion ratio for parts made from injection

foam molding is often limited to two to three-fold. In contrast to foam extrusion and injection

molding, the bead foaming technology is a manufacturing process which involves molding and

sintering of tiny foamed plastic beads into plastic foam components. This process can produce

three-dimensionally shaped foam products with ultra low densities. In this aspect, the bead

foaming technology is considered to be a highly promising alternative which possesses both the

foam expansion of extrusion foaming and the part geometry complexity of injection foam

molding [61-63].

The technology of bead foam molding, in general, comprises of two main steps: bead fabrication

and bead molding.

25

2.5.1 Bead fabrication

In principle, two approaches for the production of foamed bead exist. The first approach consists

of the production of expandable beads, which can be applied for amorphous thermoplastic resins

like polystyrene (product: EPS – expandable PS). Expandable beads are granules in which a

blowing agent (e.g. pentane) is trapped and are expanded in a second step before the welding-

process, the so-called pre-expansion. Efficient transportation of the unfoamed material and

control of density by the part-manufacturer are clear advantages compared to expanded beads,

which is the second type of beads.

Expanded beads are produced from semi-crystalline thermoplastics, since the presence of

crystalline domains prevents the storage of a blowing agent inside the solid bead [64]. Expanded

polypropylene (EPP) is produced in that way. An overview of the possible methods is given in

Fig 2.3.

Figure 2-3 Methods for the production of expandable and expanded bead foams

The most commonly used method to produce huge quantities of expandable beads of polystyrene

is the suspension-polymerization with a blowing agent. In that process, the polymerization

happens at high pressure in the presence of pentane, which leads to the incorporation of the

blowing agent inside the granules. Problems arise with additivation, since the additives are

required to be fully soluble in water to be stored in the final bead. Another drawback is the base

material itself, since not all polymers can be synthesized via suspension-polymerization.

A method to produce expandable or expanded beads is the impregnation (loading with blowing

agent) of micro-granules, which contain all required additives, with the blowing agent in an

autoclave. This is the main production process for EPP. In a first impregnation vessel the solid

PP-beads are saturated with gas at around 150 °C and then released to an expansion vessel.

26

Afterwards the beads are washed to remove any residual suspension stabilizer, which would

inhibit proper welding of the beads during steam-chest moulding [65]. For amorphous polymers,

expandable beads can be produced as well, if the saturation step takes place at temperature below

the glass-transition of the polymer-blowing agent-solution.

Alternatively, foam extrusion with under-water pelletizing allows the production of expandable

beads or already expanded beads. It is schematically shown in Fig. 2.4. In this method, gas-

loaded polymer melt is extruded through a hole-plate into a water-stream and cut by rotating

knifes. If the water-pressure is above the vapor pressure of the blowing agent, the blowing agent

is trapped within the solidifying polymer and expandable beads are produced. At low pressure,

the dissolved gas evaporates and forms bubbles resulting in expanded beads. Advantages of this

method are the exact dosing of the blowing agent(s) into the melt and a continuous and flexible

process, which allows the processing of any thermoplastic resin with additives. Variable process

parameters are temperature and pressure of the water, the rotational speed of the knives and the

temperature of the perforated plate.

Figure 2-4 Schematic of under-water pelletization as a following unit for foam extrusion

2.5.2 Bead bonding

The parts from bead foams are made in a complex, yet efficient process, which allows the

production of parts with a high geometrical degree of freedom at very low density. For the

production of parts, the expanded or expandable beads are welded together in a steam-chest

moulding machine. Therefore, the surface of the beads is partially molten or softened, which

27

leads to the inter-diffusion of chains between different beads and thereby good cohesion. Good

cohesion between the beads is necessary to ensure favorable mechanical properties [66,67].

The processing of bead foams to a part is done in a steam-chest-moulding machine in five steps.

The steps are shown on Fig. 2.5and will be explained below.

Figure 2-5 Bead foam processing in a steam-chest moulding machine: 1: closing and filling

the mould, 2: steaming, 3: cooling, 4: ejection of moulded part

1. Filling of the mould

At first, foamed beads are sucked out of a container blown into the mould by an injector, which

usually functions according to the venturi-principle. This step is critical to achieve a homogenous

distribution of the beads inside the mould.

2. Welding of the beads

After the filling process, the beads are fused together by hot steam flowing through the mould.

During steaming, the beads form physical links due to inter-diffusion of chains of neighboring

beads. To ensure high welding quality elevated temperature and a sufficient steaming time are

necessary as well as a high contact area and force between the beads. With a low contact area,

28

force is transferred only at a few points, which leads to bad mechanical properties. If the contact

force is low, the beads might not touch sufficiently thus also leading to bad welding.

For EPP, the steam has an inlet pressure between 7 and 8 bar. However, the pressure inside the

mould is lower - pressures between 2.5 and 4 bar are common [68]. Thus, a steam-temperature

up to 150 °C is achieved. In case of EPP or other semi-crystalline polymers, a part of the crystals

is molten or in the case of EPS the polymer is softened.

The steaming process consists of three steps, which are shown on Fig. 2.6. At first, the air

between the beads is purged out and the mould is pre-heated. Therefore, steam is flowing parallel

to the mould (Fig. 2.6 - 1). Secondly, the steam flows through the mould (Fig. 2.6- 2). To ensure

a temperature distribution as homogeneous as possible and thereby to ensure constant welding

quality in the whole part, the mould is steamed from both sides. This is called cross steaming.

Finally, steam is guided along the mould to improve surface quality (autoclave steaming, Fig.

2.6- 3).

Figure 2-6 Steps for steaming bead foams: 1: purging, 2: cross-steam, 3: autoclave

steaming

In EPP, a double melting-peak is essential to have a balance between a stable cellular structure

(which requires crystallinity) and proper inter-bead welding must be maintained. Therefore, the

lower melting peak ensures good bonding and the upper one keeps the structure stable and

29

prevents the collapse of cells. The creation of the double melting peak will be discussed in detail

in the chapter about EPP.

3. Cooling and stabilization

For dimensional stability of the part, cooling of the mould is a crucial step. If the part is ejected

without cooling, further expansion of the beads is possible, which leads to a deviation of the

original size. For cooling the mould is sprayed with water until a temperature of around 80 °C is

reached.

4. Ejection of the moulded part

After moulding and cooling, the part is finally ejected. Pressurized air and mechanical ejectors

are used to eject the part.

5. Post-processing of the final part

Especially at low density, shrinkage can be a challenge. For example Neopolen P can have a

shrinkage up to 2.8 % [69] (Neopolen P 8220 K, BASF SE, density 22 g/l), which comes from

the condensation of steam inside the beads that leads to a vacuum. For components requiring

high dimensional accuracy, a tempering step of the parts is necessary. For Neopolen P a

temperature of 80°C is recommended. In this step, the original shape is restored, at least

partially. Furthermore, condensed water from the steaming step is removed as well.

In principle EPP is processed in two different ways, namely the crack filling process or the

pressure filling process. Both can be combined with the so-called pre-loading step. Those

processes will be explained in this section.

In contrast to EPS, which still contains a certain amount of blowing agent, EPP-beads do not

expand any further inside the mould without special treatment. Therefore, this matter must be

dealt with process-wise.

At first, the crack-filling method will be explained. Its concept is shown on Fig. 2.7. With this

method the beads are filled into a compression-mould at ambient pressure. Before the steaming-

process, the mould is closed to its final dimensions, so that the beads are compressed. With this

technique very thin parts with a thickness even below the bead thickness can be realized. The

30

drawback of this method is an inhomogeneous density distribution, if the wall thickness is not

constant, and limitations in the part shape. An example for the application of this method is the

sun-visor in the automotive industry.

Figure 2-7 Concept of the crack filling method

Alternatively, the pressure filling method can be used, which is depicted in Fig. 2.8. Therefore,

the beads are subjected to an elevated pressure during the filling process, which leads to a

compression of the beads. After filling, the pressure is released and the beads re-expand thus

reducing macro-porosity. According to the level of filling-pressure, the compression of the beads

and thereby final density of the part can be controlled. For EPP usually filling-pressures between

1.5 and 3.5 bar are applied.

Figure 2-8 Concept of the pressure filling method

With the above-mentioned processing method only moderate densities can be achieved. To lower

the density those moulding methods must be combined with pressure pre-loading. Before the

actual moulding, the beads are subjected to pressurized hot air for several hours until the inside

pressure of the beads reaches equilibrium with the outside. The so captured air leads to additional

expansion during steaming thus allowing lower densities. Furthermore, pressure pre-loading

reduces macro-voids between the beads, which leads to better mechanical properties.

31

The working mechanism of steam-chest moulding is similar to a sintering process. The major

difference between the two processes is that the former uses high temperature steam as an

effective heating/cooling medium [70,71], while the latter normally uses hot air [72–74]. During

the steam-chest moulding process [75,76], high temperature steam is injected into the mould in

three different cycles to soften and fuse the beads. The steam vaporizes the volatile gas present in

the beads and hence causes the expansion in volume, as well as re-blowing of the beads. Through

this process, the empty space in is filled at the same time when the inter-bead fusion is created,

which results in the forming. To improve bead foaming technologies and bead-moulded

products, many researchers performed mechanical property tests of bead-moulded products

based on commercialized beads such as expanded polystyrene (EPS) and expanded

polypropylene (EPP). The formation of inter-bead bonding in EPS beads involves the diffusion

of polymer chains across the inter-bead regions during the heating process of steam-chest

moulding process. Whereas, the cooling cycle freezes the physical entanglement of the polymer

chains at the inter-bead boundaries and results in the bonding of the EPS bead foams. The steam

temperature and moulding time are two critical parameters affecting the extent of bead fusion

and significantly affects the overall mechanical properties of the moulded bead foam samples

[71,77–79]. However, if the foamed beads are steamed for too long, their cell structure might

collapse and deteriorate the surface property of the moulded product [80].

In the case of moulding of EPP beads with steam-chest moulding machine, good sintering

requires a desirable double crystal melting peak structure as shown in Fig. 2.9. The hatched area

in Fig. 2.9 represents the desirable steam temperature range between the low and high melting

peaks (Tm-low and Tm-high) of EPP beads within the steam-chest moulding machine [81–88]. When

EPP beads are processed in the steam-chest moulding machine, crystals associated with Tm-low

melt and contribute to the fusing and sintering of individual beads. The unmolten Tm-high crystals

help to preserve the overall cellular morphology and dimensional stability of the moulded EPP

product. A very narrow processing window between the two melting peaks poses a significant

challenge in setting the processing steam temperature during the molding process in steam-chest

moulding machine. A slight variation in steam temperature may cause the Tm-high crystals to get

affected and destroy the cellular morphology of the EPP beads and cause shrinkage of the

moulded EPP product. The steam can penetrate into the EPP beads during the steam-chest

moulding process. During the cooling cycle, at the end of the moulding process, the high-

32

temperature steam, which diffused into the beads, tends to condense in the cells and leads to a

negative pressure. Due to the characteristics closed cell structure of EPP beads, air cannot

penetrate into the foam within a short span of time, which results in a dramatic decrease in the

internal pressure of the foams. Consequently moulded EPP parts tend to shrink after completion

of the moulding process. An annealing process is generally used at a high temperature to enhance

the diffusion rates of steam and air and thus prevent shrinkage [67].

30 60 90 120 150 180 210-1.4

-1.2

-1.0

-0.8

-0.6

-0.4

-0.2 actual variation of steam

temperature

Tm-high

Tm-low

He

at

Flo

w (

W/g

)

Temperature (°C)

Endo

Figure 2-9 A typical double-peak melting behavior of foamed beads

The processing steam in a steam-chest moulding machine is in the superheated state and its

temperature is coupled with the processing pressure [89] according to the vapour pressure curve.

However, as the steam enters the mould cavity via small ports, the overall pressure starts

decreasing due to condensation of the steam on the beads. Furthermore, the pressure of the steam

decreases because of the resistance of the flow through the beads, which subsequently reduces

the temperature and makes it difficult to determine the actual temperature of the mould.

Moreover, considering the large volume and complicated shape of the mould cavity the

temperature distribution and thereby also density is not uniform. Hence the optimum processing

condition required for the desired properties in the moulded bead foam products can be achieved

by trial and error. Nakai et al. [84] investigated some fundamental aspects of steam-chest

moulding, such as the evaporation and condensation of steam and heat conduction, using

numerical simulation techniques. They reported reduced heat conduction to the core area of the

33

mould caused by a decrease in steam temperature as a result of drop in the steam pressure.

Generally, higher operating steam pressure is implemented to improve the heat conduction to the

core area of the mould. However, a higher operating steam pressure relates to higher operating

cost and a higher temperature leading to an increase in localized temperature near the steam

entry, and hence beads exposed to this high temperature may melt resulting in shrinkage at the

surface of the product. This dramatically deteriorates the surface property of the moulded

product.

2.5.3 Bead foam materials

2.5.3.1 Commercial bead foam materials

EPS is the most widely used bead foam material with a consumption of 4.7 Mt per year [90] due

to its low price and high availability [62]. It is heavily used for packaging applications. This

causes also major problems, because of the enormous amounts of EPS-waste. So, knowledge of

the recycling capability is very important [63]. EPS is also often used in cheaper cycling helmets,

although it offers less competitive impact properties compared to EPP, which makes the latter

material the favorite for applications with impact deformation. EPS offers slightly lower density

compared to EPP but has less favorable chemical and temperature resistance. However, transport

and storage of EPS is much cheaper. Expandable PS can be transported in huge masses, where

only small masses at the same transport volume of expanded PP can be transported due to its

foamed structure.

EPS offers good insulation capabilities, which lead – in combination with the low price - to the

second highest market shares of insulation materials after glass wool [91]. In principle, the

mechanical behavior of EPS is similar to EPP, since they posses the same basic structure.

However, EPS has a higher specific modulus and strength at the cost of elasticity. Also the

maximum temperature of usage for EPS is lower than EPP. Mechanical properties are highly

dependent on the quality of welding of the beads, which was studied in numerous publications

[71,80,92,93]. For the application in the sectors of thermal insulation and packaging, the

knowledge of creep behavior of EPS is of utmost importance [94–96]. Protective systems often

put EPS to use as a shock absorber, therefore the dynamic properties of this material were

studied in many publications [63,97,98].

34

In contrast to EPP, EPS is much less elastic, which imposes constrictions on the use of EPS for

packaging of high-value goods. This lead to the development of bead foams from PS based

blends, which offer higher elasticity, toughness at low temperature and better chemical

resistance [99].

Thermal transport in foams comprises of conduction through the solid cell walls and struts as

well as the cell gas, convection and radiation. Convection can be neglected for cell smaller than

3 mm [100], which is true for all bead foams. Thermal conduction of foams consists of the

conduction through the solid and the cell gas. The conduction of the polymeric matrix is affected

by crystallinity and orientation [101–103], which are heavily affected by the foaming process.

Either conduction through the solid or through the cell gas can be dominant, depending on

density. For low densities, the cell gas dominates the thermal transport over the solid polymer

because of its high volume fraction. For very small cells with a cell diameter in the same order of

magnitude as the free path length of an air molecule, the Knudsen-effect becomes important and

the thermal conductivity of the cell gas is reduced drastically [104].

Besides conduction, radiation is a very important transport mechanism of thermal energy in

foams. This effect is mainly dependent on cell morphology and temperature. Several workers try

to separate the total heat conductivity into its parts [100,105–110]. However, those contributions

cannot be separated in normal measurements without modelling [111], so the authors used more

or less complex models for separation.

One major drawback of the above-mentioned results of the theoretical models is the

independence of the radiative and conductive contributions. This matter was tackled by Ferkl et

al. [100] in a (at the time being) spatially one-dimensional model. No assumption on the

propagation of radiation and geometry was made. In literature, EPS is often used as a material to

investigate the thermal properties of foams in general [112,91,108]. However, the special

particle-structure of bead foams was never investigated in detail. To reduce thermal radiation

EPS bead foams are equipped with Graphite-particles, which act as reflectors for infrared

radiation thus reducing the overall thermal conductivity. An exemplary product for this kind of

EPS is Neopor (BASF SE).

EPS is well known for packaging applications. For example electronic devices are kept safe from

transport-damage using EPS crash absorbers or spacers. Also in areas, where rigorous safety

35

restrictions exist, such as helmets for cyclists or bikers or car-seats for children, EPS is used

often [113]. In the automotive industry it is used for crash-absorbers as well.

Thanks to its advantageous thermal insulation capability it is used for the insulation of houses in

form of blocks, where it is also used for acoustic insulation against footfall sound. For the

cooling of perishable goods as drugs, food or human blood EPS contributes to keep energy cost

low as insulation and makes the transportation of such goods affordable and practical.

Among particle foams (or bead foams), EPP has unique advantages, such as excellent impact

resistance, energy absorption, insulation, heat resistance and flotation. In addition, it is

lightweight and recyclable, and exhibits good surface protection as well as oil, chemical and

water resistance. Thanks to these advantages, the use of EPP is gaining increased momentum in

the automotive, packaging, and construction industries. The combination of its flexible

applicability, reasonable tooling cost, high resilience, good sound dampening at high

frequencies, and, especially, its low weight, has made EPP the material of choice for numerous

applications. For instance, EPP foams are now utilized as bumper cores, providing significantly

higher energy absorption upon impact as opposed to conventional systems. However, unlike

expandable polystyrene (EPS), which is supplied as expandable small pellets, suppliers can only

provide EPP beads in an expanded or pre-expanded form. The beads are then shipped to the parts

manufacturers for further moulding. Due to the presence of bubbles in the bead (i.e. the large

volume of the bead), the cost of storing, packaging, and transporting EPP is very high, ultimately

rendering it far more expensive than EPS or a normal PP resin. Moreover, very little research has

been conducted on EPP manufacturing, sintering behavior, and steam chest moulding process.

Consequently, when an EPP concept product is targeted, the manufacturer can only depend on

the EPP supplier to obtain a prototype, thus having little or no control over material selections

and processing conditions.

The EPP beads features high closed-cell content which is typically 95-98% as shown in Fig. 2.10

and is measured using a pycnometer in accordance to ASTM D6226. The closed-cell structure

provides high expansion force, while steam-chest moulding assists with the bonding of EPP

beads. Depending on the bulk density, EPP beads have cell diameters from 200-500 µm and cell

densities in the range of 105-10

6 cells/cm

3.

36

The batch foaming process has been successfully used to achieve the high closed-cell content in

EPP beads compared to extrusion foaming. In a batch foaming process for EPP beads, the micro-

pellets are saturated close to the melting point of the material. The high viscosity of polymer

melt provides enormous melt strength so that the cell wall can withstand the bi-axial extension

during the cell growth. The cell density, expansion ratio and crystal characteristic of the

individual EPP bead foams have a significant effect on the overall mechanical properties of the

moulded EPP bead foam product [112,114]. Guo et al. [115] investigated the critical processing

parameters to produce EPP beads in a lab-scale autoclave system. The pressure drop was

systematically controlled by using a modular die at the discharge port. The die geometry (L/D)

was decided to maintain a high enough pressure inside the chamber to prevent pre-foaming of

the gas-impregnated EPP beads. The cell density was not affected by the die geometry. On the

other hand, the volume expansion of the EPP beads slightly decreased as the die length

increased.

The saturation pressure plays a very crucial factor in achieving the high cell densities and

expansion ratios of the EPP beads processed in a autoclave bead foaming setup. In the autoclave

foaming of EPP beads with CO2, a higher saturation pressure allowed a higher CO2 content to be

dissolved into the PP pellets [116,117]. The higher CO2 content helped to reduce the energy

barrier for cell nucleation and increased the cell nucleation rate, which led to a higher final cell

density [118,119]. The volume expansion of EPP beads was also observed to increase

dramatically as the saturation pressure was increased. The higher cell density achieved at high

saturation pressure decreases the amount of gas loss from the foamed EPP beads and hence

improves the expansion ratio.

Figure 2-10 SEM micrograph of a cross-section of an EPP bead made with autoclave

foaming setup

37

The production of EPP beads with double melting peak characteristics has been well established

[120–124]. The two-peak crystal structure is generated by impregnating the PP micro-pellets

with a physical blowing agent in a autoclave chamber at elevated pressures and temperatures

around PP’s melting point over a certain period of time [121,124]. During the gas impregnation

stage, a new crystal melting peak is created at a higher temperature, Tm-high (Fig. 2.8). The newly

generated crystal peak (Tm-high) during the isothermal gas-impregnation stage of the EPP beads

stems from the perfection of the α crystal phase out of the unmelted crystals, which has a higher

orientation and hence a higher melting temperature than the original peak and is known as α2

[125,126]. The melting temperature of this peak is typically above the annealing temperature.

The Tm-low melting peak (Fig. 2.8), is generated during the rapid cooling process in autoclave

foaming chamber and is known as α1. The α1 and α2 are α forms of crystal with various degree of

perfection [127–133]. Choi et al. [125] have shown that by ramping the PP to the annealing

temperature, the less perfect crystals melt and the more perfect crystals that exist above the

annealing temperature remain unmelted. During the annealing treatment, the Tm-high from

unmelted crystals increases with a higher perfection, whereas the portion of the Tm-low peak that

forms during the cooling process decreases since more crystals are formed to the higher melting

peak. However, the work conducted by Choi et al. [125] was at ambient pressure. The actual EPP

bead manufacturing process is conducted at high pressure with blowing agent, which leads to the

dissolution of gas into the PP matrix. The dissolved gas significantly affects the crystallization

behavior of PP [134].

In the context of EPP bead foam manufacturing, the effect of dissolved blowing agent on the

generation of double crystal melting peak structure can be systematically investigated using a

high-pressure differential scanning calorimetry (HP-DSC). The plasticising effect of dissolved

blowing agent, decreases the saturation temperature required for the generation of the higher

melting peak with perfected crystals in EPP bead foams [135].

For EPP bead foams, copolymers with polypropylene (PP) as base monomer are preferred

compared to homo-PP because the latter has poor impact properties at low service temperatures

[125,136–141]. The copolymers can be binary, such as a propylene-ethylene copolymer or a

propylene-butene copolymer, or a ternary copolymer, such as propylene-ethylene-butene

copolymer [125,142]. By using branched high-melt-strength PP [82] and metallocene-catalyzed

PP [122,143], the mechanical properties and compressibility of EPP beads and their moulded

38

foam products can be improved. Other studies have shown that the mechanical properties of EPP

beads can be improved significantly by choosing an appropriate PP copolymer that will lead to

better control of the secondary crystal form [144]. For instance, to improve EPP’s in-mould

foamability, researchers have employed a PP copolymer with a lower melting temperature [145];

in another case, graphite was introduced in order to increase the heat resistance [136].

Furthermore, it has been shown that the use of PP nano-composites can also improve EPP bead

properties [146]. Efforts have also been made to produce expandable PP beads; however, the use

of either an encapsulated physical blowing agent [147] or a dispersed chemical blowing agent

[148] in the beads has not become common practice due to technical difficulties.

As mentioned earlier, for EPP foamed beads to have a good sintering during the steam-chest

moulding stage, they need to possess a double-peak (or at least broad) melting characteristic. The

ratio between the Tm-low and Tm-high peaks is thus crucial in determining the surface quality and

mechanical properties of the steam-chest moulded EPP product. If the Tm-low peak is dominant,

then the moulded EPP product may not have the same geometry as the mould. In contrast, if the

Tm-high peak is dominant, then the sintering will be weak resulting in poor mechanical properties.

The failure mechanism in moulded EPP products have been attributed to the bead boundaries and

a potential fracture path between the beads [149,150]. This is known as inter-bead bonding and it

has been reported that they tend to determine the mechanical properties of the bead products

[149,150]. Inter-bead fracture arises due to weak sintering between the EPP beads. However,

another failure mechanism occurs within the EPP beads and is known as intra-bead fracture. This

failure reflects that there is a good sintering between the EPP beads. The inter-bead and intra-

bead failure mechanism can be investigated by observing under a scanning electron microscope

as shown in Fig. 2.11.

The tensile strength of EPP samples has a strong dependency on the processing steam pressure

and corresponding temperature used during the steam-chest moulding process. The tensile

properties of EPP samples increases at higher steam pressure. A similar phenomenon was

observed in EPS bead processing, where a high tensile strength and a high degree of inter-bead

fusion was obtained at high moulding pressure [150].

The tensile strength of EPP moulded samples also increased significantly due to the development

of crystals in the inter-bead areas during the cooling cycle of the steam-chest moulding process

39

[115,151]. EPP bead size is another important parameter, which affects the inter-bead bonding

and improves the mechanical properties of moulded products [115].

EPP is in a state of constant development and getting ever closer to the customer. Previously

EPP was mainly used in the automotive industry as construction material in the application as

cores for crash bumpers or for tool boxes in the car boot. For those applications the specific

advantages of EPP as low density and good energy dissipation at impact are harvested.

Today’s trends aim towards higher functionality. For example hinges, snap fits or fasteners make

the material fit for new applications as furniture. The challenges are the steam nozzle imprints

and its technical appearance. Therefore the development of multi-material systems is facilitated

[152]. An EPP foam-core can be combined with a layer of TPE for decoration. The connection of

both components can be achieved in an online process. However, if a coating is desired, adhesion

between the coating and EPP is still challenging making a surface treatment necessary [153].

Another approach to modify the properties of EPP is hybridisation [154]. So, the EPP beads are

combined with metal bead foams in order to create a hybrid material with highly elastic behavior

at low stress (behavior dominated by EPP) and high energy dissipation at high stress (behavior

now dominated by metal foam). The purpose is to produce better crash bumpers for cars to

increase passenger safety while reducing weight. It got clear, that the development of EPP is not

at the end, but very dynamic and rapidly advancing, especially towards design and creativity.

Figure 2-11 Failure mechanism: a) inter-bead, b) intra-bead

40

2.6 Thermoplastic polyurethane

Thermoplastic polyurethane (TPU) are class of thermoplastic elastomers (TPE) that combine the

mechanical properties of vulcanized rubber with the processing characteristics of thermoplastic

polymers. The absence of the chemical networks that normally exist in rubber and instead

physical links caused by hydrogen bonding makes TPU material completely reprocessable.

It is well known that TPUs are linear segmented block copolymers of alternating soft and hard

segments. The soft segments (SS), consisting of long polymeric chains of a macro-glycol

(polyether and polyester type), are flexible and weakly polar. The hard segments (HS) are

processed by reaction between diisocyanate, e.g. diphenylmethane-4,4’-diisocyanate (MDI) and

the chain extender, e.g. butanediol. The hard segments are rigid and highly polar. At working

temperature, thermodynamic immiscibility of hard and soft segments results in phase separation

and, consequently a micro-domain structure [160]. Such a structure was first proposed by Cooper

and Tobolsky [161] and is responsible for the peculiar properties of TPUs [162,163]. The hard

segment domains behave as multifunctional tie points functioning both as a physical crosslink’s

and reinforcing fillers, whereas the soft segment form the elastomeric matrix responsible for

material flexibility. This molecular structure results in a number of interesting properties. TPUs

show a high flexibility even at low temperatures, good abrasion behavior, low compression set

and high resistance against oil, fat and solvents.

There is a large number of possible TPUs by varying in the structure of monomers, in

composition and therefore in the final properties. Due to the economic importance of TPU based

on MDI, 1,4-butanediol (BD), and a polyether or polyester macroglycol, the research efforts

have been focused on this particular class of TPUs. The final properties of the TPU are

determined not only by the chemical structure and composition, but also by the synthesis

conditions and thermal history. There is a general assumption that the changes in the thermal

history results in a different microphase structure of the TPU [164,165].

The typical polyurethane is extensively hydrogen bonded [166], the donor being the NH group of

the urethane linkage. The hydrogen-bond acceptor may be either the hard urethane segment (the

carbonyl of the urethane group) or the soft segment (an ester carbonyl or ether oxygen). The

morphological and intermolecular bondings in polyurethane block polymers have been

investigated using various thermoanalytical techniques such as dta, DSC, thermomechanical

41

analysis, and thermal expansion measurements methods. The phase separation starts again as the

cooling process is initiated. Since the mobility of the polymer chains decreases with decreasing

temperature, the phase separation process will be hindered. Different domain and crystallite

morphologies and varying degrees of phase separation can be achieved depending on the cooling

and post annealing conditions [167-172]. A very long annealing time at room temperature, or a

post annealing at temperatures near or above the glass transition temperature of the hard

segments [16,173], are necessary to approach an equilibrium state. Typically thermal transitions

observed in polyurethane elastomers may include the glass transition of either the HS or SS, a

short-range order endotherm of the HS attributable to storage of annealing effects, and

endotherms associated with the long-range order of crystalline portions of either soft [174-177]

or hard segments [178-184]. Above melting temperature of the HS crystallites, the melt becomes

homogeneous [173, 185] with the amorphous HS completely dissolved in the soft segments [167,

173]. The SS glass transition temperature can be used to qualitatively indicate the amount of hard

segment dissolved in the soft domains. A higher glass transition temperature indicates an

increased presence of hard segments dissolved in the soft domains [186,187].

The interpretation of the multiple endothermic behavior observed in TPUs have been extensively

explored and reported in many literatures. The size and position of melting endotherms have

been reported to vary with changes in composition ratio [188,189], soft segment length [176],

annealing [181,183,190,185], processing temperature [191], and mechanical deformation [192].

The endotherm occurring between 50 and 250ºC has been of considerable interest. In earlier

publication [189], the multiple endothermic behavior was attributed to either hydrogen bond

distribution effects or two types of hydrogen bonds, for example, hard segment inter-urethane

hydrogen bonds and hard segment-soft segment hydrogen bonds. However in later studies it was

reported that the observed DSC endotherm are not attributable to hydrogen bond dissociation.

Samuels and Wilkes [193] prepared polymers which employed piperazine and BDO based hard

segments that lacked available hydrogen for hydrogen bonding. However they reported similar

DSC endotherm to those of the hydrogen-bonded materials. They hypothesized that the presence

of various levels of packing order in the hard domains could contribute to the multiple

endotherms.

Seymour and Cooper [164] performed DSC annealing and variable temperature infrared studies

on a series of polyether and polyester-based polyurethanes. They supported the hypothesis

42

proposed by Samuel et al.[194]. They proposed the DSC peaks seen at TI and TII attributed to the

disruption of short and long range order respectively (due to the distribution in hard segment

lengths), and peak TIII to melting of microcrystalline order [164]. By inducing annealing the

short range ordering can be continuously improved until the merging of the I and II regions

[164].

Van Bogart et al.[183] carried an extensive DSC annealing study on several classes of TPUs and

reported that annealing at a certain temperature would invariably result in an endothermic peak

20-50ºC above the annealing temperature. Furthermore, they studied the response of an

MDI/BDO hard segment polymer to annealing, and this yielded similar results to the block

polymers containing shorter sequences of the same material. This implies that annealing-induced

ordering was an intra-domain phenomenon and is not strongly dependent on the presence of soft

segment phase.

Koberstein et al. [185] investigated annealing induced changes in polyurethane morphology

using DSC and simultaneous SAXS-DSC techniques. The relationship between composition

ratio, the presence and position of the various endotherms seen in DSC, and the nature of SAXS

data was analyzed to investigate the structure of the TPU materials. They found the existence of

three endotherm, as reported by earlier researchers.

A major limitation for the use of TPU is its middle up to high hardness. Addition of plasticizers

can achieve soft grade TPUs. However processing is much more challenging and the plasticizers

tend to migrate out of the material in long-term applications. The production of foamed TPU can

reduce the material hardness without additional plasticizers. The reduced density due to foaming

can open new fields of applications for TPU materials. TPUs can be foamed using different

techniques such as extrusion process, batch or continuous process in producing expanded bead

foams.

2.7 References

[1] Landrock AH, Handbook of Plastic Foams: Types, Properties, Manufacture and

Applications.1991, New Jersey: Noyes Publications.

[2] Khemani KC. ACS Symp. Ser. 1997.

[3] Martini-Vvedensky JE, Suh NP, Waldman FA. US Patent 4,473,665; 1984.

43

[4] Park CB, Suh NP, and Baldwin DF, 1999.

[5] Martini JE, Waldman FA, and Suh NP in SPE ANTEC. 1982.

[6] Collias, D. I., & Baird, D. G. (1995). Impact behavior of microcellular foams of

polystyrene and styrene-acrylonitrile copolymer, and single-edge-notched tensile

toughness of microcellular foams of polystyrene, styrene-acrylonitrile copolymer, and

polycarbonate. Polymer Engineering and Science, 35(14), 1178-1183.

[7] Collias, D. I., & Baird, D. G. (1995). Tensile toughness of microcellular foams of

polystyrene, styrene-acrylonitrile copolymer, and polycarbonate, and the effect of

dissolved-gas on the tensile toughness of the same polymer matrices and microcellular

foams. Polymer Engineering and Science, 35(14), 1167-1177.

[8] Matuana, L. M., Park, C. B., & Balatinecz, J. J. (1998). Cell morphology and property

relationships of microcellular foamed PVC/wood-fiber composites. Polymer Engineering

and Science, 38(11), 1862-1872.

[9] Matuana, L. M., Park, C. B., & Balatinecz, J. J. (1998). Structures and mechanical

properties of microcellular foamed polyvinyl chloride. Cellular Polymers, 17(1), 1-16.

[10] Seeler, K. A., & Kumar, V. (1993). Tension-tension fatigue of microcellular

polycarbonate - initial results. Journal of Reinforced Plastics and Composites, 12(3),

359-376.

[11] Kumar, V., & Weller, J. (1994). Production of microcellular polycarbonate using carbon-

dioxide for bubble nucleation. Journal of Engineering for Industry-Transactions of the

Asme, 116(4), 413-420.

[12] Goel, S. K., & Beckman, E. J. (1994). Generation of microcellular polymeric foams using

supercritical carbon-dioxide .1. effect of pressure and temperature on nucleation. Polymer

Engineering and Science, 34(14), 1137-1147.

[13] Handa, Y. P., & Zhang, Z. Y. (2000). A new technique for measuring retrograde

vitrification in polymer-gas systems and for making ultramicrocellular foams from the

retrograde phase. Journal of Polymer Science Part B-Polymer Physics, 38(5), 716-725.

[14] S Siripurapu, Dissertation, Blend-and surface-assisted foaming of polymers with

supercritical carbon dioxide. 2003.

