. . .. «2‘ 2...}. n4...) $.33? 94mg u.k hum. a in ha ”.55.. . . . . . w .. .fin‘mfi . 2%.”.9. .. xetzn 232. it 3.1 2 .~ it}... .1: 3 . Minna nHerf}1: . an... . l .« tau . w r. is." turf : . , 53.. .- :ch...r2x.. z? .2... .. a i. . . .ld .. r . , 3t 1.3 {Fugutrt s ._ MN.“ ‘.hfl}i¥.flfi : 1 .{t Iv fi” . I. 1‘ , n’f‘X . ‘ Erumfikgfi..." 1. Mg»?! ha... . ~..... 3 . ; . 1“ . r. 2.. . .2 . ..hr.......:. 1. a... vi» Raw: .I.I.-’~l :1~vx..~ 3.50:... xv “VI-r“ <5), 1 a 3.: 33:0... . flu}?! _. 5...: D! .r . 4 ”9.... 1335......ilexl’. i» av .Wetiium .rfllu. IIV\ I 7“ o a v is. 2251. .5. :3. lxif) :‘ . .QL :3!!! v. J... ..r I‘Otfilill THESiS ‘g't/LL' This is to certify that the dissertation entitled SYNTHESIS AND CHARACTERIZATION OF SUBSTITUTED POLYLACTIDES presented by MAO YIN has been accepted towards fulfillment of the requirements for DOCTOR degree in W 51 Major professor Date /3 S/‘v/hé 3000 MSU is an Affirmative Action/Equal Opportunity Institution 0-12771 LIBRARY Michigan State University PLACE IN RETURN Box to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE | DATE DUE DATE DUE magma 6/01 chlRC/DatoprSS-DJS SYNTHESIS AND CHARACTERIZATION OF SUBSTITUTED POLYLACTIDES By Mao Yin A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2000 sites-55$ E bfcac ABSTRACT SYNTHESIS AND CHARACTERIZATION OF SUBSTITUTED POLYLACTIDES BY Mao Yin Because of their biocompatibility and environmental degradability, polymers derived from lactic acid have long been important materials for medical applications. Once used primarily as sutures and implants, polylactide materials are now entering high volume areas such as packaging. To be successful in these new applications, polylactides must be available that exhibit a broad spectrum of physical properties while retaining the degradability of the parent polymers. This objective can be achieved by altering the substituents on the polylactide chains. A variety of substituted Iactides were synthesized and polymerized to high molecular weight polymers using solution and bulk polymerization. The polymers show a wide range of physical properties: glass transition temperatures range from —37 °C to 85 °C, hydrolytic degradation rates are 2-3 times slower than polylactide's, and degradation temperatures are around 350 °C. We established the relationship between the pendant group and the physical properties of the polymers. The kinetics and mechanism of lactide polymerization was studied. Lactide polymerizations are first-order reactions in both solution and bulk AHMAD? :I-v't' Alibi :5; 5w “R v? ‘v U ii :‘i’erenl Bali :3 rm “ «A1; Dewar, Tile a: Wis polymerization. In bulk polymerization, polymerization rates follow the order of size of substituted group, the bigger the substituted group, the slower the polymerization rate. In solution polymerizations, the polymerization rates do not follow the order of size of substituted group, which is probably caused by different rate-determining steps for the polymerizations. We proposed that lactide polymerization mechanism is similar to “atom-transfer polymerization” when initiated by Sn(Oct)2. Using it as guide, we achieved control of the molecular weights and molecular weight distributions of the polymers. We studied the influence of stereochemistry on the kinetics of polymerizations and the physical properties of the polymers. We found that racemic monomers polymerize faster than either of the optically pure monomers. The racemic polymers are usually amorphous polymers and optically pure polymers are usually crystalline polymers. Copolymerization of lactide with substituted Iactides were investigated. The architectures of polymers depend on the size difference of substituted groups between the lactide and substituted lactide. By controlling the size of substituted group, random, block and alternating copolymers can be synthesized. To my wife Tianle ACKNOWLEDGEMENTS I wish to express my deep appreciation to Professor Gregory L. Baker for his guidance, assistance and constant encouragement throughout the course of this research. A work such as this would not be possible without his assistance. It has been my privilege to have him as my mentor. I would also like to thank professor James Jackson, Milton Smith, Gary Blanchard, and Thomas Pinnavaia for their guidance and assistance. My thanks go to all Baker group members: Jun Qiao, Jun Hou, Jingpin, Yiyan, Gao, Corey, Susan, Tara, Tianqi, J, 3,, Kirk, Fadi and all the friends in Michigan State University, for making my stay an enjoyable one Finally, I want to thank my wife Tianle for her love and sacrifices. I thank my parents and sister for their love and encouragement. 138'. OF Ii US? OF Fl CST OF S llTiDDU 3.1 3.2 3.2 TABLE OF CONTENTS LIST OF TABLE ................................................................................................... ix LIST OF FIGURES .............................................................................................. xi LIST OF SCHEMES ............................................................................................ xvi INTRODUCTION .................................................................................................. 1 1. Applications .............................................................................................. 1 1.1 Medical application ................................................................................ 1 1.2 Biodegradable Packaging Therrnoplastics ............................................. 3 3. Ring-opening polymerization of Iactides and Iactones ............................ 11 3.1 Polymerizability .................................................................................... 11 3.2 Polymerization Mechanism .................................................................. 15 3.2.1 Cationic Polymerization ................................................................. 15 3.2.1.1 Lactones ................................................................................... 15 3.2.1.2. Lactides ................................................................................... 16 3.2.2 Anionic Polymerization ................................................................... 21 3.2.2.1 Lactones ................................................................................... 21 3.2.2.2 Lactide ...................................................................................... 24 3.2.3 Coordination Polymerization .......................................................... 27 3.2.3.1 Aluminum Derivatives ............................................................... 27 3.2.3.2 Tin Derivatives .......................................................................... 38 3.2.3.3 Rare Earth Metal Compounds .................................................. 46 3.2.4 Other Polymerization methods ....................................................... 48 vi 42 43 52. RESUL' 32 33 Fwy 34. 34 3.4 35. poly; 3.2.4 Other Polymerization methods ....................................................... 49 Modification of Polylactide Properties ..................................................... 51 .1 Manipulation of stereochemistry. ......................................................... 51 .2 Copolymerization ................................................................................. 55 .3 Blending ............................................................................................... 63 Degradation of the polylactide and polylactones ..................................... 64 .1 . Thermal degradation ......................................................................... 64 .2. Biodegradation .................................................................................. 66 JLTS AND DISCUSSION ........................................................................... 69 Monomer synthesis ................................................................................. 69 Bulk polymerization of substituted Iactides ...................................... . ...... 74 .1 Initiator survey ...................................................................................... 75 .2 Kinetics of Bulk Polymerization ............................................................ 83 Solution Polymerization of Substituted Lactides ..................................... 96 .1 AI(OiPr)3 as Initiator ............................................................................. 96 .2 Sn(Oct)2/ROH as initiators ................................................................. 106 .3 Comparison of AI(OiPr)3 and Sn(Oct)2lROH as lnitiators in Solution olymerization ............................................................................................ 123 4. Polymerization Mechanism ............................................................. 129 3.4.1 AI(OiPr)3 as initiator ...................................................................... 129 3.4.2 Sn(Oct)2/ROH as catalystfinitiator ................................................ 133 5. The Influence of Stereochemistry on the Kinetics of Solution olymerization ............................................................................................ 137 vii 4. Pair 5. Cap: 52 5.2. 5.2 5.3 6.C 6.2 6.3 EXPERI BIBLio-c 4. Polymer properties ..................................................................................... 150 5. Copolymerization of substituted glycolides ................................................ 161 5.1. Copolymerization through comonomers ......................................... 162 5.2 Copolymerization through asymmetric monomers. ............................ 168 5.2.1. Synthesis and Polymerization of AB Monomers. ...................... 170 5.2.2 Structure of the polymer chains. .................................................. 175 5.3 Thermal properties of the copolymer. ................................................ 179 6. Crystalline substituted polylactide. ........................................................ 183 6.1 Synthesis and polymerization of optical pure monomers ................... 183 6.2 The stereochemistry of the polymer chain ......................................... 187 6.3 The crystallinity of poly(isopropyglycolide) ......................................... 196 7. The degradation of substituted polylactides .......................................... 202 EXPERIMENTAL .............................................................................................. 212 BIBLIOGRAPHY ............................................................................................... 224 viii Table 1. Table 2. Table 3. Table I. Table 5. Table 6. Table 7. Table 8. Table 9, Table 1( Table 11 Table 1; Table 1; Table 14 Table 15 Table Ti Table 1. Table 2. Table 3. Table 4. Table 5. Table 6. Table 7. Table 8. Table 9. Table 10. Table 11 . Table 12. Table 13. Table 14. Table 1 5. Table 16. LIST OF TABLES Some Basic Physical Properties Comparisons of Poly(lactic acid) and Polystyrene .......................................................................................... 5 Comparison of Polyethylene to Poly(lactic acid) .................................. 6 Heats of combustion and Strains of Cycloalkanes per Methylene group ................................................................................................. 13 Polymerizability of cyclic ester monomers ......................................... 14 Bulk Polymerization of Ethylglycolide ................................................ 78 Bulk Polymerization of Isobutylglycolide ............................................ 79 Bulk Polymerization of Hexylglycolide ............................................... 80 The kinetic data for bulk polymerization of substituted Iactides ......... 90 Polymerization rate constants of substituted Iactides ...................... 104 Polymerization activation energies for substituted Iactides .............. 104 Polymerization rates for ethylglycolide initiated by Sn(Oct)2/BBA....114 The activation energy for ethylglycolide initiated by Sn(Oct)2/BBA..114 The activation energy of ethylglycolide initiated by Sn(Oct)2/BBA and AI(OiPr)3 .......................................................................................... 124 Experimental value and calculated value for activation energy difference. ........................................................................................ 146 Polymer Properties .......................................................................... 153 Glass transition temperature of substituted polylactides and substituted polyethylene. ................................................................. 159 Table 17. Table 18. Table 1 9. Table 20. Table 21 . Properties of copolymers ................................................................. 164 The properties of AB polymers. ....................................................... 180 Experimental and calculated values of hexad intensities in the carbonyl region of 13C NMR spectra of poly(rac-isopropyglycolide). 191 The crystallinity and heat of fusion of poly(isopropylglycolide) ........ 197 The degradation of substituted polylactide ...................................... 205 Figure I. E Figura 2. E Figure 3. I Figural. ' Figure 5, I Figure 6. : Figural. ' FigureB. F'QuraS. FlQuail). Flaurrerr. FlSure 12, FlgUTe 13. Figure 14‘ Flgllre 15 LIST OF FIGURES Figure 1. Structures and abbreviations of monomers and degradable polymers ............................................................................................................. 2 Figure 2. Structure of Al initiators ...................................................................... 27 Figure 3. Aggregation states for AI alkoxides .................................................... 30 Figure 4. The structure of porphyrin initiators ................................................... 37 Figure 5. Structure of Sn(Oct)2 and Bu28n(Oct)2 .............................................. 46 Figure 6. Stereochemistry of lactide ................................................................. 53 Figure 7. Tacticity of polylactide ........................................................................ 54 Figure 8. Copolymer architectures .................................................................... 56 Figure 9. Degradation models ........................................................................... 67 Figure 10.The structure of substituted glycolides ............................................... 71 Figure 11.‘H NMR of ethylglycolide and poly(ethylglycolide) ............................. 73 Figure 12.8qu polymerization of ethylglycolide with and without added t- butylbenzyl alcohol as co-initiator ...................................................... 81 Figure 13.Molecular weight versus conversion for the bulk polymerization of substituted glycolides ......................................................................... 82 Figure 14. Kinetics of bulk polymerization of substituted glycolides .................... 84 Figure 15. Kinetics of bulk polymerization of substituted glycolides at low conversion ......................................................................................... 85 Figure 16.The structure of the white precipitate formed during the solution polymerization of ethylglycolide using Sn(Oct)2/ROH as initiator ....... 90 xi Figure 17 Figure 18 Figure 19 Figure 20 Frgura 21 Figure 22 Figure 23 Figum 24 Filure 25 Figure 25 FiWe 28 Film 29 Figure 17.Kinetics of bulk polymerization of substituted glycolides .................... 91 Figure 18.Kinetics of bulk polymerization of substituted glycolides .................... 92 Figure 19.Polymerization/depolymerization data for the bulk polymerization of ethylglycolide ..................................................................................... 93 Figure 20.Molecular weight vesus conversion fro the solution polymerization of substituted Iactides .......................................................................... 100 Figure 21.Kinetics for the solution polymerization of substituted Iactides at 70 °C ......................................................................................................... 1 01 Figure 22.Kinetics for the solution polymerization of substituted Iactides at 90 °C ......................................................................................................... 1 02 Figure 23.Kinetics for the solution polymerization of substituted Iactides at 100 °C ..................................................................................................... 103 Figure 24.Activation energies for polymerization of substituted Iactides .......... 105 Figure 25.Kinetics for the polymerizaiton of ethylglycolide initiated by Sn(Oct)2/ROH and Sn(Oct)2 ............................................................ 109 Figure 26.Solution polymerization of ethylglycolide at different Sn(Oct)2/alcohol ratios ................................................................................................ 1 10 Figure 27.Solution polymerization of ethylglycolide at different Sn(Oct)2/alcohol ratios ................................................................................................ 111 Figure 28.Solution polymerizaiton of ethylglycolide at different monomer/initiator ratios. ............................................................................................... 112 Figure 29.Solution polymerizaiton of ethylglycolide at different monomer/initiator ratios .......................................................................................... . ...... 113 xii Figure 3 Figure! Figure 3 Figure E Figure I Figure Figure Figura FlgUfe Figure Figure FlgUre FigUfe e 30.Solution polymerization of ethylglycolide initiated by Sn(Oct)2/BBA 115 e 31.Solution polymerization of ethylglycolide initiated by Sn(Oct)2/BBA 116 e 32.The activation energy of ethylglycolide initiated by Sn(Oct)2/BBA117 e 33.IR spectra for white precipitate formed during the polymerization of ethylglyolide and by mixing 2-hydroxybutyric acid and Sn(Oct)2 ...... 120 e 34. Kinetics of solution polymerization of ethylglycolide showing the result after adding extra monomer after polymerization reached equilibrium ......................................................................................................... 1 21 e 35.1nitial polymerization rate for ethylglycolide and the decreased rate observed after adding additional monomer ...................................... 122 e 36.Solution polymerization of substituted Iactides initiated by Sn(Oct)2/BBA ................................................................................... 125 e 37.Solution polymerization of ethylglycolide initiated by AI(OiPr)3 and Sn(Oct)2/BBA ................................................................................... 126 e 38.Solution polymerization of ethylglycolide initiated by AI(OiPr)3 and Sn(Oct)2/BBA ................................................................................... 127 e 39.The activation energies for polymerization of ethylglycolide initiated by Sn(Oct)2lBBA and Al(OiPr)3 ............................................................. 128 e 40.The structure of the transition state for polymerization of substituted Iactides ............................................................................................ 132 e 41 .Polymerization of rac—Iactide and L-Iactide ...................................... 139 e 42.Ponmerization of rac-isobutylglycolide and D-isobutylglycolide ...... 140 e 43.The structure of complexes used to simulate polymer growth ......... 147 xiii Figure 4 Fgural Fgum Figure Figure Figure Fgum figure FMUm FlgUTE qu“ FQUN FlQurr Fl91m Flgun FlQUrr FiQUIT Figure 44.The calculated structure for complex with R configuration chain end and 8,8 Iactide ring .......................................................................... 148 Fgure 45. The calculated structure for the complex with S configuration chain and 8,8 Iactide ring .......................................................................... 149 Figure 46.DSC run for substituted poly(glycolide)s .......................................... 152 Figure 47.Thermal analysis results for poly(ethylglycolide) ............................. 154 Figure 48.Thermogrvimetric analysis results for poly(ethylglycolide) ............... 155 Figure 49.Thermogrvimetric analysis results for substituted poly(glycolide)s .. 156 Figure 50.The secondary interaction between polyester chains ...................... 160 Figure 51.DSC runs for Iactide and ethylglycolide copolymers. ....................... 165 Figure 52.Glass transition temperature of Iactide and ethylglycolide copolymers ......................................................................................................... 166 Figure 53.1H NM/R spectra of ethylmethylglycolide and methylphenylglycolide ......................................................................................................... 1 73 Figure 54.Structure of the diastereomers of methylphenylglycolide ................. 172 Figure 55.NOE NMR spectrum of methylphenylglycolide ................................ 174 Figure 56.Carbonyl region of ”C NMR of AB copolymers ............................... 178 Figure 57.DSC runs for AB polymers ............................................................... 181 Figure 58.Thennogrvimetn'c analysis results for AB polymers ......................... 182 Figure 59.NMR spectra of the methane and carbonyl region of poly(S- isopropylglycolide) ........................................................................... 186 Figure 60.Schematic representation of the Bernoulli and first-order Markov models ............................................................................................. 188 xiv Figure 61.05 pr Figure 5233“ figure 63} Figure 64.8 is Figure 65.x Figure 66, X Figure 67.P Figure 68,: is Figure ear Figure 70.1 C Flgme 71.1 I I Figure 73.r Figure 61.DSC runs showing the melting point for poly(isopropylgcholide)s prepared from S and rac-isopropylgycolide ..................................... 192 Figure 62.130 NMR of poly(isopropylglycolide) ................................................ 193 Figure 63.1H NMR of poly(isopropylglycolide) ................................................. 194 Figure 64. Stereosequence assignment for 13C spectrum of poly(rac- isopropylglycolide) ........................................................................... 195 Figure 65.X-ray diffraction pattern for crystalline poly(S-isopropylgcholide).... 198 Figure 66.X-ray diffraction pattern for amorphous poly(rac-isopropylgcholide)199 Figure 67.Peak deconvolution for crystalline poly(S-isopropylglycolide) .......... 200 Figure 68.Determination of heat of fusion for 100% crystalline poly(S- isopropylglycolide) ........................................................................... 201 Figure 69.The weight loss of substituted glycolide during hydrolytic degradation ......................................................................................................... 207 Figure 70.The molecular weight loss of substituted glycolide during hydrolytic degradation ...................................................................................... 208 Figure 71.The molecular weight decrease of substituted Iactides during hydrolytic degradation fitted to random chain scission model. ......... 209 Figure 72.The decrease in the degree of polymerization of substituted Iactides during hydrolytic degradation ........................................................... 210 Figure 73.Contact angles on substituted polylactides plotted as 7 (surface tension of water/methanol solutions) vs cos 0 ................................. 211 Scheme ' Scheme Schema Scheme Scheme Scheme Scheme Scheme Scheme Schem‘ Schema Schema Schem. Schem. Schilmr scheTTIr smemr s°hemr LIST OF SCHEMES Scheme 1. Synthetic route for poly(lactic acid) .................................................. 10 Scheme 2. Cherdron‘s cationic polymerization mechanism ............................... 16 Scheme 3. The cationic polymerization mechanism proposed by Penczek ....... 17 Scheme 4. Cationic polymerization of Iactide initiated by protonic acids ............ 18 Scheme 5. Polymerization of Iactide initiated by methyl triflate (traditional cationic mechanism) ....................................................................................... 19 Scheme 6. Polymerization of Iactide initiated by methyl triflate (mechanism proposed by Kricheldorf) ................................................................... 20 Scheme 7. Propagation routes for anionic polymerization of Iactones ............... 22 Scheme 8. Polymerization mechanism of B-PL .................................................. 22 Scheme 9. The polymerization mechanism of PL proposed by Jedlinski ........... 24 Scheme 10. Anionic polymerization of Iactide through deprotonation ................ 25 Scheme 11. Anionic polymerization of Iactide .................................................... 26 Scheme 12. Mechanism of Iactide polymerization using AI alkoxides ................ 28 Scheme 13. Dissociation of the AI alkoxide aggregates ..................................... 31 Scheme 14. lntramolecular transesterification ................................................... 33 Scheme 15. Intermolecular transesterification ................................................... 34 Scheme 16. Preparation of end-functioned Al initiators ...................................... 35 Scheme 17. Mechanism for tin alkoxide initiated polymerization Iactides .......... 39 Scheme 18. Mechanism for tin halide initiated polymerization of caprolactone..40 Scheme 1'. Scheme 2 Scheme 2 Scheme 1 Scheme 2 Scheme L Schema Scheme Scheme SCheme Scheme SCheme Scheme SCheme Scheme Scheme 19. Mechanism for polymerization of Iactide initiated by Sn(Oct)2 as proposed by Nijenhuis et aI.. ............................................................. 42 Scheme 20. Mechanism for polymerization of Iactide initiated by Sn(Oct)2 as proposed by Kricheldorf ..................................................................... 44 Scheme 21. Mechanism for Iactide polymerization initiated by Sn(Oct)2 as proposed by Penczek ........................................................................ 45 Scheme 22. Polymerization of Iactide initiated by rare earth catalyst ................. 48 Scheme 23. initiators made from epoxides and rare earth halides ..................... 48 Scheme 24..Zwitterionic polymerization mechanism .......................................... 50 Scheme 25. Synthesis of diblock copolymers .................................................... 59 Scheme 26. Synthesis of ABA triblock copolymers ............................................ 60 Scheme 27. Synthesis of cross-Iinkable copolymers ......................................... 62 Scheme 28. Thermal degradation via cis—elimination ......................................... 65 Scheme 29. Radical thermal degradation .......................................................... 66 Scheme 30. The synthetic routes to substituted glycolides ................................ 72 Scheme 31. The propagation step for Iactide polymerization using Sn(Oct)2 as catalyst .............................................................................................. 