‘ ‘nmfm rW. 2:7,“... yth. 1.1.2713 . . u. r. 36nd. “$1,. “1.4%.. Erma ,. 2...» . 4.. r4 .. 4%... v. , a... he». 5: . .. 0V2 .. ....:..»r..., . . . Li.- 21.14.. .3. Luna... .3. .. 4,1 LL} w? a 3.3%... . .axmungmfiv a. .W H. La. .3 ., ink .v u :71 ‘91 .M an .0 ‘ Z: and» . . L: :t a. {it .0 “a. . .f . . .u‘ .3»? . 1., 1.. .7 ... .3? v.. . L 2.: a. r . 7‘7, .. é ., :1 1.: 4 .‘ . 54.... PAP»: V ntmh. u” £28.: .\‘o >\ h v 4 .1 . . V1.1? ta Jamil] This is to certify that the dissertation entitled The Effect of Reaction Processing on the Physical, Morphological, and Barrier Properties of Poly(ethylene 2,6-naphtha1ene dicarboxylate)/ Poly(ethylene terephthalate) Blends presented by Rujida Uthaisombut has been accepted towards fulfillment of the requirements for Ph.D . degree in Packaging Major professr mama? /I. 0:2 0&/ MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 LIBRARY Michigan State Unlversity PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE Zilll 060511 6/01 cJClRC/DateDuepGS—p. 15 THE EFFECT OF REACTION PROCESSING ON THE PHYSICAL, MORPHOLOGICAL, AND BARRIER PROPERTIES OF POLY(ETHYLENE 2,6-NAPHTHALENE DICARBOXYLATE)/ POLY(ETHYLENE TEREPHTHALATE) BLENDS By Rujida Uthaisombut A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements For the degree of DOCTOR OF PHILOSOPHY School of Packaging 2001 ABSTRACT THE EFFECT OF REACTION PROCESSING ON THE PHYSICAL, MORPHOLOGICAL, AND BARRIER PROPERTIES OF POLY(ETHYLENE 2,6-NAPHTHALENE DICARBOXYLATE)/POLY(ETHYLENE TEREPHTHALATE) BLENDS By Rujida Uthaisombut PET/PEN blends have been considered as a new material, combining the economic advantages of PET with the high barrier and thermal properties of PEN, which can extend the limitatiOn of PET packaging applications. However, PET and PEN are' inherently incompatible. Extrusion of PET/PEN blends above the melting temperature of both polymers results in transesten'fication, which was shown to enhance the miscibility of the blend and improve both the clarity and barrier properties of the resultant films. This research focused on the study of the effect of the processing conditions and the blend composition on the transesterificatilon reaction. In addition, the effect of degree of transesterification, the blend composition, and the orientation on the blend characteristics was also studied. The results showed that the primary factors controlling the transesterification reaction were blending time and temperature, whereas the composition of the blends was found to have little or no effect on the interchange reaction. The PET/PEN blends with a degree of transesterification of at least 6% were miscible, homogeneous and optically clear. Further the transesterification reaction had less of an effect on the Tg, the barrier, and the mechanical properties. However, the Tm, %crystallinity, density, and molecular weight average were dependent on the degree of transesterification, regardless of how this degree of transesteiification was achieved. The blend composition was a very important factor controlling the thermal, mechanical, and barrier properties of the blends. A depression of the Tm, % crystallinity, and molecular weight was found, when a small amount of PEN was added to the PET rich phase, or vice versa. The Tm, % crystallinity and molecular weight average were lowest, when the PEN composition was 40-50 mole %. Moreover, the density of the blends decreased, as the PEN composition increased. However, the T8, the gas barrier properties and the tensile strength of the blends improved when the PEN composition increased. Orientation process enhanced the barrier and the mechanical properties of the blend films. The water vapor, the oxygen, and the carbon dioxide barrier properties of the PET/PEN blend films improved 1.5-2 times, as the blend films were 3x3 biaxially oriented. The tensile strength of the oriented blend films was 4 times greater than that of the non-oriented blend films. Moreover, oriented PEN, and the oriented blend films were flexible, while the non-oriented films were very brittle. To the memory of Dr. Jack R. Giacin (December 31, 1938 — December 12, 2000) ACKNOWLEDGEMENT I would like to take this opportunity to express my gratitude to many people who contributed in many ways to make my dissertation possible. First, I wish to express my deepest appreciation to Dr. Jack R. Giacin, who served as my advisor. I thank him for his guidance, kindness, encouragement, and effort. He always had an open door for me to ask any questions and to ask for any advises. During the hard time for him, due to his health, he never stopped giving good comments to improve this dissertation. He was the greatest teacher I have ever known. I would like to thank Dr Bruce Harte, who served as my second advisor. He always gave good advises and supported me in many ways. I thank Dr. Jayaraman, Dr. Susan Selke, and Dr. Ruben Hernandez, the members of my committee for their valuable contIibutions throughout my graduate study. I would like to thank my parents, Kwan and Wannee Leepipattanawit, for their love, support throughout my life, their encouragement to pursue higher education, and their teaching and being great examples of how to be a good person. I wish to thank my husband, A Dr. Patchrawat Uthaisombut for his love, understanding, encouragement, continuous support, and patience during my graduate study. I would like to thank Dr. Robert Rasche and Mrs. Dorothy Rasche for their support and kindness, during my stay at Michigan State University. I would like to thank many friends (Ubonrat Siripatrawan, Lee Youn Suk, Cengiz Caner, and others) for their warm hospitality and support. I thank the Center for Food & Pharmaceutical Packaging Research (CFPPR) for their financial support throughout my graduate program. I thank Mr. Michael J. Rich from the Composite Materials and Structures Center at MSU, and Mr. Johnson Kermit and Dr. Long Le fiom NMR laboratory, Chemistry Department at MSU for their kind help through my work on thermal and NMR analysis. I wish to thank Dr. Issam Dairanich and Amoco Research Center, Amoco Chemicals Company for the technical support and film production. I thank to Shell Company for the donation of PET and PEN samples. TABLE OF CONTENTS LIST OF TABLES .......................................................................................... x LIST OF FIGURES ..................................................................................... xiv Chapter 1: Introduction 1 1.1. Need for Modified PET ................................................................................................ 1 1.2. Alternative Polymer ...................................................................................................... 2 1.3. Combining PET and PEN ............................................................................................. 3 1.4. Transesterification Reactions ........................................................................................ 4 1.5. Overview of the Experiment ......................................................................................... 5 1.5.1. Processing Conditions....................... ............................................................. 5 1.5.2. Blend Composition ........................................................................................ 5 1.5.3. Film Orientation ............................................................................................. 6 1.5.4. Blend Characterization ................................................................................... 6 1.6. Objectives .......................................................................................................... 7 Chapter 2: Literature Review 8 2.1. Interesting Issues of PET .............................................................................................. 8 2.1.1. PET Structure and Properties ..................................................................... ....9 2.1.2. PET Manufacturing ...................................................................................... 10 2.1.3. PET Applications ......................................................................................... 13 2.1.4. PET Modification ......................................................................................... 16 2.2. Alternative Polymer, PEN .......................................................................................... 17 2.3. Combining PET and PEN ........................................................................................... 23 2.4. PET/PEN Blends and Copolymers ............................................................................. 25 Chapter 3:Transesterification Reactions 30 3. 1. Introduction ................................................................................................................. 30 3.1.1. Miscibility of Polymer Blends ..................................................................... 30 . 3.1.2. Transesterification Reactions ...................................................................... 31 3.1.2.1. Mechanisms of transesterification reactions ................................. 33 3.1.2.2. Factors controlling transesterification reactions ........................... 36 3.2. Materials and Methods ................................................................................................ 41 3.2.1. Materials ...................................................................................................... 41 3.2.2. Reaction Processing Parameters and Blend Composition ........................... 41 3.2.3. Extrusion ....................................................................... , .............................. 44 3.2.4. DSC Analysis ........................................................................................ ' ....... 4 5 3.2.5. ‘H-NMR Analysis ........................................................................................ 47 3.2.6. Calculation of Degree of Transesterification and D value .......................... 49 3.3. Results and Discussion .............................................................................................. 52 3.3.1. DSC Analysis ............................................................................................... 52 3.3.2. Degree of Transesterification (B) ................................................................ 55 3.3.2.1. Effect of blending time on transesterification reactions ............... 56 Screw speed ................................................................................... 56 vii Numbers of passes ........................................................................ 57 3.3.2.2. Effect of blending temperature on transesterification reactions ...58 3.3.2.3. Effect of blend composition on transesterification reactions ........ 60 3.3.3. D value ......................................................................................................... 62 Chapter 4: Characterization of PET/PEN Blends 65 4.1 . Introduction ................................................................................................................. 65 4.1.1. Thermal Properties ...................................................................................... 65 4.1.2. Molecular Weight and Molecular Weight Distribution ............................... 68 4.1.3. Density ......................................................................................................... 69 4.1 .4. Morphology .................................................................................................. 69 4.2. Materials and Methods ................................................................................................ 71 4.2.1. Film Fabrication Process .............................................................................. 71 4.2.2. Orientation Process ...................................................................................... 73 4.2.3. Analysis Methods ......................................................................................... 74 4.2.3.1. Thermal analysis ........................................................................... 74 4.2.3.2. Density measurement .................................................................... 75 4.2.3.3. Average sequence lengths of terephthalate and naphthalate units 79 4.2.3.4. Molecular weight measurement .................................................... 80 4.2.3.5. Morphological analysis ................................................................. 80 4.3. Results and Discussion .............................................................................................. 82 4.3.1. Thermal Properties of Blends ...................................................................... 82 4.3.1.1. Glass transition temperature (T3) .................................................. 83 4.3.1.2. Melting temperature (Tm) .............................................................. 89 4.3.1.3. Percent crystallinity of the blends ................................................. 93 4.3.2. Density of Blend Resins and Films .............................................................. 95 4.3.3. Average Sequence Lengths of Terephthalate and Naphthalate Units ........ 103 4.3.4. Molecular Weight of Blends ...................................................................... 103 4.3.5. Morphology ..... 110 Chapter 5: Barrier and Mechanical Properties of PET/PEN Blends 120 5.1 . Introduction ............................................................................................................ 120 5.1.1. Barrier Properties ....................................................................................... 121 5.1.2. Mechanical Properties ................................................................................ 123 5.1.2.1. Factors controlling mechanical properties ............................... 124 5.1.2.2. Tensile properties ........................................................................ 126 5.2. Materials and Methods .............................................................................................. 128 5.2.1. Measurement of Water Vapor Transmission Rate ..................................... 128 5.2.2. Measurement of Oxygen Transmission Rate ............................................. 130 5.2.3. Measurement of Carbon Dioxide Transmission Rate ................................ 131 5.2.4. Tensile Test ................................................................................................ 135 5.3.Results and Discussion .............................................................................................. 138 5.3.1. Barrier Properties ...................................................................................... 138 5.3.1.1. Water vapor permeability coefficient ......................................... 138 . 5.3.1.2. Oxygen permeability coefficient ................................................. 142 5.3.1.3. Carbon dioxide permeability coefficient .................................... 146 viii 5.3.2. Mechanical Properties ................................................................................ 150 5.3.2.1.Tensile strength ............................................................................ 150 5.3.2.2. Percent elongation at break ......................................................... 155 5.3.2.3. Young’s modulus of elasticity ................................................. 160 Chapter 6: Orientation Effect 164 6.1. Introduction 164 6.1.1. Uniaxial Orientation ................................................................................... 165 6.1 .2. Biaxial Orientation ..................................................................................... 165 6.1.3. Effect of Orientation on Polymer Properties ............................................. 166 6.1.3.1. Percent crystallinity .................................................................... 166 6.1.3.2. Barrier properties ........................................................................ 167 6.1.3.3. Mechanical properties ................................................................. 167 6.1.3.4. Effect of orientation on selected properties of PET/PEN blendsl68 6.2. Materials and Methods .............................................................................................. 169 6.3. Results and Discussion ............................................................................................ 170 6.3.1. Effect of Orientation on Thermal Properties ............................................. 172 6.3.1.1. Glass transition temperature ....................................................... 172 6.3.1.2. Melting temperature .................................................................... 173 6.3.1.3. Percent crystallinity .................................................................... 173 6.3.2. Effect of Orientation on Density of the Blends ....................................... 173 6.3.3. Effect of Orientation on Barrier Properties ............................................... 174 6.3.4. Effect of Orientation on Mechanical Properties ........................................ 174 6.4. Summary ................................................................................................................... 176 Chapter 7: Summary and Conclusions 177 7.1. Transesterification Reactions .................................................................................... 177 7.2. Blend Characteristics ................................................................................................ 179 7.2.1. Effect of Degree of Transesterification ..................................................... 179 7.2.2. Effect of Blend Composition .................................................................... 181 7.2.3. Effect of Orientation .................................................................................. 183 Appendices ................................................................................................. 186 Bibliography .............................................................................................. 219 ix List of Tables Tables page Chapter 3 3.1. Processing parameters and blend composition. .......................................................... 43 3.2. The effect of the extruder screw speed on the degree of transesterification (B). ....... 57 3.3. The effect of blending temperature and time on the degree of transesterification (B). ......................................................................................................................... 59 3.4. The effect of PEN composition on the degree of transesterification (B). ................... 61 3.5. The effect of the blending temperature and time on the degree of transesterification (B) and D-value. ....................................................................... 63 3.6. The effect of PEN composition on the degree of transesterification (B) and D-value. .................................................................................................................. 64 Chapter 4 4.1. List of the samples tested for Raman chemical imaging analysis. ............................. 82 4.2. The effect of the degree of transesterification (B) on the glass transition temperature of the PET/PEN blends. ..................................................................... 84 4.3. The effect of PEN composition on the glass transition temperature on the PET/PEN blends. ................................................................................................... 86 4.4. The comparison between calculated and measured T8 of the blend resins. ................ 88 4.5. The effect of the degree of transesterification (B) on the melting temperature (T...) of the PET/PEN blends. ......................................................................................... 90 4.6. The effect of PEN composition on the melting temperature of the PET/PEN blends. .................................................................................................................... 92 4.7. The effect of the degree of transesterification (B) on the percent crystallinity of the PET/PEN blends. ............................................................................................. 94 4.8. The effect of PEN composition on the percent crystallinity of the PET/PEN blends. .................................................................................................................... 96 4.9. The effect of the degree of transesterification on the density of the blends. .............. 98 4.10. The effect of PEN composition on the density of the PET/PEN blend resins. ....... 100 4.11. The density ofthe PET, PEN and blends ......... 101 4.12. The effect of PEN composition on the density of the PET/PEN blend films. ........ 102 4.13. The effect of the degree of transesterification (B) on the average sequence lengths of the terephthalate (1.“) and naphthalate (LnN) unit. ............................. 104 4.14. The effect of PEN composition on the average sequence lengths of the terephthalate (141T) and naphthalate (LnN) unit. .................................................. 105 4.15. The molecular weight average of the blend films ................................................... 106 4.16. The effect of the degree of transesterification on the polydispersity (MW/Mn) of the blend films ...................................................................................................... 108 4.17. The effect of the PEN composition on the polydispersity (MW/Mn) of the blend films ..................................................................................................................... 109 List of Tables (cont.) Chapter 5 5.1. The effect of the degree of transesterification on the water vapor permeability coefficient of PET/PEN blend fihns at 378°C. ................................................... 139 5.2. The effect of PEN composition on the water vapor permeability coefficient of the PET/PEN blend films at 37 .8°C. .......................................................................... 141 5.3. The effect of the degree of transesterification on the oxygen permeability coefficient of PET/PEN blend films at 25°C ....................................................... 143 5.4. The effect of PEN composition on the oxygen permeability coefficient of ' PET/PEN blend films at 25°C. ............................................................................. 145 5.5. The effect of the degree of transesterification on the carbon dioxide permeability coefficient of PET/PEN blend films at 25°C. ...................................................... 147 5.6. The effect of PEN composition on the carbon dioxide permeability coefficient of PET/PEN blend films at 25°C. ............................................................................. 148 5.7. The effect of the transesterification degree (B) on the tensile strength of PET/PEN blend films ........................................................................................... 151 5.8. The effect of PEN composition on the tensile strength of PET/PEN blend films. ...152 5.9. The effect of the transesterification degree (B) on the % elongation of PET/PEN blend films. .......................................................................................................... 156 5.10. The effect of PEN composition on the % elongation of PET/PEN blend films. ....157 5.11. The effect of the transesterification degree (B) on the modulus of elasticity of PET/PEN blend films ........................................................................................... 161 5.12. The effect of PEN composition on the modulus of elasticity of PET/PEN blend films. .................................................................................................................... 162 Chapter 7 7.1. Summary of effect of degree of transesterification and blend composition on selected properties of PET /PEN blends. .............................................................. 180 7.2. The equations showing the relations between the blend composition and the barrier and mechanical properties of the blends obtained at 300°C and 1 pass. ..182 Appendices Appendix A A1. Analysis variance of degree of transesterification of PET/PEN blend resins and films. .................................................................................................................... 187 A2. Analysis variance of D-values of PET/PEN blend resins and films. ........................ 188 A3. Analysis variance of degree of transesterification of blends with compositions of 14.5 to 50.4 mole % PEN ..................................................................................... 189 A4. Analysis variance of D-values of blends with compositions of 14.5 to 50.4 mole % PEN .................................................................................................................. 190 A5. Analysis variance of D-values of blends with compositions of 33.7 to 50.4 mole % PEN .................................................................................................................. 191 xi List of Tables (cont.) Appendix B Bl. The positions in the gradient column and the density values of the standard calibration floats ................................................................................................... 192 32. The positions and the density of the blend resins. .................................................... 193 B3. The positions and the density of the non-oriented films. .......................................... 193 B4. The positions and the density of the oriented films. ................................................. 194 Appendix C Cl. Analysis of variance of the T8 values of the blend resins with 25% (wt/wt) PEN. .. 195 C2. Analysis of variance of the Tg values of the oriented films with 25% (wt/wt) PEN. ..................................................................................................................... 196 C3. Analysis of variance of the calculated and measured T3 values of the blend resins processed at 300°C and 1 pass. ............................................................................ 197 Appendix D D1. Analysis of variance of the water vapor permeability coefficient of the non- oriented films obtained from blend resins with 25 % (wt/wt) PEN ..................... 198 D2. Analysis of varianCe of the water vapor permeability coefficient of the oriented films obtained from blend resins with 25 % (wt/wt) PEN. .................................. 199 D3. Analysis of variance of the water vapor permeability coefficient of blend films with compositions of 14.5 to 50.4 mole % PEN .................................................. 200 D4. Analysis of variance of the oxygen permeability coefficient of the non-oriented films obtained from blend resins with 25 % (wt/wt) PEN. .................................. 201 D5. Analysis of variance of the oxygen permeability coefficient of the oriented films obtained from blend resins with 25 °/o (wt/wt) PEN. ........................................... 202 D6. Analysis of variance of the oxygen permeability coefficient of the blend films with composition of 14.5 to 50.4 mole % PEN. .................................................. 203 D7. Analysis of variance of the carbon dioxide permeability coefficient of the non- . oriented films obtained from blend resins with 25 % (wt/wt) PEN ..................... 204 D8. Analysis of variance of the carbon dioxide permeability coefficient of the oriented films obtained from blend resins with 25 % (wt/wt) PEN ..................... 205 D9. Analysis of variance of the carbon dioxide permeability coefficient of the blend films with composition of 14.5 to 50.4 mole % PEN. ......................................... 206 Appendix E . E1. Analysis of variance of the tensile strength along the machine direction of the non-oriented films with 25 % (wt/wt) PEN. ........................................................ 207 E2. Analysis of variance of the tensile strength along the cross machine direction of the non-oriented films with 25 % (wt/wt) PEN. .................................................. 208 E3. Analysis of variance of the tensile strength of the oriented films obtained from blend resins with 25 % (wt/wt) PEN .................................................................... 209 xii List of Tables (cont.) Appendix F F 1. Analysis of variance of the % elongation along the machine direction of the non- oriented films with 25 % (wt/wt) PEN. ............................................................... 210 F2. Analysis of variance of the % elongation along the cross machine direction of the non-oriented films with 25 % (wt/wt) PEN. ........................................................ 211 F3. Analysis of variance of the % elongation of the oriented films obtained from blend resins with 25 % (wt/wt) PEN .................................................................... 212 Appendix G G1. Analysis of variance of the Young’s modulus along the machine direction of the non-oriented films with 25 % (wt/wt) PEN. ........................................................ 213 G2. Analysis of variance of the Young’s modulus along the cross machine direction of the non-oriented films with 25 % (wt/wt) PEN. .............................................. 214 G3. Analysis of variance of the Young’s modulus of the oriented films with 25 % (wt/wt) PEN. ........................................................................................................ 215 G4. Analysis of variance of the Young’s modulus along the machine direction of the non-oriented films with composition of 14.5 to 50.4 mole PEN % ..................... 216 G5. Analysis of variance of the Young’s modulus along the cross machine direction of the non-oriented with composition of 14.5 to 50.4 mole % PEN .................... 217 G6. Analysis of variance of the Young’s modulus of the oriented films with composition of 14.5 to 50.4 PEN mole % PEN. .................................................. 218 xiii List of Figures Figures page Chapter 2 2.1. Chemical structure of polyethylene terephthalate (PET). ............................................. 9 2.2. Polymerization of PET, (a) Ester interchange reaction and (b) Polycodensation ....... 12 2.3. Chemical structure of polyethylene naphthalate (PEN) .............................................. 17 2.4. Polymerization of PEN, (a) Ester interchange reaction and (b) Polycodensation. ..... 20 Chapter 3 3.1. NMR spectrum of the PET/PEN blend with 25% (wt/wt) PEN processed through the twin-screw extruder at 300°C and 1 pass. ........................................................ 48 3.2. DSC scan of the PET/PEN blend 25% (wt/wt) PEN processed through the single- screw extruder at 300°C and 1 pass. ...................................................................... 53 3.3. DSC scan of the PET/PEN blend 25% (wt/wt) PEN processed through the twin- screw extruder at 300°C and 1 pass. ...................................................................... 54 3.4. The effect of blending temperature and time on the degree of transesterification (B). ......................................................................................................................... 59 3.5. The effect of PEN composition on the degree of transesterification (B) .................... 61 3.6. The effect of the blending temperature and time on the degree of transesterification (B) and D value. ....................................................................... 63 3.7. The effect of PEN composition on the degree of transesterification (B) and D value. ...................................................................................................................... 64 Chapter 4 4.1. Apparatus for the gradient tube preparation. .............................................................. 78 4.2. The effect of the degree of transesterification (B) on the glass transition temperature of the PET/PEN blends. ..................................................................... 84 4.3. The effect of PEN composition on the glass transition temperature on the PET/PEN blends. ................................................................................................... 86 4.4. The comparison between calculated and measured T8 of the blend resins. ................ 88 4.5. The effect of the degree of transesterification (B) on the melting temperature (Tm) of the PET/PEN blends. ......................................................................................... 90 4.6. The effect of PEN composition on the melting temperature of the PET/PEN blends. .................................................................................................................... 92 4.7 . The effect of the degree of transesterification (B) on the percent crystallinity of the PET/PEN blends. ............................................................................................. 94 4.8. The effect of PEN composition on the percent crystallinity of the PET/PEN blends. .................................................................................................................... 96 4.9. The effect of the degree of transesterification on the density of the blends. .............. 98 4.10. The effect of PEN composition on the density of the PET/PEN blend resins. ....... 100 4.11. The density of the PET, PEN and blends ................................................................ 101 4.12. The effect of PEN composition on the density of the PET/PEN blend films. ........ 102 xiv List of Figures (cont.) 4.13. The effect of the degree of transesterification (B) on the average sequence lengths of the the terephthalate (LnT) and naphthalate(LnN) unit. ........................ 104 4.14. The effect of PEN composition on the the average sequence lengths of the terephthalate (La) and naphthalateamn) unit. ..................................................... 105 4.15. The effect of PEN composition on the molecular weight average of blend films. .106 4.16. The effect of the degree of transesterification on the polydispersity mm") of the blends. ............................................................................................................ 108 4.17. The effect of PEN composition on the polydispersity (MW/Mn) of the blend films. .................................................................................................................... 109 4.18. Dispersive Raman Spectroscopy of 300P1 25% (wt/wt) PEN polyblend samples ................................................................................................................. 113 4.19. Dispersive Raman Spectroscopy PET/PEN blend resins with 25% and 40% (wt/wt) PEN processed at 300 °C and 1 pass, and 315°C and 2 passes. .............. 114 4.20. Raman Chemical Imaging with 20X Magnification of 25% (wt/wt) PEN blend resins processed through single-screw extruder at 300 0C and 1 pass. ............... 115 4.21. Raman Chemical Imaging with 20X Magnification of 25% (wt/wt) PEN blend resins processed through twin-screw extruder at 300 0C and 1 pass ................... 116 4.22. The brightfield (a), polarized light (b), and Raman images (c) fi'om 20x magnification of the blend resins. ........................................................................ 