<<

SYNTHESIS AND CHARACTERIZATION OF POLY(SILOXANE IMIDE) BLOCK COPOLYMERS AND END-FUNCTIONAL POLYIMIDES FOR INTERPHASE APPLICATIONS by

Andrea Demetrius Bowens

Dissertation submitted to the faculty of the Virginia Polytechnic Institute and State University in partial fulfillment of the requirements for the degree of

DOCTOR OF PHILOSOPHY In Chemistry

Approved by:

Judy S. Riffle, Chair

James E. McGrath Allan R. Shultz

Neal Castagnoli John J. Lesko

December, 1999 Blacksburg, Virginia Tech

Keywords: poly(siloxane imide), block copolymer, poly(ether imide), poly(amic acid) salt, siloxane surfaces, atomic force microscopy, x-ray photoelectron spectroscopy, water contact angle, surface segregation SYNTHESIS AND CHARACTERIZATION OF POLY(SILOXANE IMIDE) BLOCK COPOLYMERS AND END-FUNCTIONAL POLYIMIDES FOR INTERPHASE APPLICATIONS

by

Andrea Demetrius Bowens

Judy S. Riffle, Chair Chemistry Department

(Abstract)

End-functional poly(ether amic acid)s and poly(siloxane imide) multiblock copolymers, comprised of 2,2’-Bis[4-(3,4-dicarboxyphenoxy)phenyl]propane dianhydride (BPADA) / meta-phenylene diamine (MPDA) and hexafluoroisopropylidene-2- bis(phthalic acid anhydride) (6FDA) / meta-phenylene diamine (MPDA) polyimide segments, have been prepared and characterized to explore possibilities for controlling interface properties. Incorporation of polydimethylsiloxane (PDMS) components into polyimide backbone structures can yield advantageous properties such as low energy surfaces and low stress interfaces. End-functional BPDA/MPDA poly(amic acid) salts and poly(siloxane amic acid) salts were prepared in methanolic or aqueous tripropylamine solutions. The polymeric salts formed stable water solutions (or dispersions) and imidized in less than 10 minutes at 260°C. The water solubility and rapid imidization times are ideal for on-line processing. Thus, these materials can be used as sizing and interface toughening agents for fiber reinforced composite manufacturing. Epoxy-polyimide networks prepared from the amine functionalized polyimide with DER 331 epoxy resin and diamino diphenylsulfone showed microphase separation (100-300 nm inclusions) by transmission electron microscopy. Slight toughening of the cured epoxy with 9 weight % imide was observed with the imide as the included phase. Epoxy bilayer films of polyimide (amine end-functional and commercial Ultem™) and poly(siloxane imide) multiblock copolymers were prepared to evaluate the -matrix interphase region. Atomic force microscopy (AFM) analysis of the bilayer films showed diffusion at the interphase for the bilayers prepared with the polyimides and the BPADA/MPDA block copolymers containing polyimide continuous phases. Poly(siloxane imide) multiblock copolymers comprised of 6FDA/MPDA polyimide structures are ideal candidates for controlling interfacial properties between substrates layered with thin films for microelectronic applications. These high Tg materials offer an approach for obtaining reduced moisture absorption and low stress interfaces. Evaluation of the refractive indices of the block copolymer films showed a decrease with increasing siloxane content thus suggesting the possibility of lower dielectric constants. The polymer-metal interfacial properties were investigated for films cast on titanium and tantalum substrates. The results suggested a correlation between the surface hydroxyl concentration of the metal oxide layer with the interfacial properties of the cast poly(siloxane imide) block copolymer films. The surface hydroxyls were thought to hydrogen bond with the PDMS component of the block copolymer. Since the titanium substrate has a higher surface hydroxyl concentration than the tantalum, higher silicon concentrations were observed. The melt imidized end-functional polyimides and poly(siloxane imide) block copolymers produced thermally stable materials with 5% weight loss temperatures well above 400°C. However, the block copolymers showed slightly lower 5% weight loss temperatures as a function of siloxane content with a significant increase in char formation. Correlation of the upper temperatures with the imide segment length was consistent with findings noted for other phase separated randomly segmented block copolymers. Incorporating PDMS into the polyimide backbone structure has an effect on the bulk and surface properties. The bulk properties of the poly(siloxane imide) block copolymers were characterized using TEM. The morphologies were consistent with classical block copolymers. Surface properties of the block copolymer films as a function of PDMS content were investigated using angular dependent X-ray photoelectron spectroscopy at take-off angles of 15, 30, and 45°. Surface enrichment of PDMS content over that of the bulk was observed at all three sampling depths. Further evidence of this siloxane enrichment in the surface was demonstrated with water contact angle analyses. With as little as 5 weight % PDMS ( = 5000 g/mol) in the block copolymer there was over a 25% increase in the water contact angle over the polyimide control. The surface topography was influenced by the degree of phase separation and was characterized using AFM. The roughness factor was used to represent the data. It was found that the surface roughness increased with increasing PDMS content. This dissertation is dedicated to my family for their unfaltering support and in memory of my grandmother, Frances Elizabeth Poole (1931-1991). Acknowledgements

I would like to extend my warmest appreciation to my advisor Dr. Judy S. Riffle for her encouragement and guidance during the research and writing of this dissertation. She also provided a great deal of help in my professional development. I am also appreciative of my other committee members Dr. James E. McGrath, Dr. Allan R. Shultz, Dr. Neil Castagnoli, and Dr. John J. Lesko. I also appreciate Dr. James P. Wightman for graciously substituting in the absence of a committee member during my final defense. He took part in many helpful discussions especially regarding the surface characterization. I would especially like to thank the excellent scientists who have assisted me in the characterization of many materials, Tom Glass for help with the Varian 400 MHz NMR, Dr. Q. Ji for size exclusion chromatography, Steve McCartney for transmission electron microscopy, Dr. Malcolm Robertson for atomic force microscopy, and Dr. T. Park for residual stress analyses. Appreciation is extended to Dr. W. Ducker and Dr. V. Venkatesan for their technical support on certain aspects of this research, and Dr. Kenneth Wynne and Steve Bullock of the Naval Research Laboratory for their discussions on Atomic Force Microscopy. There are two co-workers with whom I worked closely during portions of this research, Christy S. Tyberg and Maggie B. Bump. I would like to thank them for their contributions. A warm appreciation is extended to the Riffle research group for discussions and helping me maintain my sanity in the laboratory. Thanks are extended to the life long friends I acquired during my years in Blacksburg. I wish all of them success in their careers. Especially, I would like to thank my mother and sister, Lillian and LaCresia Bowens, and my fiancé and best friend Chester Jones for their love and inspiration. Thanks for believing in me when I did not believe in myself. This would not have been possible if it were not for the love of my savior Jesus Christ. I AM TRULY BLESSED.

I can do all things through Christ which strengtheneth me. -Philippians 4:13 Abbreviations

PDMS polydimethylsiloxane PSI poly(siloxane imide) 6FDA hexafluoroisopropylidene-2-bis(phthalic acid anhydride) BPADA 2,2’-Bis[4-(3,4-dicarboxyphenoxy)phenyl]propane dianhydride MPDA meta-phenylene diamine PMDA pyromellitic dianhydride DAPI diaminophenylindane ODA oxydianiline DCB dichlorobenzene FPSI fluorinated polyimide DDS 4-aminophenyl sulfone PS polystyrene PB polybutadiene NMP n-methylpyrolidone DMAC dimethylacetamide THF tetrahydrofuran AFM atomic force microscopy XPS x-ray photoelectron spectroscopy NMR nuclear magnetic resonance DSC differential scanning calorimetry TGA thermogravimetric analysis TEM transmission electron spectroscopy BTDA 3,3’,4,4’-benzophenonetracarboxylic dianhydride PSX polysiloxane

Tg glass transition temperature BAPP 2,2-bis[4-(4-aminophenoxy)phenyl]propane DSDA 3,3’.4,4’-diphenylsulfonetetracarboxylic dianhydride BAPSM bis[4-(3-aminophenoxy)phenyl]sulfone viii

Table of Contents

CHAPTER 1. SCOPE OF DISSERTATION...... 1

CHAPTER 2. LITERATURE REVIEW ...... 3 2.1 INTRODUCTION...... 3 2.2 POLYIMIDE SYNTHESIS...... 4 2.2.1 Polyimides from Poly(amic acid) Precursors ...... 4 2.2.1.1 Synthetic Aspects...... 4 2.2.1.2 Thermal Imidization of Poly(amic acid) Precursors ...... 5 2.2.1.3 Solution Imidization of Poly(amic acid) Precursors...... 8 2.2.1.4 Chemical Imidization of Poly(amic acid) Precursors...... 10 2.2.2 Polyimides from Poly(amic acid) Salt Precursors...... 10 2.2.2.1 Synthetic Aspects...... 10 2.2.2.2 Thermal Imidization Characteristics ...... 15 2.3 Synthetic Aspects of Polydimethylsiloxane Oligomers ...... 15 2.3.1 Polycondensation of Bifunctional Silanes...... 17 2.3.2 Ring Opening Polymerization of Cyclic Dimethylsiloxanes...... 20 2.4 BLOCK COPOLYMERS...... 25 2.4.1 Characteristics of Microphse Separated Copolymers...... 25 2.4.2 Formation of Poly(siloxane imide) Copolymers ...... 33 2.4.2.1 Randomly Segmented Copolymers...... 33 2.4.2.2 Perfectly Alternating Segmented Copolymers...... 39 2.4.2.3 Amic Ester Route to Segmented Copolymers ...... 43 2.5 CHARACTERIZATION OF PDMS CONTAINING BLOCK COPOLYMERS...... 45 2.5.1 Siloxane Surfaces ...... 45 2.5.2 Angle-Dependent X-Ray Photoelectron Spectroscopy (XPS) ...... 50 2.5.2.1 Principles of Operation ...... 50 2.5.2.2 Application of Angle-dependent XPS to Siloxane Containing Block Copolymers for Surface Characterization...... 54 ix

2.5.3 Atomic Force Microscopy (AFM) ...... 55 2.5.3.1 Principles of Operation ...... 55 2.5.3.2 Application of AFM to Block Copolymer Surface Characterization... 59 2.5.4 Molecular Weight Characterization...... 62 2.5.5 Copolymer Composition ...... 63 2.5.6 Thermal Behavior ...... 63 2.5.7 Morphology...... 65 2.5.8 Mechanical Properties...... 67 2.6 INTERPHASE APPLICATIONS ...... 67 2.6.1 Interphases for Composite Applications ...... 67 2.6.2 Interphases for Microelectronic Applications ...... 69 2.7 REFERENCES...... 75

CHAPTER 3. SYNTHESIS AND CHARACTERIZATION OF POLY(SILOXANE IMIDE) BLOCK COPOLYMERS...... 89 3.1 SYNOPSIS...... 89 3.2 EXPERIMENTAL...... 89 3.2.1 Purification of Solvents and Reagents...... 89 3.2.2 Preparation of Tetramethylammonium Siloxanolate Catalyst...... 92 3.2.3 Synthesis of Aminopropyl Terminated Polydimethylsiloxane (PDMS) Oligomers ...... 94 3.2.4 Derivatization of Aliphatic Amine Endgroups for Size Exclusion Chromatography Analysis ...... 96 3.2.5 Synthesis of BPDA/MPDA Poly(siloxane-amic acid) Block Copolymers...... 96 3.2.6 Synthesis of 6FDA/MPDA Poly(siloxane amic acid) Block Copolymers...... 98 3.2.7 Determination of Weight % Solvent in Poly(siloxane amic acid)...... 99 3.2.8 Preparation of BPADA/MPDAPoly(siloxane amic acid) Salt...... 99 3.2.9 Thermal Imidization of the Poly(siloxane amic acid)s and Poly(siloxane amic acid) Salts...... 100 x

3.3 CHARACTERIZATION...... 100 3.3.1 Molecular Weight Determination ...... 100 3.3.2 Thermogravimetric Analysis ...... 101 3.3.3 Differential Scanning Calorimetry...... 101 3.3.4 Dynamic Mechanical Analysis...... 102 3.3.5 Transmission Electron Microscopy...... 102 3.4 RESULTS AND DISCUSSION ...... 102 3.4.1 PDMS Oligomers...... 102 3.4.2 BPDA/MPDA Poly(siloxane imide) Block Copolymers ...... 107 3.4.2.1 Synthesis of BPADA/MPDA Poly(siloxane amic acid) Block Copolymers ...... 107 3.4.2.2 BPADA/MPDA Poly(siloxane imide) Block Copolymers from Poly(siloxane amic acid) Salt Precursors ...... 111 3.4.3 6FDA/MPDA Poly(siloxane imide) Block Copolymers ...... 114 3.4.3.1 Synthesis of 6FDA/MPDA Poly(siloxane amic acid) Block Copolymers ...... 114 3.4.4 Characterization of the Poly(siloxane imide) Block Copolymers...... 125 3.4.4.1 Molecular Weight Analysis ...... 125 3.4.4.2 Thermal Properties...... 125 3.4.4.3 Morphology...... 134 3.5 CONCLSIONS...... 136 3.6 RECOMMENDATIONS FOR FUTURE WORK...... 136 3.7 REFERENCES...... 137

CHAPTER 4. CONTROLLED MOLECULAR WEIGHT END-FUNCTIONAL POLYIMIDES FROM POLY(AMIC ACID) SALT PRECURSORS ...... 139 4.1 SYNOPSIS...... 139 4.2 EXPERIMENTAL...... 140 4.2.1 Materials and Their Purification...... 140 4.2.2 Synthesis of BPADA/MPDA Amine Terminated Poly(amic acid)...... 141 4.2.3 Synthesis of Hydroxy Terminated BPADA/MPDA Poly(amic acid).... 142 xi

4.2.4 Determination of Weight % Solvent in Poly(amic acid)...... 143 4.2.5 Preparation of Poly(amic acid) Salts ...... 143 4.2.6 Thermal Imidization of Poly(amic acid) Salts...... 144 4.3 CHARACTERIZATION...... 144 4.3.1 Nuclear Magnetic Resonance ...... 144 4.3.2 Molecular Weight Determination ...... 144 4.3.3 Thermogravimetric Analysis ...... 145 4.3.4 Differential Scanning Calorimetry...... 146 4.4 RESULTS AND DISCUSSION ...... 146 4.5 CONCLUSIONS...... 158 4.6 ACKNOWLEDGEMENTS...... 158 4.7 REFERENCES...... 158

CHAPTER 5. CHARACTERIZATION OF POLY(SILOXANE IMIDE) BLOCK COPOLYMER SURFACES...... 159 5.1 SYNOPSIS...... 159 5.2 EXPERIMENTAL...... 160 5.2.1 Materials...... 160 5.2.2 Preparation of Polymer Films for Surface Analysis ...... 162 5.2.3 Copolymer Composition ...... 163 5.3 CHARACTERIZATION...... 164 5.3.1 X-Ray Photoelectron Spectroscopy (XPS) ...... 164 5.3.2 Water Contact Angle Analysis...... 165 5.3.3 Atomic Force Microscopy ...... 165 5.4 RESULTS AND DISCUSSION ...... 166 5.4.1 Copolymer Composition ...... 166 5.4.2 XPS Surface Analysis of Block Copolymer Films...... 170 5.4.3 Water Contact Analysis of Block Copolymer Films...... 180 5.4.4 AFM Analysis of Block Copolymer Surfaces ...... 182 5.5 CONCLUSIONS...... 186 5.6 RECOMMENDATIONS FOR FUTURE WORK...... 190 xii

5.7 ACKNOWLEDGEMENTS...... 190 5.8 REFERENCES...... 191

CHAPTER 6. FUNCTIONAL POLYIMIDES AND POLY(SILOXANE IMIDE) BLOCK COPOLYMERS AS INTERPHASE MATERIALS ...... 192 6.1 SYNOPSIS...... 192 6.2 INTERPHASES FOR COMPOSITE APPLICATIONS ...... 194 6.2.1 EXPERIMENTAL...... 194 6.2.1.1 Materials...... 194 6.2.1.2 Preparation of Bilayer Films as Interphase Models ...... 196 6.2.1.3 Preparation of Epoxy - Polyimide Networks ...... 197 6.2.1.4 Carbon Fiber Sizing For Composite Fabrication...... 197 6.2.1.5 Measurement of Sizing Level and Uniformity ...... 198 6.2.2 CHARACTERIZATION...... 199 6.2.2.1 Transmission Electron Microscopy (TEM) and Atomic Force Microscopy (AFM) ...... 199 6.2.2.2 Fracture Toughness Analysis...... 199 6.2.2.3 X-Ray Photoelectron Spectroscopy (XPS) ...... 200 6.2.3 RESULTS AND DISCUSSION ...... 200 6.2.4 CONCLUSIONS ...... 214 6.2.4 RECOMMENDATIONS FOR FUTURE WORK ...... 214 6.2.5 REFERENCES...... 215 6.3 INTERPHASES FOR MICROELECTRONIC APPLICATIONS...... 216 6.3.1 EXPERIMENTAL...... 216 6.3.1.1 Materials...... 216 6.3.1.2 Block Polymer/Metal Interphase Study...... 218 6.3.1.3 PDMS/Silicon Wafer Interface Study...... 218 6.3.1.4 Residual Stress Analyses...... 219 6.3.2 CHARACTERIZATION...... 221 6.3.2.1 Refractive Index Analysis ...... 221 6.3.2.2 X-Ray Photoelectron Analysis...... 221 xiii

6.3.3 RESULTS AND DISCUSSION ...... 221 6.3.4 CONCLUSIONS ...... 243 6.3.5 RECOMMENDATIONS FOR FUTURE WORK ...... 243 6.3.6 ACKNOWLEDGEMENTS...... 244 6.3.7 REFERENCES...... 244 SUMMARY/CONCLUSIONS ...... 246 Vita...... 249 xiv

List of Figures:

Chapter 2

Figure 2-1. Schematic of the polymerization of 4-aminophthalic anhydride...... 4

Figure 2-2. Schematic of two-step synthesis of a polyimide from a diamine and a tetracarboxylic acid dianhydride...... 6

Figure 2-3. Schematic of the mechanism for an amine reacting with an anhydride to form the amic acid...... 7

Figure 2-4. Schematic of the equilibrium reactions occurring during imidization of a poly(amic acid)...... 9

Figure 2-5. Possible reaction mechanism for the solution imidization process...... 11

Figure 2-6. Schematic of the poly(amic acid) back reaction...... 12

Figure 2-7. Schematic of the formation of a poly(amic acid) salt using a tertiary amine...... 14

Figure 2-8. Imidization mechanisms for poly(amic acid) salts ...... 16

Figure 2-9. Structure for polydimethylsiloxane...... 17

Figure 2-10. Reaction scheme for the homofunctional polycondensation of silanediols...... 18

Figure 2-11. Schematic of polycondensation synthesis to yield a secondary amino-terminated PDMS ...... 19

Figure 2-12. Mechanism for the synthesis of PDMS using anionic equilibrium polymerization of D4 with a tetramethylammonium siloxanolate catalyst...... 22

Figure 2-13. Plot of weight fraction of cyclics in PDMS equilibrates as a function of siloxane volume % ...... 24

Figure 2-14. Schematic of block copolymer architectures...... 27

Figure 2-15. Classical microphase separated morphologies...... 28

Figure 2-16. The Meier model of a block copolymer domain...... 30 xv

Figure 2-17. Generalized modulus vs. temperature curves...... 32

Figure 2-18. Synthesis of randomly and perfectly alternating segmented copolymers ...... 34

Figure 2-19. General synthetic route of randomly segmented polyimide-polydimethylsiloxane copolymers...... 36

Figure 2-20. Plot of the polyimide phase glass transition temperature as a function of 1/xn ...... 38

Figure 2-21. Synthesis of perfectly alternating segmented engineering thermoplastics (ETP)-PDMS copolymers ...... 40

Figure 2-22. General schematic of the imide-amine transimidization reaction...... 41

Figure 2-23. Synthesis of perfectly alternating segmented poly(siloxane imide) copolymers via transimidization...... 42

Figure 2-24. Synthesis of a segmented copolymer via an amic ester precursor ...... 44

Figure 2-25. Schematic of PDMS structures depicting a flat helical chain conformation...... 47

Figure 2-26. Schematic of the near surface region of a PDMS containing block copolymer demonstrating a silicon enriched layer ...... 49

Figure 2-27. Diagram of photoelectron process and Auger process ...... 51

Figure 2-28. An illustration of the angle-dependent XPS setup...... 53

Figure 2- 29. Schematic of AFM showing the four main parts...... 57

Figure 2-30. Diagram of Tapping ModeÔ AFM...... 58

Figure 2-31. AFM height and phase image of PS-PB-PS triblock copolymer ...... 61

Figure 2-32. Model of an interphase region...... 70 xvi

Chapter 3

Figure 3-1. Synthetic scheme for the preparation of the tetramethylammonium siloxanolate catalyst...... 93

Figure 3-2. Synthetic scheme for the preparation of aminopropyl terminated PDMS oligomer...... 95

Figure 3-3. Synthetic scheme for the derivatization of alphatic amine endgroups...... 97

Figure 3-4. Size exclusion chromatograms of a 11,000 g/mole PDMS oligomer before and after stripping of the equilibrated cyclics...... 104

Figure 3-5. 1H NMR spectrum of PDMS oligomer with a targeted of 2000 g/mol...... 105

Figure 3-6. Sample calculation of the molecular weight using 1H NMR for a PDMS oligomer with targeted a of 2000 g/mol...... 106

Figure 3-7. Schematic of PDMS capping reaction with the anhydride monomer...... 109

Figure 3-8. Synthesis of BPADA/MPDA poly(amic acid) block copolymers...... 110

Figure 3- 9. Synthesis of BPADA/MPDA poly(siloxane imide) block copolymer via a poly(siloxane amic acid) salt intermediate...... 112

Figure 3- 10. Isothermal TGA analysis of poly(siloxane imide) block copolymer...... 113

Figure 3-11. Schematic of the synthesis of the fluorinated poly(siloxane amic acid) block copolymers...... 115

Figure 3-12. Size exclusion (RI) chromatograms of imidized block copolymers containing 1000 and 2000 g/mol PDMS oligomers...... 116

Figure 3-13. Size exclusion (RI and DV) chromatograms of an imidized block copolymer containing a 5000 g/mol PDMS oligomer...... 117

Figure 3-14. 1H NMR spectra monitoring the capping reaction of 5000 PDMS oligomer with 6FDA...... 119 xvii

Figure 3-15. Schematic explaining the relative nucleophilicity of the aliphatic versus the aromatic primary amine...... 120