95 Scheme 32. The first step for the Iactide polymerization using Sn(Oct)leOH . 108 Scheme 33. The mechanism for the Iactide polymerization initiated by Al(OiPr)3 ......................................................................................................... 1 30 Scheme 34. The mechanism of Iactide polymerization using Sn(Oct)2/ROH as catalyst/initiator ................................................................................ 135 Scheme 35. The mechanism for atom—transfer radical polymerization (ATRP) 136 Scheme 36. 5 Scheme 37. 1 Scheme 38. Scheme 33. Scheme 40. Scheme 41. Scheme 42. Scheme 43. Scheme 44, 81 Scheme 36. Stereochemistry of Iactide polymerization .................................... 138 Scheme 37. Kinetic scheme for Iactide polymerization .................................... 142 Scheme 38. Copolymerization of substituted glycolides .................................. 161 Scheme 39. Reactivity difference leads to “blocky” copolymers ....................... 167 Scheme 40. The comparison between two copolymerization methods ............ 169 Scheme 41. Structures of AB monomers and reactivity of different sites ......... 169 Scheme 42. Synthesis of AB glycolide monomers ........................................... 172 Scheme 43. The propagation step in the polymerization of AB monomers ...... 177 Scheme 44. The synthesis of an optically pure substituted Iactide form an amino acid precursor....... ............................................................................ 185 xviii r. Applic 1.1 Medic Be: DCTTQTycoi in Figure YET-all ar 3.5333110 surgical c The bore and lutur "399039 bl Davis ”Wynn T': de‘lelOpg A DOTOU INTRODUCTION 1. Applications 1.1 Medical application Because of their biodegradability and biocompatibility, polylactide, polyglycolide and polylactones (monomer and polymer structures are displayed in Figure 1) have been used in medical applications for wound closure,1 tissue repair and regeneration,2'4 and drug delivery.5‘7 The most successful application has been wound closure, where biodegradable surgical sutures and surgical clips have been made using polylactide, polyglycolide and polylactones. The bioresorption of sutures eliminates the possibility of foreign-body reactions and future infection. Some commercially available biodegradable sutures such as the polyglyconate sutures (glycolide/trimethylene carbonate copolymer) marketed by Davis 8. Geck as Maxon and polyglecaprone 25 (glycolide/caprolactone copolymer) have been used successfully in clinics. Tissue repair and regeneration and drug delivery applications are less developed than wound closure, but there has been exciting progress in the field. A porous biodegradable membrane made from DL-PLA was approved by the FDA for periodontal use. Clinical results show that the new biodegradable membrane, marketed in the U. S. as Guidor by the Butler Company, gives comparable if not better periodontal tissue regeneration than the nonbiodegradable Gore-Tex products made from poly(tetrafluoroethylene). Many degradable products have been developed to deliver a variety of drugs, such as anti-cancer agents, narcotic antagonists, antibiotics, fertility drugs and vaccines. Some products have been marketed such as Lupron Depot, an injectable microcapsule system marketed by Takeda-Abbot using a 70:30 DL- lactide/glycolide copolymer to deliver leuprorelin (anti-cancer drug). However, the majority of products are still in the evaluation stage and more research needs to be done. 0 We} (<50 0 m n = 1: 8-PL, 8-propiolactone n = 1: Poly(B-PL), Poly(B-propiolactone) n = 2: y-BL, y-butyrolactone n = 2: Poly(y-BL), Poly(y-butyrolactone) n = 3: 8-VL, 8-valeroiactone n = 3: Poly(8—VL), Poly(8-valerolactone) n = 4: e-CL, e-caprolactone n = 4: Poly(e-CL), Poly(e-caprolactone) .6; M. Glycolide Poly(glycolic acid) or PGA O O O \n/ik m 0 Lactides: Polylactides or PLA: (L-lactide, D-Iactide, DL-Iactide, mesa-Iactide) Poly(L-Iactide). Poly(D-Iactide). Poly(DL-lactide), Poly(meso-lactidej Figure 1. Structures and abbreviations of monomers and degradable polymers 1.2 Bior ararsv USpi plastics become Tia Ila mamm Concer wflihar mph W338. Therm lbOgg Se'r'er aha Sdlo Waier lTaie Dacgi 1.2 Biodegradable Packaging Thermoplastics There is a need for an environmentally benign packaging thermoplastic as an answer to the tremendous amounts of discarded plastic packaging materials. U. S. plastic sales in 1997 were 74.0 billion pounds of which 29% were listed as plastics in packaging.8 A significant amount of this plastic is discarded and becomes a plastic pollutant that is a blot on the landscape and a threat to marine life. Mortality estimates range as high as 1-2 million seabirds and 100,000 marine mammals per year. A further problem with disposal plastic packaging is the concern for dwindling landfill space. It has been estimated that most major cities will have used up available landfills for solid waste disposal very soon. Plastics comprise approximately 3 percent by weight and 6 percent of the volume of solid waste.9 The polymers and copolymers of lactic acid have been known for some time as unique materials since they are biodegradable, biocompatible and thermoplastic. These polymers are well-behaved thermoplastics,10-11 and are 100 percent biodegradable in an animal body via hydrolysis over a time period of several months to a year. In a wet environment they begin to show degradation after several weeks and disappear in about a year’s time when left on or in the soil or seawater. The degradation products are lactic acid, carbon dioxide and water, all of which are harmless. Conventional plastics have good reasons for their success as packaging materials. They provide appealing aesthetic qualities in the form of attractive packages, which can be quickly fabricated and filled, with specified units of products desracie meagre are vac seeder coma-rm rs veg Temper; Willa: some a Plastici t re3901 Well a Dibvrc Tor Var and Dr alarm 21 products. The packages maintain cleanliness, good mechanical properties and desirable qualities such as transparency for inspection of contents. If lactic acid polymers are to compete with conventional plastics, they should have similar or better properties.12 Table 1 comparies the properties of unplasticized 90I10 copolymer of L and DL-lactide and polystyrene. Unplasticized poly(lactic acid)s are very stiff with tensile properties similar to PS. The 90/10 copolymer is selected because of property and processing advantages. Poly(lactic acid)s containing 98 to 100% of the L-lactic acid have very high crystallinity. Fabrication is very difficult because of crazing, and need for processing at very high temperatures causes discoloration and loss of molecular weight. By using monomer, Iactide, or Iactide oligmer as plasticizers, flexible poly(lactic acid)s can be made. Using monomer and oligmer as plasticizers has some advantages. First, they are perfectly safe. Second, their incorporation as plasticizers during polymerization allows shorter cycle times in the polymerization reactor, lowers the melt viscosity during reactor handling and during compounding, thereby easing melt processing and preventing discoloration as well as maintaining molecular weight. The plasticization of poly(lactic acid)s provides a systematic attenuation of properties and a broad rang of compounds for various application. The basic tensile physical properties of a poly(lactic acid) and polyethylene are compared in Table 2. Both materials can be fine toned over a range of values for specific packaging applications. Tablr ‘ a litre Table 1. Some Basic Physical Properties Comparisons of Poly(lactic acid) and Polystyrene Poly(lactic acid) a Polystyrene ” Physical properties Oriented Unoriented Oriented Unoriented Ultimate Tensile Strengthc psi 14,700 6,900 7,400 7,015 MPa 101 48 51 48 Elastic Modulus 564,000 221,000 450,000 240,000 psi 3,889 1 ,524 3,103 1 ,655 MPa 15.4 5.8 4.0 - Impact Strength Ft—lbfrn - 0.44 0.4 0.24 MN/m - 23 21 13 Deflection temperature °F 1 81 127 200 - °c 83 53 93 - Specific Gravity 1.25 1.25 1.05 1.05 Melt Flow Rate 200°C - 46 - 3.5 1 55°C - 2 - - " 90/10 L-lactide/DL-lactide copolymer. b Crystal PS, Amoco R3. ° ASTM D882 Table 2. Comparison of Polyethylene to Poly(lactic acid) Properties Low density polyethylene Poly(lactic acid) ‘ Tensile Strength: Kpsi 2.1 8 3.19 MPa 1 5 22 Elongation, % 261 280 Tangent modulus Kpsi 54.9 36.6 MPa 379 252 100% modulus: kpsi 1.77 0.74 MPa 12 5 200% modulus kpsi 1.82 1.2 MPa 13 8 HDT 264 psi, °F 95 122 1.82 Mpa, °C 35 50 a L-Lactide plasticizer equals 19.5 wt% for lar; 5053671? these i $1 303 the cos are limi polgme DTDCESE lldmxy WlThOUg C The poly(lactic acid)s have some deficiencies that inhibit their acceptance for large-scale applications”.14 These deficiencies include high price, poor solvent resistance, insufficient physical properties, and various combinations of these factors. The selling prices of lactic acid polymers have been estimated at $1.0011b. Compared to $0.30-0.60 for polystyrene, poly(lactic acid)s are 2-3 times the cost of the conventional plastics. The major deficiencies of poly(lactic acid)s are limited physical properties. The technological development of biodegradable polymers is restricted presently by the range of these polymers that can fulfill processing and property requirements for many applications in which biodegradability would be an important materials property. Acceptance of biodegradable polymers will depend on five unknowns: (1) customer response to costs that are considerably higher than conventional polymers; (2) possible legislation (particularly water-soluble polymers); (3) the achievement of total biodegradability; (4) wider range of available physical properties; and (5) the development of an infrastructure to collect, accept, and process biodegradable polymers as a generally available option for waste disposal. 2. Synthesis of Biodegradable Polyesters Polyesters such as poly(lactic acid) can be synthesized by polycondensation from the hydroxy acid (Scheme 1). The polycondensation of hydroxy acids is generally performed by removal of water by distillation, with or without catalyst, and application of vacuum as the temperature is progressively increase mgmi moecuk polyconc Wm debnno dkmd Mend ogemng bwmde "”Q‘ODEr beam dmgm, hcemg Rir oirheSe muchras. mohOlTier. high 18mg ConCentra. The SISTEn increased.6v15 This rather simple step leads to oligomers with low molecular weight (<5,000 Daltons, Mw/Mn close to 2). There have been reports that high molecular weight poly(lactic acids) have been synthesized using polycondensation. The methods include postcondensation in organic solvents using dicyclohexylcarbodiimide (DCC),16 condensation catalyzed by distannoxane,17 and refluxing in high boiling point solvents such as diphenyl ether catalyzed by various metals and protonic acids.18 However, these methods have not been developed industrially. A better way to make high molecular weight polyesters is through ring- opening polymerization. As shown in Scheme 1, lactic acid was condensed to low molecular weight oligomers, the oligomer was cracked to the Iactide, and ring-opening polymerization of Iactide produced high molecular weight polymers. This is the method used to make poly(lactic acid) commercially.19 The drawback of this method is its high cost, which has been reduced tremendously by Cargill, Inc. using continuous processes to manufacture poly(lactic acid)s.20 Ring-opening polymerization can be carried out in bulk or in solution. Both of these methods have advantages and disadvantages. Bulk polymerization is much faster than solution polymerization because of the higher concentration of monomers and higher reaction temperature. High monomer concentration and high temperature also cause some problems. Because of high monomer concentration, a bulk polymerization system has a high viscosity, which makes the system difficult to stir. Because of concerns about viscosity and melting point oi the causes solutror scict'ror grader very lor p'bvrde disrnb'. such a amoun an ind. GT Their i0 mak of the monomers, bulk polymerizations are conducted around 200 °C, which causes some side reactions such as transesterification and discoloration. In solution polymerization, the monomers are dissolved in organic solvents, so solution polymerizations do not have the same viscosity and melting point problems as bulk polymerizations. Solution polymerizations can be conducted at very low temperature. Compared to bulk polymerization, solution polymerizations provide much better control over molecular weight and molecular weight distribution. However, solution polymerization uses large amounts of solvent such as toluene. Not only the solvents are expensive, but also dealing with large amounts of waste solvent is an environmental problem and costs money. From an industrial point of view, solution polymerizations are not economical because of their slow rates and the high cost of solvents. Thus, most industrial processes to make poly(lactic acid) use bulk polymerization. Scheme 1. Synthetic route for poly(lactic acid) Polycondensation: H 0 CH -C-COOH 330—» 3 I A O OH m Ring-opening Polymerization H O ' -H20 Waco... ____, EV} I A OH m low molecular weight oligomer l cracking O O {10)} 4 $0 m ring-opening O\n/i\ high molecular weight polymer polymerization 10 3.l 3.1 I mus‘ have act: PCiy exa: for ; nor Com war. of th for 3 lnCie 3. Ring-opening polymerization of Iactides and Iactones 3.1 Polymerizability For the ring-opening polymerization of Iactides and Iactones, one question must first be answered: are the monomers polymerizable or not? Many attempts have been made to correlate the ring-opening polymerizability of Iactones and Iactides with some basic parameter, such as ring strain, hydrolysis rate, or monomer basicity. No acceptable clear correlation has been found to date, but these attempts provide some insight into the polymerization process, and as a result some generalizations can be made. Whether a cyclic monomer can be polymerized to high molecular weight polymer depends on thermodynamic and kinetic factors of the polymerization. For a cyclic monomer to be polymerizable, the free energy of the polymerization AG must be negative under the conditions used. The effect of ring size on AG of polymerization can be readily illustrated by the use of the cycloalkanes as an example. The substitution of a heteroatom for CH2 does not significantly alter AG for polymerization as shown for various cycloalkanes, provided the heteroatom is not too dissimilar in size and bond angle, such as for O or N.21 Table 3 22 compares the heats of combustion per methylene group in these ring compounds with that of a methylene in an open chain alkane. This yields a general measure of the thermodynamic stabilities of rings of different sizes. The strain is very high for 3- and 4- membered rings, decreases sharply for 5-, 6-, 7- membered rings, increases for 8- to 11- membered rings, and then decreases again for larger 11 TEST oi: rem rings. Of course, while the presence of strain in a ring compound does not necessarily make it polymerizable, within a given class, the higher the strain, the more exothermic the polymerization, the more likely it is to be polymerizable. The thermodynamic feasibility does not guarantee the actual polymerization of Iactones. Polymerization requires that there be a kinetic pathway for the ring to open and undergo reaction at a reasonable rate. Table 4 lists the experimental results for the polymerizability of a large number of Iactones. It shows that none of the 5-membered Iactones investigated polymerized, and that substitution rendered some of the 6-membered Iactones unpolymerizable; however, hearty all of the 4-, 7-, and 8-, membered Iactones were polymerizable. Hall and coworkers23 investigated the postulate of Carothers that hydrolysis rate and polymerization tendencies should correspond with each other.24 However, they found this suggestion not to be generally valid for alkaline hydrolysis because B-propiolactone and y—butyrolactone hydrolyzed at comparable rates, while the former polymerized and the latter did not. The attempt to relate polymerizability of Iactones with basicities of Iactones by Lundberg and 00x25 also did not produce satisfactory results. 12 Table 3. Heats of combustion and Strains of Cycloalkanes per Methylene group n-member Heat of combustion per Strain per methylene cycloalkanes methylene group group (KJ/mole) (KJ/mole) 3 697.6 38.6 4 686.7 27.7 5 664.5 5.5 6 659.0 0.0 7 662.8 3.8 8 664.1 5.1 9 664.9 5.9 10 664.1 5.1 1 1 663.2 4.2 12 660.3 1 .3 13 660.7 1 .7 14 659.0 0.0 15 659.5 0.5 16 659.5 0.5 13 Table 4. Polymerizability of cyclic ester monomers Compound Ring size Polymerizability B-propiolactone 4 + B-butyrolactone 4 + y—butyrolactone 5 - y-valerolactone 5 - 8-valerolactone 6 + glycolide 6 + Iactide 6 + or-n-propyl-S—valerolactone 6 - tetramethylglycolide 6 - e-caprolactone 7 + 3-oxa-e-caprolactone 7 + cis-disalicylide 8 + di-o-cresotide 8 + trisalicylide 1 2 + + represents polymerizable, - represents unpolymerizable. 14 3.2 Pc 3.2.1 3.2.1.‘ laclbrr ‘vi 3ch 3.2 Polymerization Mechanism 3.2.1 Cationic Polymerization 3.2.1.1 Lactones The main cationic initiators used in ring-opening polymerization (ROP) of Iactones can be divided into four subgroups: (a) protonic acids (HCI, RCOOH, RSO3H, etc.), (b) Lewis acids (AICI3, BF3, FeClg, etc.), (c) alkylating agents (stabilized carbocations, e.g. CF3803CH3, Et308F4), and (d) acylating agents (eg. CH3COOCI3).26 In addition to these traditional acids and electrophiles, precursors of carbocations have also been considered, e.g., ammonium or phosphonium salts stabilized by complex counterions,27 which are transformed into the active species by thermal or photochemical processes. It is also worth mentioning diaryliodonium salts, which release an active cationic species upon reaction with a reducing agent.28 Many mechanisms have been proposed for the different cationic initiators. The polymerization mechanism proposed by Cherdron et 31.29 (Scheme 2) was accepted for a long time. It consists of an electrophilic attack on the endocyclic oxygen of the lactone and the subsequent rupture of the acyl-oxygen bond with formation of an acyl carbonium ion prone to propagate. Penczek et al. questioned Cherdron‘s mechanism. From chain-end analyses and kinetics data, they proposed another mechanism30r31 (Scheme 3), in which the electrophilic attack was on the exocyclic oxygen of lactone. 15 Kricheldorf et al.32 and Okamoto33 also proposed mechanisms for cationic polymerization, which are very similar to what Penczek has proposed. Scheme 2 Cherdron‘s cationic polymerization mechanism 0 o (\O Ké‘) R + ____> _> O 0 ll e-CL ll RO—(CH2)5 O + ———> R0— C —C C ( H2)s \Cii) Polymer 3.2.1.2. Lactides The secondary carbon atom in Iactide is less sensitive to nucleophilic attack compared to the primary carbon atom in lactones. At present, only triflic acid and methyl triflate initiate controlled cationic Iactide polymerizations; HBr and HCI also initiate cationic Iactide polymerizations, but yield low molecular weight polymers. There also are reports where Sn(ll) and Sn(lV) halogenides were used to polymerize Iactides, but it is believed that the actual initiating species are HBr and HCI, which are generated by Sn(ll) and Sn(lV) halogenides reacting with hydroxyl-containing impurities. 16 Scheme 3.The cationic polymerization mechanism proposed by Penczek. R— 05 o i 0 d} v/E‘) R + ——> L Ti 0 ' ’ RO—C—(CHQS— 0:0 ———’ Polymer A mechanism proposed by Kohn34 for the initiation and polymerization of Iactide with strong acids such as HBr and HCI is depicted in Scheme 4. Protonation at one of the two available carbonyl oxygens on Iactide is followed by a proton shift and an SN2-type transesterification reaction. In my opinion, this mechanism has flaws. The exocyclic oxygen is more nucleophilic than the endocyclic oxygen, so it is more reasonable that it should the exocyclic oxygen that attacks the carbonyl group, and not the endocyclic oxygen. Methyl triflate (gives very interesting results in Iactide polymerization. Kricheldorf found that the polymerization of L-Iactide gave polylactide that was 100% optically pure. If the polymerization proceeded with complete Walden inversion like traditional cationic polymerizations (Scheme 5), the resulting polymer must contain 50% D- and 50% L- units. The polymers should not be 17 Optic the lacti Scheme 4. Cationic polymerization of Iactide initiated by protonic acids optically active. Kricheldorf proposed a different mechanism35 (Scheme 6) for the methyl triflate initiated polymerization. The initially formed positively charged Iactide ring is opened by an SN2 attack of triflate anion accompanied by Walden inversion. The triflate ester end-group reacts with another Iactide again in an 8N2 fashion, so that a second Walden inversion occurs. The net result is perfect retention of the original configuration. Scheme 5. Polymerization of Iactide initiated by methyl triflate (traditional cationic mechanism) 0 I". S O CH: 9H: o-cre—o—co—cH—o—cr-r3 4»: R . S o’ ----- ——> 0 _..l\n/o O / Monomer Polymer 19 Scheme 6. Polymerization of Iactide initiated by methyl triflate (mechanism proposed by Kricheldorf) .OCH3 0’? ac“3 9H3 F3C—303‘ +/l\n/° __, Fac—SOZO—CH—fit—O—CH—CO—OCl-la 0 V o ‘5”3 9‘3 O + F3C— SOZO—CH— Ic— o— cr-r—co— OCH3 _> so . CH3 <1}ng t: o 2CH3 - 0’ Fae—sowo , F30— SOZOMéT/CHQ 0 CH3 & 20 3.2.2 Anionic Polymerization 3.2.2.1 Lactones Many kinds of anionic initiators have been used in the ring-opening polymerization of lactones. The most commonly used initiators are alkali metal alkoxides, such as potassium methoxide, and potassium butoxide. To improve the solubility and efficiency of the alkali alkoxides, mixtures of alkali alkoxides and crown ethers have been used as initiators. In addition to alkali alkoxides,36' 38 metal carboxylates,36 alkali metals, and cyclopentadienyl sodium39 also have been considered as initiators. The mechanism of anionic polymerization of Iactones depends on the ring size and substituents. There are two possible propagation routes in the anionic polymerization of Iactones (Scheme 7): (A) propagation with alkyl-oxygen bond scission, resulting in an active carboxylate center or (B) propagation with acyl- oxygen bond scission leading to an active alcoholate cente. e-Caprolactone (e-CL) polymerization is initiated exclusively by alcoholates. In the case of B—propiolactone (B-PL), the situation is much more complicated. According to Penzek et. al., B-PL can be initiated by alcoholates or carboxylates. When B-PL is initiated by carboxylate, the carboxylate ion is the only initiation species. When initiated by alcoholate ion, both carboxylate and alcoholate ions were produced. The two ions coexist during the initial period of polymerization. The carboxylate ions reproduce themselves every propagation step, but 50% of the alcoholate ions were changed to carboxylate ions at every 21 propagation step. After several steps, almost all the alcoholate ion vanished (Scheme 8). Scheme 7. Propagation routes for anionic polymerization of Iactones O A , \rla‘lzln ___) .W—(CHflmr—C/e J \‘0 o\ e (CH2)n __ ”('1‘ j" l —_s t/vw—E (or-r2),,,,——oe Scheme 8. Polymerization mechanism of B-PL. o i, A /’O W(CH2)2— C,\\e ___, W.(CH2)2_ c:\e A ‘0 o / / l .\ c c o B W—n—( Hzlz- 9 ——> fi—(CHQz—Oe 22 Alcoholate anions are much stronger nucleophiles than carboxylate anions. s—CL (AHp = -13.9 KJ/mol) is less strained than B-PL (AHp = -82.3 KJ/mol), so carboxylate anions only initiate polymerization of B-PL and can not initiate polymerization of e-CL. The alcoholate anions are stronger and smaller sized nucleophiles. Thus they initiate polymerization of both e-CL and B—PL, but are less discriminating. They attack at the carbonyl carbon and the carbon next to the endo-oxygen on the ring and thus both alcoholate and carboxylate ions exist during the polymerization of B-PL. Penczek’s mechanism37 seems to explain the experimental data well, but Jedlinski38r40 gave a different account about how the polymerization of B-PL proceeds. During the polymerization of B-PL, Jedlinski found unsaturated double bonds as the end groups and no traces of the initiator in the resulting polymers. This finding was supported by data from Dale and Kricheldorf“. From his data, Jedlinski proposed the mechanism shown in Scheme 9. The results above are very confusing. Both authors used the same initiator system, potassium methoxide to polymerize B-PL, but their results were totally different. The only difference in the experimental protocol was the solvent. Penczek used DMF and Jedlinski used THF. Maybe the polarity of solvents can explain the difference in the initiation of polymerization. 23 Scheme 9. The polymerization mechanism of PL proposed by Jedlinski 0 R0 010‘ R0 \ . RO'K“ —-> ———> KOH + \ll/\ (3 <3 0 f HO O'K+ - + KOH ____’ \n/\/ ___) *KO\n/\/OH 0 o - H:O/ 0 1 El Polymer Polymer O D: e. <-——— 0 3.2.2.2 Lactide The mechanism of anionic polymerization of Iactide is very complex. The polymerization conditions, such as structures of initiators, counterion, temperature, and solvent, have a large influence on the mechanism of polymerization. Krichedorf and Krieser-Saunders42 found that only strong bases such as potassium tert-butoxide (pKa = 18) and butyllithium polymerize Iactide and weak bases such as the benzoate ion (pKa = 5.7), and phenoxide ion (pKa = 9.5) were not able to initiate polymerization. In addition, polymer was racemized regardless 24 of the initiator, temperature and solvent. They also were unable to find the initiator fragment in the isolated polymer chain. From these facts, they proposed that initiation mainly involves deprotonation of Iactide (Scheme 10). Jedlinski et al.43 used potassium methoxide to initiate the polymerization of Iactide. They found the initiator fragment in the polylactide and very little raoemization. Thus, they proposed a different mechanism (Scheme 11). The methoxide anion attacks the carbonyl carbon in the monomer, with acyl-oxygen bond scission and formation of the methyl ester end group and the active Scheme 10. Anionic polymerization of Iactide through deprotonation 0 P 9’! o ’ O \n/g\ + RO‘ ———> ROH + 0% o O 25 alcoholate center. Propagation proceeds via acyl-oxygen bond scission and regeneration of alcoholate anion each propagation step. Scheme 11. Anionic polymerization of Iactide O . “he The contradictory results above may be caused by the polymerization conditions. Krichedolf et al. used a tertiary alkoxide (potassium tert-butoxide) and Jedlinski used a primary alkoxide (potassium methoxide). The difference in the nuclephilicity of anions might explain the difference in the initiation mechanism. The groups also used different solvents. The polarity of solvent might play a role in the initiation mechanism. Anionic polymerizations of Iactides and Iactones are very fast processes, but side reactions such as racemization and back-biting make it difficult to obtain high molecular weight polymer with desired properties. It is unlikely that anionic polymerization of Iactides and Iactones will have important industrial applications. 