117 4.23. The brightfield (a), polarized light(b), and Raman images (c) from 100x magnification of the blend resins. ........................................................................ 118 4.24. The brightfield (a), polarized light(b), and Raman images (c) from 20X and 100x magnification of the blend films with 25% (wt/wt) PEN processed through twin-screw extruder at 3000C and 1 pass. .............................................. 119 Chapter 5 5a. A typical stress-strain curve for elastic materials. .................................................... 127 - 5b. Example of Permatran CIV graphical data. .............................................................. 134 5.1. The effect of the degree of transesterification on the water vapor permeability coefficient of PET/PEN blend films at 37. 8°C. ................................................... 139 5.2. The effect of PEN composition on the water vapor permeability coefficient of the PET/PEN blend films at 37 .8°C. .......................................................................... 141 5.3. The effect of the degree of transesterification on the oxygen permeability coefficient of PET/PEN blend films at 25°C. ...................................................... 143 5.4. The effect of PEN composition on the oxygen permeability coefficient of PET/PEN blend films at 25°C. ............................................................................. 145 5.5. The effect of the degree of transesterification on the carbon dioxide permeability coefficient of PET/PEN blend films at 25°C. ...................................................... 147 5.6. The effect of PEN composition on the carbon dioxide permeability coefficient of PET/PEN blend films at 25°C. ............................................................................. 149 5.7. The effect of the transesterification degree (B) on the tensile strength of PET/PEN blend films ........................................................................................... 151 XV List of Figures (cont.) 5.8. The effect of PEN composition on the tensile strength of PET/PEN blend films. 152 5.9. The effect of the transesterification degree (B) on the % elongation of PET/PEN blend films. .......................................................................................................... 156 5.10. The effect of PEN composition on the % elongation of PET/PEN blend films. l 57 5.11. The effect of the transesterification degree (B) on the modulus of elasticity of PET/PEN blend films ........................................................................................... 161 5.12. The effect of PEN composition on the modulus of elasticity of PET/PEN blend films. .................................................................................................................... 162 Appendix B Bl. The calibration curve for the density gradient column at 23°C ................................. 192 xvi Chapter 1 Introduction Over the past decade poly(ethylene terephthalate) (PET) has become popular in the food packaging industry because it is a clear, light weight material with good barrier characteristics to water vapor and carbon dioxide (Davis and Howell 1993). In addition, its processibility, resealability, shatter resistance, and recyclability have made PET the packaging material of choice for various food and pharmaceutical applications. 1.1. Need for Modified PET Even though PET has found widespread acceptance, improved barrier and thermal properties are required for extending product shelf life and for packaging systems which require high oxygen barrier performance, coupled with high thermal stability and UV protection. PET has fair oxygen barrier properties. For instance, carbonated drinks can be packaged in PET bottles. However, PET also cannot be used for packaging beer. Beer rapidly loses desirable flavor and aroma characteristics when packaged in PET (Po et a1. 1996), because such moieties are very vulnerable to any amount of oxygen, which permeates through PET bottles. PET is also a low UV barrier material. Therefore, it cannot be used for products such as medicines, baby foods, and beer, which are sensitive to UV light. These products are typically packaged in glass or metal packages. The glass transition temperature (T3) of PET is approximately 74°C. Thus, a typical PET resin is not suitable for use as hot fillable or refillable packages. Several products, including hit juices, jams, tomato sauces, etc., require a hot-fill process at a temperature above 85°C (McGee and Jones 1995). Therefore, a material with a temperature resistance of greater than 85°C is required. In addition, returnable or refillable PET bottles also need to be washed with an alkaline solution at a temperature close to the T8 of PET. Thin wall PET bottles shrink under such conditions. A few techniques to overcome these problems have been proposed. For example, bottles with a wall thickness of 6-7 mm and a weight of 100 g as compared to the 3 mm and 45 g of one-way bottles, have been used to overcome this problem (Po et a1. 1996). However, cost and weight limit their use. 1.2. Alternative Polymer A new type of material, which is qualified for use under such conditions, has, therefore, been sought. One such candidate polymer, which meets the above requirements is poly(ethylene 2,6-naphthalene dicarboxylate), an analog of PET. PEN possesses oxygen permeability characteristics approximately one quarter to one-fifth that of PET (Stewart et a1. 1993). PEN also has higher temperature resistance. The T8 of PEN is approximately 120°C, 50 °C above that of PET. In addition, PEN has higher UV light barrier properties and lower shrinkage at elevated temperatures than PET (Callander and Sisson 1994). The tensile strength of PEN is also 35% higher that that of PET (Shi 1998). While PEN has decided advantages over PET for packaging oxygen sensitive and hot-fill products, PEN costs approximately five time as much as PET. In addition, the limited amount of monomer production for PEN has restricted its use (Callander and Sisson 1994; Hoffinan and Caldwell 1995). 1.3. Combining PET and PEN A potential technique for combining the economic advantages of PET and the outstanding banier and thermal properties of PEN is copolymerization and blending process of PET and PEN. PET/PEN copolymer is an optically clear material with intermediate barrier, mechanical, and thermal properties between PET and PEN. However, McGee and Jones (1995) found that PET/PEN blends had advantages over random copolymers, because blends can crystallize over a broader compositional range than copolymers. Within 15-85 mole % of naphthalene dicarboxylate content (the mole fraction of the naphthalate unit of the total terephthalate and naphthalate units, NDC), blends are semicrystalline materials while copolymers are amorphous materials. Within this range of NDC, blends also have improved oxygen barrier properties, relative to their random copolymer counterparts (McGee and Jones 1995). Therefore, a PET/PEN blend was selected for this study. However, a drawback to the blend approach is that PET and PEN are inherently incompatible. The mixture of PET and PEN yields an opaque material with non-uniform properties, and shows two T3 values (glass transition temperatures) and Tm values (melting temperatures), corresponding to one PET-rich phase and one PEN-rich phase. Transesterification reactions between PET and PEN have been studied and found to improve the compatibility and ultimately produce miscibility of the blends, which in turn results in improved clarity and uniform properties. 1.4. Transesterification Reactions Transesterification reactions are interchange reactions between hydroxyl and carboxyl groups of the PET and PEN chains (see the details in Chapter 3). Transesterification reactions are accomplished by melt mixing these two polymers at temperatures that are higher than their melting temperatures. The properties and composition of the blend change when transesterification is achieved. During this reaction processing, graft and block copolymers are initially formed, followed by formation of random copolymers. In addition, the resulting copolymers compatibilize PET and PEN, resulting in miscible blends (McGee and Jones 1995; Po et al. 1996; Shi 1998; Stewart et a1. 1993). The PET/PEN blends with a transesterification degree of greater than 5% showed a significant improvement in the blend miscibility and had only a single T8 and Tm (Hoffman and Caldwell 1995 ; McGee and Jones 1995). McGee and Jones (1995) showed that there is an inverse linear relationship between the transesterification level and the haze value (measurement of the film opacity). The authors reported that as the degree of transesterification increases, the haze value decreases. 1.5. Overview of the Experiment 1.5.1. Processing Conditions In any scheme to produce materials from melt blends of PET and PEN, a number of factors have to be dealt with. Processing conditions, as well as types of processing equipment, are primary factors controlling transesterification reactions, which in turn affect the miscibility of the blends. Stewart et a1. (1993) studied the reactive processing of PET/PEN blends with a PEN composition ranging from 50 to 80 weight % and found that the primary factors connolling transesterification are the blending time and temperature, while the composition of the blend and the residual polyester catalysts have little or no effect on the transesterification reactions. Furthermore, processing the melt polymer in a twin-screw extruder was found to achieve a higher degree of transesterification than processing in a single screw extruder (Shi 1998). This study, therefore, focuses on the processing parameters, including processing time and temperature, for accomplishing the reaction by extruding the melt polymer blend in a twin-screw extruder. 1.5.2. Blend Composition Even though blend composition has little effect on the transesterification reactions, blend composition is a primary factor controlling thermal, barrier and mechanical properties of the resultant blends. Therefore, the effect of the blend composition on blend properties was investigated. This study focused on PEN composition ranging from 10 to 40 weight percent. 1.5.3. Film Orientation Orientation, which involves some degree of stretching and heat setting, is known to enhance the properties of polymer films. Rearrangement of the polymer chains, by application of an external stress above the polymer glass transition temperature, tends to increase the percent crystallinity of semicrystalline polymers and results in improvement of mechanical, thermal and barrier properties. The influence of orientation on the thermal, mechanical, and barrier properties of blend films, with a certain degree of transesterification, was studied. 1.5.4. Blend Characterization PET/PEN blend resins with varied transesterification degree and PEN composition were processed though a twin screw extruder at temperatures between 275 and 325°C, and one and two passes though the extruder. Blend films were fabricated in a single screw extruder at 300°C, and biaxially oriented on a TM. Long film stretcher. The properties of blend resins, including mechanical, thermal, barrier, and morphology, and films then were determined. The effect of the transesterification degree, blend composition, and orientation on these properties was then evaluated. 1.6. Objectives The followings are the primary objectives for this study: 1. To study the effect of processing conditions, blending time and temperature, on the transesterification reactions and determine the optimum processing conditions to achieve the desired blend properties. 2. To establish the relationship between the transesterification degree of the PET/PEN blends and the associated thermal, barrier, and mechanical properties, with transesterification degree evaluated as a function of processing conditions. 3. To identify conditions where thermal, barrier, and mechanical properties of the blend are affected by the blend composition and transesterification reactions. 4. To determine the relationship between orientation and the thermal, mechanical and barrier properties of fihns fabricated from PET/PEN blends. Chapter 2 Literature Review In this chapter, issues dealing with PET, PEN and PET/PEN copolymers and blends are reviewed. The chemical structures, manufacturing processes, selected properties, and the applications of PET and PEN are described. Finally, a review of studies dealing with PET/PEN copolymers and blends are included. 2.1. Interesting Issues of PET Poly(ethylene terephthalate) (PET) is one of the most important commercial polyesters presently used in the packaging industry. PET is also referred to poly(oxyethyleneoxyterephthaloyl) which is its IUPAC name (Odian 1991). Some trade names for PET are Mylar, Dacron, Terylene, and Cleartuf. In 1946, PET was discovered in the UK by Whinfield and Dickson. It was first known as a textile fiber and was introduced to the textile fiber market in 1953 (Adur and Bonis 1994). Shortly thereafter, PET film was commercially introduced. Since then, films based on PET were developed and became well known in many countries, e. g. the UK, USA, France and Japan (Jadhav and Kantor 1988; J aquiss et a1. 1982; Margolis 1985). In 1966, injection-molding of PET resins was introduced to the market (Anonymous 1985; Mock 1983). Recently, PET has been widely used in many industries including textile, food and household packages, pharmaceutical and medical packages, electrical and electronic devices, and automobile parts. 2.1.1. PET Structure and Properties PET is a semicrystalline, thermOplastic polyester. It is a linear condensation homopolyester, and has repeating units as shown in the following figure, 0 O {CQW Figure 2.1: Chemical structure of polyethylene terephthalate (PET) n Characteristics of PET are affected by ester groups (—coo-), distributed along the main chain of the PET molecular structure. Moreover, PET chains contain aromatic rings which contribute to high chain stiffness and a high melting temperature. The molecular structure of PET is described by the fringed-micelle model (Bonart and Hosemann 1960; Hess and Kiessig 1943). The polymer exists as an imperfect two- phase system of interconnected crystalline and amorphous domains. The regular linear structure of PET can form an Ordered structure through chain orientation and crystallization. PET has a glass transition temperature around 78°C and a melting temperature around 255°C (Jabarin 1984; Jabarin and Chandran 1993). PET is ...r suited for the biaxial film orientation, resulting in films with outstanding mechanical and barrier properties (Bilmeyer 1962). PET exhibits high mechanical strength and toughness, fatigue resistance up to 150-1 75°C, good chemical resistance, good gas barrier properties, and thermal stability (Bilmeyer 1962; Odian 1991; Werner et al. 1988). PET with in the optimal range of molecular weight average (MW) provides good processibility and low thermal shrinkage. Moreover, with this Optimal MW, it shows excellent mechanical properties (Werner et al. 1988). PET can be used on a wide variety of film-processing equipment, due to its high tensile properties. PET films stretched along one direction (tensilized films) are used as audiocassette tapes, and typewriter or computer printer ribbons. PET exhibits excellent electrical insulating properties. Therefore, it is widely used in electrical applications. Moreover, PET exhibits high chemical resistance, and good water, oxygen, and carbon dioxide barrier properties. Therefore, it is generally used as a packaging material for food, pharmaceutical, electronic, and medical products (Briston 1989; Margolis 1985; Jadhav and Kantor 1988; Werner et al. 1988). 2.1.2. PET Manufacturing There are three steps in the PET manufacturing process including: (i) PET polymerization; (ii) film fabrication; and (iii) auxiliary processing or post-treatment (Werner et al. 1988). The PET synthesis is a two-stage ester interchange process. This is a basic manufacturing process for PET which is similar for fibers, filaments, bottles, and films (Odian 1991; P0 et al. 1996). The two-stage process involves transesterification of dimethyl terephthalate with ethylene glycol, followed by polycondensation (Odian 1991; Jadhav and Kantor 1988 ; Werner et al. 1988). The first stage begins with solution polymerization of ethylene glycol (EG) and dimethyl terephthalate (DMT) or terephthalic acid (TPA). The polymerization involves an ester interchange reaction which produces bis (2-hydroxyethyl) terephthalate (BHET) along with a small amount of larger-sized oligomers (Figure 2.2a). The reaction 10 temperature increases from 150 to 2100C. Methanol, which is a by product of this stage, is distilled off. The second stage is melt polymerization (Figure 2.2b). The reaction temperature is held at 270-280°C, which is above the melting temperature of the polymer. With this high temperature and a partial vacuum of 66-133 Pa, the ethylene glycol is removed (Odian 1991; Jadhav and Kantor 1988 ; Werner et al. 1988). Temperature control and catalysts used in this process are very important for minimizing side reactions (Odian 1991; Werner et al. 1988). Common catalysts used for the first stage include acetate salts of calcium, lithium, manganese, cobalt, and zinc. The second-stage catalysts are antimony (III) oxide, germanium compounds, and titanium compounds. An alkyl .or aryl phosphite or phosphate is often added to inactivate the . first-stage catalysts. Recently, terephthalic acid in high purity became available. Terephthalic acid and ethylene glycol are also used in the first stage process to produce bis (2-hydroxyethyl) terephthalate and the two-stage process is modified for this PET production process. The two-stage process is carried out either as a batch or continuous process. The _ molten polymer is quenched to form solid extrudate. Then the extrudate is granulated to small pellets. PET pellets can be stored, and used for film or bottle production, later. PET film fabrication is carried out using a cast extrusion process. PET pellets must be carefully dried in order to avoid hydrolytic degradation during the extrusion process (McGee and Jones, 1995; P0 et al. 1996 ; Seo and Cloyd, 1991; Stewart et al. 1993). Molten polymer is east through a flat die connected at the end of the extruder, and output sheet is biaxially stretched and heat set (Schwartz 1982; Werner et al. 1988). ll o 0 II II CH3-O—C —O—CH3 + 2HOCH2CH20H EG DMH‘ {} o 0 II II . HOCHZCHz-O—C —O—CH2CH20H + 2 CH3OH BHET (a) o o H H n HOCHZCHz-O—CQ—C—O—CHZCHZOH {} BHET O o ' H H - H HZCHZ—o—C«©{}4> <-—— ll ' PETa'C‘O‘CH2CH2 ’PETb H'O'CH2CH2‘PETb 35 and (iii) Transesterification between ester groups of PET and PEN. O O II II PETa'C‘O‘CH2CH2'PETb PETa'C'O'CH2CH2'PENb fl PENa-lC-O-CHZCHz-PENb ‘—" PENa-fi-O-CHZCHz-PETb 0 These initial interchange reactions lead to the formation of block and /or graft copolymer chains. Continued interchange reactions of the initially formed block and graft copolymers result in the formation of random copolymer chains, which accounts for the enhanced miscibility of the PET/PEN blends. 3.1.2.2. Factors controlling transesteriflcation reactions In any scheme to produce materials from melt blends of PET and PEN, a number of factors have to be deal with. These include identifying the effects of processing parameters, as well as the specific type of the processing equipment, on the interchange reactions and properties of the blends. Type of extruders and the screw speed From the results of previous studies (Shi 1998; Morton-J ones 1989) and our preliminary experiments, it was concluded that processing the blends though a twin screw extruder yielded a higher degree of transesterification than that obtained from processing though a single screw extruder, using similar processing parameters, i.e. temperature and residence time. This can be attributed to the fact that the twin-screw extruder provides 36 better mixing than a single screw extruder. Better mixing optimizes contact between the PET and PEN molecules and results in higher shear generated heat, which leads to temperature increases and enhanced rates for the transesterification reactions (Shi 1998; Morton-J ones 1989). In addition, the twin-screw extruder offers better heat transfer and a more uniform temperature, which increases the transesterification reactions (Morton- J ones 1989). In this study, we focused on processing the PET/PEN blends though the twin-screw extruder in order to achieve a varied transesterification levels. The extruder screw speed is also an important factor in controlling the transesterification reactions. Increasing the screw speed causes an increase in the shear rate and mixing effect (Shi 1998). However, increasing the screw speed also results in a decrease of the blending time or residence time. Processing parameters The effects of the processing parameters, including blending temperature, blending time, catalyst and blend composition, on the transesterifcaiton reaction have been investigated by several researchers. Blending temperature and time are the primary factors controlling the transesterification reactions, whereas blend composition and catalysts have less effect (lhm et al., 1996; McGee and Jones 1995; P0 et al.1996; Shi 1998; Stewart et al., 1993). Previous researchers found a relationship between the processing temperature and time, and the transesterification degree. Stewart et al. (1993) showed that the transesterification degree at a given temperature is a linear function of blending time, corresponding to a reaction rate constant. The authors also proposed a first-order reaction 37 rate constant model. In addition, the model showed that the reaction rate constant increases by a factor of approximately 5, as the blending temperature increases from 286 to 309 °C. An activation energy of 110 kJ mol'1 was reported. Several researchers also studied the effect of catalysts on the transesterification degree of PET/PEN blends. Stewart et al. (1993) found that adding 0-40 ppm of antimony-based catalyst to the PET/PEN blends had little effect on the transesterification reactions. Po et al. (1996) also reported similar results, in that the transesterification degree varied fiom 0.14-0.2, when 160 and 360 ppm. of Sb, 100 and 160 ppm. of Ge , 2 and 20 ppm. of Ti, and 160 and 320 ppm. of Sn were used as the catalysts. The results of this study showed only a slight effect of the type of catalyst and the quantity of catalyst on the transesterification reactions, as compared to the effect of processing time and temperature. Po et al. (1996) stated that significant effects of the catalysts might be found if substantially higher amounts of catalysts were added. In PET/PC blend systems, the catalysts were found to be a primary factor in controlling the interchange reaction. Joyce and Berzinis (1991) stated that residual . polymerization catalysts did not significantly affect the transesterification reactions of PET/PC blends, while the addition of titanium catalysts resulted in a significant increase in the transesterification degree. The transesterification rate of polyester/polycarbonate blends without titanium was found to be limited (Devaux et al.(°’°’°’°) 1982; Godard et al. 1986; Pilati et al. 1985; Smith et al. 1981). In contrast, the rapid transesterification of other polyester/polyester blends, without any catalyst, has also been reported (Kugler et al. 1987; MacDonald et al. 1991; Montaudo et al. 1992; Rarnjit and Sedgwick 1976). Stewart et al. (1993) proposed that PET/PEN blends and other polyester/polyester blends 38 had hydroxyl and carboxyl end-groups that differentiated them from the polyester/PC blends. Even though they involved similar reaction processes (alcoholysis, acidolysis, and transesterification), the controlling mechanisms for polyester/polyester blends and polyester/PC blends in the absence of titanium are different. The effect of the blend composition on the transesterification reactions was also studied. A slight effect of the blend composition on the transesterification degree was reported (Ihm et al. 1996; P0 et al. 1996; Shi 1998; Stewart et al. 1993). Ihm et a1. (1996) concluded that there was no preference between PET and PEN, when transesterification reactions occurred. However, Shi (1998) found a slight effect of blend composition on the transesterification degree. The blends with lower and higher blend compositions showed a higher transesterification degree, while the blends with intermediate levels of blend compositions (30-40% naphthalene dicarboxylate content (NDC), which is a percent mole fraction of naphthalate units based on the total moles of terephthalate and naphthalate units, had the lowest transesterification degree. This behavior was explained by the reaction kinetics theory, which showed that the activation energy should be constant for various compositions and the change of entropy should be a minimum at the equivalent blend composition of 33% NDC or 50 mole % PEN (Shi 1998). In contrast, for the PET/PC blend system, the blend composition strongly influenced the transesterification reactions observed in the PC-rich phase (Joyce and Berzinis 1991). 39 Types of PET and PEN resins Shi (1998) indicated that types of PET and PEN influenced the transesterification reactions. PET and PEN polymers made by different processing methods and parameters with different catalysts were found to differ in their molecular weight distribution and viscosity. The PET/PEN blends with higher viscosity had a lower transesterification degree, due to less mixing between reactant components. No report of the effect of molecular weight average and distribution of the PET and PEN on the transesterification reactions was shown. Sources of naphthalate. McGee and Jones (1995) proposed that the sources of the naphthalate, including PEN homopolymer and PET/PEN copolymer, influenced the time required to achieve miscible blends. Using the PET/PEN copolymer as the naphthalate source resulted in a shorter processing time to obtain clear blends. This was attributed to the pre-existence of interlinkage of PET and PEN chains from the PET/PEN copolymer. In the present study, the focus of the processing parameters is on the blending temperature and time. In addition, the composition of the blends was also studied. Even though the blend composition has little or no effect on the transesterification reactions, it is a major parameter controlling selected properties of the blends. 40 3.2. Materials and Methods 3.2.1. Materials Two homopolymers used in this study were Poly(ethylene terephthalate) (PET) and Poly(ethylene 2,6-naphthalene dicarboxylate) (PEN). PET (CTF 8406) and PEN (VRF 40046) were provided by the Shell Company (Apple Grove, WV). 3.2.2. Reaction Processing Parameters and Blend Composition For the preliminary experiment, a 25 weight % PEN blend was processed through a single-screw extruder at 303-307°C for 1 pass. The DSC and NMR analysis showed that the blend was immiscible and no transesterification reaction occurred. Blends had 2 T8 and Tm values. The 1H-NMR spectra showed a peak for PET occurring at approximately 4.8 ppm, a peak for PEN occurring at approximately 4.9 ppm, and no interaction peak occurring between the PET and PEN peaks. The blend exhibited phase separation and it appeared to be hazy or white in color (see Results Sections 3.3.1 and 3.3.2). ' Therefore, for the further experiments, all the samples were processed through a twin-screw extruder in order to increase the degree of transesterification and achieve miscible blends. The effects of blending temperature, time, and blend composition on the transesterification reactions were investigated. Blending temperatures ranging fi'om 275 to 325 °C were studied. The blending time was varied by the number of passes through a twin-screw extruder (one and two passes) and extruder screw speeds of 150 and 250 rpm. The residence times corresponding to the screw speed of 150 rpm were approximately 75 and 150 seconds for one and two passes through the extruder, respectively. The residence 41 times corresponding to the screw speed of 250 rpm were not determined. Blend compositions were varied from 10 to 40 weight percent PEN or 8.1 to 34.6 mole percent of Naphthalene Dicarboxylate (% NDC). The mole percent of naphthalene dicarboxylate content (% NDC) is the mole fraction of naphthalate units based on the total moles of terephthalate and naphthalate units expressed as percent % NDC was calculated by using the molecular weight of the naphthalate unit (242 g/mol) and terephthalate unit (192 g/mol). Since functional groups (hydroxyl and carboxyl groups) at the end of the polymer chain are the active sites of the interchange reactions, the PEN composition was also expressed as the mole percent of the end groups of the PEN in the blend (% e). The mole percent of the end groups of PEN was calculated as follows; %e = 1%WtOfPEN/MWEEE*2) (% Wt OfPEN / MWPEN * 2) + (°/o Wt OfPET / MWPET * 2) Mole percent of PEN (% PEN) was also calculated as follows; % PEN = (% Wt of PEN / Mmel (°/o wt of PEN / MWpEN) + (% Wt of PET / MWpET) Therefore, the mole percent of the end groups of PEN (% e) is equal to the mole % PEN in the blends. The weight average molecule weight of PET (MWpET) was 65,400 gram /mole, and the weight average molecule weight of PEN (MW war) was 42,900 gram /mole. The weight average molecular weights of PET, PEN, and blends were measured by gel permeation chromatography (see Chapter 4). The combinations of processing parameters and blend composition used in this study are listed in Table 3.1. These processing parameters were selected by considering the results of the previous studies as well as equipment available. 42 Table 3.1: Processing parameters and blend composition Blending Blending time Screw speed Blend composition temperature (°C) No. of passes RPM % (wt/wt) PEN 275 1 150 25 275 1 250 25 285 1 150 25 300 1 150 25 315 1 150 25 325 1 150 25 285 2 150 25 300 2 150 25 315 2 150 25 300 l 150 10 300 1 150 30 . 300 1 150 40 For the preliminary experiment, PET/PEN blend with 25% PEN (wt/wt) were processed though a single screw extruder at 303-3070C, and 1 pass. All of the experiments were performed on a twin-screw extruder. 43 3.2.3. Extrusion 3.2.3.1. Drying conditions PET and PEN resins were weighed, mixed, and dried. In order to avoid hydrolytic degradation during the extrusion process, the moisture content of the resins needs to be very low (McGee and Jones 1995; Seo and Cloyd 1991; Stewart et al. 1993). Therefore, the resins were dried at 82°C for at least 20 hours in a DBD25 vacuum dryer (Premier Pneumatics Inc., Salina, Kansas). In addition, the extrudate was redried under the same conditions before the second pass. The moisture contents of the dry resins were measured by the 720 KFS titrino Karl Fischer titration (Brinkrnann Instruments, Inc., Westbury, New York) to confirm that the samples had low moisture content (lower than 100 ppm). 3.2.3.2. Twin-screw extruder Dried resins were automatically fed through the extruder at 10 lbs per hour using an MDII weight feeder (Acrison Inc.). A Baker Perkins Model ZSK-30 co-rotating twin- screw extruder (Werner & Pfleilderer Corporation, Ramsey, New Jersey) was used to carry out the reaction processing. The dimensions of the extruder were: Outer screw diameter 30.7 mm Inner screw diameter 21.3 mm Thread depth 4.7 mm L/D ratio 26:1 44 The extruder consists of three basic heating zones: (i) the feeding zone below the feed hopper; (ii) the compression zone where polymers start to melt; and (iii) the metering zone connected to the die. There are a total of six ports along the three heating zones of the extruder. The temperatures of all ports were set equally. The melt blends left the extruder through a die head with 2 cylindrically shaped outlets. The extrudate was then immediately cooled in a water bath at approximately 50°C. After that, it was pelletized using a Conair pelletizer (J etro Division Inc., Bay City, Michigan). The blend pellets were optically clear, rod shaped, and approximately 1 mm in diameter and 4 mm long. For the preliminary experiment, the single screw extruder was used for initiating transesterification reactions. Dry PET and PEN resins (75: 35 wt/wt) were fed though a Killion single-screw extruder. Processing temperatures were 303-307°C, with 1 pass though the extruder. The residence time was 4.5 nrinutes. 3.2.4. DSC Analysis The thermal properties of the blend resins were determined by using lyfl3SC 2920 Modulated Differential Scanning Calorimetry (TA instrument Inc., New Castle, Delaware). All samples were precisely weighed with a Sartorious analytical balance (Sartorius GMBH Gottngen (Germany), Westbury, New York). Optimum weights are normally in the range of 5-10 milligrams. Whole pellets of PET/PEN blend resins were used, if their weights were in the optimum range. The blend fihns were cut and stacked to achieve the optimum weight. Use of low weight thin samples results in minimal 45 thermal gradients, which facilitates attaining steady state condition and gives accurate data. Each sample was placed in an aluminum bottom dish and a top dish was put on. The dishes were sealed by using the clamp. An empty dish was used for the compensation of the heat flow corresponding to the dishes. The polymer blends were equilibrated at 10°C for 5 minutes before being heated to 290°C (320°C for PEN) with the heating rate of 4°C per minutes. The modulated mode with an imposed modulation rate of +/- 1°C every 60 seconds was used. The modulated DSC has a superimposed sinusoidal temperature oscillation, while the conventional DSC has a linear change in temperature. Therefore, modulated DSC can separate the total heat flow into reversing and nonreversing components, which is permits accurate and easy interpretation of the results. The glass transition temperature (T8) was determined from the transition temperature of the reversing heat flow profile. Melting temperature (Tm), and heat of fusion (AHf) were determined from the endothermic event of the total heat flow profiles. The percent crystallinity of the blends was determined fi'om the heats of fusion of 100% crystalline blends, which were calculated from weight average (fi'om the mole fraction of the blend composition) of the heats of fusion of 100% crystalline PET and PEN homopolymers. The heats of fusion are 121.42 J oules/gram (29.0 calories /gram) and 103.41 Joules/gram (24.7 calories/ gram) for PET and PEN, respectively (Cheng et a1. 