Figure 3-16. Size exclusion (RI) chromatograph of a poly(siloxane imide) block copolymer prepared with 5000 g/mol PDMS oligomer (30 weight %) ...... 123

Figure 3-17. Polyimide phase glass transition temperature as a function of 1/Mn (hard block) of copolymers...... 128

Figure 3-18. TGA curves for BPADA/MPDA block copolymers prepared with the 2000 g/mol PDMS oligomer...... 130

Figure 3-19. Generalized schematic of char formation for siloxane containing ...... 132

Figure 3-20. TEM micrograph of an UltemÔ type poly(siloxane imide) block copolymer containing 70 weight % siloxane (~ 2000g/mol) and 30 weight % imide ...... 135

Chapter 4

Figure 4-1. Synthetic route to amine terminated BPADA/MPDA poly(amic acid) ...... 147

Figure 4-2. Synthetic route to phenol terminated BPADA/MPDA poly(amic acid) ...... 148

Figure 4-3. Sample calculation for residual THF weight % emaining in poly(amic acid) sample ...... 150

Figure 4- 4. Formation of BPADA/MPDA poly(amic acid) salt from poly(amic acid) precursor...... 151

Figure 4-5. Isothermal TGA analysis of an end-functional poly(amic acid) salt ...... 154

Figure 4-6. Glass transition temperature vs. 1/Mn for end-functional BPADA/MPDA Polyimides ...... 157 xviii

Chapter 5

Figure 5-1. Sample calculations for the determination of block copolymer compositions...... 163

Figure 5-2. Setup for water contact angle analyses...... 165

Figure 5-3. 1H NMR spectrum of an BPADA/MPDA poly(siloxane imide) block copolymer...... 168

Figure 5-4. 13C NMR spectra of a 6FDA/MPDA poly(siloxane imide) block copolymer ...... 169

Figure 5-5. XPS wide scans...... 172

Figure 5-6. Angle dependent XPS data depicting PDMS surface concentrations in block copolymer films as a function of PDMS bulk weight %...... 177

Figure 5-7. Angle dependent XPS data for fluorinated block copolymers containing 5 weight % (5000 g/mol) and 4 weight % (11000 g/mol) PDMS...... 178

Figure 5-8. Influence of PDMS oligomer molecular weight ( at 10 weight % bulk concentration) on the surface composition...... 179

Figure 5-9. Water contact angles of copolymer films as a function of PDMS content ...... 181

Figure 5-10. AFM phase images for fluorinated polymers with varying amounts of incorporated PDMS ( » 2000 g/mol)...... 183

Figure 5-11. AFM surface topographic images of BPADA/MPDAhomo and copolymers with PDMS weight percentages ...... 184

Figure 5-12. AFM surface topographic images of fluorinated homo and copolymers...... 185

Figure 5- 13. Diagram relating the block copolymer surface arrangement to the formation of monolayer films...... 187

Figure 5-14. Graph of bulk PDMS weight % plotted against mean roughness ...... 189 xix

CHAPTER 6

Figure 6-1. Preparation of interphase material ...... 193

Figure 6-2. Schematic of three point bend specimen...... 200

Figure 6-3. Schematic of epoxy-polyimide network synthesis...... 202

Figure 6-4. Transmission electron micrographs of epoxy- polyimide networks...... 203

Figure 6-5. Epoxy-polyimide (or poly(siloxane imide) block copolymer) bilayer films were investigated as model interphase materials...... 206

Figure 6-6. AFM images polyimide/epoxy bilayer films ...... 207

Figure 6-7. AFM image showing entire interphase region...... 208

Figure 6-8. AFM images of poly(siloxane imide) (PSI) bilayers...... 210

Figure 6-9. Depiction of lab scale sizing line ...... 212

Figure 6-10. SEM micrographs of polyimide sized carbon fibers ...... 213

Figure 6-11. Ganiometer setup used for polymer/silicon wafer interface study...... 219

Figure 6-12. Single cantilever beam setup for residual stress analyses ...... 220

Figure 6-13. A schematic diagram for the model of an oxide- covered metal showing the various layers and their compositions...... 225

Figure 6-14. XPS wide scan and Ta 4f7/2,Ta 4f5/2 photopeaks or tantalum film...... 227

Figure 6-15. XPS wide scan and the Ti 2p3/2, Ti 2p1/2 photopeaks for titanium substrate...... 228

Figure 6-16. Interfacial PDMS surface compositions for the polymer/metal interphase study ...... 230

Figure 6-17. Schematic showing the interaction of the PDMS component of the block copolymer with the metal substrate...... 231 xx

Figure 6-18. XPS wide scan and Si 2p3/2 photopeaks for the silicon wafer ...... 234

Figure 6-19. Ideal deflection curve for a polymer coated strip for several temperature cycles...... 235

Figure 6-20. Deflection curve for a 6FDA/MPDA polyimide cast on titanium...... 236

Figure 6-21. Deflection curve for a 6FDA/MPDA poly(siloxane imide) block copolymer containing 30 weight % PDMS ( = 5000 g/mol) cast on titanium...... 237

Figure 6-22. Deflection curve for a 6FDA/MPDA poly(siloxane imide) block copolymer containing 20 weight % PDMS ( = 2000 g/mol) cast on titanium...... 238

Figure 6-23. Illustration of the radius of curvature, R, determined using the single cantilever beam setup...... 241 xxi

List of Tables:

Chapter 2

Table 2-1. The yield of linear organofunctional polysiloxane oligomers in undiluted equilibrate...... 26

Table 2-2. Qualitative surface properties of PDMS ...... 46

Table 2-3. The influence of copolymer composition on the thermal stability of poly(siloxane imide) block copolymers in air ...... 64

Table 2-4. Required properties of polymers for microelectronic applications ...... 72

Table 2-5. Influence of molecular structure on the dielectric behavior ...... 74

Chapter 3

Table 3-1. Number average molecular weight, , results of aminopropyl terminated PDMS oligomers...... 108

Table 3- 2. Monomer molar concentrations for a 6FDA/MPDA poly(siloxane imide) block copolymer containing 30 weight % PDMS...... 122

Table 3-3. Copolymer composition and molecular weight of block copolymers...... 124

Table 3- 4. Thermal properties of poly(siloxane imide) block copolymers...... 126

Table 3-5. Char yield remaining at 750°C for poly(siloxane imide) block copolymers in air and nitrogen atmospheres...... 131

Table 3- 6. Binding energies for the silicon photopeak of a lock copolymer before and after exposure to 750°C in air and nitrogen atomspheres...... 133 xxii

Chapter 4

Table 4-1. Molecular weight results for polyimides from stable poly(amic acid) salt solutions ...... 152

Table 4-2. Molecular weight results for end-functional polyimides from amic acid and amic acid salt precursors ...... 155

Table 4-3. Thermal properties of end-functional BPADA/MPDA polyimides ...... 156

Chapter 5

Table 5-1. Compositions of BPADA/MPDA block copolymers used in the surface characterization study ...... 160

Table 5-2. Compositions of fluorinated block copolymers used in the surface characterization study...... 161

Table 5-3. Copolymer composition before and after extraction...... 167

Table 5-4. XPS binding energies (BE) and chemical states...... 173

Table 5-5. XPS binding energies (eV) for the polyimide homo- and copolymers referenced to C1s (285.0 eV)...... 174

Table 5-6. Elemental surface compositions compared with the bulk stoichiometries ...... 175

Chapter 6

Table 6-1. BPADA/MPDA poly(siloxane imide) block copolymer compositions used in the composite interphase study ...... 196

Table 6-2. Fracture toughness of epoxy-polyimide networks prepared with DER 331 epoxy resin...... 205

Table 6-3. Surface composition in atomic % of the BPADA/MPDA polyimide sized carbon fibers...... 211

Table 6-4. 6FDA/MPDA block copolymer compositions used in interphase study for electronic applications...... 216 xxiii

Table 6-5. Refractive indices and calculated dielectric constant for 6FDA/MPDA block copolymers polymer films...... 223

Table 6-6. XPS binding energies for the metal and metal oxide layers...... 226

Table 6-7. Surface compositions of metal substrates before coating and after film removal in atomic %...... 232

Table 6-8. Probe travel distance of cast block copolymer films...... 240

Table 6-9. Preliminary results for curvature measurements of thin films...... 242 1

CHAPTER 1. SCOPE OF DISSERTATION

The research presented in this dissertation is focused on the preparation of polyimide and poly(siloxane imide) block copolymers for fiber reinforced composite and microelectronic applications. As pictured below, the use of these materials as interphase agents not only requires knowledge of the application requirements but a fundamental understanding of the structure-property relationships, bulk morphology, surface properties, thermal properties, interfacial properties, and although not discussed herein, mechanical and adhesive properties as well. These areas are addressed in terms of the BPADA/MPDA and 6FDA/MPDA structures employed.

Structure-property relationships

Surface Properties Adhesive Properties

Interphase Applications

Mechanical Properties Thermal Properties

Bulk Morphology Interfacial Properties

In Chapter 3, the synthesis and characterization of two series of poly(siloxane imide) block copolymers are described. Fluorinated block copolymers were prepared via poly(siloxane amic acid) intermediates for microelectronic applications. The block copolymers with BPADA/MPDA structures were prepared via poly(siloxane amic acid) salt intermediates for composite applications. Chapter 4 presents the synthesis of controlled molecular weight end-functional polyimides via poly(amic acid) salt intermediates. The 2 water soluble salt precursors are ideal for sizing carbon fibers for composite applications. Chapter 5 presents the analysis of block copolymer surfaces with systematically varied compositions using various techniques. The techniques confirmed siloxane enrichment of the surface area. Chapter 6 explores the application of these materials as interphase agents. 3

CHAPTER 2. LITERATURE REVIEW

2.1 INTRODUCTION

There has been enormous commercial success in the synthesis of various copolymers comprised of different combinations of repeating units. Copolymer synthesis offers the ability to tailor the properties of a homopolymer in a desired direction by the introduction of an appropriately chosen second repeating unit. Copolymerization has been used to alter such polymer properties as flexibility, thermal stability, toughness, glass transition temperature, and crystallinity. The magnitudes and direction of the property alterations differ depending on whether statistical, alternating, or block copolymers are involved. Polyimides have become one of the most important and versatile classes of high performance polymers due to excellent mechanical and thermal properties.1-7 Incorporation of flexible polysiloxane components into the already thermally stable polyimide has been shown to yield several attractive properties while retaining many of the excellent properties of the corresponding polyimide homopolymers.3,4 These block copolymers possess good processability, low water absorption, atomic resistance, low dielectric constants, and excellent adhesion. Thus, these materials have become attractive candidates for microelectronic, printed circuit, and aerospace applications. Review articles have appeared describing the current and developing technology of this important class of polymers.5-6 This introduction will focus primarily on the areas closely related to the research topic. The literature review is divided into six sections. The first two sections discuss common synthetic methods used in the preparation of polyimides and polydimethylsiloxane oliogmers. The third section discusses some general aspects and the synthesis of microphase separated block copolymers. The fourth section covers techniques for the characterization of poly(siloxane imide) block copolymers. The final section introduces interphase applications for polyimides and poly(siloxane imide) block copolymers. 4

2.2 POLYIMIDE SYNTHESIS

Polyimides are condensation polymers derived from organic diamines and organic tetracarboxylic acids (or derivatives). The earliest report of polyimides was by Bogert and Renshaw.8 The authors reported that 4-aminophthalic anhydride evolved water at elevated temperatures with the possible formation of a “polymolecular imide” (Figure 2-1). The earlier polyimides were insoluble and infusible. The two-step method consisting of the formation of the poly(amic acid) precursor followed by cyclodehydration allowed the development of fully aromatic materials, which in selective cases remained soluble. Aromatic polyimides have become one of the most important classes of high-performance polymers. Due to their excellent electrical, thermal, and high temperature mechanical properties, aromatic polyimides have found many applications as high-temperature insulators and dielectrics, coatings, adhesives, and matrices for high- performance composites.2,3

O O C C heat O N -H2O C C H2N O O n Figure 2- 1. Schematic of the polymerization of 4-aminophthalic anhydride.

2.2.1 Polyimides from Poly(amic acid) Precursors

2.2.1.1 Synthetic Aspects

The two-step method for the preparation of polyimides consists of polymerization of a soluble poly(amic acid) intermediate, followed by dehydration of this prepolymer to yield the final polyimide. As illustrated in Figure 2-2, this 5 approach is based on the reaction of a diamine with a tetracarboxylic acid dianhydride in a polar, aprotic solvent such as dimethylsulfoxide (DMSO), dimethylacetamide (DMAC), dimethylformamide (DMF), or N-methylpyrolidone (NMP), at temperatures between 15 and 75°C.9-14 The resultant poly(amic acid) is then cyclized either thermally or chemically in a subsequent step to produce the desired polyimide. Figure 2-3 contains the poly(amic acid) formation mechanism. The formation of the poly(amic acid) is an equilibration reaction where the forward reaction is thought to start with the formation of a charge transfer complex between the dianhydride and the diamine.15-17 The more electrophilic the dianhydride, the more susceptible it is to nucleophilic attack. The reactivity of the dianhydride monomer has been related to its electron affinity. Higher values suggest greater reactivity of the dianhydride.3 Strong electron withdrawing groups activate the anhydride for nucleophilic acyl attack on the anhydride carbonyl. The reactivity of the diamine is related to its basicity. The rate constants for amidization increase as the pKa of the protonated amine increases. Highly basic amines (e.g. aliphatic amines) may form salts during the initial stages of the reaction, upsetting the stoichiometry and preventing the attainment of high molecular weight. Thus, these materials are not suitable for this reaction

18 scheme. Ideally, the diamine should have a pKa of 4.5-6.

2.2.1.2 Thermal Imidization of Poly(amic acid) Precursors

The second step of the two-step method of classical polyimide synthesis is imidization of the poly(amic acid) precursor (Figure 2-2). As with the first step, the polyimide cyclodehydration process is complex with the possibility of various side reactions. Depending on the imidization kinetics, the rate of the back reaction to the monomers may increase, possibly resulting in a reduction of the molecular weight. The water by-product produced can react with the anhydride groups to produce dicarboxylic acid groups (Figure 2-4). Additional heating 6

O O STEP 1 C C O R O + H2N R' NH2 C C Diamine O O

Dianhydride

O O H H O O C C N R' N C C OH R R HO C C OH HO C C N R' N

O O x O O H H y Poly(amic acid)

STEP 2 heat -H2O

O O C C N R N R' + H2O C C

O O n Polyimide

R = cycloaliphatic, or aromatic molecule R' = aromatic molecule

Figure 2- 2. Schematic of two-step synthesis of a polyimide from a diamine and a tetracarboxylic acid dianhydride. 7

O O

O + H N O 2 H N O O H

O O O OH H N N H O H O Figure 2- 3. Schematic of the mechanism for an amine reacting with an anhydride to form the amic acid. 8 converts the dicarboxylic acid groups to their cyclized form. Once cyclized, the dianhydrides can re-react with an amine endgroup to form an amic acid, which can then imidize, thus recovering the initial molecular weight. The most common method of converting the poly(amic acid) to the polyimide is by bulk (or melt) imidization.19-21 Films of the poly(amic acid)s are often cast from polar aprotic solvents and subsequently dried and imidized. A typical imidization cycle may proceed by the following schedule: 30 minutes at 100°C, 30 minutes at 200°C, followed by 30 minutes at 300°C. The imidization kinetics are determined by the cylization temperature and the Tg of the material.

Complete cyclization is accomplished at imidization temperatures above the Tg of the material to ensure chain mobility. Imidization temperatures close to or below the Tg of the material greatly affects the % conversion of the poly(amic acid) moieties. Investigations have shown that the rate of imidization is initially rapid and decreases over time reflecting the decrease in chain mobility.22

2.2.1.3 Solution Imidization of Poly(amic acid) Precursors

Poly(amic acid)s may also be imidized in solution. Cyclodehydration can be effected at lower temperatures, thus avoiding side reactions resulting from high temperature imidization. Solution imidization offers a particularly attractive method since one pot syntheses to cyclized polyimides are possible.23-26 The imidization process is affected by heating the poly(amic acid) NMP solution to 150°C to 180°C in the presence of an azeotroping agent such as o- dichlorobenzene, xylene or cyclohexyl pyrrolidone which is added to remove the water that is liberated during the reaction. Complete imidization can be achieved within 16 to 24 hours. Kim et al.26 investigated the kinetic and mechanistic aspects of thermal solution imidization processes at temperatures between 140° and 180°C. Concentration dependent experiments were conducted to determine the reaction order of the thermal imidization process. The data revealed that the reaction 9

O O

CH3 N N O O CH O 3 O

Polyimide n - H2O + H2O

O O OH CH3 N O O NH CH3 O O n Poly(amic acid)

O O

CH3 N O H2N O O CH O 3 O Anhydride and Amine Endgroups

- H2O + H2O

O O OH CH3 N H2N NH2 O O OH CH O 3 O Dicarboxylic Acid and Amine Endgroups

Figure 2- 4. Schematic of the equilibrium reactions occurring during imidization of a poly(amic acid). 10 mechanism was second-order and was catalyzed by the carboxylic acid groups on adjacent or nearby repeat units (Figure 2-5). This proved to be the rate determining step (r.d.s.). From the study, the authors concluded that solution imidization could be defined as a chemical process where the kinetics were determined by the chemical reactivities of the amic acid groups.

2.2.1.4 Chemical Imidization of Poly(amic acid) Precursors

Poly(amic acid)s may also be chemically imidized. This is accomplished by using chemical dehydrating agents in combination with basic catalysts.27,28 Various reagents have been employed including acetic anhydride, propionic anhydride, and n-butyric anhydride as dehydrating agents and pyridine, triethylamine, an disoquinoline as basic catalysts. Chemical imidizations carried out using triethylamine and anhydrides were reported to yield the polyimide exclusively with fast reaction rates.28

2.2.2 Polyimides from Poly(amic acid) Salt Precursors

2.2.2.1 Synthetic Aspects

The hydrolytic instability of the poly(amic acid) has been given much attention. The postulated mechanism for the poly(amic acid) back reaction to the anhydride and amine has been presented as the explanation for the hydrolytic instability of the system (Figure 2-6). Some applications require processing of the insoluble polyimides from their soluble poly(amic acid) precursors. Thus, hydrolytic stability becomes a key concern. 11

O C O H

CONH O HO OH C OH r.d.s. NH

C NH O O

-H+

O HO OH -H2O N NH

O O

Figure 2- 5. Possible reaction mechanism for the solution imidization process. 12

O O C OH O (proton transfer)

C N C N O H O H H Poly(amic acid) (ring closure to the anhydride) O O

O + H2N O

N O HO Formation of the dianhydride and amine H results in backbone cleavage.

Figure 2- 6. Schematic of the poly(amic acid) back reaction. 13

Several approaches to increasing the hydrolytic stability of the poly(amic acid) have emerged including the ester-acid route29 and formation of poly(amic ester)s intermediates30. The simplest approach is to eliminate the proton transfer in the internal acid catalyzed formation of the intermediate.31 This has been accomplished by neutralizing the pendent carboxylic acid groups with a base, such as a tertiary amine, to form a polyelectrolyte (or salt) (Figure 2-7). Several researchers have used this methodology.7,22,31-35 One of the earliest reports was in a patent by Endrey (du Pont de Nemours and Company).33 The results showed the conversion of poly(amic acid)s to poly(amic acid) salts by the addition of triethylamine to the polymerization solution. It was found that when 130% of the theoretical amount of base (based on titrated free carboxylic acid groups) was added very stable solutions could be obtained.22,37 Not only does this approach offer the advantage of a more hydrolytically stable intermediate, it also yields a water processible polyimide precursor. Water solubility can be an important processing advantage since it would allow the salts to be applied as fiber sizings from aqueous solutions or to be incorporated into waterborne resins as thermoplastic toughening agents for thermosets. This would also eliminate the need for harmful organic solvents. The water solubility of several BPADA/MPDA poly(amic acid) salts was evaluated by Facinelli et al.35 The authors prepared tetra- and trialkylammonium salts. The results showed that the water solubility of the poly(amic acid) salts followed a trend consistent with polyelectrolyte theory. As the tetraalkylammonium counterion size was increased from tetramethyl- to tetraethyl- to tetrapropylammonium, the salts became soluble in water. The authors concluded that this was because the carboxylate charges are increasingly less shielded with the larger cations, and hence, the polyelectrolyte becomes more soluble. 14

O O OH O HNR3 R3N N N O H O H Figure 2- 7. Schematic of the formation of a poly(amic acid) salt using a tertiary amine. 15

2.2.2.2 Thermal Imidization Characteristics

The thermal imidization of poly(amic acid) salts was found to be faster than the corresponding poly(amic acid). Kruez and coworkers reported that the imidization rates of tertiary salts of PMDA and 4,4’-diaminodiphenyl ether proceeded 4 to 10 times faster than the corresponding poly(amic acid). It is believed that tertiary amines promote ring closure, thus leading to faster imidization rates.22 In a study by Facinelli and coworkers35 it was reported that the poly(amic acid) salt counterion controls the mechanism by which the salt imidizes, thus controlling the properties of the final polyimide (Figure 2-8). In the case of the triethylammonium poly(amic acid) salts, imidization yielded controlled molecular weight polyimides. It was believed that an equilibrium route exists where the carboxylate anion can simply abstract the proton associated with the protonated tiethylamine, to result in the poly(amic acid) and triethylamine. The data suggested that the trialkylamines formed during the conversion to acid might act in situ to facilitate the imidization by acting as plasticizing agents and rate- increasing catalysts for imidization.

2.3 SYNTHETIC ASPECTS OF POLYDIMETHYLSILOXANE OLIGOMERS

Linear polysiloxane polymers are prepared by either polycondensation of bifunctional silanes or by ring-opening polymerization of cyclic oligosiloxanes (Figure 2-9). Historically, polysiloxanes were first synthesized in 187236; however, they did not gain commercial importance until decades later. Some of the earlier applications included fluids, elastomers, and resins. Independent of their industrial use, polysiloxanes have been extensively studied in polymer science as a model in the polycondensation process37, macrocyclization theory38, and rubber elasticity39. 16

Tetraethylammonium Salt of PAA

O (a) O H O H C C N C N O O O N heat heat -H2O C C O C OH O O O + H CH2 CH2 N CH2 CH3 3 CH2CH2 1 mole + N(CH2CH3)3 Hoffman Elmination O O H (b) O H C C N O O C N N O heat heat C -CH2CH2OH C O C OCH2CH3 O CH CH N CH CH O 3 2 2 33 O + N(CH2CH3)3 1 mole Nuleophilic Substitution

Tetramethylammonium Salt of PAA

O O H O H C C N O O C N N O heat heat C -CH3OH C O C OCH3 O H C N CH O 3 3 3 O + N(CH3)3 1 mole Nuleophilic Substitution

Triethylammonium Salt of PAA

O O H O H C C N C N O O O N heat heat -H2O C C O C OH O O O + N(CH2CH3)3 H N CH2 CH33 1 mole

Equilibration

Figure 2- 8. Imidization mechanisms for poly(amic acid) salts.35 17

Linear (PDMS) are found in a wide range of commercial applications due to their excellent properties.40 These properties include an extremely low glass temperature (-123°C), high thermal and oxidative stability, UV resistance, low surface energy and hydrophobicity, good electrical properties, high permeability to many gases, and low toxicity.