26 3.2.3 Coordination Polymerization As previous discussed, the anionic polymerization of Iactide and Iactones have side reactions such as transesterification and racemization. These undesired side reactions result from the high reactivity of the alcoholate. A decrease in the reactivity, eg. by modification of the counterion, is a possible way to eliminate, or at least delay these side reactions. During the last several decades, many research groups have worked on initiating polymerization using the alkoxides and organometallic derivative of different metals, aluminum,44r45 tin,46v47 zinc,48r49 iron50 and some rare earth metals such as yttrium,51 lanthanum.52 The most frequently used metals are aluminum and tin. 3.2.3.1 Aluminum Derivatives The most widely used aluminum initiators are aluminum triisopropoxide37, and soluble bimetallic u-oxo-alcoholates, expecially zinc and aluminum53 (Figure 2). Y Y 04 \Al-—O\ DEAF? T6 Zn\o’A( o T )1 Figure 2.Structure of Al initiators 27 Aluminum alkoxides polymerize Iactide and Iactones through a “coordination-insertion” mechanism,54‘56 which involves coordination of initiator with the exo-oxygen of the monomer and insertion of the monomer into an Al-O bond of the initiator, follwed by cleavage of the acyl-oxygen bond of the cyclic monomer (Scheme 12). Scheme 12 Mechanism of Iactide polymerization using Al alkoxides 28 The polymerization of Iactide and Iactones using Al alkoxides is a living polymerization. If all impurities are excluded, there is no termination reaction and the molecular weight of polymer increases linearly with increases in the conversion of monomer to polymer. The polymerization is first order in concentration of monomer and concentration of initiator, and has the kinetic expression shown below. —M = kperrIr dt Polymerization kinetics are influenced by polymerization conditions such as monomer and solvent.57 When using aluminum triisopropoxide to polymerize e-caprolactone (e-CL), only one of three alkoxide groups initiates polymerization. In polymerization of Iactide (LA), all three alkoxide groups initiate polymerization. This difference is caused by aggregation of the AI(OrPr)3 in the solvent. AI(OrPr)3 has two aggregation states in nonpolar solvents: a tetramer and a tn’mer, called A4 and A3 respectively.53'52(Figure 3) The exchange rate between the tetramer and trimer is small, and almost can be neglected. The two aggregates react with s-CL with different rates. Therefore, when a mixture of A4 and A3 is used to initiate the polymerization e-CL, A3 is consumed completely whereas A4 remains unreacted. This observation explains why it is assumed that only one alkoxide group from Al(OrPr)3 participates in the polymerization of s—CL. Actually, an IVA mixture was used in such a ratio that on average, one alkoxide of the initial amount of AI(OIPr)3 initiates polymerization. A3 provides all of the alkoxide groups, while A4 is inactive. Lactide is less reactive than s-CL, which gives A4 29 enough time to dissociate before the monomer polymerizes completely. Furthermore, monomer addition causes the aggregated alkoxides to dissociate to a trisolvated six-coordinate “AI(OIPr)3 - monomer” complex. The only difference between s—CL and Iactide polymerization is that for s—CL, only A3 dissociates and for Iactide, both A3 and A4 dissociate (Scheme 13). l O———Al OR OR I 74%“ O 0 RO— N/Fl l2\\Ar—0R RO A3 OR Figure 3 Aggregation states for AI alkoxides 30 Scheme 13. Dissociation of the AI alkoxide aggregates / \ A3 > R0 l OR —> Polymer .? + .P O l T O 31 The use of different solvents does not modify the average number of active alkoxide groups per AI, but according to Teyssie,53 the polymerization rate in THF is significantly smaller than that in toluene. When s—CL is polymerized, the half polymerization time 0112) is 3.5 min in toluene and 8.7 min in THF ([s-CL]=1 moi/L, [M]o/[l]o=200 and T = 0°C). The reason for the rate decrease is that THF is a much better solvating agent than toluene. The ether oxygen of THF is able to compete with the carbonyl group of Iactones and Iactide for coordination to aluminum. Because the ether oxygen on THF is more nucleophilic than the carbonyl group, it makes the formation of AI(OrPr)3o(lactone)3 unfavorable. AI(OIPr)3 is a very clean initiator for lactones. The molecular weight increases linearly with conversion and the molecular weight distribution is very narrow. But this linearity is not observed at very high conversions, or at high polymerization temperatures. The deviation from linearity is caused by transesterification reactions.55 Transesterification reactions can be intramolecular (“back-biting” Scheme 14) or intermolecular. Backbiting produces a shorter polymer chain and a cyclic oligomer from original polymer chain, and is responsible for the broadening of the molecular weight distribution and for the decrease in the number-average molecular weight. lntennolecular transesterification (Scheme 15) does not change the number-average molecular weight of the polymer, but it does increase the polydispersity of the polymer. 32 Scheme 14. lntramolecular transesterification Y VS 4 ..__._ 013%. . We ..__. / \OR 0:2}. 9M2}; cyclic °"90mer 33 Scheme 15. Intermolecular transesterification The rate of transesterification depends on the temperature, and the kind of metal center in the initiator. Kricheldorf et al.64 compared the transesterification activity of different kinds of metal alkoxides. He found that aluminum alkoxides are least likely to cause transesterification and the tin alkoxides are most likely to cause transesterification. The order for transesterification activity is AI(OIPr)3 < Zr(OnPr)4 < Ti(OnBu)4 < Bu3SnOMe < Bu26n(OMe)2. Polymerizations of Iactide and Iactones initiated by aluminum alkoxides are perfect living polymerizations and the alkoxide groups are the end-groups of the polymers. If the alkoxide group contains some functionality, the polymers will have a functional end-group. Halogens, tertiary amines, carbon-carbon double bonds, and methacrylates are typical example of functional groups that can be placed at the chain end. The initiators are easy to synthesize by mixing Al(Et)3 with the corresponding alcohol (Scheme 16). By controlling the amount of Scheme 16. Preparation of end-functioned AI initiators AllEt)a+ n HOCH2CH2X —-> StenAIrocrrzcrrzxrn 1sns3 0 9hr M . Et3-nAI(OCH2CH2X)n+/Kl( _> O O n O 0 ll X= Br, -CH=CH2, ~NEt2, -OCCH=CH2 35 alcohol, the structures of the initiators can be controlled. End-functionalized polylactides and polylactones have been used in many fields. Teyssie et al.55‘69 used them to synthesize crosslinkable polylactides and polylactones, graft copolymers and star-branched polyesters. The great versatility of end- functionalized aluminum alkoxides opens the way to the macromolecular engineering of polylactides and polylactones. This approach will dramatically increase the range of the physical properties for polylactide and polylactones. lnoue et al. reported that derivatives of tetraphenylporphinato—aIuminum are efficient and versatile initiators for the living polymerization of Iactones and Iactide. The polymerization mechanism agrees with a “coordination-insertion“ mechanism that involves acyl-oxygen cleavage. The polymerization rate can be remarkably increased by addition of sterically hindered Lewis acids. Coordination of the Lewis acid to the carbonyl oxygen of lactone or Iactide monomers makes them prone to nucleophilic attack. These porphyrin derivatives polymerize not only Iactones and Iactide (Figure 4, X=OR),70'72 but also epoxide 73:74(X=Cl), methacrylates,75 and methacylonitrile76 (X=SR, R=alkyl). The corresponding copper and iron complexes were investigated by Kricheldorf et al.77 The iron or copper porphrin complexes polymerize Iactide but produce relatively low molecular weight polymer in low yield. X = OR, Cl, Alkyl, SR, OZCR Figure 4. The structure of porphyrin initiators 37 3.2.3.2 Tin Derivatives. Many kinds of tin compounds have been used in the polymerization of Iactide and Iactones such as SnO,78 SnOz,79 SnX4, San (X = Br, Cl),80'82 Sn(OR)2. Sn(OR)4,83»86 and (RCOO’)28n.46v47337'39 The mechanism for polymerization using tin compounds is very controversial, and anionic, cationic and coordination mechanisms have all been proposed. Because of a lack of experimental evidence, none of the mechanism is very convincing. Kricheldorf90:91 investigated tin oxides and oxides of many other metals such as Mg, Sb, Pb and Ge for L-Iactide polymerization. These catalysts are heterogeneous, so the repeatability of polymerization is poor. Also, these catalysts cause monomer racemization and transesterification. There is no detailed mechanisric study of the metal oxide catalyzed polymerizations, but people believe that hydroxyl-containing impurities such as water and alcohol initiate the polymerization. Tin alkoxides perform just like aluminum alkoxides. The polymerization mechanism involves monomer coordination to the tin alkoxide, followed by insertion of monomer into the Sn-O bond (Scheme 17). Tin alkoxides are very efficient initiators. They do not cause racemization of monomer up to 150 °C, but since tin alkoxides are good transesterification catalysts for noncyclic ester groups, they may cause extensive back-biting degradation of polylactones and polylactide at elevated temperatures. The loss of molecular weight control is the reason that tin alkoxides are not as widely used as the aluminum alkoxides. 38 Scheme 17. Mechanism for tin alkoxide initiated polymerization Iactides O Bu Sn 0 —-> 3 \O/i\n/ TKOR O polymer Tin halides such as SnClz and SnBrz are very good catalysts, and high molecular weight polylactide and polylactones have been obtained using tin halide catalysts. Because they are not soluble in organic solvents and monomers, the repeatability problem still exists. The proposed mechanisms for lactone and Iactide polymerization using tin halides are very different. The mechanism of lactone polymerization is shown in Scheme 18. It includes coordination of acidic tin halides with the carbonyl oxygen, ring-opening by transfer of a halide atom from the catalyst to the (lb-carbon, followed by cleavage of the alkyl-oxygen bond. The propagation can proceed through two routes: through the carboxylate (A) or through alcoholate (B). The exact propagation species has not been identified. Lactide polymerization using tin halides acts much different. Unlike lactone polymerizations, halide end groups have not been found by NMR and elemental analysis. Lactide polymerization is not initiated by 39 transfer of halide atoms, probably because the secondary carbon on the Iactide is less sensitive to nucleophilic attack than the primary carbon of a lactone. Based on the above observations, Kricheldorf proposed that tin halides react with impurities such as alcohol or water to first form tin alkoxides, which then follow the coordination-insertion polymerization mechanism. Scheme 18. Mechanism for tin halide initiated polymerization of caprolactone SnBr4 + o ——> O BraSn\ O/U\ /Br A O 0 : (CH2)5 Br I ——> Br3Sn\ / O> (CH2)/ (CH2)5 ([3 l Polymer O l O Br ~. /u\ /Br 0 385' ° (CH2)5 B /o\ kO/lk /Br 50> Brasn (CH2)5 (cg-(2)5 40 Tin(ll) 2-ethylhexanoate (Sn(Oct)2) is the most widely used catalyst for Iactide polymerization. Commercial polylactide is synthesized using this catalyst since it provides high reaction rates, high conversions, and high molecular weights even under rather mild polymerization conditions. There are extensive studies of this catalyst system, but the mechanism is still unclear. Nijenhuis et al.92:93 proposed the cationic polymerization mechanism shown in Scheme 19. This is a complicated mechanism, but it is unlikely that the polymerization follows this path. First, the dissociation of Sn(Oct); in a non-polar environment is not energetically favorable. Second, formation of an eight membered ring during polymerization should also be unfavorable. Finally, the strong interaction between Sn and Iactide prior to reaction with the OH group is energetically unfavorable. To date, there is no spectroscopic evidence for any of the intermediates in the mechanism. 41 Scheme 19. Mechanism for polymerization of Iactide initiated by Sn(2-ethylhexanoate); as proposed by Nijenhuis et al. s..- 331 __. 14371;; H\ %‘H \xSn I Sn I R l Oct Oct \(lko o , ——> H/X\ ——> 0 9H /S_/ S" O i) R T" R0 Oct f} o O H o 0 OH H 0 OH fiO—Sn-u-O i0 <———> R0 bet 2&0 /¥0----Sn—o 3&0 RO OH _. #0 Hob—{2% .... 42 Kricheldorf et al.94 proposed the mechanism shown in Scheme 20. The first step consists of coordination of alcohol or water to Sn(Oct)2. The complexation of Iactide to the intermediate polarizes its carbonyl group, making it susceptible to nucleophilic attack from alcohol. This mechanism also has a shortcoming. Based on the assumption that the tin catalyst is active via free sp3d2 orbitals, Sn(Oct)2 can coordinate with both alcohol and Iactide. But this mechanism can not explain the activity of Bu28n(Oct)2, which is six coordinate and should not be active. In past, Bu28n(Oct)2 polymerizes Iactide, though not as well as Sn(Oct)2. The authors rationalized their results by treating the Oct group as a monodentate ligand, but it is well known92 that Oct is a bidentate ligand, and Sn(Oct)2 and Bu28n(Oct)2 have the structure shown in Figure 5. Penczek et al.95 proposed that Sn(ll) alkoxide species initiated the Iactide and lactone polymerizations shown in Scheme 21. In their work, they identified intermediates such as RO[M],,OSnOct. This mechanism is very similar to what we propose based on our experimental work. The details will be discussed later. 43 Schr Raf I R Scheme 20. Mechanism for polymerization of Iactide initiated by Sn(2-ethylhexanoate)2 as proposed by Kricheldorf (if T“ 'T1 33 )T 2,.“ RZ—Sn—R‘ + R3OH ———> R 2?"“0—H ———> a: R: 2 R3 H O 1 1 R2! 3 ”6—H R2. 2 I n’ 'u. n___,0 o\ 3 Rl’l'R V‘ 0 o ——> [21’le R O O O ___.___) Polylactide Scheme 21 Mechansim for Iactide polymerization initiated by Sn(2-ethylhexanoate); as proposed by Penczek et al. 0 \ lulu“ O Olrunnflsn“\ + ROH O O O 0 O \ + _9_, \ O\R Ounnll“sn_OR On||||"“sn 0 2n 45 O Bu 0 \ I |I|||\O \ ““‘O O||IIHI|||SnQ|H‘ 09|I||I|I||Sn‘|\\\| I \ Bu 0 O Solon); Buzsrmm)2 Figure 5. Structure of Sn(Oct)2 and Bu28n(Oct)2 There are some other proposed mechanism when Sn(2-ethylhexanoate): is used as catalyst such as the one by Vert et al39., but there is no published mechanism that can explain every experimental fact. More research is needed to fully understand this reaction. 3.2.3.3 Rare Earth Metal Compounds The 15 Ianthanide elements (La, Ce, Pr, Nd, Pm, Sm, Eu, Gd, Tb, Py, Ho, Er, Tm, Yb, and Lu) together with Sc and Y elements are called rare earth elements. They prefer tri-valent oxidation states in complex formation. Rare earth metal complexes are known to initiate the living polymerization of methyl methacrylate over a wide range of polymerization temperatures to give high molecular weight syndiotactic polymers with extremely narrow molecular weight distributions. Lactones and Iactide have also been successfully polymerized 46 using rare earth halide systems and rare earth complex systems. Yasuda et al.96-98 used [SmH(CsMe5)2]2 and LnMe(C5Me5)2(THF) (Ln=Sm, Yb, Y, Lu) to polymerize Iactide and lactones. Also, Feijen99r100 thoroughly studied the polymerization of Iactide and lactone using yttrium alkoxides. These catalysts initiate polymerization via a “coordination-insertion” reaction. The mechanism (Scheme 22) is very similar to that of the aluminum alkoxides and tin alkoxides. Shen et al.101'106 found that rare earth halides can be used to polymerize Iactide and lactones. He also found that the polymerization rate was dramatically increased by adding epoxides. The halides first react with expoxide to form the alkoxides, (Scheme 23) which are much better initiators than halides. Because of their high activity and few side reactions, rare earth catalysts will have a bright future in Iactide polymerization. 47 Scheme 22. Polymerization of Iactide initiated by rare earth catalysts. Me OOTL /%Ln‘/ __> \ O “<30 \(KO fl YLO O j/ko “is __, 5:51) “by... CEO This rife Polylactide Scheme 23. Initiators made from epoxides and rare earth halides o C) R + erx3 __, X3,,,Ln[(OCHCH2)y]X x = Cl, Br, I; R = H, CH3, CHZCI, CHZOCHZCH= CH2 48 3.2.4 Other Polymerization methods In addition to the polymerization methods mentioned above, other polymerization methods such as free-radical polymerization, active hydrogen polymerization and zwitterionic polymerization have been used with Iactide and lactones. Free radicals are usually ineffective for the polymerization of Iactones and Iactide. Molecular weights are often low, and monomer conversion limited,107 so there are few examples of radical polymerization of lactones. The polymerizations of Iactones and Iactide initiated by active hydrogen compounds, e.g., amines or alcohols, are relatively slow processes, and usually produce polyesters of low molecular weight.108 There are examples of the polymerization of Iactones initiated by ethanolamine or ethylene glycol, but polymerizations typically need 3-4 days to reach high conversion even at 180 °C. In contrast to the slow polymerization of Iactones initiated with active hydrogen compounds, zwitterionic polymerizations are fast processes that give high yields and high molecular weight polymers. Wilson and Beaman109 showed that polymerization of pivalolactone initiated with strained cyclic amines is a zwitterionic process (Scheme 24). Initiation corresponds to formation of a zwitterion as a result of nucleophilic attack by the amine at the lactone methylene group. The carboxylate anion of the zwitterion is the propagating species. Zwitterionic polymerization usually only occurs with very strained Iactones, such as B-PL. There are also some claims that the polymerization of e-CL proceeds through a zwitterionic mechanism when initiated with aniline in the presence of a protic acid. 49 Scheme 24. Zwitterionic polymerization mechanism 4;. . >$ __, 9:292:09 L Q~W%& 50 4. Modification of Polylactide Properties To be successful in areas such as packaging, polylactides must be available that exhibit a broad spectrum of physical properties while retaining the degradability of the parent polymer. Typical approaches used to modify the physical properties of polylactides include manipulation of stereochemistry, copolymerization and formation of blends. 4.1 Manipulation of stereochemistry. Lactide exists as three diastereomers: SS, RR, and RS (Figure 6). The SS diastereomer is referred to as L-lactide, the RR as D-Iactide and the RS as mesa-Iactide. An equimolar ratio of L—lactide and R-lactide is referred to as rac or D, L—lactide. Polymers prepared from different Iactides have very different physical properties. Polymers of high purity L-lactide or D-lactide are crystalline with melting points around 180 °C. The polymers of rec-Iactide are amorphous with a glass transition temperature ranging from 22-65 °C.110 The different properties are related to the tacticity of the polymers. The polymer can have three limiting kinds of tacticity: atactic, isotactic and syndiotactic (Figure 7). If all the Stereocenters on the polymer chain have the same configuration such as RRRRRRRRR or 88888888, the polymer is called isotactic. A syndiotactic polymer structure occurs when the configurations of the sterecenters alternate 51 from one repeating unit to the next such as RSRSRSRSRS. If successive stereocenters are randomly distributed, the polymer is called atactic. L-Iactide and D-lactide are known to polymerize to isotactic crystalline polymers while rec-Iactide polymerizes to atactic amphours polymer. By manipulating the stereochemistry of polymer, the properties of the polymer can be altered. For example, poly(L-Iactide) degrades slowly because of its highly crystallinity; by copolymerizing a small amount of rec-Iactide with L-Iactide, the degradation rate of the resulting polymer is much faster.111r112 The added rac- lactide introduces R stereocenters into the polymer backbone, which act as defects and interrupt the crystallinity of the polymer. Thus, the degree of crystallinity of the polymer is smaller, which results in a faster degradation rate. By controlling the amount of rec-Iactide added to L-Iactide, the rate of degradation can be controlled. Poly(L-lactide) has a melting point around 180 °C, which makes poly(L-Iactide) very hard to process without discolorization and loss of molecular weight. To solve this problem, a small amount of rec-Iactide is often copolymerized with L-Iactide to decrease the melting point. Tsuji et al.113'122 found that mixing an equimolar amount of poly(L- Iactide) and poly(D-lactide) produces a polymer with a melting temperature and crystal structure different from poly(L-lactide) and poly(D-lactide) homopolymer. For example, the homopolymer of poly(L-lactide) has a melting point of about 180 °C, but the mixture melts near 230 °C. The mixture of poly(L-lactide) and poly(D-Iactide) is termed a stereocomplex. The crystal structure of the stereocomplex has poly(L-lactide) and poly(D-lactide) chains packed side by side 52 in a 1:1 ratio of D and L monomer units. Because poly(L-lactide) and poly(D- Iactide) pack better together than either packs with itself, the racemic crystallites are more stable than crystallites from the homopolymer. O O 0 fl huh.¢ $0 0 O O (3. 3) (R. R) (R: S) L-lactlde D-lactide meso-Iactide Figure 6. Stereochemistry of Iactide 53 \erwrsee' isotactic \ir"rb MezAlO—(CI-I20H20)n— Cl'lzCI'EOMe o o ‘zzo ° 0 O , MezAI/E fio— (CH20H20),,— CHZCHQOMe (1) sn(OCt)2 o l-lO—(CH20H20)n—CHZCHQOH + oQ‘xo ____, o o 4"be O—(CH20H20)n— crecreo\[\n/L 3'21 (2) m 0 0 m Graft copolymers of Iactide and Iactones are rare compared to random copolymers and block copolymers. People mainly focus on using polyols such as poly(vinyl alcohol),150 and polysaccharides (pullulan, amylose)151r152 as initiators to polymerize Iactide and lactones. Polyols are hydrophilic, and polylactide and polylactones are hydrophobic. Combining these two segments in 60 a copolymer will provide materials very special properties such as micelle formation in solvents, which can be used to make drug delivery devices. In addition to synthesizing copolymers, the crosslinking of homopolymers and copolymers of linear polylactides and polylactones can be used to modify the physical and mechanical properties of these materials. By introducing crosslinks into polylactide, physical properties such as the crystallinity, the melting point and the glass transition temperature will be influenced. The characteristics of the degradation by hydrolysis of these biodegradable polymers will also be influenced. Three methods that have been used to crosslink polylactide and polylactones: high-energy electron beams, peroxide-induced radical crosslinking and copolymerization of functional Iactide and lactone oligomers with a multifunctional monomer. High—energy electron beams significantly degrade polylactide and polylactones, so it is not a suitable technique for crosslinked polylactide and polylactones. Peroxide crosslinking of polylactide and polylactones appears to be an effective method for affecting the thermal as well as the mechanical properties. However, in terms of the potential biomedical application of these materials, the method has two disadvantages. First, peroxides modified polylactide and polylactone chains may have undesirable degradation products. Secondly, many peroxide compounds are toxic, so crosslinked materials will need to be thoroughly extracted to remove peroxides, which is very costly.153 61 Copolymerization of functional Iactide and lactone prepolymers with other monomers is the most successful cross-linking method. As shown in Scheme 27, the multifunctional prepolymers with end-groups such as methacylate were synthesized and cured with monomers such as styrene and methyl methacrylate to form networks.154-158 Scheme 27. Synthesis of cross-linkable copolymers 0 Ho / \_ + \'/lko Sn(Octrz 4 °” WK HO M071 fl? M “W 071 flfi 62 4.3 Blending Copolymerization is an effective way to modify the properties of the polymers. However, copolymers are often hard to make and are very expensive. Blending of polymers might offer a more cost-effective way to modify polymer properties compared to copolymerization. Blends oftem have inferior properties compared to copolymers, but their ready availability and low-cost are very attractive. Blends can be made through thermal mixing and solvent mixing. In thermal mixing, the components of the blend are melted, and mixed thoroughly to make blends. In solvent mixing, the components of the blend are dissolved in a common solvent, and evaporation of the solvent yields the blend. Macroscopic properties such as impact and tensile strength, and degradation behavior can be modified by a reasonable choice of the blend components. The final properties will depend not only on the chemical composition of the blend but also on its physical characteristics, such as glass transition temperature, crystallinity and morphology, which, in turn, are a direct consequence of the compatibility between the components in the blend. The miscibility is the single most important factor for blends. If the components are miscible, the blend is a homogeneous system. If the components are immiscible, the blend will form a two-phase or multi-phase system. Most blends are phase- separated, which hinders applications of the blends. There are many reports of blends of biodegradable polymers. For instance, Langer et al.159 used poly(lactide)/pluronic blends as protein-releasing 63 matrices. Eguiburu et al.160 blended amorphous and crystalline polylactides with poly(methyl methyacrylate) and poly(methyl acrylate). The resulting blends are miscible, and have interesting thermal properties. Poly(L-Iactide) has also been blended with poly(e—caprolactone) and other polymers.161'165 5. Degradation of the polylactide and polylactones There are two ways to degrade polylactide and polylactones. They can be degraded thermally, and they can be degraded environmentally which includes hydrolytic and enzymatic degradation. 