[15] Doroudiani, S., Park, C. B., & Kortschot, M. T. (1998). Processing and characterization

of microcellular foamed high-density polyethylene/isotactic polypropylene blends.

Polymer Engineering and Science, 38(7), 1205-1215.

[16] Kumar V and Schirmer HG, 1997.

44

[17] Park, C. B., Baldwin, D. F., & Suh, N. P. (1995). Effect of the pressure drip rate on cell

nucleation in continuous processing of microcellular polymers. Polymer Engineering and

Science, 35(5), 432-440.

[18] Matuana, L. M., Park, C. B., & Balatinecz, J. J. (1998). Structures and mechanical

properties of microcellular foamed polyvinyl chloride. Cellular Polymers, 17(1), 1-16.

[19] Juntunen RP, Kumar V, and Weller JE. Journal of Vinyl and Additive Technology 2002;

6 (2): 93.

[20] Barlow C, Kumar V, and Flinn B. MD Porous, Cellular and Microcellular Materials,

ASME 2001.

[21] Barlow, C., Kumar, V., Flinn, B., Bordia, R. K., & Weller, J. (2001). Impact strength of

high density solid-state microcellular polycarbonate foams. Journal of Engineering

Materials and Technology-Transactions of the Asme, 123(2), 229-233.

[22] Kumar V. Annual Technical Conference - Society of Plastics Engineers 2002;

60th (2): 1892.

[23] Waldman FA, Ph. D. Dessertation, 1982.

[24] Kumar V, Vanderwel M, and Well J. J. Eng. Mater. Tech 1994.

[25] Arora, K. A., Lesser, A. J., & McCarthy, T. J. (1998). Preparation and characterization of

microcellular polystyrene foams processed in supercritical carbon dioxide.

Macromolecules, 31(14), 4614-4620.

[26] Cooper, A. I. (2000). Polymer synthesis and processing using supercritical carbon

dioxide. Journal of Materials Chemistry, 10(2), 207-234.

[27] Kirby, C. F., & McHugh, M. A. (1999). Phase behavior of polymers in supercritical fluid

solvents. Chemical Reviews, 99(2), 565-602.

[28] Okamoto, M., Nam, P. H., Maiti, P., Kotaka, T., Nakayama, T., Takada, M., et al.

(2001). Biaxial flow-induced alignment of silicate layers in polypropylene/clay

nanocomposite foam. Nano Letters, 1(9), 503-505.

[29] Wissinger, R. G., & Paulaitis, M. E. (1987). Swelling and sorption in polymer-CO2

mixtures at elevated pressures. Journal of Polymer Science Part B-Polymer Physics,

25(12), 2497-2510.

[30] Wissinger, R. G., & Paulaitis, M. E. (1991). Molecular thermodynamic model for

sorption and swelling in glassy polymer-CO2 systems at elevated pressures. Industrial &

Engineering Chemistry Research, 30(5), 842-851.

45

[31] Zhang Q, Xanthos M, and Dey SK. MD (American Society of Mechanical Engineers)

1998; 82 (Porous, Cellular and Microcellular Materials) 75-83.

[32] Handa, Y. P., Zhang, Z., & Wong, B. (2001). Solubility, diffusivity, and retrograde

vitrification in PMMA-CO2, and development of sub-micron cellular structures. Cellular

Polymers, 20(1), 1-16.

[33] Manninen, A. R., Naguib, H. E., Nawaby, A. V., Liao, X., & Day, M. (2005). The effect

of clay content on PMMA-clay nanocomposite foams. Cellular Polymers, 24(2), 49-70.

[34] Koros, W. J., & Paul, D. R. (1978). CO2 sorption in poly(ethylene-terephthalate) above

and below glass-transition. Journal of Polymer Science Part B-Polymer Physics, 16(11),

1947-1963.

[35] Kazarian, S. G., Vincent, M. F., Bright, F. V., Liotta, C. L., & Eckert, C. A. (1996).

Specific intermolecular interaction of carbon dioxide with polymers. Journal of the

American Chemical Society, 118(7), 1729-1736.

[36] Colton, J. S., & Suh, N. P. (1987). The nucleation of microcellular thermoplastic foam

with additives .1. theoretical considerations. Polymer Engineering and Science, 27(7),

485-492.

[37] Condo, P. D., Sanchez, I. C., Panayiotou, C. G., & Johnston, K. P. (1992). Glass-

transition behavior including retrograde vitrification of polymers with compressed fluid

diluents. Macromolecules, 25(23), 6119-6127.

[38] Peter Spitael, Christopher W. Macosko, Richard B, Mcclurg. Macromolecules, 2004, 37

(18), 6874-6882.

[39] Shimoda, M., Tsujimura, I., Tanigaki, M., & Ohshima, M. (2001). Polymeric foaming

simulation for extrusion processes. Journal of Cellular Plastics, , 37(6) 517-536.

[40] Pop-Iliev R, Liu F, G Liu, Park CB (2003). Advances in Polymer Technology, 22, 4, 280-

296.

[41] C. B. Park, D. F. Baldwin, and N. P. Suh, Polym. Eng. Sci., 35, 432-440 (1995)

[42] C. B. Park, A. H. Behravesh, and R. D. Venter, Polym. Eng. Sci., 38, 1812 (1998)

[43] A. H. Behravesh, C. B. Park, and R. D. Venter, Cellular Polymers, 17, 309 (1998)

[44] X. Xu, C. B. Park, D. Xu, and R. Pop-Iliev, Polym Eng Sci 43, 1378 (2003)

[45] W. Michaeli, Extrusion Dies for Plastics and Rubber, 52-64, Hanser (2003)

[46] H. E. Naguib, C. B. Park, P. C. Lee, and D. Xu, J. Polym Eng 26, 6 (2006)

46

[47] J. W. S. Lee, "Investigation of Foaming Behaviors In Injection Molding Using Mold

Pressure Profile," Ph.D., Mechanical and Industrial Engineering, University of Toronto,

Toronto, 2009.

[48] J. Xu, Microcellular Injection Molding: John Wiley and Sons, Inc.,, 2010.

[49] J. W. S. Lee, J. Wang, J. D. Yoon, and C. B. Park, "Strategies to achieve a uniform cell

structure with a high void fraction in advanced structural foam molding," Industrial and

Engineering Chemistry Research, vol. 47, pp. 9457- 9464, 2008.

[50] Liu, G., Park, C.B., and Lefas. J.A., “Production of Low-Density LLDPE Foams in

Rotational Molding”, Polymer Engineering and Science, 38, 1997-2009, 1998.

[51] Gogos, G., "Bubble Removal in Rotational Molding”, Polymer Engineering and Science,

44, 388-394, 2004.

[52] Abdullah, M.Z., Bickerton, S., and Bhattacharyya, D., “Enhancement of Convective Heat

Transfer to Polymer Manufacturing Molds”, Polymer Engineering and Science, 45, 114-

124, 2005.

[53] Crawford, R.J. and Throne, J.L., Rotational Molding Technology, William Andrew

Publishing/Plastics Design Library, 2002.

[54] Martin, D., Halley, P., Truss, R., Murphy, M., Jackson, O., and Kwon, O., “Polyethylene-

layered Silicate Nanocomposites for Rotational Moulding”, Polymer International, 52:11,

1774-1779, 2003.

[55] Dodge, P., Modern Plastics Encyclopedia, D-171, 1996.

[56] Pop-Iliev, R., Rizvi, G.M., and Park, C.B., “The Importance of Timely Polymer Sintering

While Processing Polypropylene Foams in Rotational Molding”, Polymer Engineering

and Science, 43, 40-54, 2003.

[57] Archer, E., Harkin-Jones, E., Kearns, M.P., and Fatnes, A., “Processing Characteristics

and Mechanical Properties of Metallocene Catalyzed Linear Low-density Polyethylene

Foams for Rotational Molding”, Polymer Engineering and Science, 44, 638-647, 2004.

[58] Needham, D.G., U.S. Patent 5,366,675, 1994.

[59] Pop-Iliev, R., and Park, C.B., “Single-step Rotational Foam Molding of Skin-Surrounded

Polyethylene Foams”, Journal of Cellular Plastics, 39:1, 49-58, 2003.

[60] Pop-Iliev, R. and Park, C.B., “Processing of Polypropylene Foams in Melt Compounding

Based Rotational Foam Molding”, Journal of Reinforced Plastics and Composites, 21:12,

1079-1100, 2002.

47

[61] Maletzko C, Keppeler U, Klaus H, de Grave I. US 6,864,298 B2 - EXPANDABLE

POLYOLEFIN PARTICLES, 2005.

[62] Tang L, Zhai W, Zheng W. Autoclave preparation of expanded polypropylene/poly(lactic

acid) blend bead foams with a batch foaming process. J Cell Plast 2011;47:429–46.

[63] Hornberger L, Might T, Mohareb N, Pirie T. Dynamic Properties of Recycled EPS Foam.

Polym Plast Technol Eng 1997;36:917–30.

[64] Doroudiani S, Park CB, Kortschot MT. Effect of the crystallinity and morphology on the

microcellular foam structure of semicrystalline polymers. Polym Eng Sci 1996;36:2645–

62.

[65] Lee EK. Novel Manufacturing Processes for Polymer Bead Foams by. University of

Toronto, 2010.

[66] Stupak PR, Frye WO, Donovan JA. The Effect of Bead Fusion of Polystyrene Foam .

1991.

[67] Zhai W, Kim Y-W, Jung DW, Park CB. Steam-Chest Molding of Expanded

Polypropylene Foams. 2. Mechanism of Interbead Bonding. Ind Eng Chem Res

2011;50:5523–31.

[68] BASF SE. Neopolen P - Technical information. 2009.

[69] BASF SE. Datasheet Neopolen® P 8220 K 2010.

[70] Svec P, Rosik L, Hoak Z, Vecerka F. styrene Based Plastics and Their Modifications.

1990.

[71] Rossacci J, Shivkumar S. Bead fusion in polystyrene foams. J Mater Sci 2003;8:201–6.

[72] Pop-Iliev R, Rizvi GM, Park CB. The importance of timely polymer sintering while

processing polypropylene foams in rotational molding. Polym Eng Sci 2003;43:40–54.

[73] Pop-Iliev R. Manufacture of Integral Skin PP Foam Composites in Rotational Molding. J

Cell Plast 2006;42:139–52.

[74] Bellehumeur CT, Tiang JS. Simulation of non-isothermal melt densification of

polyethylene in rotational molding. Polym Eng Sci 2002;42:215–29.

[75] Mills NJ. Polymer Foams Handbook. Elsevier; 2007.

48

[76] Nakai S, Taki K, Tsujimura I, Ohshima M. Numerical Simulation of a Polypropylene

Foam Bead Expansion Process 2008.

[77] Zhai W, Kim Y-W, Park CB. Steam-Chest Molding of Expanded Polypropylene Foams.

1. DSC Simulation of Bead Foam Processing. Ind Eng Chem Res 2010;49:9822–9.

[78] Zhai W, Kim Y-W, Jung DW, Park CB. Steam-Chest Molding of Expanded

Polypropylene Foams. 2. Mechanism of Interbead Bonding. Ind Eng Chem Res

2011;50:5523–31.

[79] Sands M. An Analysis of Mold Filling and Defect Formation in Lost Foam Castings.

Worcester Polytechnic Institute, 1998.

[80] Stupak PR, Frye WO, Donovan J a. The Effect of Bead Fusion on the Energy Absorption

of Polystyrene Foam. Part I: Fracture Toughness. J Cell Plast 1991;27:484–505.

[81] Mills NJ, Gilcrist A. Shear and Compressive Impact of Polypropylene Bead Foam. Cell

Polym 1999;18:157–74.

[82] Klempner D, Sendijareviʹc V, Aseeva RM. Handbook of Polymeric Foams and Foam

Technology. Hanser Verlag; 2004.

[83] Mills NJ. Time Dependence of the Compressive Response of Polypropylene Bead Foam.

Cell Polym 1997;16:194–215.

[84] Nakai S, Taki K, Tsujimura I, Ohshima M. Numerical simulation of a polypropylene

foam bead expansion process. Polym Eng Sci 2008;48:107–15.

[85] Yang C-T, Lee S-T. Dimensional Stability Analysis of Foams Based on LDPE And

Ethylene-Styrene Interpolymer Blends. J Cell Plast 2003;39:59–69.

[86] Frederick G, Kaepp GA, Kudelko CM, Schuster PJ, Domas F, Haardt UG, et al.

Optimization of Expanded Polypropylene Foam Coring to Improve Bumper Foam Core

Energy Absorbing Capability. 1995.

[87] Mahapatro A, Mills NJ, Sims GLA. Experiments and modelling of the expansion of

crosslinked polyethylene foams. Cell Polym 1998;17:252–70.

[88] Beverte I. Elastic constants of monotropic plastic foams. 2. Deformation parallel to foam

rise direction. Numerical calculations. Mech Compos Mater 1998;34:115–22.

49

[89] Viot P. Hydrostatic compression on polypropylene foam. Int J Impact Eng 2009;36:975–

89.

[90] Müller H. EPS-Herstellung: Druck erfordert Gegendruck. Kunststoffe 2010;2:20–3.

[91] Coquard R, Baillis D, Quenard D. Heat transfer in low-density eps foams. 12èmes

Journées Int. Therm., 2005, p. 291–4.

[92] Stupak PR, Donovan J a. The Effect of Bead Fusion on the Energy Absorption of

Polystyrene Foam. Part II: Energy Absorption. J Cell Plast 1991;27:506–13.

[93] Schellenberg J, Wallis M. Dependence of properties of expandable polystyrene particle

foam on degree of fusion. J Appl Polym Sci 2010;115:2986–90.

[94] Vaitkus S, Gnip I, Keršulis V, Vėjelis S. Prediction of Creep Strain of the Expanded

Polystyrene ( EPS ) in Long-term Compression 2007;13.

[95] Gnip IY, Vaitkus S, Keršulis V, Vėjelis S. Confidence intervals of long-term prediction

and synthesis of creep compliance prediction estimates for expanded polystyrene (EPS)

under permanent compressive stress. Polym Test 2008;27:688–97.

[96] Gnip IY, Vaitkus S, Keršulis V, Vėjelis S. Experiments for the long-term prediction of

creep strain of expanded polystyrene under compressive stress. Polym Test 2010;29:693–

700.

[97] Gimbel GM, Hoshizaki TB. Compressive properties of helmet materials subjected to

dynamic impact loading of various energies. Eur J Sport Sci 2008;8:341–9.

[98] Doroudiani S, Kortschot MT. Polystyrene foams. III. Structure-tensile properties

relationships. J Appl Polym Sci 2003;90:1427–34.

[99] SCHIPS C, HAHN K, HOFMANN M, RUCKDÄSCHEL H, AßMANN J, JANSSENS

G, et al. EP 2 384 354 B1 - Elastic particle foam based on polyolefin/styrene polymer

mixtures, 2013.

[100] Ferkl P, Pokorný R, Bobák M, Kosek J. Heat transfer in one-dimensional micro- and

nano-cellular foams. Chem Eng Sci 2013;97:50–8.

[101] Hansen D, Bernier GA. Thermal conductivity of polyethylene: The effects of crystal size,

density and orientation on the thermal conductivity. Polym Eng Sci 1972;12:204–8.

50

[102] Hansen D, Ho CC. Thermal conductivity of high polymers. J Polym Sci Part A Gen Pap

1965;3:659–70.

[103] Choy CL, Chen FC, Luk WH, Arle II. Thermal Conductivity of Oriented Crystalline

Polymers 1980;18:1187–207.

[104] Costeux S, Zhu L. Low density thermoplastic nanofoams nucleated by nanoparticles.

Polymer (Guildf) 2013;54:2785–95.

[105] Sundarram SS, Li W. On thermal conductivity of micro- and nanocellular polymer

foams. Polym Eng Sci 2013;53:1901–9.

[106] Xie T, He Y-L, Hu Z-J. Theoretical study on thermal conductivities of silica aerogel

composite insulating material. Int J Heat Mass Transf 2013;58:540–52.

[107] Kaemmerlen A, Vo C, Asllanaj F, Jeandel G, Baillis D. Radiative properties of extruded

polystyrene foams: Predictive model and experimental results. J Quant Spectrosc Radiat

Transf 2010;111:865–77.

[108] Placido E, Arduini-Schuster MC, Kuhn J. Thermal properties predictive model for

insulating foams. Infrared Phys Technol 2005;46:219–31.

[109] Tseng C, Kuo K. Thermal radiative properties of phenolic foam insulation. J Quant

Spectrosc Radiat Transf 2002;72:349–59.

[110] Caps R, Heinemann U, Fricke J, Keller K. Thermal conductivity of polyimide foams. Int

J Heat Mass Transf 1997;40:269–80.

[111] Coquard R, Rochais D, Baillis D. Experimental investigations of the coupled conductive

and radiative heat transfer in metallic/ceramic foams. Int J Heat Mass Transf

2009;52:4907–18.

[112] Schellenberg J, Wallis M. Dependence of Thermal Properties of Expandable Polystyrene

Particle Foam on Cell Size and Density. J Cell Plast 2010;46:209–22.

[113] Trassl C, Wörthwein H. Klasse statt Masse. Kunststoffe 2010;12:134–7.

[114] Sasaki H, Nakamura Y. EP002151470A1 - POLYPROPYLENE RESIN FOAM

PARTICLE AND MOLDED ARTICLE THEREFROM, 2010.

51

[115] Guo Y, Hossieny N, Chu RKM, Park CB, Zhou N. Critical processing parameters for

foamed bead manufacturing in a lab-scale autoclave system. Chem Eng J 2013;214:180–

8.

[116] Li G, Wang J, Park CB, Simha R. Measurement of gas solubility in linear/branched PP

melts. J Polym Sci Part B Polym Phys 2007;45:2497–508.

[117] Sato Y, Fujiwara K, Takikawa T, Takishima S, Masuoka H. Solubilities and diffusion

coefficients of carbon dioxide and nitrogen in polypropylene, high-density polyethylene,

and polystyrene under high pressures and temperatures. Fluid Phase Equilib

1999;162:261–76.

[118] Leung SN, Wong A, Guo Q, Park CB, Zong JH. Change in the critical nucleation radius

and its impact on cell stability during polymeric foaming processes. Chem Eng Sci

2009;64:4899–907.

[119] Leung SN, Park CB, Li H. Numerical simulation of polymeric foaming processes using

modified nucleation theory. Plast Rubber Compos 2006;35:93–100.

[120] Sasaki H, Ogiyama K, Hira A, Hashimoto K, Yanagisawa H, Tokoro H. US 6,838,488

B2 - PRODUCTION METHOD OF FOAMED POLYPROPYLENE RESIN BEADS,

2005.

[121] Braun F. US 6,723,760 B2 - METHOD FOR PRODUCING EXPANDED OR

EXPANDABLE POLYOLEFIN PARTICLES, 2004.

[122] Sasaki H, Sakaguchi M, Akiyama M, Tokoro H. US 6,313,184 B1 EXPANDED

POLYPROPYLENE RESIN BEADS AND A MOLDED ARTICLE, 2001.

[123] Hira A, Hashimoto K, Sasaki H. US 7,182,896 B2 - COMPOSITE FOAMED

POLYPROPYLENE RESIN MOLDING AND METHOD OF PRODUCING SAME,

2007.

[124] Schweinzer J, Fischer J, de Grave I, Kogel W. US 5703135 A - Production of expanded

polyolefin beads, 1997.

[125] Choi JB, Chung MJ, Yoon JS. Formation of Double Melting Peak of Poly(propylene- c o

-ethylene- c o -1-butene) during the Preexpansion Process for Production of Expanded

Polypropylene. Ind Eng Chem Res 2005;44:2776–80.

52

[126] Ruiyun Z, Xiaolie L, Qunhua W, Dezhu M. Melting Behavior of Low Ethylene Content

Polypropylene Copolymers with and without Nucleating Agents. Chinese J Polmyer Sci

1994;12:246.

[127] Sasaki H, Hira A, Hashimoto K, Tokoro H. US 20030162012 A1 - Expanded

polypropylene resin bead and process of producing same, 2003.

[128] Napolitano R, Pirozzi B, Varriale V. Temperature dependence of the thermodynamic

stability of the two crystalline α forms of isotactic polypropylene. J Polym Sci Part B

Polym Phys 1990;28:139–47.

[129] Turner-Jones A. Development of the γ-crystal form in random copolymers of propylene

and their analysis by dsc and x-ray methods. Polymer (Guildf) 1971;12:487–508.

[130] Janimak JJ, Cheng SZD, Zhang A, Hsieh ET. Isotacticity effect on crystallization and

melting in polypropylene fractions: 3. Overall crystallization and melting behaviour.

Polymer (Guildf) 1992;33:728–35.

[131] De Rosa C, Guerra G, Napolitano R, Petraccone V, Pirozzi B. Conditions for the α1-α2

transition in isotactic polypropylene samples. Eur Polym J 1984;20:937–41.

[132] Petraccone V, Guerra G, De Rosa C, Tuzi A. Extrapolation to the equilibrium melting

temperature for isotactic polypropylene. Macromolecules 1985;18:813–4.

[133] Hikosaka M, Seto T. The Order of the Molecular Chains in Isotactic Polypropylene

Crystals. Polym J 1973;5:111–27.

[134] Raps D, Köppl T, de Anda AR, Altstädt V. Rheological and crystallisation behaviour of

high melt strength polypropylene under gas-loading. Polymer (Guildf) 2014;55:1537–45.

[135] Nofar M, Guo Y, Park CB. Double Crystal Melting Peak Generation for Expanded

Polypropylene Bead Foam Manufacturing. Ind Eng Chem Res 2013;52:2297–303.

[136] Braun F, Glück G, Hahn K, de Grave I, Tatzel H. US 6677040 B1 - Expanded

polypropylene particles, 2004.

[137] Kogel W, de Grave I, Hahn K, Fischer J. US 5925686 A - Production of expanded

polyolefin particles, 1999.

53

[138] Sasaki H, Sakaguchi M, Akiyama M, Tokoro H. US 6077875 A - Beads comprising

uncrosslinked ethylene-propylene random copolymer having melting point of at least 140

degress celsius as base resin, 2000.

[139] Fischer J, Dietzen F-J, Ehrmann G, de Grave I, Rieger J. US 5744505 A - Prefoamed

polyolefin beads produced by extrusion, 1998.

[140] Tripathi D. Practical Guide to Polypropylene. Rapra Technology Ltd; n.d.

[141] Ito J-I, Mitani K, Mizutani Y. Annealing of commercial block polypropylene. II.

Behavior of poly(ethylene-co-propylene) component. J Appl Polym Sci 1992;46:1235–

43.

[142] Tjong SC, Xu SA. Non-isothermal crystallization kinetics of calcium carbonate-filled β-

crystalline phase polypropylene composites. Polym Int 1997;44:95–103.

[143] Sugano T, Yokkaichi SK, Yokkaichi-Shi M-S. EP 0611795 A1 - Polypropylene resin

expanded particles, n.d.

[144] Tsurugai K, Tokoro H, Oikawa M. US 5747549 A - Polypropylene resin for moldings,

1998.

[145] Yamaguchi T, Iwamoto T, Yamaguchi M, Senda K. US 6607682 B1 - Pre-expanded

polypropylene resin beads and process for producing molded object therefrom by in-mold

foaming, 2003.

[146] Ziegler M. WO 2004003063 A8 - Thermoplastische schaumstoffe mit nanostrukturierten

füllstoffen und verfahren zu ihrer herstellung, 2004.

[147] Fischer J, Rieger J, Hahn K, de Grave I, Kogel W. US 5773481 A - Polyolefin particle

foam, 1998.

[148] Park CB, Liu G, Liu F, Pop-Iliev R, D’Uva S, Zhang B. US 6103153 A - Rotational

molding with charging and rotation, 2000.

[149] Lee JWS, Wang J, Yoon JD, Park CB. Strategies to Achieve a Uniform Cell Structure

with a High Void Fraction in Advanced Structural Foam Molding. Ind Eng Chem Res

2008;47:9457–64.

[150] Barnetson A. Expanded polystyrene: development, processing, applications and key

issues. In: Eaves D, editor. Handb. Polym. Foam., Rapra Technology Limited; n.d.

54

[151] Park CB, Baldwin DF, Suh NP. Effect of the pressure drop rate on cell nucleation in

continuous processing of microcellular polymers. Polym Eng Sci 1995;35:432–40.

[152] Trassl C, Altstädt V, Schreier P. Concepts for particle foam based ultralight automotive

interior parts, 2014, p. 354–7.

[153] Schreier P, Trassl C, Altstädt V. Surface modification of polypropylene based particle

foams, 2014, p. 378–82.

[154] Rapp F, Diemert J. New Applications: Multifunctional Hybrid Foams. Kunststoffe

2012;7:60–3.

[155] Jonsson L. Foaming of TPE with thermally expandable microspheres. Thermoplast.

Elastomers, iSmithers Rapra Publishing; 2010.

[156] Sahnoune A. Foaming of Thermoplastic Elastomers with Water. J Cell Plast

2001;37:149–59.

[157] Kim SG, Lee JWS, Park CB, Sain M. Enhancing cell nucleation of thermoplastic

polyolefin foam blown with nitrogen. J Appl Polym Sci 2010:n/a–n/a.

[158] Garlotta D. A Literature Review of Poly ( Lactic Acid ). J Polym Environ 2001;9:63–84.

[159] Nofar M, Zhu W, Park CB. Effect of dissolved CO2 on the crystallization behavior of

linear and branched PLA. Polymer (Guildf) 2012;53:3341–53.

[160] P.E. Gibson, M.A. Wallace, and S.L. Cooper, Development in Block Copolymers, p.217,

I.Goodman, ed., Elsevier, London (1982).

[161] S.L.Cooper and A.V.Tobolsky, J. Appl. Polym. Sci., 10, 1837 (1966).

[162] G.Oertel, Polyurethane Handbook, Hanser Publishers, Munich (1993).

[163] C.S.Schollenberger, Polyurethane Technol., 197 (1969).

[164] R.W. Seymour and S.I. Cooper, Macromolecules, 6, 48 (1973).

[165] G. Pohl, D. Joel, H. Goering, and H.E. Carius, Plastik und Kautschuk, 40, 357 (1993).

55

[166] R.W. Seymour, G.M. Estes, and S.L. Cooper, Macromolecules, 3, 579 (1970).

[167] Y.Li, T. Gao, J.Liu, K.Linliu, C.R. Desper, and B. Chu, Macromolecules, 25, 7365

(1992).

[168] R.Bonart, L.Morbitzer, and G.J.Hentze, J. Macromol. Sci.-Phys., B3(2), 337 (1969).

[169] J. Blackwell, M.R. Nagarajan, and T.B. Hoitink, Polymer, 23, 950 (1982).

[170] J. Blackwell and C.D.Lee, J. Polym. Sci., Polym. Phys., 22, 759 (1984).

[171] R.M.Briber and E.L. Thomas, J. Macromol Sci. Phys., B22, 509 (1983).

[172] J.T.Koberstein and T.P. Russell, Macromolecules, 19, 714 (1986).

[173] Y.Li, T.Gao, and B.Chu, Macromolecules, 25, 1737 (1992).

[174] J.Foks, H.Janik, and R. Russo, Eur. Polym. J., 26, 309 (1990).

[175] R.Russo, Colloid Polym. Sci., 269, 1 (1991).

[176] K. Sreenivasan, Eur. Polym. J., 27, 811 (1991).

[177] X. Yuying, Z. Zhiping, W. Dening, and Y. Shengkang, Polymer, 33, 1335 (1992).

[178] J.Blackwell and C.D.Lee, Advances in Urethane Science and Technology, K.C. Frisch

and D. Klempner, Eds., Technomic, Stamford, Conn., (1984).

[179] Y. Camberlin, J.P. Pascault, M. Letoffe, and P.Claudy, J. Polym. Sci., Polym. Chem. Ed.,

20, 383 (1982).

[180] T.R. Hesketh, J.W.C. Van Bogart, and S.L.Cooper, Polym. Eng. Sci., 20, 190 (1980).

56

[181] W. Hu and J.T. Koberstein, J. Polym. Sci., Polym. Phys. Ed., 32, 437 (1994).

[182] J.T. Koberstein and A.F. Galambos, Macromolecules, 25, 5618 (1992).

[183] J.W.C. Van Bogart, D.A. Bluemke, and S.L. Cooper, Polymer, 22, 1428 (1981).

[184] J.W.C. Van Bogart, P.E. Gibson, and S.L. Cooper, J. Polym. Sci., Polym. Phys. Ed., 21,

65 (1983).

[185] L.M. Leung and J.T. Koberstein, Macromolecules, 19, 706 (1986).

[186] T.G. Fox, Bull. Am. Phys. Soc., 50, 549 (1956).

[187] L.A.Wood, J. Polym. Sci., 28, 319 (1958).

[188] Z.S.Petrovic and I.Javni, J. Polym. Sci., Polym. Phys. Ed., 27, 545 (1989).

[189] R.W. Seymour and S.L. Cooper, Polym. Lett., 9, 695 (1971).

[190] M.E. Kazmierczak, R.E. Fornes, D.R. Buchanan, and R.D. Gilbert, J. Polym. Sci., Polym.

Phys. Ed., 27, 2188 (1989).

[191] J.A. Miller, S.B. Lin, K.K.S. Hwang, K.S. Wu, P.E. Gibson, and S.L. Cooper,

Macromolecules, 18, 32 (1985).

[192] W. Meckel, W. Goyert, and W. Weider, Thermoplastic Elastomers, N.R. Legge, L.

Holden, and H.E. Schroeder, Eds., Hanser, Munich, New York, 1987.

[193] S.B. Clough, N.S. Schneider, and A.O. King, J. Macromol. Sci., B2, 641 (1968).

[194] S.L. Samuels and G.L. Wilkes, J. Polym. Sci., Polym. Symp., 43, 149 (1973).

57

Chapter 3 Phase Separation and Crystallization of TPU in the Presence of

Dissolved Gas:- Effects of Processing, Nano-/Micron-Sized Additives and Gas Types

3 Phase Separation and Crystallization of TPU in the Presence of Dissolved Gas

3.1 Introduction

Thermoplastic polyurethanes (TPUs) are multi-block copolymers that exhibit a unique

combination of strength, flexibility and processability due to their phase-separated

microstructure.[1,2] These properties result from a molecular structure with rigid HS domains

dispersed in the soft segment (SS) matrix. The SS is a polyol with an ester or ether group in the

main chain having a low glass transition temperature and is viscous at service temperature,

imparting flexibility to TPU. The HS is formed by the reaction of diisocyanate and short-chain

diols, which crystallizes and influences the mechanical properties in TPU such as hardness and

tear strength. As a result of this unique microstructure, TPUs exhibit very good impact properties

at low temperature, excellent chemical resistance and great flexibility over a broad service

temperature, which make them suitable for a wide range of demanding applications such as

automobile parts, construction materials, sports equipment, and medical instruments.[3]

TPU’s phase separation strongly depends on the hydrogen bonding between the HSs and its

crystallization kinetics [4,5]. Generally, the extent of phase separation is incomplete and the

microstructure of TPU consists of mixed HS and SS segmental chains. The presence of inter-

segmental mixing affects the morphology, the thermal and the mechanical properties of TPUs.

The incorporation of HSs within the SSs elevates the glass transition temperature and degrades

the TPUs elastic properties [6]. On the other hand; the inclusion of SS within the HS domains

reduces its crystallinity. Due to its high commercial value, the HS phase-separation and

crystallization behavior in TPUs based on the reaction of 4,4'-methylenediphenyl 1,1'-

diisocyanate (MDI) and butanediol (BDO) have been extensively studied[7,8]. A detailed

morphological analysis using SAXS for a series of MDI/BDO based TPUs showed that the HS

phase-separated domain structure was in agreement with a model proposed by Koberstein and

58

Stein[6]. The basis of this model is that HS chains shorter than the critical length for microphase

separation are presumed to remain dissolved within the SS microphase, while longer segments

aggregate into lamellar HS crystalline domains.

The multiple endotherms, size, and melting peak temperatures in TPUs have been studied by

numerous authors in the past, and the changes in such characteristics have been attributed to the

changes in the HS and SS ratio, thermal annealing, thermal history, processing condition, and

mechanical deformation [4,5,9-11]. Thermal annealing of TPUs leads to rearrangement of

hydrogen bonds and thus improves the formation of HS domains and their crystallization kinetics

[6,12,13]. The thermal studies on TPU have been conducted at atmospheric pressure [6-8,12-14.

There is, however, very limited study on the effects of high pressure gas on the phase-separation

and crystallization behavior of HS in the TPU microstructure. Some studies have investigated the

diffusion of gases such as oxygen, carbon dioxide and hydrogen through TPUs at high pressures;

however, none have discussed the phase-separation and/or crystallization of HSs in the presence

of the dissolved gas [15,16]. In our recent study, we demonstrated the effect of butane gas on the

phase separation and crystallization behaviors of HSs in TPU [17]. The dissolved butane acted as

a plasticizer and assisted the HS chains to phase-separate and significantly increased the overall

heat of fusion of the TPU.

The concept of utilizing dissolved gas to improve the phase separation and crystallization of HS

chains can be effectively used to develop a number of interesting technologies for TPU. One of

the technologies is to introduce cellular morphology, which would lead to density and hardness

reduction and consequently decrease the cost. The heterogeneous cell nucleation rate during

foam processing can be significantly promoted through local pressure variations [18-20], around

the HS domains and crystallites [21-13]. In addition, the surrounding areas of newly formed (or

growing) HS domains and crystals have an increased amount of gas due to gas exertion from the

phase-separated and crystallized region, which is further favorable for heterogeneous cell

nucleation[24]. It is well known that the crystallization behavior of polymers under dissolved gas

is expected to be fundamentally different from that under air at ambient pressure [25]. Dissolved

gas causes swelling of the polymer matrix which increases the molecular chain mobility [25,26].

This affects the surface tension [27-31], the viscosities [32-34], and the thermal behaviors

including the crystallization kinetics [35-38]. Overall the varying crystallization kinetics at

various pressures can significantly influence the final foam morphology.