1988; Roberts 1995). The percent crystallinity was calculated as follows: 46 The heat of fusion (AHf) of 100 % crystalline blend with X mole % NDC = (X/100 "' 103.41) + (100-X)/100 * 121.42 Joules/gram Thus; AHf of 100 % crystalline blend with 8.1 mole % NDC = 119.62 Joules/gram AHf of 100 % crystalline blend with 21 mole % NDC = 116.92 Joules/gram AH; of 100 % crystalline blend with 25.4 mole % NDC = 116.02 Joules/gram AHf of 100 % crystalline blend with 34.6 mole % NDC = 114.22 Joules/gram % crystallinity of 8.1 mole % blend = AHfGoules/gram) *V 100 / 119.62 3.2.5. ‘H-NMR Analysis The transesterification degree of the blends was measured using proton nuclear magnetic resonance spectroscopy (lH-NMR INOVA 300, Varian VNMR). Samples were prepared by dissolving 0.25 g of the blend in 5 ml of 70/30 (by weight) mixture of deuterated chloroform and trifluoroacetic acid. 1H-NMR spectra were accumulated and transformed using the standard operating software. A typical NMR spectrum for a PET/PEN blend is shown in Figure 3.1. The degree of transesterification of the blends was determined by spectra analysis using the region of the 1I-INMR spectrum between 4.8-4.9 part per million (PPM). PET and PEN peaks appear at approximately 4.8 and 4.9 PPM, respectively. In PET/PEN blends with transesterification, there are heterolinkages between the naphthalate and terephthalate units. The 1H peak corresponding to these heterolinkages appears at 4.85 PPM, between those of the PET and PEN homopolymers (Figure 3.1). The curve of the NMR spectrum was fitted by 3 normal curves using the least squares method. The intensities of the PET, PEN, and interaction peaks were 47 PET/PENpeak . l l l PENpeak\ \ I i. actualspectrum fullfit individual olot , A ‘IT'P'1""71 r' IT I V1" 5.2 5.1 5.0 4.9 4 a 4 7 as. Figure 3.1: NMR spectrum of the PET/PEN blend with 25% (wt/wt) PEN processed through the twin-screw extruder at 300°C and 1 pass. 48 determined by the area under the corresponding curves. The degree of transesterification was then calculated from these intensities using the Bemoullian statistics technique (Y arnadera and Murano 1967, and Po et al. 1991), which is described in the next section. 3.2.6. Calculation of Degree of Transesterification and D value Degree of transesterification (B) was originally described and calculated by Yamadera and Murano (1967). By assuming that the lI-INMR spectrum is a measure of the relative intensities of homolinkages and heterolinkages, the average sequence length of polymer chains and the degree of transesterification can be determined. Molar fractions of terephthalate (MT) and naphthalate (MN) are obtained from the intensities of the NMR spectrum: MT = AT-G-N /2 + Ar-G-r (1) MN = AT-G-N /2 + AN-G-N (2) where Au}; , AMLN, and A1204»; are the ratios of the intensities of the PET peak, the PEN peak, and the interaction peak of PET and PEN, respectively, and the total intensities of these three peaks. Assume that the polymer chains of PET/PEN blends are extremely long. Therefore, the number of end units is negligible compared to the number of all monomer units. Suppose that a monomer unit is randomly picked from the PET/PENiblend chains. If the selected unit is a naphthalate (N) unit, then Pm is defined as the probability that the next unit proceeding down the chain will be a terephthalate (T) unit. Pm is defined similarly, except that the roles of N and T are switched. Pm and Pm can be calculated fi'om the intensities of the NMR peaks as follows: 49 PM = ATG-N/ZMN (3) Pm = ATGN/ZMT (4) The degree of transesterification (B) is defined as: B = PNT + PTN (5) Yamadera and Murano (1967) also indicated that the T and N units obtained a random distribution when B = 1 and the probability of finding a copolyrrrer unit is based on Bernoulli statistics. Ifthese units tend to form in blocks of each unit, B is below 1. In addition, if a homopolymer mixture is found and no transesterification reaction is achieved, B is zero. On the other hand, if the homopolymer sequence length becomes shorter than would be expected by random reaction, B is greater than 1. When an alternating capolymer is formed, B is equal to two. However, a true alternating copolymer can be formed only if PET and PEN are present in equimolar amounts. For blends of PET and PEN homopolymers, percent randomness and transesterification are equal and equivalent to the transesterification degree (B) times 100. Therefore, a polymer blend without interaction or transesterification has 0 % randomness (transesterification) and B = 0, while a random copolymer has 100% randomness and B = 1. The average sequence length of T units (La) and N units (LnN) are calculated by 1411' = ZMT/ AT-G-N = l/PTN ‘ (6) Iain = 2MN / Ar-o-N = NEW (7) The fraction of the T-G-T, N-G-N and T-G-N linkages can be obtained from the intensities of the PET, PEN and PET/PEN peaks of the ‘HNMR spectrum. The 50 parameters MT, MN, Pm, Pm, LnT, LnN, and B can then be calculated from equations (1) - (7)- One should note that the B value is calculated by simply adding Pm to Pm without weighing the molar fraction of PET and PEN. Therefore, the B value is most meaningful only for a blend with equimolar amounts of PET and PEN. Therefore, the D value is proposed. Suppose that a monomer unit is randomly picked from the PET/PEN blend chains. D is defined as the probability that the next unit is different from the selected unit; i.e. the selected unit is N and the next unit is T or vice versa. D can be calculated as follows: D = I’m MN + PTN Mr (8) By substituting (3) and (4) into (8), it becomes D = (Ar-c.N/2MN)(MN) + (AT-G-N/ZMTXMT) = AT—G-N (9) The D value ranges from 0 to 1. When no transesterification reaction occurs, the D value is equal to 0. When an alternating copolymer is obtained, the D value is equal to l. Ifa random copolymer is formed, the D value is 0.5, regardless of the composition ratio. Let Dm (M) be the maximum achievable D where M is the smaller mole fraction of these two polymers. Dmax (M) can be calculated as follow: Dmax (M) = 2M (10) For instance, if the PET and PEN compositions are equal (50% mole), Dm (M) is l,and it is possible that an alternating copolymer is formed. If the PET and PEN compositions are not equal, Dm (M) is less than 1, and the alternating copolymer cannot be formed. 51 In the literature, the B value is widely used to represent the degree of transesterification. Therefore, in this study, the B value is used as well. However, the more meaningful D value is also presented. 3.3. Results and Discussion 3.3.1. DSC Analysis The blends processed though the single screw extruder (300°C and 1 pass) had two T8 values, which appeared at 73 and 119 °C, respectively (Figure 3.2). These T8 values corresponded to the PET and PEN rich phases, respectively. The melting peaks appeared as split peaks, which included the major peak at 251°C and a relatively small peak at 270°C (Figure 3.2). These melting peaks corresponded to the PET and PEN rich phases, respectively, and they seemed to merge during the DSC scan, as shown by the split peaks. Thus, the blend processed though the single screw extruder at 300°C exhibited immiscibility or phase separation. In addition, 1HNMR analysis also confirmed that no measurable amount of transesterification reactions was obtained. The DSC analysis showed that the blends processed though the twin-screw extruder had only one T8 and Tm value, which were between those of PET and PEN homopolymers. Therefore, miscible blends were obtained. A typical DSC scan of a miscible blend is shown in Figure 3.3. A summary of T8, T m and % crystallinity values for the respective blends is presented in Chapter 4, Section 4.3. 1. Therefore, for all further experiments, the twin-screw extruder was selected for processing the blends, so as to achieve varied levels of transesterification. 52 smlo: 25:75 9574:9131 cowouuo DSC FIIoz CtPEN-PET.500 Size: 93000 mg Operatorz.MJR . Method: moo 4.0 CMN TO 175 c Run Dm- 274111397 15.25 Commonlzuloouurso osc raw 303-307 c. RESIDENCE TIME 4.51m 0.00 -0.05J 73.07 cm , 69.27 °C 0 gr 77.ss°c "8'82 c (1) 221.61 °c g -0.10« “Hate seem/g 3 117.10’c ‘— m > 015- 0 o: -0.20-1 251.01 °c -025 ' T T l ' r T f V r 0 50 100 150 200 250 300 Temerature(°C) Figure 3.2: The DSC scan of the PET/PEN blend with 25% (wt/wt) PEN processed through the single-screw extruder at 300°C and 1 pass. 53 Sample: blend#3- 30001 onented film We C:...1DSC\ru;idal300010I 8120: 5.7000 mg DSC Operator. Rullda Method: IUflda memod1 Run Date; 1680099 15 38 Comment 81000113 (30001 )-onented film 1 mil - 02 a g 3 3 E E i i :l: I 3 0: --0 3 .02 i r I f l T I ' r T I ' '0 4 0 50 100 150 200 250 300 Em Up Tm (°C) UI'IW v3.0a TAInsmm Figure 3.3: The DSC scan of the PET/PEN blend with 25% (wt/wt) PEN processed through the twin-screw extruder at 300°C and 1 pass. 54 3.3.2. Degree of Transesterification (B) The results showed that no measurable transesterification reactions occurred during processing the blends though the single-screw extruder at 300 0C. The lH-NMR spectra showed a peak for PET occurring at approximately 4.8 ppm, a peak for PEN occurring at approximately 4.9 ppm, and no interaction peak occurring between the PET and PEN peaks. However, the blends processed through the twin-screw extruder at the blending temperatures of 275 to 315°C and l and 2 passes resulted in varied degrees of transesterification. The interaction peak occurred at 4.85 ppm (between the PET and PEN peaks). The degree of transesterification ranged from 0.004 to 0.53. Blend resins and blend films (non-oriented films) obtained at the same processing times and temperatures had similar ranges of degree of transesterification. Statistical analysis using single factor AN OVA (analysis of variance) showed no significant differences between the degree of transesterification of blend resins and of blend films at the confidence level of 0.05, except for 40 % of PEN by weight (Table A1 in appendix A). Therefore, it was concluded that no significant transesterification reaction occurred during film production. In addition, the effects of the processing factors and the blend composition on the degree of transesterification were studied. As the PEN composition increased, the degree of transesterification appeared to be constant. Statistical analysis was performed, and the result showed no significant difference in the degree of transesterification among blend samples with 10 to 40% PEN by weight (Table A3 in appendix A). The processing conditions of blending time and temperature appeared to be the primary factors controlling the transesterification reactions. The results are discussed in detail in the following sections. 55 3.3.2.1. Effect of blending time on transesterification reactions The blending time was varied by the extruder screw speed and the number of passes through the extruder. Screw speed The extruder screw speed is a factor controlling the blending time, as a faster screw speed yields a shorter residence time. The results showed that as the screw speed of the extruder increased, the degree of transesterification decreased (Table 3.2). The transesterification degree of the blends obtained at a 150 rpm screw speed (one pass at 275°C) was 0.004 while that of blends obtained at a 250 rpm screw speed (one pass at 27 5°C) was not measurable by NMR analysis. This is due to the limitation of lI-INMR spectroscopy. In addition, blends processed at 250 rpm and one pass at 275°C were white or opaque, which indicated the irnnriscibility of the blends. However, DSC analysis of these blends (250 rpm) showed only one T8 and one Tm value, which indicated that these blends achieved some minimal degree of transesterification. Shi (1998) found the same phenomena and concluded that the transesterification reactions lead to the creation of partial copolymers in the blend structure. This copolymer behaves like a compatibilizer, and blends appear to be more compatible, compared to a physical mixture or 0% transesterification. Stewart (1993) also showed that firrther reaction occurs during the DSC analysis. Thus, the degree of transesterification may increase during performing the DSC test, resulting in an; increase in miscibility, which leads to show only one T g and Tm value. 56 From the results obtained in the present study, it can be concluded that when the blends were processed at a screw speed of 250 rpm, the degree of transesterification achieved was minimal (see Table 3.2). Therefore, a screw speed of 150 rpm was selected as the base processing condition for the study of other processing variables. Table 3.2: The effect of the extruder screw speed on the degree of transesterification (B). The blends with 25% of PEN (wt/wt) were processed at 275°C for 1 pass. Screw speed The degree of transesterification (B) 150 rpm 0.004 250 rpm less than 0.002" "' The transesterification degree cannot be detected by NMR analysis. Numbers of passes The average degree of transesterification for the resin blends achieved from one pass and two passes at melting temperatures of 285 to 315°C and at a screw speed of 150 rpm, ranged from 0.053 to 0.513. From NMR results, as shown in Table 3.3 and Figure 3.4, the degree of transesterification increased as the number of extrusion passes increased. This confirmed the results reported by Stewart et al. (1993). The authors processed PET/PEN blends through a single-screw extruder at 295 to 315°C for 1 to 3 passes and reported a linear relationship between the number of passes and the transesterification level. Our results showed higher degrees of transesteri fication were achieved than those reported by Stewart et al. (1993), who found that degrees of transesterification ranging 57 from 0.08 to 0.14 and 0.18 to 0.21 were achieved by extruding PET/PEN blends through a single screw extruder at 305°C for one and two passes, respectively, as compared to level of transesterification between being achieved by processing though the twin screw extruder at 300°C for one and two passes. The differences in degree of transesterification observed are attributed to the fact that the twin-screw extruder used in this study gave better mixing than a single-screw extruder. 3.3.2.2. Effect of blending temperature on transesterification reactions Blends of 25% PEN (wt/wt) processed at 275, 285, 300, 315 and 325°C (150 rpm, and one pass and two passes) were investigated. The results of NMR analysis of the respective samples are shown in Table 3.3 and Figure 3.4. These results clearly demonstrate that a significant increase in transesterification was achieved at the higher processing temperatures. The relationship between the blending temperatures and the degree of transesterification appeared to follow a linear expression. From linear regression, the equations are Y = 0.0087X -2.4058 for 1 pass, and Y = 0.0117X —3.1395 for 2 passes through the extruder, where X is the temperature (°C) and Y is the degree of transesterification (Figure 3.4). The R2 values are 0.9933 and 0.9812, respectively. This conclusion is in agreement with the results of McGee and Jones (1995), and Stewart et a1. (1993). Stewart et al. (1993) also demonstrated the kinetic model of transesterification as a first-order reaction rate-constant. The PET/PEN blend resins obtained from reaction processing at a temperature of 275°C did not appear to be uniform and were mixed between clear and hazy zones. This indicated that a blending temperature of 275°C is the minimum limit to achieve a 58 Table 3.3 : The effect of blending temperature and time on the degree of transesterification (B) —Temp B of blend resins* B of blend films" (0 C) lst pass 2nd pass lst pass 2 nd pass 275 0.053 - - - 285 0.061 0.151 0.076 0.192 300 0.219 0.383 0.222 0.405 315 0.309 0.513 0.356 0.525 325 0.429 - - - "‘ Blends with 25 %PEN (wt/wt) processed through twinescrew extruder( 150 rpm) 0.7 5 2 asses , a 0.6 it p /. 3 y = 0.0117x - 3.1395 E 05 ._ R2 = 0.9812 ‘ g 0.4 7* g//// ,/i g E / g/ '45 0.3 a? ’1/ /,/ / / E 0.2 -- y’ //!/ 1 pass 8 / " ' o/ y = 0.0087x - 2.4058 2 0-1 t /’. R2=0.9933 1- §(// 0 i *r i— : 1 270 280 290 300 31 0 320 330 Temperature ( °C) O l pass-resins I lpass-films o 2 passes-resins D 2 passes-films i Figure 3.4 :The effect of blending temperature and time on the degree of transesterification (B) 59 measurable degree of transesterification. In addition, the PET/PEN resin blend obtained from reaction processing at 325 °C was brittle and difficult to pelletize because of low melt viscosity. Degradation might occur during processing blends at this temperature or at temperatures greater than 325°C. Therefore, the optimum blending temperature is considered to be in the range fi'om 285 to 315°C 3.3.2.3. Effect of blend composition on transesterification reactions The degree of transesterification achieved among PET/PEN blends with compositions ranging fiom 10% to 40% PEN (wt/wt) (14.5 to 50.4 mole % PEN) showed no significant difference (Table 3.4 and Figure 3.5). These findings were in agreement with the results of previous studies. Stewart et al. (1993), McGee and Jones (1995), and lhm et al. (1996) reported that the blend composition had little or no influence on the degree of transesterification. However, in studying blends with PEN compositions ranging from 10 to 80 weight percent, Shi (1998) found that the blends with low and high PEN levels had a slightly higher degree of transesterification than those with medium PEN levels (approximately 32 mole % NDC or 50 mole % PEN). 60 Table 3.4 : The effect of PEN composition on the degree of transesterification (B) avg B avg B Sample ID PEN mole% resins’“ films" Bavg 8 14.5 0.189 0.204 0.197— 3 33.7 0.219 0.222 0.221 9 39.5 0.181 0.208 0.195 10 50.4 0.13 0.192 0.161 * Blends processed at 300°C and 1 pass through twin-screw extruder. .0 NP 0100 I P N I The transesterification degree (8) \ 0.15 4- E 0.1 ._ 0.05 4~ 0 1 i i i t - 0.0 10.0 20.0 30.0 40.0 50.0 PEN mole % o resins I films + Bavg Figure 3.5 : The effect of PEN composition on the degree of transesterification (B) 61 3.3.2. D value The D value of blends ranged from 0.01 to 0.181. Statistically, there was no significant difference between the D values of blend resins and blend films obtained at the same processing conditions (Table A2 in Appendix A). The relationship between the D value and the blending temperature and time were found to be similar to that between the B value and the blending temperature and time (Table 3.5 and Figure 3.6). However, the PEN composition slightly influenced the D value when the PEN mole % was less than 25. Where the PEN mole % was higher than 35, the D value appeared to be constant(Table 3.6 and Figure 3.7). Statistical analysis showed no significant difference between the D values of blends with 35 to 50 mole % PEN. However, the D value of the blends with 14.5 mole % PEN was less than the D values of the other blend samples (Tables A4 and A5 in appendix A). 62 Table 3.5: The effect of blending temperature and time on the degree of transesterification and D value. Temp D average B average (0 C) D-l st pass D-2nd pass B- 1 st pass B-2 nd pass 275 0.0164 0.053 285 0.022 0.054 0.069 0.1715 300 0.074 0.130 0.221 0.394 315 0.113 0.184 0.333 0.519 325 0.148 0.450 "' Blends with 25 %PEN (wt/wt) processed through twin-screw extruder( 150 rpm) 0.7 0.7 0.6 ~r «e 0.6 a0 5 o 5 ‘6§ ' ' § 80.4 -~ -» 0.4 8% '3 %03 «r -~ 0.3 B "' 50.2 4- -~ 0.2 E 0-1 1- -_ 0.1 0 1 l l # fit 0 270 280 290 300 310 320 330 Temperature ( °C) 1 A B-lst pass A B-2 nd pass 0 D-lst pass 0 D-an pass 1 Figure 3.6 :The effect of blending temperature and time on the degree of transesterification and D value 63 The D value Table 3.6 : The effect of PEN composition on the degree of transesterification and D value Sample ID PEN mole% Davg Bavg 8 14.5 0.023 0197—“ 3 33.7 0.073 0.221 9 39.5 0.072 0.195 10 50.4 0.072 0.161 * Blends processed at 300°C and 1 pass through twin-screw extruder. 0.250 0.250 /_,,,a 0.200 .. 1/ ; «~ 0.200 E \\ u- C 0 30150 a. + 0.150 t 8 8% 1’301004» +0100 O a ' fi 3 *_§___{ C fl // , 5 0.050 .. / ~~ 0.050 e/' 0.000 t . t 4. 0.000 0.0 10.0 20.0 30.0 40.0 60.0 PEN mole °/e +Bavg '+Davg J Figure 3.7 : The effect of PEN composition on the degree of transesterification and D value The D value Chapter 4 Characterization of PET/PEN Blends 4.1. Introduction Chemisivity, molecular weight and molecular weight distribution are crucial parameters controlling other blend properties, such as thermal properties, density, morphology, gas banier properties and mechanical properties. In this chapter, we will focus on the relationships between the blend composition and the degree of transesterification (or the processing condition), and the important properties of the PET/PEN blends including thermal properties, density, molecular weight and molecular weight distribution, and morphology. In Chapter 5, the relationships between the blend composition and the degree of transesterification and the blend performance, including gas barrier and mechanical properties, will be evaluated. By combining this information, a better understanding of the basic characteristics of PET/PEN polyblends and blend performance will be provided, and predictions of barrier and mechanical properties of the resultant blend filrns can be made. 4.1.1. Thermal Properties The physical state of a polymer determines the characteristics of that polymer. The physical state in which a polymer can exist can be ideally considered as its regular polymer chains. There are three possible states: 65 (i) completely free rotation, as in a polymer at the melting state, (ii) no rotation, as in a polymer in the glassy state, at very low temperature, where its chains are frozen, and (iii) packing, as a crystalline polymer (Rodriguez 1996). Two major critical temperatures that divide the physical states of a polymer are the glass transition temperature (T8) and the melting temperature (Tm). The crystallization is also described as how polymer molecule can fit together in such a way so as to stabilize the chain arrangement in a regular lattice. A polymer at a temperature greater than its T8 is in the rubber state in which polymer chains have segmental mobility. A polymer at a temperature below its T8 is in the glassy or rigid state, in which polymer chains cannot rotate and are frozen in a specific conformation. Polymers exhibit fi'ee rotation in the melt state or at temperatures greater than its Tm. Morphology indicates whether a polymer exhibits both thermal transitions or only one thermal transition (Odian 1991). Completely amorphous polymers exhibit only a glass transition temperature, whereas semicrystalline polymers have both glass transition and melting temperatures. Crystallizable blends were selected for this study. Therefore, these blends exhibit both a glass transition temperature (T3) and a melting temperature (T m)- The physical state of a polymer is a significant parameter controlling the performance of that polymer. For instance, a polymer with a high T8 is a very rigid and strong material, and has very high gas barrier properties, due to its non (or partial) rotating polymer chains, which does not allow the gas molecules to penetrate the inner structure. The melting temperature is also an important factor controlling the processibility and flow properties of polymers, and indicates the maximum temperature 66 resistance of those polymers. There are the relationships between degree of crystallinity, and barrier and mechanical properties of polymers, as well. The crystalline matrix of a crystallized polymer has a very oriented microstructure, which results in high thermal properties and very strong materials (Odian 1991). F urtherrnore, the glass transition temperature and the melting temperature of polymer blends have been used for measuring the numbers of the blend components. PET/PEN blends without any transesterification reaction are inherently immiscible. Irnrrriscible blends exhibit both PET and PEN properties, which are not uniform, and phase separation of PET and PEN is found. The blends exhibit two T8 and two Tm values, corresponding to PET and PEN rich phases. When transesterification reactions are obtained, interlinkages between terephthalate and naphthalate units (T-N bonds) occur. This causes changes in the microstructure of the blends which, in turn, results in changes in the physical state of the blends. The transesterification reactions lead to miscible blends, which exhibit minimal sizes of the PEN domains, resulting in uniform properties of the blends. Miscible blends exhibit only one T8 and Tm, with the T8 and Tm values falling between those of PET and PEN homopolymers (Olabisi et al. 1979; MacKnight et al. 197 8; Karasz 1985; Utracki 1990). The observation of the thermal properties of the blends was done by differential scanning calorimetry (DSC). The effect of the transesterification reactions and blend composition on T1; ,Tm, and % crystallinity were studied. 67 4.1.2. Molecular Weight and Molecular Weight Distribution The molecular weight and the molecular weight distribution are measures of the average size and the distribution of the sizes of the polymer chains/molecules, respectively. The molecular weight and the molecular weight distribution are important characteristics which are responsible for several basic properties of the polymer. For thermoplastic materials, the molecular weight and molecular weight distribution strongly control the processing behavior at elevated temperatures and the mechanical properties of the therrnoplastics (Rudin 1982). In industry, plastic grades vary in the average sizes and distribution of the average sizes of the polymer chains. Polymers with high molecular weight averages exhibit higher resistance to deformation at elevated temperatures. However, this type of polymer needs an intensively higher temperature for processing or forming desired shapes (Rudin 1982). From randomness theory, Paul and Bucknall (1999) concluded that when transesterification reactions occur, the rrricrostructures of the polymer blends change. For example, for PET/PEN blends, the average sequence lengths of terephthalate and naphthalate units decrease, whereas the average length of the interlinkage T-N bond increases. The average sequence lengths can be calculated from lHNMR analysis using Yamadera and Murano’s equations (Y amadera and Murano 1967). The relationship between the degree of transesterification and the average sequence lengths of terephthalate and naphthalate units can also be determined fiom Yamadera and Murano’s equations, as shown in Section 4.2.3.2. 68 As the number of interchange reactions of the terephthalate and naphthalate units increases. the molecular weight and the molecular weight distribution of the blends will change. In the present study, determination of the molecular weight and molecular weight distribution of the blends with selected degrees of transesterification was performed by using gel permeation chromatography. The information obtained fi'om molecular weight analysis can be used to evaluate the ultimate properties of the blends, such as their melting temperatures (Brennan 1978). 4.1.3. Density The density is a typical characteristic that determines the ultimate properties of polymeric materials. For example, polymers of the same generic type with higher density tend to exhibit higher performance, such as higher gas barrier and mechanical properties. The change of the microstructure of the blends due to transesterification reactions may cause the density of the blends to change. The density of the blends was measured using the density gradient technique. The effect of transesterification reactions and blend composition on the density was studied. 4.1.4. Morphology Morphology is an important characteristic of polymers. It controls other characteristics, such as the thermal properties of the blends. Most blends, such as PET/PEN blends, consist of two or more phases. Immiscible blends exhibit two phases, a PET and PEN rich phase. Typically, PET has a lower intrinsic viscosity and exists as a continuous phase. PEN has a higher intrinsic viscosity and, therefore, exists as a 69 dispersed phase in the PET continuous phase (McGee and Jones 1995). In addition, PET/PEN immiscible blends exhibit poor optical properties, because the sizes of the PEN domains are large, and the domains scatter light. When transesterification reactions are obtained, interlinkages between terephthalate and naphthalate units (T—N bonds) occur, and the lengths of PET and PEN chains become shorter. The interchange reactions between PET and PEN chains enhance the miscibility of the blends (McGee and Jones 1995; Stewart et al. 1993). Partially miscible blends consist of PET, PEN, and terephthalate-naphthalate interlinkage phases. The blends are optically clear, due to the small size of the PEN domains, which are too small to scatter light (McGee and Jones 1995). Continued interchange reactions of the initially formed block and graft copolymers lead to the formation of random copolymer chains. Random copolymers exhibit only one phase, the T-N interlinkage phase. Therefore, changes in the morphology of the blends result in changes in the characteristics of the blends. One of the newest techniques to study the morphology of polymers is Raman chemical microscopy. Raman chemical microscope is a new type of optical microscope which is a laser confocal scanning microscope (LCSM) coupled with a filter or spectrometer (Sawyer and Grubb 1996). The Raman analysis is a powerful technique which can reveal the chemical architecture of polymeric materials without using stains, dyes, or contrast agents which might cause the interference or artifact to the analytical species. Using this technique, microstructures of immiscible blends without transesterification reaction and miscible blends with some degree of transesterification were investigated. 70 4.2. Materials and Methods 4.2.1. Film Fabrication Process Film fabrication was done by the Amoco Research Center, Amoco Chemicals Company, Illinois. The PET, PEN, and PET/PEN blends (11 samples) obtained from reaction processing were dried at 120°C overnight in a forced air-desiccating oven. The samples were crystallized and clumped together during the drying period. Therefore, the pellets were separated using a grinder after crystallization, and returned to the oven. The moisture content of the blend resins needs to be lower than 50 ppm. to avoid hydrolytic degradation. The dried resins were fed through a Killion model KL-125 single screw extruder (Werner & Pfleiderer Corporation, Ramsey, New Jersey). The extruder was connected to a 6-inch flexible lip sheet die and 3-roll temperature controlled take-off. The dimensions of the extruder were as follows: Screw diameter 31.25 mm L/D ratio 24:1 Extruder conditions for film fabrication Sample 1 (Shell 8406 PET): Screw speed = 75 rpm, Tm," = 537 °F, P = 1670-1700 psi, drive current = l3-l4amps, take-up speed = 14.9-15.0 fpm T profile (°F) = 515/525/530/530/530/500, Trolls (°F) = 100/ 100/ 100 Sample 2 (25% (wt/wt), processed at 285°C and 1 pass): Screw speed = 75 rpm, Tm.“ = 534 °F, P = 310-320 psi, 4.5-5.5 amps, 14.2-14.3 fpm r profile (F) = 515/525/530/530/530/500, Tm". (°F) = 100/100/100 71 Sample 3 (25%, processed at 300°C and 1 pass): Screw speed = 75 rpm, Tm,“ = 534 °F, P = 540-560 psi, 5-6 amps, l4.2-14.3 fpm T profile (°F) = 515/525/530/530/530/500, Trolls (°F) = 100/100/ 100 Sample 4 (25% (wt/wt), processed at 315°C and 1 pass): Screw speed = 75 rpm, Tm." = 534° F, P = 430-460 psi, 3.5-4.5 amps, 14.2-14.3 fpm T profile (°F) = 515/525/530/530/530/500, Trolls (°F) = 100/100/ 100 Sample 5 (25% (wt/wt), processed at 285°C and 2 passes): Screw speed = 75 rpm, Tm," = 532 °F, P = 260 psi, 4 amps, 14.2-14.3 fpm T profile (°F) = 515/525/530/530/530/500, Trolls (°F) = 100/100/ 100 Sample 6 (25% (wt/wt), processed at 300°C and 2 passes): Screw speed = 75 rpm, Tm," = 530-532 °F, P = 240-260 psi, 3.5-4.0 amps, l4.2-l4.3 fpm. r profile (°F) = 515/525/530/530/530/500, Tm". (°F) = 100/100/100 Sample 7 (25% (wt/wt), processed at 315°C and 2 passes): Screw speed = 75 rpm, Tm,“ = 532-533 °F, P = 300-320 psi, 3.5-4.5 amps, l4.2-14.3 fpm. T profile (°F) = 515/525/530/530/530/500, Trolls (°F) = 100/100/ 100 Sample 8 (10% (wt/wt), processed at 300°C and 1 pass): Screw speed = 75 rpm, Tm." = 538 °F, P = 560-620 psi, 3-4 amps, 14.2-14.3 fpm T profile (°F) = 515/525/530/530/530/500, Tron, °(F) = 100/ 100/ 100 Sample 9 (30% (wt/wt), processed at 300°C and 1 pass): Screw speed = 75 rpm, Tm," = 536-537° F, P = 670-740 psi, 4-5 amps, 14.2-14.3 fpm T profile (°F) = 515/525/530/530/530/500, Trolls (°F) = 100/ 100/ 100 Sample 10 (40% (wt/wt), processed at 300°C and 1 pass): 72 Screw speed = 75 rpm, Tm." = 530-532 °F, P = 740-780 psi, 5-6.5 amps, 14.6-14.7 fpm. T profile (°F) = 515/525/530/530/530/500, Trolls (°F) = 100/ 100/ 100 Sample 11 (Shell 40046 PEN): Screw speed = 75 rpm, Tm," = 580-591 °F, P = 1590-1640 psi, 14-15 amps, 14.4-14.5 fpm. T profile (°F) = 560/575/580/580/580/545, Trolls (°F) = 100/100/ 100 Melting temperature (Tmcn ) and pressure (P) were measured at the end of the extruder barrel. Temperature profiles are listed fiom the feed throat to the die. The final film thickness was approximately 10 mils (0.01 inch). The non-oriented films were found to be uniform and optically clear. The blend films were then biaxially oriented. 4.2.2. Orientation Process The eleven 10 mil films were biaxially oriented on a T.M. Long film stretcher, utilizing a unit located at the Amoco Company facility. The elongation was 300%. The 10 mil films were conditioned at 23°C and 50% RH for 40-48 hours. All of . the samples were stretched 3x3 simultaneously in both directions at a rate setting of 300% per second (6 inches per second). The stretch temperature was 100°C for all samples, except for the PEN film, which was stretched at 135°C. All samples had a soak time of 15 seconds. Twelve biaxially oriented film specimens were prepared for each sample. The final oriented film thickness was approximately 1 mil (0.001 inch). The film samples (non oriented and oriented films) were then characterized with respect to thermal, mechanical, barrier, and morphological properties as a function of composition and orientation conditions. 73 4.2.3. Analysis Methods 4.2.3.1. Thermal analysis The thermal properties of the blend resins were determined using a MDSC 2920 Modulated Differential Scanning Calorimetry (TA Instruments Inc., New Castle, Delaware). All samples were precisely weighed with a Sartorious analytical balance (Sartorius GMBH Gottngen (Germany), Westbury, New York). Optimum weights are normally in the range of 5-10 milligrams. For the blend resins, whole pellets of PET/PEN blend resins were used, if their weights were in the optimum range. For the blend films, films were cut and stacked to achieve the optimum weight. Use of low weight thin samples results in minimal thermal gradients, which facilitates attaining steady state conditions and gives a00urate data. Each sample was placed in an aluminum bottom dish and a top dish was put on. The dishes were sealed using the clamp. An empty dish was used for compensation for the heat flow corresponding to the dishes. The polymer blends were equilibrated at 10°C for 5 minutes before being heated to 290°C (320°C for PEN), with the heating rate 4°C per minute. The moderated mode with an imposed modulation rate of +/- 1°C every 60 seconds was set. A modulated DSC has a superimposed Sinusoidal temperature oscillation, while a conventional DSC has a linearchange in temperature. Therefore, a modulated DSC can separate the total heat flow into reversing and nonreversing components, which gives the benefits of accuracy and ease of result interpretation. The glass transition temperature (T8) was determined from the transition temperature of the reversing heat flow profile. The melting temperature (Tm), and the 74 heat of firsion (AHr) were determined from the endothermic events of the total heat flow profiles. The % crystallinity of blends was determined from the heats of fusion of 100% crystalline blends, which were calculated fi'om the weight average (of the mole fraction of the blend composition) of the heat of fusion of 100% crystalline PET and PEN homopolymers. The heat of firsions are 121.42 J oules/ gram (29.0 calories / gram) and 103.41 Joules/ gram (24.7 calories/gram) for PET and PEN, respectively (Cheng et al. 1988; McGee and Jones 1995). The calculation of the % crystallinity is shown in the following equations. The heat of fusion (AHf) of 100% crystalline blend with X mole % NDC e (X/100 * 103.41) + (100-X)/100 * 121.42 Joules/gram Thus, AHr of 100% crystalline blend with 8.1 mole % NDC = 119.62 Joules/gram AHf of 100% crystalline blend with 21 mole % NDC = 116.92 Joules/ gram AHf of 100% crystalline blend with 25.4 mole % NDC = 116.02 Joules/gram AHr of 100% crystalline blend with 34.6 mole % NDC = 114.22 Joules/gram % crystallinity of 8.1 mole % blend = AHfGoules/gram) * 100 / 119.62 4.2.3.2. Density measurement The density of blend resins, non-oriented films, and oriented films was determined by the density-gradient technique (ASTM Procedure D1505-85, reapproved 1990). The density analysis was performed at room temperature (23°C). 