CH3 Si O n CH3 Figure 2- 9. Structure for polydimethylsiloxane.

2.3.1 Polycondensation of Bifunctional Silanes

The polycondensation process of bifunctional silanes involves first hydrolysis followed by hydrolytic polycondensation (Figure 2-10). In the presence of an excess of water the silane is fully transformed into the silanediol and polymerization results from the polycondensation of a homofunctional silanediol.44,45 The linear and cyclic products can then equilibrate as described in the section below. The most commonly used monomer is . Burns et al.42 utilized this methodology and endcapped the resultant PDMS silanol fluid with a cyclic to produce a secondary aminoalkyl functional siloxane (Figure 2-11). Silanol condensation is carried out in the presence of a catalyst and the kinetics were found to be influenced by the reaction solvent. Numerous studies have been directed at understanding the polycondensation reaction under acidic and basic conditions in various solvents.42,44-46 The kinetic studies of the polycondensation of oligodimethylsiloxane-a,w-diol in the presence of acids revealed three processes: linear condensation, cyclization, and 18

R R

Cl Si Cl + 2H2O HO Si OH + 2HCl R R R nR Si(OH) 2 2 HO Si O H + [OSiR2]y + (n - 1)H2O

R x Figure 2- 10. Reaction scheme for the homofunctional polycondensation of silanediols. 19

CH3 CH3 CH3 Cl Si Cl + H2O HO Si OH + HCl HO Si O H n CH3 CH3 CH3

H3C CH3 Si N CH3 CH3 CH3 HO Si O H H C NHCH CHCH Si O Si CH CHCH NH 3 2 2 n 2 2 n CH3 CH3 CH3 CH3 CH3 CH3 CH3

Figure 2- 11. Schematic of polycondensation synthesis to yield a secondary amino-terminated PDMS.42 20 disproportionation. Researchers have found that the condensation of silanols is an equilibrium process.47 Hydrolysis can be detrimental to the molecular weight of the polycondensation product, thus the water must be removed from the system. High temperature in combination with the presence of water can have a serious effect on the structure.43

2.3.2 Ring Opening Polymerization of Cyclic Dimethylsiloxanes

The ring opening polymerization process allows good control over molecular weight by adjusting the stoichiometry of cyclics to endgroup reactants. These reactions can be carried out under anionic or cationic conditions and are often classified as either a kinetically or thermodynamically controlled process.

The kinetically controlled route uses hexamethylcyclotrisiloxane (D3), whereas the thermodynamically controlled route uses octamethyl- cyclotetrasiloxane (D4). Kinetically controlled polymerizations of cyclic siloxanes

CH3 CH3 H3C Si O CH3 H C O O Si CH3 3 Si Si Si O H C CH3 H3C 3 O O O Si CH Si 3 CH3 H3C CH3 CH3

D4 D3

are based on the increased reactivity of the siloxane bond in ring-strained monomers (eg. D3). The release of ring strain promotes the selective reaction of the D3 monomer (versus reaction with the Si-O bonds in the chains) using anionic conditions with lithium counterions, particularly in nonpolar solvents such as cyclohexane. Under carefully kinetically controlled conditions the polymerization 21 may approach classical living polymerization conditions.48 Thus, polymers with narrow molecular weights can be produced. On the other hand, thermodynamically controlled reactions occur when the system is allowed to reach equilibrium conditions. For polysiloxanes this results in a distribution of both linear and cyclic chains.

R2SiO R2SiO R2SiO n m n (1)

The most often used synthetic pathway for the preparation of organofunctional polysiloxanes is by anionic ring opening equilibration polymerization of D4. This area has been the extensively reviewed in the literature.47,49,51 The process includes three general steps: (1) initiation where the base catalyst attacks the silicon producing the silanolate end-group, (2) propagation-depropagation, and (3) chain equilibration. Figure 2-12 presents the synthetic route for the preparation of aminopropyl PDMS via equilibrium anionic ring opening polymerization of the cyclic tetramer

52 D4 in the presence of a siloxane chain transfer agent (end-blocker). The tetramethylammonium hydroxide generates the corresponding siloxanolate catalyst. Reaction of the catalyst with the siloxane chain transfer agent (or end- blocker) yields the base catalyzed initiator species. Equilibration of the base catalyzed initiator species with D4 results in linear chains and various cyclic species. Upon heating the reaction to 145°C the anions couple and by-products of trimethylamine and dimethylether are evolved to produce a neutralized, stable material.53 The remaining cyclics can be easily vacuum distilled. 22

CH3 CH CH Si O 3 CH CH 3 CH3 3 CH CH 3 H3C 3 3 CH O Si CH3 HO 80°C, 42 hours 3 N CH3 N N n + O Si O 4n+2 Si O CH3 Si O CH3 (coupling of H3C CH3 CH3 CH CH O Si CH3 two species) CH3 3 3 CH3 CH3 CH3 tetramethylammonium siloxanolate catalyst D4 CH3 CH3

H2N CH2 Si O Si CH2 NH2 3 3 CH3 CH3 siloxane end-blocker

CH3 CH3 CH3 CH3 CH3 CH3 H3C CH3 N O Si O Si O Si CH NH H N N 2 3 2 2 CH2 3Si O x CH3 H3C CH3 CH3 CH3 CH3 CH3 CH3 base catalyzed initiator species

(80°C, 24hours) Equilibration with D4 (145°C, 3 hours)

CH3 CH3

H2N CH2 Si O Si CH2 NH2 + Cyclics 3 n 3 CH3 CH3 Figure 2- 12. Mechanism for the synthesis of PDMS using anionic equilibrium polymerization of D4 with a tetramethylammonium siloxanolate catalyst. 23

A variety of catalysts have been successfully utilized. Protic acids are the most common initiating species used in cationic ring opening polymerizations. 54-56 Other initiating systems include Lewis acids57 and carbenium salts58. The initiating ability of protic acids increases with the acid strength. The basic Initiators used have included alkali metal hydroxides and tetraorganoammonium and phosphonium hydroxides.59 Also other strong inorganic and organic bases have been employed.60 The rate of polymerization is strongly affected by the counterion of the base. The rate increases significantly in the series of

+ + + + + + + 59,61,63 counterions: Li < Na < K < Rb < Cs » NEt4 » PEt4. This was believed to be a result of the reduction in ion pairing. The bulkier counterions loosen the ion aggregates of the silanolate centers. Kopylov et al.64 reported that the polymerization of D3 with tetramethylammmonium hydroxide showed first order kinetics with respect to the silanolate, thus, suggesting that the aggregation of the silanolate centers was insignificant. During ring opening polymerization the growing polymer chain can perform a “backbiting” maneuver where it can break the partially ionic Si-O bond present along the linear backbone resulting in the production of cyclics. Studies on the distribution of linear chains and cyclics after equilibration was first carried

61 out by Scott using [(CH3)2SiO]x (x=4-6) cyclics and was later investigated in a pioneering study by Brown and Slusarczuk in toluene65. The studies revealed that the equilibrium between the linear chains and cyclics determines the polymer yield, molecular weight and molecular weight distribution. The concentration of the equilibrium cyclics and the linear polymer is independent of the initial monomer concentration. As a consequence dilution of the system with a solvent results in a decrease in the linear chain yield so that the cyclic concentration remains the same (Figure 2-13).66 Thus, the equilibration approach can only effectively be performed in bulk or concentrated solutions. 24

0.8

0.6

0.4

0.2

Weight Fraction of Rings

100 80 60 40 20 Volume % Siloxane

Figure 2- 13. Plot of weight fraction of cyclics in PDMS equilibrates as a function of siloxane volume %. “Calculated” refers to the weight fraction of cyclics determined by the author. The graph shows the effect of dilution on the total cyclic concentration.66 25

The equilibration concentration of cyclics in bulk corresponds to the critical concentration of the system. Wright and Semlyen67 reported that the equilibrium concentration of total cyclics in bulk increases considerably with the increasing size and polar character of the substituent (Table 2-1).

2.4 BLOCK COPOLYMERS

2.4.1 Characteristics of Microphase Separated Copolymers

Block copolymers are composed of two chemically dissimilar bonded polymer segments. The sequential arrangement of the block copolymer results in five basic architectures: A-B diblock or A-B-A triblock copolymers, (A-B)n multiblock copolymers, radial or star block copolymers, and graft copolymers (Figure 2-14). 68,69 PDMS containing polyimide copolymers yield segmented or multiblock copolymers. The architecture influences such properties as the elastomeric behavior, melt rheology, and toughness in rigid materials. At a critical molecular weight of each segment, the incompatible copolymer segments phase separate similar to the behavior of incompatible blends. Unlike blends, the copolymer segments are prevented from macroscopically separating due the covalent bonds between them. This restricts the size of the phases to the microscopic scale.69,70 When phase separation occurs microphases of well defined size and shape are typically formed. A variety of ordered morphologies can be achieved depending on the segmental length and composition. As depicted in Figure 2-15, these morphologies include spheres, rods, lamellae, inverted rods, and inverted spheres.71,72 The properties of copolymers comprised of “hard” and “soft” polymer segments are determined by their relative composition. For example, copolymers with small domains of the “soft” segment embedded throughout the continuous “hard” phase behave as toughened glassy polymers.73 The inverted morphology behaves as a thermoplastic elastomer.74 26

Table 2- 1. The yield of linear organofunctional polysiloxane oligomers in undiluted equilibrate.67

R R’ Approximate yield % Me Me 82 Me Et 74

Me CF3CH2CH2 17 pH Me 70 pH pH 0 27

AA-B-B diblock diblock copolymer copolymer

TheThe addition addition of of a a third third block block producesproduces an an A A-B-B-A-A or or B B-A-A-B-B triblocktriblock copolymer. copolymer. ( )n

CouplingCoupling blocks blocks together together gives rise to [A-B]n multi- gives rise to [A-B]n multi- blockblock copolymers copolymers ThreeThree or or more more diblock diblock wherewhere n n is is greater greater than than 1. 1. copolymerscopolymers may may be be coupledcoupled at at a a single single junctionjunction point point to to formform radial radial or or star star blockblock copolymers. copolymers.

SequencesSequences of of one one monomer monomer “grafted” “grafted” onto onto a a “backbone” “backbone” of of a a second second monomer monomer result result in in graft graft copolymers. copolymers. CopolymerizationCopolymerization results results from from the the formation formation of of active active sites sites at at points points on on the the polymer polymer molecules molecules other other than than the the ends, ends, followedfollowed by by exposure exposure to to a a second second monomer. monomer.

Figure 2- 14. Schematic of block copolymer architectures.68,69 28

A A A, B B B Spheres Cylinders Lamallae Cylinders Spheres Increasing A-Content Decreasing B-Content

Figure 2- 15. Classical microphase separated morphologies.73,74 29

Many statistical thermodynamic theories have been proposed in the literature predicting the domain morphology as a function of various parameters.75-77 An illustration of the formation of a block copolymer domain is given in Figure 2-16.82 The size of these microphase separated structures is of the order of magnitude of the radius of gyration of the macromolecules. The size or periodicity of domains can be related to its molecular weight by the following equations:

Z Z d = kM and R = kM A (2) where M is the total molecular weight of the chain, d is the interdomain spacing,

MA is the molecular weight of the spherical domain forming block, R is the domain radius for a spherical domain and K is a constant. Small angle x-ray scattering (SAXS), small-angle neutron scattering (SANS), and transmission electron spectroscopy (TEM) have been used to investigate the interdomain spacings and the morphological structure of microphase separated block copolymers.79-81 Variations in casting solvent and evaporation rate have been found to alter the microphase morphology by changing the domain shape or the phase purity. Investigations using SAXS by McIntyre and Campos-Lopez82 revealed different scattering patterns for styrene-isoprene-styrene (SIS) and styrene-butadiene- styrene (SBS) triblock copolymers cast from different solvents. A diffuse interphase was observed between the styrene and diene phases cast from CCl4.

The researchers concluded the CCl4 was a good solvent for both blocks thus rendering them compatible. Similarly, Matsuo and Sagaye83 found that the bulk morphology depended upon both the copolymer composition and the solvent. Sphere or lamellar type domain morphologies could be obtained by varying either the solvent or overall composition. Noshay and McGrath69 discussed the effects of solvent evaporation rate on SBS block copolymers. The authors reported that rapidly evaporating the solvent produced a lamellar domain structure, while slowly evaporating the solvent produced spherical PS domains in a PB matrix. The most obvious method for identifying phase separated block copolymers is by visual inspection. Macroscopically phase separated 30

A B

Figure 2- 16. The Meier model of a block copolymer domain.78 31 materials appear cloudy or opaque due to the relatively large domain or phase dimensions and the different refractive indices often present in the two phases. However, the domain sizes of microphase separated materials are very small thus the films often appear transparent and could be mistaken as homogeneous materials. An alternative method of identifying phase separated materials is through characterization of the thermal transitions of the polymer film. In phase separated block copolymers the thermal properties are similar to physical blends of the same polymer segments. Thus, the thermal transitional behavior reflects that of the homopolymers at comparable molecular weights. The difference in modulus-temperature behavior between homogenous and heterogenous copolymer systems with varied compositions is illustrated schematically in Figure

2-17. The temperature at which the modulus curve drops corresponds to the Tg of the material. As evident from the schematic, the glass transition temperature of random or miscible block copolymers is determined by the copolymer composition. For the phase separated materials, the copolymer composition and degree of phase separation govern the shape of the modulus curve. Phase separated block copolymers possess two glass transition temperatures corresponding to each homopolymer. If phase mixing of one segment occurs there will be an observable shift in the Tg of that phase toward the Tg of the other phase. The magnitude of the shift correlates with the level of phase mixing. Multiblock copolymers are synthesized by condensation or step growth polymerization techniques. The two general methods for synthesizing block copolymers are referred to as the “one-prepolymer” and “two-prepolymer” methods.84 The “one-prepolymer” method yields randomly segmented block copolymers. The polymerization consists of a difunctional oligomer of known size and two difunctional monomers where the oligomer possesses the same endgroups as one of the monomers (Figure 2-18). In this method the oligomer size and composition control the average block length of the second block. This method is commonly used in the synthesis of commercially important polyurethane copolymers and polyimide siloxane copolymers.69 32

0/100 25/75 50/50 75/25 100/0 A/B A/B A/B A/B A/B

Log modulus

TgB Temperature (°C) TgA

100/0 A/B 25/75 A/B

0/100 50/50 A/B A/B Log modulus 25/75 A/B

TgB Temperature (°C) TgA

Figure 2- 17. Generalized modulus vs. temperature curves for (a) random copolymers or miscible block copolymers and (b) phase separated block copolymers.68 33

The “two-prepolymer” method involves polymerization of two different oligomers via reaction of their end groups to produce perfectly alternating segmented copolymers (Figure 2-18). In this method the average block length of each segment is known before copolymerization. There are several examples in the literature where the properties of perfectly alternating segmented copolymers are compared to randomly segmented copolymers.85-57 It has been observed that perfectly alternating segmented copolymers have a higher structural regularity which affords more defined morphologies and better phase separation. Therefore, these materials may yield higher Tg’s and improved tensile strength relative to randomly segmented copolymers

2.4.2 Formation of Poly(siloxane imide) Copolymers

Poly(siloxane imide) copolymers have been reported as early as 1966 when pyromellitic dianhydride (PMDA) was reacted with an amine terminated siloxane dimer to form the poly(amic acid) precursor followed by curing to yield the imide.88 Gerber89 prepared copolymers with siloxane segments containing up to three repeat units. Pioneering efforts by St.Clair et al.90,91 and McGrath et al.92,93 propelled these materials to the forefront of polymer synthesis. Several approaches for preparing these materials have emerged. McGrath et al. discussed the recent developments in this field in several review articles.94,95

2.4.2.1 Randomly Segmented Copolymers

The most commonly used method for the preparation of poly(siloxane imide) copolymers involves the reaction of a preformed aminopropyl terminated 34

PDMS oligomer (as discussed in the previous section) with a dianhydride and diamine monomer. The resultant poly(siloxane amic acid) copolymer is then

Randomly Segmented Copolymers

Y Y + Y X + Y Y

Y YX X Y Y X XY Y n m

Perfectly Alternating Segmented Copolymers

X X + Y Y

n Figure 2- 18. Synthesis of randomly and perfectly alternating segmented copolymers 35 cyclodehydrated by either thermal or solution imidization methods as described in section 2.2.1.1.2. This approach has been utilized by several researchers to yield randomly segmented copolymers.104,105 Figure 2-19 presents a general synthetic route. Maintaining solubility of the reactants throughout the polymerization is a major factor in obtaining high molecular weight, tough, transparent copolymers of predictable compositions. The preparation of the copolymers often requires the use of a co-solvent to overcome the differing polarities of the polyimide and polysiloxane structures. This is necessary to maintain a homogenous solution throughout the polymerization. Typically aprotic polar solvents such as dimethylacetamide (DMAc) or N-methylpyrrolidone (NMP) are used in polyimide homopolymer synthesis (section 2.2.1). Under these conditions the polysiloxane oligomers may macrophase separate during copolymerization. This was demonstrated in an investigation by Ha and coworkers.101 Tetrahydrofuran (THF) was used as a reaction solvent for the copolymerization of PDMS, oxydianiline (ODA), and trimellitic anhydride chloride. The authors reported that as the reaction proceeded the PDMS segments showed good solubility in THF, while the ODA segments did not, resulting in a rapid increase in the viscosity and a translucent solution. With the addition of NMP to the system the reaction mixture became homogenous. The NMP was thought to dissolve the ODA segments during the polymerization. Co-solvent systems of NMP or DMAc have been used in combination with THF to yield homogenous reaction conditions. The NMP or DMAc prevents precipitation of the forming poly(amic acid) while THF helps solvate the polysiloxane oligomer. Arnold et al.97 reported that a co-solvent ratio higher in THF was needed as the PDMS oligomer size and content increased. As a consequence of larger THF volumes, the solubility of the forming polymer decreased. Therefore, this method limits the amount of PDMS incorporated into the polyimide backbone. 36

O O H2N NH2 R O O +

O O

CH3 CH3

H2N CH2 Si O Si CH2 NH2

CH3 CH3

Poly(siloxane amic acid) Copolymer

O O O O CH3 CH3 R R CH2 Si O Si CH2 N N N N 3 n 3 CH3 CH3 O O O O x y

Figure 2- 19. General synthetic route of randomly segmented polyimide- polydimethylsiloxane copolymers. 37

As an alternative, the PDMS oligomers can be pre-capped with the aromatic dianhydride.96,99,100,107 Advantageously, the capped siloxanes possessed greater solubility in the reaction solvent(s). Utilizing this method, Arnold et al.97 reported the incorporation of PDMS oligomers with s ranging from 800 - 10,000 g/mol into the benzophenone tetracarboxylic dianhydride (BTDA) / 3,3’- diaminophenyl sulfone (DDS) system in amounts ranging from 5 – 70 weight %. Wescott et al.98 demonstrated that polydimethylsiloxane-triphenyl phosphine oxide containing polyimide segmented copolymers of any composition could be made with an o-dichlorobenzene (oDCB) / NMP co-solvent system by first pre- capping the PDMS oligomer. In addition to increased solubility, slowly pre- capping the PDMS oligomer prevented premature chain extension. In the aforementioned approach, the hard (or imide) segment is synthesized in the presence of the difunctional PDMS oligomer, and consequently, the segment length of the polyimide is a strong function of both the oligomer molecular weight and composition.69 This can alter the properties of the resultant copolymer as well as control the phase purity between the blocks. Fitzgerald et al.99 reported that the decrease in upper glass transition temperatures of diaminophenylindane (DAPI) / 2,2-bis(4-phthalic anhydride)hexafluoroisopropylidene (6FDA) polydimethylsiloxane segmented copolymers was due to a reduction in the imide block molecular weight. To illustrate this, the authors examined the molecular weight dependence of the polyimide phase Tg. Figure 2-20 shows a plot of the polyimide phase Tg versus

1/Xn, where Xn is the number average degree of polymerization of the polyimide blocks calculated from: 1+ r = x n 1- r

(3) r = N A N A + 2 N B where NA and NB represent the moles of dianhydride and moles of the siloxane diamine, respectively. 38

350

325

(°C) 300

g

T

275

250 0.00 0.02 0.04 0.06 0.08 0.10

1/xn Figure 2- 20. Plot of the polyimide phase glass transition temperature taken ’ l from E”( ) and tan • ( ) as a function of 1/xn, the reciprocal of the number average degree of polymerization of the polyimide segments calculated from equation (3).99 39

2.4.2.2 Perfectly Alternating Segmented Copolymers

Perfectly alternating copolymers result from the reaction of polyimide and polysiloxane oligomeric pre-polymers. This was demonstrated by Smith.107 The author prepared norbornene anhydride terminated PDMS oligomers as described by Keohan and Hallgren.104 Reaction of the anhydride terminated PDMS oligomers with amino-terminated polyimides resulted in perfectly alternating structures. The copolymers were prepared in refluxing o-dichlorobenzene using 2-hydroxypyridine as a catalyst. This synthetic method was extended to include a series of engineering thermoplastics (ETP) (Figure 2-21). An alternative method for the preparation of perfectly alternating segmented copolymers is via transimidization of a functionalized polyimide oligomer with the aminopropyl PDMS oligomer. Transimidization reactions to form high performance polyimides was pioneered by Takekoshi and coworkers.1,108 Figure 2-22 presents a general schematic of the imide-amine transimidization reaction. The researchers were successful in melt polymerizing N,N’-pyridine phthalimide monomers with appropriate diamine monomers in the presence of a transition metal catalyst. The extent of reaction was determined by the nucleophilicities of the incoming amine (nucleophile) relative to that of the leaving group. Aliphatic amine nucleophiles and aromatic amine leaving groups yielded rapid reaction rates. By contrast, aromatic amine nucleophiles exhibited much slower reaction rates and lower extents of reaction. Rogers et al.109-112 reacted polyimide oligomers capped with 2- aminopyridine or 2-aminopyrimidine with aminopropyl terminated PDMS oligomers. Copolymerization was carried out in chlorobenzene (with small amounts of chloroform or NMP to maintain solubility) and yielded high molecular weight, fully imidized, perfectly alternating segmented imide siloxane copolymers within 2 hours at temperatures of 110-115°C (Figure 2-23). Utilizing the same polyimide oligomer, Sysel et al.113,114 prepared arylamine containing perfectly alternating segmented poly(siloxane imide) copolymers from a,w-arylamine 40

O O H2N PES NH2 CH3 CH3 or Si O Si H2N PEKK NH2 O O + or CH CH 3 3 H N PI NH x 2 2 O O or H2N PEPO NH2

Dichlorobenzene 2-Hydroxypyridine 170-175°C, 24 hours

O O CH3 CH3 Si O Si N N ETP n CH3 CH3 x O O m

Figure 2- 21. Synthesis of perfectly alternating segmented engineering thermoplastics (ETP)-PDMS copolymers.107 41

O O O NH R N R + H N R' N R' + H2N R 2 NH R'

O O O Figure 2- 22. General schematic of the imide-amine transimidization reaction. 42

......