59

In this chapter, we examined the effect of melt-compounding and the presence of three different

additives on the phase separation and crystallization behavior of HS in the TPU microstructure at

atmospheric pressure using a regular DSC. These additives were nano-clay, nano-silica and

glycerol monosterate (GMS). The melt-compounding and the compounding of the additives into

TPU microstructure were done using a twin-screw micro-compounder. The effect of dissolved

CO2 at high-pressure on the phase separation and crystallization behavior of neat-TPU and TPU

with the additives was investigated using a high-pressure differential scanning calorimeter (HP-

DSC). The effect of dissolved butane on the phase separation and crystallization behavior of

neat-TPU and TPU with GMS was investigated in a specially designed saturation system. The

generated HS crystallites were analyzed with atomic force microscopy, wide-angle X-ray

diffractometer (WAXS) and small angle X-ray diffractometer (SAXS).

3.2 Experimental Procedure

3.2.1 Materials

The TPU used in this study was Elastollan manufactured by BASF with a melting temperature of

171°C, a specific density of 1.13 g/cm3, and a hardness of Shore 90A. The HSs are composed of

reaction between MDI and BDO. The SSs are polyether diols, with a high hydrolysis resistance

tendency. Glycerol monosterate (GMS) used as a diffusion retarder and also an element which

modifies the crystallization behavior of HS was Pationic 915. The nano-clay used was Cloiste

30B. The nano-silica used was Aerosil A 200. The N-butane and CO2, supplied by Linde Gas

Canada, was used as the blowing agent.

3.2.2 Sample preparation

The “as-received” TPU material (AR-TPU) was compounded in a twin screw extruder DSM

Micro-compounder. Prior to compounding, the AR-TPU was dried in a CONAIR drier at 105°C

for 4 hours to remove moisture. The compounding was implemented at a processing temperature

of 190°C for 3 min and a rpm of 50. The samples of the extruded “processed” TPU (PR-TPU)

and the AR-TPU were used in the phase separation and crystallization of HSs .

The “as-received” TPU and GMS were dry blended and then compounded in a twin screw

extruder (DSM Microcompounder). Prior to compounding, the “as-received” TPU was dried in a

CONAIR drier at 105°C for 4 hours to remove moisture. The compounding was implemented at

60

a processing temperature of 190°C for 3 min and a screw speed of 50 rpm. A series of TPU-

GMS samples with GMS contents of 0.5, 1 and 2 wt%, named as TPU-05GMS, TPU-1GMS and

TPU-2GMS, respectively, were prepared.

The “as-received” TPU material (AR-TPU) and nano-clay (Cloisite 30B) were compounded in a

twin screw extruder (DSM Microcompounder). Prior to compounding, the AR-TPU was dried in

a CONAIR drier at 105°C for 4 hours to remove moisture. A series of TPU/nano-clay sample

with nano-clay contents of 0.5, 1 & 2 wt% (TPU-05NCl, TPU-1NCl and TPU-2NCl) were

prepared.

The “as-received” TPU material (AR-TPU) and nano-silica (Aerosil A200) were compounded in

a twin screw extruder (DSM Microcompounder). Prior to compounding, the AR-TPU was dried

in a CONAIR drier at 105°C for 4 hours to remove moisture. A series of TPU/nano-silica sample

with nano-silica contents of 0.5, 1 & 2 wt% (TPU-05NSi, TPU-1NSi and TPU-2NSi) were

prepared.

3.2.3 Rheological analysis

The shear viscosities of AR-TPU, PR-TPU, TPU-GMS, TPU-NCl and TPU-NSi samples were

measured using a ARES Rheometry with a 25 mm diameter parallel plate geometry and 1 mm

gap. The samples were first heated, between the parallel plates, to the desired temperature, which

was followed by a frequency sweep test. The angular frequency ranged from 0.1 to 100 rad/s.

Dynamic oscillatory tests were carried at a strain rate of 5% corresponding to the linear visco-

elastic zone. The samples were dried in a vacuum oven prior to the experiments. The

experiments were performed in a nitrogen environment to suppress thermo-oxidative

degradation.

Rheological experiments were also performed to study the isothermal crystallization kinetics of

AR-TPU, PR-TPU, TPU-GMS, TPU-NCl and TPU-NSi samples using the similar setup as

discussed above with the ARES rheometry. Time sweep experiments were carried out at a low

frequency of 1 Hz and a strain of 5%. For all the experiments, the samples were first put between

the rheometer plates at 230oC within a nitrogen environment in a convection oven to suppress

degradation. The samples were kept at a temperature of 230oC for 3 min to eliminate the thermal

61

history of the TPU samples, and then the samples were rapidly cooled to the desired temperature

to perform the time sweep tests.

3.2.4 Atomic force microscopy

Atomic Force Microscopy (AFM) experiment was performed on samples using a Nanoscope

IIIA Multimode AFM machine. The data was collected in air in tapping mode using a diving

board TESP cantilever. The data were recorded as 512 x 512 pixel data sets at a scanning rate of

1 Hz. Prior to the AFM experiment, the samples were cryo-microtomed with a Leica UltraCut

UCT microtome machine using liquid nitrogen. The cutting speed was set to 5 mm/sec, and a

final 70 nm cut was used to get the surface.

3.2.5 Crystallization analysis of TPU at ambient pressure

To analyze the non-isothermal melt crystallization and isothermal crystallization behaviors of the

neat-TPU and TPU samples with additives (GMS, nano-clay and nano-silica) at ambient pressure

(1 bar) a regular Differential Scanning Calorimetry (DSC-Q2000) from TA Instruments was

utilized.

3.2.5.1 Non-isothermal melt crystallization analysis

The samples were heated to 230°C at a rate of 10°C/min and equilibrated for 10 min. Next, the

samples were cooled to -90°C at a rate of 10°C/min. Then the samples were reheated to 250°C at

a rate of 10°C/min. Similarly, to investigate the effects of the additives (GMS, nano-clay and

nano-silica) on the phase separation and crystallization behavior of HSs in the TPU

microstructure the melt crystallization behavior was analyzed with the method discussed above.

3.2.5.2 Isothermal crystallization analysis

The production of expanded TPU bead foams requires annealing of the material at elevated

temperatures, which would affect the HS crystalline domains. Hence to investigate the effect of

annealing, isothermal experiments were implemented at elevated temperatures under ambient

pressure (1 bar) using DSC. The samples were heated to the desired isothermal temperature at a

constant heating rate of 20°C/min. Then, the samples were annealed for 60 min. This was

followed by cooling at a rate of 20°C/min to -90°C. Subsequently a second heating step was

conducted at a rate of 10°C/min to 230°C to investigate the effects of annealing treatment.

62

3.2.6 Crystallization analysis of TPU at high-pressure with dissolved gas

As discussed earlier, during the production of expanded TPU beads, the TPU material is

annealed at elevated temperatures in the presence of gas. The dissolved gas causes swelling of

the TPU matrix, which increases the molecular chain mobility. This affects the thermal behavior

and the crystallization kinectics of the HSs chains in the TPU matrix. Overall the varying

crystallization kinetics at various pressures can significantly influence the final foam

morphology. Hence to investigate the effect of dissolved gas on the crystallization behavior of

the HS chains in the TPU matrix both the non-isothermal and isothermal crystallization behavior

of neat-TPU and TPU with different additives was investigated with CO2 and butane,

respectively.

3.2.6.1 Non-isothermal melt crystallization analysis in presence of CO2

To analyze the non-isothermal melt crystallization and isothermal crystallization behaviors of the

neat-TPU and TPU samples with additives (GMS, nano-clay and nano-silica) in presence of

dissolved CO2 a HP-DSC (NETZSCH DSC 204 HP, Germany) was utilized. The HP-DSC was

calibrated by measuring the melting points and heat of fusion for In, Bi, Sn, Pb, and Zn under

ambient and high CO2 pressures.

The samples were heated to 230°C at a rate of 20°C/min and equilibrated for 10 min in the

presence of high-pressure CO2. Next, the samples were cooled to 10°C at a rate of 20°C/min also

in the presence of high-pressure CO2. Finally, the samples were reheated to 250°C at a rate of

10°C/min in the presence of high-pressure CO2. The effects of varying CO2 pressure on the melt

crystallization behavior of the samples were investigated.

3.2.6.2 Isothermal crystallization analysis in presence of high-pressure CO2

To investigate the effect of dissolved CO2 on the isothermal crystallization behavior of AR-TPU,

PR-TPU, TPU-GMS, TPU-NCl and TPU-NSi, the samples were saturated at the desired

saturation temperature at elevated CO2 pressures by using HP-DSC (NETZSCH DSC 204 HP,

Germany). The samples were heated to the desired saturation temperature at a rate of 20°C/min

and equilibrated for 60 min. Next, the samples were cooled to 10°C at a rate of 20°C/min. Then,

CO2 was released at a very low-pressure drop rate to avoid any cell nucleation (and thereby to

63

avoid the influence of expansion on the crystallization of HS).Then, the samples were degassed

at room temperature for 2 hours and were reheated to 250°C at a rate of 10°C/min. Thereby, the

effects of isothermal saturation on the crystallization behavior of the samples were investigated

in the presence of dissolved CO2.

3.2.6.3 Isothermal crystallization analysis in presence of high-pressure butane

Our high pressure DSC (Netzsch DSC 204 HP) that has been used to study the effect of the

dissolved gas on the isothermal crystallization behaviours of polymers [39,40] could not

accommodate the liquid-state, high-pressure gas. So it was necessary to design a procedure to

study the effect of dissolved butane on the crystallization of TPU without using the high pressure

DSC.

To investigate the isothermal crystallization behaviour under butane pressure, the respective TPU

sample was saturated in the autoclave foaming chamber for 60 min at various annealing

temperatures (Figure 3.1). This procedure was similar to the foaming process, which will be

described in detail in Chapter 4. But, since expansion of bubbles may affect the TPUs

crystallization through biaxial stretching [41,42], foaming was completely prevented by rapidly

quenching the chamber in water before depressurization (i.e., after gas saturation). Then, butane

was released at a very low-pressure drop rate to avoid any cell nucleation (and thereby to avoid

the influence of expansion on the crystallization of HS). Then, the samples were degassed at the

room temperature for 48 hours and were heated from -90ºC to 230ºC at a heating rate of

10ºC/min in DSC. Thereby, the effects of isothermal saturation on the crystallization behavior of

both the AR-TPU and the PR-TPU were investigated in the presence of butane.

64

Figure 3-1 Schematic of the saturation setup with butane

3.2.7 Phase separation and crystallization analysis using X-ray diffraction

A significant problem in studies of urethane morphology is obtaining sufficient contrast in

electron micrographs to visualize the system. Since even highly phase separated systems will

possess regions of similar density, mass thickness contrast will be intrinsically low. Hence the

structure of HS crystallites have been successfully investigated by Wide-angle X-ray scattering

(WAXS) and Small-angle X-ray scattering (SAXS). Both the techniques were used in this thesis

to investigate the changes in the HS morphology and crystallinity.

3.2.7.1 Wide-angle X-ray scattering (WAXS)

The WAXS analysis was carried out using a Siemens D5000 diffractometer with Cu-Kα source

operating at 50 kV and 35 mA. The data was then processed by Siemens DiffracPlus software.

3.2.7.2 Small-angle X-ray scattering (SAXS)

The SAXS profiles were collected for samples in air and room temperature using a Bruker

SMART6000 CCD area detector with a Cu rotating anode source operating at 50kV and 90mA.

The average wavelength was 1.5418 Å. The sample to detector distance was approximately 300

mm. The raw frames were smoothed to correct for detector noise, and integrated into 1D profiles.

65

All the SAXS profiles presented have been masked in the low scattering vector region where

beam stop influenced the profile.

3.3 Results and Discussions

3.3.1 Rheological behavior of TPU and TPU nano-/micro-composites

The formation of the HS crystallite depends on the molecular mobility of the HS chains, which

are significantly affected by the viscosity [43,44]. For this purpose, the melt viscosity of both the

AR-TPU and the PR-TPU samples was measured at various conditions. Figure 3.2 shows the

viscoelastic behaviors of the AR-TPU and the PR-TPU samples at temperatures of 200°C and

210ºC, as a function of the frequency. As anticipated, both the samples showed a shear thinning

behavior. As shown in Fig. 3.2, the complex viscosity of the PR-TPU was lower than the AR-

TPU sample at both temperatures. The drop in the viscosity may be correlated with the decrease

of PR-TPU’s molecular weight and broader distribution of the HS chain segments caused by the

high mechanical shear and polymeric chain scission during the melt processing in the twin screw

extruder.

1 10 10010

1

102

103

AR-TPU- 200 C

PR-TPU- 200 C

Sh

ea

r V

isco

sity (

Pa

.s)

Frequency (rad/sec)

AR-TPU- 210 C

PR-TPU- 210 C

Figure 3-2 Complex shear viscosity plot of AR-TPU and PR-TPU

Figure 3.3 compares the viscoelastic behaviors of the AR-TPU and the PR-TPU samples with

respect to the presence of additives (1wt%GMS, 1wt% NCl and 1 wt% NSi). As shown in Fig.

3.3, the complex viscosity of TPU decreases in the presence of nano-/micron-sized additives.

The addition of GMS acts as a lubricant and hence decreases the viscosity of TPU. On the other

66

hand, the presence of nano-clay and nano-silica may affect the HS chain segment distribution

and the molecular weight resulting in the lower viscosity.

1 10 100

100

101

102

103

104

Com

ple

x v

iscosity (

Pa-s

)

Frequency (rad/sec)

AR-TPU

PR-TPU

TPU-1GMS

TPU-1NCl

TPU-1NSi

Temperature = 2100C

Figure 3-3 Complex shear viscosity plot of AR-TPU, PR-TPU, TPU-1GMS, TPU-1NCl and

TPU-1NSi

Time sweep experiments were used to study the change of crystallization under small-amplitude

oscillatory shear (SAOS) with small deformation. First, the samples were heated to 230°C and

equilibrated for 3 min to erase the thermal history. Next, the samples were cooled down rapidly

to 155°C and 165°C, respectively. After the sample reached the desired temperature, SAOS was

applied to study the quiescent crystallization. A constant frequency of 1 Hz with a strain of 1%

was selected to prevent non-linear viscoelasticity and disturbance of the evolving structures.

Figure 3.4 depicts the plots of the increase of the storage modulus (G’) during crystallization of

the AR-TPU and the PR-TPU samples. Initially, the G’ value remained constant in both the AR-

TPU and the PR-TPU samples due to less structural changes in the molten polymers. However,

at an intermediate times, the G’ values of the samples increased, which indicated a phase

transition related to the crystallization of HS. The onset of crystallization shows the properties of

the polymer to be changing from a liquid-like to a solid-like behavior as soon as the HS

crystallites became sufficiently interconnected. As shown in Fig. 3.4, the onset of crystallization

shifted clearly to an earlier time for the PR-TPU compared to the AR-TPU. Thereby, the low

viscosity can facilitate the mobility of the HSs in the PR-TPU to stack and crystallize with higher

perfection.

67

0.1 1 1010

0

101

102

103

104

105

AR-TPU-165 C

PR-TPU-165 C

G' (

Pa

)

Time (min)

AR-TPU-155 C

PR-TPU-155 C

Figure 3-4 Time sweep rheological curves of AR-TPU and PR-TPU

In the presence of additives (1 % GMS, 1 % NCl and 1% NSi), the onset of HS crystallization

further shifted to earlier time compared to PR-TPU and AR-TPU as shown in Fig. 3.5. The

decreased viscosities in presence of the additives would have further assisted the mobility of the

HSs in the TPU microstructure resulting in the observed increase in the viscosity during the time

sweep experiments.

1 10 10010

1

102

103

104

G' (

Pa

-s)

Time (min)

AR-TPU

PR-TPU

TPU-1GMS

TPU-1NCl

TPU-1NSi

Figure 3-5 Time sweep rheological curves of AR-TPU, PR-TPU, TPU-1GMS, TPU-1NCl

and TPU-1NSi

68

3.3.2 Atomic force microscopy

Figure 3.6a shows the AFM images of the AR-TPU sample, which reveals the presence of

spherical phase-separated HS crystalline domains (marked with a thick solid arrow). These HS

crystalline domains have a diameter of approximately 1 µm. This was because, the longer HS

chains could not stack to form crystallites with higher degree of perfection due to the low

molecular mobility in the AR-TPU [44,45]. As seen in Fig. 3.6b, the PR-TPU sample showed a

much smaller spherical phase-separated HS domains (marked with a thick solid arrow) ranging

between 300-500 nm in diameter, which confirms the breaking of the HS chains. However, a

high concentration of nano-sized fiber-like crystallites, marked with a thin solid arrow in Fig.

3.6b dispersed in the SS matrix was also observed. These HS crystallites are in the range of 200

to 500 nm in length. As discussed earlier, the processing provided a broad sequence distribution

of HS chains. However, the facilitated molecular mobility caused certain HS chains to stack and

crystallize with a higher perfection and thereby nano-scale crystalline structures were appeared.

Although the AR-TPU sample also showed these fiber-like crystallites marked with a thin solid

arrow in Fig. 3.6a, their concentration was much lower compared to the PR-TPU. Moreover,

both the AR-TPU and the PR-TPU showed HS spheres in the range of 50-100 nm marked with a

circle in Fig. 3.6a and Fig. 3.6b. These HS spheres are from the very short HS chains that can

hardly crystallize and hence lie dispersed in the SS matrix.

Figure 3.7 shows the AFM image of the PR-TPU sample after annealing at 160°C with the

presence of butane at a saturation pressure of 55 bar. It is clearly seen that compared to untreated

PR-TPU (Fig. 3.6b), the sample annealed with butane showed much larger HS crystallites with a

diameter of approximately 1 µm (marked with a thick solid arrow) and the mechanism for their

formation is discussed latter.

The AFM surface morphology for TPU-1GMS shows HS spherulites (marked with solid arrows

in Fig. 3.8) with approximately 500-1000 nm in diameter. The formation of α and β spherulites

have been reported in bulk morphology of TPUs as a result of high level of mobility of

MDI/BDO based HS chains due to the flexibility of BDO group[46-48]. The presence of GMS

would have assisted the HS chains to coil and fold and form spherulites, which is confirmed in

the DSC endotherms of TPU-GMS samples (Fig. 3.8).

69

Figure 3-6 AFM images: (a) AR-TPU, (b) PR-TPU; Scale: 5 μm side length in both

micrographs

Figure 3-7 AFM image of PR-TPU after saturating at 160°C with butane at 55 bar

pressure; Scale: 5 μm side length

Figure 3-8 AFM image of TPU-1GMS; Scale: 5 μm side length in the micrograph

(b

)

(a)

(b

)

70

3.3.3 Crystallization analysis of TPU at ambient pressure

3.3.3.1 Non-isothermal melt crystallization analysis with regular DSC

The isocyanate and hydroxyl end groups that form the characteristic urethane linkages in TPU

are stable in the solid state of the polymer. However, above a certain stability temperature in the

molten state, the urethane bonds start dissociating. This phenomenon is known as

“transurethanization” and it affects the sequence of the HS chain distribution [49,50]. For MDI

based TPUs, processing above 190-200 ºC increases the rate of trans-reactions [50,51].

Generally, the melt processing of MDI based TPUs leads to a polydisperse system with the

formation of both short and long HS chains. Upon cooling the melt of a polydisperse system, the

HS chains phase-separate into different crystalline domain sizes. A polydisperse distribution of

HS in TPU shows multiple melting endotherms, which can be attributed to different size and

various levels of packing order in the phase-separated HS domains and microcrystalline

structure. [50-52].

Figure 3.9 shows the cooling traces of the TPU samples. As seen, the PR-TPU showed a much

earlier on-set of the HS crystallization than the AR-TPU. These results suggest that in the PR-

TPU sample some of the HS with certain repeat units could have possibly crystallized much

faster at higher temperatures. Based on a widely accepted morphological model by Koberstein-

Stein, the most readily crystallizable sequence is estimated to be between two and four HS repeat

units for MDI/BDO based TPU [53]. However, at higher temperatures, the increased mobility

permits the crystallization of successively longer HS chains. The model also showed that BDO

residue could be present in various conformations (gauche and trans), and subsequently could

switch between these conformations at a rate that was determined by the temperature of the

sample. They concluded that the BDO residue could facilitate the folding or coiling of longer HS

chains to form crystalline lamellae. Based on this model, the widely distributed HS chains in the

PR-TPU may have better mobility to crystallize at the higher temperature with more closely

packed phase-separated HS domains, compared to the AR-TPU.

71

30 60 90 120 150 180 2100.5

0.6

0.7

0.8

0.9

1.0

1.1

Endo

183.5oC

12J/g

172oC

96.4oC

12J/g

12J/g

171oC

98.6oC

PR-TPU

AR-TPU

heating curve H

ea

t F

low

(J/g

)

Temperature (oC)

cooling curve

12J/gExo

Figure 3-9 DSC curves of the AR-TPU and PR-TPU samples

Figure 3.10 shows the DSC cooling curves for PR-TPU and TPU-GMSs samples. Overall,

addition of GMS significantly promoted the crystallization kinetics of HSs in the TPU

microstructure. In case of TPU-05GMS sample the crystallization temperature increased by

22.5°C. Further increase in GMS concentrations did not affect the crystallization temperature of

HS. Figure 3.10 also illustrates that PR-TPU and TPU-GMS samples showed double peaks in

their cooling curve. In case of PR-TPU, the high temperature peak (marked with hatched area in

Fig. 2) was observed as a shoulder peak and there was presence of sharper low temperature peak

at 98.5°C. With the addition of 0.5 wt% GMS, the high temperature shoulder peak observed in

neat-TPU changed to a broad exothermic peak. On the other hand, the low temperature peak

reduced significantly to form a shoulder peak (marked with solid dashed arrow in Fig. 3.10).

The mechanical shear from compounding would have broken the original sequence distribution

of HS chains into a much wider HS chain length distribution in both neat-TPU and TPU-GMS

samples due to the “transurethanization” reaction [54,55]. The HS chains phase-separate into

different domain sizes and microcrystalline structure upon cooling from the melt. However, as

discussed earlier some shorter HSs with repeat length below the critical phase-separation length

would remain dissolved in the SS microphase in the neat-TPU. On the other hand, in the case of

TPU-GMS samples, the lubricating effect of GMS may have promoted formation of a wide size

and perfection in HS domains due to the increase in the mobility of HS chains. The melting

endotherms confirmed the formation of wide size distribution of HS crystallites in TPU-GMS

72

samples as depicted in Fig. 3.11a. Both neat-TPU and TPU-GMS samples showed a broad

melting endotherm. However TPU-GMS (0.5 %, 1% and 2 %) showed formation of a new high

temperature shoulder peak above 180°C up to 210°C and the peak became more distinct with an

increase in the GMS concentration (hatched area in Fig. 3.11b). The highly perfected HS

crystallites, formed due to the lubricating effect of GMS, melt in the high temperature range

observed.

Figure 3-10 DSC cooling curves of PR-TPU and TPU-GMS samples

-100 -50 0 50 100 150 2000.0

0.2

0.4

0.6

0.8

1.0

172.8

171.4

74.0

74.4

TPU-05GMS

TPU-2GMS

TPU-1GMS

Heat flow

(J/g

)

Temperature (°C)

neat-TPU

72.9

70.4

171

170Endo

(a)

170 180 190 200 210 2200.05

0.10

0.15

0.20

0.25

184.3

197.4

184.1

204.7

205.9

185.4

Temperature (°C)

Heat flow

(J/g

)

neat-TPU

TPU-05GMS

TPU-1GMS

TPU-2GMS

183.9

(b)

Figure 3-11 DSC melting curves of PR-TPU and TPU-GMS samples: (a) regular plot, (b)

magnified plot for high temperatures

Figure 3.12a shows the DSC cooling curves for PR-TPU and TPU-NSi samples. Similar to the

PR-TPU samples, TPU-NSi showed double peaks in their cooling curve. The area of the high

temperature peak was observed to increase by increasing the nano-silica concentration. Thus at

73

higher nano-silica content, the HS crystallites could grow and stack into better perfection. The

melting of the bigger and/or highly perfected crystals was confirmed in the melting curve of

TPU-2NSi samples as shown in Fig. 3.12b, where a formation of a very broad high melting peak

was formed between 180°C to 210°C. In the case of addition of nano-clay to the TPU, the

crystallization was not significantly different compared to PR-TPU as shown in Fig. 3.13.

Figure 3-12 DSC curves of PR-TPU and TPU-NSi samples: (a) exotherms, (b) endotherms

Figure 3-13 DSC curves of PR-TPU and TPU-NCl samples: (a) exotherms, (b) endotherms

3.3.3.2 Isothermal crystallization with regular DSC

To investigate the effect of the annealing temperature on the HS crystallization at ambient

pressure (1 bar), an isothermal analysis was carried out using DSC. Figure 3.14 shows the DSC

endotherm of the PR-TPU after the isothermal treatments over a wide range of temperatures. It

should be noted that the AR-TPU showed a similar behavior and hence is not shown. Table 3.1

30 60 90 120 150 180 2100.0

0.1

0.2

0.3

0.4

0.5

TPU-2NSi

TPU-1NSi

98.5oC

12J/g

98oC

13J/g

98oC

13J/g

98.5oC

13J/g

He

at F

low

(J/g

)

Temperature (oC)

Cooling graphs: 10oC/min

Exo

PR-TPU

TPU-05NSi

(a)

30 60 90 120 150 180 210-0.6

-0.5

-0.4

-0.3

TPU-2NSi

TPU-1NSi

TPU-05NSi

2nd

Heating graphs: 10oC/min

Temperature (oC)

He

at F

low

(J/g

)

170oC

171oC

170oC

171oC

Endo PR-TPU

(b)

30 60 90 120 150 180 2100.0

0.1

0.2

0.3

0.4

0.5

TPU-2NCl

TPU-1NCl

TPU-05NCl

98.5oC12J/g

98oC

13J/g

98oC

13J/g

97oC

13J/g

Cooling graphs: 10oC/min

He

at F

low

(J/g

)

Temperature (oC)

Exo

PR-TPU

(a)

30 60 90 120 150 180 210-0.6

-0.5

-0.4

-0.3

TPU-2NCl

TPU-1NCl

TPU-05NCl

171oC

171oC

171oC

170oC

2nd

Heating graphs: 10oC/min

He

at F

low

(J/g

)

Temperature (oC)

Endo(b)

PR-TPU

74

summarizes the melting peaks and the heat of fusion (ΔHf) from the DSC experiments for both

the AR-TPU and the PR-TPU samples. It can be seen that the increased annealing temperature

shifted the low temperature melting peak (Tm-low) to higher temperatures generally 15-20°C

above the annealing temperature. Seymour and Copper [56] have reported this behavior in their

studies, and have related the shifting of the low melting peak to higher temperatures caused by

the growth or perfection of the smaller HS crystalline domains. At a lower annealing

temperature, the Tm-high1 and Tm-high2 melting peaks are not affected and are related to the melting

of larger and highly perfected HS crystalline domains. At an annealing temperature of 150ºC, Tm-

low merges with the Tm-high1. By further increasing the annealing temperature to 155°C, two

melting peaks began to appear as shown in Fig. 3.15. The Tm-high1 can be attributed to the melting

of the highly perfected HS crystals, which are formed due to the annealing condition. At the

same time, a new low melting peak (Tm-low), marked by arrow, is also generated as shown in Fig.

3.15. In this case, the annealing condition caused melting of the less perfect HS crystals, and the

molten HS chains re-crystallized when cooled to form the Tm-low melting peak. By increasing the

annealing temperature further to 160°C and 165°C, the Tm-high1 melting peak associated with HS

crystals, which are perfected from annealing shifts to 174.5°C and 179.8°C [57]. Thus, a new

higher melting peak is formed instead of the original melting temperature of the PR-TPU.

Although the Tm-high1 peak shifted to a higher temperature, the area under the peak decreased.

This was followed by an increase in the area of the Tm-low peak as shown in Fig. 3.15. The higher

portion of the HS crystals became molten and contributed to the formation of Tm-low melting peak

during cooling. In other words, at an increased annealing temperature, more of the original HS

crystals were melted and thereby more melt became available during the cooling, which

promoted the formation of the HS crystals with low perfection.

After annealing at 180ºC (Fig. 3.16), the Tm-high1 peaks of the AR-TPU and the PR-TPU samples

shifted to 195.2 ºC and 197°C, respectively. However, the Tm-high2 peak, which existed only in the

PR-TPU sample shifted to a very high temperature of 211°C. Thus, the PR-TPU sample had a

higher number of highly perfected HS crystals compared to the AR-TPU sample.

It is also important to observe that the total heat of fusion (ΔHT) of both the AR-TPU and the

PR-TPU samples decreased with an increase in the annealing temperature as seen in Table 3.1.

Hesketh et al. [58] studied a series of PUs after annealing at various temperatures between 120ºC

and 190ºC. They reported that the fraction of HS chains dissolved in the SS increased when

75

raising the annealing temperature. However, the ΔHT value of PR-TPU is higher in comparison

with the AR-TPU at higher annealing temperatures. The PR-TPU sample has a lower viscosity

compared to the AR-TPU sample, assisting the HS crystals that are formed during annealing to

stack into better perfection, which subsequently increases the ΔHT value.

50 100 150 200

-0.4

-0.2

0.0

0.2

Annealing

temperature

Tm-high2

Endo

PR-TPU

Tm-high1

Tm-low

180 C

160 C

150 C

140 C

100 C

120 C

80 C

60 C

w/o

annealing

He

at

flo

w (

J/g

)

Temperature C)

Figure 3-14 DSC endotherms of PR-TPU after annealing at various temperatures at

ambient pressure (1 bar)

50 100 150 2000.6

0.8

1.0

1.2

1.4Annealing Temp

Tm-low

PR-TPU

165 C

160 C

155 C

150 C

He

at

flo

w (

J/g

)

Temperature ( C)

Endo

Tm-high1

Figure 3-15 Effect of annealing at high temperature producing low temperature peak

(marked with an arrow) after cooling

76

50 100 150 200-0.6

-0.4

-0.2

13.31 J/g

15.38 J/g

197 C

195.2 C

211 C

AR-TPU

PR-TPU

He

at flo

w (

J/g

)

Temperature ( C)

Post annealing at 180 C

Endo

Figure 3-16 DSC endotherm of AR-TPU and PR-TPU after annealing at 180°C for 60 min

Table 3-1 Data of DSC measurements at ambient pressure (1bar)

AR-TPU PR-TPU

Anneal

Temp

(°C)

ΔHT

(J/g)

[SD]

Tm(oC) ΔHT

(J/g)

[SD]

Tm(oC)

TgTm-

low

Tm-

high 1

Tm-high 2 Tg Tm-low Tm-high 1 Tm-high 2

w/o

annealin

g

36.1

[0.20]-40.6

126.

0172.5 -

38.0

[0.11]-40.7 67.9 171.9 183.1

10020.2

[0.09]-45.1

126.

3174.1 -

17.6

[0.06]-42.9 118.5 171.1 183.1

12019.1

[0.02]-47.2

138.

1173.2 -

18.4

[0.07]-44.7 138.4 173.2 183.1

14017.8

[0.12]-40.1

155.

7174.7 -

19.0

[0.16]-43.8 155.9 174.8 183.1

15017.6

[0.20]-40.5

107.

7165.8 -

19.6

[0.14]-38.2 109.9 165.5 184.2

15519.6

[0.03]-35.9

114.

5170.1 -

19.8

[0.10]-42.8 118.3 170.1 184.4

16516.5

[0.14]-39.1

126.

3178.8 -

17.3

[0.20]-37.4 130.1 179.8 185.1

18013.3

[0.06]-

165.

9195.2 -

15.3

[0.10]- 163.4 197.9 211.1

At lower annealing temperatures, the presence of GMS, nano-clay and nano-silica did not affect

the isothermal crystallization behavior of TPU at ambient pressure. However at higher annealing

temperature (i.e. 180°C), the presence of GMS in particular affected the perfection in the HS

crystallites. Figure 3.17 compares the melting curve of PR-TPU and TPU-GMS samples after

annealing at 180°C. As shown the melting curve of PR-TPU consisted of three peaks. The first

77

peak is broad and has a maximum at 161.3°C. The second and third peaks are smaller but at a

much high temperatures of 197ºC and 211ºC. The broad melting peak is from the HS crystallites

formed during cooling. The two high temperature peaks are the melting of different sizes of

larger HS crystallites, which stacked into higher perfection due to the annealing. The TPU-

05GMS sample showed a broad melting peak with a maximum at 162.9°C. However the sample

showed three high temperature peaks at 189.1ºC, 199.4°C and 211.3°C respectively. Thus there

is presence of higher number of larger HS crystallites in TPU-05GMS samples. By increasing

the GMS concentration (TPU-2GMS), the broad melting peak shifted to approximately 16°C

above the melting peak observed in PR-TPU. On the other hand only one high temperature peak

is observed at 206.7°C. However the high temperature peak is larger compared to PR-TPU and

TPU-05GMS samples. Overall, the total heat of fusion (ΔHT) increased at higher concentration

of GMS.

70 140 210

-0.4

-0.2

15.4J/g

17.5 J/g

20 J/g

206.7 C

177.3 C

189.1 C199.4 C

211.3 C162.9 C

161.3 C

211 C

TPU-2GMS

TPU-05GMS

Heat flow

(J/g

)

Temperature (°C)

PR-TPU

197 C

Post annealing at 180 C

Endo

Figure 3-17 DSC endotherms of PR-TPU and TPU-GMS post annealing at 180°C

To investigate the effect of the annealing time on the HS crystallization at ambient pressure (1

bar), an isothermal analysis was carried out using DSC. Figure 3.18 shows the DSC endotherm

of the PR-TPU after the isothermal treatments over different annealing temperature and

annealing time of 60 min, 120 min and 180 min respectively. It should be noted that the AR-TPU

and TPU in presence of additives (TPU-GMS, TPU-NCl and TPU-NSi) showed a similar

behavior and hence are not shown. Overall by increasing the annealing time, the perfection

and/or the size of the HS crystallites increased, which resulted in increase in the Tm-high melting

peaks as shown in Fig. 3.18.