75 The apparatus consisted of a 2-inch diameter and 32-inch height density gradient column, a set of standard calibrated floats (Cole-Palmer Instrument Company, Vernon Hills, Illinois), toluene-carbon tetrachloride solutions, and a set of glass apparatus for gradient tube preparation (Cole-Parmer Instrument Company, Vernon Hills, Illinois). The densities of the set of 5 calibrated floats ranged fiom 1.2494 to 1.4199 g/cc, as follows: Code Color Density (g/cc) Bead 1 Large blue 1.2494 Bead 2 Blue 1.3000 Bead 3 Blue 1.3500 Bead 4 Yellow 1.3997 Bead 5 Red 1.4199 Solution preparation To obtain a density gradient ranging from 1.25 to 1.455 g/cc, a solution A with the density of 1.248 g/cc and a solution B with the density of 1.455g/cc were prepared fi'om the toluene and carbon tetrachloride. The density of the toluene and carbon tetrachloride is 0.87 and 1.59 g/cc, respectively. Solution A was a premixed solution of 375 ml of toluene and 425 ml of carbon tetrachloride, and solution B was a premixed solution of 150 ml of. toluene and 650 ml of carbon tetrachloride. Preparation of the gradient column The glass apparatus for gradient tube preparation was assembled by method C- Continuous Filling in ASTM D1505-90 (Anonymous 1990“”), as shown in Figure 4.1. 76 800 ml of solution A (the lighter solution) and 800 ml of solution B (the denser solution) were placed in volumetric flasks A and B, respectively. The calibrated floats were gently put in the empty gradient column. The gap between the end of the siphon tube and the bottom of the gradient column is very critical and should be minimal, in order to obtain the most uniform density gradient and limit bubble formation. Solution A and B were introduced to the gradient column by turning stopcocks 1 and 2, consecutively. A propeller-type stirrer in flask A was turned on and the stirring speed was adjusted to achieve the continuous mixing of solution A and B. The total flow rate of solutions A and B to the gradient column was approximately 20 ml per rrrinute, and was controlled by monitoring the level of the solutions in flask A and B, and adjusting the stopcocks as needed. The first solution delivered was the lighter end of the gradient (pure solution A with density of 1.25 g/cc). The density gradient of the solution built up as the solution being introduced into the column became progressively denser. After two thirds of the column was filled, the stopcocks were turned off, and the siphon tube was slowly removed from the column. The filled gradient column remained undisturbed for 24 hours to allow the gradient to reach steady state. The positions of the calibrated floats were then measured and the blend samples were tested. Calibration of the standard glass floats The calibrated floats, placed in the column previously, floated to constant positions. These positions corresponded to the density of each float. A calibration curve was obtained by plotting the densities of the calibrated floats against their positions in the 77 gradient column. There was a linear relationship between the positions of the floats in the gradient column and the density of the floats, with a correlation value of 0.99 (see Figure Bl , Appendix B). Testing the specimens The blend resins were dropped directly into the center of the gradient column. The non-oriented and oriented films were cut into different shapes with Size approximately 2X2 mm, for convenience of identification. The specimen samples traveled along the gradient column until they stopped at constant positions. The positions were measured and remeasured again 24 hours later to confirm that there was no further change. The density values of the samples were calculated using the calibration equation (Appendix B). Siphon tube w ./ Stopcock l Stopcock 2 — - l— —( — ‘ 1'1 IIIL Figure 4.1: Apparatus for the gradient tube preparation 78 4.2.3.3. Average sequence lengths of terephthalate and naphthalate units. The estimation of the average sequence lengths of terephthalate and naphthalate units was done using the lHNMR technique, and the Yamadera and Murano equations (see detail in section 3.2.5 ‘H-NMR analysis). Molar fractions of terephthalate units (MT) and naphthalate units (MN) are obtained from the intensities of the NMR spectrum: Mr = Ar-G-N /2 + Arc-r (1) MN - Ann: /2 + AN-G-N (2) MT + MN = 1 (3) The average sequence length of terephthalate units (La) and naphthalate units (LnN) were calculated by Lnr = ZMT / AT-o-N = l/PTN (4) LnN = 2MN / AT-G-N = “PM (5) where B = PNT + PTN (6) B = l/LnT + l/LnN (7) From equations (1) to (7), the following expressions were derived. 14.7 = 1/(1 + Mr)B (8) LnN = 1/(1 + M1013 (9) where Ann , AN-G-N, and A1204»; are the ratios of the intensities of the PET peak, the PEN peak, and the interaction peak of PET and PEN, respectively, with the total intensities of these three peaks. The fiaction of the T-G-T, N-G-N and T-G-N linkages 79 can be obtained from the intensities of the PET, PEN and PET/PEN peaks of the 1HNMR spectrum. B is the degree of transesterification. Therefore, from (8) and (9), the average sequence length of terephthalate units (Lnr) and naphthalate units (LnN) decreases, when the degree of transesterification increases. 4.2.3.4. Molecular weight measurement The molecular weight measurements were performed using the size exclusion chromatography/multi-angle laser light scattering (SEC-MALLS) technique. Chromatographic system consists of a Waters model 510 pump (0.6 ml/min), a Validyne model 7025 injector, 2 PLgel C columns, a Wyatt Dawn DSP-F multi-angle laser light scattering detector, an applied Biosystems model 757 variable wavelength ultraviolet detector, and a Waters model 410 differential refiactometer (DRI). Approximately 25 milligrams of each sample were weighed and mixed with 2.0 milliliters of a solvent of chloroform and trichloroacetic acid. Chloroform and trichloroacetic acid were used as the solvent or eluent in all experiments. The solutions ' were gently mixed overnight to ensure homogeneity. Only the solution of PEN homopolymer was heated during mixing. The solutions were then injected directly into the SEC-MALLS. The data were collected using Wyatt’s Astra software (version 4.2). 4.2.3.5. Morphological analysis The morphological analysis was performed using Raman Chemical Imaging microscopy. A Rarnan image microscope is a new type of instrument. It is an optical microscope or laser confocal scanning microscope (LCSM) with a filter or spectrometer 80 (Sawyer and Grubb 1996). The Raman imaging micr0800pe was diagrammed and described in detail elsewhere (Schaeberle and Treado 1997; Schaeberle et al. 1999; Morris et al. 1999). In this study, the Raman analysis was performed by ChemIcon Company, Pittsburgh, PA. A ChemIcon FALCON Rarnan microscope with Liquid Crystal Tunable Filter (LCTF) imaging spectrometer and Silicon Charge-Coupled Device (CCD) detector (Chemlcon Company, Pittsburgh, PA) was used to conduct the experiment. Four resin and two film samples, listed in Table 4.1, were tested. The samples were manually cross-sectioned. Final thickness was approximately 25 um. Each sample was then immobilized on a glass microscope slide and the side was placed under the Raman microscope. Laser excitation was achieved by a 532 nm Nd: YAG laser which yielded 0.2 to 1.6 watts of power at the exiting laser head. Sample excitation and Raman scatter collection was performed using a 20X (0.46 NA) and 100X (0.95 NA) objectives on the Rarnan microscope. Once the laser scatter was ejected, the Raman emission was filtered with LCTF. Raman images were collected using a 16-bit dynamic range, liquid nitrogen- cooled (-40°C), slow scan charge—coupled device (CCD) detector. The detector has 512 x 512 pixels. Pixel size is 9 pm. The CCD exposure times were 5 seconds for PET, 2 seconds f0r PEN, and 10-30 seconds for blends. 81 Table 4.1: List of the samples tested for Raman chemical imaging analysis Sample ID PEN Type of Blending Blending Types of composition extruder Temperature Time sample (% Weight) (°C) (Number of asses) . 300P1 single 25 Single 300 l Resin . 300P1 25 Twin 300 1 Film and resin .315P2 25 Twin 315 ' 2 Resin . 300Pl-40% 40 Twin 300 1 Film and resin The brightfield reflectance, polarized light, and Raman chemical images of each sample with 20X and 100X magnification were collected. The dispersive Raman spectra collected from approximately 600-3200 cm“were also acquired. 4.3. Results and Discussions 4.3.1. Thermal Properties of Blends The blend resins and films obtained from one and two passes through the extruder at temperatures of 285, 300 and 315°C showed one glass transition temperature and one melt temperature. Moreover, these T8 and T m values fell between the T8 and Tm values of pure PET and PEN, respectively. All blend samples were optically elear, which indicated that the blends were miscible. However, fiom the DSC analysis, the peaks corresponding to the melting temperatures of the blends were broader than those of PET and PEN homopolymers. According to Andresen and Zachmann (1994), the results suggest that the blends were only partially miscible. 82 The following sections will discuss the effects of the transesterification reactions and the blend composition on the T8 and Tm values and the % crystallinity of the blends. 4.3.1.1. Glass transition temperature (Tg) The results showed that the Tg values for PET/PEN blends were independent of the degree of transesterification, but were directly proportional to the molar compositions of PEN in the blend. This conclusion agrees with Stewart et al. (1993), Hoffman and Caldwell (1995), McGee and Jones (1995), P0 et al.(1996) and Shi (1998). TS and degree of transesterification DSC analysis showed that the T8 values of the blends with 25% (wt/wt) PEN were in the range of 77 to 84°C. Those blends, which were processed at 285 to 315°C, and one pass and two passes through the extruder, had a degree of transesterification ranging fi'om 0.061 to 0.525. I There were no statistically significant differences between these T8 values (Table 4.2 and Figure 4.2). The statistical analysis is shown in Appendix C, Table C3. Shi (1998) and McGee et al. (1995) found that after PET/PEN blends achieved a certain degree of transesterification, a further increase in the degree of transesterification has little effect on the resultant T8 value. Stewart et al. (1993) suggested that the broadened range of the T8 values occurred due to the partial miscibility of the polymer blends and the lack of sensitivity of DSC measurement. The partial miscibility can account for both the broad range of T8 values obtained and the film haziness. Based on 83 Table 4.2 : The effect of the degree of transesterification (B) on the glass transition temperature (T g) of the PET/PEN blends. T. (°C) sample B oreinted films non-oriented films resins PET - 86.07 73.66 76.82 #2 0.069 84.56 82.79 76.87 #3 0.221 83.55 82.79 82.04 #4 0.333 84.79 80.97 78.63 #5 0.172 85.49 80.48 82.51 #6 0.394 83.98 81.33 84.28 #7 0.519 84.38 81.51 81.77 PEN - 114.01 120.87 119.52 * Blends with 25% PEN (wt/wt) processed through a twin-screw extruder( 150 rpm) 0.6 100 3 E 95 -. g 90 ~r 5 5'" 85 T 3 § 0 g % l-a' g D '3 c v 80 7” a 3 3 3 75 .. 3 a e 70 -. :5 65 % T7 1 i i 0.0 0.1 0.2 0.3 0.4 0.5 The degree of transesterification (B) l o oriented films 0 non-oriented films A resins Figure 4.2: The effect of the degree of transesterification on the glass transition temperature of the PET/PEN blends 84 the T8 values obtained, the blends with lower levels of transesterification exhibited greater partial miscibility, while the Tg values for the polyblends which achieved a higher level of transesterification fell within a narrower temperature range. In addition, Stewart et al. (1993) found that the polymer blends further reacted during the DSC measurement. T8 and blend composition The average Tlg values of the blends with 10 to 40% (wt/wt) PEN and processed at 300°C and 1 pass were in the range of 78 to 89°C. The average T8 values of PET and PEN resins were 77 and 120°C, respectively. The average T8 values of PET/PEN blends with various blend compositions are tabulated in Table 4.3. Figure 4.3 shows the plot of T8 versus PEN composition. The average T8 values of the blend resins and non-oriented films were in the same range. Moreover, there is a linear relationship between those T8 values and the PEN composition. However, for the case of oriented films, the linear expression does not fit well. Instead, a second degree polynomial expression gives the best fit. The orientation effect will be discussed in Chapter 6. The widely known expression based on the free volume model was used 'to estimate the T8 of the copolymers and blends (DiMarzio and Gibbs 1959; Gordon and Taylor 1952). The expression is: T = wl(Tg)1+kw2(Tg)2 g wl+kw2 (10) 85 Table 4.3 : The effect of PEN composition on the glass transition temperature of the PET/PEN blends. T. (°C) sample PEN mole % oriented films non-oriented films] resins PET 0.0 86.07 73.66 76.82 #8 14.5 84.27 76.78 77.48 #3 33.7 83.55 82.79 82.04 #9 39.5 86.46 84.23 85.89 #10 50.4 84.39 88.88 89.08 PEN 100.0 117.00 120.87 119.52 * Blends processed at 300°C and 1 pass through a twin-screw extruder. The glass transition temperature (T .°C) non-oriented films y = 0.4767x + 72.841 R2=0.9989 resins y = 0.4438x + 74.481 R2 = 0.9921 go . 0 8° . oriented films 70 ' y = 0.004518 - 0.1712x + 85.881 R2 = 0.9918 60 r i ‘r i i 0 20 40 60 80 PEN mole % i 0 oriented films 0 non-oriented films A resins Figure 4.3 : The effect of PEN composition on the glass 86 transition temperature of The PET/PEN blends 100 where w), w, Ty, and T82 are the weight fractions and T1; values of pure-component l and 2, respectively, and k is a fitted constant. When k is equal to l, a linear relationship between T8 and blend composition is found (Rodriguez, 1996). The expression can be simplified as; g x l g 2 However, the above expression is applied specifically to the blends in which polymer- polymer interactions are weak. A comparison between the calculated T8 values, based on this expression and the measured Tg values were shown in Table 4.4 and plotted graphically in Figure 4.4. Even through, both calculated and measured T3 values seem to be in the same range, statistical analysis showed significant differences between both T8 values, at the confidence level of 5% (Appendix C, Table C3). The modified expression, which accounted for the effect of the polymer-polymer interaction was studied by Kwei (1984) and Kwei et al. (1987). They proposed the extended expression, such that T _ w,(Tg)l +kw2(Tg)2 +qwlw2 (12) g wl +lrw2 where q is the nature and magnitude of the polymer-polymer interactions. However, to apply this expression to the PET/PEN blend system, more information is required. 87 Table 4.4: The comparison between calculated and measured Tg of the blend resins sample PEN mole% Calculated Tg Measured Tg PET 0.0 76.8 76.8 #8 14.5 79.7 77.5 #3 33.7 84.4 82.0 #9 39.5 86.0 85.9 #10 50.4 89.6 89.1 PEN 100.0 119.5 119.5 * Blends processed at 300°C and 1 pass through a twin-screw extruder. 140 (To. °C) _. 8 8 8 . J Glass transition Temperature 0 ..A N O 3.3 N O IL Calculated T9 y = 0.4324x + 72.18 R2 = 0.9375 Measured T9 . y = 0.4429x + 70.892 r:2 = 0.9277 O 20 40 PEN mole % 60 80 100 l o Calculated Tg A Measured Tg l Figure 4.4: The conparison between the calculated and measured T8 of the blend resins 88 4.3.1.2. Melting temperature (Tm) The results showed that the degree of transesterification and the blend composition were the primary factors controlling the melting temperature of the blends. The melting temperature decreased as the degree of transesterification increased. Furthermore, depression of the Tm was found, when the polymers were blended as was expected. The minimum Tm values of the PET/PEN blends were obtained when the blend composition was 40 to 50 mole % PEN. T... and degree of transesterification The Tm values of blends with 25% (wt/wt) PEN ranged from 226 to 247°C. These blends were processed at 285 to 300°C, and l and 2 passes, and transesterification degree ranged fi'om 0.061 to 0.525. The results showed a linear relationship between the degree of transesterification and the Tm of the PET/PEN blends studied (Table 4.5 and Figure 4.5). As shown, the Tm of the blends appears to decrease as the transesterification degree increases. This melting temperature depression can be explained by the polymer interaction mode (Paul and Bucknall 1999). Miscible blends have lower chemical potentials, compared to their pure components. Thus their crystals are in equilibrium with the miscible amorphous phase, at a lower temperature. Therefore, the melting temperature of blends decreases. 89 Table 4.5 : The effect of the degree of transesterification (B) on the melting temperature (Tm) of the PET/PEN blends. Tm (°C) sample B oriented films non-orientfims resins PET - 248.71 250.07 248.54 #2 0.069 244.95 245.80 247.33 #3 0.221 237.57 237.12 240.45 #4 0.333 233.33 234.10 233.01 #5 0.172 239.66 241.38 242.23 #6 0.394 231.10 232.30 231.86 #7 0.519 226.16 225.95 226.85 PEN - 268.65 268.59 261.64 " Blends with 25% PEN (wt/wt) processed through a twin-screw extruder( 150 rpm) 250 c oriented films 245 ._ y = 40.82x + 247.54 R2 = 0.9946 240 t resins y = 46.723x + 250.78 235 , R2 = 0.9937 230 :— non-oriented films y = 42.366x + 248.65 The melting temperature (Tm,°C) 22 «— 5 R2 = 0.9732 220 . ' . 0 0.1 0.2 0.3 0.4 0.5 0.6 The degree of transesterification (B) . O oriented films 77 D noLQriented films A gins Figure 4.5 : The effect of the degree of transesterification on the melting temperature of the PET/PEN blends 90 T... and blend composition The average T... values of the blends with 10 to 40% (wt/wt) PEN, processed at 300°C and l and 2 passes, were in the range of 231 to 246°C. The average T... values of PET and PEN resins were 249 and 263°C, respectively. The results showed that the blend composition has a strong influence on the T... of the blends. The relationship between PEN molar composition and the T... is shown in Figure 4.6 and summarized in Table 4.6. The T... values of the blends were lower than those of the PET and PEN homopolymers. The minimum T... values were found to appear with the blends consisting of 40 to 50 mole % PEN. When the PEN composition was lower than 50 mole % PEN, the T... decreased with an increase in PEN composition. When the PEN composition was greater than 50 mole % PEN, the T... increased with an increase in PEN composition. Shi (1998) found a similar relationship between T... and PEN compositions, with the lowest T... occurring at 33 mole % NDC or 50 mole % PEN. The depression of the T... could be explained by the polymer interaction model (Paul and Bucknall 1999). In this model, the critical blend composition is approximately 50% wt/wt for blends with similar molecular weight, or closest to the component with the lower molecular weight. At the critical composition, blends have the lowest interaction energy. This causes the crystals to be in equilibrium with the amorphous phase at the lowest temperature. Shi (1998), and Lu and Windle (1995) also explained this phenomenon by proposing that when a small amount of PEN is added to PET, the crystalline structure of the PET rich phase is disrupted and the crystallite size is limited, resulting in the depression of T... Above 50 mole % PEN, PEN can now be considered the continuous phase, and adding PET can be 91 Table4.6: Theefi'ectofPEl‘lconpositiononthennltingtenperatureof the PET/PEN blends Tm (°C) sarrple PEN Mile % oriented filrm non-oriarted films resins PET 0.0 248.71 250.07 248.54 #8 14.5 245.06 245.91 246.42 #3 33.7 237.57 237.12 240.45 #9 39.5 235.72 234.93 235.81 #10 50.4 232.27 231.03 232.62 PEN 100.0 268.65 268.59 261.64 *Blerxlsproeessedat300°Candlpessduoughahwin—screwexmxla. non-orientedfilms y=0.0111x2-0.9378x+251.33 . R2=0.9943 oriented filrrs E y = 0.010118 - 0.8259x + 249.83 The melting temperature (Tm,°C) i 9 a e a e i 3 a: e «3 J 3 y=0.W-0.7676x+250.34 R2=0.9723 0 20 40 60 80 100 - PENmole% l Oorientedfilrrs Dnon-orientedfilrrs Aresins l fimlfifl'heefieaotPENeonpoeldononthennltim Mn ofthePETlPBlblende 92 viewed as disrupting the crystalline structure. The disruption of the crystalline structure might cause a decrease in the interaction energy, which results in reduction of T... 4.3.1.3. Percent crystallinity of the blends The results showed that the degree of transesterification and the blend composition strongly affected the % crystallinity of the blends. The % crystallinity decreased as the degree of transesterification increased. Furthermore, a reduction of the % crystallinity was found, when adding a polymer to another (Shi 1998). The minimum % crystallinity appeared in the blends with 40 to 50 mole % PEN. Percent crystallinity and degree of transesterification The % crystallinity of blends with 25% (wt/wt) PEN ranged from 23% to 38%. These blends were processed at 285 to 300°C, and 1 and 2 passes, respectively, and the transesterification degree of these blends ranged from 0.061 to 0.525. The results showed a linear relationship between the degree of transesterification and the % crystallinity of the PET/PEN blends (Table 4.7 and Figure 4.7). As shown, the % crystallinity of the blends appeared to decrease, as the transesterification degree increased. When the transesterification reactions increase, PET and PEN chains are shortened, and the number of heterolinkages of terephthalate (T) and naphthalate (N) units increases. This new structure of T-N chains would not easily form a crystalline phase, and would interfere with the formation of crystals. The shortening of the sequence length of the crystallizable components can account for the decrease in the 93 Table 4.7 : The effect of the degree of transesterification (B) on the percent crystallinity of the PET/PEN blends %Crystallinity sample B oriented films non-oriented films resins PET 40.94 32.75 50.45 #2 0.069 37.84 35.11 33.40 #3 0.221 30.83 25.60 28.88 #4 0.333 28.57 27.19 23.77 #5 0.172 34.66 31.53 28.66 #6 0.394 27.67 29.36 26.85 #7 0.519 25.70 24.46 22.88 PEN - 46.47 46.53 64.54 "' Blends with 25% PEN (wt/wt) processed through a twin-screw extruder( 150 rpm) 45 4o -- oriented films 5 35 y = -27.149x + 38.599 5 “i R2 = 0.9297 3:; 30 5‘ 25 . ‘5' § 20 " a 15 1 .3 10 «L resins non-oriented films I- y = -21.289x + 33.459 y = 48.82:! + 34225 5 «r R2 = 0.3169 R2 = 0.5937 0 t : i t . 0.0 0.1 0.2 0.3 0.4 0.5 The degree of transesterification (B) r 0 oriented films 0 non-oriented films A resins l Figure 4.7: The effect of the degree of transesterification on the percent crystallinity of the PETIPEN blends 94 0.6 ability of the blends to crystallize (Eersels and Groeninckx 1996; Eersels and Groeninckx 1997; Fakirov et al. 1996 (8" (b); Lenz and Go 1973). Therefore, crystallinity is limited when random copolymers are achieved. Random copolymers with 15 to 85 mole % NDC appear to be in the amorphous phase and no crystalline phase is found (Callander and Howell 1994; Hoffman and Caldwell 1995; McGee and Jones 1995; Shi 1998). Percent crystallinity and blend composition The % crystallinity of the blends with 10 to 40% (wt/wt) PEN and processed at 300°C and l and 2 passes is in the range of 20.5% to 36.5%. The % crystallinity of PET and PEN resins were 50% and 64.5%, respectively. The results Showed that the blend composition is a primary factor in determining the % crystallinity of the blends. The relationship between the PEN composition and the % crystallinity was expressed as a polynomial equation (Table 4.8 and Figure 4.8). This relationship is similar to the one between T... and the PEN composition. From our data, the % crystallinity of the blend is a minimum, when PEN composition is 40 to 50 mole % PEN. The % crystallinity of PET and PEN is the highest at both ends (PET and PEN homopolymers). By adding the PET to the PEN rich phase, the crystalline structure was disrupted and the % crystallinity decreased. A similar behavior occurs when adding PEN to the PET-rich phase. 4.3.2. Density of Blend Resins and Films The density of the blend resins and films was determined by the density gradient technique. The results showed that the density of the blends is dependent on the degree of transesterification and the blend composition. The density of the blends was 95 Table 4.8 : The effect of PEN composition on the percent crystallinity of the PET/PEN blends %Crystallinity sample PEN mole% oriented films non-oriented films resins PET 0.0 40.94 32.75 50.45 #8 14.5 34.54 36.49 34.24 #3 33.7 30.83 25.60 28.88 #9 39.5 27.07 27.79 23.97 #10 50.4 23.39 21.02 20.48 PEN 100.0 46.47 46.53 64.54 * Blends processed at 300°C and 1 pass through a twin-screw extruder. 100.0 70 resins oriented films 60 y = 0.0147x2 - 1.3068x + 48.089 y = 0.008x2 - 0.7422x + 40.987 3. R2 = 0.9817 R2 = 0.9816 g 50 - f. z- 40 ° 0 *c" 0 30 . 2 8 g 20 ~ 1- non oriented films 10 _ y = 0.007):2 - 0.5947x + 36.222 R2 = 0.8789 0 . a: i . 0.0 20.0 40.0 60.0 80.0 PEN mole °/e l 0 oriented films a non-oriented films A resins i Figure 4.8 : The effect of PEN composition on the percent crystallinity of the PETIPEN blends 96 considerably lower than that of the PET and PEN homopolymers. However, the density values for the blend films were slightly higher than those of the blend resins, and alter the orientation process, the density of the oriented films increased Significantly. Density and degree of transesterification The density of blends with 25% (wt/wt) PEN ranged from 1.3379 to 1.3503 g/cc (Table 4.9). These blends were processed at 285 to 300°C, and 1 and 2 passes, respectively, and the transesterification degree ranged from 0.061 to 0.525. The results showed that there is a linear relationship between the degree of transesterification and the density of the PET/PEN blends (Table 4.9 and Figure 4.9). The density of the blends appeared to decrease, as the transesterification degree increased. These results were in agreement with the effect of the transesterification reactions on the % crystallinity. Both density and % crystallinity respond to the packing ability of the polymer molecular chains. It should also be pointed out that as transesterification reactions occur, the formation of the T-N interlinkages increases. These N-T bonds would not be conductive to polymer chains packing as tightly together as those of the individual polymers. Therefore, the density of the blends decreases. 97 Table 4.9: The effect of the degree of transesterification on the density of the blends sample ID Tansesterification Density (g/cc) (B) resins non-oriented films oriented films 2 0.068 1.3389 1.3395 - 5 0.172 1.3383 1.3390 1.3503 3 0.217 1.3385 1.3390 1.3491 4 0.329 1.3384 1.3388 1.3488 6 0.389 1.3385 1.3390 1.3460 7 0.522 1.3379 1.3384 1.3459 * Blends with 25% PEN (wt/wt) processed through a twin-screw extruder( 150 rpm) Density (glee) 1.3520 oriented films 1-3500 ~~ y = -0.0129x + 1.3522 2 - 1.3480 + R 0.8455 1.3460 4L non-oriented films 13440 «i y = -0.002x +1.3395 R2 = 0.7894 1.3420 «r - 1.3400 -- _ resins 13380 a. y = -0.0017x '1' 1.3389 R2 = 0.6433 1.3360 1 : e T. 0 0.1 0.3 0.4 0.5 0.6 Transesterification degree (8) l A resins U non-oriented films 0 oriented films i Figure 4.9: The effect of the dree of transesterification on the density of the blends Density and blend composition The density of PET and PEN resins was 1.4041 and 1.3640 g/cc, respectively. The density values of the blends with 10 to 40% (wt/wt) PEN, processed at 300°C and 1 and 2 passes, were in the range of 1.3367 to 1.3535 g/ce. The density of the blended resins was significantly lower than the density of PET and PEN resins. The density of the resins seemed to decrease very slightly, as the PEN composition increased (Table 4.10, and Figure 4.10). The density of the blends decreased tremendously, following the fihn processing. The results showed that the densities of the PET and PEN non-oriented films were much lower than those of PET and PEN resins (Table 4.11 and Figure 4.11). However, afier the biaxial orientation process, the density of the PET, PEN and blend oriented films increased. There is a linear relationship between the density of non-oriented films and PEN composition (Table 4.12 and Figure 4.12). However, the relationship between the density of the oriented films and the PEN composition can be best described by a polynomial equation with a correlation coefficient (r-square value) of 0.9 (Table 4.12 and Figure 4.12). The reduction of the density of the blends by adding PEN molecules to the PET rich phase is due to the interference of PEN molecules in the regularity of the PET structure. This would result in the decrease in the ability of the polymer chains to pack close together. 99 Table 4.10: The effect of PEN composition on the density of the PET/PEN blend resins Density Sample ID PEN wt % PEN mole % of blend resins (g/cc) PET o 0.0 1.4041 ‘ 8 10 14.5 1.3385 3 25 33.7 1.3385 9 30 39.5 1.3371 10 40 50.4 1.3367 PEN 100 100.0 1.3640 "‘ Blends processed at 300°C and 1 pass through a twin-screw extruder. 1.4100 1.4000 3 .8 1.3900 ._ 1.3800 J- Density (glee) is {3 8 8 ..— 1.3500 4- 1.3400" I; ,,,,,,,,,, a... ..... 15"". 1.3300 . . .1 ; 0 20 40 60 80 100 PEN mole % Figure 4.10 : The effect of PEN composition on density of the PET/PEN blend resins 100 Table 4.11: The density of the PET, PEN and blends Blending temp. Blending time Density (g/cc) sample 1D % PEN mole (0C) (No. pass) resins non-oriented films oriented films PET 0 - - 1.4041 1.3429 1.3621 2 33.7 285 1 1.3389 1.3395 1.3481 3 33.7 300 1 1.3385 1.3390 1.3491 4 33.7 315 1 1.3384 1.3388 1.3488 5 33.7 285 2 1.3383 1.3390 1.3503 6 33.7 300 2 1.3385 1.3390 1.3460 7 33.7 315 2 1.3379 1.3384 1.3459 8 14.5 300 1 1.3385 1.3421 1.3535 9 39.5 300 1 1.3371 1.3381 1.3467 10 50.4 300 1 1.3367 1.3371 1.3483 PEN 100 - - 1.3640 1.3316 1.3446 1.4050 1.3950 1.3850 73 2 1.3750 9 z. 1 3650 g 1.3550 ‘3 1.3450 1.3350 f' oriented ilms 1'3250 7 7 non-oriented films sample ID resins Figure 4.11: The density of the blend resins and films 101 Table 4.12: The effect of PEN composition on the density of the PET/PEN blend film Densitiy (_g/ee) Sample ID PEN wt % PEN mole % non-oriented films oriented fihns PET 0 0.0 1.3429 1.3621 8 10 14.5 1.3421 1.3535 3 25 33.7 1.3390 1.3491 9 30 39.5 1.3381 1.3467 10 40 50.4 1.3371 1.3483 PEN 100 100.0 1.3316 1.3446 * Blends processed at 300°C and 1 pass through a twin-screw extruder. 1.3650 1; 1.3600 \ oriented films ‘\. . y = 4E-06x2 - 0.0006x + 1.3599 1.3550 -- ‘ R2 = 0.8935 1\ \ 3 1.3500 + \\§\\ 32 §\ E 1.3450 1 a c 8 1.3400 «- 1.3350 .. non-oriented films 1.3300 .1 y = -0.0001x + 1.3419 R2 = 0.939 1.3250 1 1r i . 0 20 40 60 80 100 ‘ PEN mole% 1 u non-oriented films 0 oriented fihns } 1 Figure 4.12: The effect of PEN composition on density of the PET/PEN blend films 102 4.3.3. Average Sequence Lengths of Terephthalate and Naphthalate Units. The average sequence lengths of the terephthalate and naphthalate units are reported in Table 4.13 and Figure 4.13. Yamadera and Murano (1967) proposed a relationship between the degree of transesterification and the average sequence lengths, which was discussed in Section 4.2.3.3. The results confirmed that as the transesterification reactions occur, the lengths of PET and PEN chains are shortened. In addition, the number of interlinkages of terephthalate and naphthalate bonds increases. The effect of the blend composition on the average sequence lengths was also studied, as shown in Table 4.14 and Figure 4.14. The results showed that as the PEN composition increased (up to 50 mole % PEN), the average sequence length of terephthalate units decreased significantly, while those of naphthalate units increased slightly. 4.3.4. Molecular Weight of Blends Only 6 samples of the blend films (non-oriented films) were selected for molecular weight measurement. The weight average (MW), number average (Mn) and 2- average molecular weight (M2) of PET, PEN, and selected PET/PEN blends are reported in Table 4.15. The results showed that the molecular weight of PET was higher than that of PEN. This is in agreement with the density analysis showing that the PET density was higher than the density of PEN (Table 4.11). Furthermore, the molecular weight averages for the blends were lower than those of PET and PEN. This is due to the effect of the 103 Table 4.13: The efeect of the degree of transesterification on the average sequence length of the terephthalate (LIT) and naphthalate(L.N) units. sample [D B-an [airman LnN—ruins LnT-films LnN-fiims 2 0.076 84.7 21.3 63.9 16.7 3 0.222 21.8 5.8 21.6 5.7 4 0.356 15.3 4.1 12.8 3.6 5 0.192 34.4 8.2 26.7 6.6 6 0.405 12.6 3.3 12.0 3.1 7 0.525 8.1 2.6 8.7 2.4 * Blends with 25% PEN (wt/wt) processed through a twin-screw extruder( 150 rpm) The average sequence length (La and Lu». units) a: .0 o 0-0 . . ‘F9—‘1'. 0 0.1 0.2 0.3 0.4 0.5 0.6 The degree of transeuerlfleatlon (B) 1 A LnT-resins A LnN-resins - LnT-fihns o LnN-films 1 Figure 4.13: The effect of the degree of transesterificaiton on the average sequence length of the terepahthalate and naphthalate units 104 Table 4.14: The effect of PEN composition on the average sequence length (1.“) of the terephthalate and naphthalate (LnN) unit sample ID PEN mole% LnT-resins 1mm... LnT-films Imam 8 14.5 69.8 5.9 81.4 5.3 3 33.7 21.8 5.8 21.6 5.7 9 39.5 23.0 7.3 19.1 6.4 10 50.4 24.9 1 1.3 15.0 8.0 * Blends processed at 300°C and 1 pass through a twin-screw extruder. 110 100~ The average sequence length (La and L"... units) 8 8 8 8 8 5‘ 8 8 10+ t I 0.0 f 10.0 20.0 30.0 PENmole% 40.0 50.0 . [ a LnT-resins A LnN-resins I LnT-films i 1 l J Figure 4.14 : The effect of PEN composition on the average sequence length of terephthalate and naphthalate units 105 60.0 Table 4.15: The molecular weight average of the blend films Figure 4.15: The effect of PEN composition on the molecular weight average of blend film (non-oriented film) 106 Blend Blend time sample ID PEN mole% Temp. (0C) no. of pass B Mn Mw Mz PET 0 46400 65400 82900 3 33.7 300 1 0.074 26600 32400 37200 4 . 33.7 315 1 0.123 27100 34400 42800 6 33.7 300 2 0.133 18200 27000 35100 10 50.4 300 1 0.192 22000 32700 40500 PEN 100 31300 42900 56100 90000 1 30000 .. M2 y =13.522x2 -1600.7x + 81619 ‘ 70000 t .. R2 =0.9515 MW :5. 60000 . i y = 9.6902x2 -1182.7x + 64653 4‘ 2 = E 50000 ,_ . R 0.9701 g "\g - - ' E 40000 -~ g 4 o .