O O O O N X X N N N Y N N N N n O O O O

Chlorobenzene CH3 CH3 110 - 115°C N2N CH2 Si O Si CH2 NH2 3 x 3 5 hours CH3 CH3

O O O O CH CH X X 3 3 N N N Y N CH2 Si O Si CH2 3 x 3 n CH3 CH3 O O O O

Figure 2- 23. Synthesis of perfectly alternating segmented poly(siloxane imide) copolymers via transimidization using a 2-aminopyrimidine terminated polyimide oligomer.1091-111 43 terminated PDMS and a,w-(3-aminophenoxy) terminated poly[oxy(dimethylsilyl)- 1,4-phenylene(dimethylsilylene)] oligomers. For randomly segmented copolymers, the length of the imide block (hard segment) is a function of both the oligomer molecular weight and composition.

This could have an effect on the imide phase Tg, as outlined earlier. For perfectly alternating copolymers the hard segment is independent of the PDMS average molecular weight and composition since the polymerization consists of reacting two preformed oligomers. Rogers et al.109 investigated the thermal properties of perfectly alternating and randomly segmented copolymers. The study revealed virtually no difference between the Tg’s of copolymers with corresponding compositions. Thus, the author concluded that the molecular weight of the imide segments was similar in the two materials.

2.4.2.3 Amic Ester Route to Segmented Copolymers

The amic ester route was developed at IBM laboratories as an alternative route to segmented poly(siloxane imide) copolymers (Figure 2-24).114-117 This route consisted of preparing a poly(amic-alkyl ester) as opposed to the poly(amic acid) intermediate. As outlined previously, the carboxylic proton has been a source of hydrolytic instability for the amic acid intermediate. Production of the amic ester intermediate yields a hydrolytically stable as well as a more soluble intermediate. This allowed the authors greater synthetic flexibility. Furthermore, the precursors could be isolated, characterized, and purified prior to imidization unlike the hydrolytically unstable poly(amic acid) analogues. 44

O O EtO EtO

+ H2N O NH2 EtO EtO O O +

CH3 CH3 N N H N CH Si O Si CH NH NH 2 2 3 x 2 3 2 H2N O O 2 CH3 CH3 N N

Poly(amic-ester)

O O O O CH3 N N N N O N N O O Si O x CH O O n O O N N 3 y

Figure 2- 24. Synthesis of a segmented copolymer via an amic ester precursor.114 45

2.5 CHARACTERIZATION OF PDMS CONTAINING BLOCK COPOLYMERS

2.5.1 Siloxane Surfaces

The copolymerization and blending of with organic polymers are of particular interest since small bulk concentrations of PDMS oligomers can result in rather dramatic surface enrichment.118-120 The unique surface behavior rendered by the addition of the low surface energy PDMS component is a direct result of the structural properties of the polymer. Table 2-2 contains the qualitative surface properties of PDMS. The combined effect of the low intermolecular forces between the methyl groups and the flexibility of the siloxane backbone have been have been used as the basis for understanding the surface properties.121 From a surface tension viewpoint, the more flexible the backbone the more readily the lowest surface energy configuration will be adopted. PDMS is arranged in a flat helical chain conformation (Figure 2-25). The backbone flexibility is related to the long Si-O (0.164 nm) and Si-C (0.190 nm) bonds. This allows the chain to adopt various polymer configurations resulting in lower steric interactions when compared to those occurring in carbon chains. The bond angles of 143° (Si-O-Si) and 110° (O-Si-O) yield an open structure and are the basic reason for the low barrier to rotation about the Si-O bond. Rotation about the siloxane bond is virtually free, with the energy being almost zero. Langmuir conceived the idea that separate surface energies exist for each of the different parts of complex molecules.122 He reported that the surface energy of the molecule was determined by the composition and orientation of the outermost groups independently of the underlying components. The siloxane backbone flexibility allows the methyl groups to be positioned in the surface. Thus, the low surface tension of PDMS is attributed to the low intermolecular forces of the pendant methyl groups on the surface of the molecule. Since the surface tension of PDMS is very low (» 20 mN/m)118,123 when compared to most organic polymers, the siloxane segments are expected to migrate to the more hydrophobic top (air) surface to form a silicon enriched layer 46

Table 2- 2. Qualitative surface properties of PDMS121

¨ Low surface tension ¨ Moderate interfacial tension against water ¨ High water repellency (hydrophobic) ¨ No surface shear viscosity ¨ Good lubrication of plastics and elastomers ¨ Good wetting, spreading and flow-out aid ¨ Soft feel 47

O O Si O O Si Me Si Me Me Me Me Me

Figure 2- 25. Schematic of PDMS structures depicting a flat helical chain conformation. 48

(Figure 2-26). Surface energetics and molecular mobility of the PDMS component govern the surface/bulk compositional differences observed. This yields such properties as reduced friction, improved gloss and feel, aid in release from molds, and increased electrical properties. Surface modification with PDMS was first investigated in 1964 by Zisman and co-workers.124 Block and graft copolymers have become an important field for the incorporation of PDMS molecules. This area was pioneered by Clark et al.23 with the surface characterization of AB block polymers of PDMS and polystyrene. The PDMS surface modification has since been consistently shown in a wide variety of PDMS containing block copolymers, including polystyrene- PDMS diblock and triblock copolymers126-127, poly(•-methylstyrene)-PDMS multiblock and starblock copolymers128, polysiloxane-polycarbonate block copolymers129,130, polyurethane-polysiloxane graft copolymers131,132, and poly(siloxane imide) block copolymers133,120 among others. Water contact angle analyses of PDMS containing block copolymer films has been used to confirm the presence of the siloxane enriched surface. This method of analysis yields information about the relative surface free energy of the material. A high water contact angle is indicative of a low energy hydrophobic surface. PDMS has a reported water contact angle of approximately 101°. Phase separated PDMS containing block copolymer films show water contact angle values very close to that of pure PDMS, thus indicating a siloxane enriched surface. Pike et al.132 demonstrated this surface phenomenon using dynamic contact angle analyses on segmented block copolymers of dimethylsiloxane- urea-urethanes. The surface of the block copolymers was initially hydrophobic indicating the PDMS enriched surface. However, it was found that with prolonged exposure to water the polymer surface rearranged to become more hydrophilic indicating the presence of the urea-urethane segments. The authors reported that the surface rearrangement was dependent upon the molecular weight of the soft block and sample history. The most likely mechanism was reorientation of the surface hydrophobic pendent PDMS methyl groups away 49

surface Silicon Enriched PDMS Surface Phase

bulk

Figure 2- 26. Schematic of the near surface region of a PDMS containing block copolymer demonstrating a silicon enriched layer.127 50 from the hydrophilic water surface. The backbone flexibility aided in reorientation when the environment of the polymer changed.134 Furukawa and coworkers investigated the surface properties of poly(siloxane imide) block copolymers.133 The study reveled that with the addition of 10 weight % PDMS ( » 750 g/mol) the contact angle increased by 25° over the homopolymer. Additional incorporation of PDMS did not contribute to an increase in water contact angle. Further investigation indicated that the surface tension of the block copolymers was almost identical with that of the PDMS homopolymer. In addition to water contact angle analyses, PDMS block copolymer surfaces have been previously studied by ion scattering spectroscopy (ISS)135,136, attenuated total reflectance Fourier transform infrared spectroscopy (ATR- FTIR)137,138, transmission electron spectroscopy (TEM)139,140, angle-dependent x-ray photoelectron spectroscopy (XPS)120, 141-145, and more recently atomic force microscopy131. For the scope of this dissertation the discussion will be limited to XPS and AFM techniques for the analysis of PDMS containing block copolymers.

2.5.2 Angle-Dependent X-Ray Photoelectron Spectroscopy (XPS)

2.5.2.1 Principles of Operation

XPS is one of the most widely used analytical techniques for studying polymer surfaces. This section will briefly discuss the main concepts of the technique, as it has been extensively covered elsewhere.144,146 Sample analysis using XPS is accomplished by irradiating a sample with monoenergetic soft x- rays (Mg Ka or Al Ka). The x-rays interact with atoms on the surface causing electrons to be emitted by the photoelectron effect. The energy of the detected electrons can be analyzed. Photoionization normally leads to two emitted electrons, a photoelectron and an Aüger electron (Figure 2-27). The Aüger electron is emitted due to 51

L2,3 or 2p

L1 or 2s Photoelectron Photon

K or 1s Auger Electron

L2,3 or 2p

L1 or 2s

K or 1s

Figure 2- 27. Diagram of photoelectron process (top) and Auger process (bottom).147 52 relaxation of the energetic ions left after photoemission. In the Aüger process, an outer electron falls into the inner orbital vacancy, and a second electron is emitted, carrying off the excess energy. It possesses kinetic energy equal to the difference between the energy of the initial ion and the doubly negatively charged final ion, and is independent of the mode of initial ionization.147 The emitted electrons have a measured kinetic energy given by KE = hn - BE - f (4) where KE is the kinetic energy of an emitted photoelectron from a certain atom, BE is the binding energy of the atomic orbital from which the electron originates, h• is the energy of the x-ray sources, and f is the spectrometer work function. Angle-dependent XPS gives compositional information as a function of depth, by varying the angle between sample surface and analyzer. This effectively limits the escape of electrons of a certain kinetic energy from a depth defined by the exit angle. Figure 2-28 shows an illustration of the measurement setup. The relationship between the sampling depth (d) and the take-off angle (q) is given by d = 3l sin(q) (5) where • is the mean free path of emitted electrons.148 Although the x-rays may penetrate many thousands of angstroms into the sample, the mean free paths of the photoemitted electrons are rather short and it is therefore only those from the outermost layers of the material which escape without energy loss. The mean free path of emitted electrons is dependent on the sample material and the kinetic energy of the photoejected electrons. At a 90° take-off angle, the relationship between sampling depth (3l) of emitted photoelectrons is defined by l(Å) = 490/KE2 + 1.1KE1/2 (6) where KE is the kinetic energy of the emitted photoelectron.149 The emitted photoelectrons from different depths do not equally contribute to the XPS signal measured. This signal attenuation effect exists because, the emitted photoelectrons (energy range of 50-1000eV) typically will travel only 20- 100 Å before they lose energy from inelastic scattering events and cannot 53

e- to analyzer e- to analyzer

x-ray x-ray d q q d

d = 3l sinq

Figure 2- 28. An illustration of the angle-dependent XPS setup. The sampling depth (d) is related to the take-off angle (q) by the relationship d=3lsinq, where l is the mean free path of emitted electrons.148 54 contribute to the photoelectron peak characteristic for a given element. At a particular take-off angle q, about 65% of the total emitted photolectrons originated within 1l sinq of the surface, 86% originated within 2••sin•, and about 95% within 3• sin• of the surface region, 149--151

2.5.2.2 Application of Angle-dependent XPS to Siloxane Containing Block Copolymers for Surface Characterization

Angle-dependent XPS has proven to be an efficient method for analyzing the surface migration of PDMS containing block copolymers. Much of the success of the technique can be attributed to its sensitivity to the near-surface region and the information obtained about the elemental composition and chemical-state of the material. Peak intensities can also be used for semiquantitative analysis of the surface region. Clark et al.125 used XPS in conjunction with contact angle measurements to study AB block copolymers of PDMS and polystyrene. The XPS results indicated that the surface region was comprised of an overlayer of PDMS ranging from approximately 13 to greater than 40 Å, depending on the solvents used to cast the film. Dwight and McGrath120,130 have also shown surface enrichment of the lower surface energy component in block copolymers of bisphenol A polycarbonate and PDMS. When low siloxane concentrations (0.01-10 weight %) were examined, a “critical” siloxane concentration of 1% bulk was found to cause the surface to be siloxane–dominated. Also, over a narrow concentration range near 50% siloxane, the surface was found to be nearly pure siloxane. Chen et al.129 investigated the surface effects of different block lengths and bulk compositions using multiblock copolymers of bisphenol A polycarbonate (BPAC) and PDMS structures. The study revealed that all the block lengths exhibited PDMS surface segregation in the surface region. However, the multiblock copolymers with longer PDMS blocks or with higher PDMS bulk 55 concentrations had higher silicon surface contents. The block copolymer structure and film preparation process determined the surface composition of the block copolymers. Furthermore, the authors found that annealing the films at elevated temperatures increased the amount of PDMS component observed in the upper surface region. Blending of PDMS containing block copolymers with organic polymers can also have a dramatic effect on the surface properties. Chen and Gardella152 investigated the surface properties of blends prepared with PDMS block copolymers and homopolymers of polystyrene, poly(•-methylstyrene), and bisphenol A polycarbonate. The study showed that when PDMS (from the block copolymer) was incorporated into the blends at 6% or less by weight, the surface PDMS concentrations of the blends would be as high as 95%. The PDMS surface concentrations did not change substantially when the bulk concentration of the block copolymer in the blend increased form 10 to 100%. The authors concluded that the amount of the block copolymer in the blend, the molecular weight of the homopolymer, and the block copolymer architecture were contributing factors that affected the surface compositions of the blends.

2.5.3 Atomic Force Microscopy (AFM)

2.5.3.1 Principles of Operation

AFM has the ability to create 3-dimensional micrographs with resolution down to the nanometer and Angstrom scales. This has made it a desirable tool for imaging surfaces. In addition to its imaging abilities, the AFM can also be employed to probe spatial variations in surface properties, such as stiffness, adhesiveness and viscoelasticity.153-155 AFM has been used in high-resolution profiling of the surface morphology and nanostructure of polymeric materials for such applications as engineering plastics, paints, coatings, packaging, fibers, and other consumer goods. 56

Figure 2-29 shows the AFM components and analysis setup. When in operating mode, the cantilever tip is scanned (or moved) over a sample surface. The feedback system enables the piezo-electric scanner to maintain the tip at a constant force or height. Maintaining the tip at a constant force enables one to obtain surface height information, whereas operating at a constant height gives force information about the surface of the sample. The forces detected between the tip and the sample surface cause the cantilever to bend. This bending is measured by a detector. A computer that drives the piezoscanner analyzes the data and converts them into an image. A map of the surface topography of the sample is produced therefrom.156,157 Although several scanning techniques are available to measure the surface topography, there are three main operating modes: Tapping Mode™ Contact Mode, and Non-Contact Mode.158 Of the three, the Tapping Mode™ is the most powerful, and provides high resolution without inducing destructive frictional forces to the sample. When the Tapping Mode™ is utilized, the cantilever is oscillated at its resonant frequency (Figure 2-30). As the cantilever is brought closer to the sample, the tip is allowed to 'tap' or have contact with the sample. By placing the tip in contact with the surface followed by lifting the tip off the surface, high resolution can be achieved while avoiding dragging. Due to the short contact time and oscillation amplitude, the likelihood of damaging either the tip or the sample, compared with the contact mode, is greatly reduced. As an extension of the Tapping Mode™, phase imaging is a technique from which qualitative information of the microstructure of the surface of a material can be obtained on the nanometer scale. The phase image is obtained from the phase lag, which is very sensitive to variations in material properties such as adhesion and viscoelasticity. Regions of a microstructure that exhibit inconsistent mechanical properties such as friction, elastic modulus, composition, or viscoelasticity are displayed in the resulting image as regions of differing contrast.159,160 57

Figure 2- 29. Schematic of AFM showing the four main parts: (1) the cantilever, (2) the piezoscanner, (3) the force sensor (includes the laser diode, mirror, and photodetector), and (4) the feedback system.158 58

Figure 2- 30. Diagram of Tapping ModeÔ AFM158 59

AFM also has the ability to distinguish components in heterogeneous samples. This ability, referred to as compositional mapping, relies on the differences in adhesive or mechanical properties of the different components of a heterogeneous sample. These differences influence the energy dissipated from the tip-sample contact, thus affecting the phase contrast when the sample is examined in the Tapping Mode™. This allows characterization of complex samples such as blends, block copolymers, and composites. The phase images obtained exhibit higher sensitivity to material properties thus allowing easy visualization of the different components. This becomes apparent when characterizing the near surface structure. The near surface phenomenon occurs because some samples often contain surface layers that may differ from their bulk composition. This may be due to different affinities to air. Hence, by tapping lightly only the topmost surface would be imaged whereas tapping harder may reveal the underlying structures. This phenomenon has been seen when imaging microphase separated block copolymers.161

2.5.3.2 Application of AFM to Block Copolymer Surface Characterization

The application of AFM to polymer surfaces was first done in 1988, shortly after its invention.162 Block copolymers have been extensively studied using AFM due to their unique properties.163-168 Grim and coworkers161 presented an interesting AFM study on polystyrene and poly-2-vinylpyridine diblock copolymers. The polymers were spun cast onto silicon wafers. Upon drying, the surface was imaged then annealed and re-imaged in a step-like manner. The authors were able to monitor the evolution of the surface morphology. They reported an initial formation of domains. During the subsequent growth stage three phenomena were observed: growth of individual domains, coalescence of neighboring domains and dissolution of the domains. After an initial increase 60 during the formation stage, the percentage of the top surface covered with islands remained constant. Berg and co-workers169 demonstrated the utilization of AFM for characterizing surface segregation of the low surface energy component of a block copolymer. The near surface phenomenon (discussed in the previous section) was seen when imaging microphase separated polystyrene-b- polybutadiene-b-polystyrene triblock copolymers (PS-PB-PS). The poly- butadiene block has a lower surface energy than the polystyrene block; thus, it dominates the copolymer/air interface as was the case in the previously discussed PDMS containing block copolymers. Light tapping on the copolymer film surface revealed a smooth film topology in the height image (Figure 2-31a) and no pronounced contrast variations in the phase image (Figure 2-31b). However, upon increasing the tapping force the height and phase image (Figure 2-31c and 2-31d) revealed a “worm-like” pattern which was consistent with the bulk morphology observed with TEM. Thus, the authors concluded that at light tapping the height images revealed the surface topography of the butadiene-rich topmost layer. Since the polybutadiene is above its glass transition temperature at ambient conditions, the AFM tip can deform it. Therefore, harder tapping reveals the more rigid underlying structure of the microphase separated triblock copolymer. Furukawa and coworkers133 investigated the surface topography of microphase separated poly(siloxane imide) block copolymers. It was found that the surface topography was greatly influenced by the composition and molecular weight of the PDMS segments because of their surface enrichment. The study revealed a relationship between the film surface and block copolymer composition. It was noted that the surface of the polyimide homopolymer film was flat and smooth, while the surfaces of the block copolymers differed from each other in terms of roughness. The polysiloxane segments segregated from the polyimide phase to form larger domains called “islands” in the “sea-island” structure. This surface roughness reached a maximum and gradually decreased with increasing PDMS concentration. The authors concluded that at small PDMS 61

(a) (b)

(c) (d)

Figure 2- 31. AFM Height (a & c) and Phase (b & d) Image of PS-PB-PS Triblock Copolymer with light and hard tapping, respectively.169 62 compositions the domains were small, but as the PDMS composition was increased the domains became larger and were readily integrated in the surface. This eventually formed a continuous phase in the surface. At that point the surface became smooth and flat as with the polyimide homopolymer.

2.5.4 Molecular Weight Characterization

Several methods can be used to determine the molecular weights of poly(siloxane imide) block copolymers. Traditionally, the relative molecular weights have been determined by intrinsic viscosity. Arnold et al.97 confirmed the synthesis of high molecular weight BTDA-DDS based poly(siloxane imide) segmented copolymers using intrinsic viscosity measurements. The authors reported values ranging from 0.50 to 1.88 dLg-1. Similar results were determined by Rogers et al.109 for perfectly alternating poly(siloxane imide) block copolymers. Gel permeation chromatography (GPC) has been used for the determination of molecular weight and molecular weight distributions. Szesztay and Ghadir170 used ultra-styragel columns calibrated with polystyrene to determine the approximate molar masses of poly(siloxane imide) copolymers. This proved to be an efficient method for copolymers with molecular weights higher than 1000 g/mol. Konas et al.171,172 established a universal calibration for soluble polyimide homopolymers. This method makes the assumption that the copolymer refractive index is the same at all molecular weights. The authors obtained normal retention behavior of the polyimides using chloroform, THF, and

NMP containing LiBr, or in NMP stirred over P2O5 before use. The addition of additives eliminated the absorption affects of the polyimide solutions. Facinelli et al.35 utilized this technique for the determination of absolute molecular weights and molecular weight distributions for controlled molecular weight poly(amic acid)s and polyimides. End-group analyses have been utilized by a number of researchers for the molecular weight determination of the PDMS and polyimide oligomers. The

s of aminopropyl terminated PDMS have been determined by several 63 researchers using NMR and potentiometric titration with HCl.110,173 NMR techniques have also been employed for the estimation of number-average molecular weights for end functionalized polyimide homopolymers up to 40,000

110,94 Mn.