5.1.Thermal degradation Understanding thermal degradation is very important for polymer processing and polymer recycling. Poly(L—lactide) has a high melting point, so processing steps such as injection molding and extrusion must be done at around 200 °C, a temperature where polylactide and polylactone begin to degrade. Degradation can be prevented if we understand the mechanism for thermal degradation. Polylactide and polylactones also can be recycled by selecting favorable thermal degradation pathways. Transesterification, cis-elimination and radical reactions165'172 have been proposed as thermal degradation mechanisms. The transesterification reaction was discussed earlier. The intra-molecular transesterification reaction generates volatile low molecular weight oligomers, which causes the polymer to degrade. The mechanism of the cis-elimination reaction is shown in Scheme 28. The elimination of a proton from methyl group produces 1-alkenes and carboxylic acids. The carboxylic acids also accelerate the degradation of the polymer. The proton on the methyl group is not very labile, so the cis-elimination reaction is lilely only at high temperatures, Usually the cis-elimination reaction cannot compete with the thermal degradation of polylactides by transesterification. Radical reactions (Scheme 29) can start with either alkyI-oxygen or acyl-oxygen homolysis. Several types of oxygen and carbon-centered macroradicals may be formed. From the radicals, a varietiy of volatile compounds can form, such as carbon dioxide, carbon monoxide and aldehydes. The thermal degradation can be influenced by many factors such as residual metal compounds and unreacted monomers. Residual metal and monomers increase the degradation rate greatly. Scheme 28. Thermal degradation via cis-elimination on 0. CH3/ \W‘sH OR C 37 ' clr 0 // ”2 \ CCHZ -——> CH3/ \¢ + crl C OH COOR OOR 65 Scheme 29. Radical thermal degradation «W @3va ** carbonyl-radical carbon-radical 1 CO, CO), aldehyde, cyclic oligomers... i acyl-racidal oxy-radical “OM °w**—» eds. age/k 5.2. Blodeg radation Biodegradation mechanisms for Iactide include hydrolytic degradation129v136r173‘175 and enzymatic degradation.176'178 The two degradation processes are very different as shown schematically in Figure 9. Enzymes are high molecular weight molecules, so it is very difficult for enzymes to diffuse into polymers. Thus, enzymatic degradation almost always happens at the surface of the polymer. The top half of the Figure 9 shows the weight loss and molecular weight loss as a function of time for enzymatic degradation. Because degradation is limited to the surface of the polymer, the surface of the polymer is eroded, but interior of the polymer is unchanged. The weight of polymer drops as polymer chains on the surface are degraded to soluble oligomers, but the molecular weight of the polymer remains almost constant until most of the polymer weight is lost. In contrast to enzymatic degradation, the Figure 9. Degradation models Weight Molecular weight Time Time Enzymatic Degradation Weight Molecular weight FF Time Time Hydrolytic Degradation 67 molecular weight drops continually and the polymer weight remains constant during hydrolytic degradation. Hydrolytic degradation ocurrs in the bulk of the polymer. Water molecules diffuse into polymer and hydrolyze the ester bonds of polylactide, causing the molecular weight of the polylactide to drop. When the molecular weight drops to where the polymer becomes soluble, the weight of polymer begin to drop. Hydrolysis of the ester bond produces a carboxylic acid, which catalyzes further degradation of the polylmers. This phenomenon is called autocatalysis. Temperature, pH of the degradation medium, the stereochemistry of the polymer, the morphology of the polymer and the physical size of the polymer all influence the degradation rate of the polymer. Because water diffuse at a slower rate in crystalline polymers, and has a lower solubility due to the higher density of chains in crystalline polymers, crystalline polylactide degrades slower than amorphous polylactide. Enzymatic degradation is sensitive to the . stereochemistry of the polymer, and enzymes often selectively degrade one isomer. The degradation process is complicated, and there are many controversial conclusions. Many people question the role of enzymes in polylactide degradation, since hydrolytic degradation always accompanies the enzymatic degradation. Much work needs to be done in order to fully understand the degradation process. 68 RESULTS AND DISCUSSION 1 . Monomer synthesis In order to synthesize substituted polylactides through ring-opening polymerization, a series of substituted glycolides, shown in Figure 10, were prepared. For simplicity, we used the glycolide structure as a template to name these compounds. Substituted glycolides have been known for over a century, with the synthesis of ethylglycolide via the dimerization of the sodium salt of 2- bromobutyric acid reported as early as 1893.179 We prepared substituted glycolides using the two routes outlined in Scheme 30. The first route mirrors the standard method used to prepare the Iactide dimer. The appropriate or-hydroxy acid was first condensed in toluene to give low molecular weight oligomers, and cracking the oligomers under reduced pressure in the presence of a transesterification catalyst such as ZnO directly yielded the volatile dimer. This method has been used industrially to produce polylactide.19 In the second route, an or-bromo acyl bromide was condensed with an or-hydroxy acid to form an ester,180 followed by ring closure to give the cyclic dimer. In this method, it is very important to remove the N(Et)3 in the work-up. Even trace amount of N(Et)3 can change dimers to low molecular weight oligomers in a short period of time. Because N(Et)3 is not nucleophilic, it is likely that Et3NHBr caused the polymerization. The first method gives the dimer in higher yield, but the second method offers more flexibility in synthesis and can be used to synthesize unsymmetrical dimers, which are discussed in a later section. 69 Both synthetic methods yield a mixture of diastereomers. Thermal cracking of oligomers consistently yields a near-statistical mixture of the R,S and R,R/S,S diastereomers. 1H NMR of a representative example, that of ethylglycolide, is shown in Figure 11. The methine protons of the 3,6- disubstituted glycolide ring appear as a doublet of doublets near 5 4.85. The signal for R,S isomer appears at 4.88, while that from the RR and 8,8 isomer is at 4.83. The near equal intensities of the peaks confirms the 1:1 mixture of R,R/S,S and R,S diastereomers. Similar results were obtained for isobutylglycolide and hexyglycolide prepared by the thermal cracking route. Non-statistical mixtures were obtained when the second route was used. The R,R/S,S isomers are about 70% of the total and the R,S isomer is about 30%, which is probably is related to the kinetics of diastereomer formation. It is likely that the R,R/S,S isomers form faster than the R,S isomer. In addition, n'ng- closing competes with oligomerization to form the linear dimer, trimer and tetramer. Because of the slower rate for forming the R,S isomer, oligomerization is probably favored to form linear oligomers rather than S,R glycolides. 70 fits rev Lactide Ethylglycolide 3,6-dimethyl-1 ,4-dioxane-2,5-dione 3,6-diethyl-1 ,4-dioxane-2,5—dione O O O Isobutylglycolide Hexylglycolide 3,6—diisobutyI-1 ,4-dioxane—2,5-dione 3,6-dihexyl-1 ,4-dioxane-2,5-dione Figure 10. The structures of substituted glycolides Scheme 30. The synthetic route to substituted glycolides O O HO p-TSOH. toluene I-IO H OH * O A R R n oligomer ZnO A O Br\'/ll\ O Br R > O RN 0 R : CH3. CH2CH3. CHZCH(CH3)2 (CHzlscHs 72 Iee_.oo>_m_§ex_oeoec oo__oo>_a_>£o Lo «:22 I. .: 2:9". add me o; m; ozs. m.~ o.m m.m o.e We 06 m.m PhhpLhLlplrhrlpp pbpb—pphbblbhh——__pr—hhlhthlrlFFlbbllhrpb—ppbb—»LI»P Edd Se 2:. £4 84 3.3 LFPFFLtht»pbb—hbbphhpbrphht—prrh 0 see me o; m; 9N m.m o.m m.m o.e me 9m m.m —~F>pr—rhblrppp—bFPF—hplubbrrbrbbbprph—pppprLrp»plphpphb»blhllerbph f< o J..— I add 09m 36 31m .I/ .I#...I.I\ I... \.f 73 2. Bulk polymerization of substituted Iactides Bulk polymerizations are solvent-free polymerizations, which are usually run at temperatures higher than the melting point of the monomer. Because of the high concentration of the monomer and high polymerization temperature typically found in bulk polymerizations, the bulk polymerization rate and polymer yield is very high. However, because of the high polymerization temperature, bulk polymerizations are usually accompanied by side reactions which limit control of the polymerization. In contrast, in solution polymerization the monomers are dissolved in organic solvents and the polymerization can be conducted at very low temperatures. Compared to bulk polymerization, solution polymerization provides better control over molecular weight and molecular weight distribution. For example, Iactide has been polymerized using both bulk polymerization and solution polymerization. The bulk polymerizations are usually run from 110 °C to 180 °C and polymerizations finish within several hours. Side reactions such as transesterification, racemization and discoloration always accompany polymerization, and thus the control over molecular weight and molecular weight distribution is poor. Solution polymerizations are usually run from 50 °C to 90 °C and polymerizations may take days to finish. Solution polymerization provides excellent control over the molecular weight and molecular weight distribution of the polymers. 74 Zilnni hdpe ofidesi variety WOWdr showr bmad POWnr show hexyl Oflen conv 533k mole 265 mor 90h bfler aye fro, 17‘ 2.1 Initiator survey. Many compounds have been used as catalysts in bulk polymerizations of Iactide including metal halides,82.181.182 metal carboxylates,46 metal oxides50»78 and many kinds of organometallic compounds.183v184 We tested a variety of catalysts for their ability to polymerize substituted glycolides. 1H NMR provides a convenient method for following the polymerization reaction. As shown in Figure 11 for ethylglycolide, the methine peak at 4.85 evolves into a broad peak at 5.05 ppm during polymerization, and thus the conversion to polymer can be calculated by integration of the two signals. Tables 5, 6, and 7 show results from the polymerization of ethylglycolide, isobutylglycolide and hexylglycolide respectively. As shown in Table 5, runs without alcohol initiators often gave near-quantitative conversion to polymer, but SnO and SnBrz gave low conversion and low molecular weights. Both SnO and SnBrz are heterogeneous systems and the poor activity is not surprising. However, the number average molecular weight obtained from soluble initiators varied greatly, ranging from 26,000 for Sn(2-ethylhexanoate)2 to > 100,000 for Ph4Sn. Given a 100:1 monomer to initiator ratio and complete participation by all initiator added to the polymerization, the Mn for these polymerization should be near 17,000. Membrane osmometry results obtained in toluene indicate that the number average molecular weight of poly(ethylglycolide) and polystyrene determined from GPC data are comparable, and thus number-average molecular weights > 17,000 are consistent with incomplete initiation. The effect of adding alcohol as 75 minu: COINS polvm pohnr ethyll Thus. of lac inhal Conn adde are l impL initiator (second row of each entry) is clear. All of initiators surveyed gave high conversions and molecular weights close to the theoretical values for M... Sn(2-ethylhexanoate)2 is the most efficient catalyst known for the bulk polymerization of Iactide. Lactide polymerizations using this catalyst have high polymerization rates and require small catalyst loads. Also, Sn(2- ethylhexanoate); produces high molecular weight polylactides in high yield. Thus, Sn(2-ethylhexanoate); was our primary choice of catalyst. Because most proposed mechanisms for solution and bulk polymerization of Iactide using Sn(2-ethylhexanoate)2 invoke participation of water or alcohol as initiators, we added t-butylbenzyl alcohol (BBA) to polymerizations to gain better control of the molecular weight and the molecular weight distribution. Without the added alcohol, the polymerization is initiated by residual moisture or alcohols that are present as impurities in the polymerization, and because the amount of such impurities vary from run to run, the initiation efficiency should also vary. The polymerization of ethylglycolide shows the importance of the added alcohol (Figure 12). The results from six Sn(2-ethylhexanoate)2 initiated polymerizations are shown, three without an alcohol initiator, and three where one equivalent of t- butylbenzyl alcohol was added. Runs with added alcohol show a linear relationship between molecular weight and conversion up to >60% conversion. The data from the three runs without alcohol are more scattered, especially at high conversion, showing that added alcohol improved the efficiency of initiation and provides better control of molecular weight. 76 poll molr thec for I one beha has attrit and I Figure 13 shows the evolution of molecular weight with conversion for the polymerization of Iactide, ethylglycolide, isobutylglycolide and hexylglycolide. The molecular weight increases linearly with conversion and nearly reaches the theoretical values, 14,000 for polylactide, 17,000 for poly(ethylglycolide), 23,000 for poly(isobutylglycolide) and 28,000 for poly(hexylglycolide). These data show one characteristic expected of a living polymerization, a linear relationship between Mn and conversion. The drop in molecular weight at the end of the run has been observed previously in Iactide polymerizations,92u185 and is usually attributed to intramolecular transesterification reactions that form cyclic products and decrease the number average molecular weight. ~4- Enlr Table 5. Bulk Polymerization of Ethylglycolide Entry Catalyst %converiona Mnx10'3 Mwan 1 Sn(Oct)2 99 26.2 2.00 95 16.0 1.73 2 Ph4Sn 95 114 1.33 96 15.5 1.88 3 SnO 7 5.4 1.15 94 14.9 1.86 4 PhD 96 31.7 1.65 97 13.9 1.90 5 SnBr4 98 37.7 1.78 98 15.2 1.89 6 SnBl'z 61 13.3 1.64 97 13.9 1.82 [ethylglycolide]:[catalyst] = 100: polymerization mn at 180 °C for 2.5 hours a. determined by 1H NMR; top and bottom entries for each catalyst are data for polymerization with no alcohol added and with neopentyl alcohol ([ethylglycolide1/[aIcohol]=100) as the initiator, respectively. 78 En hOurs data (Usab Table 6. Bulk Polymerization of Isobutylglycolide Entry Catalyst %converiona Mnx10'3 Mw/Mn 1 Sn(Oct); 97 37.1 2.00 96 22.2 1.65 2 Ph4Sn 95 50.2 1.97 96 21 .3 1.88 3 SnO 45 10.9 1.23 94 20.8 1 .84 4 PhD 96 41 .9 1.65 97 22.9 1.90 5 SnBr4 98 57.1 1.88 98 20.5 1 .78 6 SnBrz 76 31 .2 1 .84 97 21.8 1.89 [isobutylglycolide]:[catalyst] = 100: polymerization run at 180 °C for 2.5 hours a. determined by 1H NMR; top and bottom entries for each catalyst are data for polymerization with no alcohol added and with neopentyl alcohol ([isobutylglycolide]l[alcohol]=100) as the initiator, respectively. 79 Table 7. Bulk Polymerization of Hexylglycolide Entry Catalyst %converiona Mnx10'3 Mw/Mn 1 Sn(Oct); 95 53.2 1 .92 97 26.8 1.93 2 Ph4Sn 93 43.4 1.83 96 25.5 1.78 3 SnO 31 13.5 1.19 97 27.9 1.81 4 PhD 94 67.7 1.95 98 25.9 1.82 5 SnBu 95 47.1 1.88 99 26.9 1.89 6 SnBrz 32 17.3 1.54 96 23.2 1.87 [hexyllglycolide]:[catalyst] = 100: polymerization run at 180 °C for 2.5 hours a. determined by 1H NMR; top and bottom entries for each catalyst are data for polymerization with with no alcohol added and with neopentyl alcohol ([hexylglycolide]l[alcohol]=100) as the initiator, respectively. 80 35 A 30 - 25 - 4 L «9° 20 A ._ x r: 2 15 - f f”’ A x 10 " A 1’ x6 1” 5 ' I.’ ,4 fl 0 l L 1 1 0 20 4O 60 80 100 conversion (%) Figure 12. Bulk polymerization of ethylglycolide with (O) and without (A) added t-butylbenzyl alcohol as co-initiator. Polymerization conditions: 130 °C, [Sn(Oct)2]/[t-butylbenzyl alcohol] = 1, [Monomer]l[lnitiator] = 100. 81 30 O 20 40 60 80 100 Conversion (%) Figure 13. Molecular weight versus conversion for the bulk polymerization of substituted glycolides. (A) Iactide, (I) ethylglycolide, (O) hexylglycolide, (O) isobutylglycolide. Polymerization condition: 130 °C, [Sn(Oct)z]/[t-butylbenzyl alcohol] = 1, [Monomer]/[lnitiator] = 100. 82 2.2 Kinetics of Bulk Polymerization Ring-opening polymerizations of Iactides and Iactones typically follow first order kinetics that can be expressed by Eq. (1) __d[M]= R— T: kp[M][l] .......................... (1) where [M] and [I] are the concentration of monomer and initiator, and k,, is the rate constant for propagation. For the case of a living polymerization, [I] is constant and integration of Eq. (1) yields: [M L — ln( [M 10 ):kp[1]0t ............................. (2) where [M], is the concentration of the monomer at time t, [Mjo is the initial monomer concentration (at t=0) and [ljo is the initial concentration of initiator. For living polymerizations, [I] is constant and plots of -ln([M],/[M]o) vs. t are linear. Figure 14 shows kinetic data for several substituted Iactides. After a fast initial polymerization, the rate slows as the conversion exceeds 80% (- In([M]./[M]o)=1.6). The data for low conversions are shown in Figure 15, and from each data set, kp[l] values were extracted from slopes of the linear portion of the the plots. However, the data in Figure 14 also show strong deviation from linearity at conversions above 80%, and that the conversion saturated near 97% conversion. 83 4 ‘ A A . I 3- ‘ - ’ I C 3 o 5 $5 22“ 2 ‘c’ ‘- ' o 11 'O o ‘: I 0 . o 20 40 Time (min) Figure 14. Kinetics of bulk polymerization of substituted glycolides. (A) Iactide, (I) ethylglycolide, (O) hexylglycolide, (O) isobutylglycolide. Polymerization conditions: 130 °C, [Sn(Oct)2]/[t-butylbenzyl alcohol] = 1, [Monomer]l[lnitiator] = 100 3 E. 2' ,2 E E 1. A O I I o 5 1o 15 Time (min) Figure 15. Kinetics of bulk polymerization of substituted glycolides at low conversion. (A) Iactide, (I) ethylglycolide, (O) hexylglycolide, (O) isobutylglycolide. Polymerization conditions: 130 °C, [Sn(Oct)2]/[t-butylbenzyl alcohol] = 1, [Monomer]l[lnitiator] = 100 85 From Figure 14 and Figure 15, we Ieamed that bulk polymerizations of substituted Iactides are first order in the concentration of monomer and first order in the concentration of initiator at low conversion. At high conversion, the polymerization rate begins to drop and finally polymerization stops without reaching 100% conversion. This phenomenon has been reported in the literature.92.93 Witzke et 31.37 explained it using an equilibrium polymerization model. However, we think there are two possibilities that can influence the polymerization kinetics. First, the polymerization is an equilibrium reaction, second, the initiator continuously degrades during the polymerization. The possibility of catalyst decay came from the solution polymerization of ethylglycolide, where we recovered a precipitate from the polymerizations. We believe it has the structure shown in Figure 16. This precipitation may represent catalyst lost from the solution polymerization (discussed later). We used these two models to fit the kinetic data that we obtained from bulk polymerization. If the polymerization is an equilibrium reaction, the kinetics of the polymerization can be expressed using following equations.87 [M], = [MM +([M]0 -[M]eq)e_kpm’ ............................... (1) 86 where [Mleq is the monomer concentration at equilibrium, X is the conversion (([Mo-[M],)/[M]o), and kp is rate constant for polymerization. We fit the data using equation (2). The results shown in Figure 17 show that the fit was good. If the initiator continuously degrades during the polymerization, the kinetics of the polymerization can be expressed as: _d[M] = -kdr dt kp[M][I]e ................................... (3) p: where kd is the rate constant for catalyst decay. Integration of equation (3) yields [M]: kP 4" -ln——— =—I (1- d .................................. 4 ([M]0) kd[ ]o e ) () We also used equation (4) to fit the kinetic dat the result is shown in Figure 18. Unfortunately, the data also fits equation (4) well. Thus, we still can not decide which model correctly describes the kinetics of the Iactide polymerization. To decide which model is appropriate, we studied the depolymerization reaction. If the Iactide polymerization is indeed an equilibrium reaction, depolymerization should give us the same [Mleq as polymerization. We scrupulously purified the polymers to remove all monomer in the polymer, and we conducted polymerization and depolymerization using the same conditions. The 87 results are shown in Figure 19. The polymerization and depolymerization indeed reached the same [M]9q, about 3%. The white precipitate recovered from solution polymerization was tested as a catalyst for the bulk polymerization of ethylglycolide. The polymerization also reached 97% conversion. These experiments proved that the equilibrium model is the right model. At low conversion, the polymerization is first order in monomer concentration with characteristics of a living polymerization. For a first order reaction, the kinetics can be expressed as: d M R =-—[-—]-=kp[M][1] ............................ (5) dt where [M] and [I] are the concentration of monomer and initiator, and kp is the rate constant for propagation. For the case of a living polymerization, [l] is constant and integration of Eq. (5) yields _ LML _ 111([M]0 ) _ k p[I]0t ........................ (6) where [M], is the concentration of the monomer at time t, [M10 is the initial monomer concentration (at t=0) and [I] is the concentration of initiator From Figure 15, we know that the plot of -In([M],/[M]o) vs. tis linear at low conversion. The slope is kp[I], which is the rate constant for Iactide polymerization. The correct fit to the kinetic model should produce the same kp[l]. 88 The results (Table 8) show that the kp[I] values obtained from the equilibrium model are consistent with those obtained from the fit to the first order reaction, and that the catalyst degradation model dramatically overestimates the k,,[l]. These k,,[l] data indirectly support that equilibrium model as the correct model to describe the kinetics of Iactide polymerization. 89 H—0\ 0 C—C’ \' .x O§C C’ / \\ "s- \ / \ R 0 I: O O’H C.) I H’ O I ,0 \c_c< \,Sn: >C C\’ R/ \O/ i O O—H 5 Sn Figure 16. The structure of the white precipitate formed during the solution polymerization of ethylglycolide using Sn(Oct)2/ROH as initiator Table 8. The kinetic data for bulk polymerization of substituted Iactides Kpm (sec‘1)><103 Monomer from from fit to from fit to catalyst conversion equilibrium degradation vs. time data model model Iactide 6.2 6.7 15.1 ethylglycolide 4.8 4.7 10.8 hexylglycolide 3.7 3.7 6.7 isobutylglycolide 3.0 3.0 5.3 90 100 80 § 3 60 .9 9 (D t): 40 8 A Poly(lactide) I Poly(ethylglycolide) 20 e Poly(hexylglycolide) O Poly(isobutylglycolide) 0 l l l 0 1 0 20 30 40 time (min) Figure 17 Kinetics of bulk polymerization of substituted glycolide. (A) Iactide, (I) ethylglycolide, (O) hexylglycolide, (O) isobutylglycolide. Polymerization condition: 130 °C, [Sn(Oct)2]l[t-butylbenzyl alcohol]=1, [Monomer]l[lnitiator]=100. The data were fitted using an equilibrium model. 91 .. / . e 3 2 5‘1. 2. 2 ’ 1? 1 _ A Poly(lactide) I Poly(ethylglycolide) ,, e Poly(hexylglycolide) I O Poly(isobmylglycolide) 0 . 0 20 40 Time (min) Figure 18. Kinetics of bulk polymerization of substituted glycolide. (A) Iactide, (I) ethylglycolide, (I) hexylglycolide, (O) isobutylglycolide. Polymerization condition: 130 °C, [Sn(Oct)2]/[t-butylbenzyl alcohol]=1, [Monomer]l[lnitiator]=100. The data were fitted using the initiator degradation model. 92 conversion 100 95 90 _ o .e- .p_e. - _ _ _ 3 ______ o . ‘ . g I 9 _ e polymerization - o depolymerization 0 10 20 time (minutes) Figure 19. Polymerization/depolymerization data for the bulk polymerization of ethylglycolide. Conditions: 200 °C [Sn(Oct)z]l[BBA]=1 [monomer]l[catalyst]=100 93 Ring opening polymerization of substituted Iactides is facile, though a bit slower than Iactide itself. The propagation step of glycolide polymerization is depicted in Scheme 31. The glycolide ring is arbitrarily drawn in a planar conformation with the R group in an equatorial position, although a more realistic representation would need to account for the boat-like conformation of the ring seen from x-ray studies and the mixture of glycolide diastereomers used in the polymerization. The polymerization rate of glycolide is faster than Iactide (R = CH3), because the steric bulk of the methyl group hinders nucleophilic attack at the ring carbonyls. Increasing the size of the ring substitutent should decrease the polymerization rate further. Returning to Figure 15, we see that the rates of polymerization follow the expected trend: lactide> ethylglycolide> hexylglycolide> isobutylglycolide. A survey of Iactide monomers by Hall186 showed that ring substitution plays a major role in defining the polymerizability of Iactide. For example, 3,3,6,6-tetramethyl-1,4-dioxane-2,5-dione, obtained by adding two methyl groups to the Iactide ring, does not undergo ring-opening polymerization. Presumably, nucleophilic attack by either the initiator or the growing polylactide chain is too hindered to lead to polymer. Scheme 31. The propagation step for Iactide polymerization using Sn(2-ethylhexanoate); as catalyst R' O \o I ,o Snz'V 0/5 “"0 R b 95 3. Solution Polymerization of Substituted Lactides Solution polymerizations exhibit different polymerization behavior than bulk polymerizations. Because of the added solvent, the concentration of the monomer is lower and the propagation rate, klele'], is reduced. Also, solution polymerizations are run at lower temperatures than bulk polymerization, usually lower than boiling point of the solvent. Because of the lower polymerization temperature, control of molecular weight and the molecular weight distribution is much better and there are fewer side reactions. The solvents used in solution polymerizations of Iactide include toluene, THF and 0112012. The most commonly used solvent is toluene. We choose Al(OiPr)3 and Sn(2-ethylhexanoate)2/alcohol as initiators since Al(OiPr)3 is known to be a good initiator for solution polymerization of Iactide, while the Sn(2-ethylhexanoate)2/alcohol system is commonly used for bulk polymerization. We wanted to evaluate the Sn system under solution polymerization conditions, since studying the more controlled solution polymerization can help us understand more about the polymerization mechanism when Sn(2-ethylhexanoate); is used as catalyst for bulk polymerization. 3.1 Al(OiPr); as Initiator Al(OiPr)3 is well-known to initiate living Iactide polymerization, where the molecular weight grows linearly with conversion and the polymerization follows first order kinetics. Plots of —In([M],/[M]o) versus t should be linear. The substituted Iactides have structures similar to Iactide, so we should expect the polymerization of substituted Iactides also to be living and follow first-order kinetics. The molecular weight versus conversion data are shown in Figure 20 and a plot of —In([M],/[M]o) versus t is shown in Figure 21. As expected, the polymerization of substituted Iactides using Al(OiPr)3 is a living polymerization and follows first order kinetics. From Figure 20 we can see that Al(OiPr)3 initiates three chains. The molecular weight of polymers can be predicted from the equafion: = [M] xM " 3x[1] ° where M, is the number-average molecular weight of polymer, [M] is concentration of the monomer, [I] is the concentration of the initiator, M0 is molecular weight of monomer and n is the number of initiating species for the Al(OiPr)3 complex. Since the observed molecular weight is one third that predicted by the WW] ratio, each isopropoxide must initiate one polymer chain. Based on the bulk polymerization results, we expected that the rates of the polymerization should follow the order: Iactide > ethylglycolide > hexylglycolide > isobutylglycolide, because the steric bulk of the ring substitutents increase in that order. However, the solution polymerization rates do not follow the expected order. Lactide, ethylglycolide and hexylglycolide do follow the expected trend, but isobutylglycolide has the bulkiest side group and the fastest polymerization rate. 97 To understand this phenomenon, we calculated the activation energy for solution polymerization of substituted Iactides. The activation energy can be calculated from the Arrhenius equation: -152 k = Ae RT Ink = lnA— Ea RT where k is the rate constant, A is a constant of integration called the frequency factor, E, is the activation energy, R is the gas constant and T is the temperature in Kelvin. When —In R is plotted against 1/T, the activation energy and frequency factor can be extracted from slope and intercept, respectively. Polymerizations were run at three temperatures: 70 °C, 90 °C and 100 °C. The results are shown in Figure 21, Figure 22 and Figure 23 respectively. The rate constants are shown in Table 9. The activation energies were obtained from the plots shown in Figure 24 and the results are listed in Table 10. The results are totally unexpected. We believed that bulkier side groups should hinder attack at the carbonyl group on the Iactide ring, and lead to a higher the activation energy. Our expectation that the bulkier side groups should result in the higher activation energies is based on the assumption that the nucleophilic attack on the carbonyl group is the rate-determining step. This might not be the case. The polymerization of Iactide includes several mechanistic steps: coordination, nucleophilic attack, and ring opening. Nucleophilic attack might be rate limiting 98 for linear alkyl substituents, but may not be the rate-determinating step for more bulky or highly substituted Iactides. 