78

Figure 3-18 DSC endotherms of PR-TPU after annealing at different saturation

temperature and time

3.3.4 Crystallization analysis of TPU in presence of high-pressure dissolved gas

3.3.4.1 Non-isothermal melt crystallization analysis with high-pressure CO2

Figure 3.19 compares the non-isothermal melt crystallization behavior of PR-TPU at ambient

pressure (1 bar) and CO2 pressure (45 bar) with different cooling rates from the melt. Overall,

with a decrease in the cooling rate the crystallization temperature shifted to higher temperatures

for both the samples cooled in ambient pressure and in the presence of CO2 (1 bar).

Thus at lower cooling rate, the HS chains have longer time to stack and form crystallites with

better perfection. This behavior was observed in the samples cooled at ambient pressure with an

increase in the total heat of crystallization (ΔHC) (J/g). Similar behavior was observed for the

samples saturated with CO2 at 45 bar and cooled with 20oC/min, 10

oC/min and 5

oC/min.

However, when the samples were cooled at 2oC/min, the ΔHC was observed to decrease

compared to samples cooled at ambient pressure (1 bar).

Figure 3.20 compares the ΔHC of PR-TPU cooled with different cooling rates from the melt at

ambient pressure (1 bar) and different CO2 pressure’s. At 15 bar CO2 pressure, the samples

cooled with 50C/min and 20

0C/min showed a sudden increase in the ΔHC compared to those

cooled with 20C/min and 10

0C/min. With increase in the CO2 pressure to 30 bar, the ΔHC at all

the cooling rates decreased. However sample cooled with 200C/min showed highest decrease in

30 60 90 120 150 180 210 240-1.0

-0.9

-0.8

-0.7

-0.6

-0.5

-0.4

-0.3Iso 220

oC

Iso 180o

C

Iso 160o

C

He

at

flo

w (

J/g

)

Temperature (0C)

180 min

120 min

60 min

Iso 100o

C

Tm-high

79

the ΔHC value. Further increase in the CO2 pressure to 45 bar resulted in increase in the ΔHC at

all the cooling rates.

Figure 3-19 Non-isothermal melt crystallization behavior of TPU at different cooling rates:

(a) ambient pressure (1 bar), (b) CO2 pressure (45 bar)

Figure 3.21 compares the non-isothermal melt crystallization behavior of TPU with additives (1

% GMS, 1% NCl and 1 % NSi) at ambient pressure (1 bar) and in the presence of CO2 pressure

(45 bar). Overall, the presence of additives and high-pressure CO2 did not affect the

crystallization and phase-separation of HSs in the TPU microstructure. However in the case of

TPU-1GMS, the crystallization temperature slightly shifted further to higher temperature. Hence

the presence of GMS and the plasticization effect of CO2 may have further increased the

mobility of HS chains, which assists the chains to stack and form crystallites with higher degree

of perfection.

30 60 90 120 150 180 210 2400.0

0.3

0.6

0.9

20oC/min

10oC/min

5oC/min

2oC/min14J/g

13.5J/g

12J/g

11J/g

He

at F

low

(J/g

)

Temperature (oC)

Atmospheric pressure (1bar)(a)

30 60 90 120 150 180 210 2400.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

8.5J/g

9.5J/g

11.2J/g

20oC/min

10oC/min

5oC/min

2oC/min

He

at F

low

(J/g

)

Temperature (oC)

45 bar CO2 pressure

7J/g

(b)

80

Figure 3-20 Heat of crystallization of PR-TPU samples at different CO2 pressure and

cooled from the melt with different cooling rates

Figure 3-21 Non-isothermal melt crystallization behavior of TPU in presence of different

fillers and in the presence of CO2 pressure (45 bar)

3.3.4.2 Isothermal crystallization analysis with high-pressure CO2

Figure 3.22 depicts the HS crystal melting behavior of the PR-70A and PR-90A samples

saturated at ambient pressure (1 bar) and different CO2 pressures (28, 60 and 103 bar). The

saturation time was 30 min for both samples, and the saturation temperature was set 140°C for

PR-70A and, 160°C for PR-90A. Overall, in both the annealing settings, the sharp high

temperature melting peak (Tm-high) related to the melting of highly perfected HS crystals existed

in both the PR-70A and PR-90 samples. But the position of the Tm-high was affected differently

for both the PR-70A and the PR-90A samples with the change in the CO2 pressure. For the PR-

90A sample saturated at 28 bar and 60 bar, the Tm-high shifted to slightly lower temperatures due

to the plasticizing effect of CO2. However, by increasing the saturation pressure to 83 bar, the

0 10 20 30 40 50 602

4

6

8

10

12

14

16

18

20

Heat of C

rysta

llization (

Hc)

(J/g

)

CO2 pressure (bar)

2oC/min

5oC/min

10oC/min

20oC/min

PR-90A

30 60 90 120 150 180 2100.0

0.3

0.6

0.9

Temperature (oC)

He

at F

low

(J/g

)

TPU-1wt% GMS

TPU-1wt% NSi

TPU-1wt% NCl

45bar CO2

Atmospheric Pressure (1bar)

81

Tm-high increased by approximately 10°C to 182°C. Thus, at a higher saturation pressure, the

mobility of the existing HS crystalline domains may have increased, which may have assisted in

their higher degree of stacking and perfection, which resulted in a higher melting temperature.

The PR-70A sample depicted increase in the Tm-high at all the investigated CO2 pressures (28, 60

and 83 bar) as shown in Fig. 3.22b. On the other hand, saturation with CO2 induced a broader

low melting peak (Tm-low) in both the PR-70A and the PR-90A samples at all the pressures.

Although the Tm-high peak shifted to a higher temperature with an increase in the CO2 pressure,

the area under the peak decreased for both the PR-70A and the PR-90A samples. This was

followed by an increase in the area of the Tm-low peak as shown in Fig. 8. The higher portion of

the HS crystals became molten due to the plasticization effect of CO2 and contributed to the

formation of Tm-low melting peak during cooling. In other words, at an increased CO2 pressure,

more of the original HS crystals were melted and thereby more melt became available during the

cooling, which promoted the formation of the HS crystals with low perfection. Thus it can also

be concluded that CO2 induced a high degree of HS crystal nucleation in the TPU microstructure.

Further, there is a significant increase in the total crystallinity (Tcrys) of both the PR-70A and PR-

90A samples after annealing with CO2, compared to the annealing at ambient pressure (Fig.

3.22). This must have been due to the plasticization effect of the dissolved CO2, which may have

facilitated the HS chain mobility that further promotes the HS phase separation (i.e.

crystallization and crystal nucleation).

30 60 90 120 150 180-0.6

-0.5

-0.4

-0.3

-0.2

-0.1

0.0

0.1(a)

Tcrys

=8.5%

107.5153.6

1 bar

Heat flow

(J/g

)

Temperature (°C)

saturation time = 30 minsaturation temp = 140 C

PR-70A

Tcrys

=18.8%

73.5118.6

160.883 bar

Tcrys

=19.8%

72.5118.3

158.5

60 barT

crys=23.2%

73.7

155.6

28 bar

30 60 90 120 150 180 210 240

-0.6

-0.5

-0.4

-0.3

-0.2

-0.1

0.0

0.1

0.2(b)

Tcrys

=15.0%

133.0

174.7

1 bar

He

at

flo

w (

J/g

)

Temperature (°C)

saturation temp = 160 Csaturation time = 30 min

PR-90A

Tcrys

=28.5%

70.0

Tcrys

=29.5% 151.3

182.0

83 bar

60 bar

174.2

133.4Tcrys

=32.3%

72.0

69.7

133.8

173.128 bar

Figure 3-22 DSC melting endotherms after annealing over a range of CO2 pressures at a

fixed saturation temperature and time for (a) PR-70A and (b) PR-90A

Figure 3.23 compares the melting behaviors of the PR-70A and PR-90A after the samples were

isothermally treated over a range of temperatures with the presence of CO2 (60 bar). Overall, for

82

both PR-70A and PR-90A samples, the formation of double melting peak shifted to a lower

annealing temperature due to the plasticization effect of CO2. Furthermore, there is a significant

increase in the total crystallinity (Tcrys) of the HS in both PR-70A and the PR-90A after

saturating with CO2.

30 60 90 120 150 180 210-0.5

-0.4

-0.3

-0.2

-0.1

0.0

(a) saturation pressure = 60 bar

Tcrys

=20.8%

Tcrys

=20.9%

Endo

PR-70A

118.3 C

139.1 C

68.8 C

122.6 C

68.7 C

He

at

flo

w (

J/g

)

Temperature (°C)

Annealing

temp

140 C

120 C

100 C

Tcrys

=18.5%

72.5 C

158.5 C

saturation time = 30 min

30 60 90 120 150 180 210 240

-0.5

-0.4

-0.3

-0.2

-0.1

0.0

0.1

(b)

170.2 C

Tcrys

=28.8%

Tcrys

=27.2%

Tcrys

=25.3%

155.4 C

70.1 C

72.0 C

133.0 C

173.2 C

He

at

flo

w (

J/g

)

Temperature (°C)

Annealing

temp

165 C

160 C

140 C

PR-90A

Endo

55.5 C

142.9 C

177.1 C

saturation time = 30 minsaturation pressure = 60 bar

Figure 3-23 DSC melting endotherms after annealing at 60 bar CO2 pressure for 30 min at

a range of saturation temperatures for (a) PR-70A and (b) PR-90A

Three different saturation times were investigated, and the results are shown in Fig. 3.25.

Overall, with an increase in the saturation time, the unmelted HS crystalline domains in both the

PR-70A and PR-90A samples rearranged to form more perfect crystals with a higher melting

temperature. It was also observed that the Tcrys of both the PR-70A and PR-90A samples

increased with a higher saturation time.

0 30 60 90 120 150 180 210 240-0.3

-0.2

-0.1

0.0

0.1

0.2

0.3

0.4(a)

Tcrys

=24%

Tcrys

=20.3%

Tcrys

=20.1%

139.1

141.9

142.7

68.5

68.8

PR-70A

120 min

60 min

30 min

Heat flow

(J/g

)

Temperature (°C)

saturation temp = 120 Csaturation pressure = 60 bar

69.4

0 30 60 90 120 150 180 210 240

-0.3

-0.2

-0.1

0.0

0.1

0.2

0.3

0.4(a)

133.0

120.0

121.5

Heat flow

(J/g

)

Temperature (°C)

saturation temp = 160 Csaturation pressure = 60 bar PR-90A

120 min

60 min

30 min

Tcrys

=32.4%

68.8

Tcrys

=30.9%

175.467.5

Tcrys

=27.2%

174.172.0

173.2

Figure 3-24 DSC melting endotherms after annealing over a range of saturation times at a

fixed saturation pressure and temperature for (a) PR-70A and (b) PR-90A

3.3.4.3 Isothermal crystallization analysis with high-pressure butane

Figure 3.25a compares the melting behaviors of the AR-TPU and the PR-TPU after the samples

were isothermally treated at 165°C at ambient pressure (1 bar), and with the presence of butane

83

(55 bar). Overall, in both the annealing settings, a sharp Tm-high1 melting peak existed in the AR-

TPU and the PR-TPU samples. But after annealing in the presence of butane, the Tm-high1 shifted

to a lower temperature by approximately 5°C to 6°C for both samples due to the plasticizing

effect of butane.

On the other hand, saturation with butane induced a broader low melting peak (Tm-low) in both

samples. Further, there is a significant increase in the heat of fusion (ΔHTm-low) values of both

samples, after annealing with butane compared to annealing at ambient pressure as shown in Fig.

3.25a. There is also a significant increase in the total heat of fusion (ΔHT) value after annealing

with butane compared to annealing at ambient pressure. This must have been due to the

plasticization effect of butane, which may have facilitated the HS chain mobility (i.e., HS

flexibility) that further promotes the HS phase separation (i.e., the crystallization). However, the

PR-TPU sample showed a much higher increase in the ΔHT value compared to the AR-TPU

sample over a wide range of annealing temperatures as seen in Fig. 3.25b. By increasing the

butane pressure to 103 bar (Fig. 3.25b), the ΔHT value of the PR-TPU was further increased, due

to the increased flexibility of the HS chains to form crystalline domains.

The heat of fusion related to the Tm-high1 melting peak (ΔHTm-high1) was roughly estimated as

shown by the area marked with the dashed lines in Fig. 3.25a. Overall, the ΔHTm-high1 of both the

AR-TPU and the PR-TPU samples increased after annealing with butane (55 bar) compared to

annealing at ambient pressure (1 bar). However, the PR-TPU showed a much higher increase in

the ΔH Tm-high1 (Fig. 3.26) compared to the AR-TPU over a wide range of annealing temperatures.

Hence, the microstructure of the PR-TPU has a greater number of highly perfected HS crystals

compared to the AR-TPU.

84

-50 0 50 100 150 200 2500.0

0.2

0.4

0.6

0.8

1.0 1 bar (ambient)

Tm-low

Tm-low

HTm-high1

HT=16.5J/g

HT=40.4J/g

HT=17.3J/g

Tm-high1

Tg

173.8 C

174.2 C

Tm-high1

Saturation time= 60 min

HT=44.1J/g

Temperature ( C)

He

at flo

w (

J/g

)

Tg

Saturation temp= 165 C

PR-TPU

AR-TPU

55 bar (butane)

Endo

(a)

150 155 160 1650

10

20

30

40

50

AR-TPU (55 bar)

PR-TPU (55 bar)

AR-TPU (1 bar)

PR-TPU (1 bar)

T (

J/g

)

Saturation temperature ( C)

AR-TPU (103 bar)

PR-TPU (103 bar)

Saturation time= 60 min(b)

Figure 3-25 .(a) Comparison of DSC endotherm of AR-TPU and PR-TPU after annealing

at atmospheric pressure (w/o butane) and 55 bar butane; (b) total heat of fusion of AR-

TPU and PR-TPU after saturation with butane.

Figure 3.27 compares the Tg values of the AR-TPU and the PR-TPU after annealing at different

temperatures at ambient pressure, and in the presence of butane at 55 bar. Overall, the Tg of both

samples decreased with butane treatment as expected. Because of its plasticizing effect, the

dissolved butane must have increased the intermolecular distance. According to Table 3.2 and

Fig. 3.25, the HS crystallinity was significantly increased by gas dissolution for both melting

peaks at Tm-low and Tm-high1 due to the increased mobility. Overall, the SS purity may have

improved significantly due to the increased phase separation (i.e., the crystallization).

85

150 155 160 1650

3

6

9

12

15

18

Saturation time= 60 min

PR-TPU (1 bar)

AR-TPU (55 bar)

PR-TPU (55 bar)

Saturation temperature ( C)

m-h

igh (

J/g

)

AR-TPU (1 bar)

Figure 3-26 Comparison of DSC endotherm of AR-TPU and PR-TPU after annealing at

atmospheric pressure (w/o butane) and 55 bar butane

150 155 160 165

-70

-60

-50

-40

-30

-20

Tg

(C

)

Saturation temperature ( C)

AR-TPU (1 bar)

PR-TPU (1 bar)

AR-TPU (55 bar)

PR-TPU (55 bar)

Figure 3-27 Tg after annealing in ambient pressure (1 bar) and in the presence of butane

(55 bar).

Table 3-2 Comparison of PR-TPU’s DSC measurements at ambient pressure (1 bar) and

butane pressure (55 bar)

Anneali

ng

Temp

(°C)

Ambient pressure (1 bar) Butane pressure (55 bar)

ΔHTm-

low (J/g)

ΔHTm-

high1

(J/g)

Tm (°C)ΔHTm-

low (J/g)

ΔHTm-

high1

(J/g)

Tm (°C)

Tg Tm-low

Tm-

high1

Tg Tm-low Tm-high1

150 7.2 12.4 -38.2 109.9 165.5 30.3 15.3 -51.2 97.8 162.5

155 7.7 12.1 -42.8 118.3 170.1 34.9 12.9 -55.2 102.7 165.3

165 10.6 6.7 -37.4 130.1 179.8 33.8 10.3 -57.0 119.5 173.8

86

It is well known that annealing TPUs at lower temperatures at ambient pressure (1bar) increases

the size of the smaller HS crystallites and shifts the low temperature melting peak (Tm-low) always

~15-20°C higher than the corresponding annealing temperature. Figure 3.28 compares the

melting endotherms of the annealed PR-TPU and TPU-1GMS samples at 150°C under ambient

pressure (1 bar) and butane pressure of 55 bar. The saturation with butane induced a broad low

melting peak (Tm-low) in both samples. With the introduction of butane, not only a significant

increase in the heat of fusion of ΔHTm-low was observed (Fig. 3.28), there was also a significant

raise in the total heat of fusion (ΔHTot) in both the PR-TPU and TPU-1GMS samples. The

presence of dissolved butane facilitated the HS chain mobility (i.e., HS flexibility), improving

the HS crystallization kinetics. Furthermore, the presence of GMS together with butane caused a

much higher increase in the ΔHTot value of the TPU-1GMS sample, compared to that of the TPU

without GMS (Fig. 3.29).

-50 0 50 100 150 200 250

-0.4

-0.2

0.0

HTot

= HTm-low

+ HTm-high1

HTm-low

Tm-low

HTm-high1

Tg

Endo

neat-TPU

Tmhigh1He

at

Flo

w (

J/g

)

Temperature (°C)

butane pressure (55 bar)

ambient pressure (1 bar)

Tmhigh1

TPU-1 GMS

Tg

Tm-low

Figure 3-28 Comparison of DSC melting endotherm of PR-TPU and TPU-1GMS after

annealing at ambient pressure (1bar) and in the presence of butane (55 bar) at 150°C for

60 min

Figures 3.30a and 3.30b compares the ΔHTot and ΔHTm-low values of PR-TPU and TPU-1GMS

samples, treated isothermally at 150°C under ambient pressure (1 bar), and various butane

pressures. At a lower butane pressure (i.e., 23 bar), the ΔHTot of PR-TPU increased as compared

to annealing at ambient pressure (1 bar). However at higher butane pressures (53 bar and 103

bar), the ΔHTot of PR-TPU was not significantly affected. In contrast to PR-TPU, the ΔHTot of

TPU-1GMS sample continuously increased with an increase in the butane pressure, exhibiting

threefold raise at 103 bar. Similar trends were observed in the ΔHTm-low values of the PR-TPU

and TPU-1GMS samples as shown in Fig. 3.30b. The synergy caused by the lubricating nature of

87

GMS and the butane’s plasticizing effect may have induced a larger number of less perfect,

small-sized HS crystals during cooling, which resulted in the increase of the ΔHTm-low values of

TPU-1GMS and a higher ΔHTot values compared to PR-TPU.

150 155 160 1650

10

20

30

40

50

60

To

tal h

ea

t o

f fu

sio

n (

HT

ot)

(J/g

)

PR-TPU (55 bar)

TPU-1GMS (1 bar)

neat-TPU (1 bar) TPU-1GMS (55 bar)

Saturation temperature ( C)

Figure 3-29 ΔHTot of PR-TPU and TPU-1GMS after annealing at ambient pressure (1bar)

and in the presence of butane (55 bar) for 60 min over a range of annealing temperature’s

0 20 40 60 80 10015

20

25

30

35

40

45

To

tal h

ea

t o

f fu

sio

n (

HT

ot)

(J/g

)

Saturation pressure (bar)

PR-TPU

TPU-1GMS(a)

0 20 40 60 80 100

5

10

15

20

25

TPU-1GMS

m-low (

J/g

)

Saturation pressure (bar)

PR-TPU(b)

Figure 3-30 (a) Total heat of fusion (ΔHTot) of PR-TPU and TPU-1GMS over range of

butane pressure after annealing at 150°C for 60 min, (b) ΔHTm-low values of PR-TPU and

TPU-1GMS over range of butane pressure after annealing at 150°C for 60 min

The heat of fusion related to the Tm-high1 melting peak (ΔHTm-high1) was estimated as shown by the

area marked with the dashed lines in Fig. 3.28. Overall, annealing under butane pressure resulted

in a higher ΔHTm-high1 in both PR-TPU and TPU-1GMS samples compared to annealing at

ambient pressure (1 bar). However, the increase in the ΔHTm-high1 was more pronounced in TPU-

1GMS samples (Fig. 3.31) over a wide range of annealing temperatures. Hence, the

88

microstructure of the TPU-1GMS has a greater number of larger phase-separate HS domains as

well as highly perfected HS crystals as compared to PR-TPU. However, the Tm-high1 of both the

PR-TPU and the TPU-1GMS samples (Fig. 3.32) decreased at butane pressures beyond 23 bar

which suggests the formation of less perfected HS crystals due to the excessive plasticizing

effect of butane [59-61]. The maximum observed Tm-high1 depression was around 5-6°C under

103 bar and 55 bar butane pressures for PR-TPU and TPU-1GMS samples, respectively.

150 155 160 1650

3

6

9

12

15

18

PR-TPU (55 bar)

PR-TPU (1 bar)

TPU-1GMS (1 bar)

m-h

igh1 (

J/g

)

Saturation temperature ( C)

TPU-1GMS (55 bar)

Figure 3-31 ΔHTm-high1 of PR-TPU and TPU-1GMS annealed under ambient pressure and

butane pressure of 55 bar over a range of annealing temperature’s for 60 min

0 20 40 60 80 100165

170

175

180

185

TPU-1GMS

Tm

-hig

h1 (

oC

)

Saturation pressure (bar)

neat-TPU

Figure 3-32 The Tm-high1 variations of PR-TPU and TPU-1GMS samples versus butane

pressures saturated at 165°C for 60 min

Figure 3.33 compares the changes in the Tg of PR-TPU and TPU-1GMS samples after annealing

at 150°C at ambient pressure and various butane pressures. As expected, the Tg of both samples

decreased with butane treatment. Because of its plasticizing effect, the dissolved butane must

89

have increased the intermolecular distance. However, the TPU-1GMS sample showed much

lower Tg values compared to PR-TPU. According to Table 3.3 and Figs. 3.30, 3.31 and 3.32, the

HS crystallinity was significantly increased by gas dissolution for both melting peaks at Tm-low

and Tm-high1 due to the increased mobility. The crystallinity was further increased due to the

lubricating effect of GMS. Overall, the SS purity may have improved significantly due to the

increased HS phase separation resulting in lower Tg. It should be noted that this decrease in the

Tg is different from the Tg depression in the presence of the dissolved gas in which case the Tg

will decrease steadily with the dissolved gas content, i.e., the pressure [61].

0 20 40 60 80 100

-60

-50

-40

-30

Tg (

C)

Saturation pressure (bar)

neat-TPU

TPU-1GMS

Saturation temperature= 150 C

Figure 3-33 Tg of PR-TPU and TPU-1GMS after annealing at ambient pressure and

various butane pressures

Table 3-3Comparison of TPU-1GMS sample DSC measurements at ambient pressure (1

bar) and butane pressure (55 bar)

Anneali

ng

Temp

(°C)

Ambient pressure (1 bar) Butane pressure (55 bar)

ΔHTm-

low (J/g)

ΔHTm-

high1

(J/g)

Tm (°C)ΔHTm-

low (J/g)

ΔHTm-

high1

(J/g)

Tm (°C)

Tg Tm-low

Tm-

high1

Tg Tm-low Tm-high1

150 7.2 12.4 -38.2 109.9 165.5 30.3 15.3 -51.2 97.8 162.5

155 7.7 12.1 -42.8 118.3 170.1 34.9 12.9 -55.2 102.7 165.3

165 10.6 6.7 -37.4 130.1 179.8 33.8 10.3 -57.0 119.5 173.8

3.3.5 WAXS analysis

The XRD profiles of PR-TPU and TPU-1GMS are depicted in Fig. 3.34. While the PR-TPU

showed a broad, amorphous scattering halo, the TPU-1GMS sample showed a sharper peak halo

90

between 10° and 30° of 2θ. The previous XRD studies based on MDI/BDO TPU crystals have

shown similar peak [16].

0 10 20 300

1000

2000

3000

4000

TPU-1GMSIn

ten

sity (

a.u

)

2 (degrees)

PR-TPU

Figure 3-34 Comparison of XRD profiles of PR-TPU and TPU-1GMS

Figure 3.35 shows the diffracted X-ray intensity as a function of the scattering angle (2θ),

measured in the structural changes in the TPU-1GMS sample induced by annealing at 150°C at

ambient pressure (1 bar) and with butane saturation at pressures of 55 bar and 103 bar. As

discussed earlier, the plasticizing effect of butane tends to increase the molecular mobility of HS

chains. Thus, the HS chains stack into a more stable and perfected crystallites and hence the

overall crystallinity increases. It is evident from the traces that by increasing the butane pressure,

the sharpness and broadness of XRD halo increases (55 bar and 103 bar data are shown in Fig.

3.35). These results indicate on the development of HS crystallites, which occurred during the

saturation process in the presence of dissolved butane.

91

12 14 16 18 20 22 24 26 28 30 32 34 36 38 400

1000

2000

3000

4000

5000

1 bar

55 bar

Sca

tte

rin

g in

ten

sity

Scattering angle (2 )

103 bar

Figure 3-35 Comparison of XRD profiles of TPU-1GMS annealed at ambient pressure

(1bar) and various butane pressures at a saturation temperature of 150°C

3.3.6 SAXS analysis

Figure 3.36 shows the SAXS profiles of TPU-1GMS after the samples were annealed at 165°C at

ambient pressure (1 bar) and in the presence of butane at two pressures of 55 bar and 103 bar,

respectively. The horizontal axis represents the scattering vector (q) defined by Equation 3.1.

(Eq 3.1)

where, λ is the wavelength and θ is the scattering angle. The peaks were observed in all the

SAXS profiles. Based on the AFM phase image (Figs. 3.6, 3.7 and 3.8), the HS domains are

dispersed in the SS and hence these SAXS peaks are attributed to the inter-distance between the

HS domains.

Overall, the SAXS intensities increased and the width of the scattering curve decreased after

increase in the butane pressure compared to the case of annealing at ambient pressure. The

increase in the SAXS intensity with the increase in butane pressure signifies improvement in the

phase-separation of HSs within the TPU microstructure. The width of the scattering curve

decreases after annealing with butane, which is related to the increase in the average HS domain

size [62-64].

92

The inter-distance between the HS domains (d-spacing) was estimated by using Bragg’s

Equation, d=2π/qm. In the case of the TPU-1GMS sample annealed at ambient pressure (1 bar),

the maximum peak position (qm) was located at a value much below 1 A-1

. For the TPU-1GMS

sample annealed with butane (55 bar and 103 bar), the qm shifted to 2.1 A-1

and 1.9 A-1

,

respectively. For the TPU-1GMS samples annealed with butane (55 bar and 103 bar), the d-

spacing was 2.99 Å and 3.30 Å, respectively. Based on the qm value, the d-spacing for the TPU-

1GMS sample annealed at ambient pressure (1 bar) is much higher than the samples annealed

with butane. Since the volume fraction of HSs in all the samples is constant, the decrease in the

d-spacing data indicates that TPU-1GMS samples showed a higher degree of phase separation

and an increase in the domain size after saturation with butane. This observation supports the

formation of the Tm-low peak as shown in Fig. 3.28. The lubricating effect of GMS and

plasticization of butane resulted in a higher degree of phase separation of HS chains. The HSs

dissolved in the SSs matrix would have also phase-separated and formed HS domains.

1 2 3 4 5 6 7 8 9

0.0

2.0x104

4.0x104

6.0x104

8.0x104

1.0x105

1.2x105

1.4x105

1.6x105

1.8x105

qm (< 1 A

-1)-1 bar

qm (2.1 A

-1)- 55 bar butane

Inte

nsity (

I q,

a.u

.)

q (A-1)

qm (1.9 A

-1)-103 bar butane

Figure 3-36 SAXS profiles of TPU-1GMS samples after annealing at different pressure’s at

150°C

3.4 Conclusions

The phase separation and crystallization behavior of TPU is very sensitive to the processing

conditions. There has been extensive research work published in the literature regarding the

phase separation and crystallization behavior of TPU at atmospheric pressure (1 bar). However

there has not been any research work reported in the literature to investigate the effect of high-

93

pressure dissolved gas on the crystallization behavior of TPU. In this PhD work, for the first time

the crystallization behavior of TPU in the presence of dissolved gas has been systematically

investigated and published. The crystallization behavior of dissolved CO2 was investigated using

a HP-DSC. However, to investigate the effect of aliphatic hydrocarbon (butane), which cannot be

used in a HP-DSC, a specially designed high-pressure saturation system was developed. It was

observed that the presence of dissolved CO2 and butane induced a large number of less perfected

HS crystallites, which was a result of increase in the HS crystal nucleation mechanism during the

cooling from the annealing temperature after the completion of the annealing process. The

blending of GMS with TPU significantly improved the phase separation and crystallization

behavior of TPU. The presence of GMS acted as a lubricating agent and assisted the HS chains

to stack into higher degree of perfection and also assisted in the growth of HS crystallites to form

spherulitic crystals. Thus the overall crystallinity of the TPU was significantly improved after

annealing with CO2, butane and GMS. The presence of nano-clay and nano-silica did not

significantly affect the HS phase separation and crystallization behavior in the TPU

microstructure both independently and in synergy with dissolved gas. Another interesting

observation was the effect of melt-compounding, which resulted in the breakage of HS chains

and assisted the chains to stack and form HS crystallites with higher degree of perfection.

Overall the increase in the phase-separation and crystallization of HSs due to annealing with

dissolved gas and with GMS resulted in improved SS purity, which was observed with decrease

in the glass transition temperature. Thus the SS elasticity is also improved as a result the

annealing with dissolved gas and GMS compared to annealing at ambient pressure.

3.5 References

[1] Gibson PE, Wallace MA, Cooper AL. Development in Block Copolymer. London;

Elsevier; 1982.

[2] Cooper SL, Tobolsky AV. J Appl Polym Sci 1966; 10:1837-1844.

[3] Koberstein JT, Galambos AF. Macromolecules 1992; 25:5618-5624.

[4] Schollenberger CS. Polyurethane Technol 1969; 197-214.

[5] Scholl A, Wever W De. Kunststoffe 1998; 4:556.

[6] Yilgor E, Yilgor I, Yurtsever. Polymer 2002; 43: 6551-6559.

94

[7] Garrett JT, Runt J, Lin JS. Macromolecules 2000; 33: 6353-6359.

[8] Leung LM, Koberstein JT. Macromolecules 1986; 19: 706-713.

[9] Schneider NS, Paik Sung CS, Matton RW, Illinger JL. Macromolecules 1975; 8: 62-67.

[10] Seymour RW, Cooper SL. Macromolecules 1973; 6: 48-53.

[11] Coleman MM, Lee KH, Skrovanek DJ, Painter PC. Macromolecules 1986; 19: 2149-2157.

[12] Skrovanek DJ, Howe SE, Painter PC, Coleman MM. Macromolecules 1985; 18: 1676-

1683.

[13] Camargo RE, Macosko CW, Tirrell M, Wellinghoff ST. Polymer 1985; 26: 1145-1154.

[14] Brunette CM, Hsu SL, Rossman M, MacKnight WJ, Schneider NS. Polym Eng Sci 1981;

21: 668-674.

[15] Bonart R, Morbitzer L, Hentze GL. Macromol Sci, Phys 1969; B3: 337-356.

[16] Blackwell J, Lee CD. Adv. Urethane Sci Technol 1984; 9: 25-46.

[17] Koberstein JT, Russell TP. Macromolecules 1986; 19: 714-720.

[18] Benegir A, Michielsen S, Pourseyhimi B. J Appl Polym Sci 2009; 111: 1246-1256.

[19] Tocha E, Janik H, Debowski H, Vancso GJ. J Macromol Sci Phys 2002; B41: 1291-1304.

[20] Wagner KB. Macromolecules 1992; 25: 5591-5595.

[21] Miller JA, Lin SB, Hwang KKS, Wu KS, Gibson PE, Cooper SL. Macromolecules 1985;

18: 32-44.

[22] Martin DJ, Meijis GF, Gunatillake PA, Renwick GM. J Appl Polym Sci 1997; 63: 803-817.

[23] Briscoe BJ, Kelly CT. Polymer 1996; 37: 3405-3410.

[24] McBride JS, Massaro TA, Cooper SL. J Appl Polym Sci 1997; 23: 201-214.

[25] Wang C, Leung SN, Park CB. Ind Eng Chem Res 2010; 49(24): 12783-12792.

[26] Leung SN, Wong A, Wang C, Park CB. J Supercrit Fluids 2012; 63: 187-198.

[27] Wong A, Park CB. Chem Eng Sci 2012; 75(1): 49-62.

[28] Wong A, Guo Y, Park CB. J Supercrit Fluids 2013; 79: 142–151.

95

[29] Lips PAM, Velthoen IW, Dijkstra PJ, Wessling M, Feijen J. Polymer 2005; 46: 9396-9403.

[30] Mihai M, Huneault MA, Favis BD. Polym Eng Sci 2010; 50(3): 629-642.

[31] Taki K, Kitano D, Ohshima M. Ind Eng Chem Res 2011; 50: 3247-3252.

[32] Li YG, Park CB, Li HB, Wang J. Fluid Phase Equilibria 2008; 270: 15-22.

[33] Li YG, Park CB. Ind Eng Chem Res 2009; 48(14): 6633-6640.

[34] Park H, Park CB, Tzoganakis C, Tan KH, Chen P. Ind Eng Chem Res 2006; 45(5): 1650-

1658.

[35] Park H, Thompson RB, Lanson N, Tzoganakis C, Park CB, Chen P. J Phys Chem B 2007;

111: 3859-3868.

[36] Park H, Park CB, Tzoganakis C, Chen P. Ind Eng Chem Res 2007; 46: 3849-3851.

[37] Park H, Park CB, Tzoganakis C, Tan KH, Chen P. Ind Eng Chem Res 2008; 47(13): 4369-

4373.

[38] Liao X, Li YG, Park CB, Chen P. J Supercrit Fluids 2010; 55: 386-394.

[39] Nofar M, Zhu W, Park CB. Polymer 2012; 53: 3341-3353.

[40] Nofar M, Tabatabaei, Park CB. Polymer 2013; 54: 2382–2391.

[41] Zhai WT, Kuboki T, Wang L, Park CB, Naguib HE. Ind Eng Chem Res 2010; 20: 9834-

9845.

[42] Keshtkar M, Nofar M, Majithiya K, Park CB, Carreau P. “Extruded PLA/Clay

Nanocomposite Foams Blown with Supercritical CO2”. Macromolecular Bioscience,

submitted 2013.