2.5.5 Copolymer Composition

NMR techniques have been widely used for the determination of the composition for soluble copolymer systems. Rogers and coworkers110 observed quantitative incorporation of PDMS oligomers into poly(siloxane imide) copolymer systems. Dry NMR solvents are used for the analysis. It was found that the presence of water caused micellization of the block copolymers. The more hydrophilic polyimide segments surrounded the hydrophobic PDMS aggregates. This caused an inaccurate measurement by yielding lower PDMS content values.109

2.5.6 Thermal Behavior

The thermal stability of the poly(siloxane imide) block copolymers is usually determined by dynamic thermogravimetric analysis and is often reported as 5% weight loss temperatures. Poly(siloxane imide) block copolymers exhibit slightly lower 5 weight % loss temperatures than that of the polyimide homopolymer. This is due to the fact that the PDMS blocks are less thermally stable than polyimides.97 Table 2-3 contains the 5 weight % loss temperatures of a series of ODPA-Bis P based poly(siloxane imide) block copolymers.109 The thermal stability of the copolymers decreases slightly with increasing siloxane content. The char yield in air at 700° C increases with the incorporation of siloxane moieties. This has been attributed to the retardation of cyclic generation and formation of organosilicates at high temperature in air.97 The formation of the ceramic like silicate char forms a protective barrier, thus preventing further 64

Table 2- 3. The influence of copolymer composition on the thermal stability of poly(siloxane imide) block copolymers in air.109

Polyimide Oligomer Mn PDMS Mn (kg/mol) Weight % PDMS 5 weight % loss by Char (kg/mol) TGA (C°) Yield 700°C (wt %) ODPA-Bis P-PA 40,000 ------550 9 ODPA-Bis P-PA 4.1 1.09 21 461 16 ODPA-Bis P-PA 4.1 2.55 37 467 37 ODPA-Bis P-PA 4.1 4.5 52 428 11 ODPA-Bis P-PA 11.0 1.09 9 499 11 ODPA-Bis P-PA 11.0 2.55 18 491 20 ODPA-Bis P-PA 11.0 4.5 28 441 20 65 erosion of the bulk material. This has been shown to interrupt the burning process with a limited production of smoke. Microphase separated poly(siloxane imide) block copolymers exhibit a two phase morphology, thus the glass transition temperatures of both components can be observed. The upper Tg corresponds to the glass transition temperature 99 of the PDMS phase, which occurs between –115 and –123°C. The upper Tg corresponds to the glass transition temperature of the polyimide phase. Differential scanning calorimetry and dynamic mechanical behavior have been used to evaluate these transitions.

As outlined in section 2.4.1.1, the polyimide phase Tg of randomly segmented poly(siloxane imide) copolymers may be considerably depressed relative to the value expected for high molecular weight homopolymer. It has been found that the size of the polyimide block significantly influences its Tg. As 99 the length of the polyimide block decreases the upper Tg decreases. This was later addressed using perfectly alternating block copolymers by Rogers et. al109 as described in section 2.4.1.2. The authors found that the resulting transition temperatures were a strong function of the original number average segment lengths.

The depression in the polyimide phase Tg has also been attributed to partial phase mixing at the siloxane-polyimide block interface. Investigations by Arnold and coworkers97 revealed that block copolymers prepared with low molecular weight polysiloxanes ( = 900 g/mol) exhibited phase mixing. It was believed that the polysiloxane phases mixed with the polyimide phases.

2.5.7 Morphology Microphase separation is expected to occur in poly(siloxane imide) block copolymers due to the large difference in solubility parameter of the siloxane and imide segments. The phase separated materials have distinct morphologies that are dependent on the block size, copolymer composition, and chemical composition of the polyimide phase.140 These materials display morphologies 66 consistent with other block copolymer systems. TEM has proven to be an efficient method for bulk morphology analysis of phase separated block copolymers.99,120,128,140 Morphological investigations carried out by York140 for compression molded poly(siloxane imide) block copolymers revealed a semi-continuous or rod-like morphology for a 15 weight % siloxane sample. Solution cast films of the sample copolymer showed spherical siloxane in a continuous polyimide matrix. The author attributed the shift from the semi-continuous texture to the spherical texture to the casting solvent (NMP) acting as a preferential solvent. The polyimide chains are more soluble in the NMP. This swells the chains and favors a shift to the more discrete spherical structure. The author also observed spherical siloxane phases in a continuous polyimide matrix for compression molded copolymers prepared with a less polar polyimide and analogous copolymer compositions. This suggested that decreasing the polarity of the polyimide phase favors the more discrete microdomain structure with more surface to volume ratio. The PDMS domain size in spherical morphologies is related to its molecular weight according to equation (3). TEM analysis revealed the presence of discrete siloxane spheres approximately 2 nm in diameter for a BTDA-DDS based block copolymer containing 10 weight % of a 2,000 g/mol PDMS oligomer. Increasing the PDMS oligomer molecular weight to 10,000 g/mol yielded siloxane domains of about 10 nm in diameter. Furthermore, the PDMS spherical domains became more defined indicating a higher degree of phase separation with an increasing siloxane molecular weight.99 Samseth and coworkers174,175 showed that solution cast block copolymer films containing at least 40 weight % siloxane exhibited a siloxane continuous phase. Investigations by Rogers et al.109 revealed that block copolymers prepared with 20 weight % siloxane ( = 1000 g/mol) exhibit a very fine microphase separated structure. Increasing the siloxane concentration to 50 weight % ( = 4500 g/mol) yielded a co-continuous type morphology. 67

2.5.8 Mechanical Properties

The incorporation of PDMS into a polyimide backbone greatly impacts the mechanical behavior of the resultant block copolymer. As discussed in the previous section, block copolymers containing low PDMS weight percentages exhibit morphologies in which the polysiloxane exists as spheres in a continuous polyimide matrix. The polysiloxane serves to toughen the copolymer relative to the polyimide homopolymer with little effect on the tensile modulus or strength. Fracture toughness values indicated over a 50% increase with the addition of just

5 weight % PDMS ( = 4500 g/mol) for BTDA-MDA based block copolymers.176 The effects of morphology can also be observed in the stress-strain behavior of the block copolymer. As more polysiloxane is incorporated, the morphology changes until the siloxane becomes the continuous phase at approximately 40 weight % siloxane. At this point mechanical properties reflect the properties of the polysiloxane segment. The elongation to break was less than 100% with a failure stress of about 40 MPa for a BTDA-DDS based block copolymer containing 40% PDMS.

2.6 INTERPHASE APPLICATIONS

2.6.1 Interphases for Composite Applications

Fiber-reinforced composites are increasingly being used in structural applications because of their high strength to weight ratios and corrosion resistance. In general, the composites are prepared with theromset matrices, and the resultant composites are often brittle in nature. Thus, much attention has been devoted to toughening these composites. Several methods have been pursued which include (1) switching to high-performance, tougher thermoplastic 68 matrices177,179, (2) using rubber180,181 or thermoplastic182,183 toughened thermosetting matrices, (3) using interpenetrating network matrices that offer improved toughness and easy processing184-188, and (4) applying a tough interfacial interlayer or interphase onto the fibers before the matrix impregnation189-193. The latter approach offers several potential advantages which include: (a) absorption of crack propagation energy, (b) relieving stress concentration around the fiber (resulting from curing or an external load), and (c) healing fiber surface flaws and protecting the fibers from breakage. Using a tough, high Tg interphase material may provide improved composite toughness and better temperature stability.194

As thermally stable high Tg materials, polyimides and poly(siloxane imide) block copolymers are attractive candidates for composite interphase applications. It was noted that the incorporation of small amounts of PDMS greatly improved the toughness of the polyimide with little effect on the thermal stability or the upper glass transition temperature. Polyimide and poly(siloxane imide) block copolymers should lead to tougher fiber-matrix interphases regions therefore yielding tougher composites. The addition of such engineering thermoplastics has been shown to improve the toughness of thermosetting resins without significant sacrifice of properties.195 Bucknall and Gilbert196 obtained toughened epoxy resins with polyetherimide blends. The authors reported a linear relationship between the fracture toughness and the polyetherimide content up to 25 weight %. This was accompanied by only a small reduction in Young’s modulus. Investigations by Wimolkiatisak and Bell194 demonstrated that an electrocopolymerized 3-carboxyphenylmaleimide-styrene interphase greatly toughened AS4/Epon 828-MDA composites. The optimum interphase thickness was found to be about 0.16 mm. At this thickness the researchers observed 100% improvement in the double-cantilever beam (DCB) Mode I critical strain energy release rate and an improvement of 60% in the notched Izod impact 69 strength. Furthermore, the apparent interlaminar shear strength remained close to the value of the control composite. The chemistry of the polymer coating or interphase material (between the fiber and composite matrix) can be altered to form strong physical and/or chemical bonds with the matrix to improve adhesion between the fiber and matrix. The adhesion of carbon fibers to thermosetting polymers has been investigated by Bascom and coworkers.1971 Madhukar and Drzal198 showed that by altering fiber surface characteristics the mechanical performance of composite laminates was drastically altered. As depicted in Figure 2-32, the interphase region should contain a gradient composition of the polymer coating and matrix materials extending outward from the fiber surface. Judicious placement of the ductile polymer specifically at the interface may enhance composite toughening using only very small amounts of the modifier material. Typically, the interphase material would comprise less than 1 weight % of the total composite. Robertson and coworkers199 showed an improvement in composite durability from 20 ksi to 32 ksi using this approach with only 0.8 weight % thermoplastic sizing in carbon fiber reinforced vinyl ester composites. By constrast, 15-20% (based on matrix weight) of thermoplastic or rubber tougheners added to the matrix resin would be needed. Thus, by placing the modifier directly on the fiber surface, as opposed to dissolving it in the matrix resin, the viscosity of the matrix resin can be maintained low.

2.6.2 Interphases for Microelectronic Applications

As electronic devices become smaller, the structural demands on their components become larger. Polymers have become prime candidates as interlayer dielectrics and passivation layers for multichip packaging technologies. These materials offer such advantages as lower signal propagation delays and the possibility of higher wiring densities for interchip connections.200 Due to the 70

INTERPHASE REGION

THERMOSETTING MATRIX CARBON FIBER POLYMER COATING

Figure 2- 32. Model of an interphase region which could form by first applying a thermoplastic polymer coating onto a fiber surface, then adding a miscible thermosetting matrix resin and curing. The interphase region contains a gradient composition of the polymer coating and matrix material. 71 nature of fabrication processes for these devices, the polymers must meet fairly rigorous requirements (Table 2-4). When compared to the inorganic alternatives of alumina or silicon oxides, polymers have some obvious shortcomings with respect to withstanding severe thermal, chemical, and mechanical stresses. However, they can offer lower dielectric constants and ease of processing which may lead to an overall improvement in device performance. Polyimides have become the leading class of polymers for microelectronic applications.2,201 The preferred use of these materials as interlayer dielectrics for thin film wiring to interconnect chips, EC pads and other components on multichip modules can be attributed to their excellent thermal and electrical properties which satisfy the criteria for microelectronics fabrication and performance. Both semirigid and rigid aromatic polyimides have proven suitable for such applications. Even more attractive are the fluorine-containing polyimides.

Generally, fluorinated polyimide materials have lower dielectric constants and reduced moisture absorption relative to their non-fluorine containing counterparts.202-207 In addition, these materials show good processability and good optical properties. Thus, fluorinated polyimides have been investigated by several researchers for microelectronic applications.208-213 Auman208 showed dielectric constants between 2.5 and 2.8 for polyimides incorporating various fluorine containing moieties. These materials also exhibited less than 3% water absorption and low coefficient of thermal expansion (CTE) values. Common flexible chain polyimides derived from non-linear aromatic diamines and/or dianhydride precursors have a CTE typically in the range of 25 to 40ppm/°C. This is relatively high when compared to 2-3ppm/°C in the case of the commonly used inorganic materials and 6-25ppm/°C for commonly used metals. When these materials are placed in multilayer structures the difference in CTE between the polymer, metal, and ceramic causes thermal stresses to build-up in the interface region during fabrication. Excessive interfacial stress can cause film cracking and/or delamination resulting in device malfunction or 72 failure. Thus, many researchers have focused on improving the compatibility of the polyimide within this system.214,215 Table 2- 4. Required properties of polymers for microelectronic applications

Polymer Properties High Glass Transition Low Dielectric Constant (<3) Temperatures (>300°C)

High Degradation Good Processibility From Solution Temperatures (>350°C)

Good Mechanical Properties, High Good Adhesion to the Substrate Modulus (10% elongation)

Low CTE (5-9x10-5/°C) Low Moisture Absorption 73

Hedrick and coworkers114 proposed using microphase separated block copolymers as a means of promoting rapid stress relaxation in poly(pyromellitic dianhydride-oxydianiline) films to yield low stress materials. PDMS was chosen as a copolymer block to provide a large modulus mismatch with the polyimide. The results showed that a block copolymer containing 20 weight % PDMS of molecular weight 5400 gmol-1 had no residual thermal stress after curing at 350°C. The combination of fluorine containing polyimides with polydimethylsiloxanes appears to be a promising area for microelectronic applications. Arnold et al.97 demonstrated the influence of molecular structure on the dielectric behavior of polyimide homopolymers and poly(siloxane imide) segmented copolymers (Table 2-5). As expected, the fluorine containing structures yielded lower dielectric constants. A slight decrease was observed with the addition of just 10 weight % PDMS into the polyimide structure. 74

Table 2- 5. Influence of molecular structure on the dielectric behavior.84

Polyimide System Dielectric Constantb PMDA-ODAa (Kapton, ambient) 3.32 PMDA-ODAa (dry) 3.24 PMDA-Bis Pa 2.70 PMDA(50 mol%)/6F(50 mol%)-Bis Pa 2.62

BTDA-DDSa 3.17 BTDA-DDS-10% PSX 800 3.09 BTDA-Bis Pa 2.77 BTDA-Bis A 3.10

6F-DDSa 2.87 6F-Bis P 2.62 6F-Bis P-10% PSX 800a 2.61 6F-Bis P-10% PSX 800a 2.54 a Bulk imidized b Dielectric constant at 15 GHz 75

2.7 REFERENCES

1. Polyimides, ed. Wilson, D.; Stenzenberger, H. D.; Hergenrother, P. M. New York: Chapman and Hall, 1990. 2. Polyimides, ed. Mittal, K. L., vol. 1 and 2, New York: Plenum Press, 1984. 3. Feger, C.; Khojasteh, M. M.; McGrath, J. E. ed. Polyimides: Chemistry, Materials, and Characterization, Elsevier: 1989. 4. Sroog, C. E. Prog. Polym. Sci. 16 (1991): 561-694. 5. Abadie, M. J. M; Sillion, B. Polyimides and Other High-Temperature Polymers, Elsevier: 1991. 6. Takekoshi, T. In New Polymer Materials, Advances in Poly. Sci., vol. 94, ed. C. G. Overberger, Berlin: Springer-Verlag 1990, pp 1-25. 7. Volkesen, W. In High Performance Polymers; ed. Hergenrother, P. M., Berlin: Springer-Verlag 1994; Vol. 117; pp 11-164. 8. Bogert, T. M.; Renshaw, R. R. J. Am. Chem. Soc. 30 (1908): 1135. 9. Jones, J. I; Ochynski, F. W.; Rackley, R. A. Chem. Soc. 30 (1908): 1135. 10. Frost, L. W.; Kesse, J. J. Appl. Polym. Sci. 8 (1964): 1039. 11. Sroog, C. E.; Endrey, A. L.; Abramo, S. V.; Berr, C. E.; Edwards, W. M.; Olivier, K. L. J. Polym. Sci.: Part A (1963) 1: 3135. 12. Bower, G. M.; Frost, L. W. J. Polym. Sci.: Part A 1 (1963): 3135. 13. Adrova, N. A.; Bessonov, M. I.; Laius, L. A.; Rudakov, A. P. Polyimides: A New Class of Thermally Stable Polymers; Connecticut: Technomic., 1970. 14. Dine-Hart, R. A.; Wright, W. W. J. Appl. Polym. Sci. 11 (1967): 609-627. 15. Frost, L. W.; Kesse, J. J. Appl Polym. Sci. 8 (1964): 1039. 16. Svetlichnyi, V. M.; Kalnin’sh, K. K.; Kudryavtsev, V. V.; Koton, M. M. Dokl. Akad. Nauk SSSR (Eng. Trans.) (1977), 237, 693. 17. Kalnin’sh, K. K.; Solov’eva, G. I.; Belen’kii, B. G.; Kudryavtsev, V. V.; Koton, M. M. Dokl. Akad. Nauk SSSR (Eng. Trans.) 204 (1972): 473. 18. Hoclgkin, J. H. J. Polym. Sci.: Polym. Chem. 14 (1976): 409. 19. Synder, R. W.; Thomson, B.; Bartges, B.; Czerniwski, D.; Painter, P. C. Macromolecules 22 (1989): 4166. 76

20. Scroog, C. E. in Encyclopedia of Polymer Science and Technology, N. M. Bikales, Ed.; New York: Interscience, 1969; Vol. 11; pp 247. 21. Cassidy, P. C.; Fawcett, N. C. in Encyclopedia of Chemical Technology; New York: Wiley, 1982, pp 704. 22. Kruez, J. A.; Endrey, A. L.; Gay, F. P.; Scroog, C. E. J. Polym. Sci.: Part A 4 (1966): 2607. 23. Waldbauer, R. O.; Rogers, M. E.; Arnold, C. A.; York, G. A.; Kim, Y. J.; McGrath, J. E. Polym. Prep. 31 (1990): 432. 24. McGrath, J. E.; Rogers, M. E.; Arnold, C. A.; Kim, Y. J. Hedrick, J. C. Makromol. Chem., Macromol. Symp. 51 (1991): 103. 25. Rogers, M. E.; Woodard, H.; Brennan, A.; Cham, P. M.; Marand, H.; McGrath, J. E. Polym. Prep. 33(1) (1992): 461. 26. Kim, Y. J.; Glass, T. E.; Lyle, G. D.; McGrath, J. E. Macromolecules 26 (1993): 1344-1358. 27. Kailani, M. H.; Sung, C. S. P.; Haung, S. Macromolecules 25 (1992): 3751. 28. Angelo, R. J.; Golike, R. C.; Tatum, W. E.; Kruex, J. A. Mid-Hudson Chapter of Soc. Plas. Eng., (1987) : 67. 29. Moy, T. M.; DePorter, C. D.; McGrath, J. E. Polymer 34 (1993): 819. 30. Synthesis of Polyamic Ester by Reaction of Polyamic Acid Salts with Alkyl Halides; ed. Flaim, T. D.; Horter, B. L.; Mors, M. G., New York: Elsevier Science, 1988, pp 279-291. 31. Schmitt, K.; Gude, F.; Brandt, S. to United States 3,882,085, (1975). 32. Kojima, M.; Noda, Y.; Suzuki, Y.; Okamoto, T. to United States 3,925,313, (1975). 33. Endrey, A. L., to United States 3,242,136, (1966). 34. Reynolds, R. J. W.; Seddon, J. D. J. Plym. Sci.: Part C 23 (1668): 45-56. 35. Facinelli, J. V.; Gardner, S. L.; Dong, L.; Sensenich, C. L.; Davis, R. M.; Riffle, J. S., Macromolecules 29(23) (1996): 7343-7350. 36. Ladenburg, A., Ann., 164 (1872): 300. 37. Jacobson, H. and Stockmayer, W. H., J. Chem. Phys. 18 (1950): 1600. 38. Brown, J. F. and Slusarczuk, G. M. J., J. Amer. Chem. Soc. 87 (1965): 931. 77

39. Mark, J. E. Acc. Chem. Res. 18 (1985): 202. 40. Orrah, D. J., Ph.D. Thesis, Univ. of York, 1987. 41. Chojnowski, J.; Kazmierski, K.; Rubinsztajn, S.; Stanczyk, W., Makromol. Chem., 187 (1986): 2039. 42. Speier J. L.; Roth, C. A.; Ryan, R. W., J Org. Chem. 36 (1971): 3120. Burns, G. T.; Decker, G. T.; Roy, A. K. (1994) US Patent 5 290 901 to Dow Corning. 43. Vondracek, P. and Gent, A. N., J. Appl. Polymer Sci., 27 (1982): 4517. 44. Grubb, W. T., J. Amer. Chem. Soc., 76 (1954): 3408. 45. Lasocki, Z. and Chrzczonowicz, S., J. Polym. Sci, 59 (1962): 259. 46. Silanes, Surfaces and Interfaces, ed., Pohl, E. R. and Osterholtz, F. D. in Leyden, D. E., New York: Gordon and Breach, 1986, p 481. 47. Saam, J. C., in Zeigler, J. M., and Fearon, F. W. S. (eds.), Silicon Base Polymer Science, American Chemical Society, Washington D. C., 1990. P 71. 48. Szwarc, M., Carbanions, Living Polymers and Electron Transfer Process, New York : Wiley, 1968. 49. Advances in Polymer Chemistry, ed. Kireev, V. V., in Korshak, V. V., Moscow: Mir, 1986, p 199. 50. Wright, P. V. and Beevers, M. S., in Cyclics Polymers, ed., Semlyen, J. A. New York: Elsevier, 1986, p 85. 51. Yilgor, I. and McGrath, J. E., Adv. Polym. Sci., 86 (1988): 1. 52. Riffle J. S; Yilgor, I.; Tran, C.; Wilkes, G. L; McGrath, J. E.; and Banthia, A. K., “Elastomeric Polysiloxane Modifiers for Epoxy Networks: Synthesis of Functional Oligomers and Network Formations Studies in Epoxy Resin Chemistry II,” ed. Bauer, R. S., ACS Symposium Series # 221, ACS, Washington D.C., 1983. 53. Senger, C. L. Elsbernd, Synthetic, Kinetic and Characterization Studies on Siloxane Homopolymers and Segmented Copolymers, Ph.D. Dissertation, Virginia Tech, 1988. 78