- 7‘ I I a 30000 \X/ 20000 —» Mn ‘ y = 6.6629x2 - 817.81x + 46435 10000 " R2 = 0.9999 0 1 1 ‘r a 0 20 40 60 80 100 PEN mole "lo 1 0 Mn - MW 3 Mz y transesterification reactions. The transesterification reactions between the hydroxyl or carboxyl groups of PET and PEthauses a decrease in the lengths of the polymer chains. When comparing the average molecular weight values of blends processed at 300°C, the results showed that blends processed through the extruder twice had lower average molecular weight values than the blends processed through the extruder only once. The plot between the mole % PEN and the'respective molecular weight of the blends is shown in Figure 4.15. Polynomial expressions were found to give the best fit. This shows a similar behavior as the effect of PEN composition on Tm, % crystalinity, and the density. The minimum molecular weight values of the blends were found at a blend composition of 45-50 mole % PEN. This confirmed that adding one polymer to. another polymer rich phase would disrupt the blend structure and cause a decrease in molecular weight. The effect of the blend composition on the molecular weight will be crucial only when the transesterification reactions exist. The polydispersity, or the ratio of M“, and Mn, which is a measure of the breadth of the molecular weight distribution, is reported in Tables 4.16 and 4.17, and presented graphically in Figures 4.16 and 4.17 . The polydispersity of the PET and PEN films were 1.41 and 1.37, respectively. The blend films processed at 300 and 315°C and 1 pass had a lower polydispersity than those of PET and PEN homopolymers. However, the polydispersity of blend films processed at 300 OC and 2 passes was higher than those of the PET and PEN. This might be due to the degradation of the blends during the processing of the second pass through the extruder. 107 Table 4.16: The effect of the degree of transesterification 0n the polydispersity (MW/Mn) of the blend films Blending Blend time sample ID Temp. (0 C) no. of pass B Man PET 1.41 3 300 1 0.074 1.22 4 315 1 0.123 1.27 6 300 2 0.133 1.48 PEN 1.37 * Blends with 25% PEN (wt/wt) processed through a twin-screw extruder( 150 rpm) r" O) O .3.; N3 CO 9 .0 o: o The polydispersity ("w/Mn) p o «5 on O 0 .° N o 8 0.074 0.123 0.133 The degree of transesterification (8) Figure 4.16: The effect of the degree of transesterification 0n the polydispersity of the blend films 108 Table 4.17: The effeet of PEN composition on the polydispersity (Winn) of the blend films PEN sample 10 mole% Mwan Mn Mw PET 0 1.41 46400 65400 3 33.7 1.22 26600 32400 10 50.4 1.49 22000 32700 PEN 100 1.37 31300 42900 * Blends processed at 300°C and 1 pass through a twin-screw extruder. .4.-tee 810:503 C00 The Polydispersity (WM) 9 9.0.0.0 8888 0 33.7 50.4 100 PEN mole % Figure 4.17: The effect of PEN composition on the polydispersity (MW/Mn) of the blend films 109 Comparing the polydispersity of the blends with the blend composition of 33 and 50 mole % PEN, the results showed no relationship between those values. Therefore, the polydispersity of the blends appears to be independent of the blend composition. However, replicate experiments are required to confirm this conclusion. 4.3.5. Morphology The dispersive spectra of PET, PEN and selected blends are shown in Figures 4.1.8 and 4.19. The brightfield reflectance, polarized light and Raman chemical images of each sample with 20X and 100X magnification were obtained and are shown in Figures 4.20 to 4.24, respectively. Dispersive Raman spectroscopy provided unique information for differentiation between PET and PEN homopolymers. This included a shift of the aromatic C=C feature from 1608 cm'1 (PET) to 1629 em"(PEN). Raman spectroscopy of the blend samples showed the combined spectral features from both PET and PEN. Moreover, they revealed a variation in the relative spectral peak height in localized PET and PEN domains. The samples were tested as fihn and resin form. Testing on film and resin samples seems to give similar results. For the sample 300P1-sing1e (blend processed through single screw extruder at 300°C and 1 pass), the brightfield reflectance, polarized light and Raman chemical image revealed an agreeable result and showed evidence of sample heterogeneity (Figure 4.20). The images from both low (20X) and high (100X) magnification yielded molecular contrast between PET and PEN components. From Figure 4.20, the blue region, corresponding to the Raman shifi of 1608 em'l, was indicative of the PET domain, and 110 the green region, corresponding to the Raman shift of 1629 cm", was indicative of the PEN domain. These images suggested that PEN is a minor or dispersive component with a submicron average domain size within the PET continuous phase. This result is consistent with previous results, in that the blend processed through the single screw extruder was immiscible and exhibited phase separation. Thermal analysis of this sample indicated the phase separation of the blends by showing 2 T g and Tm values. This also confirmed that PEN has a higher viscosity than PET (McGee 1995), therefore PEN is likely to be the dispersive domain within the PET continuous domain. For other blend samples processed through the twin-screw extruder, 300P1, 315P1 and 300P1-40% (Figures 4.21 to 4.24), the brightfield reflectance, polarized light and Raman chemical images showed evidence of sample homogeneity, at the spatial resolving power of this microscope technique (approximately 200 nm). The images from both low and high magnification indicated the same results. The results are in the agreement with the thermal analysis indicating that those blends were miscible and they exhibited only one T8 and Tm. From the NMR analysis, the transesterification degree of samples 300P1, 315P1, and 300P1-40% varied fiom 0.161 to 0.520. In conclusion, the morphological analysis revealed that the transesterification reactions yielded miscibility of the blends, and caused a reduction in size of the PET and PEN chains. At the spatial resolving power of this microscope technique (approximately 200 nm), an occurrence of total miscibility or a single phase cannot be concluded. This is because Utracki (1990) reported that phase separation could still be found when the domain sizes of blend components are smaller than 10 A°. Moreover, blends with a single T8 do not always imply that blends are totally miscible (U tracki 1990). Single Tm 111 peaks from the DSC analysis indicated that miscibility is obtained. However, the peaks were broad, which suggest that blends were only partially miscible, not totally miscible. In this experiment, the size of the PEN domain of the blends processed through the twin-screw extruder cannot be determined. In order to improve the resolution of the Raman Chemical Imaging microscope, a microtomed sample is required. The microtomed sample can minimize the contribution from out-of-plane components. Another recommendation is to use the Transmission Electron Microscope (TEM), which can give a higher spatial resolving power. 112 Offset Intensity 300P1 25%PEN MM 300P1 25%PEN Bead inf __ g . N M 1000 1500 2000 2500 3000 Raman Shift (cm") Figure 4.18: Dispersive Raman Spectroscopy of 300P1 25% (wt/wt) PEN polymer samples. 113 Offset Intensity 315P2 (25% PEN) - - _ AW 300P1-Single (25% PEN) MM . A —M— 300P1-40% PEN 1“?” m... , 5 -‘W 1100 1600 2100 2600 3100 Raman Shift (cm-1) Figure 4.19: Dispersive Raman Spectroscopy PET/PEN blend resins with.25% and 40% (wt/wt) PEN processed at 300 °C and 1 pass, and 315°C and 2 passes. 114 Polarized Light Image Raman Chemical Brightfield Image LCTF Raman Spectra PER Blue = 1608 cm‘1 \A Green = 1629 cm“ PET 'D 0 .t‘ E 22 .9? 39 >2 I l l l l I 15% 1590 1a!) 162) 1640 1%0 Raman Shllt (cm") Figure 4.20: Raman Chemical Imaging with 20X magnification of 25% (wt/wt) PEN blend resins processed through single-screw extruder at 300°C and 1 pass. 115 Brightfield Image Polarized Light Image Raman Chemical Image LCTF Raman Spectra 3 Blue = 1608 cm" g Green = 1629 cm'1 E 25 9 E >3 e , , , , . fl 1580 1600 1620 1640 1660 1680 Raman Stilt (em'l) Figure 4.21: Raman Chemical Imaging with 20X magnification of 25% (wt/wt) PEN blend resins processed through twin-screw extruder at 300°C and 1 pass. 116 300P1single c Figure 4.22: The brightfield (a), polarized light (b), and Raman images (0) from 20x magnification of the blend resins Blue region (PET) = 1608 cm", and green region (PEN) = 1629 cm'1 117 300P1$1ng|e c Figure 4.23: The brightfield (a), polarized light(b), and Raman images (0) from 100x magnification of the blend resins . Blue region (PET) = 1608 cm], and green region (PEN) = 1629 cm'1 118 Figure 4.24: The brightfield (a), polarized light (b), and Raman images (c) from 20X and 100x magnification of the blend films with 25 % (wt/wt) PEN processed through twin-screw extruder at 300°C and 1 pass. Blue region (PET) = 1608 cm", and green region (PEN) = 1629 cm‘1 119 Chapter 5 Barrier and Mechanical Properties of PET/PEN Blends 5.1. Introduction Many foods and beverages were once packed in glass containers or metal cans. However, many of them are now packed in plastic containers due to their light weight, convenience of production, use and disposal, low cost of storage and handling, recyclability, etc. A main conversion has been the replacement of glass by polymers such as PET. However, a limitation of PET applications is its only fair barrier properties, and developing techniques for extending applications for PET has been the subject of recent studies. A modified PET with a higherbarrier performance has been sought for extending product shelf life, and for packaging systems which require high gas barrier performance. PET/PEN blends are an altemative which combines the high barrier and mechanical performance of PEN with the economic advantage of PET. Previous studies show that the oxygen barrier of PEN is approximately four times higher than that of PET (Stewart et a1. 1993) and the tensile strength of PEN is also 35% higher than that of PET (Shi 1998). However, there are few reports on the barrier and mechanical properties of PET/PEN blends. Therefore, in this chapter, these important properties of the PET/PEN will be studied. 120 5.1.1. Barrier Properties The barrier properties are very important parameters in selecting packaging materials for consumer products, especially in the food industry. This is due to the fact that the barrier properties, including water, oxygen, carbon dioxide, and organic vapor barrier, are primary factors controlling the shelf life and the quality of the packaged products. The general theory of gas permeation through a polymer matrix involves two basic parameters, namely the diffusion and solubility coefficients. The diffusion and solubility coefficients are often independent of one another (Nemphos et a1. 1986; Salarne 1967). There are four steps in the gas permeation process, which include solubility of the permeant in the polymer matrix, diffirsion though the polymer matrix, building up of a concentration gradient across the polymer matrix, and desorption of the permeant in the packaging system from the inner wall. The permeability of a polymer in the packaging system is dependent on several internal and external factors. The internal factors include the chemical structure and the . functional groups of the polymer, magnitude of the interchain forces (secondary forces), the packing ability of the chain segments, the degree of crystallization, and orientation (Brennan et al. 1996; Holsti-Miettinen et al. 1995; Nemphos et al. 1986). The external factors include the nature of the permeant, and the environmental condition such as temperature and relative humidity. In this study, the external factors were kept constant, while the internal factors were varied by adjusting the transesterification reactions and the blend composition. 121 When blending PET and PEN, the physical, mechanical, and barrier properties of this blend change, compared to its pure components. As discussed in the previous chapter, the transesterification reactions cause morphological changes. In particular, the average sequence lengths of the terephthalate and naphthalate units decrease, and the number of N-T bonds increases. Consequently, this causes changes in the volume fraction of both the matrix phase (PET) and the dispersed phase (PEN). From the volume fractions of all components, the permeability coefficient of the polymer blend can be predicted (Barrer et al. 1983; Fricke 1924; Holsti-Miettinen et al. 1995; Kit at e1. 1995; Michaels and Bixler 1961; Paul and Buclcnall 1999). Paul and Bucknall (1999) mentioned that the presence of a impermeable dispersed phase in the continuous matrix domain increases the tortuosity of the path a permeant molecule must travel through the polymer matrix. Maxwell proposed an expression for the tortuosity factor (7), which is the effective path length divided by actual thickness of the film. The expression is based on the volume fraction (¢d) of spherical, impermeable or nonconducting particles. The expression is: ¢d TEl+-2— (1) The permeability coefficient of a blend can be calculated as follows (Barrer et a1. 1983; Micheals and Bixler 1961): (2) where P; and P”. are the permeability coefficients of the blend and matrix polymer, respectively, and ¢,,. is the volume fraction of the matrix polymer. 122 Another parameter controlling the barrier properties of the blends is the blend composition. As the PEN composition increases, the barrier properties have been shown to improve. This behavior can be explained as follows. PET and PEN have similar chemical structures. However, PEN has naphthalate rings or double-aromatic rings, while PET has single-aromatic rings. The barrier properties of PEN are better than those of PET since the aromatic rings of naphthalates are stiffer than the single aromatic rings of terephthalates. The high stiffness of the PEN chains results in less segmental mobility of the polymer chains. Therefore, the free volume in the polymer matrix is fixed, which is difficult for a permeant to travel across the polymer matrix. Therefore, by increasing the PEN composition in the blend, the number of naphthalate units in the main chains increased. This results in improvement of the gas barrier properties of the blends (McGee and Jones 1995; Stewart et a1. 1993). However, as discussed in Chapter 4, as a small amount of PEN is added to the PET rich phase, the PEN will interfere with the crystalline formation of the PET. This will reduce the degree of crystallinity of the blends and might decrease the barrier properties of the blends. In this chapter, the effect of the transesterification reactions and PEN composition . on the barrier properties will be discussed. 5.1.2. Mechanical Properties The study of the mechanical properties of polymers is an irnportant area in applied polymer science. By understanding the mechanical behavior and the factors controlling it, polymeric materials can be properly and efficiently utilized. 123 5.1.2.1. Factors controlling mechanical properties The mechanical properties of the polymer blends are dependent on internal and external factors. The internal factors refer to the morphological structure or the chemical nature of the polymer blends. The primary internal factors are the following (Paul and Bucknall 1999; Rudin 1982): (a) for matrix materials (i) molecular weight average (ii) branching, cross-linking, entanglement density (iii) degree of crystallinity (iv) plasticizers, fillers, and additives (v) orientation (vi) other consequences of the processing history, and the thermal history of the polymer (b) for dispersed-phase materials (i) composition (ii) particle size For example, as the molecular weight, entanglement density, and % crystallinity of the polymer increase, the tensile strength and the modulus of elasticity of the polymers increase, while the flexibility or % elongation decrease. The thermal properties are also important factors in determining the brittleness and the ductility of a polymer. T8 relates to the segmental mobility of the molecular chains, and it also determines whether chain-straightening and/or molecular slippage will 124 occur. The chain-straightening refers to the elongation of the molecular chain from the equilibrium position to a new dimension under stress, while molecular slippage refers to the molecule movement past adjacent molecules due to the application of stress (Schwartz 1982). As polymers are placed under stress, they may exhibit both phenomena, depending on their morphological structure, such as branching, cross- linking, degree of crystallinity, etc. At a temperature below T8, those phenomena do not occur appreciably. Therefore, the polymers are hard and brittle. At a temperature above T3, polymers exhibit the chain-straightening and molecular slippage. Therefore, the polymers are soft and flexible (Rudin 1982). The external factors which affect the mechanical properties of polymers include test temperature, annealing, and hydrostatic pressure (Paul and Bucknall 1999; Rudin 1982). As ambient temperature increases, the stress at a given strain always decreases, while the elongation at break increases. As the hydrostatic pressure increases, Young’s modulus, and the yield stress of semicrystalline materials increase (Schultz 1974). Furthermore, the annealing process enhances the strength of semicrystalline polymers (Rudin 1982; Schultz 1974). In this study, all external factors were fixed. The study focused on the effect of the transesterification reactions and the blend composition on the mechanical properties. As discussed above, the morphology of PET/PEN blends depend on the transesterification reactions and the blend composition, and the mechanical properties are dependent on the morphology. 125 5.1.2.2. Tensile properties The tensile test is the most important technique for measuring the mechanical properties of polymers. It determines the force necessary to pull the specimen apart, and the stretch of the deformed plastics before breaking (Schwartz 1982). The general terms used in the tensile test are stress and strain. The stress is the amount of load applied to a unit area, and the strain is the ratio of the extension to original length. The proportional limit is the greatest stress that a material can withstand, without deviating from proportionality of stress to stain (Hooke’s law), and the elastic limit is the greatest stress the polymer membrane can be placed under, without exhibiting permanent deformation (ASTM D638-9land D882-90). A typical stress-strain curve for a polymeric material is shown in Figure 5a. This plot introduces 5 descriptive parameters: 1) tensile strength: the maximum stress that the material can withstand before breaking 2) yield strength: the stress beyond which the material exhibits a deviation from the proportionality of stress to strain, and becomes nonelastic 3) elongation: the extension under stress 4) modulus of elasticity: the ratio of the stress and strain in the interval of linear proportionality. 5) toughness: the total energy absorbed by the material under stress before breaking. 126 Stress Modulus of Elasticity (slope of the linear portion) Tensile Strength ............ ............... .............. ................. ................. .................... .................... ............................. ............................... ..................................... ........................................ .......................................... V ........................................... .......................................... ........................................... .......................................... ............................................ .......................................... ............................................ ........................................... ............................................ ........................................... .............................................. ........................................... .............................................. ............................................... ............................................... ................................................ .............................................. ................................................ ............................................... ................................................ ............................................... Elongation at Break Figure 5a. A typical stress-strain curve for elastic materials 127 In this chapter, the effect of the transesterification reactions and blend composition on the tensile strength, % elongation, and the modulus of elasticity will be described. 5.2. Materials and Methods 5.2.1. Measurement of Water Vapor Transmission Rate The Permatran-W 3/31 developed by the MoCon Co. (7500 Boone Avenue North, Minneapolis, Minnesota 55428) was employed to measure the water vapor transmission rates. The isostatic testing procedure was performed following the ASTM F -1249-90 standard test method (Anonymous 1990 m) for water vapor transmission rate through plastic film and sheeting using a modulated infrared sensor. The water vapor transmission rate tests were performed at 378°C (100°F) and 90% relative humidity (RH). The Permatran-W3/31 consists of two diffusion cells (Cell A and Cell B). Two film samples can be tested at the same time. Each diffusion cell is divided into two chambers. The front chamber behaves as the water vapor environment, and the rear chamber simulates the dry environment (0% RH). The water vapor is generated by introducing the carrier gas (nitrogen) through a water reservoir located next to the diffusion cells. The water vapor is then delivered directly to the front chamber producing a concentration gradient. The level of the relative humidity is controlled by adjusting the pressure of the carrier gas. Dry carrier gas (0 % RH) is circulated through the rear chambers. Film samples are inserted between the front and rear chamber of the diffusion cell. The water vapor traveling from the front chamber is absorbed, diffused through the film wall, and is then desorbed to the rear chamber. 128 The following test parameters were used; The exposed area 50 cm2 The examination time 10 minutes The number of rezero cycles 4 cycles The partial pressure of the water vapor in the front chamber was 44.24 mmHg at 378°C and 90% RH. The rear chamber was dry; thus it was assumed to have partial pressure for water vapor of 0 mmHg. The conditioning period for the non-oriented films (10 mil films) and oriented films (1 mil films) were 5 hours and 1 hour, respectively. During the examination period of 10 minutes, the rear chamber of a diffiision cell is connected to a fixed wavelength infrared sensor, which is interfaced to a computer system. The computer software receives the signal from the infrared sensor, and converts the signal output to the transmission rate value. For each test cycle, the sample in a diffusion cell istested for 10 minutes and the Permatran system will then automatically switch to test the sample in another diffusion cell. The rezero stage is the testing with dry nitrogen or cganier gas for 10 minutes every 4‘h cycle, in order to confirm that no water vapor is contaminating the carrier gas system. The film samples were tested until equilibrium was reached. The water vapor transmission rate values were the average of duplicate samples. The water vapor permeability coefficients were calculated using the following equation: WVTR*l P40 =7 where Pup , WVT R, I, and Ap are the permeability coefficient (g mth m2 mmHg), the (3) water vapor transmission rate (g/h m2), fihn thickness (mil), and the differential pressure 129 or the driving force (mmHg). The water vapor permeability coefficients were later converted to SI units (kg m/s m2 Pa). 5.2.2. Measurement of Oxygen Transmission Rate The oxygen transmission rate was determined with an Oxtran 200 permeability tester (MoCon Co., 7500 Boone Avenue North, Minneapolis, Minnesota 55428). The isostatic testing procedure was performed following the ASTM D3985-81 stande test method (Anonymous 1981) for oxygen gas transmission rate through plastic film and sheeting using a coulometric sensor. The oxygen transmission rate tests were performed at 25°C (70°F) and dry conditions (0 % RH). The Oxtran 200 permeability tester is similar to the Permatran-w3/31, in that the Oxtran 200 system consists of two diffusion cells and two samples can be tested at the same time. Instead of the infrared sensor, a coulometric sensor is connected to the Oxtran 200. In this experiment, compressed air with 21% oxygen was used as the test gas or source of the oxygen, and dry nitrogen was used as the carrier gas. The film samples were placed between the upper and lower chamber of the diffusion cell. The test gas was introduced through the upper chamber, while the carrier gas was purged through the lower chamber. A computer system is connected to the Oxtran 200, therefore one can control the system via the computer. The computer converts the signal strength from the sensor to the oxygen transmission rate. 130 The following test parameters were used; The exposed area 50 cm2 The examination time 20 minutes The number of rezero cycles 4 cycles The driving force was 0.21 atmosphere of oxygen between the upper and the lower chambers. The conditioning period for the non-oriented films (10 mil films) and oriented films (1 mil fihns) were 15 hours and 1 hour, respectively. The film samples were tested until equilibrium was reached. The oxygen transmission rate values were the average of duplicate samples. The oxygen permeability coefficients were calculated using the following equation: (4) where P02, 027‘ R, l, and Ap are the oxygen permeability coefficient (cc mil/h m2 atrn), the oxygen transmission rate (cc/h m2), film thickness (mil),land the differential pressure or driving force (atrn). The oxygen permeability coefficients were later converted to SI units (m3 m/ s m2 Pa and kg m/s m2 Pa). 5.2.3. Measurement of Carbon Dioxide Transmission Rate The carbon dioxide transmission rate was measured using a Permatran CIV permeability tester (MoCon Co., 7500 Boone Avenue north, Minneapolis, Minnesota 55428). The dynamic accumulation technique was used to perform the carbon dioxide 131 transmission rate test. (See the manufacturers manual for the Permatran CIV). The tests were performed at 25°C and dry conditions (0% RH). The Permatran CIV permeability tester consists of five diffusion cells and an infrared sensor. Each diffusion cell is divided into upper and lower chambers. The upper chambers are purged with 100% carbon dioxide (test gas). The lower chambers are purged with nitrogen gas (carrier gas). The lower chambers are connected to the infrared sensor with sensitivity at a particular wavelength that can detect carbon dioxide (C02). The first diffusion cell is used for calibration. The other four diffusion cells are test chambers. A film sample is inserted between the upper and lower chambers of each test cell. The sample is conditioned for a sufficient amount of time to establish a steady state of carbon dioxide transmission rate. The oriented l-mil films were conditioned at 25°C for 24 hours. The thicker film samples were conditioned at the accelerated condition (45°C) for a longer times since these samples required a longer time to reach steady state. The non-oriented 10-mil films were conditioned for 72 hours at 45°C. In addition, the non-oriented 10—mi1 films were kept at the test temperature (25°C) for 24 hours after the conditioning period and before the testing to allow the sample to acclimate to the test environment. When the test begins, the lower chamber of a particular diffusion cell and the sensor are connected and form a recirculating loop. The gradual accumulation of the amount of carbon dioxide permeated through the film samples is monitored by the sensor. The chart recorder displays the resultant signal. As the concentration of C02 increases with time, the corresponding output voltage increases. The carbon dioxide transmission 132 rate can be calculated from the rate of change of the CO2 concentration in the recirculating loop. Each cell was allowed to accumulate C02 in this manner for 15 minutes, before the loop was opened. The sensor was then purged with dry carrier gas for 3 minutes to remove residual C02 from the sensor. The instrument was then prepared for the next test cycle. It should be noted that while a chamber is being tested, the other three chambers can be conditioned at the same time. The calibration cell (the first diffusion cell) is loaded with an aluminum sheet or a very low-permeated film. The very low-permeated fihn will prevent the transfer of C02 from the upper chamber to the lower chamber. During the calibration step, the lower chamber of the calibration cell and the sensor are connected and form a recirculating ' loop. A known amount of C02 is then transferred from a reservoir directly into the loop. The output from the chart recorder corresponds to the known amount of C02, and is used for calibration (the amount of C02 (cc) per response height (cm.) of the chart recorder). The following test parameters were used; The exposed area 50 cm2 The accumulation time 15 minutes The purging time 3 minutes The driving force was 1 atmosphere of C02 at the upper chamber. The samples were tested every 6 or 12 hours until there was no further increase in the transmission rate, which means that the samples had reached steady state. The carbon dioxide transmission rate values were calculated using the following equations (see Figure 5b): 133 pu--¢---- Vd steward (0.0213 cc.) calibration I L RP V J \W—j accumulation period purging period 15 minutes 2 minutes Figure 5b. Example of Permatran CIV graphical data ' 134 ,5 Rv (5) V COZTR = t—j (6) where t stands for the time for 0.0213 cc of carbon dioxide to have permeated. RP and R. stand for the recorder paper advance (cm), and the chart recorder speed (20 cm/h), respectively. CO2T R is the carbon dioxide transmission rate (cc/m2 h). V4 and A are the standard amount of C02 (0.0213 cc) and the test areas (50 cmz), respectively. The carbon dioxide permeability coefficient was calculated using the following equation: C0 TR *1 Pco. = 4A7— (7) where PCO2 , C02T R, I, and Ap are the carbon dioxide permeability coefficient (cc mil/m2 h atrn), the carbon dioxide transmission rate (cc/m2 h), film thickness (mil), and the differential pressure or driving force (1 atrn). The carbon dioxide permeability coefficients were later converted to SI units (m3 m/ s m2 Pa and kg m/s m2 Pa). 5.2.4. Tensile Test Tensile tests were performed using an Instron Universal Tester (Instron Inc., Canton, Massachusetts) following the ASTM D882-90 and ASTM D638-91 standard method (Anonymous 1990‘“); Anonymous 1991). The tensile strength, Young’s modulus, and percent elongation at break were determined. The tests were performed at room temperature (approximately 25 °C), and 50% RH. The tensile test is a technique which involves placing a sample under tensile stress, and monitoring the resistance to the deformation of that sample. This test method 135 is used to measure mechanical properties such as the strength and flexibility of the samples. 5.2.4.1. Test samples All 11 samples of the non-oriented 10-mil films and the oriented l-mil films were tested. Each oriented sample was tested with at least 8 replications. For the non- oriented films, each sample was tested along the machine direction (MD, the direction in which film samples came out from the flat die) and cross machine direction (CMD, the direction perpendicular to the machine direction) with at least 8 replications. All samples were cut to 1X5 inches by a JDC Precision Sample cutter (Thwing-Albert Instrument Company, Philadelphia). The thickness of each sample was measured by a model 549M micrometer (Testing Machines Inc., Amityville, New York). The results showed that thickness of the non-oriented and oriented film samples were 0.01 i 15% and 0.001 1 15% inches, respectively. Therefore, cross sectioned areas of 0.01 and 0.001 in2 were used for non-oriented and oriented samples, respectively. 5.2.4.2. Instron machine The Instron Universal tensile tester is a testing machine which employs a constant rate of crosshead movement. It conSists of a fixed member carrying a grip to hold one side of the sample, and a movable member carrying a second grip to hold the opposite side of the same sample. Samples are held vertically by those grips. A gap between the two grips is very important and is always set to a fixed length. The sample should be aligned perfectly along the pulling direction, and not slip fiom the grips. When the load 136 is applied, the film sample is pulled in the vertical direction, and is stretched under the applied stress. It will deform and eventually break. The following parameters were used for all tests; Load capacity 1000 pounds cross-head speed 20 inches per minute gap length 2 inches. The load and extension profile is recorded by a chart recorder connected to the Instron machine. A typical stress-strain chart is showed in Figure 5a. All the important parameters, including the tensile strength, % elongation, and Young’s modulus of elasticity, can be calculated as follows: tensile strength = maximum load / cross sectioned area (lb/m2) % elongation = the extension/ original gap length *100 modulus = the slope of the linear portion of the stress-strain chart (lb/m2) Tensile strength and modulus were reported in SI units (MPa), using the following unit conversion: PSI (lb/in2) * 6.8948/1000 = MPa 137 5.3. Results and discussion 5.3.1. Barrier Properties The respective permeants, including water vapor, oxygen and carbon dioxide, showed the same trends. The barrier properties were found to be dependent on the PEN composition. In contrast, the degree of transesterification had little or no effect on these properties. The permeability coefficient values were feund to decrease as the PEN composition increased. The relationship between the permeability coefficient values and the PEN composition of the blends appeared to be linear. However, there was no significant difference between the permeability coefficients of the blend fihns with various degrees of tranSesterification. The results are summarized in Tables 5.1-5.6 and presented graphically in Figures 5.1-5.6, respectively. A detailed discussion of the water vapor, oxygen and carbon dioxide permeability coefficient values of the PET/PEN blend fihns is included in the following sections. 