99 10 \I’ .3 /A O I T I I 0 20 40 60 80 100 conversion (%) Figure 20. Molecular weight versus conversion for the solution polymerization of substituted Iactides. (A) Iactide, (I) ethylglycolide, (O) hexylglycolide, (O) isobutylglycolide. Polymerization conditions: toluene as solvent, Al(OiPr)3 as initiator, 90 °C, [Monomer]l[lnitiator] = 100 100 1.2 - E E 0.8 f E E E A 0.4 ’ ‘ I I o 41 I l l 0 5 10 15 20 25 time (h) Figure 21 Kinetics for the solution polymerization of substituted Iactides at 70 °C. (A) Iactide, (I) ethylglycolide, (O) hexylglycolide, (O) isobutylglycolide. Polymerization conditions: toluene as solvent, Al(OiPr)3 as initiator, [Monomer]l[lnitiator] = 100 101 -|n([M]t/[M]0) Time (h) Figure 22 Kinetics for the solution polymerization of substituted Iactides at 90 °C, (A) Iactide, (I) ethylglycolide, (I) hexylglycolide, (O) isobutylglycolide. Polymerization conditions: toluene as solvent, Al(OiPr)3 as initiator, [Monomer]l[lnitiator] = 100 102 1.6 ’3 '— 1. a 2 2. E .E 0.8 0.4 Time (h) Figure 23 Kinetics for the solution polymerization of substituted Iactide 100 °C. (A) Iactide, (I) ethylglycolide, (O) hexylglycolide, (O) isobutylglycolide. Polymerization condition: toluene as solvent, Al(OiPr)3 as initiator, [Monomer]l[lnitiator] = 100 103 Table 9. Polymerization rate constants of substituted Iactides kpx1 03 (LImoI-s) Monomers 70 °C 90 °C 100 °C Iactide 10.5 38.5 73.8 ethylglycolide 8.03 25.0 54.2 hexylglycolide 6.15 18.2 32.6 isobutylglycolide 22.2 40.7 56.6 Table 10. Polymerization activation energies for substituted Iactides Monomers E. (kJImol) lnA Iactide 68.8 21 .6 ethylglycolide 66.2 20.4 hexylglycolide 58.6 17.4 isobutylglycolide 32.9 9.7 -1.5 - ’3 5 E -2.5 - _3.5 L l 1 0.0026 0.0027 0.0028 0.0029 0.003 1IT Figure 24. Activation energies for polymerization of substituted Iactides (A) Iactide, (I) ethylglycolide, (O) hexylglycolide, (O) isobutylglycolide 105 3.2 Sn(Oct)2IROH as initiators The Sn(Oct)2/ROH system is a very good catalyst/initiator system for bulk polymerization, but Its use in solution polymerizations is rarely mentioned in the literature. Also, solution polymerizations have a lower polymerization rate than bulk polymerizations, so it is relatively easy to study the kinetics and mechanism of the Sn(Oct)2/ROH system in solution. That ROH is the initiator and Sn(Oct); is the catalyst often has been claimed in the literature, but direct evidence is lacking. We compared solution polymerization of ethylglycolide initiated by Sn(Oct)2 and by Sn(Oct)leOH, where ROH is neopentyl alcohol. The results are shown in Figure 25. The diamonds of Figure 25 are the data for polymerization initiated by only Sn(Oct); where moisture and impurities have been scrupulously excluded. The rate of polymerization (slope of curve) is very low, and the polymerization barely reached 10% conversion. If moisture and impurities are not carefully excluded, polymerizations often showed appreciable conversion. The squres at the top of Figure 25 show results for a polymerization where one equivalent of neopentyl alcohol was added as initiator. Both higher polymerization rates and conversions are routinely obtained when ROH is used. Hydroxyl-containing compounds such as ROH and H20 are truly the initiator for the polymerization. The 10% conversion achieved in the bottom trace is likely caused by impurities. No matter how hard you try to exclude moisture and impurities, small-scale polymerizations always are affected by trace impurities. In the bottom trace, the impurities were carefully excluded, which caused the low polymerization rate and low conversion. 106 To further understand the role of ROH in polymerization, we ran solution polymerizations at different Sn(Oct)2/ROH ratios, while keeping the concentration of ethylglycolide and Sn(Oct)2 constant. The results are shown in Figure 26 and Figure 27. There are two important results. The polymerization follows first order kinetics up to about 70% conversion, and the higher the ROH/Sn(Oct)2 ratio, the faster the polymerization. However, as shown in Figure 27, the polymerization rate did not increase by five when the ROH/Sn(Oct)2 ratio was increased by five, which means that the kinetics of the polymerization are not first order in ROH. We ran two solution polymerizations with ROH/Sn(Oct)2=1, and [M]/[Sn(0ct)2]=100 and 200, respectively. The results are shown in Figure 28 and Figure 29. Figure 29 is the linear part of Figure 28. The results show that when the amount of both Sn(Oct); and ROH double, the rate of polymerization also doubles, which proves that Sn(Oct)2 and ROH together decide the kinetics of the polymerization. An explanation for the results is that the first step of the polymerization is an equilibrium between Sn(Oct)2 and ROH as shown in Scheme 32. The true initiatior is intermediate (A). The concentration of (A) can be calculated using equation below: K ___ [0ctSn0R][H0ct] [Sn(Oct)2][R0H] where K is the equilibrium constant. 107 When [Sn(Oct2]/[ROH]=1, the concentration of initiator (OctSnOR) can be calculated using following equation: [OctSnOR] = qusnmct), ][ROH] so the concentration of initiator (OctSnOR) will double if [Sn(Oct)2] and [ROH] both double. If [ROH] increases, [OctSnOR] will increases, but at a level defined by equation above. Scheme 32. The first step for Iactide polymerization using Sn(Oct)2/ROH o O—Sn—O + ROH o 0 HO 4. O—Sn— OR 0 A 108 0001:1013 DO U 22" D E 1: 2 1:: g a 5. l1. 0 D O§'M”OOOO 1 0 100 200 300 400 Figure 25 Kinetics for the polymerization of ethylglycolide initiated by Sn(Oct)2/ROH and Sn(Oct)2. (Cl) Sn(Oct)2/neopentyl alcohol (9) Sn(Oot)2. Polymerization temperature is 90 °C. [Ml/[Sn(Oct)2]=100, [ROH]/[Sn(0ct)2]=1 109 2.5 A QIAI O.‘A 21- .. ‘ O . A A 5°45- .'A l—l .‘ 2. - $1.. E I ' .5 0.5- 0 f l l i 50 100 150 200 time (h) 0 Figure 26 Solution polymerization of ethylglycolide at different Sn(Oct2)/alcohol ratios. Polymerization temperature is 90 °C. (0) [Sn(OctzlllROH]=1/10, (I)[Sn(0ct2]/[ROH]=1/5, (A) [Sn(Octfl/[ROH]=1 . ROH is neopentyl alcohol. [M]=0.2 mollL, [Ml/[Sn(Oct)2]=100. 110 -|n([Mlt/[M]0) 0 4L l 0 20 40 60 time (h) Figure 27 Solution polymerization of ethylglycolide at different Sn(Oct2)/alcohol ratios. Polymerization temperature is 90 °C. (0) [Sn(Oct21/[ROH]=1/10, (I)[Sn(Oct2]/[ROH]=1/5, (A) [Sn(Oct2]l[ROH]=1. ROH is neopentyl alcohol. [M]=0.2 mollL, [Ml/[Sn(Oct)2]=100. 111 3 09 e .9 D a 1!? o. D 22 ' . DD .5... e P 2. 0 DD .2.. . C D _'1 - 1:] CI QC] El OP 1 L L 0 100 200 300 400 time(h) Figure 28 Solution polymerization of ethylglycolide at different monomer/initiator ratio. ([3) [M]![I]=200 (Q) [M]/[l]=100. Polymerization temperature is 90 °C [M]=0.2 mollL, [ROH]/[Sn(0ct)2]=1 112 1.5 - E E. E .— 1 - E —? o 0.5 - O l A l o 20 4o 60 80 time (h) Figure 29 Solution polymerization of ethylglycolide at different monomer/initiator ratio. (0) [M]/[l]=200 (Q) [M]/[|]=100. Polymerization temperature is 90 °C [M]=0.2 moI/L, [ROH]/[Sn(0ct)2]=1 113 We also ran solution polymerizations of ethylglycolide at three temperatures 70 °C, 90 °C and 110 °C. The results are shown in Figure 30 and Figure 31. Figure 31 shows the linear portion of Figure 30. These figures provide polymerization rates for ethylglycolide initiated by Sn(Oct)2/BBA at three temperatures (Table 11). From these rates, we calculated the activation energy for ethylglycolide polymerization (Figure 32 and Table 12). Table 11 Polymerization rates for ethylglycolide initiated by Sn(Oct)2IBBA k, x 103 (mollL-s) Monomer 70 °c 90 °c 110 °c ethylglycolide 0.31 2.43 13.1 [Sn(Oct)2]/[BBA]=1, BBA is t-butyl benzyl alcohol. [M]/[Sn(0ct)2]=100 Table 12 The activation energy for ethylglycolide initiated by Sn(Oct)2/BBA monomer E. (KJImoI) lnA ethylglycolide 75.68 21 .3 114 AA 2.5 - A I I I AA I. ’3 2 ' A I. H I E '5 .A I . e ... U I C I . O -. .. u 1 I ... I O 0.5 0' O .0 o l l 0 200 400 600 time (h) Figure 30 Solution polymerization of ethylglycolide initiated by Sn(Oct)2/BBA. [Sn(Oct)2]/[BBA]=1, BBA is t-butyl benzyl alcohol. (A) 110 °C (I) 90 °C (0) 70 °C [Ml/[Sn(Oct)2]=100 115 1.2 0.8 - -|n([Mlt/[M]0) O 0.4 o 0.2 O 0 1 1 0 50 100 150 time (h) Figure 31 Solution polymerization of ethylglycolide initiated by Sn(Oct)2/BBA. [Sn(Oct)2]/[BBA]=1, BBA is t-butyl benzyl alcohol. (A) 110 °c (I) 90 °C (9) 70 °c [Ml/[Sn(Oct)2]=100 116 ln(kp) _5.5 l l I l 2.5 2.6 2.7 2.8 2.9 3 1rrx103 Figure 32 The activation energy of ethylglycolide initiated by Sn(Oct)2lBBA 117 As mentioned above, we isolated a white precipitate during the solution polymerization of ethylglycolide. Vert et al.89 reported a similar precipitation during polymerization of Iactide using Sn(Oct)2.This white precipitate does not dissolve in organic solvents, so we could not directly identify it using NMR. Indirect methods were used to identify this white precipitate. Dissolution of the precipitation in dilute HCI solution, followed by extracting with ether, gave one organic compound, 2-hydroxybutyric acid. Also, when 2-hydroxylbutyric acid was added to Sn(Oct)2, a white precipitation formed instantly. These two precipitates have almost identical IR spectra (Figure 33). How does this white precipitate influence the kinetics of the solution polymerizations? Is it umimportant, as was shown for bulk polymerizations, or does it remove active catalyst from the polymerization and have a significant influence on the polymerization kinetics? To answer these questions, we set up the following experiment. A solution polymerization of ethylglycolide was run until polymerization stopped. A second volume of ethyllgycolide solution equal to that initially used was added to polymerization system. If formation of the precipitate does not influence the kinetics of the solution polymerization, the polymerization should return to its equilibrium state with polymerization rate halved. If all of the catalyst has degraded to the white precipitate, and can not polymerize ethylglycolide, then the polymerization would not return to its equilibrium state and will remain near 50% conversion. The result is shown in Figure 34. The initial polymerization reached a plateau at 93% conversion at about 54 hours. Additional monomer solution was added, and the conversion dropped to about 118 50%. The polymerization then slowly retured to the plateau at >90% conversion. Thus, we can say that the polymerization is indeed an equilibrium reaction. However, we can not ignore the effect of catalyst degradation. As shown in Figure 35, the polymerization rate at the beginning of the polymerization and after the monomer solution was added are very different. The initial polymerization rate is almost ten times faster than the polymerization rate after the monomer was added. As mentioned before, if there is no effect from catalyst degradation, the polymerizationrate after adding monomer should be half the initial polymerization rate. Since the observed rate is 5 times lower than expected, we conclude that ~80% of the catalyst had precipitated by the time the second portion of monomer was added, and does not participate in polymerization. 119 white precipitate from 2-hydroxy acid : .2 a % white precipitate from .3 polymerization «l 4500 3500 2500 1 500 500 wave number (cm'1) Figure 33. IR spectra for white precipitates formed during polymerization of ethylglycolide and by mixing 2-hydroxylbutyric acid and Sn(Oct)2. 120 3 - <— monomeradded 2.5 - 1:1 I E 2 ' :1 a - a a :1 H15 P E E . D I 1 b: [553 I 0.5 ' 0 1 1 1 I 0 50 100 150 200 250 time (h) Figure 34. Kinetics of solution polymerization of ethylglycolide showing the result after adding extra monomer after the polymerization reached equilibrium. Polymerization temperature is 90 °C. [M]/[Sn(0ct)z]=100, [ROH]I[Sn(Oct)2] = 1 121 1.6 1.2 .. E. g F! 0.8 '- E. E 0.4 b O l l 0 50 100 150 Time (h) Figure 35. Initial polymerization rate for ethylglycolide and the decreased rate observed after adding addtional monomer. Polymerization temperature is 90 °C. [Ml/[Sn(Oct)2]=100, [ROH]I[Sn(Oct)2] = 1 122 Earlier, we saw that for solution polymerizations using Al(OiPr)3, the polymerization rates for substituted Iactides did not follow the order predicted by the steric hindrance of the substituent on the glycolide ring (Figure 21, 22 and 23). To determine whether the effect is initiator-related, we ran solution polymerizations of substituted Iactides using Sn(Oct)2/BBA instead Al(OiPr)3. The results are shown in Figure 36. The results are slightly different, but again the polymerization rates did not follow the order expected based on steric hinderance. For polymerizations using Al(OiPr)3, at 90 °C, the order of the polymerization rate is isobutylglycolide > Iactide > ethylglycolide >hexylglycolide, but for polymerization using Sn(Oct)2/BBA, the order is Iactide > isobutylglycolide > ethylglycolide > hexylglycolide. These results reflect activation energy differences for the two initiator systems. 3.3 Comparison of Al(OiPr)3 and Sn(Oct)2lROH as Initiators in Solution Polymerization Sn(Oct)2/ROH is the best initiator for bulk polymerization of Iactides, but it is less effective in solution polymerizations. Figure 37 shows a comparison of the Al(OiPr)3 and Sn(Oct)2/ROH initiators for solution polymerization of ethylglycolide at 70 °C. The polymerization initiated by Al(OiPr)3 is almost 10 times faster than the polymerization initiated by Sn(Oct)2/ROH. Considering that an Al(OiPr)3 molecule initiates three chains and Sn(Oct)2IBBA only initiates one chain, Iactide chains in the Al(OiPr)3 system grow almost three times faster at 70 °C than those initiated by Sn(Oct)2/BBA. Figure 38 shows that the same rate difference holds 123 at 90 °C. The activation energies calculated from the solution polymerizations are shown in Figure 39 and Table 13. The activation energy for Al(OiPr)3 catalyzed polymerization is ~10% less than for the Sn(Oct)2/BBA system. Theoretically, lnA from both initiators should be the same because lnA represent the steric factors of the monomer. The values of lnA calculated from the experiments are fairly close. Table 13 The activation energy of ethylglycolide initiated by Sn(Oct)2/BBA and Al(OiPr)3, initiator E. (KJlmol) lnA Al(OiPr)3 66.2 20.4 Sn(Oct)2/BBA 75.7 21 .3 124 1.4 -|n([M]lI[M]0) 80 time (h) Figure 36 Solution polymerization of substituted Iactides initiated by Sn(Oct)2/BBA. Polymerization temperature is 90 °C, (A) Iactide, (I) ethylglycolide, (I) hexylglycolide, (O) isobutylglycolide [Sn(Oct)2]l[BBA]=1, BBA is t-butyl benzyl alcohol. [Ml/[Sn(Oct)2]=100. Toluene is the solvent. 125 1.4 -In([M]tI[M]0) 0 50 100 time (h) Figure 37. Solution polymerization of ethylglycolide initiated by Al(OiPr)3 (O) and Sn(Oct)2/BBA (I). [M]o=0.2 mol/L. [M]![l]=100, [Sn(Oct)2]/[BBA]=1 polymerization temperature is 70 °C and toluene is the solvent. 126 1.6 b A 1.2 " ,2. E 2'. g 0.8 - 5 u I 0.4 0 [ 1 1 1 0 20 40 60 80 time (h) Figure 38. Solution polymerization of ethylglycolide initiated by Al(OiPr)3 (O) and Sn(Oct)2/BBA (I). [M]o=0.2 mol/L. [M]![l]=100, [Sn(Oct)2]/[BBA]=1 polymerization temperature is 90 °C and toluene is the solvent. 127 ln(kp) _6 J L l l 2.5 2.6 2.7 2.8 2.9 3 1rrx103 Figure 39. The activation energy for polymerization of ethylglycolide initiated by Sn(Oct)2/BBA (I) and Al(OiPr)3 (9), 128 3.4. Polymerization Mechanism 3.4.1 Al(OiPr); as initiator We think there are two possible scenarios that cause abnormal kinetic behavior. First, the polymerization is a three-step process as shown in Scheme 33. It includes coordination (A), nucleophilic attack (B) and ring-opening (C). When we expect polymerization rate to drop with increasing size of substituted group, we assume that the nucleophilic attack step (B) is the rate-determining step. However, the nucleophilic attack step (B) may not be the rate-determining step and identification of the rate-determining step has not been addressed in literature. It is possible that step B and step C are rate-determining. In step C, the ring opens up to form a linear chain. In a ring system, the substituents on the ring and the initiator are more crowded than that on a linear chain. The larger the group, the more crowded the ring and the easier it is to open the ring. If step C is rate-determining, larger substituents on the ring should lead to faster polymerization. From the activation energy data from Table 10, we see that the activation energy decreased with increases in the size of the substituents. We believe that step C is the rate-determining step for isobutylglycolide polymerization. The frequency factor lnA, which represent the steric effect of the reaction, also is in the right order. The larger the size of the substituted group, the smaller the value of lnA. 129 Scheme 33. The mechanism for the Iactide polymerization initiated by Al(OiPr)3 R1 R1 Al(OR2)3 #0 ———> 0 R A 3 R20 R20\A 0 O ‘—C—_ R1 R20 0 R1 R20 A second scenario is based on the coordination number of Al during Iactide polymerization. As mentioned in the Introduction, Al(OiPr)3 exists as a trimer and a tetramer (shown in Figure 3) in solution. During Iactide POlymerization, the aggregated complex is broken up by monomer to form Complex (A) shown in Figure 40. The coordination number on Al is 6. During the POlymerization of Iactide, ethylglycolide and hexylglycolide, complex (A) is f0"fled. In the case of isobutylglycolide, it is may not possible to accomodate three monomers on Al because of the large size of substituted group. It is may be Possible to form the 4-coordinate complex (B) shown in Figure 40, with only one 130 monomer molecule coordinated to the Al atom. The 4-coordinate Al is more reactive than 6-coordinate AI, and perhaps that is why isobutylglycolide polymerized faster than Iactide, ethylglycolide and hexylglycolide. However, in polymerizations of substituted Iactides initiated by Sn(Oct)2/ROH, the polymerization rates also do not follow the order of the size of subsitituents, and Sn(Oct)2/ROH does not aggregate. Thus, it is unlikely that this hypothesis is right. 131 Figure 40. The structure of the transition state for polymerization of substituted Iactides 132 3.4.2 Sn(Oct)2IROH as catalyst/initiator Based on the kinetic data, we propose the mechanism shown in Scheme 34. The first step is an equilibrium between Sn(Oct)2 and ROH to form RO-Sn- Oct and HOct. Because of the non-polar environment, it is unlikely that either Sn(Oct); or ROH will dissociate appreciably. We were unable to observe the RO- Sn-Oct intermediate using low temperature NMR. One reason for our failure could be that the equilibrium favors Sn(Oct)2 and thus the equilibrium concentration of RO-Sn-Oct is below the detection limits of the NMR. A different method must be developed to identify the intermediate. Recently, Penczek et al. reported detection of RO-Sn-Oct during polymerizaiton using MALDI Mass spectrometry. The second step is initiation. The actual initiator is RO-Sn-Oct, a tin alkoxide. Initiation by tin alkoxides has been well documented in the literature. Though the equilibrium amount of RO-Sn-Oct is small, a fast equilibrium leads to fast initiation. We know from kinetic data that polymerizations using Sn(Oct)2/ROH are living polymerizations below 80% conversion. One characteristic of living polymerization is a fast initiation reaction. The first two steps of the polymerization are much faster than the third step of the polymerization: propagation. The propagation of Iactides initiated by metal alkoxides have been well studied in systems such as Al(OiPr)3 and Sn(OMe)4. It is a coordination-insertion process. The fourth step of the polymerization is the regeneration of catalyst and initiator. HOct reacts with 133 active chains to regenerate Sn(Oct); and a polymer chain with a hydroxyl chain end. The hydroxyl chain end can re-initiate polymerization. The last step of the polymerization is catalyst degradation. As mentioned before, Sn(Oct)2 degrades to an insoluble white powder, the salt of Sn and the 2-hydroxy acid. In bulk polymerization, this white powder can also polymerize Iactide, so it does not influence the kinetics of the polymerization. However, in solution polymerization, this white powder is insoluble, which means the catalyst is no longer in the polymerization system and causes the polymerization rate to drop dramatically. This reaction mechanism is similar to atom-transfer polymerization (ATRP) (Sheme 35). The initiator reacts with catalyst to generate an active species, followed by initiation and propagation, and then initiator and catalyst are regenerated. ATRP is a living polymerization. Lactide polymerization initiated by Sn(Oct)2/ROH is also a living polymerization at low conversion, and both have termination steps. 134 Scheme 34. The mechanism of Iactide polymerization using Sn(Oct)2/ROH as catalystflnitiator ROH SnOctz ROMnOH \J HOct M, kd V HOct ROSnOct M, k, ROMSnOct (SnM)x polymer (insoluble) 135 Scheme 35. The mechanism for atom-transfer radical polymerization (ATRP) R'——Cl + ’i‘ R'~[ACH2—C‘}CH2—CXY =R-c1 Y n ('3' CUC'Z H2C:CXY X I Y n kdxi X I_ I_ Y n 136 3.5. The Influence of Stereochemistry on the Kinetics of Solution Polymerization The stereochemistry of the monomer has a big influence on the kinetics of solution polymerizations of substituted Iactides. As shown in Figure 41 and Figure 42, racemic Iactide polymerized twice as fast as L-lactide, and racemic isobutylglycolide polymerized three times as fast as D-isobutylglycolide. Lactide monomers have two stereocenters, so racemic Iactides include an equimolar amounts of the (R,R) and (S,S) stereoisomers. As shown in Scheme 36, when L-lactide or D-lactide are polymerized, there is only one propagation reaction, an R active center reacting with (R,R) monomer with a rate constant ka/RR, or an S active center reacting with (S,S) monomer with a rate constant MKS/3,3). When racemic Iactides polymerize, four propagation reactions must be considered, the active center with R stereochemistry reacting with (R,R) monomer with a rate constant kmmg), the active center with R stereochemistry reacting with (S,S) monomer with a rate constant mpg/3,3), the active center with S stereochemistry reacting with (R,R) monomer with a rate constant kmsmm and the active center with S stereochemistry reacting with (S,S) monomer with a rate constant kp(s,s,s,. Usually, kmmg) is equal to MKS/5,3) and kpayss, is equal to kp kp(s,s.s) and kpmm R), the racemic monomer will polymerize faster than the pure D or L monomers, which is called syndiotactic preference. 137 Scheme 36. Stereochemistry of Iactide polymerization o O 1 R2 O R o kMRIRJRI \ -———> ‘0 A110R3>2 + o R‘ R1 0 R (R. R) o O 1. R2 o R “(mo ”M's, ‘0 ‘AI<0R3>2 + o —" R‘ 'R1 0 R (S. S) O S (R. R) O O 1 RM. R2 /u\/ iii/[LO was,” ‘0 ‘AI(0R3)2 + (kn/l. —__’ - "'R1 R1 0 s (5.3) 138 0.8 - ’a "5" 0.6 L E E. :5 0.4 - O 0.2 - 0 1 1 0 2 4 6 time (h) Figure 41. Polymerization of rec-Iactide (I) and L-lactide (O). Polymerization temperature is 90 °C. [M]0=0-2 mollL, initiator is Al(OiPr)3, [M]/[l]=100. Toluene is the solvent. 139 I I 1.5 ~ I E E E H 1 l- E 2.5 0.5 - 0 1 0 5 10 time (h) Figure 42. Polymerization of rec-isobutylglycolide (I) and D-isobutylglycolide (O). Polymerization temperature is 90 °C. [M]o=0.2 mollL, initiator is Al(OiPr)3, [M]l[|]=100. Toluene is the solvent. 140 As shown in Figure 41 and Figure 42, polymerization rates of L-lactide and rac-lactide are different. Similar results were obtained for the polymerization rate of L-isobutylglycolide and rec-isobutylglycolide. Because of the stereochemistry of the monomer and the chain end, there are four different reactions during the polymerization rec-polylactides as shown in Scheme 36. The kinetics of rac-lactide polymerization can be expressed as shown in Scheme 37. Equations A and C in Scheme 37 produce isotactic chains, so k5,“) = [QR/RR) = k,~, while B and D produce syndiotactic chains, so k(s,R,R) = kHz/3,3) =ks. It has been found that k. < ks in Iactide polymerization, which means syndiotactic chain placement is favored in Iactide polymerization. Munson et al.187 found k. lks =0.6 for the bulk polymerization of Iactide initiated by Sn(Oct); at 180 °C and Kasperczyk188 found k; lks = 0.32 for solution polymerization of Iactide initiated by tert-butoxide at 20 °C in THF. The method used to calculate k; Iks was monitoring ratios of stereosquences by NMR. Because the NMR assignment of the stereosquence is still a controversial topic, and the resolution of NMR is not sufficent to clearly separate all stereosquences and account for side reactions such as transesterification and racemization which will alter the stereosquence, the ki lks ratio reported by Munson and Kasperczyk may not be very accurate. By studying the kinetic behavior of rac-lactide and L-lactide polymerization, we obtained more accurate measures of kilks. 141 Scheme 37. Kinetic scheme for Iactide polymerization “918135) —s' + $3 —_> —sss‘ A kMSIRR) —S' + RR ——> —SRR B , kpima) , —— R + RR ——-> — RRR c kp(RIS.S) — R. 1' SS —-> —SRR' D 142 From Scheme 37, we can write the kinetic equation as: d M 6 0 I I “157] = k....,.,is iiSi+k..,..,.,[S 11R1+k....,..[R ][R]+k(R/s_.-,[R 1181-...(7) Since k(S/S.S) = k(R/R,R) = ki k(R/S,S) = k(S/R,R) = ks We re-write equation (7) as: —i"[—M—i = k,[Si][S]+ks[Si][R]+k,-[Ri][R]+ kS[R'][S] ................................... (8) In rac-lactide polymerization: _ JM [SI—[RI- 2 Equation (8) can be re-written as —i[(%4—] = O.5k,-[Si][M]+ 0.5kS[Si][M]+ O.5k,-[Ri][M]+ 0.51., [MM] ........... (9) — gig—i = 0.5k,([5‘]+ [Ri])[M] + 0.5k,([5* ] + [Ri])[M] .................................. (10) Also: [S']+[R']=[1] 143 Equation (10) can be re-written as: _ aw] .11 = (051-, + 0.51.5 )[I][M] ........................................ (11) lntergration of (11) yields: [M]. -1 —= 0.5k.