[43] Biemond GJE, Feijen J, Gaymans RJ. Macromol Mater Eng 2009; 294: 492-501.

[44] Holden G, Legge NR, Schroeder HE. Thermoplastic elastomers: a comprehensive review;

Munich: Hanser Publishers; 1987.

[45] Armistead JP, Wilkes GL. J Appl Polym Sci 1988; 35: 601-629.

[46] Okamoto M, Nam PH, Maiti P, Kotaka T, Hasegawa N, Usuki A. Nanoletters 2001; 1: 295-

298.

[47] Yuan M, Song Q, Turng LS. Polym Eng Sci 2007; 47: 765-779.

[48] Martini-Vvedensky JE, Waldman FA, Suh NP. SPE-ANTEC 1982; 28: 674-676.

96

[49] Koberstein JT, Gancarz I, Clark TC. J. Polym Sci Part B- Polym Physics 1986; 24: 2487-

2498.

[50] Eisenbach CD, Nefzger H. Multiphase macromolecular systems, contemporary topics in

polymer science; New York: Plenium Publ Corp; 1989.

[51] Eisenbach CD, Baumgartner M, Guenter C. Advances in elastomer and rubber elasticity;

New York: Plenum Publ Corp; 1986.

[52] Gaymans RJ. Prog Polym Sci 2011; 36: 713-748.

[53] Koberstein JT, Stein RS. J Polym Sci, Polym Phys Ed 1983; 21: 1439-1472.

[54] Koberstein JT, Gancarz I, Clark TC. J Polym Sci Part B- Polym Phys 1986; 24: 2487-2498.

[55] Eisenbach CD, Nefzger H. Multiphase macromolecular systems, contemporary topics in

polymer science; New York: Plenium Publ Corp; 1989.

[56] Seymour RW, Cooper SL. Macromolecules 1973; 16: 48-53.

[57] Van Bogart JWC, Bluemke DA, Cooper SL. Polymer 1981; 22 :1428-1438.

[58] Hesketh TR, Van Bogart JWC, Cooper SL. Polym Eng Sci 1980; 20(3): 190-197.

[59] Hossieny NJ, Barzegari MR, Nofar M, Mahmood SH, Park CB. Polymer 2014; 55, 651-

662.

[60] Nofar M, Zhu W, Park CB. Polymer 2012; 53: 3341-3353.

[61] Nofar M, Tabatabaei A, Park CB. Polymer 2013; 54: 2382–2391.

[62] Van Bogart JWC, Bluemke DA, Cooper SL. Polymer 1981; 22:1428-1438.

[63] Van Bogart JWC, Gibson PE, Stuart L. J Polym Sci Polym Phys Ed 1983; 21: 65-95.

[64] Leung SN, Wong A, Park CB, Zong J. J Appl Polym Sci 2008; 108(6): 3997-4003.

97

Chapter 4 Foaming Behavior of TPU in Simulation Foaming Setup:- Effects of HS Crystallites, Nano-/Micro-Sized Additives, Blowing Agent

Types and Foaming Methods

4 Foaming Behavior of TPU in Simulation Foaming Setup

4.1 Introduction

In recent years, researchers have identified the effect of polymer crystallites on bubble nucleation

in foam processing - the crystalline domain in semi-crystalline polymers can promote cell

nucleation through local pressure variations [1-3] around the crystals [4-7] while the surrounding

area of newly formed (or growing) crystals have a supersaturated condition with the gas released

from the crystals, favorable for cell nucleation [7].

In this context, some studies have reported the crystallization kinetics of various polymers in the

presence of dissolved gas to understand better the role of crystals in foaming. The dissolved gas

causes swelling of the polymer matrix [8,9], which in turn increases the chain mobility, and

thereby, affects the surface tension [10-14], the viscosities [15-17], and the thermal behaviors

including the crystallization kinetics [18-20]. Therefore, the varying crystallization kinetics at

various gas contents can influence the final foam morphology.

Aliphatic hydrocarbons with a low boiling temperature (such as n-pentane, n-butane, etc.) are

commonly used blowing agents due to their high solubility and low diffusivity, which make

foam processing easier in industry [21,22]. Gendron et al. [23] reported production of

microcellular polycarbonate foams using n-pentane as the blowing agent, which must be due to

the crystals formed by the dissolved gas [24]. Tang et al. [25] produced microcellular foams

from blends of PP and PLA using n-pentane as the blowing agent. However, both the studies

reported non-uniform cell morphology with bimodal cell size distribution. In terms of the crystal

effect in polymer foaming, there has not been any study on the crystallization kinetics of a

polymer with the presence of dissolved butane.

98

There have been few reports on microcellular TPU foams [26,27]. Recently, Yeh et al. [28]

reported production of microcellular TPU foams using nanoparticles as bubble nucleating agent.

Nanoparticles behave as effective bubble nucleating agents [29-32] due to the existence of local

pressure variations [2-4] around the nanoparticles, and thereby they can be advantageous for

manufacturing of microcellular foams [33]. It should also be noted that the nucleation efficiency

of nanoparticles is highly dependent on the particle dispersion [29,30,34,35], the particle aspect

ratio [36], and the particle surface treatment [30]. In many instances, it was found that better

particle dispersion resulted in higher nucleation efficiency, higher cell density, and smaller cell

size. On the contrary, uniformly intercalated nanoclay particles with high rigidity turned out to

be more effective in cell nucleation than the well exfoliated nanoclay particles with pliability,

because of the higher local pressure variations around the more rigid nanoparticles [35]. Also,

the non-uniform dispersion of nanoparticles may lead to a bimodal cell size distribution [37].

Hence, incorporating nanoparticles as nucleating agents in foaming technology presents many

challenges.

In this chapter, different techniques of producing microcellular TPU bead foams in a simulation

bead foaming system is discussed. The results and processing parameters will be utilized to

manufacture TPU beads in a lab-scale autoclave bead foaming setup. In the first technique, we

present a novel technology to produce microcellular TPU foams by redistributing the HSs with

the dissolved butane without addition of any micro and/or nanoparticles as a bubble nucleating

agent. Moreover, we elucidated a wider processing window to produce microcellular TPU foams

using widely distributed nano-sized phase-separated HS crystalline domains. This was done by

reproducing the TPU material through a twin-screw extruder, which resulted in the breakage of

HS chains and hence formation of a broad distribution of HS crystalline domains. These well

distributed crystalline domains behaved as effective heterogeneous cell nucleating sites to

achieve microcellular TPU foams even at a moderate saturation pressure of butane. The foaming

was further improved by controlling the HS crystallites in the presence of GMS, which resulted

in a large number of highly perfected HS crystallites in the TPU microstructure. The third

technique discusses the effect of water, which is used to evenly distribute the heat during the

saturation step whole processing of TPU in lab-scale autoclave foaming process. It was observed

that water significantly plasticized TPU and results in high degree of perfection or growth of HS

crystallites. The growing HS crystallites in synergy with the presence of nano-clay increased the

99

heterogeneous nucleation rate and resulted in microcellular TPU nanocomposite foams at a very

mild processing condition with CO2.

4.2 Experimental Procedure

4.2.1 Materials

The TPU used in this study was Elastollan manufactured by BASF with a melting temperature of

171°C, a specific density of 1.13 g/cm3, and a hardness of Shore 90A. The HSs are composed of

reaction between MDI and BDO. The SSs are polyether diols, with a high hydrolysis resistance

tendency. Glycerol monosterate (GMS) used as a diffusion retarder and also an element which

modifies the crystallization behavior of HS was Pationic 915. The nano-clay used was Cloiste

30B. The N-butane and CO2, supplied by Linde Gas Canada, was used as the blowing agent.

4.2.2 Sample preparation

The “as-received” TPU material (AR-TPU) was compounded in a twin screw extruder DSM

Micro-compounder. Prior to compounding, the AR-TPU was dried in a CONAIR drier at 105°C

for 4 hours to remove moisture. The compounding was implemented at a processing temperature

of 190°C for 3 min and a rpm of 50. The samples of the extruded “processed” TPU (PR-TPU)

and the AR-TPU were used in the foaming experiments.

The AR-TPU and GMS were dry blended and then compounded in a twin screw extruder (DSM

Microcompounder). Prior to compounding, the “as-received” TPU was dried in a CONAIR drier

at 105°C for 4 hours to remove moisture. The compounding was implemented at a processing

temperature of 190°C for 3 min and a screw speed of 50 rpm. A series of TPU-GMS samples

with GMS contents of 0.5, 1 and 2 wt%, named as TPU-05GMS, TPU-1GMS and TPU-2GMS,

respectively, were prepared.

The AR-TPU and nano-clay (Cloisite 30B) were compounded in a twin screw extruder (DSM

Microcompounder). Prior to compounding, the AR-TPU was dried in a CONAIR drier at 105°C

for 4 hours to remove moisture. A series of TPU/nano-clay sample with nano-clay contents of

0.5, 1 & 2 wt% (TPU-05NCl, TPU-1NCl and TPU-2NCl) were prepared.

100

4.2.3 Butane sorption experiment

The sorption behaviors of butane in the AR-TPU and the PR-TPU samples were measured using

a Magnetic Suspension Balance (MSB) from Rubotherm GmbH. The swelling ratio was obtained

by means of an in-house PVT visualization setup [38]. A complete description of the

experimentation process is described elsewhere [39]. Approximately 0.3 g of a disk shaped

sample of TPU was placed in an aluminum container. The melt was heated at a desired

temperature (T) and in a vacuum (P=0), and a weight reading was obtained from the balance

readout W (0, T). Butane was then supplied at 20.7 bar to the system, and an appropriate time

was given to reach the saturation stage, after which another reading was obtained from the

balance readout, W (P, T).

4.2.4 Foaming setup and procedure

4.2.4.1 Foaming setup and procedure

The AR-TPU, PR-TPU and TPU-GMS samples were foamed in an autoclave foaming chamber.

The overall setup of the batch foaming process is shown in Fig. 4.1. Samples were first placed

inside the high-pressure vessel. The pressure vessel was then vacuumed to remove residual

moisture. Subsequently, butane was fed into the pressure vessel using a Teledyne ISCO high-

pressure syringe pump, and then maintained at a constant saturation pressure. The system was

heated to various saturation temperatures and kept for 60 min. The saturation temperature range

used in this study was between 150oC-170

oC and the selected saturation pressures were 53 bar

and 103 bar. The pressure was then rapidly released by opening a ball-valve connected to the

vessel and the chamber vessel was cooled in a water bath.

101

Figure 4-1 Schematic of the simulation foaming setup with butane

4.2.4.2 Foaming with CO2 and water

The processing of expanded polymer bead foams in an autoclave bead foaming setup requires a

media to evenly distribute heat to the material during the saturation process close to the melting

point the material. Water is a popular and cheap media to distribute the heat and avoid the

material from sticking together during the bead foaming process. However water may also affect

the foaming behavior of a material. In this setup the PR-TPU and the TPU-NCl nanocomposite

samples were foamed in an autoclave foaming chamber with the presence of water and CO2. The

overall setup of the batch foaming process is shown in Fig. 4.2. The system was heated to

various saturation temperatures and kept for 60 min. The pressure was then rapidly released by

opening a ball-valve connected to the vessel and the sample was foamed.

102

Figure 4-2 Schematic of the TPU and TPU nanocomposite foaming setup with water and

CO2

4.2.5 Foam characterization

The morphology of the foams was observed with a JOEL JSM-6060 scanning electron

microscope (SEM). The samples were fractured in liquid nitrogen, mounted on stubs, and sputter

coated with Au/Pd.

An image analysis on the SEM micrograph was conducted to obtain the average cell size and the

cell density using Image J (from the National Institute of Health). A micrograph showing more

than 100 bubbles was chosen, and the software determined the number of cells in the

micrographs. By analyzing the area of the micrographs, the cell density of each sample was

estimated using Equation 4.1. The density of the TPU foam was evaluated using a water-

displacement technique (ASTM D792-00). Using this information, the volume expansion ratio

(VER) of the samples was then evaluated as shown in Equation 4.2.

(Eq. 4.1)

(Eq. 4.2)

103

4.3 Results and Discussion

4.3.1 Sorption of butane in TPU

The increase in the molecular weight of the SS chains and the decrease in the concentration of

HSs affect the amount of gas impregnation of the TPUs [40]. The HS domains formed via

hydrogen bonding between the urethane groups have very limited diffusion of gas. Figure 4.3

compares the sorption of butane in the AR-TPU and the PR-TPU samples at temperatures of

150ºC and 190ºC. The saturation pressure of butane was 20.7 bar. Overall, the two TPU samples

showed a very similar sorption of butane at the investigated saturation temperatures.

0.000

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

190 CSo

lub

ility

(g

of b

uta

ne

/ g

of p

oly

me

r AR-TPU

PR-TPU

150 C

Figure 4-3 The solubility of butane in AR-TPU and PR-TPU at 20.7 bar

4.3.2 Effect of HS crystallites on foaming of TPU with butane

Figures 4.4 and 4.5 show the SEM micrographs of the foamed AR-TPU and the PR-TPU

samples saturated at different temperatures, and at saturation pressures of 55 bar and 103 bar,

respectively. As shown, the PR-TPU depicted a higher cell density than the AR-TPU in all the

foaming conditions. Moreover, the PR-TPU showed microcellular morphologies with fine cell

sizes in the range of 2 μm to 10 μm, and cell densities between 109

cells/cm3

and 1011

cells/cm3

as shown in Fig. 4.6. The microcellular morphologies in the PR-TPU were also observed over a

wider range of temperatures compared to the AR-TPU.

104

(a) 150

oC (b) 160

oC (c) 165

oC

(d) 150

oC (e) 160

oC (f) 165

oC

Figure 4-4 Foam morphology of TPU prepared at 55 bar and 150ºC, 160ºC, and 165ºC: (a),

(b), and (c) AR-TPU; (d), (e), and (f) PR-TPU; Scale bars: 10 µm

(a) 150

oC (b) 160

oC (c) 165

oC

(d) 150

oC (e) 160

oC (f) 165

oC

Figure 4-5 Foam morphology of TPU prepared at 103 bar and 150ºC, 160ºC and 165ºC:

(a), (b), and (c) AR-TPU; (d), (e), and (f) PR-TPU; Scale bars: 10 µm

One of the critical requirements for the production of microcellular foams is a very high degree

of thermodynamic instability generated from the super-saturation of a gas with a low solubility

[41]. It is very interesting to note, that the manufactured microcellular TPU foam was achieved

using butane that has a much higher solubility than the inert gas blowing agents such as CO2 and

N2 that are commonly used for microcellular foaming [39,42,43]. This indicates that despite the

lower volatility of butane itself, the generated large pressure variations around the formed HS

105

crystals with the help of dissolved butane must have generated enough thermodynamic

instability. Similar observations were made by Gendron et al. [23] from the polycarbonate foams

processed with pentane. We believe the local pressure variations [2, 4] around the crystals of the

polycarbonate matrix generated from a dissolved blowing agent [24] were responsible for the

microcellular cell density from this case.

The high cell density achieved in microcellular foams is largely determined by the cell

nucleation power during the foam processing. Nucleation is a classical phenomenon which

requires molecules to overcome an energy barrier and gather together (via local density and

energy fluctuation) to form embryos of the new phase. The free-energy barrier is often lower if a

bubble is nucleated on the surface of a second phase (heterogeneous nucleation as seen in

Equation 4.4), such as solid additives and impurities, when compared to the case where the

bubble is nucleated in the bulk phase of the polymer-gas mixture (homogeneous nucleation as

seen in Equation 4.3) [44-47].

(Eq. 4.3)

(Eq. 4.4)

where is the geometrical factor that relates to the surface geometry of the nucleating

agents, is the contact angle between the bubble surface and the solid surface measured in the

liquid phase. Based on the Equations 3 and 4, the increase in the saturation pressure, is expected

to reduce the free-energy barrier for bubble nucleation, and hence increase the bubble nucleation

rate. Thus, as anticipated with an increase in the saturation pressure to 103 bar, the cell densities

of both the AR-TPU and the PR-TPU increased significantly as shown in Fig. 4.6 (cell density

vs. T graph).

106

150 155 160 165 1700

5

10

15

20

25

30

35

40

45

50

PR-TPU (103 bar)

AR-TPU (103 bar)

PR-TPU (55 bar)

Ave

rag

e c

ell

siz

e (

m)

Temperature ( C)

AR-TPU (55 bar)(a)

150 155 160 165 17010

110

110

210

310

410

510

610

710

810

910

1010

1110

1110

12

PR-TPU (103 bar)

PR-TPU (55 bar)

AR-TPU (55 bar)

Ce

ll d

en

sity (

ce

lls/c

m3)

Temperature ( C)

AR-TPU (103 bar)

(b)

Figure 4-6 Characterization of AR-TPU and PR-TPU foams: (a) average cell size and (b)

cell densities

The free-energy barrier is also affected via localized stress variations ( , around the solid

fillers, which influence the degree of supersaturation [3].The presence of crystals also induces

stresses in the polymer matrix, which may cause bubbles to form via heterogeneous nucleation

mechanism (Equation 4.4). In the system consisting of dissolved butane in the TPU matrix (Fig.

4.7), the mobility of HS chains, including the HS crystalline domains, is increased due to the

swelling of the matrix [8,9]. Consequently, the HS crystalline domains are increasing, and the

surrounding area of the growing HS crystallites has a supersaturated condition with the gas

released from the crystals [7]. Further, the SS chains in the vicinity of these newly growing HS

crystallites are constrained because of the connection of surrounding molecules with the crystals

[35]. The SS chains that are constrained by the HS crystallites would generate locally varying

shear, compressive and tensile stresses. In the area where a tensile stress is applied to the SS

matrix, a negative occurs. This reduces the activation energy for cell nucleation, which

leads to the increase in the heterogeneous nucleation rate (Equation 4.4).

107

Figure 4-7 Schematic of TPU/butane morphology displaying the possible broad HS length

distribution

The cell nuclei density in the PR-TPU was observed 1-2 orders of magnitude higher than in the

AR-TPU. It is believed that two possible mechanisms are responsible for this difference. Firstly,

more broadly distributed HS domains in the PR-TPU would create a higher number of the

crystals, which can affect heterogeneous cell nucleation, compared to the narrowly distributed

HS domains in the AR-TPU, mathematically speaking. Secondly, the PR-TPU has a larger

number of well-dispersed fiber-like nano-crystalline HS with a large aspect ratio that melt at

much higher temperature according to Fig. 3.6b (Chapter 3). These nano-crystals would act as

heterogeneous cell nucleating agents in the PR-TPU sample [35].

Figure 4.8 shows the expansion ratio of the AR-TPU and the PR-TPU foamed samples. The

expansion ratios of both the AR-TPU and the PR-TPU increased as the temperature increased

from 150°C to 165°C at both saturation pressures. More crystals melted at a higher saturation

temperature, causing the SS to be more flexible, and hence, the TPU samples expanded easily.

This is similar to the case of high stiffness governing on the expansion ratio as the temperature

increases in the typical mountain shape observed in the extrusion foaming [48]. On the other

hand, at 170°C (55 bar), which is close to the melting point of the TPU, the expansion ratio of

AR-TPU was decreased, whereas the expansion ratio of the PR-TPU increased further. The large

number of fiber-like nano-crystalline HS domains of the PR-TPU shown in Fig. 3.6b, which

SS +

butane

SS +

butane

SS +

butane

SS +

butane

SS +

butane

108

seem to be highly perfected and thereby were not melted at 170°C, must have acted as a physical

branching network with high melt strength favorable for cell growth. In contrast, the AR-TPU

did not have these highly compacted HS domains, and therefore, a much smaller number of

crystals existed in the polymer matrix. Consequently, the volume expansion ratio decreased and

it must have been governed by the loss of the blowing agent at elevated temperature [48].

150 155 160 165 1700

1

2

3

4

5

6

7

8

9

10

PR-TPU (103 bar) AR-TPU (55 bar)

PR-TPU (55 bar)

Exp

an

sio

n R

atio

Temperature ( C)

AR-TPU (103 bar)

Figure 4-8 Expansion ratios of AR-TPU and PR-TPU foams

Figures 4.9 and 4.10 show the SEM micrographs and the cell density of the PR-TPU and TPU-

GMS samples, respectively, foamed after saturation at a pressure of 55 bar and at various

temperatures. Both samples showed microcellular morphologies, i.e., with cell sizes less than 10

μm and cell densities greater than 109 cell/cm

3. Moreover, with the addition of only 0.5% GMS

(TPU-05GMS), the foam morphology improved with the formation of fine cells in the range of

2-10 μm and an increase in the cell density above 1010

cells/cm3. At a saturation temperature of

170°C, the PR-TPU foams showed bigger cells (> 10 μm) with thinner cell walls, while the TPU-

GMS sample showed finer cells (≤ 10 μm) with thicker cell walls.

109

(a) 150°C (b) 160°C (c) 165°C (d) 170°C

(e) 150°C (f) 160°C (g) 165°C (h) 170°C

(i) 150°C (j) 160°C (k) 165°C (l) 170°C

Figure 4-9 Foam morphology under 55 bar butane pressure at different saturation

temperatures. (a-d) PR-TPU; (e-h) TPU-05GMS; (i-l) TPU-1GMS

150 155 160 165 17010

6

107

108

109

1010

1011

TPU-05GMS

TPU-1GMSCe

ll D

en

sity (

ce

lls/c

m3)

Saturation Temperature ( C)

PR-TPU

Figure 4-10 Cell densities of PR-TPU and TPU-GMS foams

During cell nucleation, molecules have to overcome the energy barrier to form embryos

of the new phase. From a thermodynamic perspective, the free-energy barrier is often lower if a

bubble is nucleated on the surface of a second phase (heterogeneous nucleation as seen in

Equation 4.4), such as solid additives and impurities [37,41-42].

110

where is the geometrical factor that relates to the surface geometry of the nucleating

agents, is the contact angle between the bubble surface and the solid surface measured in the

liquid phase. is expressed as seen in Equation 4.5, and considers the surface of the

nucleating agents to be rough due to formation of agglomerates [49]. Thus nucleating agents

such as inorganic fillers, organic phases and nanoparticles have been commonly employed to

reduce and induce a high degree of nucleation to achieve microcellular foams.

(Eq. 4.5)

The free-energy barrier is also affected via localized stress variations ( , around the solid

fillers, which influence the initial degree of super-saturation [49,50]. The presence of crystals

also induces around growing bubbles in the polymer matrix, which reduces the critical

radius and facilitates heterogeneous nucleation (Equation 4.4) [51].

In the TPU-GMS matrix consisting of dissolved butane and GMS, the mobility of HS chains

including the existing HS crystalline domains, is significantly increased due to the swelling of

the matrix [1,2]. Consequently, the phase separated HS crystalline domains are increasing, and

the surrounding area of the growing HS crystallites has a supersaturated condition with the gas

released from the crystals [1]. Further, the SS chains in the vicinity of these newly growing HS

crystallites are constrained because of the connection of surrounding molecules with the crystals

[13,50]. The SS chains could generate locally varying shear, compressive and tensile stresses,

causing potentially negative . This reduces the activation energy for cell nucleation,

which leads to the increase in the heterogeneous nucleation rate (Equation 5). The spherulites in

the TPU-GMS matrix would also reduce the nucleation energy due to its rough surface (Equation

5). Hence, the overall cell density in the TPU-GMS foams was higher than that in the PR-TPU

foams (Fig.4.10).

111

Figure 4.11 shows the expansion ratio of the PR-TPU and the TPU-GMS foams. In all the cases,

the expansion ratio increased as the temperature was increased from 150°C to 165°C. However,

the expansion ratio of the TPU-GMS foams was higher than that of the PR-TPU foams in the

temperature range of 150-160°C. The improved elasticity of the SS from better phase separation

of HS chains would have assisted with higher expansion of the TPU-GMS foamed samples

compared to the PR-TPU samples.

150 155 160 165 170

0

2

4

6

8

TPU-05GMS

TPU-1GMS

Exp

an

sio

n R

atio

Saturation Temperature ( C)

PR-TPU

Figure 4-11 Expansion ratios of PR-TPU and TPU-GMS foams

4.3.3 Foaming of TPU and TPU nano-clay nanocomposites with CO2 and water

4.3.3.1 Isothermal crystallization analysis with CO2 and water

To investigate the effect of dissolved CO2 and water (CO2+water) on the isothermal

crystallization behavior of TPU-1NCl, the sample was saturated at 150oC with 55 bar CO2

pressures in the high-pressure batch setup shown in Fig.4.2. The samples were heated to the

desired saturation temperature at a rate of 20°C/min and equilibrated for 60 min. Next, the

samples were cooled by quenching the chamber in water bath. Then, CO2 and water mixture was

released at a very low-pressure drop rate to avoid any cell nucleation (and thereby to avoid the

influence of expansion on the crystallization of HS).Then, the sample was degassed at room

temperature for 48 hours and was heated to 250°C at a rate of 10°C/min in DSC. Thereby, the

effects of isothermal saturation on the crystallization behavior of the samples were investigated

in the presence of dissolved CO2 and water.

112

Figure 4.12 compares the melting behaviors of the TPU-1NCl after the samples were

isothermally treated at 150°C at ambient pressure (1 bar), with the presence of CO2 (55 bar), and

with the presence of CO2 and water (CO2+water). It should be noted that the PR-TPU showed a

similar behavior and hence is not shown. Overall, in all the annealing settings, the Tm-high melting

peak existed in the TPU-1NCl samples. But after annealing in the presence of CO2, the Tm-high

shifted to a higher temperature by approximately 5°C to 6°C. On the other hand, saturation with

CO2+water, the Tm-high shifted further to a much higher temperature by approximately 11°C and

also a new high melting peak at 207.4°C was generated as shown in Fig. 4.12. Hence the

plasticization effect of water caused in the increased mobility of the existing un-melted HS

crystals and resulted in much better perfection. There is also significant increase in the total heat

of fusion (ΔHT) value after annealing with CO2 and also CO2+water compared to annealing at

ambient pressure. However the ΔHT of the sample annealed with CO2+water was slightly higher

compared to the sample annealed with CO2 as shown in Fig. 4.12. This must have also been due

to the plasticization effect of water, which may have facilitated the HS chain mobility (i.e., HS

flexibility) that further promotes the HS phase separation (i.e., the crystallization).

0 30 60 90 120 150 180 210 240-0.6

-0.4

-0.2

0.0

0.2

19.0 J/g

46.4 J/g

118.5

64.9

67.4

155.3 207.4

182.8

171.4

Saturation time= 60 min

55 bar- CO2+ water

55 bar- CO2

He

at flo

w (

J/g

)

Temperature (°C)

1 bar

Endo

Saturation temp= 150 C

165.6 (Tm-high

)

48.2 J/g ( HT)

Figure 4-12 Comparison of DSC melting endotherm of TPU-1NCl after annealing at

ambient pressure (1bar), in the presence of CO2 (55 bar) and in the presence of CO2 and

water at 150°C for 60 min

113

4.3.3.2 Isothermal crystallization analysis with CO2 and water

Figures 4.13 and 4.14 show the SEM micrographs of foamed PR-TPU and the TPU-1NCl

samples prepared at 150°C, and a saturation pressure of 55 bar with CO2 and CO2+ water. As

shown, both the PR-TPU and the TPU-1NCl samples depicted a higher cell density after foaming

with CO2+water compared to foaming with only CO2. Although, nano-clay behave as an

effective bubble nucleating agents [52], it is observed that the nucleation efficiency of the nano-

clay was not very high in the case of the samples foamed using CO2. However in the TPU-1NCl

system consisting of dissolved CO2 and water, the mobility of HS chains and the HS crystalline

domains, is increased due to the plasticization effect as discussed earlier. The HS crystalline

domains act as a heterogeneous nucleation site and increase the nucleation rate. This mechanism

is confirmed as observed with the increase in the cell density of PR-TPU as shown in Fig. 4.13b.

The nucleation rate is higher in TPU-1NCl caused due to synergistic effects of the nano-clay

particles and the HS crystalline domains acting as nucleation sites [53].

(a) (b)

Figure 4-13 Foam morphology of PR-TPU prepared at 55 bar and 150°C: (a) CO2 and (b)

CO2+water

(a) (b)

114

Figure 4-14 Foam morphology of TPU-1NCl prepared at 55 bar and 150°C: (a) CO2 and

(b) CO2+water

4.4 Conclusions

In this study, a novel technique of utilizing the HS domains in the TPU microstructure was used

to prepare microcellular TPU foams using butane as the foaming agent over a wide range of

foaming conditions. Since butane has not been used for microcellular plastics because of its low

volatility and high solubility, and thereby low thermodynamic instability generated from the

rapid solubility drop, it is interesting to note the microcellular cell nucleation induced with the

butane used in this study. Although butane generates a relatively low thermodynamic instability,

its impact on the crystallization caused microcellular nucleation. It was observed that the melt

processing of AR-TPU caused a breakage of the HS chains. Thus, the PR-TPU sample showed

broad distribution of HS domains, which also included some highly ordered HS nano-crystals

with very high melting temperature. Moreover, the saturation temperature and butane’s

plasticizing impact significantly induced larger content of HS domains with higher perfection in

the PR-TPU. Consequently, without addition of any nucleating agents, the cell nucleation was

promoted in the vicinity of the largely distributed and perfected HS domains over a wide

saturation temperature range of 150°C-170°C at a saturation pressure of 55 bar. Overall, the PR-

TPU showed very a high nucleation rate compared to the AR-TPU due to the presence of broad

HS domains in their microstructure.

The crystallization kinetics of TPU was significantly improved in the presence of GMS and

dissolved butane, which resulted in the formation of large number of less perfect HS crystallites

dispersed in the SS matrix whereas some highly perfected HS crystals are also formed. Unlike its

low volatility and high solubility, butane was successfully utilized in the fabrication of

microcellular TPU foams. This was facilitated through the impact of butane on the crystallization

of HSs. The HS crystallites acted both as heterogeneous nucleating sites as well as reinforcement

leading to the microcellular morphology with a high expansion ratio in TPU-GMS samples.

Consequently, without addition of any nucleating agents, cell nucleation was promoted in the

vicinity of the largely distributed and perfected HS domains over a wide saturation temperature

range of 150-170°C at a saturation pressure of 55 bar. Overall, the TPU-GMS showed very high

nucleation rates compared to the PR-TPU.

115

This study also investigated the effect of water and super-critical CO2 as co-blowing agents for

the production of PR-TPU and TPU nano-clay nanocomposite microcellular foams at a moderate

CO2 pressure of 55 bar and saturation time of 60 min. The cell density increased significantly

due to the synergistic effects of nano-clay particles and the HS crystalline domains acting as

bubble nucleation sites.

4.5 References

[1] Wang C, Leung SN, Park CB. Ind Eng Chem Res 2010; 49(24): 12783-12792.

[2] Leung SN, Wong A, Wang C, Park CB. J Supercrit Fluids 2012; 63: 187-198.

[3] Wong A, Park CB. Chem Eng Sci 2012; 75(1): 49-62.

[4] Wong A, Guo Y, Park CB. J Supercrit Fluids 2013; 79: 142–151.

[5] Lips PAM, Velthoen IW, Dijkstra PJ, Wessling M, Feijen J. Polymer 2005; 46: 9396-9403.

[6] Mihai M, Huneault MA, Favis BD. Polym Eng Sci 2010; 50(3): 629-642.

[7] Taki K, Kitano D, Ohshima M. Ind Eng Chem Res 2011; 50: 3247-3252.

[8] Li YG, Park CB, Li HB, Wang J. Fluid Phase Equilibria 2008; 270: 15-22.

[9] Li YG, Park CB. Ind Eng Chem Res 2009; 48(14): 6633-6640.

[10] Park H, Park CB, Tzoganakis C, Tan KH, Chen P. Ind Eng Chem Res 2006; 45(5): 1650-

1658.

[11] Park H, Thompson RB, Lanson N, Tzoganakis C, Park CB, Chen P. J Phys Chem B 2007;

111: 3859-3868.

[12] Park H, Park CB, Tzoganakis C, Chen P. Ind Eng Chem Res 2007; 46: 3849-3851.

[13] Park H, Park CB, Tzoganakis C, Tan KH, Chen P. Ind Eng Chem Res 2008; 47(13): 4369-

4373.

[14] Liao X, Li YG, Park CB, Chen P. J Supercrit Fluids 2010; 55: 386-394.

[15] Lee M, Park CB, Tzoganakis C. Polym Eng Sci 1999; 39(1): 99-109.

[16] Ladin D, Park CB, Park SS, Naguib HE, Cha SW. J Cell Plast 2001; 37(2): 109-148.

[17] Wang J, James DF, Park CB. J Rheology 2010; 54(1): 95-116.

116

[18] Naguib HE, Park CB, Song SW. Ind Eng Chem Res 2005; 44: 6685-6691.

[19] Nofar M, Zhu W, Park CB. Polymer 2012; 53: 3341-3353.

[20] Nofar M, Tabatabaei, Park CB. Polymer 2013; 54: 2382–2391.

[21] Yu L, Dean K, Li L. Prog Polym Sci 2006; 31: 576-602.

[22] Pillin I, Momtrelay N, Grohens Y. Polymer 2006; 47: 4676-4682.

[23] Gendron R, Daigneault LE. Polym Eng Sci 2003; 43(7): 1361-1377.

[24] Li YG, Park CB. J Appl Polym Sci 2010; 118(5): 2898-2903.

[25] Tang L, Zhai W, Zheng W. J Cell Plast 2011; 47 (5): 429-446.

[26] Dai X, Liu Z, Wang Y, Yang G, Xu J, Han B. J Supercrit Fluids 2005; 33: 259-267.

[27] Ito S, Matsunaga K, Tajima M, Yoshida Y. J Appl Polym Sci 2007; 106: 3581-3586.

[28] Yeh SK, Liu YC, Wu WZ, Chang KC, Guo WJ, Wang SF. J Cell Plast 2013,

DOI:10.1177/0021955X13477432.

[29] Han X, Zeng C, Lee LJ, Koelling KW, Tomasko DL. Polym Eng Sci 2003; 43(6): 1261-

1275.

[30] Zeng C, Han X, Lee LJ, Koelling KW, Tomasko DL. Advanced Materials 2003; 15(20):

1743-1747.