5.3.1.1. Water vapor permeability coefficient Degree of transesterification and water vapor permeability coefficient The water vapor permeability coefficient (Pup )values of the blend films with various degrees of transesterification and 25% (wt /wt) PEN are shown in Table 5.1 and Figure 5.1. The Pup of the non-oriented and the oriented blend fihns with 25% (wt/wt) PEN ranged 66m 2.29 x 10'l5 to 2.42 x 10'”, and 1.38 x 101510 1.64 x 10'15 kg m/s m2 Pa, respectively. The degree of transesterification and the processing condition had 138 Table 5.1: The effect of the degree of transesterification on the water vapor permeability coefficient of PET/PEN blend films at 37.8°C. average PHZO x 10''5 (Kg m/ m2 5 Pa) Sample ID PEN wt % B non-oriented films oriented films 2 25 0.076 2.42 1.43 3 25 0.222 2.40 1.52 4 25 0.356 2.42 1.38 5 25 0.192 2.41 1.49 6 25 0.405 2.29 1.43 7 25 0.525 2.35 1.64 "' Blends with 25% PEN (wt/wt) processed through a twin-screw extruder( 150 rpm) 3.00E-15 2505-15 ._ 2005-15 «~ 1505-15» a g: a 1 r: 1.005-15~~ 5005-16 4: The water vapor permeability (kg mlsee-mz-Pa) 0.00E+00 r ‘ % 0 0.1 0.2 0.3 0.4 0.5 The degree of transesterlflcatlon (B) 1 I D non-oriented films 0 oriented fihns j Figure 5.1: The effect of the degree of transesterification on the water vapor permeability coefficient of PET/PEN blend films at 37.8°C 139 little or no effect on the Pup of the blend fihns. There was no statistically significant difference between the Pup values of the blend films with different degrees of transesterification. The details of the statistical analysis of the Pup values as performed are shown in Tables D1 and D2, in Appendix D. PEN composition and water vapor permeability coefficient The Pup values of the PET and PEN non-oriented fihns were 2.96 x 10'15 and 1.15 x 10'15 kg m/s m2 Pa, respectively (Table 5.2). The Pap value of non-oriented PET film was approximately 2.6 times as high as that of the non-oriented PEN fihn. The PHI, values of the oriented PET and PEN fihns were 1.65 x 10.15 and 0.494 x 10'15 kg m/s m2 Pa, respectively (Table 5.2). The P” 10 value of the oriented PET film was approximately 3.3 times as high as that'of the oriented PEN fihn. The lower Pup value of the PEN film is due to the naphthalate rings in the main chains of the PEN. The naphthalate rings cause a high stiffness of the molecular chains, and this results in less segmental mobility and higher barrier properties (Brennan et al., 1996; Holsti-Miettinen et al. 1995; Nemphos et al. 1986). The P” 10 values of the blend fihns processed at 300°C and 1 pass through the extruder, with a PEN concentration ranging from 10 to 40% (wt/wt) PEN, were in the range of 2.76 to 2.12 x 10‘15 kg m/s m2 Pa for non-oriented films, and 1.70 to 1.28kg m/s m2 Pa for oriented films, respectively. There is a statistical difference between 140 Table 5.2: The effect of PEN composition on the water vapor permeability coefficient of the PET/PEN blend films at 37.8°C DI Pnzo x 10"5 (Kg m/ m2 5 Pa) Sample 1 PEN mole% non-oriented films oriented films 1 0.0 2.96 1.65 8 14.5 2.76 1.70 3 33.7 2.40 1.52 9 39.5 2.37 1.26 10 50.4 2.12 1.28 l 1 100.0 1.15 0.494 '1' Blends processed at 300°C and 1 pass through a twin-screw extruder. 3.505-15 3.005-15 g y = -2E-17x + 35-15 D 2 _ g a. 2.505-15 «I R ‘ 0'99“ E 9.- 8} 2.005-15 1» 8 3 «\- g 1' 1505-15 -~ {\0\ b E- n 3 3" 100515 \ P a 4.- - 4- xxx, ; . g y.= -1E-17x + 2E-15 \\ .— 5.00E‘16 “” R2 = 0.9358 9 0.005+00 . r . 0 20 40 60 80 100 PEN mole % F a non-oriented films 0 oriented films Figure 5.2: The effect of PEN composition on the water vapor permeability coefficient of PET/PEN blend films at 37.8°C 141 the P1120 values of blend films with various PEN compositions at a confidence level (p) of 0.05 (see Table D3 in Appendix D). The results showed that the water vapor barrier properties of the PET/PEN blend films improved as the PEN composition increased. The relationship between the PEN composition and the Pflzo values for the non-oriented and oriented films was found to be linear, as shown graphically in Figure 5.2. This behavior can be attributed to the fact that as the‘PEN composition increases, the numbers of the naphthalate units in the main chains of the blend increases, resulting in an increase in chain stiffness and a decrease in the PH20 values. As shown in Table 5.2, the Pflzo values of the oriented films were 50% of those of the non-oriented films. The orientation effect is discussed in Chapter 6. 5.3.1.2. Oxygen permeability coefficient Degree of transesterification and oxygen permeability coefficient The oxygen permeability coefficient (P02 ) values of the blend films with various degrees of transesterification and 25% (wt/wt) PEN are shown in Table 5.3 and presented graphically in Figure 5.3. The P02 of the non-oriented and oriented films with 25% (wt/wt) PEN ranged from 4.99 x 10‘19 to 5.31 x 10'”, and 2.82 x 10'19 to 3.90 x 10'19 kg m/s m2 Pa, respectively. The P02 of these non-oriented blend films and oriented blend films were approximately 20% and 35% lower than those of non-oriented and oriented PET films, respectively. There was no significant difference between the P02 of the blend films with different degree of 142 Table 5.3: The effect of the degree of transesterification on the oxygen permeability coefficient of PET/PEN blend films at 25° C average Po2 x 10"9 (kg 111/1112 5 Pa) P02 x 10''9 (m3 m/mz s Pa) Sample ID B non-oriented films oriented films non-oriented films oriented films 2 0.076 5.22 3.32 3.66 2.33 3 0.222 5.31 2.82 3.72 1.98 4 0.356 5.22 3.55 3.66 2.49 5 0.192 5.11 3.45 3.58 2.42 6 0.405 4.99 3.51 3.50 2.46 7 0.525 5.07 3.90 3.55 2.73 " Blends with 25% PEN (wt/wt) processed through a twin-screw extruder( 150 rpm) 6.E-19 5‘ 5.5-19- a: §§ i i ¥ g .. 8 .. £ 4.5-19 -- , c" 3.5 5 1 ' ' i E ., 35-19 0 5 o 9 a 3 E i‘ 5': 25-19 L 0 e. O S I- 1.5-19- 0.5+00 1 1 1 1 5 0 0.1 0.2 0.3 0.4 . 0.5 The degree of transesterification (B) o non-oriented films 0 oriented films j Figure 5.3 : The effect of the degree of transesterification on the oxygen permeability coefficient of the PET/PEN blend films at 25°C I43 0.6 transesterification. Statistical analysis of the P02 values was performed and the results are summarized in Tables D4 and D5, in Appendix D. Based on the results of statistical analysis, it was concluded that degree of transesterification, or the processing conditions, had little or no effect on the P02 of the blend films. PEN composition and oxygen permeability coefficient The P01 values of the non-oriented PET and PEN films were 6.53 x 10'19 and 3.00 x 10‘19 kg m/s m2 Pa, respectively (Table 5.4). The P02 values of non-oriented PET films were approximately 2.2 times as high as those of the non-oriented PEN films. The Po2 of the oriented PET and PEN films were 4.23 x 10'19 and 8.66 x 10’" kg m/s rn2 Pa, respectively (Table 5.4). The P02 value of oriented PET film was approximately 5 times as high as that of the oriented PEN film. The P02 values of the blend films processed at 300 °C and 1 pass through the extruder, with PEN composition ranging from 10 to 40% (wt/wt) PEN, were in the range of 5.89 x 10'19 to 4.69 x 10"9kg m/s m2 Pa for non-oriented films, and 3.76 x 10"9to 2.96 x 10'19kg m/s in2 Pa for oriented films, respectively. Statistical analysis was performed and the results showed a significant difference between the P02 values of the blend films with various PEN compositions (see Table D6 in Appendix D). The P02 values of the blends fall between the P02 values of PET and PEN films. A linear relationship between the PEN composition and the P02 values of the non-oriented and oriented films was found, as shown in Figure 5.4. The results showed that as the PEN 144 Table 5.4: The effect of PEN composition on the oxygen permeability coefficient of PET/PEN blend films at 25°C Po2 x 10"9 (kg rn/rn2 5 Pa) Po2 x 10"9 (rrr3 m/rn2 s Pa) Sample ID PEN mole% non-oriented films oriented films non-oriented films oriented films 1 0.0 6.53 4.23 4.58 2.96 8 14.5 5.89 3.76 4.13 2.64 3 33.7 5.31 2.82 3.72 1.98 9 39.5 5.06 3.05 3.55 2.14 10 50.4 4.69 2.96 2.29 2.07 11 100.0 3.00 0.87 2.10 0.61 "' Blends processed at 300°C and 1 pass through a twin-screw extruder. 8.E-19 7.5-19 1- y = -3E-21x + 6E-19 R2 = 0.9985 6.E- A CD 1 V a C 2 5 § .. .e a: E “5 5.5-19 1- “ I E 3 3 0 4.5-19 a E. D 5 5 3.5-19 «- 2 ’0‘ 2.5-19 .2 y = -3E-21x + 45-19 '- 1'5'19 “ R2 = 0.9618 7* OE+00 -‘ ¢ 1 : ; Y 1 t i 0 10 20 30 40 50 60 70 80 90 100 PEN mole % D non-oriented films 0 oriented fihns Figure 5.4: The effect of PEN composition on the oxygen permeability coefficient of PET/PEN blend films at 25° C 145 composition increased, the oxygen permeability coefficient of the PET/PEN blend films decreased. Furthermore, the P02 values of the oriented films were 50% of those of the non- oriented films. The orientation effect is discussed in Chapter 6. 5.3.1.3. Carbon dioxide permeability coefficient Degree of transesterification and carbon dioxide permeability coefficient The carbon dioxide permeability coefficient (PCO2 ) values of the blend films with various degrees of transesterification and 25% (wt/wt) PEN are shown in Table 5.5 and presented graphically in Figure 5.5. The Pa,2 values of the non-oriented and oriented films with 25% (wt/wt) PEN ranged from 3.71 x 10'18 to 4.03 x 10'”, and 2.17 x 10'18 to 2.70 x 10'18 kg m/s m2 Pa, respectively. The Pa,2 of these non-oriented blend films and oriented blend fihns were approximately 27% and 30% of those of non-oriented and oriented PET films, respectively. There was no significant difference between the FCC: values of the blend films obtained from various processing conditions. Statistical analysis of the Pa,2 values as a function of processing conditions was performed and the results are summarized in Table D7 and D8, in Appendix D. Therefore, the degree of transesterification, or the processing conditions, had little or no effect on the Pa,2 of the blend films. 146 Table 5.5: The effect of the degree of transesterification on the carbon dioxide permeability coefficient of PET/PEN blend films at 25°C sample In B PC02 x 10'18 (kg ml m2 5 Pa) PC02 x 10'18 (m3 m/ m2 s Pa) non-oriented films oriented films non-oriented films oriented films PET 0 5.49 3.08 2.80 1.57 #2 0.076 3.71 2.52 1.89 1.29 #3 0.222 4.03 2.17 2.06 1.1 1 #4 0.356 3.98 2.41 2.03 1.23 #5 0.192 3.84 2.66 1.96 1.35 #6 0.395 3.71 2.70 1.89 1.38 #7 0.53 3.94 2.45 2.01 1.25 #PEN 0 1.42 0.54 0.73 0.28 * Blends with 25% PEN (wt/wt) processed through a twin-screw extruder( 150 rpm) 5.E-18 “+— 4.E-18 4.5-18 9’ r.“ —l a) 1.E-18 5.E-19 Carbon diox1de permeablllty coefficient (kg.mlm2.sec. Pa) N N (a) m m rp a 66 63 —+——a —k——e —1———1 1——+~1— 0.E+00 9 o 0.1 1 1' 0.2 1 0.3 0.4 0.5 Degree of Transesterification (B) : . 1 D non-onented films 0 oriented films Figure 5.5: The effect of the degree of transesterification on the C02 permeability coefficient of PET/PEN blend films at 25°C 147 PEN composition and carbon dioxide permeability coefficient The Pco, values of the non-oriented PET and PEN films were 5.49 x 10'18 and 1.42 x 10'18 kg m/s m2 Pa, respectively (Table 5.6). The FCC: value of non-oriented PET film was approximately 4 times as high as that of non-oriented PEN film. The Pm: values of the oriented PET and PEN films were 3.08 x 10'18 and 0.54 x 10'18 kg m/s m2 Pa, respectively (Table 5.6). The P092 value of oriented PET film was approximately 5.7 times as high as that of the oriented PEN film. The PCO2 values of the blend films processed at 300°C and 1 pass through the extruder, with the composition ranging from 10 to 40% (wt/wt) of PEN, were in the range 0f4.62 to 3.25 x 10'18 kg m/s rn2 Pa for non-oriented films, and 2.92 to 1.63 x 10"“kg m/s m2 Pa for oriented films, respectively. Statistical analysis was performed and the results showed a significant difference between the Pa,2 values of the blend films with various PEN composition (see Tables D9, appendix D). The Pm2 values of the blends fall between the Pa,2 values of PET and PEN films, respectively. A linear relationship between the PEN composition and the Pco, values of the non-oriented and oriented films was found, as shown in Figure 5.6. The results showed that the barrier properties of the PET/PEN blend films improved as the PEN composition increased. The PCO2 values of the oriented films were approximately 50% of those of the non-oriented films. The orientation effect is discussed in Chapter 6. 148 Table:5.6: The effect of PEN composition on the C02 permeability coefficient of PET/PEN blend filns at 25°C sample 11) pm mol% Pcozx 10‘18 (kg m/ mfs Pa) Pcozx 10'1§(m3 ml m2 s Pa) non-oriented films oriented films non—oriented films oriented films PET 0.0 5.49 3.08 2.80 1.57 #8 14.5 4.62 2.92 2.35 1.49 #3 33.7 4.03 2.17 2.06 1.11 #9 39.5 3.83 1.95 1.95 0.99 #10 50.4 3.25 1.63 1.66 0.83 #PEN 100.0 1.42 0.54 0.73 0.28 * Blends processed at 300°C and 1 pass through a twin-screw extruder. 6.E-18 E! g E 5-5‘13 ‘ y = -4E-20x + 5E-18 g "5 R2 = 0.9938 g . 4.E-18 «- h 0 3 8 g d, 3.E-18 X x .9 r: ‘O C c .92 2.E-18 ~~ o .2 1 0 0 1-E-18 -- y = -3E-20x 1 35-18 R2 = 0.9814 0.5100 1 1 1 1 . 0 20 40 60 80 101 PEN mole °/o D non-oriented films 0 oriented films Figure 5.6: The effect of PEN composition on the C02 permeability coefficient of PET/PEN blend films at 25 °C 149 5.3.2. Mechanical Properties The tensile tests of non-oriented and oriented films were performed at room temperature. The results showed that the non-oriented PEN and PET/PEN films were very brittle, whereas the non-oriented PET films were very flexible. After biaxial orientation, PEN and blend films were more flexible, and the %elongation at break of these samples increased dramatically. There was no relationship between the degree of transesterification and the tensile strength, %elongation, and Young’s modulus of elasticity of the non-oriented, and oriented blend films. However, the blend composition seems to control the tensile strength and %elongation of the non-oriented and oriented films. The Young’s modulus of elasticity of the non-oriented films was independent of the blend composition, while the Young’s modulus of elasticity of the oriented films seems to be slightly influenced by the blend composition. A detailed discussion of the effect of the degree of transesterification and blend composition on the mechanical properties of the non-oriented and oriented films is included in the following section. 5.3.2.1. Tensile strength The tensile strengths of the PET, PEN, and PET/PEN blend films are summarized in Tables 5.7 and 5.8, respectively. The results showed that the tensile strength along the machine direction (MD) and cross machine direction (CMD) of non-oriented PET films were 45.6 and 41.8 MPa, respectively (Table 5.8). The tensile strength of the non- oriented PEN films was 61.5 MPa (Table 5.8). Therefore, the tensile strength of non- 150 Table 5.7: The effect of the transesterification degree (B) on the tensile strength of PET/PEN blend films Tensile strength (MPa) non-oriented -films oriented films Sample ID Bavg MD CMD 0.076 50.89 45.35 153.2 3 0.222 47.33 48.08 182.7 4 0.356 50.27 40.84 170.0 5 0.192 47.05 44.74 148.1 6 0.405 44.89 44.90 126.0 7 ' 0.525 51.42 50.99 148.8 * Blends with 25% PEN (wt/wt) processed through a twin-screw extruder( 150 rpm) 250 . 200 -- (I E E I g db *2 150 § § { g f 3 100 + '- 50 “ W ‘ I I E I I 0 1 1 1 1 1 0 0.1 0.2 0.3 0.4 0.5 - 0.6 Degree of transesterlflcation (B) 1:1 non-oriented films-MD , A non-oriented films-CMD o oriented films Figure 5.7: The effect of the transesterification degree on the tensile strength of PET/PEN blend films 151 Table 5.8 : The effect of PEN composition on the tensile strength of PET/PEN blend films Tensile strength (MPa) non-oriented -fi1ms oriented films Sample ID PEN mole % MD CMD 1 O 45.6 41.8 241.4 8 14.5 42.2 45.5 186.7 3 33.7 47.3 48.1 182.7 9 39.5 50.7 45.3 190.7 10 50.4 53.9 50.3 201.2 11 100 61.5 61.5 269.3 * Blends processed at 300°C and 1 pass through a twin-screw extruder. 120 300 110 oriented films 270 100 y = 0.0245x’ - 2.021411 + 29.25 1; ., ‘ R2 = 0.8908 240 .. n. E 90 a - 210 °- E 1: 80 E E 3 70 180 g 5 60 15 g g .g 0 g . , 55,-- 2 £50 /../' 120 3 '3 c 40 .. 90 = 5 3 30 ~- - 3 1- .. MD CMD .. 50 .2 20 y = 0.187311 1 42.753 y = 0.1917x 1 41.149 10 1 R2 = 0.8943 R2 = 0.9306 30 0 1 1L 1 1 o 0 20 40 60 80 100 PEN mole % 1 L o oriented films 0 non-oriented films-MD A non-oriented films-CMD Figure 5.8: The effect of PEN composition on the tensile strength of PET/PEN blend films 152 of oriented films oriented PEN film was approximately 40-50% higher than those of non-oriented PET film. Shi (1998) also studied the tensile strength of the PET/PEN blends, and reported a similar result, but with a slightly lower percent difference (35%) between the PET and PEN films. The average tensile strength along the machine direction of the non-oriented films was slightly higher that those along the cross machine direction. The orientation process enhanced the tensile strength of the PET, PEN, and blend films. The tensile strengths of the non-oriented PET, PEN and blend films were in the range of 41 to 62 MPa while those of oriented films were in the range of 108 to 269 MPa, (Tables 5.7 and 5.8). Therefore, the tensile strength of the PET, PEN, and blend films improved approximately 4 to 5 times afier orientation. The orientation effect on the tensile strength of blend films will be discussed in Chapter 6. Degree of transesterification and tensile strength The results showed no relationship between the tensile strength of the blend films processed at the various degrees of transesterification (Figure 5.7). This was in agreement with the permeability and the T8 studies, in that these parameters were also independent of the degree of transesterification. Statistical analysis of the tensile strength of the non-oriented and oriented films with the same PEN composition of 25% (wt/wt) of PEN and processed at various conditions was performed as shown in Appendix E. There was no statistically significant difference between the MD tensile strength values of the non-oriented blend films, whereas a significant difference between the CMD tensile strength values of non- oriented blend films was found (Table El and E2 in Appendix E). Moreover, statistical 153 analysis showed a significant difference between the tensile strength values of the oriented films processed at different conditions (Table E3 in Appendix E). PEN composition and tensile strength A relationship between the tensile strength of the blend films and the PEN composition was found. The MD and CMD tensile strength values of the non-oriented films were directly proportional to the PEN composition (Figure 5.8). For the oriented films, the relationship between the PEN composition and the tensile strength was best described by a second order polynomial expression (Figure 5.8). This enhancement of the tensile strength occurred due to the inclusion of a number of the naphthalate groups in the main structure of the blends, resulting in an increase in the stiffness of the polymeric chains and a corresponding increase in the glass transition temperature of the blends, even though, when the PEN composition was increased, a depression of melting temperature and percent crystallinity were found. Above the glass transition temperature, the tensile strength and the modulus is dependent on the percent crystallinity of the polymer, while below the glass transition temperature, the effect of crystallinity on the tensile strength and modulus is relatively small (Paul and Bucknall 1999) Therefore, in this study, the effect of the naphthalate rings seems to be the predominant factor controlling the tensile strength. Moreover, the results showed that the tensile strength values of the non-oriented blend were in the range between those of PET and PEN. Therefore, the tensile strength of the non-oriented blend films improved by adding PEN to PET. However, the tensile strength values of the oriented blend films were lower than those of PET and PEN 154 oriented films. The effect of orientation on the tensile strength of PET and PEN films seems to be stronger than on blend films. A detailed discussion of the orientation effect on the tensile strength of the blends is included in Chapter 6. 5.3.2.2. Percent elongation at break The non-oriented PEN and blend films were very brittle, relative to PET films. The % elongation values of the non-oriented films with 10 to 40 % (wt/wt) PEN and processed at 285 to 315 °C, and l and 2 passes ranged from 2.7 to 3.4 %, and those of PEN were in the range of 4% to 5% (Tables 5.9 and 5.10). The PET non-oriented films had a considerably higher elongation 500 to 600% (Table 5.10). I For the PET and PEN films, the % elongation along the machine direction was greater than that along the cross machine direction. For the non-oriented blend films, the % elongation along both directions appeared to be similar, while non-oriented films typically showed differences in % elongation between the machine and cross machine direction (DeVries et a1. 1977; Park and Mount 1988). This might be due to the very low percent elongation of the blends and limitations in the sensitivity of the Instron machine. After the orientation process, the % elongation of the PET fihns decreased to 213% (Table 5.10). Ray et a1. (1977) studied the mechanical properties of PET films and reported similar results, where the % elongation of oriented PET films (60 to 100%) was lower that that of non-oriented films (250%). The decrease in the % elongation of the PET films might be the outcome of the orientation effect, which in leads to an increase in the % percent crystallinity. As presented in Table 4.8 (Chapter 4), afier the orientation process, the % crystallinity of PET increased. From the literature, 155 Table 5.9: The effect of the transesterification degree (B) on the % elongation of PET/PEN blend films % Elongation non-oriented films oriented films Sample ID Bavg MD CMD 2 0.076 3.4 2.99 148.8 3 0.222 3.03 3.06 130.4 4 0.356 3.5 2.7 149.5 5 0.192 3.0 2.87 148.6 6 0.405 3.0 2.86 127.7 7 0.525 3.0 3.14 140.2 * Blends with 25% PEN (wt/wt) processed through a twin-screw extruder( 150 rpm) 5 300 4.5 -~ v 4 280 a 4 -_ 1- 260 E . c E 3.5 ~- In 0 - 240 O a g 3 .. g i i fl g i «- 220 gg 2.5 «~ ~» 200 E g 2 «~ L 180 ° 5 1.5 —» 1- 160 e 1 1 ~ i } «~ 140 0.5 - } ~~ 120 0 1 4 1 1 a 100 0 0.1 0.2 0.3 0.4 0.5 0.6 Transesterificatlon (B) 1 1 Figure 5.9: The effect of the transesterification degree on the % elongation of PET/PEN blend films 156 [El non-oriented films-MD A non-oriented films-CMD 0 oriented films} %Elongation for oriented films Table 5.10: The effect of PEN composition on the % elongation of PET/PEN blend films %Elongation non-oriented films oriented films Sample ID PEN mole % Nfl) CMD l 0 604.00 506.43 213.1 8 14.5 3.28 3.0 181.6 3 33.7 3.03 3.06 130.4 9 39.5 3.08 2.9 144.1 10 50.4 3.28 3.26 134.6 11 100 5.33 4.07 107.8 * Blends processed at 300°C and 1 pass through a twin-screw extruder. 1 a 14 lb oriented films 1_ 200 g .. y=0.0135x’-2.3691x+211.27 g 12 1’ R2=0.9489 c ”D 1— 150 § 10 fl y=0.0005x2-0.0341x+3.6487 é 8 I R2=0.9984 o «1. E CMD 100 .2 6 1 y=0.0002:8-0.0054x+3.0041 1= 5 R2=0.9456 C 2 m 2 ._ a! 0 1 1 ; 1 o 0 20 40 60 80 100 PEN mole% D non-oriented films-MD A non-oriented films-CMD <> oriented films Figure 5.10: The effect of PEN composition on the % elongation of PET/PEN blend films 157 %Elongatlon for oriented films crystallinity seems to reduce the ability of chain-straightening, and molecular slippage, which corresponds to a decease in the extensibility of the polymer chains (Rudin 1982; Schultz 1974; Schwartz 1982). In contrast to the PET films, the % elongation of the oriented PEN and blend films improved tremendously. The elongation value of the oriented PEN film was 108% and those of oriented blend films were in the range of 128 to 182 %, respectively (Table 5.9 and 5.10). The results indicated that the PET, PEN and blends exhibited different stress-strain behaviors, when they were oriented under the following orientation conditions: a stretch temperature of 100°C for PET and blend samples, 135°C for the PEN films, and a soak time of 15 seconds for all samples. Moreover, the results fiom Chapter 4 (Table 4.8) showed that following the orientation process, the % crystallinity of PEN films remained the same, and those of the blends increased slightly, relative to that of PET. This suggests that PET, PEN, and blend films have a different mechanism of strain-induced crystallization. This mechanism involves the promotion of chain- straightening, and molecular slippage (Jenkins 1995; Shi 1998). The details of the effects of orientation on the % elongation of the PET, PEN and blend films are presented in Chapter 6. Degree of transesterification and percent elongation at break The results showed no relationship between the % elongation of the blend films with various degrees of transesterification (Figure 5.9). These results are in agreement with the tensile strength study and can be explained as mentioned previously. 158 Statistical analysis of the % elongation of the blend films with the same PEN composition of 25% (wt/wt) of PEN and processed at the various conditions was performed, as shown in Appendix F. There was no statistically significant difference in the % elongation along the machine direction of the non-oriented blend films, whereas a significant difference in the % elongation along the cross machine direction of the non- oriented films was found (Table F1 and F2 in Appendix F). Additionally, statistical analysis showed no significant different in the % elongation of the oriented films processed at different conditions (Table F3 in Appendix F). PEN composition and percent elongation at break The results showed that there was a relationship between the % elongation of the film samples and the PEN composition (Table 5.10 and Figure 5.10). For the non- oriented films, the % elongation appeared to increase slightly as the PEN composition increased. Due to the very low range of the % elongation of these samples, it is recommended that the experiment be repeated to confirm the results. However, these results seem. to be reasonable. The results are in agreement with the effectof the PEN composition on % crystallinity, density, and molecular weight average, which are the primary factors controlling the elongation of the polymers. These properties decreased as the PEN composition increased (Chapter 4). The reduction in % crystallinity, density, and molecular weight average promote the flexibility of the polymers (Paul and Bucknall 1999; Rudin 1982; Schultz 1974; Schwartz 1982). In contrast to the non-oriented films, the % elongation of the oriented films decreased as the PEN composition increased. This might be due to the fact that the 159 orientation effect outweighs the effect of the PEN composition, and results in a decrease in the ability for extension under the application of stress. 5.3.2.3. Young’s modulus of elasticity Young’s modulus of elasticity values for the PET, PEN, and PET/PEN blend films are shown in Tables 5.11 and 5.12. The results showed that the modulus of non- oriented PET, PEN, and blend films were in the range of 1554 to 1739 MPa (Tables 5.11, and 5.12), respectively. There was no statistically significant different between the modulus of those samples (Figures 5.11 and 5.12). However, after orientation, the modulus of PET, PEN, and blend films increased (Table 5.12 and Figure 5.12). The average moduli of oriented PEN films (2574 MPa) was slightly higher than that of oriented PET films (2304 MPa). However, the moduli of oriented blend films were lower than either of PET or PEN oriented films. Degree of transesterification and Young’s modulus of elasticity The results showed no relationship between the modulus of elasticity of the blend films with various degrees of transesterification (Figure 5.11). These results were in agreement with the barrier studies and the tensile and elongation results, in that those properties of the blend films did not depend on the degree of transesterification. Statistical analysis was performed and showed that there was no significant difference between the moduli of the blend films with 25% (wt/wt) of PEN, processed 160 Table 5.11: The effect of the transesterification degree (B) on the modulus of elasticity of PET/PEN blend films Young's Modulus (MPa) non-oriented films oriented films Sample ID 331% Nfl) CMD 2 0.076 1569.95 1549.95 1785.8 3 0.222 1718.18 1618.90 1823.6 4 0.356 1714.74 1560.98 1679.5 5 0.192 1688.54 1549.95 1693.9 6 0.405 1627.86 1520.30 1697.8 7 0.525 1739.07 1586.49 1673.6 * Blends with 25% PEN (wt/wt) processed through a twin-screw extruder (150 rpm) The modulus of elasticity (MPa) 2200 2000 + 1800 .1 1600 1- 1400 .- 1200 1- 1000 0.1 0.2 0.3 0.4 Transesterificatlon (B) 0.5 i U non-oriented films-MD A non-oriented films-CMD o oriented filmsj Figure 5.11: The effect of the transesterification degree on the modulus of elasticity of PET/PEN blend films 161 0.6 Table 5.12: The effect of PEN composition on the modulus of elasticity of PET/PEN blend films Young's Modulus (MPa) non-oriented films oriented films Sample ID PEN mole % MD CMD 1 0 1648 155.4 2304.3 8 14.5 1684 1653 1932.4 3 33.7 1718 1619 1823.6 9 39.5 1556 1617 1901.1 10 50.4 1692 1532 2049.2 11 100 1619 1685 2573.5 * Blends processed at 300°C and 1 pass through a twin-screw extruder. 3000 a“! 2500 E .é‘ o 2000 '5 ll) 2 .2 1500 O n 2 g 1000 .. E oriented films 53 500 1- y = 0.1958x2 - 15.874x + 2223.1 R2 = 0.8993 0 r $ 1 1r 0 20 40 60 80 100 . PEN mole % — a — non-oriented films-MD ----A--- non-oriented filrns-CMD _ o oriented films 1 Figure 5.12 : The effect of PEN composition on the modulus of elasticity of PET/PEN blend films 162 under various conditions of temperatures and blending times (Table G1, G2 and G3, in Appendix G). PEN composition and Young’s modulus of elasticity There was no relationship between the MD and CMD moduli of the non-oriented blend films, and the PEN composition. Statistical analysis showed (Appendix G) no significant difference in MD moduli of the non-oriented blend films with different compositions, whereas statistically significant differences between the CMD moduli of non-oriented films and PEN composition were found (Tables G4 and G5, in Appendix G). The significant difference between the CMD modulus values for the non-oriented films might be due in part to the low sensitivity of the Instron machine, and may also , reflect a slight effect of the PEN composition on the modulus of elasticity of non-oriented films. However, a relationship between the modulus values of the oriented films and the PEN composition was found, and was best described by a second order polynomial expression (Figure 5.12). Statistical analysis showed a significant difference between these samples as well (Table G6 in Appendix G). This relationship was similar to the relationship found between the tensile strength and the PEN composition. A non- equivalent effect of the orientation process on PET, PEN and blends was found, in that the orientation effect on the modulus of PET and PEN films seems to be greater than for the blend films. Thus, the modulus of the oriented blend films was lower than those of PET, and PEN oriented films. A detailed discussion of the orientation effect on the modulus of elasticity is included in Chapter 6. 163 Chapter 6 The Orientation Effect 6.1. Introduction Orientation is an important process to enhance the performance of fabricated polymers. Most commercial semicrystalline polymers are partially oriented in order to yield advantageous properties, e. g., mechanical strength, barrier properties, and dielectric polarizability (Peterlin 1988). Products with designed-in orientation, including fabricated sheets and filaments with uniaxial orientation, and films and bottles with biaxial orientation, are increasingly important in the polymer industry (White 1988). Therefore, in-line processes for uniaxial and biaxial orientation have been developed for a variety of molding operations. Molecular orientation generally refers to the alignment and rearrangement of segments of polymeric chains by an application of an external stress at a particular temperature. Hence, it involves some degree of stretching and heat setting. In general, the orientation process is a rapid stretching of a polymer at a temperature above the T‘ for amorphous polymers and at a temperature between the T8 and the Tm for semicrystalline polymers, and a subsequent quenching of the polymer (Park and Mount 1988). 164 6.1.1. Uniaxial Orientation Uniaxial orientation promotes high performance of the polymers in one direction, usually along the machine direction. Melt polymer or extrudate is released from a flat die and passed through a set of chill rolls which force the hot plastic to stretch under a controlled temperature. This film fabrication process produces some degree of uniaxial orientation. In addition, the quenching rate is controlled to ensure that the orientation is not lost by molecular relaxation (Park and Mount 1988) Uniaxial orientation along the cross machine direction can be obtained using the tenter frame or melt inflation. However, these two techniques are seldom employed (Park and Mount 1988). The relationship between physical properties and uniaxial orientation of the polymer has been studied. As the orientation occurs, the c-axis of the molecular chains in the amorphous and crystalline phases of the semicrystalline polymer is aligned in the orientation direction. This alignment of the c-axis chains allows the strong covalent bond along the chain backbone to carry the load applied in the orientation direction. In the perpendicular direction, the weaker intermolecular Van der Waals forces are predominant, which can only carry a lower load (DeVries et al. 1977; Hoshino et. al. 1962; Seferis and Samuels 1979). Therefore, uniaxial orientation promotes the mechanical properties only along the orientation direction. 6.1.2. Biaxial Orientation Biaxial orientation substantially improves the physical properties of polymers in both the machine and the cross machine directions. Commercial equipment for biaxial orientation includes tenter-frame, double-bubble, and blown film machines (Park and 165 Mount 1988). In the tenter-frame and double-bubble process, polymers are biaxially oriented in the solid state at a temperature below the crystalline melting point. In the blown film process, polymers are oriented in the melt state and are rapidly quenched (DeVries et al. 1977; Middleman, S. 1977; Samuels 1974). Biaxially oriented films achieve superior improvement in physical properties compared to uniaxially oriented films. This superior improvement is due to the redistribution of the c-axis chains by the biaxial orientation (DeVries et a1. 1977). 6.1.3. Effect of Orientation on Polymer Properties Orientation influences the morphological, physical and thermal properties of polymers, and consequently improves the mechanical, barrier, optical, and electrical ' properties. 6.1.3.1. Percent crystallinity Strain-induced crystallization is one of the results of the orientation (White 1988). Orientation tends to enhance. crystallinity by bringing the long axis of the molecular chains close together and parallel to each other (Rudin 1982). The amount of crystallization induced by the orientation process is controlled by the stretch rate, stretch temperature, and the stretch ratio (J abarin 1992). However, the difference between orientation and percent crystallinity is that crystallinity requires a regular placement of molecular chains, whereas orientation requires only the alignment of molecular chains, regardless of the location of the 166 molecules (Rudin 1982). Therefore, orientation can occur in both amorphous and semicrystalline polymers. 6.1.3.2. Barrier properties In general, orientation promotes an enhancement of the barrier properties of glassy polymers (El-Hibri and Paul 1985; O’Brien et al. 1987). This improvement in the barrier properties of semicrystalline polymers is due to the rearrangement of the larnellar crystalline domains, in a direction perpendicular to the direction of the penetrant flow. In addition, simple crystallinity acts as a tortuous blocking which lengthens the path of penetrants in their diffusion across the film bulk phase. This behavior, therefore, leads to a significant reduction of the gas permeability of polymeric films. Furthermore, the effect of orientation on highly crystalline materials is stronger than that on an amorphous or less crystalline material (Brady et a1. 