+0.5k It .......................................... 12 “Wk; ( . .)[] ( ) Plots of -ln([M],/[M]o) vs tgive (0.5K,- + 0.5 ks)[I] as the slope. In L-Iactide polymerization, there is only one monomer, so the kinetics of the polymerization can expressed using equation (13): M — 111% = k,[1]t ............................................................. (13) k,- [I] can be obtaied from plot of —ln([M],/[M]o) vs t. If a L-lactide and a rac- lactide polymerization is run under the same conditions, [I] is the same for both polymerizations. From these two slopes, k. lks can be solved. For polymerization of Iactide at 90 °C, k. lks = 0.29, when initiated by Al(OiPr)3 in toluene, which is close to the value (0.32) reported by Kasperczyk. For polymerization of isobutylglycolide, k; lk8 =0.24 under the same conditions. From k. Iks, the difference in activation energy can be calculated using the Arrhenius equation. 144 polyn iacti the con an Syl ple Fi ic —(AE,.—A_ES) _i_:e RT k S (AE,-AES)=—RT1n% S For polymerization of Iactide, the energy difference is 3.7 KJ/mol. For polymerization of isobutylglycolide, the energy difference is 4.2 kJ/mol. We also used semi—empirical molecular mechanics calculation to simulate Iactide polymerization. The calculations were done using SPARTAN. We set up the two complexes shown in Figure 43. Complex A has a chain end with R configuration and a coordinated S,S Iactide ring. Complex B has a chain end with an S configuration and a coordinated S,S Iactide ring. Complex A leads to syndiotactic chain placement and complex B will produce isotactic chain placement. The optimized structures of complex A and complex B are shown in Figure 44 and Figure 45 (Images are presented in color). In Figure 44, the O1 and AI bond is almost perpendicular to the Iactide ring, so it is in the right position for nucleophilic attack. We drove 01 close to C10 on the Iactide ring until the new O1-C10 bond was formed and monitored the energy of the complex along the pathway. The transition state energy of nucleophilic attack can be obtained from this energy trajectory. From the energies of complex A and the transition state, the activation energy of nucleophilic attack can be calculated, which is the activation energy for syndiotactic chain placement. Using the same method, the activation energy for producing isotactic chain placement was also calculated. 145 betwee dose. Ta user 0011 the diff. is Ste between calculated value and experimental value is shown in Table 14. It is fairly close. Table 14. Experimental value and calculated value for activation energy difference. AE, - AE. (KJ/mol) Iactide isobutyglycolide Experimental value 3.7 4.2 Calculated value a 4.5 5.1 a. Calculated using SPARTAN. The SPARTAN calculation has strong limitations. First, the calculation is used to simulate gas phase, which is very different from actual reaction conditions. Second, since a semi-empirical method was used in the calculation, the accuracy of calculation may not be good enough to judge the small energy difference shown in Table 14. The preference for syndiotactic chain placement for Iactide polymerization is cause by small differences in activation energy. which is probably caused by steric hindrance at polymer growing site. 146 O H H 0 CH5. \ O ""‘CH3 0 -. W‘- O ,.I' CH3 CH3 l O H O A R chain end and S S Iactide ring \ o O H: H ' 0 CH3 \ O “HICHG O’AIIIIHIO O _ O o I H /OH H3 H ° ‘1’ S chain end and S S Iactide ring Figure 43. The structure of complexes used to simulate polymer growth 147 as 07 0 H27 2 H25 , ‘21. 111- , llz . 4 f 01 c3 H1 02 1 - H3 ()9 "23 \ 011 I 10 7 H12 llfl "‘..._~ 0‘ '9. ' (:13 013 H22 32 HS a0 v H30 c5 r c1 " \ 0” .,. ‘ in? H28 cs H5 H15 Figure 44. The calculated structure for the complex with R configuration chain end and 8,8 Iactide ring. 148 H15 11 c5 "7 I‘- 05 I .3 H22 3 1121 03 )10 i 0,2 cm 013 011 . ’ ' H [:0 1117 ,. c17 . ' ' H18 C15 ' H27 Figure 45. The calculated structure for the complex with S configuration chain end and 8,8 Iactide ring. 149 4. Pol give 1 0111111 0013 5qu mi pol) wei $111 of ii] 4. Polymer properties To determine the properties of the polymers, polymerizations were run to give molecular weights near 50,000 g/mol (Table 15). The crude polymers were purified by precipitation into methanol and dried to constant weight. All polymers obtained from the polymerization of substituted glycolides are colorless, and are soluble in solvents ranging from toluene to CHCI3. Flexible films can be either melt pressed at 150 °C or cast from solvent. Because of the low Tgs for the polymers, the films tend to be somewhat tacky regardless of the molecular weight, and characterization of the polymers by polarizing optical microscopy showed that none of the polymers is crystalline. Given that the monomers used in the polymerizations is a mixture of RR, 8,8, and R,S diastereomers, the lack of crystallinity is not surprising. DSC scans were used to measure the glass transition temperatures for the polymers, and the results appear in Figure 46 and Table 15. Overall, thermal analysis data show that Tgs of the substituted glycolides range from —37 °C for poly(hexylglycolide) to 66 °C for polylactide. For polymers substituted with linear alkyl groups, The Tgs decreased as length of the alkyl group increased. For these polymers, the flexible pendant group reduces T9 by acting as “internal diluent”, lowering the frictional interaction between chains. Conversely, the branched pendant group of poly(isobutylglycolide) hinders rotation of the polymer backbone, resulting in a higher T9. 150 and 010 line We further investigated the properties of poly(ethylglycolide) using thermal and dynamic mechanical analyses. DSC runs show a T9 near 12 °C, with no sign of crystallinity. A T9 of 12 °C is lower than that of polylactide itself (66 °C), and in line with our expectation that increasing the length of the side chain should lead to a decrease in T9. DMA runs support the Tg assignment made from DSC data. As shown in Figure 47, the tan 5 trace for the polymer shows a peak near 12 °C which is nearly identical to the baseline inflection seen in the DSC data. In addition, the DMA probe position trace shows the expected behavior for a polymer heated to above T9, expansion followed by penetration of the probe through the sample. The decomposition of polymers measured using TGA define the limiting use temperature of the polymer. Figure 48 shows the TGA plots for the decomposition of poly(ethylglycolide) in air and under N2. The two data sets show only slight differences, an indication that the decomposition is probably dominated by depolymerization of the polymer to monomer. As shown in Figure 49, the TGA plots for all polymers are similar, with the onset for decomposition shifting to higher temperatures as the size of the alkyl group increases. We found that all polymers decomposed to their monomers, and thus the shift likely reflects the lower volatility of monomers that have large substituents. In trial depolymerization reactions run on 0.1 g scales, we recovered over 93% of polymer mass as monomers. 151 Poly(lactide) _.——-—'—'/‘-— Poly(isobutylglycolide) Poly(ethylglycolide) Poly(hexylglycolide) heat flow (endo) —> l j l l l L l I -80 -60 -40 -20 0 20 40 60 80 100 temperature (°C) Figure 46 DSC runs (second heating after flash quenching from 100 °C) for substituted poly(glycolide)s. Heating rate: 10 °C/min under helium. 152 Table 15. Polymer Properties Polymer Mnx10'3 3' MwIMn T, °C b polylactide 35.2 1 .89 22-65 Poly(ethylglycolide) 45.6 1.78 15 Poly(hexylglycolide) 43.2 1 .91 ~37 Poly(isobutylglycolide) 47.3 1 .83 22 a. measured by GPC in THF using polystyrene as standard b. measured by DSC under He at a rate of 10 °C/min 153 DSC probe position -55 -35 -15 5 25 temperature (°C) Figure 47. Thermal analysis results for poly(ethylglycolide). The bottom 3 traces are DMA results for a sample in a parallel plate geometry. Heating rate: 10 °C/min under helium (DSC) or N2 (DMA). 154 100 - 80- sample weight (%) 0 100 200 300 400 500 600 temperature (°C) Figure 48 Thermogravimetric analysis results for poly (ethylglycolide). Heating rate: 40 °C/min. 155 100 \" Poly(hexylglycolide) 30 r " S -o-0 Poly(lactide) —> I: 60 - .9 o 3 .23 40 r Poly(ethylglycolide) E :0 Cl) 20 - Poly(isobutylglycolide) 0 l L 200 250 300 350 400 450 500 temperature (°C) Figure 49 Thermogravimetric analysis results for substituted poly(glycolide)s run in air. Heating rate: 40 °Clmin. 156 00ml trent chai ternl brar cha than incr 9101 dist cha res sut Qro Sn1 gka the Ski Ser 9r0 The glass transition temperatures of the substituted polylactides and comparable substituted polyethylenes are shown in Table 16. There is a similar trend in the glass transition temperature with changes in the structure of the side chain. For polymers substituted with linear alkyl groups, the glass transition temperature decreased as length of the alkyl group increases. Conversely, branched pendant groups result in higher glass transition temperatures. The glass transition temperature is dependent on the flexibility of polymer chain and secondary forces between polymer chains. When the group changes from methyl to ethyl to hexyl group, the cross-sectional area of the group increases slightly. However, because of the linear nature of these substituted groups, they are very flexible. They act as “internal plasticizer" to increase the distance between polymer chains and lower the frictional interaction between chains. The net effect is to reduce the rotational barriers of the backbone, which results in a decrease of the glass transition temperature. For branched substituents, such as isobutyl and isopropyl, the cross sectional size of the groups is larger than the in linear counterparts, and the flexibility of the group is smaller, which result in a higher rotational barrier for the backbone and a higher glass transition temperature. Compared to isopropyl, the extra methylene unit of the isobutyl group increases the flexibility of the group, which decreases the glass transition temperature. There is one important difference between the substituted polylactide series and the polyethylene series. Polylactides with branched substituted groups, poly(isobutylglycolide) and poly(isopropylglycolide), have lower glass 157 transiti01 poiy(isol tempera betweer strong 0' This int transitio strong, further iSOprop interact DOIy(isc barrier weaker transitii the int; of ihe the p°iypri transition temperatures than polylactide. In the substituted polyethylene series, poly(isobutylethylene) and poly(isopropylethylene) have higher glass transition temperatures than poly(propylene). The difference lies in the secondary forces between polymer chains in both series. Polylactides are polyesters, which have strong dipole-dipole interactions between polymer chains as shown in Figure 50. This interaction hinders rotation of the backbone, which increases the glass transition temperature of polymer. When R is small, the chain-chain interaction is strong, but when R is large as in the isobutyl and isopropyl cases, the chains are further separated and the interaction is weak. The larger size of isobutyl and isopropyl groups increases the rotational barrier, but the weakened secondary interaction between polymer chains decreases the rotational barrier. Thus, for poly(isopropylglycolide) and poly(isopropylglycolide), the increase in rotational barrier gained from increasing the size of the substituted group can not offset the weakened interaction between polymer chains, which results in a lower glass transition temperature than polylactide. For the substituted polyethylene series, the interaction between polymer chains is very weak because of the low polarity of the polymer backbone. The rotational barrier is mainly decided by the size of the substituted group. That is why poly(isopropyethylene) and poly(isobutylethylene) have higher glass transition temperatures than polypropylene. 158 Table 16. Glass transition temperature of substituted polylactides and substituted polyethylene. polymer T9 (°C) polymer To (°C) polylactide 45-65 Poly(propylene) -15 - -3 Poly(ethylglycolide) 12 Poly(1-butene) -50 Poly(hexylglycolide) -37 Poly(hexylethylene) -65 Poly(isobutylglycolide) 23 Poly(isobutylethylene) 29 Poly(isopropylglycolide) 50 Poly(isopropyethylene) 50 159 Figure 50. The secondary interaction between polyester chains 160 5. Cop< twor0' togefli asymi copol 5. Copolymerization of substituted glycolides The copolymerization of substituted glycolides has been done using the two routes shown in Scheme 38. In route 1, the two comonomers were mixed together and heated up with catalyst to make copolymers. In route 2, the asymmetrically substituted glycolides were synthesized and polymerized to make copolymers. Scheme 38 Copolymerization of substituted glycolides o o R‘ #TLO Sn(Oct)2/ROH R1 + \n/kRz A o o RTE]: Sn(Oct)2/ROH \n/iRz ROW/RC0 (2) 161 5.1. copc PIOI E0. 001' MC inl si 5.1. Copolymerization through comonomers Ethylglycolide (EG) and Iactide (LA) have very similar structures. Their copolymers should retain the biodegradability, but have different physical properties than their homopolymers. A series of copolymers (EG/LA=1I5, EG/LA=1/3, EG/LA=1, EG/LA=3, EG/LA=5) were synthesized by bulk polymerization at 130 °C catalyzed by Sn(Oct)2/BBA. All polymers have molecular weights >40,000 g/mol. The DSC plots are shown in Figure 51. The DSC runs show one glass transition temperature for each copolymer, an indication that the copolymer is not phase separated. Considering the structural similarity of the monomers, it is not surprising that copolymers are single phase. The glass transition temperatures listed in Table 17 fall those of polylactide and poly(ethylglycolide) and increase with an increase the mole fraction of the Iactide comonomer. The glass transition temperatures of copolymers can be predicted using the Fox equation: where T9, Ty. and T93 are glass transition temperature of the copolymer, and the pure homopolymers derived from monomers A and B respectively, and w represents the mass fraction of the polymer. The Fox equation successfully predicts the glass transition temperatures of random copolymers and plasticized systems. An underlying assumption is that the polymer is homogeneous in composition, and that the comonomers are distributed randomly along the chain. 162 As sh equa‘ or nr 5805 and far As shown in Figure 52, the glass transition temperature of copolymers fit the Fox equation fairly well. However, whether the copolymer is truly a random copolymer or not must be defined by a characterization technique such as NMR, which is sensitive to structure at the molecule level. Based on copolymerizations of Iactide and glycolide]:189 the distribution of monomer units in the polymers is probably far from random because differences in reactivity for the monomers. As illustrated in Scheme 39, these effects are likely due to differences in the steric bulk at the carbon or to the carbonyl. Lactide and glycolide share the same ring structure, but the absence of lactide’s methyl groups makes glycolide far more reactive and thus “blocky” polymers are obtained from copolymerizations. However, transesterification reactions can increase the randomness of copolymers. 163 Table 17. Properties of copolymers Polymers M.,x10'a 3 "Jun T9 °C b Polylactide 35.2 1 .89 66 EGILA=1/5 41.5 1.85 42. EGILA=1I3 47.8 1.76 38 EGILA=1 41.7 1.88 30 EGILA=3 50.8 1 .92 23 EGILA=5 43.6 1.85 19 Poly(ethylglycolide) 45.6 1 .78 15 a. measured by GPC in THF using polystyrene as standard b. measured by DSC under He at a rate of 10 °Clmin 164 Polylactide _.—-v—'—"—'/—_— EGILA=1/5 ._.__._»/—‘_‘—_ EGILA=1/3 Jm EGILA=1 f/h— endo —> Poly(ethylglycolide) f l l -50 0 50 100 Temperature ( °C) Figure 51. DSC runs (second heating after flash quenching from 100 °C) for Iactide and ethylglycolide copolymers. Heating rate: 10 °Clmin under helium. 165 70 50 1 curve from A Fox uation 0 eq 0 ‘5 I'- 30 5 10 I U U I 0 0.2 0.4 0.6 0.8 1 W1/(W1'i'W2) Figure 52. Glass transition temperatures of Iactide and ethylglycolide copolymers. 166 Scheme 39. Reactivity difference leads to “blocky” copolymers. RO‘ less more hindered hindered polymerizes preferentially glycolide rich Iactide rich 167 5.2 5.2 Copolymerization through asymmetric monomers. Simple copolymerization has some shortcomings. If one of the comonomers is not polymerizable, the copolymer cannot be made. Also, the mismatch in reactivity between the comonomers leads to inhomogeneous incorporation of the comonomers into the polymer chain. These problems can be solved by designing the AB monomers shown in Scheme 40. A variety of AB monomers have been synthesized. As shown in Scheme 41, each monomer contains two sites for ring opening, and to a first approximation the two carbonyls of the ring can be treated independently. One site should have a reactivity similar to Iactide, while the other should more closely resemble that of a substituted Iactide. Thus we expect that the active end of growing Iactide chain will attack the least hindered site on the AB monomer (the Iactide carbonyl) at a rate somewhat slower than for Iactide, and ring will open to give the substituted lactic acid residue at the growing chain end. AB monomers also have limitations. AB monomers only produce copolymers whose composition is 50% A and 50% B. Using mixtures of comonomers can produce copolymers with any composition of A and B. 168 Sci Scheme 40. The comparison between two copolymerization methods 111$. 2.311941. Because monomer A is not polymerizable, the compolymer can not be synthesized O O _.__’ O 0% O n 0 Using an AB monomer, the copolymer can be synthesized Scheme 41 . Structures of the AB monomers and reactivity of different sites 0 o O O O O \[I/‘V o O O O tflmgthngchofldg ethylmethylglycolide methylphenylglycolide hindered site ( for ring -opening 0 1L 0 kg— preferred site of ring-openning 169 5.2.1 CODI syn trar (I) 5.2.1. Synthesis and Polymerization of AB Monomers. As shown in Scheme 42, the AB monomers were prepared by condensation of a—bromopropionyl bromide and the desired lactic acid. In this synthesis, it is important to remove all of the NEta during the work-up, since even trace amounts of NEta will polymerize the monomer to low molecular weight oligomers. Racemic a-bromopropionyl bromide and racemic lactic acids were used in the syntheses. As shown in Figure 53, the monomers are not statistical mixtures of R,R, 8,8 and R,S diastereomers. In the case of ethylmethylglycolide, the diastereomers form in a 3:1 ratio. For methylphenylglycolide, the selectivity is so high that the minor diastereomer is barely seen in the 1H NMR spectrum. After recyrstallization, the pure diastereomer can be obtained. To identify the major isomer, we obtained the Nuclear Overhauser Effect (NOE) difference NMR spectra of methyphenylglycolide. As shown in Figure 54, methines H8 and Hi, can be attached to the same side of the ring or opposite site of ring. If Ha and H, are on the same side of ring, the signal of the Hb will be enhanced if H. is radiated. As shown in Figure 55, the intensity of H), increased when H, was radiated. Thus, the major diastereomer is a mixture of the R,R and 8,8 isomers, and the minor isomer is the R,S, or SR isomer. The kinetics of ring-closure probably favor the formation of the R,R and 8,8 isomers, and ring-closing competes with formation the oligomers. The selectivity can directly result from the ring closing step, or the selectivity can arise from different rates for conversion of the linear dimers to Iactides or oligomers. 170 high stow) knov ope trim poi Ethylmethylglycolide and methylphenylglycolide can be polymerized to high molecular weight polymer easily. However, trimethylglycolide polymerizes slowly. Even after 24 hours at 180 °C, the conversion is only about 75%. It is known that the more substituents ring has, the more difficult it is for the ring to open. Lactide and other substituted Iactides all are di-substituted, and trimethylglycolide is tri-substituted. One extra substituent makes it hard to polymerize. 171 Scheme 42. Synthesis of AB glycolide monomers O OH + Br R2 R1)\[r Rz/Kn/ —Z—’ >—-"/kRi O 0 Br 0 acetone, NEt3 O RK‘JK O 0\n/i\ R1 0 O H O 0 HI, 0 CH3 O—goicm 0—<\ 011., (R, R) (R, S) Figure 54 Structure of the diastereomers of methylphenylglycolide 172 Minor iSt Ha Major isomer Hb / i J Minor isomer __., i a 2’ - 111. Major isomer Minor isomer 0 ' ii 1 ‘i I 0% 11 . ‘. i " 1 . Ii ’ ,I: 1": . ii. 15‘”. .' ll ii 0 ~ V: L ’ r r. ‘1. l iv \/ =4. ”"J;J J / ".J ,i 1'11 1 WflJw/“L-I' TrTTYTIFTTrIITfFFIT—Tl T I I T l T—TTT—Tfi'T—fl 5.1 5.0 4.9 4.8 T T I I T T T T l T I T Y i T I I l T T T 1 T I T T T I Y I I I Figure 53. 1H NMR spectra of ethylmethylglycolide and methylphenylglycolide 173 ou__oo>_m_>cocq_>£oE ho E4583 mzz m0: .3 239.... I'ti bl? 1') [7' L It Iii t iii | 174 5.2.2 Str Who reactions 3 reiative siz smaII, 101 °°P0iymer dominates If either F reaction I form. is not Ia know [I po‘Yietr R1 is m imp. mande DOIy(e1 dEieCt piliyml reSidu 5.2.2 Structure of the polymer chains. When AB monomers polymerize, there are four possible propagation reactions as shown in as Scheme 43. Which reaction dominates depends on the relative size of the R1 and R2 side group. If the size difference of R1 and R2 is small, four reactions co-exist in the polymerization leading to random copolymers. If the size difference of R1 and R2 is large, one of the four reactions dominates, and an alternating (AB)n polymer with head to head defects will form. If either R1 or R2 is so large that attack from active center is totally blocked, one reaction exists in the polymerization, and a perfectly alternating copolymer will form. For ethylmethylglycolide, R1 is methyl and R2 is ethyl. The size difference is not large, and from the homopolymerization of Iactide and ethylglycolide, we know the reactivity difference of two monomers is not large. We expect that poly(ethylmethylglycolide) is a regio-random polymer. For methylphenylglycolide, R1 is methyl and R2 is phenyl. The phenyl group is much larger than the methyl group, but from studies of the polymerization of mandelide, we know that the mandelic acid residue should have some reactivity. We expect that poly(ethylmethylglycolide) is an alternating polymer with some head to head defects. For trimethylglycolide, we know that tetramethyglycolide is not polymerizable, so we expect that active center can only attack the lactic acid residue and give poly(trimethylglycolide) as a perfectly alternating polymer. 175 The Tra 001 ad; an 16 till T1 si tc fr We used 13C NMR to determine the chain structure of the AB copolymers. The carbonyl region of the 13C NMR of AB copolymer is shown in Figure 56. Trace A is the 13C NMR spectrum of poly(ethylmethylglycolide), which is very complicated. There are signals from two different carbonyl groups: one adjacent to methyl group with chemical shifts from 169 ppm to 169.8 ppm and one adjacent to the ethyl group with chemical shifts from 168.2 ppm to 169 ppm. Both carbonyl groups show multiple peaks, which are caused by the randomness of the chain structure and the stereochemistry of the chain. Trace B is the 130 NMR spectrum of poly(trimethylglycolide), which shows a simpler pattern. Again, there are signals from two different carbonyl groups: one adjacent to the methyl group with chemical shifts from 168.7 ppm to169.6 ppm and one adjacent to the dimethyl groups with chemical shifts from 170.3 ppm to 171.3 ppm. Since tetramethylglycolide is not polymerizable, we think that the nucleophilic attack ocurred exclusively at the Iactide residue and that poly(trimethylglycolide) may be a perfectly alternating polymer. Although the carbonyl resonances also show multiple peak patterns, this may be caused by the stereochemistry of the chain, because racemic monomer was used. Trace C is the 13C NMR of poly(methylphenylglycolide). Both carbonyl groups have multiple peak patterns, although the resolution is not very good. Because the monomer used in synthesis of AB copolymers are not optically pure, the spectra contains the information of both stereochemistry of chain and structure of chain. It is very difficult to determine the structure of chains. To 176 U03 are Sch unambiguously determine the structure of the chain, optically pure monomers are needed. The synthesis of optically pure monomers is still ongoing. Scheme 43. The propagation step in the polymerization of AB monomers 1 ___> head to tail ——2—> head to head O 4 O R‘ ——-> head to head R2 w 0—2: R2 _:L, head to tail [Mi]: 3 177 T I l V V V I T f fi T I V V V I ' V I T V V V l V fi V V I" f r I V V l 170.5 169.5 168.5 167.5 ppm it . i l T". I ,1; i j t. 1 i i " a r. i l _\" \ ’ v’ \ \ "x W,va.. A A/J VA] \w‘—/ \WW ’J’J \W N/ J M) \\m l I I l I 1 l I I Y I Y I I I Y I ' ' ‘ 171.5 170.5 169.5 168.5 ppm .1 C i ‘1 I l . J 1 . I i / ‘v’ . i \\ a" i «I: it i i I \_ ‘ii .‘ VA~ f] j " (I '1 rj/ \ H’l 2" J 1‘ x ,r \m .nrvx/‘mm_~~ Maw" ‘ww/‘N/V'W’r'm ‘NvW-I‘MAW'KNM/y \‘\k.-.V~W.AN T T I T T Y T T T T 1' Y I I I Y Y I I f f T I I I I T I T 1 7 T I 7 I I Y 7 7 Y I I I Y Y 1 170.5 169.5 168.5 167.5 166.5 ppm Figure 56. Carbonyl region of 13C NMR of AB copolymers. A is poly(ethylmethylglycolide), B is poly(trimethylglycolide) and C is Dolv(methvl0henvlolvcolide) 178 5.3 The E copolyr greatly Poly(rr Aithou its sh 4,000 transi expe1 isom How. from Optic the is v. or i W Ch: de dif tht Ti 5.3 Thermal properties of the copolymer. DSC was used to measure the glass transition temperatures of the copolymers. The results are shown in Figure 57 and Table 18. Bulky side chains greatly increase the glass transition temperature of polymers. Poly(methylphenylglycolide) has a glass transition temperature at 85 °C. Although, poly(trimethylglycolide) shows a glass transition temperature at 50 °C, its should be higher, since the molecular weight of polymer was low (about 4,000). Polymers with molecular weights over 20,000 should have higher glass transition temperatures. None of the polymers were crystalline, which was expected, since the ethylmethylglycolide was a mixture of 8,8, R,R, and R,S isomers. Trimethylglycolide and methylphenylglycolide were also racemic. However, we made (S,S)-methylphenylglycolide and (R,R)—methylphenylglycolide from L-mandelic acid and D-mandelic acid, and the copolymers from these optically pure AB monomers are still amorphous. There are two possibilities for the lack of crystallinity in these copolymers. First, the proton or to the phenyl ring is very reactive, and at high temperature, this proton may be lost to form a radical or ion. In either case, the polymers are going to be racemized, which inhibits crystallization. During bulk polymerization at 180 °C, the reaction mixture changed from colorless to dark brown very quickly. We think it was caused by the deprotonation. Second, because of transesterification and because the reactivity difference for attack at the hindered and less hindered side is not large enough, there are many head-to-head and head to tail placements in the polymer chain. Thus, the polymer chain is not ordered enough to crystallize. 