[31] Okamoto M, Nam PH, Maiti P, Kotaka T, Hasegawa N, Usuki A. Nanoletters 2001; 1:

295-298.

[32] Yuan M, Song Q, Turng LS. Polym Eng Sci 2007; 47: 765-779.

[33] Martini-Vvedensky JE, Waldman FA, Suh NP. SPE-ANTEC 1982; 28: 674-676.

[34] Lee YH, Wang KH, Park CB, Sain M. J Appl Polym Sci 2007; 103(4): 2129-2134.

[35] Wong A, Wijnands SFL, Kuboki T, Park CB. J Nanoparticle Res 2013; 15: 1815-1829.

[36] Chen L, Ozisik R, Schadler LS. Polymer 2010; 51 (11): 2368-2375.

[37] Zeng C, Hossieny N, Zhang C, Wang B. Polymer 2010; 51 (3): 655-664.

[38] Li YG, Park CB, Li HB, Wang J. Fluid Phase Equilibria 2008; 270: 15-22.

[39] Hasan MM, Li YG, Li G, Park CB. J. Chem Eng Data 2010; 55: 4885-4895.

117

[40] Briscoe BJ, Kelly CT. Polymer 1996; 37: 3405-3410.

[41] Guo Q. PhD thesis. Mechanical Engineering Department, University of Toronto; 2007.

[42] Li G, Gunkel F, Wang J, Park CB, Altstädt V. J Appl Polym Sci 2007; 103(5): 2945-2953.

[43] Li G, Wang J, Park CB, Simha R. J Polym Sci Part B- Polym Physics 2007; 45(17): 2497-

2508.

[44] Leung SN, Wong A, Park CB, Zong J. J Appl Polym Sci 2008; 108(6): 3997-4003.

[45] Leung SN, Wong A, Guo Q, Park CB. Chem Eng Sci 2009; 64: 4899-4907.

[46] Gibbs JW. The scientific papers of J. Willard Gibbs. In Thermodynamics, vol. 1;

Woodbridge, Connecticut: Ox Bow Press; 1993.

[47] Ward CA, Tucker AS. J Appl Phy 1975; 46: 233-238.

[48] Naguib HE, Park CB, Reichelt N. J Appl Polym Sci 2004; 91(4): 2661-2668.

[49] Tocha E, Janik H, Debowski H, Vancso GJ. J Macromol Sci Phys 2002; B41: 1291-1304.

[50] Wagner KB. Macromolecules 1992; 25: 5591-5595.

[51] Martin DJ, Meijis GF, Gunatillake PA, Renwick GM. J Appl Polym Sci 1997; 63: 803-

817.

[52] Cooper SL, Tobolsky AV. J Appl Polym Sci 1966; 10:1837-1844.

[53] Koberstein JT, Galambos AF. Macromolecules 1992; 25:5618-5624.

118

Chapter 5 Modification of Steam-Chest Molding Technology

5 Modification of Steam-Chest Molding Technology

5.1 Introduction

Expanded polymeric bead foams are popular materials used in packaging, thermal and sound

insulation applications [1, 2]. Expandable polystyrene (EPS), expanded polyethylene (EPE), and

expanded polypropylene (EPP) are widely used modern moldable bead foams. The successful

commercialization of EPP has caused the application of polymeric bead foams into more

advanced applications in areas such as automotive production [3]. Currently, there is an

increasing interest in investigating the processing behavior and mechanical properties of EPP 4-

8], because it has a higher service temperature and better mechanical properties compared to

those of EPS and EPE. In addition, EPP has some other advantages such as excellent impact

resistance, energy absorption, insulation, heat resistance, and flotation. Furthermore, it is

lightweight and recyclable, exhibits good surface protection and high resistance to oil, chemical,

and water. Due to these advantages, the use of EPP is gaining increased momentum in the

automotive, packaging, and construction industries [2, 4-8]. For instance, EPP molded foams are

utilized as bumper cores, providing significantly higher energy absorption upon impact as

opposed to conventional systems [3]. EPP bead foams have also been moving into more complex

applications in such areas as energy management, acoustic preference, and structural support [2-

9].

For all applications of EPP, the physical and mechanical properties of EPP bead foams are

influenced mainly by inter-bead bonding, because the bead boundaries usually develop into

fracture paths when a force is applied [10]. Inter-bead bonding is highly dependent on the

temperature of the medium transferring heat, and inter-bead bonding management is essential for

quality control [11].

Steam-chest molding technology is a commercially available and utilized high-temperature

steam to cause sintering of EPP beads. The processing steam temperature in a steam-chest

molding machine is coupled with the steam pressure [12]. The EPP bead foam has a high melting

119

peak of about 150-170ºC, and hence high steam temperatures and pressures are required for

processing, which causes a higher operating cost. The final physical and mechanical properties

of EPP molded product depend on the strength of the inter-bead bonding, which is significantly

affected by the molding conditions such as the steam pressure, steam temperature and molding

time. During processing, however, the steam pressure varies because of the resistance of the flow

through the beads, which makes it difficult to determine the actual temperature in the mold.

Moreover, considering the large volume and complicated shape of the mold cavity, the

temperature distribution across the mold cavity is not uniform. Nakai et al. [13] reported reduced

heat conduction to the core area of the mold caused by decrease in steam temperature due to

decrease in steam pressure. Zhai et al.[14, 15] also showed that both the degree of inter-bead

bonding and the tensile strength had a direct relationship with the steam pressure/temperature.

Other studies also reported that inter-bead bonding strength normally increased with the molding

pressure and time [11-16], and that improved the tensile and compressive strengths and fracture

toughness [17]. However, if beads are steamed for a too long time, their cell structure might

collapse [18]. Furthermore, a higher operating steam pressure relates to higher temperature

leading to an increase in localized temperature near the steam entry and hence beads exposed to

this high temperature may melt resulting in shrinkage at the surface of the product. This

dramatically deteriorates the surface property of the molded product.

In this study, the existing steam-chest molding machine was modified with the introduction of

hot air in an attempt to reduce the sensitivity of the decrease in the steam temperature with a

pressure drop. The hot air conditions were optimized using different critical parameters such as

the hot air flow rate, hot air temperature and hot air pressure. Also, the effects of adding hot air

on the process heating time, surface quality, thermal property, and tensile properties of the

molded EPP products are thoroughly investigated.

5.2 Theoretical Background

A double-peak melting behavior (Fig. 5.1) is required for EPP beads to have good sintering

during steam-chest molding. The hatched area in Fig. 5.1 represents the desirable steam

temperature range between the low and high melting peaks of EPP within the steam-chest

molding machine. When the foamed beads are processed in the steam-chest molding machine,

crystals associated with the low melting temperature (Tm-low) melt and contribute to the fusing

120

and sintering of individual beads. Meanwhile, the un-melted high melting temperature (Tm-high)

crystals help to preserve the overall cellular morphology of the bead foams [19]. A very narrow

processing window between the two melting peaks poses significant challenge in setting the

processing steam temperature. The steam temperature is sensitive and depends on the

corresponding steam pressure. The slight variation in steam temperature may cause the Tm-high

crystals to get affected and destroy the cellular morphology of the beads and hence cause

shrinkage of the molded product. The ratio between the low and high melting peaks is thus

crucial in determining the surface quality and mechanical properties of bead foam products [20].

The phenomenon of creating multiple crystal melting peaks for semi-crystalline polymers was

reported in earlier studies [21, 22]. The appearance of a new peak can be attributed to various

crystal structures, crystal sizes and their arrangement and perfection during the heating or

annealing treatments. In the case of EPP, the Tm-high peak originates from the perfection of

crystals during the gas-saturation stage in an autoclave at an elevated temperature around the

melting point (Tm) of polypropylene (PP) [14, 19]. The less perfect PP crystals are allowed to

partially melt and re-stack. During the prolonged gas impregnation stage, the remaining crystals

behave as crystallization nuclei that grow and become more perfect crystals due to the

rearrangement of the polymer molecular chains. The Tm-high melting peak is typically 15-20oC

above the annealing temperatures [14,19]. The Tm-low melting peak is created during the

subsequent foaming and rapid cooling stage.

30 60 90 120 150 180 210-1.4

-1.2

-1.0

-0.8

-0.6

-0.4

-0.2 actual variation of steam

temperature

Tm-high

Tm-low

He

at

Flo

w (

W/g

)

Temperature (°C)

Endo

Figure 5-1 Double-peak melting behavior of EPP foamed beads

121

The steam supplied for molding of EPP beads in steam-chest molding machine is in the

superheated state, which flows via small ports into the mold cavity. As the steam flows from the

surface toward the core of the cavity, its pressure decreases due to resistance from the EPP

beads. Overall, the steam temperature decreases with a decline in the pressure. Thus at a low

pressure, the superheated steam changes to saturated steam and finally condenses to saturated

water. The decrease in the steam temperature with the pressure decline can be estimated by

considering a throttling process caused by the resistance of the passage amongst the closely

packed beads. Both the fixed and moving mold in the steam-chest molding machine was

insulated, and hence for simplicity we can consider the process to be adiabatic. The decline in

steam temperature with a decrease in pressure can be described using Joule-Thompson

coefficient (μJ)[23]. Considering a steady-state throttling process with a steady flow across a

restrictor, Joule-Thompson coefficient is given by

(Eq. 5.1)

where T is the temperature, P is the pressure and denotes a partial derivative at constant

enthalpy. Goodenough computed values of μJ for superheated steam covering a wide range of

pressure and temperatures [24]. For a range of temperature between 121°C and 176°C and

pressure of 2.44 atm, the μJ of superheated steam was 13 °C/atm [25]. This range of pressure and

temperature is very similar to the actual condition of steam used for melting the Tm-low crystals

and create sintering of EPP bead foams in a steam-chest molding machine. As observed from μJ

of superheated steam, the steam temperature is very sensitive to the steam pressure and decreases

significantly with a decline in pressure. The actual steam temperature during sintering of EPP

beads in a steam-chest molding machine varies over a broad range between 120°C and 167°C

(Fig. 5.1) due to a decrease in the pressure. Due to the broad temperature range, proper heat

transfer would not occur in the core area of the molded EPP part, which results in poor sintering

of EPP beads and hence leads to poor mechanical properties. In order to maintain a high

temperature in the core area, an undesirable high steam pressure (i.e., high temperature) will

therefore be required on the surface. This will melt high temperature peak crystals (Tm-high) on the

surface, and thereby causes a non-uniform morphology with a high operating cost.

122

In our new design, we propose to use a mixture of steam and hot air to reduce the resultant μJ.

Compared to superheated steam, μJ of hot air at the same pressure and temperature is

significantly low at 0.01°C/atm [25]. Hot air can potentially present a very attractive, cost-

effective method to fundamentally reduce the sensitivity of a decrease in the steam temperature

with a drop in pressure. However, hot air is a very poor heat conductor and has a thermal

conductivity value of 36.6×10-3

W/m.C [26]. On the other hand, steam has a very high thermal

conductivity of 32.1×103 W/m.C.

27 Hence, using a mixture of hot air and steam can provide a

synergistic effect of low μJ of hot air and high thermal conductivity of saturated steam. Overall,

much superior heat transfer can be achieved at the core of the mold by supplying a mixture of hot

air and steam during the steam-chest molding of EPP beads. The introduction of hot air in the

steam-chest molding process may result in EPP bead products with improved surface quality,

enhanced mechanical properties and shortened cycle time resulting in a reduced operating cost.

5.3 Modifications on Steam-Chest Molding Machine to Incorporate Hot Air

The steam chest molding machine was modified to accomplish the following main functions: (i)

preventing steam condensate from entering the mold during the heating cycle, (ii) supplying hot

air into the steam injection pipe during the heating cycle, and (iii) monitoring the processing

temperature and pressure of steam and hot air mixture entering both the moving and fixed mold

channels. Figure 5.2 shows a schematic of the modified steam chest molding machine. To

facilitate the first function, i.e., preventing the steam condensate from entering the mold, the

steam supply piping was redirected to enter from the bottom of both the fixed and moving molds.

To maintain the steam above its condensation temperature, band heaters with proportional-

integral-derivation (PID) feedback control (Omega, CN7833) was located on the steam supply

piping and special heaters were inserted on the mold surface. Furthermore, all the exposed steam

supply piping and the metal surface of the fixed and moving molds were insulated. The steam

could be supplied in a wide range of pressure from 0 MPa to maximum working steam pressure

of 0.4 MPa. To supply hot air into the steam injection line, special heaters using coiled copper

piping were designed to heat the supplied compressed air. The compressed air could be heated to

approximately 200°C (T2 and T4 in Fig. 5.2). The compressed pressure could be controlled over

a wide range from 0 to 0.69 MPa using a pressure controller as seen in Fig. 5.2. The flow rate of

the supplied hot air could be varied from 0 to 120 l/min using a flow control valve as seen in Fig.

123

5.2. To achieve a good mixing of hot air and steam, hot air was introduced into the center of the

steam supply line to create annular flow of the steam into the hot air (marked with circle in Fig.

5.2). To monitor the processing temperature of steam and hot air mixtures, thermocouples were

located inside both the molds (T1 and T3 in Fig. 5.1). Similarly, pressure gauges were located to

monitor the pressure during processing (P1-P4 in Fig. 5.2).

Figure 5-2 A schematic of modified steam chest molding machine with hot air supply

5.4 Experimentation

5.4.1 Materials

The EPP beads, APPRO 5415 were supplied by JSP International. The beads have an expansion

ratio of 15 with bulk density of 60.9 g/L. The melting behavior of the EPP beads was examined

by DSC (TA Instruments, Q2000). The melting behavior of the beads showed a double peak

melting characteristics with a low melting (Tm-low) and high melting (Tm-high) peaks at 141.2°C

and 160.9°C, respectively (Fig. 5.1).

124

5.4.2 Steam-chest molding setup and experimental design

A laboratory-scale steam-chest molding equipment (DABO Precision, Korea) was used in this

study. The dimensions of the mold cavity were 15 cm × 6 cm × 5 cm. The steaming process in

the steam-chest molding included: 1) steam injection from the fixed side of the mold (1st

steaming cycle), 2) steam injection from the moving side of the mold (2nd

steaming cycle), and 3)

steam injection from both sides of the mold (3rd

steaming cycle). The 1st and 2

nd steaming cycles

were conducted to create the fusion between the EPP beads. The 3rd

steaming cycle was used to

remove pores on the surface of the molded EPP part. In the 1st steaming cycle, the steam was

flushed from the fixed side, passed through the bed of EPP beads in the mold cavity, and exited

from the moving side. During the 2nd

steaming cycle, the process is reversed and the steam was

flushed from the moving mold side. For the 3rd

steaming cycle, the steam was flushed from both

fixed and moving sides of the mold. The hot air was introduced during all three steam injection

cycles. For each set of experiments, the sample cooling time was remained unchanged.

To investigate and optimize the effect of hot air on the surface quality and the tensile properties

of the molded EPP, the temperatures of air and air flow rate were varied at three levels as shown

in Table 5.1. Since the inter-bead bonding usually increases with the steam pressure and the

heating time [11, 14, 16], the steam pressure was kept constant at 0.38 MPa (gauge pressure).

The unit of steam pressure/gauge pressure used in this study is the relative pressure in MPa,

which is 0.1 MPa lower than the absolute pressure. The corresponding steam temperature at the

gauge pressure of 0.38 MPa was 151°C from the steam table. The hot air pressure was also kept

constant at 0.41 MPa. Table 5.2 shows the complete experimental matrix.

To investigate the effect of air pressure on the surface quality and the tensile properties of the

molded EPP, the air pressure was varied at two levels of 0.41 MPa and 0.69 MPa as seen in

Table 5.1. At the lower air pressure, the air heaters have a higher possibility of getting damaged

and hence only two pressures could be investigated. The hot air temperature and the flow rate

were kept constant at 160°C and 80 l/min, respectively.

125

Table 5-1 Experimental parameters and design variables

Fixed Parameters Variables

Steam

pressure

(MPa)

Steam

temperature

(°C)

Air flow rate

(liters/min)

Air temperature

(°C)

Air pressure

(MPa)

0.38

151

80

100

120

110

160

200

0.41

0.69

Table 5-2 Experimental matrix

Run Air temperature

(°C)

Air flow rate

(l/min)

Steam

pressure

(MPa)

Hot air pressure

(MPa)

1 110 80

2 110 100

3 110 120

4 160 80

5 160 100 0.38 0.41

6 160 120

7 200 80

8 200 100

9 200 120

5.4.3 Surface quality characterization

Line scans were performed over 10 mm at six different locations on the fixed and moving

mold side of each sample (Fig. 5.3) using an optical profilometer (Nanovea ST 400,

Microphotonics Inc., Irvine, CA, USA) to measure the surface profile, and thereby, the surface

roughness of the samples. The locations on the moving mold side are designated by M1 to M6.

The sampling rate was 500 data points per mm in all the scans. ISO 4287 standard was adapted

in the calculations. The surface quality was characterized by the roughness values of Ra and Rz,

the waviness value of Wa, and the surface roughness profile. The roughness values of Ra and Rz

are calculated using Eqs. 5.2 and 5.3[28]:

(Eq. 5.2)

where, yi is the vertical distance from the mean line to the ith

data point. The roughness profile

contains n ordered, equally spaced points along the trace.

126

(Eq. 5.3)

Rz is the average distance between the highest peak and lowest valley in each sampling length. s

is the number of sampling lengths and Rti is Rt for the ith

sampling length. Surface profiles were

measured using similar line scans with 5 μm intervals between each line scan. In order to capture

the surface irregularities with spacing greater than the roughness sampling length (2 μm), the

waviness value (Wa) was used. Wa was calculated using the same equation as for Ra (Eq. 5.2) but

by using data over a wider sampling length (i.e., 20 μm) [29]. The morphologies of the molded

EPP samples were also observed by SEM (JEOL JMS 6060).

Figure 5-3 Rectangular area showing the location of line scans to characterize the surface

property on fixed mold and moving mold surface of molded EPP sample

5.4.4 Tensile property characterization

The tensile strength of molded EPP samples was measured using a Micro tester (Instron 5858) at

a crosshead speed of 5 mm/min. Rectangular specimens were cut from three different locations

across the thickness of the molded samples as shown in Fig. 5.4. Typical dimensions of the

specimens were as follows: thickness = 14 mm, width = 19 mm, and height = 155 mm. At least

five specimens were tested at each condition.

127

Figure 5-4 Schematic of specimen preparation for tensile tests

5.4.5 Thermal property characterization

The thermal history of molded EPP samples was analyzed by using Differential Scanning

Calorimeter, DSC (Q2000, TA Instruments), calibrated against characterized indium. The

thermal behavior was investigated at three different locations of the fix mold and moving mold

surface of the molded EPP samples. A temperature ramp process from 20°C to 230°C at a

heating rate of 10°C/min was carried out to investigate the melting behavior of the EPP samples.

The degree of crystallinity was calculated from the integration of the DSC melting peaks by

using 290 J/g as the heat of fusion (ΔHm) of 100 % crystallized PP [30].

5.5 Results and Discussion

5.5.1 Effect of hot air on the steaming time

In order to increase the productivity and reduce the operating cost, it is necessary to shorten the

processing time. One of the major impediments to shorten the processing time is the time

required to build-up the steam pressure to flow through the EPP beads in the mold during the

steaming cycles. The introduction of hot air may affect the build-up time of the steam pressure

and hence the overall steaming time, which will ultimately affect the processing time.

The steam pressure supplied to the equipment was 0.45 MPa. However, the desired processing

steam pressure (0.38 MPa) during the individual steaming cycle was controlled by using a

compound gauge. The gauge measured the pressure inside the mold cavity during the steaming

128

cycle using a pressure transducer and signaled to start the subsequent cycle after the desired

processing pressure was achieved. As discussed earlier the 1st and 2

nd steaming cycles were

crucial for the overall sintering of the EPP part, and hence, the time required to complete these

cycles with pure steam and steam mixed with hot air was recorded. The 3rd

steaming cycle was

set for 10 sec for all the experiments. The total steaming time to complete the molding of one

EPP part was calculated by adding the times for the three steaming cycles. Figure 5 depicts the

total steaming time for pure steam and steam mixed with hot air with various flow rates and

fixed temperature and pressure of 160 °C and 0.41 MPa, respectively. Overall, the total steaming

time decreased with the increase of the hot air flow rate and the highest air flow rate of 120 l/min

resulted in a time decrease of approximately 32 %. The reduction in the total steaming time

shows the effective use of hot air to reduce the overall processing time. The total steaming time

to complete the molding of one EPP part by increasing the hot air pressure to 0.69 MPa showed

similar result as observed at the hot air flow rate of 120 l/min.

56

60

64

68

72

76

Pure

steam

80 100 120

To

tal

ste

am

ing

tim

e (

se

c)

Hot air flow rate (l/min)

Figure 5-5 Effect of hot air and its flow rate on the total steaming time

5.5.2 Effect of hot air on the total processing temperature

Figures 5.6a and 5.6b compare the final processing temperatures after the completion of 1st and

2nd

steaming cycles for pure steam and steam mixed with hot air having various flow rates. The

processing temperature was measured at locations T1 and T3 as shown in Fig. 5.6c. Two

important observations can be made from the actual temperatures measured at T1 and T3 and

from the difference between the temperatures of T1 and T3 (ΔTcycle1 and ΔTcycle2 in Figs. 5.7a

and 5.7b). First, the introduction of hot air resulted in the decrease of ΔTcycle1 and ΔTcycle2. The

129

ΔTcycle1 and ΔTcycle2 values for EPP part molded with pure steam was 47°C and 48°C,

respectively. The introduction of hot air at low flow rate of 80 l/min resulted in ΔTcycle1 and

ΔTcycle2 values of 31°C and 22°C, respectively. By increasing the hot air flow rate, ΔTcycle1 and

ΔTcycle2 further decreased and at the highest flow rate of 120 l/min, the ΔTcycle1 and ΔTcycle2

reached 3°C and 4°C, respectively, which corresponded to roughly 94% and 92% reduction,

compared to those of pure steam. The high ΔTcycle1 and ΔTcycle2 values of pure steam suggest that

the process of expansion and sintering of EPP beads restricted the flow of steam and caused a

decrease in its pressure. Due to the high Joule-Thompson coefficient of steam, there was a

significant decrease in the steam temperature leading to very poor heat transfer across the mold.

With the introduction of hot air, however, the heat transfer across the mold improved

significantly and thus resulted in a more uniform temperature profile.

The second observation is that the localized source temperatures (T1 and T3) at the end of the

steaming cycles decreased with the introduction of hot air. It can be observed that after the

completion of the 1st and 2

nd steaming cycles with pure steam, the temperature at T1 and T3 was

167°C and 166°C, respectively, which were approximately 11.5°C and 10.5°C higher than the

supplied steam temperature of 155.5°C at 0.45 MPa. This can be understood considering the

quick sintering of EPP beads on the surface exposed to the high temperature steam. The sintered

beads started restricting the flow of steam and created a plugging behavior. Consequently, the

latent heat caused an increase of the steam temperature, and this further aggravated the

temperature gradient; i.e., further overheating on the surface whereas the core has not received

the enough heat to sinter each other because of the lowered temperature of the steam flowing at a

lower pressure.

But by introducing hot air, the temperature in the core could be maintained high to be able to

cause more uniform sintering across the thickness. So instead of causing a premature sintering on

the surface and an increase in the source temperatures (T1 and T3), the temperature of the beads

became more uniform. At a low hot air flow rate of 80 l/min, the temperatures at T1 and T3

decreased slightly to 164°C and 162°C, respectively. By increasing the hot air flow rate to 120

l/min, the temperatures further decreased to 151°C and 155°C, respectively. Thus with the

introduction of hot air, the flow of steam is improved significantly and the source temperature

increase (i.e., T1 and T3) due to the blockage of the flow could be successfully decreased.

130

120

130

140

150

160

170T

em

pe

ratu

re (

C)

Temperature T1

Temperature T3

Hot air flow rate (l/min)

Pure

steam80 100 120

(a) 1st steaming cycle

Tcycle1

Air temperature = 160 CSteam temperature = 151 C

110

120

130

140

150

160

170

(b)

Te

mp

era

ture

(C

)

Temperature T1

Temperature T3

Pure

steam

80

Hot air flow rate (l/min)

100 120

2nd

steaming cycle

Tcycle2

Steam temperature = 151 CAir temperature = 160 C

Figure 5-6 Effect of hot air and its flow rate on the processing temperature during (a)1st

steaming cycle and (b) 2nd steaming cycle. (c) A schematic illustrating the locations where

the processing temperatures of T1 and T3 were measured.

Controlling the pressure of the hot air resulted in a similar trend to the case of controlling the

flow rate of hot air. Figures 5.7a and 5.7b compare the final processing temperatures after

completion of the 1st and 2

nd steaming cycles for pure steam and steam mixed with hot air at two

pressures of 0.41 MPa and 0.69 MPa. The introduction of hot air at a pressure of 0.41 MPa

resulted in the decrease of ΔTcycle1 and ΔTcycle2 values by 53% and 58% to 31°C and 27°C,

respectively. By further increasing of the hot air pressure to 0.69 MPa, the ΔTcycle1 and ΔTcycle2

values significantly decreased and reached to only 2°C, which accounted for about 95%

reduction. The variation in the hot air temperature did not significantly change the processing

temperatures and hence is not discussed.

131

120

130

140

150

160

170

Te

mp

era

ture

(C

)

Pure

steam0.41 0.69

Hot air pressure, MPa

(a) Temperature T1

Temperature T31

st steaming cycle

Tcycle1

Steam temperature = 151 C

Air temperature = 160 C 120

130

140

150

160

170

Temperature T1

Temperature T3(b)

Te

mp

era

ture

(C

)

Pure

steam0.41

Hot air pressure, MPa

0.69

2nd

steaming cycle

Tcycle2

Steam temperature = 151 C

Air temperature = 160 C

Figure 5-7 Effect of hot air and its pressure on the processing temperature during (a) 1st

steaming and (b) 2nd steaming cycles

5.5.3 Effect of hot air flow rate on surface properties

Figure 5.8 compares the actual profile data from the line scans performed using the optical

profilometer on the EPP parts molded using pure steam and steam mixed with hot air. The data

was measured at six different locations (shown in Fig.5.3) on the surface of the molded EPP part

on moving mold side and designated from M1 to M6. As seen in Fig. 5.8a, for the EPP part

molded with pure steam, the variation of the line profile over the scan length of 10 mm spanned

a range of 195 μm between -120 μm and 75 μm. However, by introducing hot air, the variation

of line profile decreased significantly and spanned within a range of 75 μm between -50 μm and

25 μm (Fig. 5.8b). To obtain more quantitative information on surface quality, the surface

roughness (Ra and Rz) and waviness values (Wa) were calculated and the results are discussed.

Figure 5.9 compares the surface roughness values (Ra and Rz) of EPP parts molded using pure

steam and steam mixed with hot air. The roughness was measured for the surfaces of the molded

EPP part on the fixed and moving mold sides, since these surfaces are exposed to the steam and

hot air entrance ports on the molds. To investigate the effect of the hot air flow rate, the

temperature and pressure were kept constant at 160°C and 0.41 MPa, respectively. The flow rate

was varied from 80 l/min to 120 l/min. Overall, the introduction of hot air improved the surface

quality. It is observed that at a low hot air flow rate of 80 l/min, the surface roughness was not

significantly improved compared to the pure steam case with similar standard deviations. But by

132

increasing the hot air flow rate to 120 l/min, the surface roughness values decreased by

approximately 50 % reaching an Ra value of only about 1 μm, which is considered a soft touch

finish and thus a significant improvement. Furthermore, the high hot air flow rate resulted in

very similar surface roughness values on both the surfaces indicating improved uniformity in the

surface quality. Both Ra and Rz roughness values showed similar dependency on the hot air flow

rate.

0 2000 4000 6000 8000 10000

-100

-50

0

50

100

Lin

e p

rofile

(m

)

M1

M2

M3

M4

M5

M6

Line scan length ( m)

Pure steam (a)

0 2000 4000 6000 8000 10000

-100

-50

0

50

100

Steam + hot air

Line scan length ( m)

Lin

e p

rofile

(m

)

M1

M2

M3

M4

M5

M6

(b)

Figure 5-8 Comparison between actual line profile values measured over the scan length of

EPP parts molded with (a) pure steam and (b) steam mixed with hot air at 120 l/min

.

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

Hot air flow rate, l/min

Ro

ug

hn

ess v

alu

e [

Ra

] (

m)

Fixed surface

Moving surface

Pure

steam80 100 120

(a)

0

4

8

12

16

20

24

120

Fixed surface

Moving surface

Ro

ug

hn

ess v

alu

e [

Rz]

(m

)

Pure

Steam80 100

Hot air flow rate, l/min

(b)

Figure 5-9 Effect of hot air and its flow rate on (a) Ra and (b) Rz surface roughness

parameters

Figure 5.10 shows the waviness values (Wa) of the EPP parts molded using pure steam and steam

mixed with hot air. Similar to the roughness values, the EPP parts molded with mixture of steam

133

and hot air possessed lower waviness values. The waviness values decreased proportionally with

an increase in the hot air flow rate. The improvement in waviness is visualized in Fig. 5.11 by

the surface profiles. In both Figs 5.11a and 5.11b, the solid arrows show the surface topography

of a single bead. The surface height within a single bead of the samples molded with steam

spanned a range of 150 µm as shown in Fig. 5.11a. But with the introduction of hot air, the

surface height varied within a range of only 50 µm (Fig. 5.11b).

0

9

18

27

36

45

120 100 80

Wa

vin

ess v

alu

e [W

a]

(m

) Fixed surface

Moving surface

Pure

steam Hot air flow rate, l/min

Figure 5-10 Effect of hot air and its flow rate on the waviness (Wa) values of molded EPP’s

surface

µm

0

25

50

75

100

125

150

175

200

225

250

275

300

325

350

0 2000 4000 6000 8000 µm

µm

0

1000

2000

3000

4000

5000

6000

7000

8000

µm

0

25

50

75

100

125

150

175

200

225

250

0 2000 4000 6000 8000 µm

µm

0

1000

2000

3000

4000

5000

6000

7000

8000

(a) (b)

Figure 5-11 Fixed mold surface micro-topography of EPP bead molded products using (a)

pure steam and (b) steam mixed with hot air with an air flow rate of 100 l/min

Molded EPP samples produced using pure steam and steam mixed with hot air at

different flow rates were cut directly with a sharp knife. SEM micrographs of the cut surfaces of

134

the samples are shown in Fig. 5.12. The samples were prepared from the fixed mold side of the

molded part. The sample molded with pure steam showed a high degree of cell collapse in the

structure of the EPP beads at the surface, which caused formation of a thick skin (marked by

arrow) as shown in Fig.5.12a. Overall, the introduction of hot air reduced the cell collapse of

EPP beads. It is observed that at a low hot air flow rate of 80 l/min (Fig. 5.12b), the cell collapse

improved slightly compared to the pure steam case. But by increasing the hot air flow rate to 120

l/min (Fig. 5.12c), the cell collapse of EPP beads decreased significantly.

(a) (b)

(c)

Figure 5-12 SEM micrographs of the cut surfaces of fixed mold surface of EPP samples

produced using steam and steam mixed with hot air at different flow rates (a) pure steam,

(b) 80 l/min, and (c) 120 l/min

135

5.5.4 Effect of hot air temperature on surface properties

To investigate the effect of hot air temperature, the air pressure and flow rate were kept constant

at 0.41 MPa and 100 l/min, respectively, and three temperatures of 110ºC, 160ºC, and 200ºC

were investigated. Figures 5.13a and 5.13b show the Ra and Rz roughness values measured for

the surfaces of the molded EPP part on the fixed and moving mold sides. It can be noted that

varying the hot air temperature did not cause any significant change in the surface quality of the

molded parts. Compared to the roughness values of EPP molded part with pure steam, the

molded part with a mixture of steam and hot air at 110ºC and 160ºC showed slight improvement

in the overall surface property. However, at a higher temperature of 200 ºC, the overall Ra value

became more inconsistent.

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

Fixed surface

Moving surface

Pure

steam110 160 200

Ro

ug

hn

ess v

alu

e [

Ra

] (

m)

Hot air temperature, C

(a)

0

4

8

12

16

20

24

Fixed surface

Moving surface

Pure

steam110 160 200

Ro

ug

hn

ess v

alu

e [

Rz]

(m

)

Hot air temperature, C

(b)

Figure 5-13 Effect of hot air temperature on (a) Ra and (b) Rz surface roughness

parameters

5.5.5 Effect of hot air pressure on surface properties

To investigate the effect of the hot air pressure, the air temperature and flow rate were kept

constant at 160°C and 80 l/min, respectively, and two pressures of 0.41 MPa and 0.69 MPa were

investigated. Figures 5.14a and 5.14b show the Ra and Rz roughness values measured on the

fixed and moving mold side surfaces of the molded EPP part. It is seen that by introducing hot

air at a low flow rate of 80 l/min and a pressure of 0.41 MPa, the surface roughness became more

inconsistent with an increase by about 9% on the fixed mold surface. However, by increasing the

hot air pressure to 0.69 MPa, the surface roughness decreased by approximately 50% reaching an

Ra value of about only 1.12 μm and thus a significant improvement. Furthermore, a hot air

pressure of 0.69 MPa resulted in very similar surface roughness values on both the surfaces

136

indicating an improved uniformity in the surface quality. Both Ra and Rz roughness values

showed similar dependency on the hot air pressure.

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

R

ou

gh

ne

ss v

alu

e [

Ra

] (

m)

Air flow rate = 80 l/min

Air temperature = 160 C

Fixed surface

Moving surface

0.41 0.69

Hot air pressure, MPa

Pure

steam

(a)

0

4

8

12

16

20

24

R

ou

gh

ne

ss v

alu

e [

Rz]

(m

)

Pure

steam0.41 0.69

Hot air pressure, MPa

Fixed surface

Moving surface

Air flow rate = 80 lts/min

Air temperature = 160 C

(b)

Figure 5-14 Effect of hot air pressure on (a) Ra and (b) Rz surface roughness parameters

Figure 5.15 shows the waviness values (Wa) of the EPP parts molded using pure steam and steam

mixed with hot air at two different air pressures. It is seen that by introducing hot air at lower

pressure of 0.41 MPa, the waviness value (Wa) decreased by 9% and 45% at the fixed and

moving molds surfaces, respectively. By increasing the hot air pressure to 0.69 MPa, the

waviness value (Wa) showed a further decrease by 28% and 55% at the fixed and moving molds

surfaces, respectively. Furthermore, at both hot air pressures, both the surfaces indicated

significant uniformity in the surface quality.