54. Kogan, E. V., Ivanova, A. G., Reikhsfeld, V. O., Smirnov, N. I., and Grubber, V. N. in Siloxane Polymers, eds. Clarson, S. J. and Semlyen, J. A., New Jersey: PTR Printice Hall, 1993, p. 40. 55. Ishizuka, S. and Aihara, T., J. in Siloxane Polymers, ed. Clarson, S. J. and Semlyen, J. A., New Jersey: PTR Printice Hall, 1993, p. 40. 56. Chojnowski, J.; Rubinsztajn, S. and Wilczek, L., in Siloxane Polymers, ed. Clarson, S. J. and Semlyen, J. A., New Jersey: PTR Printice Hall, 1993, p. 40. 57. Tchernyshev, A. I. and Yastrebov, V. V. in Siloxane Polymers, Clarson, S. J. and Semlyen, J. A. (eds.), New Jersey: PTR Printice Hall, 1993, p. 40. 58. Chojnowskki, J., Mazurek, M., Scibiorek, M., and Wilczek, L., Makromol. Chem. 175 (1974): 3299. 59. Gilbert, A. K., Kantor, S. W., J. Polym Sci. 40 (1959): 35. 60. Kendrick, T. C., Parbhoo, B. M., and White, J. W., in Comprehensive Polymer Science, eds. Allen, G.; Bevington, J. C.; Eastmond, G. C.; Ledwith, A.; Russo, S.; and Sigwalt, P., Oxford: Pergamon, 1989, vol. 4, p 459. 61. Scott, D. W., J. Am. Chem. Soc. 68 (1946): 2294. 62. Hurd, D. T.; Osthoff, R. C.; and Corrin, M. L. J. Amer. Chem. Soc. 76 (1954): 249. 63. Boileu, S. in McGrath J. E. ed. Ring Opening Polymerization, American Chemical Society Symposium Series #286, Washington, D.C., 1985, p 23. 64. Kopylov, V. M.; Prikhodko, P. L.; and Koviazin, V. A. Vysokomol. Soed. A24 (1982): 1751. 65. Brown, J. F. and Slusarczuk, G. M. J., J. Poly. Sci., A2 (1964): 523. 66. Semlyen, J. A. and Wright, P. V., Polymer 10 (1969): 543. 67. Wright, P. V. Semlyen, J. A. Polymer 11 (1970): 462. 68. McGrath, J. E., Journal of Chemical Education 58 (1981): 914. 69. Noshay, A., McGrath, J.E. Block Copolymers: Overview and Critical Survey New York: Academic Press, 1977. 70. Meier, D. J., J. Polym. Sci., C, 26 (1969): 81. 79

71. Molau, G. E., in Block Polymers, S. L. Aggarwal, Ed., New York-London: Plenum Press, 1970, Chapter 1. 72. Quirk, R. P.; Kinning, D. J. and Fetters, L. J. in “Comprehensive Polymer Science,” 7(1) (1987). 73. Zelinski, R. Z. and Childers, C. W., Rubber Chem. Techn. 41(1) (1968): 161. 74. Legge, N. R.; Holden, G.; Schroeder, H. E., (eds.), “Thermoplastic Elastomers: A Comprehensive Review,” New York: Hamser Publishers, 1987. 75. Meier, D. J.; Hashimoto, T.; Shbayama, M. and Kawai, H., Macromolecules, 18 (1985): 1855. 76. Henderson, C.P. and Williams, M. C. Polymer 26 (1985): 2022. 77. Hefland, E. and Wasserman, Z. R, in Developments in Block Copolymers – 1, I Goodman (ed.), Applied Science Publishers, London (1982). 78. Encyclopedia of Polymer Science and Engineering, 2nd Edition, 9, New York: John Wiley and Sons, 1990, 760-784. 79. Torikai, N.; Takabayashi, N.; Noda, I.; Koizumi, S.; Moru, Y. Matasushita, Y., Macromolecules 30 (1997): 5698-5703. 80. Bates, F. S. and Hartney, M. A., Macromolecules 18 (1985): 2478. 81. Roovers, J. Methods of Exp. Phys. 16c (1988): 275. 82. McIntyre, D. and Campos-Lopez, in “Block Polymers,” ed. S. Aggarwal, New York: Plenum Press, 1970, pp. 19-30. 83. Matsuo, M. and Sagaye, S., in “Collodial Morphological Behavior of Block and Graft Copolymers,” ed. Movau, G. E., New York: Plenum Press, 1971, pp 1-19. 84. Odian, G. Principles of Polymer Science, 3rd Edition, John Wiley & Sons, Inc., New York, 1991, pp.145-149. 85. Kajiyama, M.; Kakimoto, M. and Imai, Y., Macromolecules, 23 (1990): 1244, 86. Oishi, Y.; Nakata, S.; Kajiyama, M.; Imai, Y., J. Polym. Sci.: Part A, 30 (1992): 2357 87. Oishi, Y.; Nakata, S.; Kakimoto, M.; Imai, Y., J. Polym. Sci.: Part A, 31(1993): 933 80

88. Kuckertz, V. H., Die. Makro. Chem., 98 (1966): 101. 89. Gerber, V. G., Die. Angew. Makro. Chem., (1968) 415, 212. 90. Maudgal, S.; Clair, St. Clair, T. L (1984) SAMPE Quartely 16(1), 6-12. 91. Maudgal, S.; Clair, St. Clair, T. L., Int J Adhes. (1984), 4(2), 87. 92. Yilgor, I; Yilgor, E; Johnson, B. J.; Eberle, J.; Wilkes, G. L.; McGrath, J. E. Polymer Preprints 24(2) (1983): 78. 93. Yilgor, I; Yilgor, E; Eberle, J.; Steckle, W Jr, Johnson, B. C.; Tyagi, D., Wilkes, G. L.; McGrath, J. E., Polymer Prints 24(1) (1983): 167. 94. McGrath, J. E.; Dunson, D. L.; Mechum, S. J.; and Hedrick, J. L., Adv. Polym. Sci., 140 (1999): 61. 95. Yilgor, I. And McGrath, J. E., “Polysiloxane Containing Copolymers: A Survey of Recent Developments,” Adv. Polym. 86 (1988): 1. 96. Johnson, B. C.; Yilgor, I.; McGrath, J. E., Polymer Preprints, 25(2) (1984): 54. 97. Arnold, C. A.; Summers, J. D.; Chen, Y. P.; Bott, R. H.; Chen, D. and McGrath, J. E., Polymer 30 (1989): 986. 98. Wescott, J. M.; Yoon, T. H.; Kiefer, L.; Rodrigues, D.; Konas, M.; Wilkes, G. L.; McGrath, J. E., Polymer Preprints, 34(1) (1993): 308. 99. Fritzgerald, J. J.; Tunney, S. E. and Landry, M. R., Polymer, 34(9) (1993): 1823. 100. Tian, S. B.; Park, Y. S.; Xu, G. J. Polym. Sci.: Part B 32 (1994): 2019. 101. Seong, Y. H.; Boo-Keun, O.; Young, M. L.; Polymer 36(18) (1995): 3549. 102. Fukakawa, N; Masatoshi, Y.; Kimura, Y., J. Polym. Sci.: Part A 36 (1998): 2237-2245. 103. Summers, J. D. and McGrath, J. E., Polymer Prep. 28(2) (1987): 230. 104. Keohan, F. L.; Hallgren, J. E., “Silicon Based Polymer Science,” Adv. Chem. Series 224 (1990): 165. 105. Nakata, S.; Kawata, M.; Kakimoto, M.; and Imai, Y., J. Polym. Sci.: Part A, 31 (1993): 3425-3432. 106. Arnold, C. A.; Summers, J. D.; McGrath, J. E., Polym. Eng. Sci, 29(20), (1989): 1413. 81

107. Smith, S. D., Synthesis and Characterization of Perfectly Alternating Segmented Copolymers Composed of PDMS and Engineering Thermoplastics, M.S. Thesis, VPI&SU (1991). 108. Takekoshi, T. and Kochanowski, E. J. US Patent 3,850,558 (1974) to General Electric. 109. Rogers, M. E., “Synthesis and Characterization of High Performance Polyimide Homo and Copolymers,” Ph.D. Dissertation, VPI&SU (1993). 110. Rogers, M. E.; Glass, T. E.; Mecham, S. J.; Rodrigues, d. Wilkes, G. L.; McGrath, J. E., J. of Polymer Sci. A: Poly Chem 32 (1994): 2663. 111. Rogers, M. E.; Rodrigues, D.; Brennan, A.; Wilkes, G. L.; McGrath, J. E. in Contemporary topics in Polymer Science, ed., Salamone, J. S and. Riffle, J. S., New York: Plenum Press 7 (1992): 47-55 112. Sysel, P.; Babu, J. R.; Konas, M.; Riffle, J. S.; McGrath, J. E., Polymer Preprints 33(2) (1992): 218. 113. Sysel, P. and Ouipicky, D., Polymer International 401 (1996): 275-279. 114. Hedrick, J. L.; Brown, H. R.; Volksen, W.; Sanchez, M.; and Plummer, C. J. G.; and Hilbom, J. S.; Polymer 38(3) (1997): 605-613. 115. Volksen, W.; Yoon, D. Y.; Hedrick, J. L., IEEE Trans. Components, Hybrids Manufact. Technol. 15 (1992): 10. 116. Volksen, W.; Yoon, D. Y.; Hedrick, J. L.; Hofer, D. C.; Mat. Res. Soc. 227 (1992): 23. 117. Volksen, W.; Hedrick, J. L.; Russell, T. P.; Swanson, S., J of Applied Polymer Science 66 (1997): 99. 118. Chen, X. and Gardella Jr., J. A. Macromolecules 27 (1994): 12. 119. Zhuang, H.; Gardella, Jr.; Incavo, J. A.; Rojstacze, R. S.; Rosenfeld, J. C., J. Adh., 63(1) (1997): 199 120. Dwight, D., McGrath, J. E.; Lawson, G.; Patel, N.; York, G., “Surface and bulk microphase separation in siloxane-containing block copolymers and their blends: The roles of composition and kinetics from multiphase macromolecular systems,” Contemp. Topics in Plym, Sci., Vol. 6, B. M. Culbertson, Ed., Plenum, pp.265-288. 82

121. Clarson, S. J. and Semlyen, J. A., eds., Siloxane Polymers, New Jersey: PTR Prentice Hall, 1993. 122. Langmuir, I., Amer. Chem. Soc. 38 (1916): 2286. 123. Roe, R. J., J. Phys. Chem. 72 (1968): 2013. 124. Jarvis, M. L.; Fox, R. B.; and Zisman, W. A.; in “Contact Angle, Wettability and Adhesion,” Advances in Chemistry 43, American Chemical Society, Washington, D.C., 1964, p.317. 125. Clark, D. T.; Peeling, J.; and O’Malley, J. M.; J. Polym. Sci. Polym. Chem. Ed., 17 (1976): 543 126. Chen, X.; Gardella, J. A., Jr.; Kumler, P.L. Macromolecules 25 (1992): 6621. 127. Chen, X.; Gardella, J. A., Jr.; Kumler, P.L. Macromolecules 25 (1992): 6631. 128. Chen, X.; Gardella, J. A., Jr.; Kumler, P.L. Macromolecules 26 (1993): 3778. 129. Chen, X.; Lee. H. F.; Gardella, J. A., Jr., Macromolecules 26 (1993): 4601. 130. Dwight, D.W.; McGrath, J.E.; Riffle J. S.; Smith, S.D.; and York G. A. Journal of Electron Spectroscopy and Related Phenomena 52 (1990): 457. 131. Ttezuka, Y.; Ono, T.; and Imai, K.; Journal of Colloid and Interface Science, 136(2) (1990): 408. 132. Pike, J.; Ho, T.; Wynne, K. J. Chem. Mater. 8 (1996): 856-860. 133. Furukawa, N; Yuasa, Y.; Omori, F. and Yamada, Y., J. Adhes., 59 (1996): 281. 134. Andrade, J. D.; Gregonis, D.E.; and Smith, L.M; in Andrade, J. D. ed., Surface and Interfacial Aspects of Biomedical Polymers, Vol.1, New York: Plenum Press, 1985, p. 15. 135. Schmitt, R.L.; Gardella J. A., Jr.; Magill, J.H.; Salvati L., Jr.; Chin, R. L. Macromolecules 18 (1985): 2675. 136. Gardella J.A.; Hercules, D. M., Anal. Chem. 53 (1981): 1879-1884. 137. Mittlefelht, E.R.; Gardella, J.A., Jr.; Appl. Spectrosc. 191 (1986): 227. 138. Mittlefelht, E.R.; Gardella, J.A., Jr.; Appl. Spectrosc. 1989, 43, 1172. 139. Hasegawa, H. and Hashimoto, T. Macromolecules 1985, 18, 589. 140. York, G. A. “Structure Property Relationships of Multiphase Copolymers,” Ph.D. Dissertation, VPI&SU, 1990. 83

141. Gorelova, M.; Levin V. Makromol. Chem., Macromol. Symp. 44 (1991): 317. 142. Thomas, H.R.; O’Malley, J.J. ACS Symp. Series, Photo Electron and Ion Probers of Polymer Structure and Properties 162 (1981): 320. 143. Pertsin, A.J.; Gorelova, M. M.; Levin, V. Y.; and Makarova, L. I. J. Appl Polymer Sci. 45 (1992): 1195. 144. Briggs, D. and Seah, M. P., eds. Practical Surface Analysis, vol. 1, Wiley, 1990. 145. Hunt, B.J. and James, M.I., eds., Polymer Characterisation, Glasgow: Blackie Academic & Professional, 1993. 146. Dilks, A., ACS Symposium Series 162 (1981): 294. 147. Wagner, C.D.; Riggs, W.M.; Davis, L.E., Handbook of Photoelectron Spectroscopy, ed. Mondla, J. F. and Muilengerg, G.E., Eden Prarie, MN: Perkin-Elmer Corporation Physical Electronics Division,1979. 148. Ratner, B. D. and Horbett, T. A., Coll. Interf. Sci. 83 (1981): 630. 149. Seah, M.P.; Dench, W.A. Surf. Interface Anal. 1(1) (1979): 2. 150. Payter, R.W. Surf. Interface Anal. 3(4) (1981): 186. 151. Powell, C.J. J. Electron Spectrosc. Related Phenom. 47 (1988): 197. 152. Chen, X.; Gardella, Macromolecules 27(12) (1994): 3778. 153. Hansma, P. K.; Cleveland, J. P.; Radmacher, M. D.; Walters, A.; Hillner, P. et al., Appl. Phys. Lett. 64 (1994): 1738. 154. Magonov, S. N.; Elings, V.; Whangbo, M.-H., Surf. Sci. Lett. 375 (1997) L385. 155. Whangbo, M.-H.; Magonov, S. N.; Bengel, H., Probe Microsc. 1 (1997): 12. 156. Bonnell, D. A., ed., “Scanning Tunneling Microscopy and Spectroscopy, Theory, Techniques, and Applications,” New York: VCH Publishers, 1993, pp. 191 – 249. 157. Wiesendanger, R. “Scanning Probe Microscopy and Spectroscopy,” Cambridge: Cambridge University Press, 1994, pp. 210-251. 158. Digital Instruments website: http://www.di.com. 159. Chernoff, D.A. DI Nanovations, V3 N1 (1996): 6-9. 160. Magonov, S. DI Application Notes, 1995. 84

161. Grim, P. C. M.; Psarros, M.; ten Brinke, G.; and Hadziioannou, G., Polymer Prep. 35(1) (1994): 125 162. Albrecht, T. R.; Dovek, M. M.; Lang, C. A.; Grutter, P.; Quate, C. F.; Kuan, S. N. J.; Frank, C. W.; Pease, R. F. W. J. Appl Phys. 64 (1988): 1178. 163. Leclére, Ph.; Lazzaroni, R.; Brédas, J. L.; Yu, J. M.; Dubois, Ph.; and Jérôme, R. Langmuir, 12(18) (1996): 4318. 164. Annis, B. K.; Noid, D. W.; Sumpter, B. G.; Reffner, J. R. and Wunderlich, B., Makromol. Chem., Radid Commun. 13 (1992): 169. 165. Saraf, R. F. Macromol. 26 (1993): 3623. 166. Annis, B. K.; Schwark, D. W.; Reffner, J. R.; Thomas, E. L.; Wunderlich, B.; Makromol. Chem. 193 (1992): 2589. 167. Berg, R. van den; Groot, H. de; Dijk, M. A. van; and Denley, D. R., Polymer 35(26) (1994): 5778. 168. Motomatsu, M.; Mizutani, W.; and Tokumoto, H., Polymer 38,(8) (1997): 1779. 169. Berg, R. van den; Groot, H. de; van Dijk, M. A.; and Denley, D. R., Polymer 35 (1995): 5778. 170. Szesztay, M.; Ghadir, M., Angew Makromol. Chem 209 (1993): 11. 171. Konas, M.; Moy, T. M.; Rogers, M. W.; Shultz, A. R.; Ward, T. C.; McGrath, J. E., J. of Polymer Sci., Part B: Polymer Physics 33 (1995): 1429. 172. Konas, M.; Moy, T. M.; Rogers, M. W.; Shultz, A. R.; Ward, T. C.; McGrath, J. E., J. of Polymer Sci., Part B: Polymer Physics 33 (1995): 1441. 173. Schauer, J.; Sysel, P.; Marousek, V.; Pentka, Z.; Pokorny, J.; Miroslav, B.; Journal of Appl. Polym. Sci. 61 (1996): 1333. 174. Samseth, J.; Spontak, R. J.; Mortensen, K., J. Polym. Sci.: Part B 31 (1993): 467. 175. Samseth, J..; Mortensen, K., J.; Burns, T. L; spontak, R. J., J. Appl. Polym. Sci. 44 (1992): 1245. 176. Tesoro, G. C.; Rajendran, G. P.; Uhlmann, D. R.; Park, C.-E., Ing. Eng. Chem. Res. 26(8) (1987): 1673. 85

177. Carlin, D. M., Proc. 47th SPE Ann. Tech. Conf., Vol. XXXV, May 1-4, (1989), p. 14447. 178. Gaur, U.; Desio, G.; and Miller, B.; Poc. 47th SPE Ann. Tech. Conf., Vol. XXXV, May 1-4, (1989), p. 1513. 179. Lustiger, A., Proc. 47th SPE Ann. Tech. Conf., Vol. XXXV, May 1-4, (1989), p. 1493. 180. Encyclopedia of Polymer Science and Engineering, vol. 12, New York: John Wiley & Sons, 1988, p. 445. 181. Bucknall, C. B., Toughened Plastics, London: Applied Science, 1977. 182. Hedrick, J. L.; Patel, N. M.; and McGrath, J. E., “Toughening of Epoxy Resin Networks with Functionalized Engineering Thermoplastics in Toughened Plastics I: Science and Engineering,” ACS Adv. Chem. Series 233, Riew, C. K. and Kinloch, A. J. eds., ACS, Washington, D. C., pp. 293-304. 183. Yoon, T. H.; Priddy J., D. B.; Lyle, G. D.; and McGrath, J. E.; Macromolecules, 98 (1995): 673. 184. Pater, R., Proc. 47th SPE Ann. Tech. Conf., Vol. XXXV, May 1-4, 1989, p. 1434 185. Hanky, A. O. and St. Clair, T. L., SAMPE J., 21(4) (1985): 40. 186. Raghava, R. S., J. Polym. Sci., Part B, Polym. Phys., 26 (1988): 65. 187. Delvigs, P., Polym. Compos., 7(2) (1986): 101. 188. Pater, R. H., SAMPE J. 26(5) (1990): 19 189. Bell, J. P.; Chang, J.; Rhee, H. W.; and Joseph, R., Polym. Compos. 8(1) (1987): 46. 190. Chang, J.; Bell, and Shkolnik, S., J. Appl. Polym. Sci. 34(1987): 2105. 191. Dagli, G. and Sung, N. H.; Proc. ACS, Polym. Materials: Sci. Eng. 56 (1987): 410. 192. Crasto, A. S.; Own, S. H.; and Suramanian, R. V., in Composite Interfaces, Ishida, H. and Koening, J. L. ed., New York: North-Holland, 1986. 193. Kardos, J. L., ACS, Div. Polym. Chem. Polym. Prep., 24 (1983): 185. 194. Wimolkiatisak, A. S., Bell, J. P.; J. Appl. Polym. Sci., 56(11) 1992: 1899. 86

195. Yoon, T. H.; Liptak, S. C.; Priddy Jr., D. B.; and McGrath, J. E.; J. Adhesion, 45 (1993): 191-203. 196. Bucknall, C. B; Gibert, A. H., Polymer 30 (1989): 213-217. 197. Bascom, W. D.; Yon, K.-J.; Jensen, R. M.; Cordner, J. Adhes. 34 (1991): 79-98. 198. Madhukar, M. S.; and Drzal, L. T. Journal of Composite Materials 25 (1991): 932-957. 199. Robertson, M. A.; Bump, M. B.; Verghese, S. R. McCartney, Lesko, J. J.; Riffle, J. S.; Kim, I.-C.; and Yoon, T.-H. Journal of Adh. Sci., in press. 200. Tummala, R. R. and Rymaszewski, E. J. “Microelectronics Packaging Handbook,” Chapter 1, New York: Van Nostrand Reinhold, 1989. 201. Lai, J. H., Ed., Polymers for Electronic Applications, Boca Raton, Florida: CRC Press, 1989. 202. Rogers, M. E., Brink, M. H.; McGrath, J. E., Brennan, A., Polymer 34(4) (1993): 849. 203. Rogers, M. E.; Woodward, M. H.; Brennen, A; Cham, P. M.; Marand, H.; and McGrath, Polymer Prep. 33(1) (1992): 461. 204. Matsuura, T.; Ishizawa, M.; Hasuda, Y.; and Nishi, S, Macromolecules, 25 (1992): 3450. 205. Eashoo, M.; Shen, D.; Wu, Z.; Lee, C. J.; Harris, F. W.; and Cheng, S. Z. D., Polymer 15 (1993): 3209. 206. Chuang, K. C.; Kinder, J. D.; Hull, D. L.; McConville, D. B.; Youngs, J. W.; Macromolecules, 1997, 30, 7183-90. 207. Rogers, H. G.; Gaudiana, R. A.; Hollinsed, W. C.; Kalyanaraman, P. s.; Manello, J. S.; McGowan, C.; Minns, R. A.; Sahatjian, R., Macromolecules, 18 (1985): 1058. 208. Auman, B., Mat. Res. Soc. Symp. Proc. 381 (1995): 19. 209. Dunson, D. L.; Sankarapandian, M.; Glass, T. E.; Oyama, H.; Farr, I. V.; and McGrath, J. E., Polymer Preprint 39(2) (1998): 796. 210. Nagarkar, P. V.; Lu, J. P.; Volfson, D.; Jensen, K. F.; Senturia, S. D., Mat. Res. Soc. Symp. Proc. 264 (1995): 263. 87

211. Lee, Y. K.; Jeng, S.-P.; Auman, B., Mat. Mat. Res. Soc. Symp. Proc. 264 (1995): 31. 212. Hendricks, N. H., Lau, K. S. Y.; Smith, A. R.; Wan, W. B., Mat. Mat. Res. Soc. Symp. Proc. 264 (1995): 59. 213. Carter, K. R.; Cha, H. J.; Dipietro, R. A.; Hawker, C. J.; Hedrick, J. L.; Labadie, J. W.; McGrath, J. E.; Russell, T. P.; Sanchez, M. I.; Swanson, S. A.; Volksen, W.; and Yoon, D. Y., Mat. Mat. Res. Soc. Symp. Proc. 264 (1995): 79. 214. Sachdev, H. S.; Khojasteh, M. M.; and Feger, C., eds., Advances in Polyimides and Low Dielectric Polymers, Proceedings of the 6th International Conference on Polyimides and other Low K Dielectrics, 1999. 215. Kander, R. G. and Chen, Q.-Q., Polymer 34(4) (1993): 783. 89

CHAPTER 3. SYNTHESIS AND CHARACTERIZATION OF POLY(SILOXANE IMIDE) BLOCK COPOLYMERS

3.1 SYNOPSIS

This chapter discusses the synthesis and characterization of polydimethylsiloxane (PDMS) containing block copolymers via an amic acid salt intermediate. The PDMS oligomers were prepared by ring opening equilibration in the presence of an aminopropyl functional end-blocker. The block copolymer synthesis was carried out via a two step process by a homogeneous method. The first step consisted of endcapping the PDMS oligomer with an excess of anhydride endgroups, then this was reacted with an aromatic diamine. The method presented represents a particularly attractive route to poly(siloxane imide) block copolymers because the amic acid intermediates are soluble in moderately polar solvents such as tetrahydrofuran. The imidization can easily be effected in the melt above the upper Tg. This circumvents problems with solubility in related synthetic methods associated with the fact that the PDMS oligomeric starting materials are not soluble in the dipolar aprotic solvents (eg. NMP) normally utilized for polyimidizations.