179 The Polyttrimeti; the low m about 6% l the impurit thermally é Poly Poly The decomposition temperatures of the polymer are shown in Figure 58. Poly(trimethylglycolide) has the lowest the decomposition temperature, due to the low molecular weight of the polymer. Poly(methylphenylglycolide) shows about 6% residue at 370 °C, which then disappears at 600 °C. It is not caused by the impurity such as catalyst, it is well-known that aromatic polymers often form thermally stable compounds when thermally decomposed. Table 18. The properties of AB polymers. Polymer M..x10'3 a Min/Mn T, °C ” Poly(ethylmethylglycolide) 31 .2 1 .81 85 Poly(trimethylglycolide) 35.6 1 .85 50 i Poly(methylphenylglycolide) 4.2 1.25 22 a. measured by GPC in THF using polystyrene as standard b. measured by DSC under He at rate of 10 °Clmin 180 methyl phenyl glycolide trimethylglycolide J ethyl methyl glycolide T I T -10 30 70 110 temperature ( °C) ’uhH’ endo —-. Figure 57. DSC runs (second heating after flash quenching from 100 °C) for A8 polymers. Heating rate: 10 °Clmin under helium. 181 100 - -— Poly(methylphenylglycolide) \ _Poly(ethylmethylglycolide) (D O t O) O A O - Poly(trimethylglycolide) weight percentage (%) N O ‘r v O ' “ 0 200 400 600 temperature ( °C) Figure 58. Thermogravimetric analysis results for AB polymers run in air. Heatina rate: 40 °Clmin. 182 6.1 Syr I polyme polyme crystal availal amino optica Synth with t This meth litera fOUn 88p; qua- Was last 180' rag She 6. Crystalline substituted polylactide. 6.1 Synthesis and polymerization of optical pure monomers All of the substituted polylactides mentioned above are amorphous polymers because the monomers used to make the polymers are racemic. If we polymerize optically pure monomers, the resulting polymers should show some crystallinity. Since most optically pure 2-hydroxyacids are either not readily available or are very expensive, we prepared optically pure 2-hydroxyacids from amino acids. Optically pure amino acids are easily available and cheap, and the optically pure monomers can be obtained through simple chemical reactions. The synthesis of a monomer from valine is shown in Scheme 44. Valine was treated with NaNOz in acidic aqueous solution to yield the corresponding 2-hydroxyacid. This reaction proceeds with retention, and thus L-valine yields L-2-hydroxy-3- methylbutyric acid. The synthetic procedure used is a modified version of a literature preparation. The literature preparation uses HCI as the acid, but we found that about 10% of the product is the 2-chloroacid, which is very difficult to separate from the 2—hydroxyacid. Switching to dilute H2804 gave nearly quantitative yields of the 2-hydroxyacid. The dimer of 2-hydroxy-3-methylbutyric acid is isopropylglycolide, which was synthesized using the same method used to synthesize other substituted lactides such as ethylglycolide. Because we wanted to obtain optically pure isopropylglycolide, the high temperature cracking process used in making racemic monomers is inappropriate since it leads to some epimerization of the stereocenter. 2-Hydroxy-3-methylbutyric acid was condensed in toluene using p- 183 toluenesulfonic acid as catalyst to give a mixture of the dimer and low molecular weight oligomers. The resulting glycolide was crystallized directly from toluene to give pure S-isopropylglycolide. rac—2-Hydroxy-3-methylbutyric acid gave a mixture of R,R, 8,8 and R,S isomers, and after recyrstallization from ether, only pure rac-isopropylglycolide (1 :1 mixture of R,R and 8,8 isomers) was obtained. Melt polymerizations carried out at 180 °C using Sn(Oct)2/BBA as the catalyst/initiator reached 95% conversion after about one hour. At the beginning of the polymerization, the mixture was an easy to stir liquid, but after several minutes, the mixture became very viscous. After 10 minutes, the polymerization mixture solidified and was impossible to stir. The solidification was caused by crystallization of the polymer, which has a melting point >180 °C. The details of crystallization will be discussed later. The molecular weight distribution of the resulting polymer is very broad, with a polydispersity about 3, probably due to crystallization of polymer. Because of crystallization, the polymerization reaction can become diffusion controlled. If the diffusion rate of monomer is much slower than the polymerization rate, monomer is not evenly distributed in the mixture. Thus, polymerization sites have different propagation rates leading a mixture of long and short polymer chains and a broad molecular weight distribution. NMR spectra of poly(S-isopropylglycolide) are shown in Figure 59. The top trace shows the methine proton region for the homo decoupled 1H NMR spectrum, and the bottom trace shows the carbonyl region of the 13C NMR spectrum. We conclude that there is almost no racemization during the polymerization, because only one peak is seen in both spectra. 184 Scheme 44. The synthesis of an optically pure substituted Iactide from an amino acid precursor 185 TT'UIITTTTTTITWTTTTVTII'YUTl'UI'T'lrl’Ill'TrUllU . . I . . . . 5.2 5-1 5.0 4.9 4.8 ppm VIVWTTYI‘FTYV [ I V H... ......, “1-... ....,...f, 168.9 168.7 168.5 168.3 168.1 ppm Figure 59. NMR spectra of the methine (top) and carbonyl (bottom) region of poly(S-isopropylglycolide) 186 6.2 The stereochemistry of the polymer chain A number of the physical properties of poly(lactide) are linked to its stereosequence distribution in the polymer chain. For example, pure isotactic poly(L-Iactide) crystallizes at faster rate and to a larger extent than when L- lactide is polymerized with small amount of either D—lactide or meso—lactide."90 in the 1H and 13C NMR spectra of polylactides, the observed resonances can be assigned to various stereosequence combinations in the polymer. The assignments are designated as various combination of “i” isotactic pair-wise relationships (-RR- and -SS-) and “s” syndiotactic pair relationship (-RS- and - SR-). In the NMR spectra, the -RR- and -SS- diads and -RS- and -SR- diads are indistinguishable and have identical chemical shifts. The assignment of resonances for polylactides up to hexads has been made by several research groups.137:191'194 Bernoulli and first-order Markov models have been used to rationalize the assignments. In Figure 60 are shown the building of a polymer chain by a Bernoulli and a first-order Markov process. In the Bernoulli model, the stereochemistry of chain end is not important, which means that stereochemistry of the chain end does not influence the addition of the monomer to the chain end. In the first—order Markov model, the stereochemistry, which may be i or s is strongly influenced on the adding monomers to the chain end. 187 My. ' ‘- i.0 /d|> 6' \ . First-order Markov model Figure 60. Schematic representation of the Bernoulli and first-order Markov model. 188 The physical properties of poly(isopropylgycolide) are also dependent on the stereosequence of the polymer chain. As shown in Figure 61, pure poly(S- isopropyglycolide) has a melting point around 230 °C. The melting point of poly(isopropyglycolide) decreases when S-isopropylglycolide is polymerized with a small amount of rec-isopropyglycolide. From measurements of the heat of fusion, we found that the extent of crystallization also decreased. As was done with polylactide, we would like to assign the stereosquences of poly(isopropyglycolide) to understand the relationship between the physical properties of poly(isopropyglycolide) and stereochemistry of the polymer chain. The stereochemical assignments were made by comparing the trends observed in the spectra of a number of poly(isopropylglycolide)s to the probability distribution expected on the basis of their Iactide feed composition. Shown in Figure 62 and Figure 63 are the carbonyl region of the 13C NMR of poly(isopropyglycolide) and the methine region of 1H NMR of poly(isoproplyglycolide). Trace A in each figure is spectrum of the polymer prepared from 100% S-isopropyglycolide. The single peak in both spectra can be identified as resulting from an iiiii sequence. The B traces are spectra for the polymers prepared from 85% R-isopropylgycolide and 15% racemic isopropyglycolide. In additon to the iiiii peak, several new peaks appeared. Because the amount of S-isopropyglycolide is small, probability dictates that these peaks should be isiii and iiisi sequences. By comparing the spectra from samples with different ratios of S-isopropylglycolide and rac-isoproylglycolide, we assigned all 11 hexads of the polymer chain. The result is shown in Figure 64. 189 Because the resolution of 1H NMR spectrum is not as good as for the 13C NMR, we only assigned stereosequences in the 130 NMR spectrum. The unusual intensitry of the iiiii hexad in the poly(rac-isopropyglycolide) spectrum suggests that the polymerization follows non-Bernoullian statistics and stereoselection occurs during the polymerization of rec-isopropyglycolide. We used the first-order Markov model to describe such a process. We assume that the probabilities for an RR monomer adding to an RR chain end and an SS monomer adding to an SS chain end are p, and the probabilities of an RR monomer adding to an SS chain end and an SS monomer adding to an RR chain end are (1-p). It is possible to calculate the expected intensity values of the individual sequences as shown below: iiiii = p3 + 0.5p2(1-p) iiiis = siiii = iisii = o. 5p2(1-p ) iiisi = isiii = 0.5p2(1-p) + 0.5p(1-p)2 iisis = siiis = sisii = o. 5p(1-p)2 isisi = 0.5p(1-p)2 + 0.5(1-p)3 sisis = 0.5(1-p)3 We compared the experimental values and calculated values using equation 2[(lca.-lexp)/lexp]2, where lea. is calculated peak intensity and lexp is experimental peak intensity. We found that when p=0.61, the lea. is closest to law. The result is shown in Table 14. It indicates a preference for addition RR 190 monomer to the RR chain end and SS monomer addition to SS chain end, which enhances contribution of isotactic segments in poly(isopropylglycolide). This result is very different from that obtained for polylactide, where several authors reported that the Iactide polymerization favors syndiotactic chain segments (p<0.5). The reason for this difference is not clear and is still under investigation. Table 19. Experimental and calculated values of hexad intensities in the carbonyl region of 13C NMR spectra of poly(rac-isopropygcholide) iiiii+siii iiisi sisii isiii-I-iisii iisis+sisis isisl i+iiiis +siiis Experimental 46.2 11.8 4.4 22.3 7.07 8.25 values Calculated values 44.4 11.9 4.6 23.8 7.61 7.61 (p=0.61) 191 100:0 93:7 endo ——> 85:15 100 150 200 250 temperature (°C) Figure 61. DSC runs showing the melting point for poly(isopropyglycolide)s prepared from S and rac-isoproplyglycolide. The ratios are the Szrac content in the polymers. 192 t.’ C O C rac-isopropylglycolide H >§\ n B 85% S-isopropyglycollide 15% rac-isopropyglyoolide A 100% S-isopropyglycolide A 168.9 168.7 168.5 168.3 168.1 ppm Figure 62. 13C NMR of poly(isopropylglycolide) 193 O C rac-is0propylglycolide {L H n z-.- ,4,-_-_._J”\/x_ ,,- ._ ht r __..fi 8 85% S-isopropylglyoolide 15% rac-isopropylglycolide A 100% S-isopropyglycolide A L lelllTTlTlTTITTTTITITTIFTTTITTTTITTIIITIIIIITTl—I 5.20 5.15 5.10 5.05 5.00 4.95 4.90 4.85 4.80 ppm Figure 63. ‘H NMR of poly(isopropylglycolide) 194 _ L $2_oo>_m_>ao.aom_.oflv>_oq Co 828on on. .8 “cascgmmm 85:389on .3 23m.“— EQQ m.me odoa h.mmfi m.me m.mmH 9me .7me N.moH m.mmH v.mmH m.mmH odofl ~2me m.mwa «3de LhP-Frbphpp—Pblhh—i—i—pp—--h—PPPL.—bL~P_F-PFF-uph—phhp—thh—rhb-p—PF>»~»FL——bpthpph EocoQEoo .m:u_>__uc_ ME» .5wa 9w: 528on 8.936% 830on .mBom 195 6.3 The crystallinity of poly(isopropyglycolide) We used X-ray diffraction to determine the crystallinity of poly(R- isopropyglycolide). This method has been used by Herrnans and Weidinger to measure the crystallinity of polyethylene,195 polypropylene,196 and polystyrene.197 The X-ray diffraction pattern of crystalline poly(isopropylglycolide) is shown in Figure 65. A valid procedure for subtracting the amorphous scattering from the total scattering is always the first and most essential step in any attempt to derive a measure of crystallinity from X-ray diffraction. However, unavoidably, it always involves certain uncontrollable assumptions. The X-ray diffraction pattern for amorphous poly(isopropyglycolide) made from rac-isopropyglycolide is shown in Figure 66. The amorphous sample shows a broad peak with maximum at about 12 degrees. We assumed that the amorphous fraction in the crystalline sample should have the same maximum and the same line-shape. To make the problem easier, we selected a 10-degree window, 6-16 degrees, for the calculation. PeakSolve software was used to deconvolute the diffraction pattern into amorphous and crystalline components. The result is shown in Figure 67. The area under amorphous peak is Os.m and the area under three crystalline peaks is O”. The degree of crystallinity is calculated by comparing the results from two or more samples of the same polymer with crystalline fraction X1 and X2, where X1-X2 is as large as possible. Then, X1 and X2 can be described using following equafion: X1/X2=Ocr1/Ocr2 196 (1‘X1)/(1'X2)=Oam1/Oam2 The crystallinity of several different poly(isopropylglycolide) samples are shown in Table 20. We combined the crystallinity data and the heat of fusion measured using DSC to estimate the heat of fusion for a 100% crystalline sample. By plotting the heat of fusion vs. crystallinity of sample, we use data from Figure 68 to estimate the heat of fusion for 100% crystallinity. The result, 43.6.J/g, is about half that of polylactide’s 100 J/g. Table 20. The crystallinity and heat of fusion of poly(isopropylglycolide) Samples Crystallinity ' (%) Heat of fusion (AH Jig) 1 46.9 22.1 2 35.2 17.9 3 27.8 14.3 a. estimated from X-ray analysis of samples 197 4500 4000 - 3500 - 3000 - 2500 - 2000 - 1500 - intensity/second 1000 - 500 - 0 l l I I 0 10 20 3O 40 50 26 Figure 65. X-ray diffraction pattern for crystalline poly(S-isopropylglycolide) 198 350 300 - 250 - 200 - 150 F intensitry/second 100 - 50- o l l l 0 1 0 20 30 40 26 Figure 66. X-ray diffraction for amorphous poly(rac-isopropylglycolide) 199 4000 h (D O O O I 2000 ’ amorphous peak . Intensity/second 1000 26 Figure 67. Peak deconvolution for crystalline poly(S-isopropylglycolide) 200 25 AH1oo = 43.6 Jlg 3 3 20 - r: .2 in I .2 h o H a 0 15 r I 10 l 1 20 30 40 50 crystallinity (%) Figure 68. Determination of heat of fusion for 100% crystalline poly(isopropylglycolide) 201 7. The degradation of substituted polylactides The degradability of the substituted polylactides was evaluated by hydrolytic degradation in phosphate buffer solution (0.01M KHzPOJKzHPO4 solution, pH=7.4) at 55 °C. To exclude the influence of autocatalysis. we chopped the polymer samples into small pieces. The auotcatalytic effect is caused by the increase in acidity when the degradation products are acidic and have limited diffusion. If polymer samples are small enough, the diffusion rate of buffer solution into the bulk is higher than the polymer chain scission and the hydrolytic degradation can proceed uniformly in the bulk of the specimen. The morphology of the polymer also influences the degradation of polymer. The crystalline polymer degrades much slower than the amorphous polymer. The much tighter packing of the crystalline phase makes water much harder to penetrate into the bulk of the polymer. To study the influence of chemical structure of polymer on degradability, we have to eliminate the influence of morphology. To do that, we used amorphous substituted polylactides prepared from racemic monomers. All polymers were tested using DSC to ensure that the polymers were amorphous. The weight loss and molecular weight decrease with degradation time are two of most important properties of biodegradable polymers. The weight and molecular weight loss versus time are shown in Figure 69 and Figure 70, respectively. The sample weights were constant for a period of time, and then began to drop. The molecular weight of polymer dropped immediately once the degradation begins. These are typical behavior for hydrolytic degradation. 202 To calculate the degradation rate constant, we used the random scission model to fit the data. For the random scission model, the average number of bond cleavages per polymer molecule (N) was calculated according to: N=[Mn(0)/Mo(t)-1]= Kail"n(0)t where Mn(0) and mm are the number-average molecular weights of polyesters at time 0 and time t. kd is the rate constant for hydrolytic degradation, and P..(0) is the number-average of degree of polymerization at time 0. K, was obtained from plots of [Mn(0)/Mo(t)-1]/P,,(0) versus t. The slopes of the plots are the rate constants for hydrolytic degradation. The results are shown in Figure 71. From the figure, we can see that polylactide degrades fastest, poly(ethylglycolide) and poly(hexylglycolide) have almost the same degradation rate and slower than poly(lactide), and poly(isobutylglycolide) degrades the slowest. An interesting phenomenon showed up when remaining weight fraction was plotted vs. degree of polymerization, as shown in Figure 72. The weights of all polymers began to drop at almost the same degree of polymerization, which is about 30. We think this is critical molecular weight for entanglement. When the degree of polymerization is higher than 30, almost all of the polymer chains are entangled, so polymers will not dissipate into the water and there is almost no weight loss for the polymers. When the degree of polymerization of polymer is 203 lower than 30, some of polymer chains are not entangled, so these chains will dissipate into water and there is weight loss for the polymers. How the substituent influences the degradation rate of substituted polylactides is the question we need to answer. We think there are two factors that influence the degradation rate: the glass transition temperature and the surface energy of the polymers. At the degradation temperature, the free volume will be larger in polymers with a relatively low glass transition temperature than those with high glass transition temperatures. Thus, low T9 polymers will have a higher water content, which causes the higher degradation rate. If a polymer has a very low surface energy (hydrophobic), the solubility of water will be low and the degradation rate will be low. The glass transition temperatures of the substituted lactides are shown in Table 21. The order of glass transition temperature is Iactide > isobutylglycolide > ethylglycolide > hexylglycolide. Based on glass transition temperature, we would expect the order of the degradation rate is hexylglycolide > ethylglycolide > isobutylglycolide > Iactide. However, besides the glass transition temperature, we also need to consider the surface energy of the polymers. We used contact angle method measurements to characterize the surface energy of the substituted polylactides. We make a series of water/methanol solutions with different water/methanol ratios. The surface tensions of these solution were measured, and then these solution were dropped on the polymer films. The contact angle between the solution and polymer film 204 were measured, and the surface energy of the film were calculate using the Zisman equation: 0089 = 1 + marc) where 9 is contact angle, 71. is surface tension of testing solution (water/methanol solution). yo is the surface energy of the polymer surface, and B is constant. Values of cost) vs yL were plotted for each substituted polylactide shown in Figure 73, and then the lines were extend to cost) =1. The intercept 7., value is the surface energy of polymer (Table 21). Table 21 . The degradation of substituted polylactide polymer Glass transition Surface energy Degradation temperature(°C) y (dynlcm) rate (day'1) x103 polylactide 65 21 .6 3.7 Poly(ethylglycolide) 1 5 19.9 2.6 Poly(isobutylglycolide) 23 18.5 2.4 Poly(hexylglycolide) -37 16.2 1 .0 205 The length of linear alkyl substituted group increases from Iactide to ethylglycolide to hexylglycolide, which causes the glass transition temperature and surface energy to decrease. However, these two factors have opposite effects. Decreasing glass transition temperature increases the degradation rate, and decreasing the surface energy decrease the degradation rate. So a combination these two factors may explain the observed order of degradation For poly(isobutylglycolide), the branched the side chain gives the polymer a relatively high glass transition temperature and low surface enerQY; these two factors work in concert to slow the degradation rate. It is not surprising that poly(isobutylglycolide) has slowest the degradation rate. 206 1 1 . o O O A . ' O I I O t s c \a .3 0.8 . \\ 0 t! 5 IL :5, 0.6 - 0 3 g) A“ N 8 , o 9‘ 0.2 - e\ 0 1 ‘ 1 0 20 40 60 Time (days) Figure 69. The weight loss of substituted glycolide during hydrolytic degradation. (A) Iactide, (I) ethylglycolide, (O) hexylglycolide, (O) isobutylglycolide. 207 7. 1501;01:286me T‘ ‘ [Poly(etl'iylglycolide) ; 0.8 - oPoly(isobutylglycolide) ‘ o Poly(hexylglycolide) l a 0.6 - ‘E’ 2 2. O t; c 2 0.4 A 0.2 ~ 0 \‘ ‘u o 1 ME 0 10 20 30 40 Time (days) Figure 70. The molecular weight loss of substituted glycolide during hydrolytic degradation. (A) Iactide, (I) ethylglycolide, (O) hexylglycolide, (O) isobutylglycolide. 208 0.1 I 0.08 - 8 ‘c’ a /‘ c 0.06 - ‘ I ,L a: E: e a 004 ’ / 7; I a A 0.02 - ‘ 0 l L l 0 10 20 30 40 Time (days) Figure 71. The molecular weight decrease of substituted lactides during hydrolytic degradation fitted to the random chain scission model. (A) Iactide, (I) ethylglycolide, (O) hexylglycolide, (O) isobutylglycolide 209 "=30 A :0 I ‘0”- .0 oo F’ 5: -b O) Remaining Weight Fraction (3 iv 1000 100 10 1 Pn Figure 72. The decrease in the degree of polymerization of substituted Iactides during hydrolytic degradation. (A) Iactide, (II) ethylglycolide, (I) hexylglycolide, (O) isobutylglycolide 210 60 40- 10- 0 I l I l 0 0.2 0.4 0.6 0.8 1 cos(6) Figure 73. Contact angles on substituted polylactides plotted as 7 (surface tension of water/methanol solutions) vs cos 9. (A) Iactide, (I) ethylglycolide, (O) hexylglycolide, (O) isobutylglycolide 211 EXPERIMENTAL 1. General Unless otherwise specified, ACS reagent grade starting materials and solvents were used as received from commercial suppliers without further purification. Proton nuclear magnetic resonance (‘H NMR) and carbon nuclear magnetic resonance (13C NMR) analyses were carried out at room temperature in deuterated chloroform (CDCI3) on a Varian Gemini-300 and Varian Vax-SOO spectrometers with the solvent proton signals being used as chemical shift standards. Mass spectral analyses were carried out on a VG Trio-1 Benchtop GC-MS. The molecular weights of polymers were determined by gel permeation chromatography (GPC) using a PLgel 20m Mixed A column and a Waters R401 Differential Refractometer detector at room temperature. THF was used as the eluting solvent at a flow rate 1 mL/min, and monodisperse polystyrene standards were used to calibrate the molecular weights. The concentration of the polymer solutions used for GPC measurements was 1 mg/mL. Differential scanning calorimetry (DSC) analyses of the polymers were obtained using a Perkin Elmer DSC 7. Samples were run under a helium atmosphere at a heating rate of 10 °Clmin, with the temperature calibrated with an indium standard. Thermogravimetric analyses (T GA) were run in both air and under nitrogen at a heating rate of 10 °Clmin using a Perkin Elmer TGA 7. Measurements of the 212 mechanical properties of polymer samples were made using a Perkin Elmer DMA 7 Dynamic Mechanical Analyzer at a heating rate of 10 °Clmin. X-ray Powder Diffraction (XRD) patterns were recorded on a Rigaku rotaflex 2008 diffractometer equipped with a rotating anode, Cu Kgx-ray radiation (k: 1.541838A) and a curved crystal graphite monochromator. The x-ray was operated at 45 KV and 100 mA. Diffraction patterns were collected at 001° intervals between 5 and 45° values of 20 at a scanning rate of 01° per minute and DS and SS slit widths of 1/2. Powder samples were prepared by spreading solid samples on the window of the glass sample holder with a spatula, or by taping polymer film made from solvent casting on to the window of the sample holder. 2. Synthesis Symmetric Monomer Synthesis, general procedure. (Route 1 in Scheme 3) A mixture of 10 g of the appropriate a-hydroxy acid and 0.2 g of p- toluenesulfonic acid in 700 mL of toluene was heated at reflux for four days, with the water removed azeotropically using a Barrett trap. The toluene Solution was then cooled, washed with sat. NaHCOa, and dried over M9804. After removing the toluene, about 0.1 g of ZnO was added and the residue was distilled under reduced pressure using a Kugelrohr distillation apparatus.19 3,6-Diethyl-1,4-dioxane-2,5-dione (Ethylglycolide) (1). Ethylglycolide was collected at 180 °C (100 mtorr), and was dissolved in the minimum amount 213 of ether needed to dissolve the product. The solution was cooled to —30 °C, and petroleum ether was added drop-wise until the solution turned cloudy. The colorless crystals were collected by cold filtration and dried under vacuum to give 5.2 g (63%) of ethylglycolide as a colorless oil. 1H NMR indicates that the product is a statistical mixture of diastereomers. 1H NMR (300 MHz, CDCI3): 6 4.88 (dd), 4.83 (dd, 1H total for the signals at 4.88 and 4.83), 2.08 (m, 2H), 1.15 (tt, 3H). 130 NMR (75 MHz, cocna) 5: 167.62, 166.58, 76.4, 75.6. mp 19.5 - 20.5 °c Iit.179 21-22 °c; MS (EI) m/z = 173.4 (M+1). Preparation via or-bromoacyl bromides.180 Under a nitrogen atmosphere, 0.90 g (8.7 mmol) of 2- hydroxybutyric acid and 2.1 g (8.7 mmol) of 2-bromobutyryl bromide were heated at 80 °C until HBr evolution ceased (0.5-2 h). The solution was cooled and 100 mL of dry acetone were added followed by the drop-wise addition of 2.2 mL (17 mmol) of triethylamine. The solution was heated to reflux for 3 h, and then cooled to room temperature. After removing the acetone under reduced pressure, the residue was washed with sat. NaHCOa, extracted with ether, and the ether layer was dried over M9804. After removing the ether, the dimer was purified by recrystallization as described above to give 0.62 g (41%) of ethylglycolide as a colorless oil. The product is approximately a 4:1 mixture of the R,R/8,8 and R,S diastereomers. 3,6-Diisobutyl-1,4-dioxane-2,5-dione (isobutylgycolide) (2). Isobutylglycolide was collected at 120 °C (50 mtorr), and was recrystallized from ether. The white crystals were collected by filtration and dried under vacuum to give 6.1g (71%) of isobutylglycolide as white solid. 1H NMR indicates that the 214 product is a statistical mixture of the R,R/8,8 and R,S diastereomers. 1H NMR (300 MHz, CDCI3): 5 4.91 (dd), 4.89 (dd), (1H total for the signals at 4.91 and 4.89), 1.8-2.0 (m, 3H), 0.95 (m, 6H). 13C NMR (75 MHz, CDCI3) 8: 167.28, 166.20, 74.88, 74.09, 40.37, 38.82, 24.07, 23.84, 23.02, 22.84, 21.29, 21.25. MS (mlq) 229.4 (M+1); mp 167-167.5 °c ("L198 169-170 °C). 3,6-Dihexyl-1,4-dioxane-2,5-dione (hexylgycolide) (3). Hexylglycolide was collected at 145 °C (50 mtorr), and was recrystallized from ether. The white crystals were collected by filtration and dried under vacuum to give 5.8 g (65%) of hexylglycolide as a white solid. 1H NMR indicates that the product is a statistical mixture of the R,R/S,S and R,S diastereomers. 