1974; Koros and Hellums 1988). O’Brien et a1. (1987) studied the effect of orientation on the C02 and CH4 permeability of uniaxially and biaxially oriented polyirnide samples of pyromellitic dianhydride and oxydianiline. Biaxial orientation showed a stronger effect on the solubility, diffusivity, and permeability, than that of the uniaxial orientation. El-Hibri and Paul (1985) also agreed that orientation produces a significant reduction in solubility and diffusivity, resulting in a decrease in the permeability of polymeric materials. 6.3.1.3. Mechanical properties Uniaxial orientation strongly affects mechanical properties, in that the tensile strength and the Young’s Modulus of elasticity increase whereas the elongation at break 167 decreases (Nadella et a1. 1978). Therefore, uniaxially oriented fihns are brittle in the cross machine direction (Matsumoto et a1. 1981). In contrast to uniaxial orientation, biaxial orientation enhances the tensile strength, Young’s modulus, and elongation at break, in all directions in the plane of the films (White 1988). For instance, biaxially oriented bottles exhibit high mechanical properties in both the bottle-axis direction and the hoop direction. However, this relationship between the tensile strength and Young’s modulus, and orientation are only found at a low degree of orientation. When crystallites are completely oriented, further stretching enhances the tensile strength and Young’s modulus (Nadella et al. 1978; White et al. 1974; Samuels 1974). Amorphous and semicrystalline polymers behave differently, when uniaxially. - oriented. Semicrystalline polymers exhibit necking or an abrupt change of the thickness. In contrast, amorphous polymers can be stretched to any thickness, if properly supported (Samuels 1974). 6.1.3.4. Effect of orientation on selected properties of PETIPEN blends A difference between the stress-strain behavior of PET and PEN was found (Cakrnak et 31.1996; Murakami et al. 1996; Shi 1998) PEN shows neck formation, when it is uniaxailly oriented at temperature between T8 and he (cold crystallization temperature), at a stretch ratio of less than 5 mm/min (Shi 1998). However, for PET the neck behavior usually occurs at T8 or below. Therefore, a difference in the mechanism and the temperature dependency of strain-induced crystallization of PET and PEN was considered (Shi 1998). 168 Bauer (1997) showed that the density of biaxially oriented films of PET/PEN blends increased with an increase in the stretching ratio (2.5x2.5, 3.0x3.0 and 3.5x3.5) and heat setting temperature (120, 150 and 180°C). The increased density indicates an increase in the level of crystallinity. Oxygen permeability also decreased as the stretching ratio or orientation temperature was increased. The oxygen permeability of the 3.5x3.5 biaxially oriented films of a polymer blend with 25% wt/wt of PEN was about 40% lower than the corresponding PET films. Hoffman and Caldwell (1995) studied the oxygen permeability of PET/PEN blends and copolymers of compositions varying from 0-100 weight percent PEN, and found that the permeability coefficient decreased to approximately half when fihns were 4x4 biaxially oriented at 20°C above T8. In addition, Bauer (1997) mentioned that for pre-heated samples, the density was independent of the stretching ratio. At the same heat set temperature but different draw ratios, samples reached almost the same range of density. In this chapter, the effect of biaxial orientation on selected properties of the PET, PEN and PET /PEN blends will be discussed. A comparison of the orientation effect on the blend films with various degree of transesterification and various compositions will be discussed. 6.2. Materials and Methods The eleven PET, PEN, and PET /PEN blend films (IO-mil, non-oriented films) were biaxially oriented on a T.M. Long film stretcher. All of the samples were stretched 3x3 simultaneously in both directions. The rate setting was 300% per second. The stretch temperature was 100°C for all samples, except for the PEN films, which were 169 stretched at 135 °C. Twelve biaxially oriented film specimens were prepared from each sample. The final oriented film thickness was approximately 1 mil (0.001 inch). The details of the orientation process are discussed in Section 4.2.2. The thermal properties of the oriented blend films were determined using a DSC 2920 Modulated Differential Scanning Calorimetry (MDSC). The glass transition temperatures (T8), melting temperatures (Tm), and the % crystallinity were determined. The details of the thermal analysis procedure are presented in Section 4.2.3. 1. The density gradient technique was used to determine the density of the non- oriented and oriented fihns (Section 4.2.3.2). The Permatran-W 3/31 was employed to measure the water vapor transmission rates. The oxygen and carbon dioxide permeability values were determined with an Oxtran 200 permeability tester and the Permatran CIV permeability tester, respectively. All procedures had followed ASTM standards, where applicable (Sections 5.2.1, 5.2.2, and 5.2.3). . Tensile tests were performed using an Instron Universal Tester, following the ASTM standard methods (Section 5.2.4). Tensile strength, Young’s modulus, and % elongation at break were determined. 6.3. Results and Discussion The T8, Tm, and % crystallinity of all non-oriented and oriented film samples were reported in the Results and Discussion section of Chapter 4 (Section 4.3.1). The density of all samples was presented in Section 4.3.2. The mechanical and barrier properties of 170 the oriented and non-oriented film samples were summarized in Sections 5.3.1 and 5.3.2, respectively. To facilitate the discussion, the term “orientation-induced increase (011)” of a particular numerical property of a polymer has been defined to be the difference between the value of that property of a 3x3 biaxially oriented polymer and that for the non- oriented one. For instance, the tensile strength of 3x3 biaxially oriented PET film was 45.6 MPa, and the tensile strength of the non-oriented PET film was 241.4 MPa (Table 5.8). Therefore, the 011 of the tensile strength of this PET film was 241.4 - 45.6 = 195.8 MPa. The result showed that orientation affected the T3, % crystallinity, density, barrier properties, and mechanical properties of both PET and PEN, as well as their respective blends. However, orientation seems to have less of an effect on the Tm of the blends, compared to the effects of degree of transesterification and PEN composition. For the PET films, the orientation-induced increase (011) of all properties, except for the Tm and the Young’s modulus, was higher than that of PEN and blend films. In PET rich phase systems (PEN composition lower than 50 mole % PEN), the 011 of all properties, except for Tm and Young’s modulus, were higher than that of blends with a higher PEN composition. There was no significant difference of the 011 of all properties (except for the % crystallinity and the density) of the blends with different degrees of transesterification. However, the 011 of the % crystallinity and density of the blends slightly decreased as the degree of transesterification increased. 171 Therefore, the degree of transesterification had less effect on the 011 of most properties of the blends than PEN composition. The effect of OH on particular properties of the PET, PEN and blend films is discussed in the next section. 6.3.1. Effect of Orientation on Thermal Properties 6.3.1.1. Glass transition temperature The results showed that the T, of the PET and the blend films increased about 13 °C after 3x3 biaxial orientation. In contrast, the T, of the PEN film decreased 6°C after it was oriented (Table 4.3 and Figure 4.3). For the blends with different degrees of transesterification, the T, increased between 2-5°C after orientation (Table 4.2 and Figure 4.2). Moreover, the T, values of oriented blend films were approximately the same for different degrees of transesterification. These blend films had a composition ranging from 0'to 50.4 mole % PEN. No experiments for the blends with PEN compositions of more than 50.4 mole % PEN were performed. Therefore, the predicted T, values of blend films with high PEN compositions were obtained by best fitting the data with a mathematical expression. The relationship between the composition of the oriented blend films and the T, values was found to be a second order polynomial, while that of the non-oriented blend films and the T, values was linear. The result showed that in PET rich phase systems (PEN composition lower than 50 mole % PEN), the 011 of the Tg of the blend with the lower PEN composition was higher than the 011 of blend with the higher PEN composition (Table 4.3 and Figure 4.3). 172 6.3.1.2. Melting temperature The results showed that orientation had little effect on the Tm of the film samples (Tables 4.5 and 4.6, and Figures 4.5 and 4.6). The Tm of the PET and the blend films decreased slightly after orientation. The Tm of PEN films did not change after orientation. 6.3.1.3. Percent crystallinity After orientation, the % crystallinity of the PET films increased fi‘om 33% to 41% while the % crystallinity of the PEN films did not change (Table 4.8 and Figure 4.8). For the blends, regardless of the degree of transesterification, the % crystallinity increased slightly afier. orientation (Table 4.8 and Figure 4.8). The result showed that in PET rich phase systems, the 011 of the % crystallinity of blends with a lower PEN composition was higher than that of blends with a higher PEN composition (Table 4.8 and Figure 4.8). In the PEN rich phase systems, the % crystallinity did not change afier the blends were oriented. The difference in the orientation effect on PET, PEN and the blends is due to the difference in the mechanism of strain-induced crystallization and the temperature dependency of the strain-induced crystallization for the pure components and the blends (Jenkins 1995; Shi 1998). 6.3.2. Effect of Orientation on Density of the Blends Orientation slightly but predictably affected the density of the PET, PEN and blend films (Tables 4.10 and 4.12, and Figures 4.10 and 4.12). The 011 of the density of blends with lower degees of transesterification was higher than that of the ones with 173 higher degrees of transesterification (Table 4.10 and Figure 4.10). Moreover, the 011 of the density of the PET fihn was higher than both those of PEN and blend films. In PET rich phase systems, the 011 of the density of the blends with lower PEN composition was higher than those with higher PEN compositions (Table 4.12 and Figure 4.12). 6.3.3. Effect of Orientation on Barrier Properties The water vapor, oxygen, and carbon dioxide barrier properties of PET, PEN and blend films improved considerably when these films were 3x3 biaxially oriented (Tables 5.1 to 5.6, and Figures 5.1 to 5.6). The results showed no significant difference in the permeability coefficients of the oriented blend films with different degree of transesterification. The 011 of the permeability coefficients of PET film were higher than those of PEN and blend films. In addition, the OH in the permeability coefficients of the blends with a lower PEN composition were higher than those with a higher PEN composition. 6.3.4. Effect of Orientation on Mechanical Properties 6.3.4.1. Tensile strength The tensile strength of PET, PEN, and PET/PEN blends significantly increased after the films were biaxially oriented (Tables 5.7 and 5.8, and Figures 5.7 and 5.8). The OH of the tensile strength of the blends, regardless of the degree of transesterification, was approximately the same (Table 5.7 and Figure 5.7). The 011 of the tensile strength of the PET and the PEN films were higher than those of the blend films. A linear relationship between the PEN composition and the tensile strength of the non-oriented 174 film was found. However, for the oriented films, a polynomial expression was found to best describe the relationship between the PEN composition and tensile strength. Moreover, there was a slight difference of the OH of the tensile strength of the blend films with different PEN compositions (Figure 5.8). 6.3.4.2. Percent elongation at break The % elongation at break of the PEN and the PET/PEN blend films significantly increased as the films were biaxially oriented (Tables 5.9 and 5.10, and Figures 5.9 and 5.10). However, the % elongation at break of the PET films decreased after the PET films were oriented (Table 5.10). The OH in the % elongation of the blends, regardless of the degree of transesterification, was approximately the same. The OH in the % elongation of the blends decreased as the PEN composition increased (Table 5.10 and Figure 5.10). Moreover, the OH in the elongation of the blend films were higher than that of the PEN film. 6.3.4.3. Young’s modulus The Young’s modulus of elasticity of the PET and the PEN films improved dramatically afier the films were 3x3 biaxially oriented. The OH of the modulus of the PEN film was higher than that of the PET film (Table 5.12 and Figure 5.12). The Young’s modulus of the blend films increased slightly after they were oriented (Tables 5.11 and 5.12, and Figures 5.11 and 5.12). Both degree of transesterification and blend composition had a little or no influence on the OH of the modulus of the blends. 17S 6.4. Summary The degree of transesterification does not influence the orientation-induced increase (OH) of most blend properties. After blends achieved a certain degree of transesterification, the blend properties do not depend on the degree of transesterification. However, blend composition is a major factor influencing the OH of most blend properties. The 011 of PET was greater than that of PEN and the blends. In addition, the OH of the blends with the lower PEN composition were greater than that of the higher PEN composition. I This might be due to the difference between the temperature dependency and the mechanism of orientation-induced crystallization of PET and PEN. In this study, PET, PEN and blends were 3x3 biaxailly oriented under the same stretch rate but different stretch temperatures. The stretch temperature for PET and the respective blends was 100°C, and for PEN was 135°C. The stretch temperatures of PEN and blends might not be sufficient or equivalent to that of PET. Moreover, the stretch rates for PET, PEN and blends to achieve sufficient orientation results might be different. Shi (1998) reported similar results. PET and PEN exhibited different stress- strain behaviors when PET and PEN were uniaxially oriented under equivalent ‘ conditions, at a temperature of 15°C above the polymer T, (Shi 1998). Therefore, future investigation of the orientation conditions for PET, PEN and blends to achieve the maximum orientation effect is recommended. 176 Chapter 7 Summary and Conclusions 7.1. Transesterification Reactions PET and PEN are inherently immiscible. The mixture of PET and PEN yields an opaque material with non-uniform properties, and shows two T, and Tm values corresponding to one PET-rich phase and one PEN-rich phase. Transesterification reactions are interchange reactions between hydroxyl and carboxyl groups of the PET and PEN chains. T ransesterification reactions are accomplished by melt mixing these two polymers at a temperature which is higher than their melting temperatures. The transesterification reactions enhance the miscibility of the blend, and cause a reduction in size of the PET and PEN chains. Morphological analysis confirmed that blends which achieved a minimal degree of transesterification exhibited sample homogeneity, at the spatial resolving power of Dispersive Raman spectroscopy technique (approximately 200 nm), while the blends without a minimal degree of transesterification showed phase separation. The properties and composition of the blends change when transesterification is achieved. During this reaction processing, graft and block copolymersvare formed. The PET/PEN blends with a transesterification degree of at least 6% showed a significant improvement in the blend miscibility and had only a single T, and Tm. These blends were optically clear and uniform. However, the DSC analysis of the blends showed single, broad Tm peaks, suggesting the occurrence of partial miscibility. 177 The processing conditions including the blending temperature and the blending time are primary factors controlling the degree of transesterification, while blend composition had little or no effect. There was a linear relationship between the degree of transesterification and the blending temperature. Moreover, the degree of transesterification increased as the blending time (number of pass) increased. However, there was no relationship between the degree of transesterification of the blends and the blend composition. In the range of the processing conditions used in this study, the optimum blending temperature was found to be between 285 and 315°C. The PET/PEN blend resins obtained fiom reaction processing at a temperature of 275°C did not appear to be uniform and were mixed between clear and hazy zones. According to DSC and lHNMR analysis, a blending temperature of 27 5°C is the minimum limit to achieve a measurable degree of transesterification. In addition, the PET/PEN blend resin obtained fi'om reaction processing at 325 °C was brittle and had low melt strength. Degradation might occur during processing at this temperature, or at temperatures greater than 325°C. In addition, the D value was proposed to represent the degree of transesterification instead of the existing B-value. The D value is more meaningful than the B-value. However, in this dissertation the B-value is used to represent the degree of transesterification, so that the results can be compared with other studies. 178 7.2. Blend Characteristics Table 7.1 shows a summary of the effects of transesterification and blend composition on the blend characteristics including: (i) thermal properties, (ii) % crystallinity, (iii) density, (iv) molecular weight, (v) banier properties, and (vi) mechanical properties. 7.2.1. Effect of Degree of Transesterification The PET/PEN blends with a degree of transesterification of at least 6% were miscible, homogeneous and optically clear. ’ Further, the transesterification reactions had little effect on the T,, the banier, and the mechanical properties. However, the Tm, % crystallinity, density, and molecular weight average were dependent on the degree of transesterification, regardless of how this degree of transesterification was achieved. The "1’m , % crystallinity, and density of the blends decreased linearly, as the degree of transesterification increased. This is due to the fact that the transesterification reactions lead to the formation of terephthalate-naphthalate (N -T) bonds, which might interfere with crystal formation. The T, did not change as the degree of transesterification increased. This is in agreement with the water vapor, oxygen, and carbon dioxide barrier properties and the mechanical properties, including tensile strength, % elongation at break, and Young’s modulus of elasticity, in that they were independent of the degree of transesterification. 179 Table 7.1: Summary of the effect of the degree of transesterification and blend composition on selected properties of the PET/PEN blends. The effect of degree of The effect of degree of PEN transesterification (B) composition (mole % PEN) Resins Films Oriented Resins Films Oriented films films T, N N N L+ L+ P+ Tm L- L- L- U U U %Crystallinity L- L- L- U U U Density L- L- L— n/a L— P- Permeability N/a N N n/a L- L- PH,01P0,,1PCO, Tensile strength N/a N N n/a L+ U %Elongation N/a N N n/a n/a P- at break Young’s N/a N N n/a N U modulus L linear relationship N no relationship P second order polynomial expression U parabola expression + the preperty increases as B or mole % PEN increases the property decreases as B or mole % PEN increases 180 7.2.2. Effect of Blend Composition The blend composition is an important parameter which controls all the characteristics and pr0perties of the blends. A depression of the T,,, , % crystallinity, and molecular weight was found, when a small amount of PEN was added to the PET rich phase, or vice versa. The Tm , % crystallinity and molecular weight average were lowest when the PEN composition was 40-50 mole %. Moreover, the density of the blends decreased, as the PEN composition increased. The depression of these properties is a consequence of the disruption of the crystalline structure, as a small amount of one polymer is added to another. However, the decrease in Tm , % crystallinity, molecular weight and density seems to have a less influence on the barrier and mechanical properties of the blends. Since the blends were only partially miscible, characteristics of PET and PEN are still exhibited by the blends. When the PEN composition increased, the T,, the gas barrier properties, and the tensile strength of the blends improved. Increase of the number of naphthalate units leads to a decrease in the segmental mobility of the molecular chains, resulting in higher chain stiffness and improvement in gas barrier properties and tensile strength. The relationship between the blend composition and the above properties can be expressed as a series of mathematical equations as shown in Table 7.2. Therefore, the properties of the blends with different blend compositions can be predicted using those equations. However, it should be pointed out that the equations presented apply only to the blends obtained by the processing conditions used in this study. 181 Table 7.2: The equations showing the relations between the blend composition and the barrier and mechanical properties of the blends obtained at 300°C and 1 pass. Properties Equation (C = PEN composition, % mole of PEN) Non-oriented films Oriented films Barrier properties Pup =-2E-17C+3E-15 Pup =-lE-17C+ 213-15 1. Water vapor permeability coefficient, P ”20 :: - .. + - = - - - 2. OXYgen permeability P°= 3E 21C 65 19 P0. 3E 21C+ 4E 19 coefficient, P0, 3- Carbon dioxide Pa, = 415200 + 513.1 8 Pa, = -3E-20C + 312-1 8 permeability coefficient, P €02 Mechanical Properties 1, Tensile strength, 7, rsavm) = 0.1873C +42.753 Ts = 0.0245C’-2.0214C+229.25 Ts(CMD) = 0.1917C +4l.149 2. %elongation at break, 5,0110) = 0.000502 —0.034rc E.= 0.013502 -2.3591c +211.27 E +3.6487 B E,(CMD) =0.0002C’ —0.0054C _ +3.0041 N/A _ g=0.1958C" —15.874C+2213.l 3. Young’s modulus of elastics, a The unit of permeability coefficients-is kg ,m / s m2 Pa. The unit of the tensile strength and the Young’s modulus is kPa. 182 7.2.3. Effect of Orientation Orientation is a significant process employed to enhance the barrier and mechanical properties of polymer blends. The water vapor, oxygen, and carbon dioxide barrier properties of the PET/PEN blend films improved 1.5-2 times, when the blend films were 3x3 biaxially oriented. The tensile strength of the oriented blend films was 4 times greater than that of the non-oriented blend films. Moreover, oriented PEN, and the oriented blend films were flexible, while the non-oriented films were very brittle. To facilitate the discussion, the term “orientation-induced increase (011)” of a particular numerical property of a polymer has been defined to be the difference between the value of that property for a 3x3 biaxially oriented polymer and that of the non- oriented one. For instance, the tensile strength of 3x3 biaxially oriented PET film was 241.4 MPa, and the tensile strength of the non-oriented PET film was 45.6 MPa (Table 5.4). Therefore, the OH of the tensile strength of this PET film was 241.4 - 45.6 = 195.8 MPa. The orientation-induced increase (OH) values of the T,, Tm, barrier, and mechanical properties were independent of the degree of transesterification. However, the 011 of the %crystallinity and the density showed a slight decrease as the degree of transesterification increased. Orientation had a non-equivalent effect on selected properties of PET, PEN and blend fins. For example, for the PET films, the relative % increases in the T,, % crystallinity, density, barrier properties, tensile strength, and % elongation were higher than those of the PEN and blend films. In contrast, the OH of the Young’s modulus of the PEN films was higher than that for the PET films. In addition, the OH of those 183 properties (except for the Young’s modulus) of the blend films appeared to be dependent on the blend composition. The OH of the respective properties of blends with a higher PET rich phase seemed to be greater than those of the blends with a higher PEN rich phase. This might be due to the difference in the mechanism of strain-induced crystallization and the temperature dependency of the strain-induced crystallization for PET, PEN, and their respective blends. The results from this study showed that PET/PEN blends are a fair alternative for modified PET. However, transesterification reactions between PET and PEN are required to enhance the miscibility of the blends. The blend composition is a very important factor in controlling the thermal, mechanical, and barrier properties of the blends. Therefore, selection of the blend composition is very crucial, and a compromise between the requirements of the blend performance and the cost of the PEN needs to be considered. The results of this study showed that the barrier properties and the tensile strength of the blends were significantly improved over the corresponding PET values. The blends were as brittle as PEN. However, after biaxial orientation, the blends are much more flexible. It is recommended that further studies on PET/PEN blends for bottles and other containers be carried out. Thermal treatment and orientation are important processes in bottle production. These processes are also important factors controlling the characteristics of the blends. In this study, the film fabrication process for PET, PEN and blends was very similar in terms of the equipment and technique, except for the processing temperature and time. For bottle production, it is expected that similar equipment and techniques are also applicable for PET, PEN, and blends. Moreover, for 184 the blow molding process, the viscosity of the melted resins and heat resistance of the blend bottles are important and should be considered as well. 185 Appendices 186 Appendix A Statistical analysis of the degree of transesterification The analysis of variance (ANOVA) was performed by using a single factor AN OVA on the microcomputer statistical program in Microsoft 2000. Table A1: Analysis of variance of degree of transesterification of the PET/PEN blend resins and films. Source of variance Blend resins Blend films F-value Fem“. Mean Mean Sample ID 2 285°C 1 pass 0.061 0.076 2.6052" 10.128 3 ' 300°C 1 pass 0.219 0.222 0.0526* 6.6079 4 315°C 1 pass 0.309 0.356 9.8932" 16.2581 5 285°C 2 passes 0.151 0.192 5.1675“ 98.502 6 300°C 2 passes 0.383 0.405 4.4185" 5.9874 7 315°C 2 passes 0.513 0.525 3.6570“ 7.7086 8 10% PEN wt. 0.189 0.204 0.3065" 10.128 9 30% PEN wt . 0.181 0.208 5.2177* 10.128 10 40% PEN wt 0.13 0.192 40.9731 10.128 * Insignificant difference at the confidence level (p) of 0.05, where F -value is less than Fmtjca]. ** Insignificant difference at the confidence level (p) of 0.01, where F—value is less than Fmfica]. 187 Table A2: Analysis of variance of D-values of the PET/PEN blend resins and films. Source of variance Blend resins Blend films F-value Fen-“ca. Sample ID Mean Mean 2 285°C 1 pass 0.020 0.025 2.8285“ 10.1279 3 300°C 1 pass 0.073 0.074 0.0126* 6.60787 4 315°C 1 pass 0.103 0.123 8.0671" 21.1975 5 285°C 2 passes 0.047 0.061 3.92* 18.5127 6 300°C 2 passes 0.126 0.133 0.6508“ 5.98737 7 315°C 2 passes 0.187 0.180 2.5312“ 7.7086 8 10% PEN wt. 0.027 0.024 0.5504‘ 10.1279 9 30% PEN wt 0.066 0.079 4.6035“ 10.1279 10 40% PEN wt 0.056 0.087 44.3538 10.1279 * Insignificant difference at the confidence level (p) of 0.05, where F-value is less than Fcritical- ‘ "Insignificant difference at the confidence level (p) of 0.01, where F-value is less than Fcritical- 188 Table A3: Analysis of variance of degree of transesterification of blends with varied composition of 14.5 to 50.4 mole % PEN. Source of variance Mean Sample ID 8 14.5 mole % PEN 0.197 3 33.7 mole % PEN 0.221 9 39.5 mole % PEN 0.195 10 50.4 mole % PEN 0.161 ANOVA: single factor Source of Degree of Mean Variation SS freedom square F P-value Fain-ca. Between Groups 0.008291 3 0.002764 4.666937"' 0.01394 5.09192 Within Groups 0.010659 18 0.000592 Total 0.01895 21 *Insignificant difference at the confidence level (p) of 0.01, where F-value is less than Fcritical- 189 Table A4: Analysis of variance of D-values of blends with varied composition of 14.5 to 50.4 mole % PEN. Source of variance Mean Sample ID 8 14.5 mole % PEN 0.025 3 33.7 mole % PEN 0.074 9 39.5 mole % PEN 0.072 10 50.4 mole % PEN 0.072 ANOVA: single factor Source of Degree of Mean Variation SS freedom square F P-value Fcn'fica] Between Groups 0.009503 3 0.003168 31.0989“ 2.47E-07 5.09192 Within Groups 0.001833 18 0.000102 Total 0.011336 21 *Significant difference at the confidence level (p) of 0.01, where F -va1ue is larger than Fcritical- 190 Table A5: Analysis of variance of D-values of blends with varied composition of 33.7 to 50.4 mole % PEN. Source of variance Mean Sample ID 3 33.7 mole % PEN 0.074 9 39.5 mole % PEN 0.072 10 50.4 mole % PEN 0.072 AN OVA: single factor Source of Degree of Mean Variation SS freedom square F P-value Fen-“cal Between Groups 2.48E-05 2 1.24E-05 0.099373“ 0.906038 3.73889 Within Groups 0.001747 14 0.000125 Total 0.001772 16 *Insignificant difference at the confidence level (p) of 0.05, where F-value is less than Fcritical- 191 Appendix B Calibration curve for the density analysis Calibration curve for the standard calibration floats The solvent system: Toluene and Carbon tetrachloride. The calibration equation: Y = -0.0038X + 1.4336, where Y is the density (g/cc) and X is the position hour the bottom of the gradient column. Table B1 : The positions in the gradient column and the density values of the standard calibration floats. Code Color Density Position lg/cc) Bead 1 Large blue 1.2494 49.5 Bead 2 Blue 1 1.300 34.5 Bead 3 Blue 2 1.3500 22.0 Bead 4 Yellow 1.3997 9.0 Bead 5 Red 1.4199 4.0 i ; Figure 13.1. The calibration curve I for the standard floats at 23° C 1 1.4400 g 1.4200 - 1.4000 1 ’13 1.3800 4 g 9 1.3600 . Q 1.3400 1 § 1.3200 4 . '3 1.3000 1 1 .1: . I" 1:200 ’ y = -0.0038x + 1.4336 I ' 0° ‘ R’ = 0.999 1 1 1.2400 « ! ' 1.2200 . . . . . i 0 10 20 30 40 50 60 The Position : 192 The positions of the blend samples and the calculated density based on calibration equation. Table B2.The positions and the density of the blend resins Sample ID Position Posiiton Posiiton Position Average density test 1 test 2 test 3 test 4 (g/cc) PET 8 8 8 8 1.4041 2 25.6 25.6 25.65 25.6 1.3389 3 25.75 25.75 25.7 25.75 1.3385 4 25.75 25.75 25.75 25.75 1.3384 5 25.75 25.8 25.8 25.8 1.3383 6 25.75 25.7 25.75 25.75 1.3385 7 26 26 25.8 25.8 1.3379 8 25.75 25.75 25.7 25.7 1.3385 9 26.1 26 26.15 26.15 1.3371 10 26.2 26.2 26.25 26.25 1.3367 PEN 18.8 18.8 18.9 18.9 1.3640 Table B3. The positions and the density of the non-oriented films sample ID Position Position Average density test 1 test 2 (g/cc) PET 24.5 24.6 1.3429 2 25.6 25.3 1.3395 3 25.6 25.6 1.3390 4 25.6 25.7 1.3388 5 25.6 25.6 1.3390 6 25.6 25.6 1.3390 7 25.75 25.75 1.3384 8 24.8 24.7 1.3421 9 25.8 25.9 1.3381 10 26.1 26.1 1.3371 PEN 27.6 27 .6 1.3316 193 Table B4. The positions and the density of the oriented films Sample ID Position Position Average Density test 1 test 2 (g/cc) PET 18.5 19 1.3621 #2 22.25 22.4 1.3485 #3 22.4 21.75 1.3494 #4 21.8 22 1.3501 #5 21.2 21.6 1.3520 #6 23 23 1.3459 #7 22.8 23 1.3463 #8 20.4 21.3 1.3541 #9 23 22.6 1.3467 #10 22 23.25 1.3473 PEN 23.2 23.3 1.3450 194 Appendix C Statistical analysis for the glass transition temperature of the blends The analysis of variance (AN OVA) was performed by using a single factor AN OVA on the microcomputer statistical program in Microsoft 2000. Table Cl: Analysis of variance of the T g values of the blend resins with 25 % (wt/wt) PEN Source of variance Mean Sample Processing Degree of condition transesterification T, (°C) (B) 2 285°C /1 pass 0.069 76.87 3 300°C /1 pass 0.221 82.04 4 315°C /1 pass 0.333 78.63 5 285°C /2 pass 0.172 82.51 6 300°C /2 pass 0.394 84.28 7 315°C /2 pass 0.519 81.77 ANOVA Degree Source of of Variation SS freedom Mean square F P-value Fen-m1 Between Groups 77.96479 5 15.59296 1.766286" 0.207538 3.325837 Within Groups 88.28105 10 8.828105 Total 166.2458 15 V *Insignificant difference at the confidence level (p) of 0.05, where F-value is less than Fcritical~ 195 Table C2: Analysis of variance of the Tg values of the oriented films with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of condition transesterificatio n T0 (°C) (B) 2 285°C /1 pass 0.069 84.56 3 300°C /1 pass 0.221 83.55 4 315°C /1 pass 0.333 84.79 5 285°C /2 pass 0.172 85.49 6 300°C /2 pass 0.394 93.98 7 315°C /2 pass 0.519 84.38 ANOVA Degree Of Source of Variationl SS freedom lMean square F P-value F "ML Between Groups 4.488835 5 0.897767 0.505588" 0.764399 4.387374 Within Groups 10.65414 6 1.77569 Total 15.14297 1 l *Insignificant difference at the confidence level (p) of 0.05, where F-value is less than Fcritical- 196 Table C3: Analysis of variance of the calculated and measured T, values of the blend resins processed at 300°C and 1 pass. Source of variance Mean Sample % (wt/wt) PEN T, (°C) 8 10 78.6 3 25 83.2 9 30 86.0 10 40 89.4 . Degree Source of Of Mean Variation SS freedom square F P-value F min-ca. Bet-ween Groups 124.5782 3 41.52606 31.66289“ 0.00302 6.591392 Within Groups 5.246023 4 1.311506 Total 129.8242 7 *Significant difference at the confidence level (p) of 0.05, where F-value is greater than Fcritical- 197 Appendix D Statistical analysis of the permeability coefficient The analysis of variance (ANOVA) was performed by using a single factor AN OVA on the microcomputer statistical program in Microsoft 2000. Table D1: Analysis of variance of the water vapor permeability coefficient (Pflzo ) of the non-oriented films obtained from blend resins with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of P” 10 condition transesterification (Kg ml, m2 Pa) (B) 2 285°C /1 pass 0.069 2.42E-15 3 300°C /1 pass 0.221 2.40E-15 4 315°C /1 pass 0.333 2.42E-15 5 285°C /2 pass 0.172 2.41E-15 6 300°C /2 pass 0.394 2.29E-15 7 315°C glass 0.519 2.35E-15 ANOVA Degree Source of of ' Variation SS freedom Mean square F P-value Fmfic,_.__ Between Groups 2.457E-32 5 4.91E-33 5.29312“ 0.033139 8.745928 ' Within Groups 5.569E-33 6 9.28E-34 Total 3.014E-32 1 l *Insignificant difference at the confidence level (p) of 0.01, where F -value is less than F critical- - 198 Table D2: Analysis of variance of the water vapor permeability coefficient (PHI, ) of the oriented films obtained from blend resins with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of P1120 condition transesterification (Kg ml, m2 Pa) (B) 2 285°C /1 pass 0.069 1.43E-15 3 300°C /1 pass 0.221 1.52E-15 4 315°C /1 pass 0.333 1.38E-15 5 285°C /2 pass 0.172 1.49E-15 6 300°C /2 pass 0.394 . 