0

9

18

27

36

45

Air flow rate = 80 l/min

0.69

Fixed surface

Moving surface

Pure

steam0.41

Hot air pressure, MPa

Wa

vin

ess v

alu

e [W

a]

(m

)

Air temperature = 160 C

Figure 5-15 Effect of hot air pressure on the waviness (Wa) values of molded EPP’s surface

137

5.5.6 Thermal properties of molded EPP samples

Figure 5.16 shows the DSC thermographs for the surfaces of the molded EPP part on the fixed

and moving sides of the mold. Table 5.3 also lists the melting behavior and crystallinity of the

molded EPP samples at different conditions. As seen in Fig. 5.16, the molded EPP samples

exhibited three melting peaks from high to low temperatures, denoted as Tm-high, Tmi, and Tmc,

respectively. The Tm-high was the original high melting peak of the EPP beads (Tm-high = 160.4ºC),

which remains constant at all the processing conditions. The lowest melting peak, Tmc in the DSC

curve (marked with arrow in Fig. 5.16) was reported to be the melting peak of crystals formed

during the cooling process [15]. Generally, this temperature was reported to be slightly lower

than the original low melting temperature of EPP beads (Tm-low =140.6ºC) [14], and a similar

behavior was observed in the melting endotherm of samples from the surface of the molded EPP

part on the fixed mold side in the presence of pure steam and steam mixed with hot air at

different conditions. The total crystallinity (XT) of the EPP samples molded with pure steam and

steam mixed with hot air is also shown in Table 5.3. It appears that the XT value decreased with

an increase in the hot air flow rate. The samples of the EPP part on the moving mold side showed

similar decreasing trend in XT after the introduction of hot air. The increase in XT is caused by

treatment at higher temperatures causing melting of original crystals and subsequent formation

during cooling [14]. The formation of crystals during cooling is related to Tmc [14]. The

crystallinity during cooling was estimated by the shaded area from the DSC plots (Fig. 5.16) and

their corresponding values (Xc) are listed in Table 5.3. Overall, Xc decreased with the

introduction of hot air with various flow rates. This further indicated that the melting of original

crystals was lower on the surface of samples molded using steam mixed with hot air.

The melting peak Tmi in the DSC curve (dashed line in Fig. 5.16) was reported to be created by

melting of the crystals that had possibly been induced by the fast heating and annealing treatment

that followed [14]. As seen in Table.5.3, Tmi decreased from 151.2ºC for pure steam to 148.4°C

with the introduction of hot air at the flow rate of 120 l/min. Another important observation is

that the Tmi peak was the weakest in the case of pure steam and became more pronounced with

the increase in the hot air flow rate for the surfaces of the EPP part, on both the fixed and moving

mold sides. This further confirms the decrease in the processing temperatures with hot air which

reduces the annealing temperature on the surface of the molded EPP part. Zhai et al. [14] also

found and reported that Tmi tends to become weak or even disappears at higher processing

138

temperature. They also reported that the Tmi was very sensitive and increased linearly with

increased treatment temperature. This strongly confirms that the improved surface quality seen

with an increase of hot air flow rate was due to the reduced local surface temperature, which

ultimately caused less melting of original crystals in the EPP beads.

The EPP beads molded at different hot air temperatures showed similar correlation between their

thermal behaviors and surface properties. As also discussed earlier, the hot air temperature does

not cause any significant effect on the surface properties and hence the thermal behaviors at

different hot air temperatures are not discussed in detail.

60 90 120 150 180

0

1

2

120 l/min

100 l/min

80 l/min

He

at

flo

w (

W/g

)

Temperature (°C)

Pure

steam

Fixed Mold Surface

Endo

(a)

60 90 120 150 180

0

1

2

He

at

flo

w (

W/g

)

Temperature (°C)

Pure

steam

80 l/min

100 l/min

120 l/min

Endo

Moving mold surface(b)

Figure 5-16 DSC thermographs of molded EPP samples (a) fixed mold surface and (b)

moving mold surface

Table 5-3 Melting points and crystallinity of molded EPP samples at fix and moving mold

surface at different processing conditions of pure steam and steam with hot air

Fixed Mold Surface Moving Mold Surface

Pure

steam

80

lts/min

100

lts/min

120

lts/min

Pure

steam

80

lts/min

100

lts/mi

n

120

lts/min

T mc (ºC) 139.3 139.9 138.9 139.1 142.4 141.9 138.2 138.6

T mi (ºC) 151.2 151.1 149.1 148.4 152.7 152.2 150.3 149.7

T m-high

(ºC)

159.7 159.4 159.3 159.2 159.6 159.1 159.4 159.4

XC (%)

[cooling]

25.7 28.2 21.5 20.1 29.0 27.1 24.8 19.7

XT (%)

[total]

38.0 37.2 35.5 34.6 38.0 38.1 34.6 31.1

139

5.5.7 Effect of hot air on tensile properties

As discussed earlier, the surface roughness and thermal property of the molded EPP parts

showed high sensitivity to the flow rate and the pressure of the hot air. The hot air temperature

however did not cause any significant change on the properties of the molded EPP part.

Figure 5.17 compares the tensile strength of EPP molded parts using pure steam and steam

mixed with hot air at different flow rates. The tensile strength was measured for samples from

the surfaces of the molded EPP part on the fixed mold side, center and moving mold side. The

tensile strength measured at the center of the sample molded with pure steam was approximately

12% and 20% lower than the corresponding values of samples of the molded EPP part on the

fixed and moving mold side. This was caused due to the reduced heat flow to the core of the

sample caused by decrease in the steam pressure. Overall, the introduction of hot air improved

the tensile strength in the center of the molded part. The tensile strength became very uniform

over the entire molded part with the introduction of hot air. At a hot air flow rate of 80 l/min, the

tensile strength of the samples from the surface of EPP part at the fixed and moving mold side

did not change much as compared to those of the part molded with pure steam. However, the

tensile strength in the center improved by approximately 20%. By increasing the hot air flow rate

to 100 l/min and 120 l/min, the tensile strength at the center increased by 14% and 16%,

respectively.

0.0

0.2

0.4

0.6

0.8

Fixed surface

Center

Moving surface

Hot air flow rate, l/min

Te

nsile

str

en

gth

(M

Pa

)

Pure

steam

80 100 120

Air pressure = 0.41 MPa

Air temperature = 160 C

Figure 5-17 Tensile strengths of molded EPP samples produced with pure steam and steam

mixed with hot air at different flow rates

140

Figure 5.18 compares the tensile strength of EPP molded parts using pure steam and steam

mixed with hot air at different air temperatures. Three temperatures of 110ºC, 160ºC, and 200ºC

were investigated. Overall, the tensile strength was seen to be consistent over the entire molded

part at all the investigated air temperatures. It can be seen that by introduction of hot air at

temperature of 100ºC, the tensile property of the samples from surfaces of the molded EPP part

on the fixed mold and moving mold side increased by 16% and 8 %, respectively as compared to

those of the parts molded with pure steam. With increase of the air temperature to 160ºC, the

tensile strength remained approximately unchanged, compared to the case of pure steam. Further

increase in the hot air temperature to 200ºC, did not change the tensile strength at the fixed mold

surface, but it decreased the tensile strength of the moving mold surface by approximately 10 %,

compared to the pure steam case. On the other hand, at all three hot air temperatures of 110ºC,

160ºC and 200ºC, the tensile property of the center of the molded sample increased by 27%, 14%

and 20%, respectively, compared to corresponding values of the pure steam samples. As

discussed earlier, the steam temperature did not play a major role in the overall quality of the

molded EPP parts. However, the observed improvement in the tensile property of the molded

samples in the center originated from the improved heat flow caused by the hot air flow rate.

0.0

0.2

0.4

0.6

0.8

T

en

sile

str

en

gth

(M

Pa

)

Fixed surface

Center

Moving surface

Pure

steam

110 160 200

Hot air temperature, C

Air flow rate = 100 l/minAir pressure = 0.41 MPa

Figure 5-18 Tensile strengths of molded EPP samples produced with pure steam and steam

mixed with hot air at different temperatures

Figure 5.19 compares the tensile strength of EPP molded parts using pure steam and steam

mixed with hot air at different pressures. Overall, the uniformity of the tensile strength across the

molded sample increased with the increase of air pressure. By introduction of hot air at pressure

of 0.41 MPa, the tensile strength at fixed mold surface, center and moving mold surface

141

increased by 4%, 6% and 20% respectively, compared to their corresponding values in pure

steam case. Further increase in hot air pressure to 0.69 MPa resulted in further improvement in

tensile strength. The tensile strength at fixed mold surface, center and moving mold surface

increased by 19%, 12% and 34%, respectively compared to their corresponding values in pure

steam case.

0.0

0.2

0.4

0.6

0.8

Fixed surface

Center

Moving surface

Pure

steam

0.41 0.69

Hot air pressure, MPa

Te

nsile

str

en

gth

(M

Pa

)

Air temperature = 160 C

Air flow rate = 80 l/min

Figure 5-19 Tensile strengths of molded EPP samples produced with pure steam and steam

mixed with hot air at different pressures

5.6 Conclusions

In this study, the existing steam-chest molding machine was modified to accommodate the

application of hot air in an attempt to reduce the sensitivity of the steam temperature decrease

with a pressure drop. The introduction of hot air was optimized to investigate the effect of

different parameters such as the hot air flow rate, the hot air temperature and the hot air pressure,

while the steam pressure was kept constant. The steaming time decreased by 32% and the local

temperature at entry port decreased by 8% at the highest available air flow rate of 120 l/min. The

overall heat transfer improved significantly with an increase in the hot air flow rate. The surface

roughness values (Ra and Rz) decreased by approximately 50% at the hot air flow rate of 120

l/min. An increase in the hot air pressure also showed a decrease in the surface waviness (Wa) by

approximately 50%. However, varying the hot air temperature did not cause any significant

change on the surface property.

The use of hot air with steam showed significant improvement in the overall consistency of the

tensile property across the molded EPP part as compared to the samples molded with pure steam.

142

The corresponding consistency was achieved due to an improved heat flow to the core of the

molded sample. This is possible due to the synergistic effect of the high thermal conductivity of

steam and the low Joule-Thompson coefficient of hot air. With either an increase in the hot air

flow rate or in the pressure, the heat flow is improved leading to an overall improvement in the

tensile property. Hence the results of this work reveal the potential application of hot air in the

steam-chest molding process to produce EPP bead products with improved surface quality,

enhanced mechanical properties and a shortened cycle time resulting in a reduced operating cost.

5.7 References

(1) Eaves, D. Handbook of Polymer Foams; Rapra Technology: Shawbury, Shrewbury, U.K,

2004.

(2) Schut, J.H. Expandable Bead Molding goes High-Tech. Plast. Technol. 2005, 51, 68.

(3) Sopher, S.R. Advanced Development of Molded Expanded Polypropylene and

Polyethylene Bead Foam Technology for Energy Absorption. SPE ANTEC Tech. Pap.

2005, 2577.

(4) Avalle, M.; Belingardi, G.; Montanini, R. Characterization of Polymeric Structural Foams

under Compressive Impact Loading by Means of Energy-Adsorption Diagram. Int. J.

Impact Eng. 2001, 25, 455.

(5) Beverte, I. Deformation of Polypropylene Foam Neopolen®P in Compression. J. Cell.

Plast. 2004, 40, 191.

(6) Bouix, R.; Viot, P.; Lataillade, J.L. Polypropylene Foam Behavior under Dynamic

Loading. Int. J. Impact Eng. 2009, 36, 329.

(7) Bureau, M.N.; Champagne, M.F.; Gendron, R. Impact-Compression-Morphology

Relationship in Polyolefin Foams. J. Cell. Plast. 2005, 41, 73.

(8) Viot, P. Hydrostatic Compression on Propylene Foam. Int. J. Impact Eng. 2009, 36, 975.

(9) Britton, R. Update on Mouldable Particle Foam Technology; Rapra Technology:

Shawbury, Shrewsbury, UK, 2009.

(10) Mills, N.J.; Kang, P. The Effect of Water Immersion on the Fracture Toughness of

Polystyrene Foam used in Soft Shell Cycle Helmets. J. Cell. Plast. 1994, 30, 196.

(11) Rossacci, J.; Shivkumar, S. Bead Fusion in Polystyrene Foams. J. Mater. Sci. 2003, 38,

201.

143

(12) Mills, N.J. Polymer Foams Handbook: Engineering and Biomechanics Application and

Design Guide; Butterworth Heinemann: Oxford, 2007.

(13) Nakai, S.; Taki, K.; Tsujimura, I.; Oshima, M. Numerical Simulation of a Polypropylene

Foam Bead Expansion Process. Polym. Eng. Sci. 2008, 48, 107.

(14) Zhai, W.; Kim, Y.W.; Jung, D.W.; Park, C.B. Steam-Chest Molding of Expanded

Polypropylene Foams. 1. DSC Simulation of Bead Foam Processing. Ind. Eng. Chem. Res.

2010, 49, 9822.

(15) Zhai, W.; Kim. Y.W.; Jung, D.W.; Park, C.B. Steam-Chest Molding of Expanded

Polypropylene Foams. 2. Mechanism of Interbead Bonding. Ind. Eng. Chem. Res. 2011, 50,

5523.

(16) Sands, M. An Analysis of Mold Filling and Defect Formation in Lost Foam Castings; M.S.

Thesis: Worcester Polytechnic Institute, Worcester, MA, 1998.

(17) Stupak, P.R.; Donovan, J.A. The Effect of Bead Fusion on the Energy Absorption of

Polystyrene Foams, Part II: Energy Absorption. J. Cell Plast. 1991, 27, 506.

(18) Stupak, P.R.; Frye, W.O.; Donovan, J.A. The Effect of Bead Fusion on the Energy

Absorption of Polystyrene Foam. Part I: Fracture Toughness. J. Cell. Plast. 1991, 27, 484.

(19) Nofar, M.; Guo, Y.; Park, C.B. Double Crystal Melting Peak Generation for Expanded

Polypropylene Bead Foam Manufacturing. Ind. Eng. Chem. Res. 2013, 52, 2297.

(20) Guo, Y.; Hossieny, N.; Chu, R.K.M.; Park, C.B.; Zhou, N. Critical Processing Parameters

for Foamed Bead Manufacturing in a Lab-Scale Autoclave System, Chemical Engineering

Journal. 2013, 214, 180.

(21) Choi, J.B.; Chung, M.J.; Yoon, J.S. Formation of Double Melting Peak of Poly(propylene-

co-ethylene-co-1-butane) during the Pre-expansion Process for Production of Expanded

Polypropylene, Ind. Eng. Chem. Res. 2005, 44, 2776.

(22) Sharudin, R.W.B.; Ohshima, M. CO2-induced Mechanical Reinforcement of Polyolefin-

based Nanocellular Foams, Macromol. Mater. Eng. 2011, 296, 1054.

(23) Van Wylen, G.J.; Sonntag, R.E., Fundamentals of Classical Thermodynamics, third ed.;

John Wiley and Sons: New Jersey, 1986.

(24) Goodenough, G.A. Thermal Properties of Steam: University of Illinois Bulletin. 75, 1914.

(25) Roebuck, J.R. The Joule-Thomson Effect in Air. Am. Acad. of Arts and Sciences, 1925,60,

537.

144

(26) Kadoya, K.; Matsunaga, N.; Nagashima, A. Viscosity and Thermal Conductivity of Dry

Air in the Gaseous Phase, J. Phys. Chem. Ref. Data. 1985, 14, 947.

(27) Keyes, F.G.; Vines, R.G. The Thermal Conductivity of Steam, Int. J. Heat Mass Transfer.

1964, 7, 33.

(28) Degarmo, E.P.; Black, J.T.; Kohser, R.A. Materials and Processing in Manufacturing,

ninth ed.; John Wiley and Sons: New Jersey, 2002.

(29) Whitehouse, D. Handbook of Surface Nanometrology, second ed.; CRC Press: Florida,

2011.

(30) Wunderlich, B. Macromolecular Physics, Vol. 1, Crystal Structure, Morphology, Defects,

first ed.; Academic Press: New York, 1973.

145

Chapter 6 Processing of TPU Bead Foams In Lab-Scale Bead Foaming System and Sintering Mechanism With Steam-Chest Molding

Technology

6 Production and Sintering of E-TPU Beads

6.1 Introduction

Thermoplastic polyurethanes (TPUs) are multi-block copolymers that exhibit a unique

combination of strength, flexibility and processability due to their phase-separated

microstructure [1,2]. These properties result from a molecular structure with rigid HS domains

dispersed in the soft segment (SS) matrix. The SS is a polyol with an ester or ether group in the

main chain having a low glass transition temperature and is viscous at service temperature,

imparting flexibility to TPU. The HS is formed by the reaction of diisocyanate and short-chain

diols, which crystallizes and influences the mechanical properties in TPU such as hardness and

tear strength. As a result of this unique microstructure, TPUs exhibit very good impact properties

at low temperature, excellent chemical resistance and great flexibility over a broad service

temperature, which make them suitable for a wide range of demanding applications such as

automobile parts, construction materials, sports equipment, and medical instruments. A major

limitation for the use of TPU is its middle up to high hardness. Addition of plasticizers can

achieve soft grade TPUs. However processing is much more challenging and the plasticizers tend

to migrate out of the material in long-term applications. The production of foamed TPU can

reduce the material hardness without additional plasticizers. The reduced density due to foaming

can open new fields of applications for TPU materials. TPUs can be foamed using different

techniques such as extrusion process, batch or continuous process in producing expanded bead

foams.

Recently, expanded TPU bead foams (E-TPU) that can be molded into complex three-

dimensional products have been developed [3]. At the present, industry utilizes soft grade TPUs

to process E-TPU beads in order to make sintering of the beads more effective and easier during

the steam-chest molding process. However, softer grade TPUs has less concentration of HS (i.e.,

crystallinity) and hence suffers from lower mechanical properties and lower service

146

temperatures. Further, the soft TPUs may suffer from severe dimensional instability from

exposure to high temperature steam during the steam-chest molding process [4]. The expanded

TPU beads can also suffer from a high degree of shrinkage after foaming due to the loss of gas.

Glycerol esters are predominant additives used commercially which provides anti-collapse

protection by forming a barrier on the surface of the foam, slowing the egress of the blowing

agent. This allows time for air to enter the cells, replace the blowing agent, and prevent collapse

of the foam. After acting as a plasticizer, GMS eventually migrates to the surface of the bubbles

within the polymer matrix. Hence the amount of GMS collected on the skin of the foams is

minimal. The use of glycerol esters may also affect the crystallization kinetics of a polymer.

Naguib et al. reported increase in the crystallization temperature and the degree of crystallinity of

linear and branched polypropylene in presence of GMS [5]. The crystals generated during the

foaming process can lead to the production of high quality foams with fine cell size and high cell

density. The crystals can also be effectively used in the sintering of beads during the steam-chest

molding process. In this context, investigating the crystallization behavior of TPU with the

presence of GMS and butane can provide new strategies on the processing and molding of E-

TPU bead foams and their products.

In this chapter, a lab-scale autoclave bead foaming setup was used to manufacture E-TPU beads

with desirable crystalline structure. Furthermore, the processed E-TPU beads were sintered using

a steam-chest molding machine into rectangular three dimensional samples. The effect of

varying HS concentration on the bubble nucleation and the cell density of the processed E-TPU

beads was investigated. The processing techniques to produce E-TPU beads were also

investigated. The E-TPU beads were characterized to investigate the morphology and the

expansion ratio. The thermal behavior was also investigated to characterize the effect of foaming

on the development of HS crystalline domains in the E-TPU foams. Finally, the tensile property

of the moulded sample was studied to investigate the sintering behavior of the E-TPU beads.

6.2 Materials and Experimental Procedure

6.2.1 Materials

Three types of commercially available TPUs (Elastollan) from BASF were selected to

manufacture E-TPU bead foams. The densities of the selected TPU resins were 1.08 g/cm3, 1.11

g/cm3 and 1.13 g/cm

3, with a hardness of Shore 70A, Shore 80A and Shore 90A, respectively.

147

The higher hardness is caused due to the higher concentration of HS. The E-TPU beads based on

the base TPU materials are designated as E-TPU-70A, E-TPU-80A and E-TPU-90A

respectively. CO2 with a 99 % purity produced by Linde Gas was used as the impregnation gas.

6.2.2 Lab-scale bead foaming setup

Figure 6.1 shows a schematic of the overall lab-scale autoclave bead foaming system that has

been designed and constructed at our laboratory. The autoclave consists of a cylindrical chamber,

a guided cylinder which is welded on the lid of the chamber, and a rotary shaft driven by a DC

motor with three propellers mounted on it. The chamber has a exit valve situated at the bottom

through which the samples are discharged and the foaming of the material occurs.

Figure 6-1 A schematic of autoclave bead foaming set-up

6.2.3 Expanded TPU (E-TPU) bead foaming procedure

Three processing techniques were used to manufacture E-TPU bead foams.

6.2.3.1 Pressure-drop method without water

To conduct foaming experiments, 10 grams of TPU pellets was put into the autoclave chamber.

The chamber was then maintained at designated CO2 pressure and temperature for a period of

time to impregnate the TPU pellets with CO2. A wide range of saturation temperature (Tsat)

ranging between 140°C to 170°C was utilized during the impregnation stage. The saturation time

148

(tsat) was fixed as 60 min based on the simulation experiments described in Chapter 4. After the

saturation process, the shut-off valve was opened and the TPU pellets were discharged from the

chamber. Once the saturated TPU pellets exited the chamber, foaming occurred due to the

thermodynamic instability to form expanded TPU bead foams.

6.2.3.2 Pressure-drop method with water

In this method, the chamber was filled with 750 ml of water and 30 grams of TPU pellets. The

water was used as a mixing media and to uniformly distribute the heat to the TPU pellets. Next,

CO2 was supplied at the desired pressure and the chamber was heated to the desired saturation

temperature. The Tsat was selected based on the results from previous method and the tsat was 60

min. The samples were saturated for a certain saturation time for the impregnation of CO2.

Subsequently, the depressurization was accomplished by opening the shut-off valve and the

saturated TPU pellets, water and CO2 were discharged from the chamber. Once the saturated

pellets exited the chamber, foaming occurred due to thermodynamic instability.

6.2.3.3 Temperature-jump method

In this method, the chamber was filled with 30 grams of TPU pellets and saturated with CO2 at

room temperature for a desired time (tsat). After the saturation process, the CO2 was released

from the autoclave and the impregnated TPU pellets was transferred to a hot oil bath set at the

desired foaming temperature (Tfoam) to induce the thermodynamic instability. The foaming time

(tfoam) was fixed at 60 sec and then the expanded TPU beads were removed. The E-TPU beads

were washed prior to the characterization process.

6.2.4 Thermal behavior of E-TPU beads

The thermal behavior of processed E-TPU beads was measured in a differential scanning

calorimetry (DSC 2000, TA Instruments) by heating the foamed samples to 230oC at a rate of

10oC/min.

6.2.5 Gel Permeation Chromatography (GPC)

Although the TPU used in the experiments are polyether based with high hydrolysis resistance,

the chance of hydrolysis increases during the annealing experiments at high temperature in

presence of water. Hence the Mw of the foamed beads was measured using a gel permeation

149

chromatography (GPC) (experiments were conducted at Nike, Beaverton, USA). The Mw was

analyzed relative to linear polystyrene standards with RI detection in THF mobile phase.

6.2.6 Water up-take analysis

The water-uptake percentage was measured by saturating the E-TPU-90A beads over a range of

temperatures. Wetted E-TPU-90A beads were obtained by entirely immersing in water for

different saturation times. Then the E-TPU beads were wiped with paper towel, and immediately

weighed using a digital scale to measure the water up-take to 0.001 g accuracy (ASTM D570).

The water up-take percentage was calculate based on Equation 6.1.

Water-uptake rate = [(Wt – W0)/W0] x 100% Eq. 6.1

where Wt is the weight of the water saturated E-TPU bead and W0 is the initial mass of the

sample. The E-TPU beads with water were also used in the sintering process with the steam-

chest molding machine to investigate the effect of water on the sintering process.

6.2.7 Foam characterization

The morphology of the E-TPU bead foams was observed with a JOEL JSM-6060 scanning

electron microscope (SEM). The samples were fractured in liquid nitrogen, mounted on stubs,

and sputter coated with Au/Pd.

An image analysis on the SEM micrograph was conducted to obtain the average cell size and the

cell density using Image J (from the National Institute of Health). A micrograph showing more

than 100 bubbles was chosen, and the software determined the number of cells in the

micrographs. By analyzing the area of the micrographs, the cell density of each sample was

estimated using Equation 6.2. The density of the E-TPU foam was evaluated using a water-

displacement technique (ASTM D792-00). Using this information, the volume expansion ratio

(VER) of the samples was then evaluated as shown in Equation 6.3.

(Eq. 6.2)

(Eq. 6.3)

150

6.2.8 Steam-chest molding of E-TPU beads

A lab-scale steam-chest molding machine commercially manufactured by DABO Precision

(DPM-0404VS) from Korea was used for the sintering of the processed E-TPU bead foams. The

mold cavity was 15 cm x 6 cm x 2.5 cm. The mold consists of a fixed side and a moving side.

Both the mold surfaces have ports for injection of the steam into the mold cavity. The basic

process of steam-chest molding process consists of three main steps. Figure 6.2 summarizes

these steps. In the first step, the E-TPU beads are filled into the mold cavity. In the second step,

the steam is injected from the fixed mold at the desired processing steam pressure (P1). Then

steam is injected from the moving mold (P2). Subsequently, the steam is injected from both

molds (P3) followed by depressurization and holding to stabilize the sample. In the third step, the

mold and the sample is cooled with water and followed by vacuuming to remove the water.

Finally the sample is ejected from the mold. The unit of steam pressure used in this study is the

gauge pressure in bar, which is 1 bar lower than the absolute pressure.

6.2.9 Mechanical property measurement

Dog-bone shaped specimens were prepared from the molded part for tensile test experiments.

The dimensions of the specimen were based on the ASTM D3574-11 standard test for flexible

cellular materials such as slab, bonded and molded urethane foams. Tensile strengths of the

specimens were measured using a Zwick Roell tensile tester at a crosshead speed of 100

mm/min.

151

Figure 6-2 Steam-chest molding procedure

6.3 Results and Discussions

6.3.1 Foaming behavior of E-TPU beads

As per the earlier discussion in chapter 4, the HS crystals play a very important role as

heterogeneous nucleating agents to increase the cell density of TPU beads foams. By controlling

the HS crystalline domains the overall morphology and expansion ratio of the TPU beads can be

effectively controlled. In the lab-scale autoclave foaming of TPU beads, the effect of different

processing techniques (pressure drop method and temperature jump method) and processing

parameters (effect of water, saturation temperature and saturation pressure), which ultimately

affects the HS crystalline domains and hence the overall morphology of TPU beads foams is

systematically investigated.

152

6.3.1.1 Effect of water on foaming behavior of E-TPU beads

Figure 6.3 depicts the SEM morphologies of the AR-TPU-90A beads processed without water

and with water. The tsat and CO2 pressure were 30 min and 55 bar, respectively. The AR-TPU-

70A and AR-TPU-80A beads showed similar results and hence have not been included in the

results. Overall foaming with water significantly improved the foaming behavior of TPU beads.

It can also be observed that foaming in presence of water decreased the saturation temperature

(Tsat) to manufacture TPU bead foams. The Tsat decreased by approximately 20°C after

processing TPU beads in the presence of water.

(a) (b)

(c) (d)

Figure 6-3 Morphology of AR-TPU-90A beads at 55 bar CO2 pressure: (a), (b) without

water; (c), (d) with water

6.3.1.2 Effect of water on foaming behavior of E-TPU beads

Figures 6.4 and 6.5 shows the SEM morphologies of TPU beads processed without water and

with water at different Tsat and CO2 pressures (Psat). Two interesting observations can be made

from the morphologies of TPU beads shown in Figs 6.4 and 6.5. First, the effect of saturation

pressure is quite visible in the TPU foam morphologies processed without water and with water.

By increasing the Psat from 55 bar to 82 bar, the cell density significantly increased. At higher

153

saturation pressure, the concentration of CO2 in the TPU matrix increases, which decreases the

Rcr (Equation 6.4) and results in higher nucleation rate.

Eq. 6.4

The second observation is the effect of a critical saturation temperature (Tcritical) at which the

foaming behavior of TPU beads improved significantly. In the case of TPU beads processed

without water, the Tcritical was at 160°C at both the investigated saturation pressures as shown in

Figs. 6.4c and 6.4f. The foaming of TPU beads at Tcritical is observed to have improved

dramatically. However, the Tcritical for the TPU beads processed with water decreased to 145°C as

shown in Figs. 6.5b and 6.5d.

(a) (b) (c)

(d) (e) (f)

Figure 6-4 Morphology of AR-TPU-90A beads processed without water: (a), (b), (c) 55 bar

CO2; (d), (e), (f) 83 bar CO2

154

(a) (b)

(c) (d)

Figure 6-5 Morphology of AR-TPU-90A beads processed with water: (a), (b) 55 bar CO2;

(c), (d) 83 bar CO2

6.3.1.3 Effect of processing methods on foaming behavior of E-TPU beads

Figures 6.6a and 6.6b compares the AR-TPU-70A beads processed by pressure-drop and

temperature-jump method described in section 6.2.3.2 and section 6.2.3.3, respectively.

Similarly, Figs. 6.7a and 6.7b compares the AR-TPU-90A beads processed by pressure-drop and

temperature-jump method. Overall, in both AR-TPU-70A and AR-TPU-90A beads the

temperature-jump method significantly increased the cell density and reduced the cell size. In the

temperature-jump method the TPU pellets are saturated with CO2 at the room temperature and

the thermodynamic instability is achieved by suddenly increasing the temperature to the foaming

temperature. On the other hand, in the pressure-drop method the TPU pellets are saturated at the

foaming temperature and the thermodynamic instability is achieved by sudden drop in the

pressure of the system. Two possible reasons may have resulted in the observed difference in the

foaming mechanism with the two different methods. First, at the same saturation pressure (55 bar

CO2) the solubility of CO2 is much higher at lower temperature compared to saturation done at

higher temperature. Hence higher concentration of CO2 will reduce the critical radius and result

in higher nucleation. Second at higher saturation temperature more of the existing HS crystals

155

are molten and only a few remaining HS crystals may have contributed as heterogeneous

nucleating sites compared to a large number of crystals present in the temperature-jump method.

(a) (b)

Figure 6-6 Morphology of AR-TPU-70A beads processed with CO2 pressure of 55 bar at

110°C: (a) pressure-drop method (b) temperature-jump method

(a) (b)

Figure 6-7 Morphology of AR-TPU-90A beads processed with CO2 pressure of 55 bar at

140°C: (a) pressure-drop method (b) temperature-jump method

6.3.2 Characterization of TPU

Figure 6.8 compares the expansion ratio of AR-TPU-70A and AR-TPU-90A beads processed

over a range of saturation temperatures with different methods using 55 bar CO2 pressure.

Overall the expansion ratio of both AR-TPU-70A and AR-TPU-90A increased with an increase

in the saturation temperature. However, the expansion ratio of samples processed with pressure-

drop method and in the presence of water was higher compared to pressure-drop without water

and the temperature-jump method.

156

120 140 1600

2

4

6

8

10

12

14

16 Pressure drop- without water

Temp Jump

Exp

an

sio

n r

atio

Saturation temperature (oC)

Pressure drop- with water(a) AR-TPU-70A

100 120 140 160 1800

2

4

6

8

10

12

14

16

18

Expansio

n R

atio

Saturation temperature (oC)

AR-TPU-90A(b)

Pressure drop- without water

Temp Jump

Pressure drop- with water

Figure 6-8 Expansion ratio of E-TPU beads produced with different methods: (a) AR-TPU-

70A, (b) AR-TPU-90A

Figure 6.9 compares the expansion ratios of AR-TPU-70A, AR-TPU-80A and AR-TPU-90A

beads processed over a range of foaming temperature with the temperature-jump method.

Overall the expansion ratio of all the samples increased with an increase in the foaming

temperature. However, the expansion ratio of samples processed with pressure drop method and

in the presence of water was higher compared to pressure drop without water and the

157

temperature jump method. More HS crystals melted at a higher foaming temperature, causing the

SS to be more flexible, and hence, the TPU bead foams expanded easily. It is also interesting to

compare the expansion ratios of the beads based on the type of TPU. The HS concentration (i.e

crystallinity) increases with higher hardness (AR-TPU-70A < AR-TPU-80A<AR-TPU-90A).

Hence AR-TPU-70A with lower HS concentration is more elastic and had a higher expansion

compared to AR-TPU-80A and AR-TPU-90A at lower foaming temperatures. However at higher

foaming temperature the expansion ratio of AR-TPU-90A was higher compared to AR-TPU-70A

and AR-TPU-80A bead foams as shown in Fig. 6.9. This is similar to the case of high stiffness

governing on the expansion ratio as the temperature increases in the typical mountain shape

observed in the extrusion foaming [6]. The presence of higher amount of HS crystallites in AR-

TPU-90A assisted with higher expansion at higher foaming temperatures.