1H NMR (300 MHz, CDCI3): 8 4.88 (dd), 4.83 (dd), (1 H total for the signals at 4.88 and 4.83), 2.01 (m, 2H), 1.4-1.6 (br m, 2H), 1.2-1.4 (br m, 6H), 0.85 (t, 3H). 13C NMR (75 MHz, CDCI3) 6: 166.95, 165.85, 76.42, 75.65, 31.99, 31.46, 31.41, 30.15, 28.74, 28.56: MS (m/z) 285.4 (M+1), mp 78-80 °C . 3,6-DiisopropyI-1,4-dioxane-2,5-dione (isopropylgycolide) (4). lsopropylglycolide was synthesized from DL-2-hydroxy-3-methyl butyric The crude product was recrystallized directly from toulene after removing of toluene. The white crystal was collected by filtration and dried under vacuum. Yield 47%. 1H NMR (300 MHz, CDCI3): 6 4.95 (d, 1H), 2.30 (m, 1H), 1.03 (d, 3H), 1.01 (d, 3H). 13C NMR (75 MHz, CDCI3) 5 166.38, 79.55, 29.36, 19.55, 15.81. MS (m/z) 201.4 (M+1), mp 137-138 °c [iit.199 136 °C]. 215 3S,6$-DiisopropyI-1,4-dioxane-2,5-dione (S-isopropylgycolide) (5). S- isopropylgycolide was synthesized from L-2-hydroxy-3-methyl butyric The S- isopropylgycolide was recrystallized from toluene. The white crystals were collected by filtration and dried under vacuum. Yield 43%, [alzo = -264.0 (c=1, THF). 1H NMR (300 MHz, CDCI3): 84.95 (d, 1H), 2.30 (m, 1H), 1.03 (d, 3H), 1.01 (d, 3H). 13C NMR (75 MHz, CDCI3) 8: 166.38, 79.55, 29.36, 19.55, 15.81. MS (m/z) 201.4 (M+1), mp 140-141 °C. The S-isopropylglycolide is 99% ee was determined by hydrolyzing S-isopropylglycolide to 2-hydroxy-3-methyl butyric acid and comparing the optical rotation value the acid with known value ([a]20=+19.0 c=1 CHCI3) from Aldrich catalog. 2-hydroxy-3-methylbutyric acid. (6) A solution of 60 mL concentrated H2SO4 in 1L r water was cooled in an ice-bath. To the cooled solution was added 60 g of valine follwed by the dropwise of 144 g NaNOz dissolved in 1L water. After the solution was stirred at 0 °C for overnight, the solution was extracted with ether (6 x 300 mL), and the ether layer was dried over M9304. After removing the ether, the product was purified by recrystallized from toluene. The white crystals were collected by filtration and dried under vacuum to give 45 g (75%) of 2-hydroxy-3-methyl butyric acid as a white solid. DL-2-hydroxy-3-methyl butyric acid was obtained from DL-valine and L-2-hydroxy-3-methyl butyric acid was prepared form L-valine. The L-2-hydroxy—3-methyl butyric acid was obtained in 99% ee based on measurement of its optical rotation ([or]20=+18.9 c=1 CHC'a). 216 Unsymmetric Monomer synthesis, general procedure. Under a nitrogen atmosphere, one equivalent of a 2-hydroxy acid and one equivalent of a 2-bromoacyl bromide were heated at 80 °C until HBr evolution ceased (0.5-2 h.). The solution was cooled and 200 mL of dry acetone was added for each gram of acid, followed by the dropwise addition of two equivalent of triethylamine. The solution was heated to reflux for 6 h., and then was cooled. After removing the acetone, ethyl acetate was added to dissolve the residue. The solution was washed with 2N HCI, water, then washed with sat. NaHCOa, and the organic layer was dried over MgSO4. After removing solvent, the monomers was purified by recrystallization and distillation. 3-ethyl-6-methyI-2,5-dloxane-1,4-dione (7) was synthesized from 2- hydroxybutyric acid and 2-bromopropionyl bromide. It was purified by distillation (45 °C/50 mtorr) to give a colorless oil. Yield: 53%. 1H NMR (CDCla) 8 5.03 (q), 4.99 (q, 1H for signal at 5.03 and 4.99 ppm), 4.86 (dd), 4.83 (dd, 1H for signal at 4.86 and 4.83), 1.90-2.20 (m, 2H), 1.69 (d), 1.66 (d, 3H for signals at 1.69 and 1.66 ppm). 1.15 (t). 1.14 (t, 3H for signals at 1.15 and 1.14ppm). 13C NMR (CDCI3) 8 167.6, 166.8, 166.3, 165.6, 77.7, 76.6, 72.5, 72.2, 25.3, 23.3, 17.5, 15.7, 9.1, 8.7. MS (m/q) 159.4 (M+1), The product is approximately a 3:1 mixture of the R,R/S,S and R,S diastereomers. 3-dimethyl-6-methyI-2,5-dioxiane-1,4-dione (8) was synthesized from 2- hydroxyisobutyric acid and 2-bromopropionyl bromide. Recrystallization from ether give white crystal. mp 66-69°C. Yield: 43%. 1H NMR (CDCI3) 8 5.07 (q, 1H), 217 1.68 (s, 6H). 1.65 (d, 3H). 130 NMR (00013) 8 168.59, 166.64, 6055,7292, 26.25, 25.31, 17.45 MS (m/q) 159.4 (M+1) 3-methyI-6-phenyl-2,5-dioxane-1,4-dione (9) was synthesized by using mandelic acid and 2-bromopropyionyl bromide. Recrystallization from toluene gave white crystals. mp 153-156 °C. Yield: 33%. 1H NMR (CDCI3) 8 7.43 (m, 5H), 5.92 (s, 1H) 5.17 (q, 1H), 1.63 (d, 3H). 130 NMR (00013) 5 166.90, 165.55, 131.24, 129.95, 128.92, 127.44, 77.74, 72.77, 16.42. MS (m/q) 207.4 (M+1), S-3-methyl-S-6-phenyl-2,5-dloxiane-1,4-dione (10) was synthesized from S-mandelic acid and 2-bromopropionyl bromide. Recrystallization from toluene gave white crystals. mp 160-163 °C, yield: 41%. [0020 = +301.0 1H NMR (CDCI3) 8 7.43 (m, 5H), 5.92 (s, 1H) 5.17 (q, 1H), 1.63 (d, 3H). 13C NMR (CDCla) 8166.90, 165.55, 131.24, 129.95, 128.92, 127.44, 77.74, 72.77, 16.42. MS (m/q) 207.4 (M+1). R-3-methyl-R-6-phenyl-2,5-dioxiane-1,4-dione (11) was synthesized from R-mandelic acid and 2-bromopropionyl bromide. Recrystallization from toluene gave white crystals. mp 160-163 °C, yield: 35%. [alzo = -302.0. 1H NMR (CDCla) 8 7.43 (m, 5H), 5.92 (s, 1H) 5.17 (q, 1H), 1.63 (d, 3H). 13C NMR (CDCI3) 8166.90, 165.55, 131.24, 129.95, 128.92, 127.44, 77.74, 72.77, 16.42. MS (m/q) 207.4 (M+1). 218 3. Bulk polymerization of substituted glycolides. Solvent-free polymerizations were carried out in sealed tubes prepared from 3/8 inch diameter glass tubing. A representative polymerization is described below. In an oxygen and moisture-free dry box, a solution of initiator in toluene (z 0.01 M) and 0.19 of monomer were added to the tube. The amount of initiator solution added was determined by the desired monomer/initiator ratio. For runs using initiators that are insoluble in toluene, the initiator was added directly to the tube and the walls of the tube were washed with solvent to ensure that all of the initiator was added to the monomer. The solvent was removed in vacuum, and the tube was sealed and immersed in an oil bath at 130 °C. At the end of the polymerization, the tube was cooled, opened, and the polymer was dissolved in THF. A portion of the sample was evacuated to dryness and analyzed by NMR for conversion. After removal of the solvent, the polymer was dissolved in toluene and precipitated into methanol to remove residual initiator. Typical yields of poly(ethylglycolide) were >85%. For kinetic runs, multiple tubes were prepared and individual tubes were removed from the heating bath at predetermined intervals and were cooled in ice, opened, and the contents dissolved in THF. A portion of the sample was analyzed by GPC for molecular weight, and the remainder was evacuated to dryness and analyzed by NMR for conversion. Polymerizations that used alcohols as co-initiators were set up as described above, except that the appropriate amount of alcohol was added to the toluene solution of initiator just prior to adding the initiator solution to the tube. 219 For insoluble initiators, tubes were first loaded with monomer and initiator, and the alcohol co-initiator was directly added to the tube as a toluene solution. 4. Solution polymerization of substituted glycolides. The reaction flask was charged with 2 mmol of monomer and dried under vacuum (diffusion pump) at room temperature overnight. Toluene (10 mL ) was added to the solvent flask through a rubber septum with a syringe, and the toluene was purified by initiating an anionic polymerization of styrene. The solvent was transferred under vacuum to the reaction flask, and the initiator Al(OiPr)3 or Sn(Oct); was added into reaction flask with a syringe through a rubber septum. The amount of initiator solution added was determined by the desired monomer/initiator ratio The polymerization was carried out at 70 °C, 90 °C and 100 °C. After the polymerization finished, the reaction was terminated with 1 mL 2NHC| solution then washed with distilled water until PH=7. The polymer was precipitated into cold methanol, filtered, and dried under vacuum. For kinetic studies, small samples were removed at predetermined times using a syringe through the rubber septum. The samples were analyzed by NMR for conversion and GPC for molecular weight. 220 5. The surface energy of polymers Polymer films. The polymers were dissolved in small amount of toluene, then the solutions were spread on the clean glass slides. The glass slides were dried in air for 3 days, then dried under vacuum for one day. Surface tension of solutions. A series of HzolCHaoH solutions were made. The densities of solutions were measured using hydrometer. The surface tensions of solutions were measured using pendant drop method200 by FTA 200 contact angle analyzer, which was calibrated using de-ionized water. Surface tension of polymers. The H20ICH30H solutions were dropped on the polymer films. The contact angles between the solutions and polymer films were measured using FTA 200 contact angle analyzer, which was calibrated using de-ionized water. From change of contact angles with change of surface tensions of solutions, the surface energies of polymers can be calculated by Zisman equation: 6089 = 1 1' BWLJYc) where 9 is contact angle, yL is surface tension of testing solution (water/methanol solution). yo is the surface energy of the polymer surface, and [3 is constant. 221 6. The degradation of polymers. About 50 mg of polymer sample was accurately weighed and chopped to 1mm by 1mm pieces. The chopped sample was immersed in to 15 mL 0.1M phosphate buffer solution (pH=7.4). The degradation was carried out at 55 i 0.1 °C. After a predetermined period of time, the sample was taken out, washed thoroughly with distill water and dried under vacuum. The sample was weighed to determine the weight loss and molecular weight loss was determined by GPC. 222 BIBLIOGRAPHY 223 10. 11. 12. 13. 14. 15. BIBLIOGRAPHY Zhang, X. C.; Wyss, U. P.; Pichora, D.; Goosen, M. F. A.; J. Macromol. Sci.- Pure Appl. Chem. 1993, A30, 933-947. Han, D. K.; Hubbell, J. A.; Macromolecules 1997, 30, 6077-6083. Kricheldorf, H. R.; KreiserSaunders, l.; Macromol. Symp. 1996, 103, 85- It: 102. ' Chen, X. H.; McCarthy, S. P.; Gross, R. A.; Macromolecules 1997, 30, 4295-4301 . * Nobes, G. A. R.; Marchessault, R. H.; Maysinger, D.; Dmg Deliv. 1998, 5, 167-177. Fukuzaki, H.; Yoshida, M.; Asano, M.; Kumakura, M.; Biomaten'als 1990, 11 , 441446. Breitenbach, A.; Kissel, T.; Polymer 1998, 39, 3261-3271. McCoy, M.; C&EN 1998, 76, 71-72. Sinclair, R. G.; J. M. S. - Pure Appl. Chem. 1996, A33, 585-597. Meinander, K.; Niemi, M.; Hakola, J. S.; Selin, J. F.; Macromol. Symp. 1997, 123, 147-153. Grijpma, D. W.; Nijenhuis, A. J.; Vanwijk, P. G. T.; Pennings, A. J.; Polym. Bull. 1992, 29, 571 -578. Spinu, M.; Jackson, C.; Keating, M. Y.; Gardner, K. H.; J. Macromol. Sci.- Pure Appl. Chem. 1996, A33, 1497-1530. Vert, M.; Schwarch, G.; Coudane, J.; J. Macromol. Sci.-Pure Appl. Chem. 1995, A32, 787-796. Bastioli, C.; Macromol. Symp. 1998, 135, 193-204. Fukuzaki, H.; Yoshida, M.; Asano, M.; Kumakura, M.; Polymer 1990, 31, 432. 224 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. Buchholz, B.,German patent 443,542,A2 (1991) Otera, J.; Kawada, K.; Yano, T.; Chem. Lett. 1996, 225-226. Yamaguchi, A.; Suzuki, K.; Enomoto, K.; Ajioka, M.; Bull. Chem. Soc. Jpn. 1995, 68, 2125-2131. Deane, D. D.; Hammond, E. G.; J. Dairy Sci. 1960, 43, 1421. Guber, P. R.; Benson, R. D.; lwen, M. L.,U. S. patent 5,141,023 (1992) Johns, D. B.; Lenz, R. W.; Luecke, A. Ring-Opening Polymerization; lvin, K. J. and Saegusa, T., Ed.; Elsevier Applied Science Publishers, 1984; Vol. 1, pp 461. Odian, G. Principle of Polymerization; Third ed.; John Wiley & Son:, 1991. Hall, H. K., Jr.; Brandt, M. K.; Mason, R. W.; J. Am. Chem. Soc. 1958, 80, 6420. Carothers, W. H.; Borough, G. L.; Van Natta, R. W.; J. Am. Chem. Soc 1932, 54, 761. Lundberg, R. 0.; Cox, E. F. Ring-Opening Polymerization; Marcel Dekker: New York, 1969. Kricheldorf, H. R.; Michael Jonte, J.; Dunsing, R.; J. Macromol. Soc. 1986, A23(4), 495. Rozenberg, D.; Makromol. Chem, Macromol. Symp. 1990, 32, 267. Crivello, J. V.; Lockhart, T., P.; J. polym. Sci, Polym. Chem. Ed. 1983, 21 , 97. Cherdron, H.; Korte, F.; Makromol. Chem. 1962, 56, 179. Penczek, S.; Hofman, A.; Szymanski, R.; Slomkowski, S.; Makromol. Chem. 1984, 185, 655-667. Hofman, A.; Stomkowski, S.; Penczek, S.; Makromol. Chem. 1987, 188, 2027-2040. 225 32. 33. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. Kricheldorf, H. R.; Michael Jonte, J.; Dunsing, R.; Makromol. Chem. 1986, 187, 771-785. Okamoto, Y.; Makromol. Chem, Macromol. Symp. 1991, 42/43, 117. Kohn, F. E.; Van Ommen, J. G.; Feuen, J.; European Polymer Journal 1983, 19, 1081-1088. Kricheldorf, H. R.; Dunsing, R.; Makromol. Chem. 1986, 187, 1611-1625. Jedlinski, Z.; Kurcok, P.; Kowalczuk, M.; Matuszowicz, A.; Dubois, P.; Jerome, R.; Kricheldorf, H. R.; Macromolecules 1995, 28, 7276-7280. Penczek, S.; Duda, A.; Libiszowski, J.; Macromol. Symp. 1998, 128, 241- 254. Kurcok, P.; Kowalczuk, M.; Hennek, K.; Jedlinski, Z.; Macromolecules 1 992, 25, 201 7-2020. Yuan, M. L.; Xiong, C. D.; Deng, X. M.; J. Appl. Polym. Sci. 1998, 67, 1273- 1276. Kurcok, P.; Penczek, J.; Franek, J.; Jedlinski, Z.; Macromolecules 1992, 25, 2285-2289. Kricheldorf, H. R.; Scharnagl, N.; Jedlinski, 2.; Polymer 1996, 37, 1405- 141 1. Kricheldorf, H. R.; Kreiser-saunders, l.; Makromol. Chem-Macro. Chem. Phys. 1990, 191, 1057-1066. Jedlinski, Z.; Walach, W.; Kurcok, P.; Adamus, G.; Makromol. Chem.- Macro. Chem. Phys. 1991, 192, 2051-2057. Emig, N.; Nguyen, H.; Krautscheid, H.; Reau, R.; Cazaux, J. B.; Bertrand, G.; Organometallics 1998, 17, 3599-3608. Leborgne, A.; Vincens, V.; Jouglard, M.; Spassky, N.; Makromol. Chem, Macromol. Symp 1993, 73, 37-46. 226 46. 47. 48. 49. 50. 51. 52. 53. 55. 56. 57. 58. 59. 60. 61. 62. Kowalski, A.; Duda, A.; Penczek, S.; Macromol. Rapid Commun. 1998, 19, 567-572. Witzke, D. R.; Narayan, R.; Abstr. Pap. Am. Chem. Soc. 1998, 216, U11- U12. Schwach, G.; Coudane, J.; Engel, R.; Vert, M.; Polym. Int. 1998, 46, 177- 182. Kricheldorf, H. R.; Damrau, D. 0.; Macromol. Chem. Phys. 1997, 198, 1753-1766. Sodergard, A.; Stolt, M.; Macromol. Symp. 1998, 130, 393-402. Simic, V.; Girardon, V.; Spassky, N.; Hubert-Pfalzgraf, L. G.; Duda, A.; Polym. Degrad. Stabil. 1998, 59, 227-229. McLain, S. J.; Ford, T. M.; Drysdale, N. E.; Abstr. Pap. Am. Chem. Soc. 1992, 204, 188-POLY. Bero, M.; Kasperczyk, J.; Adamus, G.; Makromol. Chem, Macromol. Symp 1993, 194, 907-912. Degee, P.; Dubois, P.; Jerome, R.; Macromol. Symp. 1997, 123, 67-84. Dubois, P.; Jacobs, C.; Jerome, R.; Teyssie, P.; Macromolecules 1991, 24, 2266-2270. Montaudo, G.; Montaudo, M. S.; Puglisi, C.; Samperi, F.; Spassky, N.; Le Borgne, A.; Wisniewski, M.; Macromolecules 1996, 29, 6461 -6465. Duda, A.; Macromolecules 1996, 29, 1399-1406. Kowalski, A.; Duda, A.; Penczek, S.; Macromolecules 1998, 31, 2114-2122. Ropson, N.; Dubois, P.; Jerome, R.; Teyssie, P.; Macromolecules 1993, 26, 6378-6385. Duda, A.; Penczek, S.; Macromolecules 1995, 28, 5981-5992. Duda, A.; Penczek, S.; Macromol. Rapid Commun. 1995, 16, 67-76. Dubios, P.; Jerome, R.; Teyssie, P.; Polymer Perprint 1994, 35, 536-537. 227 67. E 68. ' 69. 70. 71 72 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. Ropson, N.; Dubois, P.; Jerome, R.; Teyssie, P.; Macromolecules 1995, 28, 7589-7598. Kricheldorf, H. R.; Boettcher, C.; Tonnes, K. U.; Polymer 1992, 33, 2817- 2824. Degee, P.; Dubois, P.; Jerome, R.; Teyssie, P.; Macromolecules 1992, 25, 4242-4248. Dubois, P.; Jerome, R.; Teyssie, P.; Abstr. Pap. Am. Chem. Soc. 1990, 199, 16-POLY. Barakat, l.; Dubois, P.; Jerome, R.; Teyssie, P.; J. Polym. Sci. Pol. Chem. 1993, 31, 505-514. Tian, D.; Dubois, P.; Jerome, R.; Teyssie, P.; Macromolecules 1994, 27, 4134-4144. Barakat, |.; Dubois, P.; Jerome, R.; Teyssie, P.; Goethals, E.; J. Polym. Sci. Pol. Chem. 1994, 32, 2099-3110. Trofimoff, L.; Aida, T.; lnoue, 8.; Chem. Lett. 1987, 991-994. Endo, M; Aida, T.; lnoue, S.; Macromolecules 1987, 20, 2982-2988. lsoda, M.; Sugimoto, H.; Aida, T.; lnoue, S.; Macromolecules 1997, 30, 57- 62. Sugimoto, H.; Aida, T.; lnoue, S.; Macromolecules 1990, 23, 2869-2875. Sugimoto, H.; Kawamura, C.; Kuroki, M.; Aida, T.; lnoue, S.; Macromolecules 1994, 27, 2013-2028. sugimoto, H.; Aida, T.; lnoue, S.; Macromolecules 1993, 26, 4751-4755. Mogstad, A. L.; Waymouth, R. M.; Macromolecules 1994, 27, 2313-2315. Kricheldorf, H. R.; Boettcher, C.; Makromol. Chem-Macro. Chem. Phys. 1993, 194, 463-473. Kricheldorf, H. R.; Meierhaack, J.; Makmmol. Chem-Macro. Chem. Phys. 1993, 194, 715-725. 228 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. Sandner, B.; Steurich, 8.; Gopp, U.; Polymer 1997, 38, 2515-2522. Kricheldorf, H. R.; Sumbel, M.; Eur. Polym. J. 1989, 25, 585-591. Kricheldorf, H. R.; Mahler, A.; Polymer 1996, 37, 4383-4388. Kricheldorf, H. R.; WeegenSchulz, B.; Polymer 1995, 36, 4997-5003. Kricheldorf, H. R.; Lee, S. R.; Scharnagl, N.; Macromolecules 1994, 27, 3139-3146. Kricheldorf, H. R.; Sumbel, M. V.; Kreisersaunders, l.; Macromolecules 1991, 24, 1944-1949. Kricheldorf, H. R.; Eggerstedt, S.; Macromolecules 1997, 30, 5693-5697. Kricheldorf, H. R.; Lee, S. R.; Macromolecules 1995, 28, 6718-6725. Witzke, D. R.; Narayan, R.; Kolstad, J. J.; Macromolecules 1997, 30, 7075- 7085. Leenslag, J. W.; Pennings, A. J.; Makromol. Chem-Macro. Chem. Phys. 1987, 188, 1809-1814. Schwach, G.; Coudane, J.; Engel, R.; Vert, M.; J. Polym. Sci. Pol. Chem. 1997, 35, 3431-3440. Kricheldorf, H. R.; Boettcher, C.; J. Macromol. Sci-Pure Appl. Chem. 1993, A30, 441-448. Kricheldorf, H. R.; Damrau, D. 0.; J. Macromol. Sci-Pure Appl. Chem. 1998, A35, 1875-1887. Nijenhuis, A. J.; Grijpma, D. W.; Pennings, A. J.; Macromolecules 1992, 25, 6419-6424. Du, Y. J.; Lemstra, P. J.; Nijenhuis, A. J.; Vanaert, H. A. M.; Bastiaansen, C.; Macromolecules 1995, 28, 2124-21 32. Kricheldorf, H. R.; Kreisersaunders, l.; Boettcher, C.; Polymer 1995, 36, 1253-1259. Kowalski, A.; Duda, A.; Penczek, S.; Macromolecules 2000, 33, 689-695. 229 96. Yasuda, H.; lhara, E.; In Advances in Polymer Science; Advances in Polymer Science 133, 1997. 97. Yasuda, H.; lhara, E.; Bull. Chem. Soc. Jpn. 1997, 70, 1745-1767. 98. Yamashita, M.; Takemoto, Y.; lhara, E.; Yasuda, H.; Macromolecules 1996, 29, 1798-1806. 99. Stevels, W. M.; Ankone, M. J. K.; Dijkstra, P. J.; Feijen, J.; Macromolecules 1996, 29, 8296-9303. 100. Stevels, W. M.; Ankone, M. J. K.; Dijkstra, P. L.; Feijen, J.; Macromol. Chem. Phys. 1995, 196, 1153-1161. 101. Huang, O. H.; Shen, Z. 0.; Zhang, Y. F.; Shen, Y. Q.; Shen, L. F.; Yuan, H. 2.; Polym. J. 1998, 30, 168-170. 102. Shen, Y. Q.; Zhu, K. J.; Shen, Z. 0.; Yao, K. M.; J. Polym. Sci. Pol. Chem. 1996, 34, 1799-1805. 103. Shen, Y. Q.; Shen, Z. 0.; Zhang, F. Y.; Zhang, Y. F.; Polym. J. 1995, 27, 59-64. 104. Shen, Y. Q.; Shen, Z. 0.; Zhang, Y. F.; Yao, K. M.; Macromolecules 1996, 29, 8289-8295. 105. Shen, Y. Q.; Shen, Z. 0.; Shen, J. L.; Zhang, Y., F.; Yao, K. M.; Macromolecules 1996, 29, 3441 -3446. 106. Shen, Y. Q.; Sheri, Z. 0.; Zhang, Y. F.; Hang, G. H.; J. Polym. Sci. Pol. Chem. 1997. 35, 1339-1352. 107. Ohse, H.; Cherdron, H.; Korte, F.; Makromol. Chem. 1965, 86, 312. 108. Saotome, K.; Kodaira, Y.; Makromol. Chem. 1965, 82,41. 109. Wilson, D. R.; Beaman, R. G.; J. Polym. Sci, Part A-1 1970, 8, 2161. 110. Perego, G.; Cella, G. D.; Bastioli, C.; J. Appl. Polym. Sci. 1996, 59, 37-43. 111. Macdonald, R. T.; McCarthy, S. P.; Gross, R. A.; Macromolecules 1996, 29, 7356-7361. 230 112. :113. 114. 115. '116. 117. °118. " 119. .120. 121. 122. 123. 124. 125. 126. 127. 128. Grijpma, D. W.; Pennings, A. J.; Macromol. Chem. Phys. 1994, 195, 1633- 1647. Tsuji, H.; Horii, F.; Hyon, S. H.; lkada, Y.; Macromolecules 1991, 24, 2719- 2724. lkada, Y.; Tsuji, H.; Macromolecules 1993, 26, 6918-6926. lkada, Y.; Hyon, S. H.; Tsuji, H.; Macromolecules 1991, 24, 5651-5656. Tsuji, H.; lkada, Y.; J. Appl. Polym. Sci. 1994, 53, 1061-1071. Tsuji, H.; lkada, Y.; Hyon, S. H.; Kimura, Y.; Kitao, T.; J. Appl. Polym. Sci. 1994, 51, 337-344. Tsuji, H.; lkada, Y.; Macromolecules 1993, 26, 6918-6926. Tsuji, H.; lkada, Y.; Macromolecules 1992, 25, 5719-5723. Tsuji, H.; Hyon, S. H.; lkada, Y.; Macromolecules 1992, 25, 2940-2946. Tsuji, H.; Hyon, S. H.; lkada, Y.; Macromolecules 1991, 24, 5651-5656. Tsuji, H.; Hyon, S. H.; lkada, Y.; Macromolecules 1991, 24, 5657-5662. HiljanenVainio, M. P.; Drava, P. A.; Seppala, J. V.; J. Biomed. Mater. Res. 1997, 34, 39-46. Lostocco, M. R.; Huang, S. J.; Abstr. Pap. Am. Chem. Soc. 1996, 212, 245- PMSE. Stevels, W. M.; Ankone, M. J. K.; Dijkstra, P. J.; Feijen, J.; Macromol. Symp. 1996, 102, 107-113. Veld, P.; Velner, E. M.; VanDeWitte, P.; Hamhuis, J.; Dijkstra, P. J.; Feijen, J.; J. Polym. Sci. Pol. Chem. 1997, 35, 219-226. Duda, A.; Biela, T.; Libiszowski, J.; Penczek, S.; Dubois, P.; Mecerreyes, 0.; Jerome, R.; Polym. Degrad. Stabil. 1998, 59, 215-222. Gallardo, A.; Roman, J.; Dijkstra, P. J.; Feijen, J.; Macromolecules 1998, 31, 7187-7194. 231 129. 130. 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144. 145. King, E.; Cameron, R. E.; Polym. Int. 1998, 45, 313-320. Storey, R. F.; Herring, K. R.; Hoffman, D. C.; J. Polym. Sci. Pol. Chem. 1991, 29, 1759-1777. Nakayama, A.; Kawasaki, N.; Maeda, Y.; Arvanitoyannis, |.; Aiba, S.; Yamamoto, N.; J. Appl. Polym. Sci. 1997, 66, 741 -748. Cal, J.; Zhu, K. J.; Yang, S. L.; Polymer 1998, 39, 4409-4415. Ruckenstein, E.; Yuan, Y. M.; J. Appl. Polym. Sci. 1998, 69, 1429-1434. Kim, Y. J.; Adamson, D. H.; Abstr. Pap. Am. Chem. Soc. 1998, 216, U24- U24. Lee, D. S.; Choi, S. W.; Jeon, B. M.; Kim, S. W.; Abstr. Pap. Am. Chem. Soc. 1998, 216, U878-U878. Li, S. M.; Anjard, S.; Rashkov, l.; Vert, M.; Polymer 1998, 39, 5421-5430. Song, C. X.; Feng, X. D.; Macromolecules 1984, 17, 2764-2767. Choi, Y. R.; Bae, Y. H.; Kim, S. W.; Macromolecules 1998, 31, 8766-8774. Grijpma, D. W.; Joziasse, C. A. P.; Pennings, A. J.; Macromol. Rapid Commun. 1993, 14, 155-161. Joziasse, C. A. P.; Veenstra, H.; Topp, M. D. C.; Grijpma, D. W.; Pennings, A. J.; Polymer 1998, 39, 467-474. Li, Y. X.; Kissel, T.; Polymer 1998, 39, 4421-4427. Eguiburu, J. L.; Berridi, M. J. F.; Sanroman, J.; Polymer 1995, 36, 173-179. Jerome, R.; Mecerreyes, D.; Tian, D.; Dubois, P.; Hawker, C. J.; Trollsas, M.; Hedrick, J. L.; Macromol. Symp. 1998, 132, 385-403. Jacobs, 0.; Dubois, P.; Jerome, R.; Teyssie, P.; Macromolecules 1991, 24, 3027-3034. Vanhoorne, P.; Dubois, P.; Jerome, R.; Teyssie, P.; Macromolecules 1992, 25, 37-44. 232 146. 147. 148. 149. 150. 151. 152. 153. 154. 155. 156. 157. 158. 159. Reeve, M. S.; McCarthy, S. P.; Gross, R. A.; Macromolecules 1993, 26, 888-894. Xiong, C. D.; Cheng, L. M.; Xu, R. P.; Deng, X. M.; J. Appl. Polym. Sci. 1995, 55, 865-869. Tortosa, K.; Miola, C.; Hamaide, T.; J. Appl. Polym. Sci. 1997, 65, 2357- 2372. Wang, H.; Dong, J. H.; Qiu, K. Y.; J. Polym. Sci. Pol. Chem. 1998, 36, 695- 702. Huang, S. J.; Onyari, J. M.; Nicolais, L.; DelNobile, S.; Mensitieri, G.; Ambrosio, L.; Abstr. Pap. Am. Chem. Soc. 1997, 213, 126-CELL. Ohya, Y.; Maruhashi, S.; Ouchi, T.; Macromolecules 1998, 31 , 4662-4665. Ohya, Y.; Maruhashi, S.; Ouchi, T.; Macromol. Chem. Phys. 1998, 199, 2017-2022. Nijenhuis, A. J.; Grijpma, D. W.; Pennings, A. J.; Polymer 1996, 37, 2783- 2791. Aoyagi, T.; Miyata, F.; Nagase, Y.; J. Control. Release 1994, 32, 87-96. Grijpma, D. W.; Kroeze, E.; Nijenhuis, A. J.; Pennings, A. J.; Polymer 1993, 34, 1496-1503. Sandner, B.; Steurich, S.; Wartewig, S.; Macromol. Symp. 1996, 103, 149- 162. Storey, R. F.; Warren, S. 0.; Allison, C. J.; Puckett, A. D.; Polymer 1997, 38, 6295-6301. Grijpma, D. W.; Pennings, A. J.; Macromol. Chem. Phys. 1994, 195, 1649- 1663. Cook, A. D.; Pajvani, U. B.; Hrkach, J. S.; Cannizzaro, S. M.; Langer, R.; Biomaterials 1997, 18, 1417-1424. 233 160. 161. 162. 163. 164. 165. 166. 167. 168. 169. 170. 171. 172. 173. 174. 175. 176. Eguiburu, J. L.; lruin, J. J.; Fernandez-Berridi, M. J.; Roman, J. S.; Polymer 1998, 39, 6891-6897. Lostocco, M. R.; Borzacchiello, A.; Huang, S. J.; Macromol. Symp. 1998, 130, 151-160. Yang, J. M.; Chen, H. L.; You, J. W.; Hwang, J. C.; Polym. J. 1997, 29, 657-662. Tsuji, H.; lkada, Y.; J. Appl. Polym. Sci. 1996, 60, 2367-2375. Blumm, E.; Owen, A. J.; Polymer 1995, 36, 4077-4081. Liu, X.; Dever, M.; Fair, N.; Benson, R. S.; J. Environ. Polym. Degrad. 1997, 5, 225-235. Wachsen, 0.; Platkowski, K.; Reichert, K. H.; Polym. Degrad. Stabil. 1997, 57, 87-94. Albertsson, A. C.; Sjoling, M.; J. Macromol. Sci-Pure Appl. Chem. 1992, 29, 43-54. Zhang, X. C.; Wyss, U. P.; Pichora, D.; Goosen, M. F. A.; Polym. Bull. 1992, 27, 623-629. Sodergard, A.; Nasman, J. H.; Polym. Degrad. Stabil. 1994, 46, 25-30. Albertsson, A. C.; Palmgren, R.; J. Macromol. Sci-Pure Appl. Chem. 1993, A30, 919-931. Wachsen, 0.; Reichert, K. H.; Kruger, R. P.; Much, H.; Schulz, G.; Polym. Degrad. Stabil. 1997, 55, 225-231. Kopinke, F. D.; Mackenzie, K.; J. Anal. Appl. Pyrolysis 1997, 40-1, 43-53. Li, S. M.; McCarthy, 8.; Biomaterials 1999, 20, 35-44. Hyon, S. H.; Sen-I Gakkaishi 1998, 54, 527-531. Li, S. M.; Garreau, H.; Vert, M.; Petrova, T.; Manolova, N.; Rashkov, |.; J. Appl. Polym. Sci. 1998, 68, 989-998. lwata, T.; Doi, Y.; Macromolecules 1998, 31 , 2461-2467. 234 177. Abe, H.; Doi, Y.; Hori, Y.; Hagiwara, T.; Polymer 1998, 39, 59-67. 178. Gajria, A. M.; Dave, V.; Gross, R. A.; McCarthy, S. P.; Polymer 1996, 37, 437-444. 179. Bischoff, C. A.; Walden, P.; Chem. Ber. 1893, 26, 262-264. 180. Schollkopf, V. U.; Hartwig, W.; Sprotte., U.; Angew. Chem 1979, 91. 181. Kricheldorf, H. R.; Weegenschulz, B.; J. Polym. Sci. Pol. Chem. 1995, 33, 2193-2201. 182. Albertsson, A. C.; Eklund, M.; J. Polym. Sci. Pol. Chem. 1994, 32, 265-279. 183. Deng, X. M.; Zhu, Z. X.; Xiong, C. D.; Zhang, L. L.; J. Appl. Polym. Sci. 1997, 64, 1295-1299. 184. Kricheldorf, H. R.; Lee, S. R.; Macromolecules 1996, 29, 8689-8695. 185. Kasperczyk, J.; Bero, M.; Makromol. Chem-Macro. Chem. Phys. 1993, 194, 913-925. 186. Hall, H. K.; Schneider, A. K.; J. Am. Chem. Soc. 1958, 80, 6409-6412. 187. Thakur, K. A. M.; Kean, R. T.; Hall, E. S.; Kolstad, J. J.; Munson, E. J.; Macromolecules 1998, 31, 1487-1494. 188. Kasperczyk, J. E.; Macromolecules 1995, 28, 3937-3939. 189. Liu, G.; Fang, Y. E.; Shi, T. Y.; Chem. J. Chin. Univ.-Chin. 1997, 18, 486- 488. 190. Huang, J.; Lisowski, M. S.; Runt, J.; Hall, E. S.; Kean, R. T.; Buehler, N.; Lin, J. S.; Macromolecules 1998, 31, 2593-2599. 191. Thakur, K. A. M.; Kean, R. T.; Zell, M. T.; Padden, B. E.; Munson, E. J.; Chem. Commun. 1998, 1913-1914. 192. Zell, M. T.; Padden, B. E.; Paterick, A. J.; Hillmyer, M. A.; Munson, E. J.; Thakur, K. A. M.; Kean, R. T.; Abstr. Pap. Am. Chem. Soc. 1998, 215, U310-U310. 235 193. Thakur, K. A. M.; Kean, R. T.; Hall, E. S.; Doscotch, M. A.; Munson, E. J.; Anal. Chem. 1997, 69, 4303-4309. 194. Thakur, K. A. M.; Kean, R. T.; Zupfer, J. M.; Buehler, N. U.; Doscotch, M. A.; Munson, E. J.; Macromolecules 1996, 29, 8844-8851. 195. Weidinger, A.; Hermans, P. H.; Makromol. Chem. 1961, 44-46, 24. 196. Weidinger, A.; Hermans, P. H.; Makromol. Chem. 1961, 50, 98. 197. Weidinger, A.; Hermans, P. H.; Makromol. Chem. 1962, 52, 169. 198. Takayama, Y.; Bull. Chem. Soc. Jpn. 1933, 8, 173-176. 199. Schmidt, E.; Sachtleben, R.; Justus Liebigs Ann. Chem. 1878, 193, 112. 200. Adamson A. 6., Physical Chemistry of Surfaces; Fifth ed.; John Wiley & Son:, 1990. 28-30. 236