3.2 EXPERIMENTAL

3.2.1 Purification of Solvents and Reagents

The utilization of pure reagents and solvents is vital for the synthesis of high molecular weight poly(dimethylsiloxane imide) block copolymers. Therefore, the solvents and reagents were carefully purified and dried before use. 90

Tetrahydrofuran (THF) THF (EM Sciences) was refluxed over sodium with a small amount of benzophenone as an indicator for dryness (which formed the purple sodium benzophenone ketyl). The solvent was distilled immediately prior to use (b.p. 67°C). O

Toluene Toluene (Aldrich) was dried over calcium hydride overnight and distilled immediately prior to use (b.p. 110°C).

CH3

2,2’-Bis[4-(3,4-dicarboxyphenoxy)phenyl]propane dianhydride (bisphenol- A dianhydride, BPADA) BPADA (General Electric) was recrystallized from toluene (molar ratio 4:1 toluene:BPADA) and acetic anhydride (molar ratio 3:1 BPADA:acetic acid) under nitrogen. After filtering, the crystals were refluxed to dissolve them in toluene (molar ratio 4:1 toluene:BPADA) and acetic anhydride (molar ratio 15:1 BPADA:acetic anhydride). The second recrystallization process was repeated twice then the crystals were washed with fresh toluene and n-heptane. The purified BPADA was dried in a vacuum oven at 140°C overnight immediately prior to use (m.p. 189.1-191.1°C). O O CH3 O C O O O CH3

O O 91

Hexafluoroisopropylidene-2-bis(phthalic acid anhydride) (6FDA) Purified, monomer grade, 6FDA (Hoechst Celanese) was vacuum dried at 180°C overnight to ensure complete anhydride cyclization immediately prior to use (m.p. 247°C).

O O CF3

O O CF3

O O

Meta-phenylene diamine (MPDA) MPDA, (Aldrich) was sublimed using an 80°C oil bath and stored under nitrogen in the dark prior to use (m.p. 64-66°C).

H2N NH2

Octamethylcyclotetrasiloxane (D4)

D4 (Petrarch) was stirred over crushed calcium hydride under a nitrogen atmosphere overnight in order to remove any water present. A vacuum distillation of the D4 was carried out at 60°C using a short path distillation apparatus. The dried D4 was stored until used under a nitrogen atmosphere in a round bottom flask sealed with a rubber septum (r = 0.956 g/ml). 92

CH3 CH3 CH3 Si O O Si CH3

CH3 Si O O Si CH3 CH3 CH3

Bis(3-aminopropyl)tetramethyldisiloxane Bis(3-aminopropyl)tetramethyldisiloxane (Petrarch) was stirred over crushed calcium hydride under a nitrogen atmosphere. The material was distilled at 120-130°C under a low vacuum, and stored under a nitrogen atmosphere until used (r = 0.897 g/ml).

CH3 CH3

H2N CH2 Si O Si CH2 NH2 3 3 CH3 CH3

3.2.2 Preparation of Tetramethylammonium Siloxanolate Catalyst

To prepare the tetramethylammonium siloxanolate catalyst, the apparatus equipped with a mechanical stirrer, nitrogen inlet running through the solution, and Dean Stark trap was flame dried. Tetramethylammonium hydroxide pentahydrate (0.475 g, 0.00452 moles) was added to the reaction vessel followed by addition of D4 (9.56 g) via syringe. The reaction mixture was stirred with nitrogen bubbling through it for 42 h at 80°C. The nitrogen flow was high enough to aid in the removal of water to the Dean Stark trap. Following the reaction, the mixture was cooled to room temperature and diluted with D4 (10 ml) to reduce the viscosity. The catalyst was used almost immediately to avoid moisture absorption of atmospheric water. Figure 3-1 contains the synthetic 93

CH3 CH CH Si O 3 3 CH3 O Si CH3 HO + N CH3 CH3 Si O CH3 CH3 O Si CH3 CH3 CH3

80°C, 42 hours

CH 3 CH3 CH CH CH3 H3C 3 3 CH N 3 O Si O Si O Si O N H3C x CH3 CH3 CH3 CH3 CH3 CH3

Figure 3- 1. Synthetic scheme for the preparation of the tetramethylammonium siloxanolate catalyst. 94 scheme for the preparation of the tetramethylammonium siloxanolate catalyst. The catalyst concentration was approximately 6.82 x 10-5 mol/g of solution.

3.2.3 Synthesis of Aminopropyl Terminated Polydimethylsiloxane (PDMS) Oligomers

The synthetic route used was an equilibration of the cyclic tetramer D4 and a difunctional siloxane endblocking reagent in the presence of a siloxanolate catalyst (Figure 3-2).1-4 An example of the equilibration procedure is discussed for the preparation of a targeted 1000 oligomer. To a flame dried 500 ml one-neck round bottomed flask equipped with a magnetic stir bar and septum was added bis(3-aminopropyl)tetramethyldisiloxane (49.704 g, 0.12 mol, 55 ml) followed by D4 (150.3 g, 0.507 mol, 155 ml). The reaction mixture was heated to

80°C and 0.04 mole % based on moles of D4 (2.5 ml) of the siloxanolate catalyst solution was added. The reaction temperature was maintained for 24 hours to allow the mixture to equilibrate. The reaction mixture was subsequently heated at 140°C for 3 hours while bubbling nitrogen through it to decompose the catalyst and evolve the trimethylamine by-product. Then the reaction mixture was heated at 125°C under vacuum (1-2 mm pressure) for 5 hr. to remove the cyclics from the equilibrium mixture. The polymer was stored under nitrogen in a flame dried bottle until used. PDMS oligomers were prepared with targeted number average molecular weights of 2000, 5000, and 11000 g/mol using a similar procedure. Size exclusion chromatography was used to verify the removal of cyclics resulting from equilibration. Potentiometric titration (based on the amount of amine endgroups present) and 1H NMR were used to verify the molecular weights of the PDMS oligomers. 95

CH3 CH CH Si O 3 3 CH3 CH3 O Si CH3 + H2N CH2 Si O Si CH2 NH2 CH3 Si O 3 3 O Si CH3 CH3 CH3 CH3 CH3

Tetramethylammonium Siloxanolate Catalyst 80°C, 24 hours 0.04 mole % 140°C, 3 hours

CH3 CH3

H2N CH2 Si O Si CH2 NH2 + Cyclics 3 n 3 CH3 CH3

Figure 3- 2. Synthetic scheme for the preparation of aminopropyl terminated PDMS oligomer. 96

3.2.4 Derivatization of Aliphatic Amine Endgroups for Size Exclusion Chromatography Analysis

Derivatization of the aliphatic amine endgroups were carried out to avoid possible interference with the ultrastyragel columns used in size exclusion chromatography (Figure 3-3).5 Approximately 2 g of the polysiloxane oligomer was placed in a 20 ml vial with 1.2 mol of benzophenone per mole of amine end- groups. Activated molecular sieves were added to absorb the water evolved from the reaction. This also aided in driving the reaction to completion. The vial was capped with a rubber septum with a needle outlet. The vial was placed in a 120°C oil bath overnight. The samples were filtered before size exclusion chromatography analysis.

3.2.5 Synthesis of BPDA/MPDA Poly(siloxane-amic acid) Block Copolymers

A sample procedure is discussed for an 80 weight % BPDA/MPDA polyimide-siloxane block copolymer. The reaction vessel for this procedure was a dry two-neck 250 ml round bottom flask containing a magnetic stir bar and fitted with an addition funnel and stopper. BPADA (4.920 g, 9.5 mmol) was weighed into the reaction vessel and purged with nitrogen. Dry THF (100 ml) was added to the flask via cannula to dissolve the contents. Aminopropyl terminated PDMS oligomer (1.473 g, 1.4 mmol) with an of 1050 g/mol was added via syringe into a dry, 50 ml, round bottom flask sealed with a rubber septum. Dry THF (10 ml) was added to the aminopropyl terminated PDMS and stirred. The PDMS solution was transferred to an addition funnel via syringe and added dropwise to the stirring BPADA solution. The 50 ml PDMS flask was washed twice with THF (10 ml), and the washings were added to the reaction in the same manner. The reaction mixture was allowed to stir 2-3 hr. MPDA (0.216 g, 0.002 moles) was weighed into a dry 500 ml round bottom flask containing a 97

O CH3 CH3 C + H2N CH2 Si O Si CH2 NH2 2.4 3 n 3 CH3 CH3

OH H CH3 CH3 H OH

N CH2 Si O Si CH2 N 3 3 CH3 n CH3

CH3 CH3 C C N CH2 Si O Si CH2 N + 2 H2O 3 3 CH3 n CH3

Figure 3- 3. Synthetic scheme for the derivatization of aliphatic amine endgroups. 98 magnetic stir bar. The flask was sealed with a rubber septum and purged with nitrogen. Dry THF (20 ml) was added to the flask via cannula and stirred. The anhydride terminated PDMS solution was transferred to the stirring amine solution. The 250 ml flask was also washed with THF. A sufficient amount of THF was added to the reaction to provide a solution of 6 weight % solids. The reaction mixture was stirred for 24 hours. The poly(siloxane-amic acid) solution was exposed to a brisk nitrogen flow at room temperature to evaporate most of the solvent. After two days the poly(siloxane-amic acid) was placed in a vacuum oven at 25°C to remove more solvent. The BPADA/MPDA poly(siloxane-amic acid) block copolymers were analyzed by 1H NMR to determine residual solvent content.

3.2.6 Synthesis of 6FDA/MPDA Poly(siloxane amic acid) Block Copolymers

A sample procedure is discussed for a 90 weight % 6FDA/MPDA poly(siloxane imide) block copolymer. The reaction vessel for this procedure was a dry two-neck 250 ml round bottom flask containing a magnetic stir bar and fitted with an addition funnel and stopper. 6FDA (8.88g, 20mmol) was weighed into the reaction vessel and purged with nitrogen. Dry THF (100 ml) was added to the flask via cannula to dissolve the contents. Aminopropyl terminated PDMS oligomer (2.691g, 1.3mmol) with an of 2070 g/mol was added via syringe into a dry, 50 ml, round bottom flask sealed with a rubber septum. Dry THF (10 ml) was added to the aminopropyl terminated PDMS and stirred. The PDMS solution was transferred to an addition funnel via syringe and added dropwise to the stirring BPADA solution. The 50 ml PDMS flask was washed twice with THF (10 ml), and the washings were added to the reaction in the same manner. The reaction mixture was allowed to stir 2-3 hr. MPDA (2.0196g, 18.7mmol) was weighed into a dry 500 ml round bottom flask containing a magnetic stir bar. The flask was sealed with a rubber septum and purged with nitrogen. Dry THF (20 99 ml) was added to the flask via cannula and stirred. The anhydride terminated PDMS solution was transferred to the stirring amine solution. The 250 ml flask was also washed with THF. A sufficient amount of THF was added to the reaction to provide a solution of 10 weight % solids. The reaction mixture was stirred for 24 hours. The poly(siloxane-amic acid) solution was exposed to a brisk nitrogen flow at room temperature to evaporate most of the solvent. After two days the poly(siloxane-amic acid) was placed in a vacuum oven at 25°C to remove more solvent. The fluorinated poly(siloxane-amic acid) block copolymers were analyzed by 1H NMR to determine residual solvent content.

3.2.7 Determination of Weight % Solvent in Poly(siloxane amic acid)

To reduce the possibility of error in molecular weight determination and poly(siloxane amic acid) salt preparation, the weight % solvent in the poly(amic acid) samples was determined utilizing 1H NMR on a Varian 400 MHz instrument. The peak area due to the 4 THF protons at 3.6 ppm was compared to the peak area resulting from the 2 amide protons per poly(siloxane amic acid) repeat unit at 10.4 ppm. The amount of residual THF corresponded to 1 THF molecule per 1 –COOH group.

3.2.8 Preparation of BPADA/MPDAPoly(siloxane amic acid) Salt

The tripropylamine base was titrated with 0.025 N HCl that was previously standardized against Na2CO3. The salt was prepared in a basic methanolic solution. Tripropylamine (3.4 mmol) was dissolved in methanol (20 ml) in a 50 ml beaker. The poly(siloxane amic acid) block copolymer (2 g, 1.4 mmol repeat units) was weighed into a 50 ml round bottom flask. The basic solution was added slowly to the copolymer while stirring. The volume of tripropylamine added was a 10 mole % excess relative to the carboxylic acid groups. The 100 mixture was stirred until homogeneous, indicating complete conversion to the salt. The solvent was then evaporated under a nitrogen flow. The poly(siloxane amic acid) salts were further dried under vacuum at 20°C for 48 hours.

3.2.9 Thermal Imidization of the Poly(siloxane amic acid)s and Poly(siloxane amic acid) Salts

The BPADA/MPDA poly(siloxane amic acid) salt block copolymers were imidized under nitrogen at 260°C for 10 min, and the fluorinated poly(siloxane amic acid) block copolymers were imidized at 280°C for 15 minutes. The poly(siloxane imide) block copolymers were analyzed by size exclusion chromatography, differential scanning calorimetry, thermogravimetric analysis, nuclear magnetic resonance, dynamic mechanical analysis, and transmission electron microscopy.

3.3 CHARACTERIZATION

3.3.1 Molecular Weight Determination

The values of the propylamine terminated PDMS oligomers were determined by 1H NMR and potentiometric titration of the amine endgroups with

0.025 N HCl (standardized against Na2CO3). A MCI GT-05 (COSA Instruments Corp.) automatic potentiometric titrator was used in the analysis. Molecular weights were established after the equilibration catalyst and equilibrium cyclics had been removed. The molecular weights of the block copolymers were determined by size exclusion chromatography in N-methylpyrrolidone (NMP) (dried over phosphorus pentoxide) at 60oC. A Waters 150C instrument equipped with differential refractive index and viscosity (Viscotek Model 100 Differential Viscometer) 101 detectors in parallel was utilized. Absolute molecular weight averages were calculated by the “universal calibration” procedure based on the hydrodynamic volumes of solute molecules appearing at the same elution volume. Since differential refractive indices were used to determine solute concentration, it was further assumed that the average refractive indices of the block copolymer molecules appearing at a given elution volume were the same as those appearing at any other elution volume. This assumes that the composition distribution of the copolymers across the molecular weight distribution range was constant.

3.3.2 Thermogravimetric Analysis

The weight losses versus temperature were determined using a Perkin Elmer TGA 7 thermogravimetric analyzer. This provides some indication of the thermal and thermo-oxidative stability of the materials. Samples of the polyimides obtained via the amic acid route and those obtained via the amic acid salt route were placed in platinum pans connected to an electric microbalance. The samples were heated at a rate of 10°C per minute in a flowing air atmosphere and in a nitrogen atmosphere. Weight losses of the samples were measured as a function of time and temperature. The data were reported as 5 weight % loss temperatures. In addition, isothermal TGA was performed on BPADA/MPDA poly(siloxane-amic acid) salts to assess their relative imidization times. The weight losses of the amic acid salts were measured as a function of time and temperature.

3.3.3 Differential Scanning Calorimetry

Differential scanning calorimetry (DSC) was used to determine the upper glass transition temperatures of the block copolymers using a Perkin-Elmer DSC 102

7. The samples were run using a heating rate of 10°C per minute from 25°C to 250°C for the BPADA/MPDA block copolymers and 25°C-350°C for those prepared with 6FDA. The Tg values were obtained from the second heating cycle after rapid cooling from above the Tg of the polymer. The values were taken as the midpoint temperatures of the change in slope of the DSC curves.

3.3.4 Dynamic Mechanical Analysis

The upper glass transition temperatures of some of the block copolymers prepared with 6FDA were determined using a Perkin-Elmer DMA7. The samples were run in extension mode with a heating rate of 5°C per minute from 25°C to

350°C. The Tg values were obtained from the tan delta peaks.

3.3.5 Transmission Electron Microscopy

The morphologies of the block copolymers were examined using transmission electron microscopy on a Philips 420T TEM at 100kV. The block copolymer films were cryo-microtomed at -120°C using a Reichert-Jung Ultra-E ultramicrotome equipped with a FC-4D cryo attachment. Approximately 500-700 Å thick sections were cut and collected on 600 mesh copper grids and analyzed.

3.4 RESULTS AND DISCUSSION

3.4.1 PDMS Oligomers

Synthesis of PDMS oligomers. Aminopropyl terminated PDMS oligomers with targeted values of 1000, 2000, 5000, and 11000 g/mole 103

were prepared via the well established equilibration polymerization of D4 with 1,3- bis(aminopropyl)tetramethyldisiloxane as the endblocking reagent.1-4 The reaction proceeded by an anionic mechanism and involved an equilibration of linear polymer chains and low molecular weight cyclics. The cyclics are a result of backbiting where a growing polymer chain breaks the Si-O bond along the linear backbone. Removal of the cyclics after equilibration was carried out using vacuum distillation. This was necessary in order to obtain accurate stoichiometries in the copolymer synthesis and to avoid cyclic impurities in the resultant block copolymers. Size exclusion chromatography proved to be an efficient method for monitoring the removal of the equilibrated cyclics.5 The PDMS oligomers were derivatized with benzophenone to form ketimine endgroups prior to SEC analysis. This was carried out to avoid the adsorption of the aliphatic amines on the ultrastyragel columns of the SEC chromatograph. Figure 3-4 shows the size exclusion chromatographic traces of the 11000 g/mol derivatized aminopropyl terminated PDMS oligomer before and after removal of the cyclics. Approximately 10-12 weight % cyclics remained in the equilibrated samples. This is in agreement with results reported by Riffle et al. of 10-15 weight %.1 Molecular Weight Characterization of PDMS oligomers. The molecular weights of the stripped PDMS oligomers were established by two methods: 1H NMR and potentiometric titration of the aliphatic amine endgroups. Figure 3-5 contains the 1H NMR spectrum of the PDMS oligomer with a targeted of 2000 g/mol. The protons assigned to the methylene groups (labeled b and d) were used in calculating the molecular weights (Figure 3-6). 104

-96.0 before after

-96.5

-97.0 Response (mV)

-97.5 10 20 30 Retention Volume (mV)

Figure 3- 4. Size exclusion chromatograms of a 11,000 g/mole PDMS oligomer before and after stripping of the equilibrated cyclics. 105

a a CH CH 3 3 b c c d H2N CH2 Si O Si CH2 CH2 CH2 NH2 3 n 163.34 CH3 CH3

d c,e b 4.00 8.14 3.74

3.0 2.5 2.0 1.5 1.0 0.5 0.0 PPM

Figure 3- 5. 1H NMR spectrum of PDMS oligomer with a targeted of 2000 g/mol. 106

Sum of the peak areas due to 8 –CH2 protons (b and d) = 7.74

7.74 = 0.9675mol 8

Area due to –CH3 protons (a) = 144.57 163.34 = 27.223RU 6

= [RU/mol * MW (RU)] + MW (EG) – 16

é27.223 ù = * 74.1545 + 116 - 16 = 2186g / mol ëê0.9675 ûú

Figure 3- 6. Sample calculation of the molecular weight using 1H NMR for a PDMS oligomer with targeted a of 2000 g/mol. EG, RU, and MW represent endgroup, repeat unit, and molecular weight, respectively. 107

Table 3-1 contains a summary of the results obtained from the two methods investigated. The experimental results for the PDMS oligomers with targeted s of 1000, 2000, and 5000 g/mol were in excellent agreement with the targeted molecular weights. However, the 11000 PDMS oligomer yielded experimental results somewhat higher than the targeted value. An explanation for these results can not be offered. The titration values were used in the subsequent block copolymer calculations.