100 110 120 130 140 150 160 170

1

2

3

4

5

6

7

8

9

10

Exp

an

sio

n r

atio

Saturation temperature (0C)

AR-TPU-70A

AR-TPU-80A

AR-TPU-90A

Figure 6-9 Expansion ratio of different TPU foam beads processed with temperature-jump

method

6.3.3 Thermal behavior of E-TPU bead foams

6.3.3.1 Effect of water on thermal behavior of E-TPU beads

Figure 6.10 compares the DSC melting curves of AR-TPU-90A after annealing at 150°C for 30

min with different annealing conditions. The first curve at the bottom is the heating curve of the

sample, which was annealed at ambient pressure (1 bar). The second curve from the bottom is

158

the sample that was annealed in HP-DSC in presence of 55 bar CO2 pressure and without the

effect of foaming. The third curve depicts the melting behavior of the TPU bead foams processed

without water with 55 bar CO2. Finally, the fourth curve at the top shows the melting behavior of

the TPU bead foams processed in the presence of water with 55 bar CO2. Compared to annealing

at ambient pressure (1 bar), annealing with CO2 resulted in the decrease of the high temperature

(Tm-high ) melting peak from 165°C to 163.7°C due to the plasticizing effect of CO2. However the

formation of a new low temperature melting peak (Tm-low) at 62°C is observed after annealing

with CO2. Hence, CO2 assisted with nucleation of less perfected HS crystals. After foaming

without water, the Tm-high shifted to 167.2°C and the Tm-low shifted to 64°C. The foaming action

would have caused extensional stress and hence resulted in perfection of HS crystals and hence

the melting peaks shifted to higher temperatures. Further, foaming with water resulted in the

shifting of the Tm-high by approximately 20oC to 185

oC. The Tm-low also shifted to higher

temperature by 11oC. Also a new melting peak was formed at 159°C. The presence of water may

have acted as a plasticizer and assisted the mobility of HS crystals. On the other hand, the

extensional stress caused by the foaming action resulted in the perfection or growth of the HS

crystals forming the new very high melting temperature peak. Whereas, during the cooling action

after the foaming, new HS crystallites may have nucleated forming the low temperature melting

peaks. The total heat of fusion (ΔHT) (J/g) also increased after foaming with water compared to

foaming without water and annealing at different conditions. Thus the overall crystallinity of the

TPU foams increases after foaming.

The observed improvement in the foaming of TPU beads in the presence of water, which was

discussed in the section 6.3.1.1 can also be attributed to the formation and perfection of HS

crystallites that decreases the Rcr (Equation 6.3) by inducing local pressure variation (ΔPlocal) in

the amorphous SSs.

159

Figure 6-10 DSC melting curves of AR-TPU-90A after annealing at 150°C for 30 min with

different annealing conditions

6.3.3.2 Effect of annealing temperature on thermal behavior of E-TPU beads

Figure 6.11 depicts the DSC melting curves of AR-TPU-90A bead foams processed at different

saturation temperature’s and saturation time of 30 min in the presence of 55 bar CO2. These

beads are processed with pressure-drop method with water. Overall at all the annealing

temperatures, the bead foams showed three distinct melting peaks. A very low temperature

melting peak (Tm-low) was observed at 75°C. A new high melting peak (Tmc) was formed in the

range of 150°C to 160°C. This peak is related to the melting of the HS crystals formed during the

cooling phase after the saturation step and the foaming was completed. The third peak (Tma) is

observed at very high temperature in the range of 178 °C to 185°C depending on the saturation

temperature. This peak is related to the melting of the HS crystals, which are perfected during the

annealing process and also due to the extensional stress caused by the foaming action.

However it is interesting to observe the sudden increase in the Tmc and Tma melting peaks after

foaming at saturation temperature of 145°C. The Tmc and Tma temperature’s increased by

approximately 9°C and 6°C by changing the saturation temperature from 140°C to 145°C.

Further increase in the saturation temperature to 150oC did not affect the Tmc and Tma melting

30 60 90 120 150 180 210 240-0.5

0.0

0.5

165 (Tm-high

)

1.2 J/g

16 J/g

2.8 J/g

11.7 J/g

167.2

163.7

62 (Tm-low

)

64

75

159

foam-with water

foam-w/o water

Unfoamed

Heat flow

(J/g

)

Temperature (°C)

1 bar

1859.6 J/g

Endo

160

peaks. The overall crystallinity also increased significantly with a sudden increase in the total

heat of fusion (ΔHT) to 28.5 J/g as shown in Fig. 6.9. The sudden change in the overall HS

crystalline domains at this critical saturation temperature also affected the foam morphology,

which was shown in Figs. 6.4 and 6.5 also discussed in section 6.3.1.2.

30 60 90 120 150 180 210 240-0.4

-0.2

0.0

0.2

0.4

20.5 J/g

28.5 J/g

79.9

76.1

Tmc

=151

160

159

185

184

150 C

145 CHe

at flo

w (

J/g

)

Temperature (°C)

140 CT

ma=178

HT=23.1 J/g

Saturation

temp

72

Endo

Figure 6-11 DSC melting curves of AR-TPU-90A bead foams processed with pressure-drop

method with water over a range of saturation temperature with 55 bar CO2 pressure

6.3.3.3 Effect of processing method of thermal behavior of E-TPU beads

Figure 6.12 depicts the melting curves of the AR-TPU-70A bead foams manufactured with the

different processing techniques at temperature of 120°C and 55 bar CO2 pressure. The formation

of highly perfected HS crystallites as a result of the extensional stress caused by the expansion

can be observed in the bead foams processed with pressure-drop method with the presence of

water compared to the beads processed with temperature-jump method.

161

0 50 100 150 200 2500.0

0.2

0.4

16.5 J/g

16.3 J/g

151.1oC

68.2oC

141.7oC

65.7oC

166.8oC

136.1oC

He

at

flo

w (

J/g

)

Temperature (°C)

Press-drop-withwater

Press-drop-w/owater

Temp-Jump

66.9oC

13.6 J/g

AR-TPU-70A

Figure 6-12 DSC melting curves of AR-TPU-70A bead foams processed with different

methods

6.3.4 GPC analysis

Figure 6.13 compares the molecular weight of the AR-TPU-70A and AR-TPU-90A bead

foams processed in water with the unfoamed TPU materials. As seen in the Fig. 6.13, the

saturation of TPU at high temperature in the presence of water causes breakage of SS chains due

to hydrolysis and results in decrease of the overall molecular weight. However the AR-TPU-90A

foams had much lower decrease in the molecular weight. This might be due to the higher

concentration of HS in AR-TPU-90A beads, which may have acted as filler strengthening the

overall material.

162

Figure 6-13 Average molecular weight of the E-TPU beads processed with pressure-drop in

the presence of water: (a) AR-TPU-70A, (b) AR-TPU-90A

6.3.5 Sintering of E-TPU beads with steam-chest molding machine

To investigate the sintering behavior of TPU beads processed with the temperature-jump method

with the steam-chest molding machine, three beads were selected based on the cell morphology,

expansion ratio and processing method. Table 6.1 shows the different TPU bead foam materials

processed with the temperature-jump method used for the molding experiments, the expansion

Unfoamed-70A Foamed-70A

0.0

5.0x104

1.0x105

1.5x105

2.0x105

2.5x105

3.0x105

Mw

(g

/mo

l)

(a) Tsat

= 115oC

tsat

= 60 min

Psat

= 800 psi

Unfoamed- 90A Foamed-90A0.0

5.0x104

1.0x105

1.5x105

2.0x105

2.5x105

Tsat

= 145oC

tsat

= 60 min

Psat

= 800 psi

Mw

(g

/mo

l)

(b)

163

ratio and the processing conditions (steam pressure/time) used in the steam-chest molding

machine to produce the E-TPU samples from the respective beads. To investigate the sintering

behavior of TPU beads processed with the pressure-drop method with water, E-TPU-90A beads

were selected. Table 6.2 shows the different processing conditions (steam pressure/time) used in

the steam-chest molding machine to produce the molded samples. Figure 6.14 depicts the actual

beads and their respective SEM images showing their cellular morphologies. It can be observed

that the beads processed with the temperature-jump method showed very glossy surface finish as

a result of the microcellular morphology achieved in their microstructure compared to the beads

processed with pressure-drop method.

Table 6-1 Different E-TPU beads and conditions (steam pressure/time) used to produce

molded E-TPU samples Sample Processing

method

Expansion

ratio

Fixed mold Moving mold Both Molds

Pressure

(bar)

Time

(sec)

Pressure

(bar)

Time

(sec)

Pressure

(bar)

Time

(sec)

E-TPU-

70A

Temp-

jump

14 1.6 25 1.6 5 1.6 15

E-TPU-

80A 8 1.6 25 1.6 5 1.6 15

E-TPU-

90A 8 3.8 75 3.8 75 3.8 20

Table 6-2 Different conditions (steam pressure/time) used to produce molded E-TPU-90A

samples

Sample Processing

method

Expansion

ratio

Fixed mold Moving mold Both Molds

Pressure

(bar)

Time

(sec)

Pressure

(bar)

Time

(sec)

Pressure

(bar)

Time

(sec)

E-TPU-

90A

Pressure-

drop with

water

13

1.5 25 1.5 5 1.5 15

2 25 2 5 2 15

2.2 25 2.2 5 2.2 15

2.4 25 2.4 5 2.4 15

164

(a) (b)

(c) (d)

(e) (f)

Figure 6-14 Actual E-TPU beads and their cellular morphologies: (a), (b) E-TPU-70A; (c),

(d) E-TPU-80A; (e), (f) E-TPU-90A

Figure 6.15 shows the molded E-TPU-90Abeads processed with the pressure-drop method with

water over range of steam pressure in the steam-chest molding machine. By increase in the steam

pressure, the overall sintering of the E-TPU-90A beads improved. By increasing the steam-

pressure to 2.2 bar (Fig. 6.15c), the overall dimensional stability of the molded sample started to

decrease and bead shrinkage was observed. At a steam pressure of 2.4 bar (Fig. 6.15d), the

molded part completely collapsed due to high degree of shrinkage in the beads. There is an

optimal pressure, which resulted in the best bonding of the E-TPU-90A beads while maintaining

the overall dimensional stability of the product. However, it is interesting to observe the

fractured E-TPU-90A molded part in Fig. 6.16, which was processed at 2 bar steam pressure and

165

showed the best overall sintering and dimensional stability. Although, the surface of the beads

was deformed due to the high temperature steam, the sintering of the beads was not very

effective. The fractured sample cracked very easily as a result of inter-bead failure, which is a

cause of very poor bead-to-bead bonding.

(a) (b)

(c) (d)

Figure 6-15 E-TPU-90A beads molded over range of steam pressure; (a) 1.5 bar, (b) 2 bar,

(c) 2.2 bar, (d) 2.4 bar

Figure 6-16 Fractured E-TPU-90A bead foam molded part manufactured with 2.2 bar

steam pressure

166

There might be two possible reasons for the observed bead shrinkage after the steam-chest

molding procedure at steam pressure 2.2 bar and above. First, the high steam temperature may

cause excessive melting which results in the percolation of the HS crystals. However, at the

steam pressure of 2.2 bar, approximately 50 % of the HS crystals still exists in the E-TPU bead

foams and hence it suggests that the shrinkage might not have been caused due to the percolation

of the HS crystals. The second possible reason for the shrinkage might be due to the thermal

stress induced in the beads during the bead foam processing step at high annealing temperature

of 145oC. The thermal stress is released during the molding step and due to the elastomeric

nature of TPU there is excessive shrinkage of the E-TPU beads. One possible technique would

be to use water laden beads during the steam-chest molding procedure. The water trapped in the

beads would vaporize due to the high temperature steam and cause the beads to expand and thus

reduce the effect of shrinkage and improve sintering.

To investigate the effect of water during the steam-chest molding process, the processed E-TPU

beads were immersed completely in water at different temperatures and soaking times and the

up-take amount of water was measured. Figure 6.17 shows the uptake percentage of water

absorbed in the E-TPU beads. At soaking temperatures of 25°C and 50°C, there was not

significant intake of water in the E-TPU beads. By increasing the soaking time to 70°C, the

percentage of water uptake increased significantly. With increase in the soaking time to 90°C,

the percentage of water uptake increased even further. However it should be noted that there was

also significant shrinkage after the E-TPU beads soaked with water at 90°C were exposed to

atmosphere. This again may have been as a result of the high thermal stress induced in the E-

TPU beads at higher soaking temperature. The “Req” value signifies the time of soaking and it

can be observed in Figure 6.17 that the water uptake percentage reaches a plateau after certain

time. The time that the percentage of water uptake reaches saturation is 3 hours.

To investigate the effect of water on the sintering behavior of E-TPU-90A beads, the beads

soaked at 50°C and 70°C for 3 hours were selected as marked by the box in Figure 6.17. Figure

6.18 depicts the E-TPU-90A beads soaked with water at 50°C and 70°C, and subsequently

molded at steam pressure of 2 bars. Both the samples still showed high degree of shrinkage. The

E-TPU-90A beads molded without any water intake showed much better overall dimensional

stability at the same molding condition as shown in Figure 6.15b. Thus processing of E-TPU

beads at higher processing temperatures might not be the best method to manufacture beads as

167

they tend to induce high degree of thermal stress and result in shrinkage during the molding

process.

0 2 4 6 8 10 12 14 16 18 200

10

20

30

40

50

60

70

80

90

Wa

ter

Up

take

%

Req

250C

500C

700C

900C

Figure 6-17 Water uptake percentage in E-TPU-90A beads over a range of temperature’s

and times

(a) (b)

Figure 6-18 E-TPU-90A beads soaked with water molded at 2 bar steam pressure ; (a) 50°C

water temperature, (b) 70°C water temperature

168

Figure 6.19 shows the E-TPU-70A, E-TPU-80A and E-TPU-90A beads molded into rectangular

parts. The steam pressure used to mold the beads was 1.6 bar. Qualitatively, it can be observed

that there is a very effective sintering of the three beads. Furthermore, the dimensional stability

of the molded part also is very good. To get a more quantitative data on the sintering behavior of

the molded samples, the tensile properties of the molded products was measured. Figure 6.20

shows the overall process from the loading to the final fracture of the sample being tested for the

tensile property. The samples extended by approximately 350% until the fracture.

(a) (b) (c)

Figure 6-19 Steam-chest molded E-TPU bead foams: (a) E-TPU-70A, (b) E-TPU-80A, (c)

E-TPU-90A

(a) (b)

Figure 6-20 Tensile property testing of E-TPU-70A molded sample: (a) loaded sample, (b)

fractured sample

Figure 6.21 shows the stress v/s strain plot for a series of E-TPU-70A and E-TPU-80A samples.

Further, Fig. 6.22 shows the Young’s modulus and tensile strength’s of the E-TPU-70A and E-

169

TPU-80A samples and the values are compared with EPP and EPLA molded bead foam parts

with the same density. The Young’s modulus of E-TPU-70A and E-TPU-80A is much lower

compared to the EPP and EPLA samples as shown in Fig. 6.22a. The lower Young’s modulus is

due to the elastomeric property of the TPU. However, the Young’s modulus of E-TPU-80A is

slightly higher than the E-TPU-70A sample. The tensile strength of the E-TPU beads was also

observed to be higher than EPP and EPLA beads as shown in Fig. 6.22b. Furthermore, the tensile

strength of E-TPU-80A was significantly higher than E-TPU-70A as shown in Fig. 6.22b. The E-

TPU-80A has higher concentration of HSs and hence the overall crystallinity of the material is

higher compared to E-TPU-70A sample. The higher HS crystalline domains would have acted as

filler and hence increased the overall tensile strength of the material.

Figure 6-21 Stress v/s strain curves of the samples: (a) E-TPU-70A, (b) E-TPU-80A

170

0

5

10

15

20

25

30

EPLA

EPP

E-TPU-80A

Yo

un

g's

Mod

ulu

s (

MP

a)

E-TPU-70A

(a)

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

EPLAEPP

E-TPU-80A

Te

nsile

Str

en

gth

(M

Pa

)

(b)

E-TPU-70A

Figure 6-22 Comparison of Young’s modulus and tensile strength of E-TPU, EPP and

EPLA molded samples: (a) Young’s modulus, (b) Tensile strength

The bead-to-bead sintering was further investigated by observing the SEM images of the molded

part. The surface, cut surface and fractured surface of the molded E-TPU-70A and E-TPU-80A

were observed and the results are shown in Fig. 6.23. As seen, the surface of the molded E-TPU-

70A and E-TPU-80A samples showed a good surface quality with inter-bead bonding between

the TPU beads. Qualitatively, the inter-bead bonding between the TPU beads can be visualized

in the cut surface of the molded sample. The beads are effectively sintered with well-defined

171

bead-to-bead surface. However, the fractured surface images give the best indication on a very

good bonding between the beads. Both, the E-TPU-70A and E-TPU-80A samples showed a

complete intra-bead failure which indicates a strong bead-to-bead sintering quality. The intra-

bead failure observed in E-TPU-70A and E-TPU-80A is much higher than other fractured

surface of other polymer bead foam parts such as EPP and EPLA.

172

Figure 6-23 SEM micrographs of the surfaces, the cut surfaces, and the fracture surfaces of

molded E-TPU-70A and E-TPU-80A samples

173

In order to investigate the mechanism of sintering between the E-TPU beads, it would be

interesting to compare the thermal behaviors of the molded E-TPU beads after the steam-chest

molding process. Figures 6.24 compares the first heat melting curve, the foamed bead melting

curve and the molded sample melting curve for the E-TPU-70A, E-TPU-80A and E-TPU-90A

bead foams. As shown in Fig. 6.24a, after processing E-TPU-70A beads at a foaming

temperature of 130°C a very broad melting behavior is formed with two distinct melting peaks at

68.2°C and 135.6°C, respectively compared to the first heat melting behavior of TPU-70A. The

melting temperature decreases after foaming due to the plasticization effect of dissolved CO2,

which decreases the melting temperatures. However, the total heat of fusion (ΔHT) increased

from 16.8 J/g to 18.5 J/g after foaming compared to the neat-TPU-70A sample. After molding

the beads in the steam-chest molding machine at a steam temperature of 133°C, a new high

melting temperature peak is formed at 160.9°C and a very broad low melting peak is formed at

68.1°C. The high temperature melting peak is the melting of the perfected HS crystallites formed

during the steaming cycle, which results in annealing of the beads. Whereas, the low melting

temperature peak are the smaller of less-perfected HS crystallites formed during the cooling

cycle. The ΔHT also increased significantly after molding in the steam-chest molding machine as

seen in Figure 6.24b. The heat of fusion of the low melting peak was also observed to increase

significantly from 10.5 J/g to 20.2 J/g after molding the E-TPU-70A compared to the foaming

stage. Similar behavior was observed for E-TPU-80A and E-TPU-90A beads as shown in

Figures 6.24 b and 6.24 c.

The sintering mechanism of the E-TPU beads can be explained after the results discussed above.

The sintering is achieved by the melting and diffusion of the less perfected HS crystallites in the

E-TPU bead foams during the molding process. The HS chains along the adjacent bead surface

diffuse across each other and form less perfected crystallites during the cooling stage to form the

strong sintering behavior. On the other hand, the perfection of the existing HS crystallites

forming the new high melting temperature peak, which takes place as a result of the annealing

step during the steam-chest molding helps in maintaining the overall geometry of the molded E-

TPU bead foam part.

174

40 60 80 100 120 140 160 180 2000.0

0.1

0.2

0.3

0.4

0.5

Molded- 133 C

Foam- 130 C

68.10C

160.90C

7.9J/g20.2 J/g

156.9oC

he

at

Flo

w (

J/g

)

Temperature (°C)

121.7oC

16.8 J/g

10.52 J/g

68.20C

135.60C

8.0 J/g

Tsteam

= 1330C

E-TPU-70A Sintering Mechansims

First Heat

(a)

-60 -30 0 30 60 90 120 150 180 210 2400.0

0.5

1.0

E-TPU-80A Sintering Mechansims

Mold- 1330C

Foam- 1000C

Tsteam

= 133.30C

Tg=-44.0

Tg=-46.3

7.2 J/g0.9 J/g

3.5 J/g5.2 J/g

86.6

0.6 J/g10.6J/g

153.166.1

134.268.7

156.9

Heat F

low

(J/g

)

Temperature (°C)

86.6

Tg=-49.2

First Heat

(b)

175

-30 0 30 60 90 120 150 180 210 240-0.5

0.0

0.5

1.0

E-TPU-90A Sintering Mechansims

32.0 J/g

101.90C

179.30C

160.50C90.9

0C

19.3 J/g29.9 J/g

35.6 J/g

1720C

Mold- 1650C

Foam- 1650C

He

at

Flo

w (

J/g

)

Temperature (°C)

First Heat

1260C

6.8 J/g

(c)

Tsteam

= 1650C

Figure 6-24 DSC melting peak comparisons of neat-TPU, foamed E-TPU beads and molded

E-TPU beads: (a) E-TPU-70A, (b) E-TPU-80A, (c) E-TPU-90A

176

6.4 Conclusions

In a lab-scale autoclave bead foaming setup, E-TPU beads with desirable crystalline structure

was manufactured and the beads were sintered using a steam-chest molding machine. The effect

of varying HS concentration on the bubble nucleation and the cell density of the processed E-

TPU beads was investigated. It was observed that the perfection in the existing HS crystalline

domains and the new HS crystallites developed during the saturation process induced a higher

degree of bubble nucleation which resulted in high cell density with smaller cell size in the E-

TPU bead foams. The processing techniques to produce E-TPU beads were also investigated. It

was observed that processing with pressure-drop method in the presence of water significantly

improved the overall foaming behavior (cell density and expansion ratio) of E-TPU beads

compared to the pressure-drop method without water. The presence of water resulted in higher

plasticization of the TPU matrix, which caused formation of some highly perfected HS crystals.

On the other hand, a large number of less-perfected HS crystallites (i.e. nucleation) also were

formed after processing E-TPU beads with water. These HS crystallites improved the cell

nucleation of E-TPU beads. Another interesting observation was the significant increase in the

cell nucleation at the critical saturation temperature, which was also reduced by 15-20ºC due to

the plasticization effect of water.

Compared to the pressure-drop methods, the temperature-jump method was found to be more

effective to achieve microcellular E-TPU beads due to the existence of large number of less-

perfected HS crystallites, which can be utilized as heterogeneous bubble nucleating agents. The

E-TPU beads produced via temperature-drop method also showed very good bead-to-bead

sintering using the steam-chest molding machine compared to the samples produced with the

pressure-drop method with water. It was also observed that the TPUs with lower HS

concentrations were more effective for sintering of the E-TPU beads using steam-chest molding

machine.

177

6.5 References

[1] Gibson PE, Wallace MA, Cooper AL. Development in Block Copolymer. London;

Elsevier; 1982.

[2] Cooper SL, Tobolsky AV. J Appl Polym Sci 1966; 10:1837-1844.

[3] Bonart R, Morbitzer L, Hentze GL. Macromol Sci, Phys 1969; B3: 337-356.

[4] Blackwell J, Lee CD. Adv. Urethane Sci Technol 1984; 9: 25-46.

[5] Park H, Park CB, Tzoganakis C, Chen P. Ind Eng Chem Res 2007; 46: 3849-3851.

[6] Naguib HE, Park CB, Reichelt N. J Appl Polym Sci 2004; 91(4): 2661-2668.

178

Chapter 7 Conclusion and Future Recommendations

7 Conclusion and Future Recommendations

7.1 Summary of Major Contributions

In this thesis, we successfully developed expanded TPU (E-TPU) bead foams with desirable hard

segment (HS) crystal melting peak structure for the sintering of the beads using steam-chest

molding machine. The effect of melt processing, micro/nano-additives and different gases on the

crystallization and phase separation behavior of the HSs in the TPU microstructure was

systematically investigated using regular DSC, high-pressure differential scanning calorimetry

(HP-DSC) and a specially designed saturation system for liquid-state hydrocarbons. It is shown

that the phase-separation and crystallization that can be induced in presence of different gas at

different pressures and also in the presence of micro/nano additives can significantly influence

the TPU’s foaming behavior (i.e. cell nucleation and expansion behavior). The HS crystallites

were also successfully utilized to create a very strong sintering of the E-TPU beads into three

dimensional rectangular parts using the steam-chest molding machine. The important

conclusions are given below.

7.1.1 Effect of processing, nano-/micro-sized additives and dissolved gas on the phase separation and crystallization behavior of TPU

The phase separation and crystallization behavior of TPU is very sensitive to the processing

conditions. There has been extensive research work published in the literature regarding the

phase separation and crystallization behavior of TPU at atmospheric pressure (1 bar). However

there has not been any research work reported in the literature to investigate the effect of high-

pressure dissolved gas on the crystallization behavior of TPU. In this PhD work, for the first time

the crystallization behavior of TPU in the presence of dissolved gas has been systematically

investigated and published. The crystallization behavior of dissolved CO2 was investigated using

a HP-DSC. However, to investigate the effect of aliphatic hydrocarbon (butane), which cannot be

used in a HP-DSC, a specially designed high-pressure saturation system was developed. It was

observed that the presence of dissolved CO2 and butane induced a large number of less perfected

HS crystallites, which was a result of increase in the HS crystal nucleation mechanism during the

179

cooling from the annealing temperature after the completion of the annealing process. The

blending of GMS with TPU significantly improved the phase separation and crystallization

behavior of TPU. The presence of GMS acted as a lubricating agent and assisted the HS chains

to stack into higher degree of perfection and also assisted in the growth of HS crystallites to form

spherulitic crystals. Thus the overall crystallinity of the TPU was significantly improved after

annealing with CO2, butane and GMS. The presence of nano-clay and nano-silica did not

significantly affect the HS phase separation and crystallization behavior in the TPU

microstructure both independently and in synergy with dissolved gas. Another interesting

observation was the effect of melt-compounding, which resulted in the breakage of HS chains

and assisted the chains to stack and form HS crystallites with higher degree of perfection.

Overall the increase in the phase-separation and crystallization of HSs due to annealing with

dissolved gas and with GMS resulted in improved SS purity, which was observed with decrease

in the glass transition temperature. Thus the SS elasticity is also improved as a result the

annealing with dissolved gas and GMS compared to annealing at ambient pressure.

7.1.2 Effect of HS crystallites on the foaming behavior of TPU

In this study, a novel technique of utilizing the HS domains in the TPU microstructure was used

to prepare microcellular TPU foams using butane as the foaming agent over a wide range of

foaming conditions. Since butane has not been used for microcellular plastics because of its low

volatility and high solubility, and thereby low thermodynamic instability generated from the

rapid solubility drop, it is interesting to note the microcellular cell nucleation induced with the

butane used in this study. Although butane generates a relatively low thermodynamic instability,

its impact on the crystallization caused microcellular nucleation. It was observed that the melt

processing of AR-TPU caused a breakage of the HS chains. Thus, the PR-TPU sample showed

broad distribution of HS domains, which also included some highly ordered HS nano-crystals

with very high melting temperature. Moreover, the saturation temperature and butane’s

plasticizing impact significantly induced larger content of HS domains with higher perfection in

the PR-TPU. Consequently, without addition of any nucleating agents, the cell nucleation was

promoted in the vicinity of the largely distributed and perfected HS domains over a wide

saturation temperature range of 150°C-170°C at a saturation pressure of 55 bar. Overall, the PR-

TPU showed very a high nucleation rate compared to the AR-TPU due to the presence of broad

HS domains in their microstructure.

180

The crystallization kinetics of TPU was significantly improved in the presence of GMS and

dissolved butane, which resulted in the formation of large number of less perfect HS crystallites

dispersed in the SS matrix whereas some highly perfected HS crystals are also formed. Unlike its

low volatility and high solubility, butane was successfully utilized in the fabrication of

microcellular TPU foams. This was facilitated through the impact of butane on the crystallization

of HSs. The HS crystallites acted both as heterogeneous nucleating sites as well as reinforcement

leading to the microcellular morphology with a high expansion ratio in TPU-GMS samples.

Consequently, without addition of any nucleating agents, cell nucleation was promoted in the

vicinity of the largely distributed and perfected HS domains over a wide saturation temperature

range of 150-170°C at a saturation pressure of 55 bar. Overall, the TPU-GMS showed very high

nucleation rates compared to the neat-TPU.

This study also investigated the effect of water and super-critical CO2 as co-blowing agents for

the production of PR-TPU and TPU nano-clay nanocomposite microcellular foams at a moderate

CO2 pressure of 55 bar and saturation time of 60 min. The cell density increased significantly

due to the synergistic effects of nano-clay particles and the HS crystalline domains acting as

bubble nucleation sites.

7.1.3 Effect of HS crystallites on the foaming behavior of TPU

Steam is a powerful medium for transferring heat rapidly and therefore it is commonly used in

polymer bead foam sintering. But because of the thermodynamic property, the pressure loss

unavoidable during flow through the beads causes a temperature decrease and thereby negatively

affects the sintering behavior and the mechanical properties in the core of steam-chest molding.

In order to reduce the sensitivity of the temperature to the pressure variation inside the mold, hot

air was added to the steam line. The introduction of hot air was optimized to investigate the

effect of different parameters such as the hot air flow rate, the hot air temperature and the hot air

pressure, while the steam pressure was kept constant. The steaming time decreased by 32% and

the local temperature at entry port decreased by 8% at the highest available air flow rate of 120

l/min. The overall heat transfer improved significantly with an increase in the hot air flow rate.

The surface roughness values (Ra and Rz) decreased by approximately 50% at the hot air flow

rate of 120 l/min. An increase in the hot air pressure also showed a decrease in the surface

181

waviness (Wa) by approximately 50%. However, varying the hot air temperature did not cause

any significant change on the surface property.

The use of hot air with steam showed significant improvement in the overall consistency of the

tensile property across the molded EPP part as compared to the samples molded with pure steam.

The corresponding consistency was achieved due to an improved heat flow to the core of the

molded sample. This is possible due to the synergistic effect of the high thermal conductivity of

steam and the low Joule-Thompson coefficient of hot air. With either an increase in the hot air

flow rate or in the pressure, the heat flow is improved leading to an overall improvement in the

tensile property. Hence the results of this work reveal the potential application of hot air in the

steam-chest molding process to produce EPP bead products with improved surface quality,

enhanced mechanical properties and a shortened cycle time resulting in a reduced operating cost.

7.1.4 Lab-scale autoclave processing of E-TPU beads and sintering with steam-chest molding machine

In a lab-scale autoclave bead foaming setup, E-TPU beads with desirable crystalline structure

was manufactured and the beads were sintered using a steam-chest molding machine. The effect

of varying HS concentration on the bubble nucleation and the cell density of the processed E-

TPU beads was investigated. It was observed that the perfection in the existing HS crystalline

domains and the new HS crystallites developed during the saturation process induced a higher

degree of bubble nucleation which resulted in high cell density with smaller cell size in the E-

TPU bead foams. The processing techniques to produce E-TPU beads were also investigated. It

was observed that processing with pressure-drop method in the presence of water significantly

improved the overall foaming behavior (cell density and expansion ratio) of E-TPU beads

compared to the pressure-drop method without water. The presence of water resulted in higher

plasticization of the TPU matrix, which caused formation of some highly perfected HS crystals.

On the other hand, a large number of less-perfected HS crystallites (i.e. nucleation) also were

formed after processing E-TPU beads with water. These HS crystallites improved the cell

nucleation of E-TPU beads. Another interesting observation was the significant increase in the

cell nucleation at the critical saturation temperature, which was also reduced by 15-20ºC due to

the plasticization effect of water.

182

Compared to the pressure-drop methods, the temperature-jump method was found to be more

effective to achieve microcellular E-TPU beads due to the existence of large number of less-

perfected HS crystallites, which can be utilized as heterogeneous bubble nucleating agents. The

E-TPU beads produced via temperature-drop method also showed very good bead-to-bead

sintering using the steam-chest molding machine compared to the samples produced with the

pressure-drop method with water. It was also observed that the TPUs with lower HS

concentrations were more effective for sintering of the E-TPU beads using steam-chest molding

machine.

7.2 Summary of Major Contributions (Publications)

Published Journal Articles

1- Hossieny, N., Barzegari, M.R., Nofar M., Mahmood S.H., Park, C.B., “Crystallization of

Hard Segment Domains With the Presence of Butane for Microcellular Thermoplastic

Polyurethane Foams”, Polymer, 2014, 55, 651-662. (Chapter 3 and 4)

2- Hossieny, N., Ameli, A., Park, C.B., “Characterization of Expanded Polypropylene Bead

Foams With Modified Steam-Chest Molding”, Industrial & Engineering Chemistry Research,

2013, 52 (24), 8236-8247. (Chapter 5)

Submitted and Ready to Submit Articles

1- Hossieny, N., Shaayegan, V., Ameli, A., Saniei, M., Jahani, D., Park, C.B., “Effects of

Glycerol Monosterate and Butane on Phase Separation of Hard Segment and Its Impact on

Microcellular Thermoplastic Polyurethane Morphology”, RSC Advances, submitted in May

2014 and minor revision received. (Chapter 3 and 4)

2- Hossieny, N., Raps, D., Altstädt, V., Park, C.B., “Past and Present Developments in Bead

Foams and Bead Foaming Technology”, Polymer, submitted in July 2014. (Chapter 2)

3- Hossieny, N., Ameli, A., Saniei, M., Park, C.B., “Expanded Thermoplastic Polyurethane

Beads: Thermal, Foaming and Sintering Behaviors”, ready and will be submitted in August

2014. (Chapter 5)

183

7.3 Recommendations for Future Research

The processing of E-TPU bead foams with the conventional pressure-drop bead foaming method

provided some very interesting insight on the effect of water. The water plasticized the TPU and

decreased the processing temperature. The HS crystalline domains were also affected with the

presence of water. However the cell morphology was not seen to improve with the changes in the

HS crystalline domain. Especially, considering other bead foams such as EPP and EPLA, which

successfully utilized the crystals generated during the bead processing to produce microcellular

morphologies. The HS crystalline domains during E-TPU processing was not used effectively to

produce microcellular morphologies. Hence for future work, it would be very interesting to

systematically investigate using the simulation system described in Chapter 4 and decouple the

effect of water on the HS crystalline domains and their subsequent effect on the foam

morphology of TPU.

Another area to explore based on the scientific knowledge of the E-TPU bead foams generated

from this research is to develop other bead foam materials based on elastomeric materials .

Polyether-block-amide (Pebax) is a very good material to initially investigate based on the E-

TPU beads developed in this research. The processing of the beads can also be investigated using

continuous efficient process with extruder and an underwater pelletizer. Further, the sintering of

the investigated and developed elastomeric beads materials can be systematically investigated

using the modified steam-chest molding machine, which was developed in this research. The use

of hot air can reduce the moulding time and energy consumption, which would significantly

promote the use of expanded elastomeric materials for a wide range of applications. The

mechanical properties of the parts can also be improved with hot air thus making it very

attractive in a variety of industrial applications.