1.43E-15 7 315°C /2 pass 0.519 1.64E-15 ANOVA Degree Source of of Variation . SS freedom Mean square F P-value Fain-cap, Between Groups 1.47E-31 5 2.93E-32 1.283578" 0.333081 3.105875 Within Groups 2.74E-31 12 2.29E-32 Total 4.21E-3l l7 *Insignificant difference at the confidence level (p) of 0.05, where F-value is less than Fcritical- 199 Table D3: Analysis of variance of the water vapor permeability coefficient (PHI, ) of blend films with composition of 14.5 to 50.4 mole % PEN. Source of variance Mean of Pfizo (kg m/ m2 s Pa) Sample ID Non-oriented films Oriented films 1 PET 2.96E-15 1.65E—15 8 14.5 mole % PEN 2.76E-15 1.7OE-15 3 33.7 mole % PEN 2.40E-15 1.52E-15 9 39.5 mole % PEN 2.37E-15 1.26E-15 10 50.4 mole % PEN 2.12E-15 1.28E-15 11 PEN 2.15E-15 0.49E-15 Blends processed at 300°C and 1 pass though the extruder. AN OVA: single factor/non-oriented films Source of Degree of Mean Variation SS freedom square F P-value melfi Between Groups 8.9E-31 4 2.224E-31 296.7477 3.98E-06 5.192163 Within Groups 3.7SE-33 5 7.494E-34 Total 8.93E-31 9 *Significant difference at the confidence level (p) of 0.05, where F-value is greater than I:"critical- AN OVA: single factor/oriented films Source of ' Degree of Mean ‘ Variation SS fi'eedom square F P-value Fmfl Between Groups 4.06E-3l 4 1.02E-31 7.178281 0.007013 3.63309 Within Groups 1.27E-3l 9 1.42E-32 Total 5.34E-31 l3 *Significant difference at the confidence level (p) of 0.05, where F -value is greater than Fcritical- 200 Table D4: Analysis of variance of the oxygen permeability coefficient (P02 ) of the non- oriented films obtained from blend resins with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of P02 condition transestagfication (Kg ml, m2 Pa) 2 285°C /1 pass 0.069 5.22E-l9 3 300°C /1 pass 0.221 5.31E-19 4 315°C /1 pass 0.333 5.22E-l9 5 285°C /2 pass 0.172 5.11E-19 6 300°C /2 pass 0.394 4.99E-19 7 315°C /2 pass 0.519 5.07E-19 ANOVA Degree Source of of Variation SS freedom Mean square F P-value Fain-0.1 Between Groups 2.80686E-39 5 5.6lE-40 0.71319* 0.621525122 2.77285 Within Groups 1.41682E-38 18 7.87E-40 Total 1.69751E-38 23 . *Insignificant difference at the confidence level (p) of 0.05, where F-value is less than l:"critical- 201 Table D5: Analysis of variance of the oxygen permeability coefficient (P0z ) of the oriented films obtained from blend resins with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of PO2 condition transestgfication (Kg ml, m; h) 2 285°C /1 pass 0.069 3.32E-19 3 300°C /1 pass 0.221 2.82E-l9 4 315°C /1 pass 0.333 3.55E-19 5 285°C /2 pass 0.172 3.45E-19 6 300°C /2 pass 0.394 3.51E-19 7 315°C /2 pass 0.519 3.90E-19 ANOVA Degree Source of of Variation SS freedom Mean square F P-value Fcritical_ Between Groups 7.96641E-38 5 1.59E—38 2.3561* 0.163235 4.387374 Within Groups 4.05733E-38 6 6.76E-39 Total 1.20237E-37 l 1 *Insignificant difference at the confidence level (p) of 0.05, where F-value is less than Fcfiticglo 202 Table D6: Analysis of variance of the oxygen permeability coefficient (P02 ) of the blend films with composition of 14.5 to 50.4 mole % PEN. Source of variance Mean of P02 (cc mil/ m2 d atrn) Sample ID Non-oriented films Oriented films 1 PET 157.83 102.16 8 14.5 mole % PEN 142.36 90.84 3 33.7 mole % PEN 128.32 68.21 9 39.5 mole % PEN 122.29 73.61 10 50.4 mole % PEN 113.29 71.48 1 1 PEN 72.45 20.91 Blends processed at 300°C and 1 pass though the extruder. ANOVA: single factor/ non-oriented films Degree Source of of Variation SS freedom Mean square F P-value Famed Between Groups 4915.988 4 1228.997 21.1345 5.05E-06 3.055568 Within Groups 872.2683 15 58.15122 Total 5788.257 19 *Significant difference at the confidence level (p) of 0.05, where F-value is greater than Fcritical- ANOVA: single factor/oriented films Degree Source of of Variation SS freedom Mean square F P-value Fem“. Between ' ' Groups 1516.824 4 379.2061 23.6317 0.004811 6.388234 Within Groups ' 64.18599 4 16.0465 Total 1581.01 8 *Significant difference at the confidence level (p) of 0.05, where F-value is greater than Fcritical- 203 Table D7: Analysis of variance of the carbon dioxide permeability coefficient (PCO2 ) of the non-oriented films obtained fi'om blend resins with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of Pm2 condition transesterification (Kc ml, m2 Pa) (B) 2 285°C /1 pass 0.069 3.71E-18 3 300°C /1 pass 0.221 4.03E-18 4 315°C /1 pass 0.333 3.98E-18 5 285°C /2 pass 0.172 3.84E-18 6 ’ 300°C /2 pass 0.394 - 3.71E-18 7 315°C /2 pass 0.519 3.94E-18 ANOVA Degree Source of - of ' variation 88 freedom Mean square F P-value Fm'fiu] Between Groups 1.89E-37 5 3.77E-38 824643" 0.01158 8.745928 Within Groups 2.74E-38 6 4.57E-39 Total 2.16E—37 l 1 *Insignificant difference at the confidence level (p) of 0.01, where F-value is less than Feritical- 204 Table D8: Analysis of variance of the carbon dioxide permeability coefficient (Pa, ) of the oriented films obtained from blend resins with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of PCO2 condition transest(eBr;fication (Kg ml, m; Pa) 2 285°C /1 pass 0.069 2.52E-18 3 300°C /1 pass 0.221 2.17E-18 4 315°C /1 pass 0.333 2.41E-18 5 285°C /2 pass 0.172 2.66E-18 6 300°C /2 pass 0.394 2.70E-18 7 315°C /2 pass 0.519 2.45E-l8 ANOVA Degree Source of of Variation SS freedom Mean square F P—value Fcriticpl_ Between Groups 2.79E-37 5 5.58E—38 5.23702“ 0.101738 9.013434 Within Groups 3.2E-38 3 1.07E-38 Total 3.11E-37 8 *Insignificant difference at the confidence level (p) of 0.05, where F-value is less than Fcritical- . 205 Table D9: Analysis of variance of the carbon dioxide permeability coefficient (Pm: ) of the blend films with composition of 14.5 to 50.4 mole % PEN. Source of variance Mean of Poo, (cc mil/ cm2 d atm) Sample II) Non-oriented Oriented films films 1 PET 9.65E-02 5.41E-02 8 14.5 mole % PEN 8.1 113-02 5.13E-02 3 33.7 mole % PEN 7.08E-02 3.81E-02 9 39.5 mole % PEN 6.74E-02 3.43E-02 10 50.4 mole % PEN 5.71E-02 2.87E-02 l 1 PEN 2.50E-02 9.52E-03 Blends processed at 300°C and 1 pass though the extruder. ANOVA: single factor/ non-oriented films Degree Source of of Variation SS fieedom Mean square F P-value Fem“. Between Groups 0.001787 4 0.000447 204.5253 5.192163 Within Groups 1.09E-05 5 2.18E-06 Total 0.001798 9 *Significant difference at the confidence level (p) of 0.05, where F -value is greater than Fcritical- ANOVA: Source of V Total F 47.47457 5.192163 *Significant difference at the confidence level (p) of 0.05, where F-value is greater than Fcritical- 206 Appendix E Statistical analysis of the tensile strength The analysis of variance (ANOVA) was performed by using a single factor AN OVA on the microcomputer statistical program in Microsoft 2000. Table E1: Analysis of variance of the tensile strength along the machine direction of the non-oriented films with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of Tensile strength condition transesterification (MPa) (B) 2 285°C /1 pass 0.076 50.89 3 300°C /1 pass 0.222 47.33 4 315°C / 1 pass 0.356 50.27 5 285°C /2 pass 0.192 47.05 6 300°C /2 pass 0.405 44.89 7 315°C /2 pass 0.525 51.42 ANOVA Degree of Source of Variation SS freedom Mean Square F P-value F m-ticL Between Groups 4807141 5 9614282 3.0392"I 0.02136 3.557915 Within Groups 11704607 37 3163407 Total 1651 1748 42 *Insignificant difference at the confidence level (p) of 0.01, where F-value is less than Fcfifica]. 207 Table E2: Analysis of variance of the tensile strength along the cross machine direction of the non-oriented films with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of Tensile strength condition transesterification (MPa) (13) 2 285°C /1 pass 0.076 45.35 3 300°C /1 pass 0.222 48.08 4 315°C /1 pass 0.356 40.84 5 285°C /2 pass 0.192 44.74 6 300°C /2 pass 0.405 44.9 7 315°C /2 pass 0.525 50.99 ANOVA Degree Source of of Variation SS freedom Mean square F P-value PM“, Between Groups 8242377 5 1648475 30748” 1.52E-12 2.455828 Within Groups 2090876 39 53612.2 Total 10333253 44 ‘Significant difference at the confidence level (p) of 0.05, where F-value is greater than Feritical- 208 Table E3: Analysis of variance of the tensile strength of the oriented films obtained from blend resins with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of Tensile strength condition transesterification (MPa) (B) 2 285°C /1 pass 0.076 153.2 3 300°C /1 pass 0.222 182.7 4 315°C /1 pass 0.356 170.0 5 285°C /2pass 0.192 148.1 6 300°C /2 pass 0.405 126.0 7 315°C /2 pass 0.525 148.8 AN OVA Degree Source of of Variation SS freedom Mean square F P-value Fwy-c..— Between Groups 12054.5 2410.89 15.086* 8.16E-08 2.49361 Within Groups 5433.47 34 159.808 Total 17487.9 39 *Significant difference at the confidence level (p) of 0.05, where F-value is greater than Fcritical- 209 Appendix F Statistical analysis of the Percent elongation The analysis of variance (ANOVA) was performed by using a single factor AN OVA on the microcomputer statistical program in Microsoft 2000. Table Fl: Analysis of variance of the % elongation along the machine direction of the non-oriented films with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of % Elongation condition transesteri fication (B) 2 285°C /1 pass 0.076 3.4 3 300°C /1 pass 0.222 3.03 4 315°C /1 pass 0.356 3.5 5 285°C /2 pass 0.192 3.0 6 300°C /2 pass 0.405 3.0 7 315°C /2 pass 0.525 3.0 ANOVA Degree of Source of Variation SS freedom Mean Square F P-value F critical Between Groups 1.63185 5 0.32637 2.1918" 0.076632 2.477165 Within Groups 5.360398 36 0.1489 Total 6.992248 41 *Insignificant difference at the confidence level (p) of 0.05, where F-value is less than Fcritical- 210 Table F2: Analysis of variance of the % elongation along the cross machine direction of the non-oriented films with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of % Elongation condition transesterification (B) 2 285°C /1 pass 0.076 2.99 3 300°C /1 pass 0.222 3.06 4 315°C /1 pass 0.356 2.7 5 285°C /2 pass 0.192 2.87 6 300°C /2 pass 0.405 2.86 7 315°C /2 pass 0.525 3.14 ANOVA Degree Source of of Variation SS freedom Mean square F P-value Fain-“L Between Groups 0.77159 5 0.154318 7.9954“ 2.91E-05 2.455828 Within Groups 0.75273 39 0.019301 Total 1.52432 44 *Significant difference at the confidence level (p) of 0.05, where F-value is greater than Fcritical- 211 Table F3: Analysis of variance of the % elongation of the oriented films obtained from blend resins with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of % Elongation condition transesterification (B) 2 285°C /1 pass 0.076 148.8 3 300°C /1 pass 0.222 130.4 4 315°C /1 pass 0.356 149.5 5 285°C /2 E58 0.192 148.6 6 300°C /2 pass 0.405 127 .7 7 315°C /2 pass 0.525 140.2 ANOVA Degree Source of of Variation SS freedom Mean square F P-value FmtjcaL Between Groups 3074.44 614.888 2.6826" 0.038469 3.63048 Within Groups 7563.92 33 Total 10638.4 38 *Insignificant difference at the confidence level (p) of 0.01, where F-value is less than Fcritical- 212 Appendix G Statistical analysis of the Young’s modulus The analysis of variance (AN OVA) was performed by using a single factor AN OVA on the microcomputer statistical program in Microsoft 2000. Table G1: Analysis of variance of the Young’s modulus along the machine direction of the non-oriented fihns with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of Young’s modulus condition transesterification (MPa) (13) 2 285°C /1 pass 0.076 1569.95 3 300°C /1 pass 0.222 1718.18 4 315°C /1 pass 0.356 1714.74 5 285°C /2 pass 0.192 1688.54 6 300°C /2 pass 0.405 1627.86 7 315°C /2 pass 0.525 1739.07 ANOVA Degree of Source of Variation SS fieedom Mean Square F P-value. F crimp Between Groups 2772.832 5 554.5664 25854" 0.04254 3.574399 Within Groups 7721.813 36 214.4948 Total 10494.64 41 *Insignificant difference at the confidence level (p) of 0.01, where F-value is less than Fcritical- 213 Table G2: Analysis of variance of the Young’s modulus along the cross machine direction of the non-oriented films with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of Young’s modulus condition transesterification (MPa) (B) 2 285°C /1 pass 0.076 1549.95 3 300°C /1 pass 0.222 1618.90 4 315°C /1 pass 0.356 1560.98 5 285°C /2 pass 0.192 1549.95 6 300°C /2 pass 0.405 1520.30 7 315°C /2 pass 0.525 1586.49 ANOVA Degree Source of of Variation _ SS freedom Mean square F P-value Fcfitica_l_ Between Groups 981.8469 5 196.3694 1.8280* 0.131293 2.469648 Within Groups 3974.564 37 107.4207 Total 4956.41 1 42 *Insignificant difference at the confidence level (p) of 0.05, where F-value is less than l:cr'itical- 214 Table G3: Analysis of variance of the Young’s modulus of the oriented films with 25 % (wt/wt) PEN. Source of variance Mean Sample Processing Degree of Young’s modulus condition transesterification (MPa) (B) 2 285°C /1 pass 0.076 1785.8 3 300°C /1 pass 0.222 1823.6 4 315°C /1 pass 0.356 1679.5 5 285°C /24>ass 0.192 1693.9 6 300°C /2 pass 0.405 1697.8 7 315°C /2 pass 0.525 1673.6 ANOVA Degree Source of of Variation SS freedom Mean square F P-value Fain-m; Between Groups 136299 5 27259.9 0.9082* 0.485661 2.45583 Within Groups 1170557 39 30014.3 Total 1306856 44 *Insignificant difference at the confidence level (p) of 0.05, where F-value is less than Fcriticai- 215 Table G4: Analysis of variance of the Young’s modulus along the machine direction of the non-oriented films with composition of 14.5 to 50.4 mole % PEN. Source of variance Mean Sample ID Young’s modulus (MPa) 1 PET 1648 8 14.5 mole % PEN 1684 3 33.7 mole % PEN 1718 9 39.5 mole % PEN 1556 10 50.4 mole % PEN 1692 ll PEN 1919 Blends processed at 300°C and 1 pass though the extruder. ANOVA: single factor Source of Degree of Mean Variation SS freedom square F P-value Fem-“L Between Groups 2318.794 5 463.7587 1.69459“ 0.163425 2.502631 Within Groups 9031.084 33 273.6692 Total 11349.88 38 *Insignificant difference at the confidence level (p) of 0.05, where F -value is less than Fcritical- 216 Table G5: Analysis of variance of the Young’s modulus along the cross machine direction of the non-oriented with composition of 14.5 to 50.4 mole % PEN. Source of variance Mean Sample ID Young’s modulus (MPa) 1 PET 1554 8 14.5 mole % PEN 1653 3 33.7 mole % PEN 1619 9 39.5 mole % PEN 1617 10 50.4 mole % PEN 1532 11 PEN 1685 Blends processed at 300°C and 1 pass though the extruder. ANOVA: single factor Source of Degree of Mean Variation SS freedom square F P-value Fcritiglr Between . . GI‘OUE 2763.005 5 552.601 7.68659" 3.58E-05 2.443429 Within Groups 2947.554 41 71.89155 Total 5710.559 46 *Significant difference at the confidence level (p) of 0.05, where F-value is greater than Fcritical- 217 Table G6: Analysis of variance of the Young’s modulus of the oriented films with composition of 14.5 to 50.4 mole % PEN. Source of variance Mean Sample ID Young’s modulus (MPa) 1 PET 2304.3 8 14.5 mole % PEN 1932.4 3 33.7 mole % PEN 1823.6 9 39.5 mole % PEN 1901.1 10 50.4 mole % PEN 2049.2 11 PEN 2573.5 Blends processed at 300°C and 1 pass though the extruder. ANOVA: single factor Source of Degree of Mean Variation SS freedom square F P-value Fwy Between Groups 2693506 5 5387013 16.1716"I 1.55E-08 2.462549 Within Groups 1265834 38 33311.43 Total 3959341 43 *Significant difference at the confidence level (p) of 0.05, where F -value is greater than Feritical- 218 Bibliography 219 Bibliography Adur, A. M., and L. J. Bonis. 1994. PET-LCP Compatibilized Alloys: A New Unique Development. Future-Pak 1994. Proceedings 11th International Schroeder Conference on Packaging Innovation, Westin Hotel, Chicago, IL. Nov 16-17. Anonymous. 1981. ASTM D3985-81 Standard Test Method for Oxygen Gas Transmission Rate Through Plastic Film and Sheeting Using a Coulometric Sensor. Anonymous. 1985. PET Economical Choice for Engineering Uses. Plastics World. August, p.122-123. Anonymous. 1990 (a). ASTM D882-9O Standard Test Method for Tensile Properties of Thin Plastic Sheeting. Anonymous. 1990“). ASTM D1505-90 Standard Test Method for Density of Plastics by the Density-Gradient Technique. Anonymous. 1990“). ASTM F 1249-90 Standard Test Method for Water Vapor Transmission Rate Through Plastic Film and Sheeting Using a Modulated Infrared Sensor. Anonymous. 1991. ASTM D638-91 Standard Test Method for Tensile Properties of Plastics. Anonymous. 1994 (a). PEN: Assessing the Potential. European Plastics News. 21(10):32. Anonymous. 1994 0”. Enter a Super Polyester. Textile Month. September, p. 30. Anonymous. 1995. PEN the Hero in Hot-fill Breakthrough. Packaging Week. 10(44): 1. Anonymous. 1999. Online Literature: HiPERTUFTM Polyester Resin. http://www2.shellchemical.com. Appel, O. H. 1996. Barrier Properties of PET and PEN Bottles. Kunststoffe Plast Europe. May, p. 13-14. Bakkar M., Encyclopedia of Packaging Technology, 1986. Wiley and Sons Inc., New York. Barrer, R.M., J.A. Banie, and MG. Rogers. 1983. Journal of Polymer Science, Part A. 1:2565. 220 Bauer, C.W., 1997. The Effects of Degree of Stretching on the Properties of Heat Set Biaxially Stretched Naphthalate-Containing Polyester Films. Plastic Film & Sheeting. 13:336-347. Bilrneyer, F .W., J r., 1962. Textbook of Polymer Science. Wiley-Interscience, New York. p. 237. Bonart, R., and R. Hosemann. 1960. Zur Analyse der Langperiodeninterferenzen. Z. Elelctrochem. 64:314. Brady, T. E., S. A. J abarin, and G.W. Miller. 1974. Permeability of Plastic Films and Coatings to Gases, Vapors, and Liquids. Plenum Press, New York Brennan, D.J., J .E. White, A.P. Haag, S.L. Kram, M.N. Mang, S.Pikulin, and ON. Brown. 1996. Poly(hydroxy amide ether): New High-barrier Thermoplastics. Macromolecules. 29:3707-3716. Brennan, W.P. 1978. Characterization and Quality Control of Engineering Thermoplastics by Thermal Analysis in Thermal Analysis application. Perkin-Elmer Corporation, Norwalk, Connecticut. Briston, J. H. 1989. Plastics films. 3rd. Longrnan Scientific and Technical, Harlow Essex CM20 2JE, England. Buchner, S., D. Wiswe, and H. G. Zachmann. 1989. Kinetics of Crystallization and Melting Behavior of Poly(ethylene naphthalene-2,6-dicarboxylate)‘. Polymer. 30:480. Cakmak, M, J .C. Kim, and SW. Lee. 1996. Neck Formation and its Elimination in Uni and Biaxial Defamation of PEN and its Blends. Polymer Preprint. American Chemical Society, Division of Polymer Chemistry. p. 233. Cakmak, M., Y.D. Wang, and M. Sirnharnbhatla. 1990. Processing Characteristics, Structure Development, and Properties of Uni and Biaxially Stretched PEN Films. Polymer Engineering and Science. 30(12): 721-733. Callander, D and B. Howell, 1994. High Performance Naphthalate Based Packaging Resins part H. Proceedings Bev-Pak Asia 1994, Hong Kong. Callander, D and E. Sisson 1994. High Performance Naphthalate Based Packaging Resins. Proceedings Bev-Pak America 1994, Tarpon Springs, FL. Castle, D. 1996. PEN Homopolymer Safe for Food Contact. Packaging Week. 13(45):5. Cheng, S.Z.D., R. Pan, H. S. Bu, M. Cao, and B. Wunderlich. 1988. Thermodynamic Properties in the Liquid State of High Melting Polymers Containing Phenylene and Naphthylene Groups. Macromolecular Chemistry. 189:1579-1594. 221 Crawshaw, R., and E. Sisson. 1995. “HiPERTUF” Polyester Resins-refining the Packaging Envelope. Proceedings Inno Bev, Mainz, Germany, October 3. Cruz, C.S., F. J. Balta Calleja, H. G. Zachmann, and D. Chen. 1992. Mechanical Properties and Structure of Glassy and Sernicrystalline Random Copolymers of PET and PEN. Journal of Material Science. 27:2161-2164. Davis, V.A., and R.W. Howell, 1995. HiPERTUF 1'” Resins for High Performance Containers. Polyester Technical Center, Shell Chemical Company, Akron, OH, 44305. Devaux, J ., P. Godard, and J .P. Mercier. 1982(3). Bisphenol-A Polycarbonate- Poly(butylene Terephthalate) Transesterification. , 1. Theoretical Study of the Structure and of the Degree of Randomness in F our-Component Copolycondensates. J oumal of Polymer Science, Polymer Physics Edition. 20:1875-1880. Devaux, J ., P. Godard, J .P. Mercier, R. Touillaux, and J.M. Dereppe. 1982“”. Bisphenol- A Polycarbonate-Poly(butylene Terephthalate) Transesterification. H. Structure Analysis of the Reaction Products by IR and 1H and 13C NMP. Journal of Polymer Science, Polymer Physics Edition. 20: 1881-1894. Devaux, J ., P. Godard, and JP. Mercier. 1982“). Bisphenol-A Polycarbonate- Poly(butylene Terephthalate) Transesterification. IH. Study of Model Reactions. Journal of Polymer Science, Polymer Physics Edtion. 20:1895-1900. Devaux, J ., P. Godard, and J .P. Mercier. 1982“”. Bisphenol-A Polycarbonate- Polbeutylene Terephthalate) Transesterification. IV. Kinetics and Mechanism of the Exchange Reaction. Journal of Polymer Science, Polymer Physics Edition. 20: 1901- 1907. De Vries, A. J ., C. Bonnebat, and J. Beautempts. 1977. Uni-and Biaxial Orientation of Polymer Films and Sheets. Journal of Polymer Science: Polymer Symposium. 58:109- 156. DiMarzio, E. A. and J. H. Gibbs. 1959. Glass Temperature of Copolymers Journal of Polymer Science. 40: 121-131. Eersels, K.L.L., and G. Groeninckx. 1996. Influence of Interchange Reactions on the Crystallization and Melting Behaviour of Polyarnide Blends as Affected by the Processing Conditions. Polymer. 37(6): 983-989. Eersels, K.L.L., and G. Groeninckx. 1997. Thermal Analysis of Polyarnide Blends as Obtained by Reactive Melt-Extrusion: Influence of Blend Composition on Crystallization and Melting Behavior. Journal of Applied Polymer Science. 63:573-580. 222 El-Hibri, M. J. and D. R. Paul. 1985. Effects of Uniaxial Drawing and Heat-Treatment on Gas Sorption and Transport in PVC. Journal of Applied Polymer Science. 30:3649-3678 Fakirov, S., M. Sarkissova, and Z. Denchev. 1996‘“). Melting- and Crystallization- induced Sequential Reordering in Immiscible Blends of Poly(ethylene terephthalate) with Polycarbonate or Polyarylate. Macromolecular Chemistry and Physics. 197 :2837-2867. F akirov, S., M. Sarkissova, and Z. Denchev. 1996“”. Sequential Reordering in Condensation Copolymers, 3a) Miscibility-Induced Sequential Reordering in Random Copolyesterarnides. Macromolecular Chemistry and Physics. 197:2889-2907. Forcinio, H. 1993. Plastic Packaging, The Heat is on. Prepared Food. November, p. 86- 88. Fricke, H. 1924. A Mathematical Treatment of the Electric Conductivity and Capacity of Disperse Systems. Physical Review. 24:575-587. Gerding T. K., F. Van Den Berg, and N. de Kruijf. 1995. Trend in Food Packaging: Arising Opportunities and Shifting Demands. Proceeding 9th International Association of Packaging Research Institutes (IAPRI) World ConferenCe on Packaging. September 1995. 1(1): 1-14. Godard, P., J .M. Dekoninck, V. Devlesaver, and J. Devaux. 1986. Molten Bisphenol-A Polycarbonate-Poly(ethylene Terephthalate) Blends. H. Kinetics of the Exchange Reaction. Journal of Polymer Science: Part A: Polymer Chemistry. 24:3315-3324. Gordon, M. and J. S. Taylor. 1952. Journal of Applied Chemistry. 2: 493. Guo, M. 1995. ‘3 C NMR Studies on Macroinitiator and Block Copolymers Derived from 13 C enriched Difunctional Initiator. Polymer Preprint. American Chemical Society, Division of- Polymer Chemistry. p. 825. Guo, M., and HG. Zachmann 1993. Intermolecular Cross-polarization Nuclear Magnetic Resonance Studies of the Miscibility of PET/PEN Blends. Polymer. 34(12): 2503-2507. Hanna, 8., and A. H. Windle. 1988. Geometrical Limits to Order in Liquid Crystalline Random Copolymers. Polymer. 29:207-223. Heisey, C. L., DC. Hoffinann, and J .A. Zawada. 1996. Crytallization of NDC-containing Blends and Copolyrner. Polymer Preprint. American Chemical Society, Division of Polymer Chemistry. p. 231. Hess, K., and H. Kiessig. 1943. Zur Kenntnis der Feinstruktur der Polymidfasem. Die Naturwissenschaften. 3 1 : 171 . Hoffman, D.C., J .K, Caldwell, 1995. Prediction of Barrier and Glass Transition Temperature Enhancements in PET/PEN Blends. Proceedings Specialty Polyesters 1995. 223 Holsti-Miettinen, R.M., K.P. Perttila, J.V. Seppala, and MT Heino. 1995. Oxygen Barrier Properties of Polypropylene/Polyarnide 6 blends. Journal of Applied Polymer Science. 58(9) 1551-1560. Hoshino, S., J. Powers, D. G. Legrand, H. Kawai, and R. S. Stein. 1962. Orientation Studies on Drawn Polyolefins. Journal of Polymer Science. 58:185-204. Ihm, D.W., S. Y. Park, C. G. Chang, Y. S. Kim, H. K. Lee. 1996. Miscibility of Poly(ethylene terephthalate)/poly(ethylene naphthalate) Blends by Transesterification. J oumal of Polymer Science: Part A: Polymer Chemistry. 34:2841. J abarin, SA. 1984. Orientation Studies of Poly(Ethylene Terephthalate). Polymer Engineering and Science. 24(5): 376-384. J abarin, S.A., 1992. Strain-Induced Crystallization of Poly(Ethylene Terephthalate) Polymer Engineering and Science. 32(18): 1341-1349. J abarin, SA. and P.Chandran. 1993. Biaxial Orientation of PET, Part 1: Nature of the Stress-strain Curves. Advances in Polymer Technology. 12(2): 119-132. J adhav, J .Y., and SW. Kantor. 1988. Polyesters, Thermoplastic in Encyclopedia of Polymer Science and Engineering. 2nd edition. John Wiley and Sons Inc., New York. 12:193-256. Jaquiss, D.B.G., W.F.H. Borman, and R.W. Campbell, 1982. Encyclopedia of Chemical Technology, M. Grayson ed. John Wiley and Sons Inc., New York. 18:549. Jenkins, SD. 1995. PET/PEN Copolymers and Blends for Hot-fill and High Barrier Packaging Applications. Proceddings Specialty Polyesters 1995. ICI Chemicals and Polymers Ltd. p. 197. Joyce, R. P. and A. P. Berzinis. 1991. Proceedings Compalloy 91. Schotland Business Research, Princeton. p. 65. Karasz, F. E. 1985. Glass Transitions and Compatibility, Phase Behavior in Copolymer Containing Blends. In Polymer Blends and Mixtures. D. J. Walsh, J .S. Higgins, and A.Maconnachie (eds). Martinus Nijhoff Publishers, Boston. Kelley, F .N., and F. Bueche. 1961. Viscosity and Glass Temperature Relations for Polymer-Diluent Systems. Journal of Polymer Science. 50:549-556. Kimmel, R. 2000. The Challenge of the Plastic Beer Bottle. Presentation at 2000 [FT Annaul Meeting Technical Program, held at Dallas, Texas, USA. June 11-14. 224 Kit, K.M., J .M. Schiltz, and R. Gohil. 1995. Morphology and Barrier Properties of Oriented Blends of Poly(Ethylene Terephthalate) and Poly(Ethylene 2,6-Naphthalate) with Poly(Ethylene-co-Vinyl Alcohol). Polymer Engineering and Science. 35(8): 680- 692. Koros, W. J ., and M. W. Hellums. 1988. Transport Properties in Encyclopedia of Polymer Science and Engineering. 2nd edition. John Wiley and Sons Inc., New York. Supplement volume. p. 788-799. Kotliar, A.M. 1981. Interchange Reactions Involving Condensation Polymers. Macromolecular Reviews. 16:367-395. Kugler, J ., J .W. Gilrner, D. Wiswe, H.G. Zachmann, K. Hahn, and E.W. Fischer. 1987. Study of Transesterification in PET by Small-Angle Neutron Scattering. Macromolecules. 20:1 1 16. Kwei, T.K. 1984. The Effect of Hydrogen Bonding on the Glass Transition Temperatures of Polymer Mixtures. Journal of Polymer Science: Polymer Letters Edition. 22(6): 307- 3 14. Kwei, T.K., E.M. Pearce, J .R. Pennacchia, and M. Charton. 1987. Correlation Between the Glass Transition Temperatures of Polymer Mixtures and Intermolecular Force Parameters. Mracromolecules. 20:1174-1176. Lenz, R.W., and S. Go. 1973. Crystallization-Induced Reactions of Copolymers. HI. Ester Interchange Reorganization of Poly(cis/trans-l ,4-Cyclohexylenedimethylene Terephthalate). Polymer Science: Polymer Chemistry Edition. 11:2927-2946. Light, R. R. and R. W. Seymour. 1982. Effect of Sub-T, Relaxation on the Gas Transport Properties of Polyesters. Polymer Engineering and Science. 22(14): 857-864. Lu, X., and AH. Windle. 1995. Crystallization of Random Copolymers of PET and PEN. Polymer. 36(3): 451. MacDonald, W.A., A.D.W. McLenaghan, G. Mclean, R.W. Richards, and SM. King. 1991. A Neutron Scattering Investigation of the Transesterification of a Main-Chain Aromatic Polyester. Macromolecules. 24: 6164. MacKnight, W. J ., F. E. Karasz, and J. R. Fried. 1978. Solid State Transition Behavior of Blends. In Polymer Blends. D.R. Paul and S. Newman (eds.). Academic Press. New York. Margolis, J .M. 1985. Engineering Thermoplastics: Properties and Applications. Marcel Dekker, Inc., New York. 225 Matsumoto, K., J. F. Fellers, and J. L. White. 1981. Uniaxial and Biaxial Orientation Development and Mechanical Properties of Polystyrene Films. Journal of Applied Polymer Science. 26:85-96. McGee, T.M., and AS. Jones. 1995. The Effect of Processing Parameters on Physical Properties of PET/PEN Blends for Bottle Application. Eastman Chemical Company, Kingsport, TN. . Mencik, Z. 1976. Krystalicka Struktura Polykondenzatu Kyseliny Nafialen-2,6- dikarbonove Etylenglykolem. Chemisky Prumysl. 17(2): 78-81. Michaels, AS, and H.J. Bixler. 1961. Flow of Gases Through Polyethylene. Journal of Polymer Science. 50:413-439. Middleman, S. 1977. Fundamentals of Polymer Processing. McGraw-Hill Inc., New York Mock, J. A. 1983. Advancing Technology to Spur Big Growth for Plastic Composites During the ‘808. Plastic Engineering. 39(2): 13-19. Montaudo, G., M. S. Montaudo, E. Scamporrino, and D. Vitalini. 1992. Mechanism of Exchange in Polyesters. Composition and Microstructure of Copolymers Formed in the Melt Mixing Process of PET and Poly(ethylene adipate). Macromolecules. 2225099. Morris, H. R., J. F. Turner II, B. Munro, R.A. Ryntz, and P. J. Treado. 1999. Chemical Imaging of Thermoplastic Olefin (TPO) Surface Architecture. Langmuir. 15:2961-2972. Morton-J ones, DH. 1989. Polymer Processing. Chapman and Hall Ltd. Murakami, S., M. Yamakawa, M.Tsuji, and S. Kohjiya. 1996. Structure Development in the Uniaxial-Drawing Process of Poly(ethylene naphthalate). Polymer. 37(17): 3945- 3951. Nadella, H. P., J. E. Spruiell, and J. L. White. 1978. Drawing and Annealing of Polypropylene Fibers: Structural Changes and Mechanical Properties. Journal of Applied Polymer Science. 22:3121-3133. Nemphos, S. P., M. Salame, and S. Steingiser. 1986. Barrier Polymers. In Encyclopedia of Packaging Technology. M. Bakkar (eds). Wiley & Sons Inc., New York. O’Brien, K.C., W.J. Koros, and GR. Husk. 1987. Influence of Casting and Curing Conditions on Gas Sorption and Transport in Polyimide Films. Polymer Engineering and Science. 27(3): 211-217. Odian, G. 1991. Principles of Polymerization. 3rd. John Wiley and Sons Inc., New York. 226 Olabisi, O., L. M. Robeson, and MT. Shaw. 1979. Polymer-Polymer Miscibility. Academic Press Inc, New York. Park, H.C and EM. Mount HI. 1988. Films, Manufacture in Encyclopedia of Polymer Science and Engineering. 2nd edition. John Wiley and Sons Inc., New York. 7:96-107. Paul, DR. and CB. Bucknall. 1999. Polymer Blends. Volume 2: Formulation. John Wiley and Sons Inc., New York. Peterlin, Anton. 1988. Morphology in Encyclopedia of Polymer Science and Engineering. 2nd edition. John Wiley and Sons Inc., New York. 10:72-94. Pidgeon, R. 1996. Amoco Unviels Details of PEN Container Range. Packaging Week. 10(45): 937. ' Pilati, F., E. Marianucci, and C. Berti. 1985. Study of the Reactions Occurring during Melt Mixing of Poly(ethylene Terephthalate) and Polycarbonate. Journal of Applied Polymer Science. 30: 1267-1275. Po, R., E. Occhiello, G. Giannotta, L. Pelosini, and L.Abis. 1996. New Polymeric Materials for Containers Manufacture Based on PET/PEN Copolyesters and Blends. _ Polymers for Advanced Technologies. 7: 365-373. Po, R., P. Cioni, L. Abis, E. Occhiello, F. Garbassi. 1991. Sequence Analysis of Aromatic/ethylene Glycol Copolyesters Using 1H and 13C NMR of Oxyethylenic Groups. Polymer Communications. 32(7): 208-212. Porter, R.S., and L.Wang. 1992. Compatibility and Transesterification in Binary Polymer Blends. Polymer. 33(10): 2019-2030. Ramj it, H.G. and RD. Sedgwick. 1976. The Kinetics of Ester-Ester Exchange Reactions by Mass Spectrometer. Journal of Macromolecular Science-Chemistry A. 10:815. Ray, D. D., C. B. Shriver, and R. J. Gartland. 1977. Here’s Why Polyethylene Terephthalate is the Maj or Competitor for Beverage Container Applications. Part 1- Properties and Applications. Plastics Design & Processing. 17(8): 47-52. Rodriguez, F. 1996. Principles of Polymer Systems. Taylor & Francis, Bristol, PA 19007. Rudin A. 1982. Basic Principles of Polymer Molecular Weights. In: The Elements of Polymer Science and Engineering. Acedemic Press, Inc., San Diego, California. Salame,M. 1967. A Correlation between the Structure and Oxygen Permeability of High Polymers. Polymer Preprint. American Chemical Society, Division of Polymer Chemistry. 8:137. 227 Samuels, RI. 1974. Structured Polymer Properties. John Wiley & Sons Inc., New York, New York. Sawyer, LC, and D. T. Grubb. 1996. Polymer Microscopy, 2nd. Chapman & Hall. London, UK. Scarlett, AC. 1958. US. Pat. 2,823,421 (to E.I. do Pont de Nemours & co., Inc. on Feb 18, 1958). Schaeberle, M. D., and P. J. Treado. 1997. Proceedings Pittsburgh Conference, Atlanta, GA. March, p. 913. Schaeberle, M. D., H. R. Morris, J. F. TurnerII, and P. J. Treado. 1999. Raman Chemical Imaging Spectroscopy. Analytical Chemistry News & Features. March 1, p. 175A-181A. Schultz, J. 1974. Polymer Material Science. Prentice-Hall, Inc., Englewood Cliffs, New Jersey. Schwartz, S. S. 1982. Plastics Materials and Processes. Van Nostrand Reinhold Company, New York. Seferis, J. C. and R J. Samuels. 1979. Coupling of Optical and Mechanical Properties in Crystalline Polymers. Polymer Engineering and Science. 19(14): 975. Sec, KS, and J .D. Cloyd, 1991. Kinetics of Hydrolysis and Thermal Degradation of Polyester Melts. Journal of Applied Polymer Science. 42:845-850. Shepherd, RA. and KL. Ronald. 1991. US. Pat. 5,006,613 (to Eastman Kodak Company); - Shi, Y. 1998. Study on PET/PEN blends: Transesterification Reaction and Properties. Ph.D. dissertation. The University of Toledo. Toledo, Ohio. Smith, W. A., J. W. Barlow, and DR. Paul. 1981. Chemistry of Miscible Polycarbonate- copolyester Blends. Journal of Applied Polymer Science. 26: 4233-4245. Stewart, M.E., A.J. Cox, and D.M. Naylor. 1993. Reactive Processing of PET/PEN Blends. Polymer. 43(19): 4060-4067. Swallow, J .C. and D.K. Baird. 1948. Brit. Pat. 609,797 (to ICI on Oct. 6, 1948). Tibbitt, J. M. 1994. Isophthalic Acid and 2,6 Naphthalene Dicarboxylate Modified Polyesters for Packaging Application. Proceedings 11th International Schroeder Conference on Packaging Innovation, Future-Pak 1994, Westin Hotel, Chicago, Illinois. November 16-17, 1994 228 Toth, J ., C. Rusch, and L. Desoulter. 1994. PET/PEN Blends, Enhancing the Performance of Polyester Containers-Outlook. Proceedings Bev-Pak Asia 1994, Hong Kong. September 22-23, 1994. Tyne M., B. Howell, and E. Sisson. 1995. Redefining the Polyester Packaging Envelope for High Temperature and Chemically Aggressive Applications, Bev-Pak Americas 1995, Innisbrook, FL. Utracki, L. A. 1990. Polymer Alloys and Blends; Thermodynamics and Rheology. Hanser Publishers, New York. Wang, CC. and Y.M. Sun. 1994 (a). Studies on the Formation of PEN. 1. Kinetic Aspects of Catalyzed Reaction in Transesterification of Dimenthylnaphthalate with Ethylene Glycol. Journal of Polymer Science: Part A: Polymer Chemistry. 32: 1295-1304. Wang, CC. and Y. M. Sun. 1994“”. Studies on the Formation of PEN. H. Polymerization of Bis(Hydroxyethyl)naphthalate by Various Metallic Catalysts. J oumal of Polymer Science: Part A: Polymer Chemistry. 32: 1305-1315. Werner, E., S. J anocha, M. J. Hopper, and K. J. Mackenzie. 1988. Polyesters, Films in Encyclopedia of Polymer Science and Engineering. 2nd edition. John Wiley and Sons Inc., New York. 12:193-216. Whinfield, J .R. and J .T. Dickson. 1946. Brit. Pat. 578,079 (June 14, 1946). White, J. L. 1988. Orientation in Encyclopedia of Polymer Science and Engineering. 2nd edition. John Wiley and Sons Inc., New York. 7:595-616. White, J. L., K. C. Dharod, and E. S. Clark. 1974. Interaction of Melt Spinning and Drawing Variables on the Crystalline Morphology and Mechanical Properties of I-IDPE and LDPE Fiber. Journal of Applied Polymer Science. 18:2539-2568. Yamadera, R., and M. Murano. 1967. The Determination of Randomness in Copolyesters by High Resolution NMR. Journal of Polymer Science. 5: 2259-2268. 229 IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII muttWImultilwulyunull