3.4.2 BPDA/MPDA Poly(siloxane imide) Block Copolymers

3.4.2.1 Synthesis of BPADA/MPDA Poly(siloxane amic acid) Block Copolymers

The poly(siloxane amic acid) block copolymers were prepared by a two step method in THF. The first step consisted of pre-capping the PDMS oligomer with the dianhydride monomer (Figure 3-7). This was followed by chain extension with the aromatic amine. The capping reaction was carried out slowly by adding the PDMS oligomer dropwise to minimize chain extension. Conventional methods of poly(amic acid) preparation have consisted of adding the solid dianhydride monomer to a stirring solution of diamines.18 This often resulted in inhomogeneous reaction mixtures as the polymer began to form. Pre- capping the PDMS oligomer has been reported to result in better solubility of the reactants throughout the polymerization process.6-11 The reaction mixtures were allowed to stir for 1h at room temperature before chain extension with MPDA to form the poly(siloxane amic acid) block copolymers (Figure 3-8). 108

Table 3-1. Number average molecular weight, , results of aminopropyl terminated PDMS oligomers

PDMS PDMS PDMS (g/mol) (g/mol) (g/mol) (Predicted) (Titration) (1H NMR) 1000 1052 1038 2000 2036 2186 5000 5094 5304 11000 11868 12250 109

O O CH3 CH3 R1 H2N CH2 Si O Si CH2 NH2 + O O 3 3 n CH3 CH3 O O (excess)

THF, 25°C

O O O O CH3 CH3

R1 N CH2 Si O Si CH2 N R1 O 3 n 3 O OH CH3 CH3 HO

O O O O

+ BPADA

CH3 R1 = O O

CH3 Figure 3-7. Schematic of PDMS capping reaction with the anhydride monomer 110

O O O O CH3 CH3

R N CH2 Si O Si CH2 N R O 1 3 n 3 1 O OH CH3 CH3 HO

O O O O

H2N NH2 THF, 25°C

(+ unreacted dianhydride)

O O O O CH3 CH3

CH Si O Si CH N R1 N N R N 2 3 n 2 1 HO OH CH3 CH3 OH HO

O O x O O y

CH3 R1 = O O CH3 Figure 3-8. Synthesis of BPADA/MPDA poly(amic acid) block copolymers 111

3.4.2.2 BPADA/MPDA Poly(siloxane imide) Block Copolymers from Poly(siloxane amic acid) Salt Precursors

The poly(siloxane amic acid)s were exposed to a brisk nitrogen flow at room temperature to evaporate most of the THF solvent, then dried further under vacuum. Since the drying temperature was well below the Tg of the polymer residual THF remained. The weight % of solvent remaining was determined by 1H NMR to reduce the possibilities for weighing errors in the poly(amic acid) salt preparation. The values obtained matched closely to the stoichiometric calculation based on 1 THF per 1 –COOH group. Similar relationships of solvent/poly(amic acid) complexes have been previously reported.12-14 The poly(siloxane amic acid)s were converted to salts by addition of the amic acid powders to basic aqueous or methanolic solutions (Figure 3-9). These intermediates were of interest for fiber/matrix composite interphase materials in composite applications. Coating fibers with these materials from aqueous solutions was a practical method for applying sizings. Tripropylammonium salts of the BPADA/MPDA block copolyamic acids formed clear aqueous and methanolic solutions. After solvent evaporation, the salts were melt imidized at 260°C under nitrogen over 10-15 min (Figure 2-9). TGA confirmed that imidization under these conditions was complete since no weight losses for the resultant materials were observed until above 400°C. In addition, TGA indicated that imidization proceeded relatively rapidly at temperatures just above Tg. One of the precursor salts was heated at 200°C/min up to 270°C where it was held isothermally (Figure 3-10). Rapid weight loss due to imidization and tripropylamine volatilization commenced near 200°C and was complete within about 1.5 minutes prior to reaching 270°C. 200°C was just above the Tg for the imide block in this copolymer which contained 90 weight % imide and 10 weight % dimethylsiloxane (~2000 g/mol PDMS oligomer). It is reasonable to hypothesize that the improved mobility of the molecules in the vicinity of Tg contributed to the rapid imidization at this temperature. 112

O O O O CH3 CH3

CH Si O Si CH N R1 N N R1 N 23 n 2 OH CH3 CH3 HO OH HO O O O x O y

Tripropylamine, N(Pr)3 [1.1 moles / COOH] MeOH (or H2O)

O O O O CH3 CH3

CH Si O Si CH N R1 N N R N 2 3 n 2 1 O CH3 CH3 O O O HN(Pr)3 HN(Pr) (Pr)3NH (Pr)3NH 3 O O O O x y

Heat

(-H2O)

O O O O CH3 CH3 R1 R1 CH Si O Si CH N N N N 2 3 n 2 3 CH3 CH3 O O x O O y

Figure 3- 9. Synthesis of BPADA/MPDA poly(siloxane imide) block copolymer via a poly(siloxane amic acid) salt intermediate. 113

Temperature (°C) 0 100 200 270 Constant at 270 °C

100

80

60 W t. %

40

20

0 0 2 4 6 8 10 Time (minutes)

Figure 3- 10. Isothermal TGA analysis of poly(siloxane imide) block copolymer demonstrating complete imidization within about 1-2 min prior to reaching 270°C. 114

3.4.3 6FDA/MPDA Poly(siloxane imide) Block Copolymers

3.4.3.1 Synthesis of 6FDA/MPDA Poly(siloxane amic acid) Block Copolymers

The methodology used in the preparation of the BPADA/MPDA poly(siloxane imide) block copolymers was later applied to the fluorinated materials. Thus, the PDMS oligomers were pre-capped with the 6FDA monomer by allowing them to react for 1h before chain extension with MPDA and unreated 6FDA (Figure 3-11). The fluorinated block copolymers were imidized directly from the poly(amic acid)s (as opposed to being first converted to salts). The resultant fluorinated poly(siloxane imide) block copolymers were of particular interest for electronic applications, and thus, aqueous solutions were avoided. Size exclusion chromatography analyses of the imidized block copolymers in NMP showed unimodal, approximately gaussian molecular weight distributions for those polymers prepared with 1000 and 2000 g/mol PDMS oligomers (Figure 3-12). However, the block copolymers prepared with the higher molecular weight PDMS oligomers were found to have non-uniform composition distributions across the molecular weight range (Figure 3-13). The RI chromatogram of a block copolymer prepared with the 5000 g/mol PDMS oligomer showed a negative peak on the high molecular weight end of the curve. The refractive index of the NMP SEC solvent was between the refractive indices of the polyimide and PDMS block structures. Pure PDMS appears as a negative peak whereas the polyimide results in a positive peak. Thus, it was believed that the negative peak in the block copolymer chromatogram on the high molecular weight end was a result of abnormally high concentrations of PDMS in the high molecular weight fraction. 115

O CF3 O CH3 CH3 C C H2N CH2 Si O Si CH2 NH2 + O O 3 n 3 CF CH3 CH3 C 3 O O

THF, 25oC

O CF3 O CH3 CH3 O CF3 O C C N CH2 Si O Si CH2 N C C O H 3 n 3 H O CH3 CH3 CF3 C C CF3 O O OH HO O O

H N NH 2 2 THF, 25oC

(+ unreacted 6FDA)

O CF3 O O CF3 O CH3 CH3 C C NH HN CH2 Si O Si CH2 NH HN 3 n 3 CF3 CF OH HO 3 OH CH3 CH3 HO O O O x O y soft segment hard segment Figure 3-11. Schematic of the Synthesis of the Fluorinated Poly(siloxane amic acid) Block Copolymers. 116

-85

-90

-95 1000 g/mol Respose (mV) 2000 g/mol -100 10 20 30 Retention Volume (mL)

Figure 3-12. Size exclusion (RI) chromatograms of imidized block copolymers containing 1000 and 2000 g/mol PDMS oligomers. The PDMS capping reactions were conducted for 1h before chain extension. 117

20

15

10

5 DV 0 RI Response (mV)

-5 10 20 30 Retention Volume (mL)

Figure 3-13. Size exclusion (RI and DV) chromatograms of an imidized block copolymer containing a 5000 g/mol PDMS oligomer. The PDMS capping reaction was conducted for 1h before chain extension. 118

It was believed that the non-uniform composition distributions observed when higher molecular weight PDMS oligomers were used resulted from insufficient pre-capping with the 6FDA monomer. During chain extension these reaction mixtures temporarily phase separated (clouded), then subsequently cleared. It is possible that during this period the block copolymers were polymerizing in two phases, one rich in siloxane and the other rich in MPDA. Therefore, the negative peak in the RI chromatogram was associated with the high molecular weight PDMS rich areas of the block copolymer. The DV chromatogram indicated a higher viscosity for the same area of the RI chromatogram (Figure 3-13). In an effort to better understand the polymerization chemistry associated with the pre-capping step, a 5000 g/mol aminopropyl terminated PDMS oligomer was reacted with a calculated excess of 6FDA and monitored using 1H NMR. The peaks associated with the methylene protons adjacent to the amine group (»2.5 ppm) diminished with the progression of the capping reaction. Thus, these protons were compared to the methylene protons adjacent to the silicon (» 0.5ppm) which remained constant throughout the reaction. Analysis of the reaction mixture over time indicated that the PDMS oligomer was 90% capped after 2.5 hours; however, the methylene peak did not completely diminish until over 4.5 hours (Figure 3-14). This demonstrated that the 1h reaction time allowed for pre-capping the 5000 g/mol PDMS oligomer was not sufficient to provide fully capped oligomers. These results indicated that the amidization reaction in THF of the aliphatic amine (of a 5000 g/mol PDMS oligomer) was considerably slower than that of the aromatic amine as reported by Facinelli et al.15 The authors observed rapid amidization rates in THF with high molecular weights achievable in less than 20 minutes for BPADA/MPDA structures. This is surprising since aliphatic amines (as used in this reaction) are normally much stronger nucleophiles than aromatic amines (as used in Facinelli’s study). The nitrogen lone pair of the aromatic amine can be delocalized through the benzene ring, and the aliphatic amine lone pair is localized (Figure 3-15). Thus, the aliphatic amine should be a 119

after 30 min

after 2.5 hrs

after 4.5 hrs

after 8.5 hrs

3.0 2.8 2.6 2.4 2.2

Figure 3- 14. 1H NMR spectra monitoring the capping reaction of 5000 PDMS oligomer with 6FDA. (NMR precision is approximately +/- 3%) 120

Stronger Nucleophile:

CH2 CH2 CH2 NH2 Lone pair of the aliphatic primary amine has no resonance forms, thus it is localized and should be a better nucleophile.

Weaker Nucleophile:

NH NH NH2 2 NH2 2

Lone pair of the aromatic primary amine can be delocalized through the ring.

Figure 3- 15. Schematic explaining the relative nucleophilicity of the aliphatic versus the aromatic primary amine. 121

better nucleophile than the aromatic amine, therefore leading to faster amidization reactions. The reaction kinetics are governed by the reactant concentrations through the relationship, Rate = k [reactant 1][reactant 2], where k is the rate constant. The higher molecular weight oligomers were considerably more viscous, thus more solvent was added. This reduced the concentration of the endgroups. The inability of the reactive ends to “find each other” in the reaction mixture may lead to lower reaction rates. Table 3-2 contains the reaction concentrations for the monomers used to form the 30 weight % PDMS block copolymers containing different oligomer sizes. The larger PDMS oligomer molar concentrations were considerably lower than those for the smaller oligomers. Unfortunately, the reaction kinetics could not be evaluated for this system since [PDMS] changed over the reaction time. Thus, it is not clear if the aminopropyl endgroups present after 30 min are a result of the last PDMS addition. The slower reaction times may also be a result of salt formation during the initial stages of the reaction. As discussed in section 2.2.1.1, the rate constants for amidization increase as the pKa of the protonated amine increases. The highly basic primary aliphatic amine has a pKa of approximately 10.7 whereas an aromatic amine has a pKa of approximately 4.6. It is probable that the aminopropyl endgroups will abstract the acidic proton of the poly(amic acid), thus reducing the nucleophile concentration in solution. The reaction mixture is in equilibrium, therefore given sufficient reaction time the salt may be reverted back to the poly(amic acid) to complete the polymerization. Figure 3-16 shows the RI chromatogram of a poly(siloxane imide) block copolymer prepared with 30 weight % of the 5000 g/mol PDMS oligomer. The pre-capping reaction mixture was allowed to stir for 6 hours before the addition of the MPDA monomer. The RI chromatogram did not exhibit a negative peak in the high molecular weight region, as previously observed. The expected approximately gaussian shape suggested a uniform copolymer composition over the molecular weight distribution range. 122

Table 3- 2. Monomer molar concentrations for a 6FDA/MPDA poly(siloxane imide) block copolymer containing 30 weight % PDMS

PDMS [PDMS] [6FDA] MW (g/mol) mol/L mol/L 1000 0.0280 0.1093 2000 0.0093 0.1429 5000 0.0057 0.1274 11000 0.0025 0.1250 123

15.8

11.5

7.2

3.0

Response (mV)

-1.3 10.0 14.4 18.8 23.2 27.6 32.0 Retention Volume (mL)

Figure 3- 16. Size exclusion (RI) chromatograph of a poly(siloxane imide) block copolymer prepared with 5000 g/mol PDMS oligomer (30 weight %). PDMS pre- capping reaction was reacted for 6 hrs at 25°C before chain extension. 124

Table 3-3. Copolymer composition and molecular weight of block copolymers

Copolymer PDMS Weight % (g/mol)

PDMS (g/mole) 1000 20 8,000 1000 10 17,000 BPADA/MPDA 2000 70 10,700 2000 50 11,700 2000 10 23,000

1000 30 10,600 1000 10 24,100 2000 30 20,100 6F-DA/MPDA 2000 20 37,900 2000 10 22,700 5000 30 34,100 5000 10 31,100 11000 50 31,000 11000 10 29,900 125

3.4.4 Characterization of the Poly(siloxane imide) Block Copolymers

3.4.4.1 Molecular Weight Analysis

The amic acid precursors were melt imidized at 280°C under nitrogen for 10-15 minutes. The molecular weights of the imidized materials were analyzed using SEC (Table 3-3) in NMP. The values of the block copolymers ranging from 8,000 to 34,000 g/mole indicate the successful attainment of copolymer high molecular weights by the synthetic procedures employed.

3.4.4.2 Thermal Properties

Table 3-4 displays the imide block values, the upper glass transition temperatures (imide block Tgs), and the temperatures at 5% weight loss for the block copolymers formed from given PDMS oligomers and weight % imide compositions.

The values for the imide blocks were calculated from the and the weight composition of the PDMS oligomer, the m-phenylene diamine, and BPADA or 6FDA feed. The following relation was employed:

wh x = (1) ws Ms where x is the of the hard segment (imide block), Ms is the of the soft segment (PDMS), and wh and ws are the weight % of hard segment and soft segment, respectively. All of the materials exhibited thermo-oxidative stabilities well above 400°C by dynamic TGA, but the copolymers showed 5% weight loss temperatures somewhat lower than that of the polyimide homopolymer (Table 3-4). The degree of reduction depended on the amount of siloxane present; the larger the amount of siloxane, the larger the reduction in thermal stability. This reduction in 126

Table 3- 4. Thermal properties of poly(siloxane imide) block copolymers

Weight % PDMS Weight % Imide Block Upper Tg (°C) 5% wt. loss

(g/mol) Imide Block Length (Mn) (°C) (in air) BPA-DA/m-PD 100 5900 190 521 1000 38 818 not visible 436 1000 50 1174 161 435 1000 80 5250 168 458 1000 90 9309 179 460 2000 30 1077 161 450 2000 80 8000 197 476 2000 90 23000 205 502

6F-DA/m-PD 100 27000 287 501 1000 70 2125 168 406 1000 80 2703 184 409 1000 90 6692 258 436 2000 70 4061 186 435 2000 80 7524 214 432 2000 90 13385 285 435 5000 70 11886 228* 423 5000 90 45846 269* 456 11000 50 11863 257* 418 11000 90 106767 260* 440

*Values were determined from the tan delta peak using Dynamic Mechanical Analysis with a heating rate of 5°C/min. 127 thermal stability was believed to be due to the initial degradation of the terminal n-propyl linkages of PDMS.16

The upper Tg’s of the block copolymers, i.e. the Tg’s of the imide block domains, were found to be slightly lower than that of a high molecular weight polyimide homopolymer (Table 3-4). Similar results were reported by McGrath et al.4 for polydimethylsiloxane-polyimide copolymers containing short PDMS blocks

(Mn » 900 g/mol). It was postulated that the lower transition temperatures were due to partial phase mixing. The reduction could also be attributed to decreased imide block molecular weights in the copolymers, as reported by Fitzgerald et

16 al. It is evident from the data (Table 3-4) that the upper Tg’s increase with increasing predicted polyimide block lengths in each series. Assuming the imide blocks relax independently from the dimethylsiloxane blocks, the upper glass transitions (Tg of the imide block domains) should 17 decrease linearly with 1/ of the imide blocks. Figure 3-17 presents plots of 4 Tg vs. 10 / (imide block) for the dimethylsiloxane-imide block copolymers prepared from a 2000 g/mol PDMS oligomer. The BPADA /MPDA imide block

4 copolymer Tg in the limit of infinite imide block (intercept at 10 / = 0)

14 is 203°C. Facinelli et al. reported the Tg of UltemÔ type poly(etherimide)s in the limit of infinite molecular weight to be ~237°C.14 The lower value of the transition temperature of the imide blocks in the present poly(siloxane imide) block copolymers suggests the likelihood that there is some phase mixing of the lower molecular weight siloxane blocks into the imide block domains. Definite evidence of phase mixing was noted in the UltemÔ type block copolymers prepared with a PDMS oligomer having ~1000 g/mole . Its extrapolated upper Tg was only 175°C. A similar extrapolation of the 6FDA/MPDA imide block Tg data (Figure 3-17) for block copolymers prepared with the 2000 g/mole PDMS oligomer yielded a glass transition temperature of 303°C. For the 6FDA/MPDA block copolymer series prepared 128

350

300 303°C

250

200 BPADA Block Copolymer

Tg (°C) 203°C 150 6FDA Block Copolymer

100

50

0 0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012 1/Mn (x104)

Figure 3-17. Polyimide phase glass transition temperature as a function of 1/Mn

(hard block) of copolymers prepared with PDMS having an Mn of ~ 2000 g/mol. 129

with the 1000 g/mole PDMS oligomer the extrapolated Tg was 299°C. Farr reported a Tg value of 295°C for a high molecular weight 6FDA/MPDA 19 polyimide. Although this Tg was not that of an infinite molecular weight polyimide, it appears that siloxane block mixing into 6FDA/MPDA block domains is not as great as siloxane block mixing into BPADA/MPDA block domains. Figure 3-18 shows the dynamic TGA curves in air and nitrogen atmospheres for BPADA/MPDA block copolymers prepared with the 2000 g/mol PDMS oligomer. The results show decreasing thermal with increasing siloxane content as indicated by their initial weight loss temperatures. The curves also indicated an increase in char with an increase in the amount of incorporated PDMS. The formation of char was even more pronounced in a nitrogen atmosphere (Table 3-5). The development of char has been reported to be the result of chain scission (between the silicon and carbon bonds) with simultaneous crosslinking reactions. A generalized view of char formation is shown in Figure 3-19. It has been reported that in the presence of air the silicon undergoes degradative reactions to produce white silicate-like derivatives. XPS analysis of the char resulting from degradation in air confirmed the presence of SiO2. The silicon binding energies reported in Table 3-6 shows a shift of 0.8 eV from the PDMS silicon binding energy to the SiO2 silicon binding energy. The black char resulting from degradation in a nitrogen atmosphere was also characterized using XPS (Table 3-6). The results indicated relatively no change in the silicon binding energy. These results suggest that the degradation follows a different mechanism in air versus nitrogen atmospheres.23 It was speculated that tying up the ends of the siloxane chain allows the material to withstand higher temperatures than the PDMS oligomer thus allowing conversion to the silicate- like structures.24 130

100

80

60 10%, N2 50%, N2

Weight % 50%, Air 40

70%, Air 70%, N2 20

0 10%, Air

0 100 200 300 400 500 600 700 800 Temperature (°C)

Figure 3- 18. TGA curves for BPADA/MPDA block copolymers prepared with the 2000 g/mol PDMS oligomer. 131

Table 3-5. Char yield remaining at 750°C for poly(siloxane imide) block copolymers in air and nitrogen atmospheres.

Polymer PDMS Air Nitrogen

Structure Weight % () Weight % Weight % Polyimide Control 0.04 28.84 BPADA/MPDA 10(2k) 1.69 48.97 20(2k) 6.56 46.68 10(2k) 4.73 52.16 20(2k) 6.12 47.40 50(2k) 11.67 20.64 70(2k) 12.5 22.45

10(1k) 4.0 41.7 6FDA/MPDA 20(1k) 5.8 38.5 10(2k) 1.9 39.7 20(2k) 3.1 42.3 132

Heat Chain Scission Further crosslinking to form char

Crosslinking to form char

Further chain scission to form low molecular weight products

Figure 3-19. Generalized schematic of char formation for siloxane containing polymers.24 133

Table 3- 6. Binding energies for the silicon photopeak of a block copolymer before and after exposure to 750°C in air and nitrogen atmospheres

Silicon Photopeak Binding Energy (eV) Film 102.5 Air Char 103.3 Nitrogen Char 102.6 134

3.4.4.3 Morphology

The morphological properties of the solvent cast block copolymer films were investigated using TEM. Due to higher electron density relative to the polyimide, the siloxane enriched areas appear slightly darker than the imide enriched areas. Figure 3-20 shows a typical TEM micrograph of a block copolymer containing 70 weight % of an ~2000 g/mole polydimethylsiloxane and 30 weight % of the polyimide. The micrograph shows a very fine microphase separated structure with the siloxane as the continuous phase and the imide as the included phase. The domain size can be related to the molecular weight of the spherical domain forming block.20 The imide inclusions were approximately 10 nm in diameter which is larger than what has been reported in the literature for similar structures (~2 nm).21-22 A partial explanation of this might be the possibility of siloxane block mixing into the imide block domains; however, this could not possibly explain the 125-fold difference in domain volumes. 135

100nm

Figure 3- 20. TEM micrograph of an UltemÔ type poly(siloxane imide) block copolymer containing 70 weight % siloxane (~ 2000g/mol) and 30 weight % imide. (The siloxane phase is slightly darker and is the continuous phase and the imide is the included phase). 136

3.5 CONCLSIONS

The two-step poly(amic acid) synthesis in THF proved to be an efficient method of preparation for the systematically varied polydimethylsiloxane- polyimide block copolymers containing BPADA/MPDA and fluorinated structures. The BPADA/MPDA poly(siloxane imide) block copolymers were prepared from water soluble precursors. The poly(siloxane amic acid) salts can be imidized quickly (<10 min. at 260°C), thus suggesting that the potential for online processing in composite manufacturing is promising. The salts of these materials are water soluble and their aqueous solutions are stable in air. Block copolymers containing fluorinated structures were prepared via melt imidization of their amic acid precursors. Both block copolymer series produced microphase separated materials with thermal stabilities well above 400°C. It was demonstrated that the thermal stability and char formation of the block copolymers was influenced by the amount of PDMS incorporated. The reduction in the upper Tg of the copolymers is thought to be due to the lowering of the imide phase molecular weight. This is not completely ignoring the possibility of phase mixing to some extent, especially with those copolymers prepared with 1000g/mol PDMS oligomers.

3.6 RECOMMENDATIONS FOR FUTURE WORK

An extension to this research would include an evaluation of the mechanical properties of the block copolymer systems. Other areas might include the application of the synthetic approach employed to other systems. 137

3.7 REFERENCES

1. Riffle J. S; Yilgor, I.; Tran, C.; Wilkes, G. L; McGrath, J. E.; and Banthia, A. K., “Elastomeric Polysiloxane Modifiers for Epoxy Networks: Synthesis of Functional Oligomers and Network Formations Studies in Epoxy Resin Chemistry II,” R. S. Bauer Ed., ACS Symposium Series # 221, ACS, Washington D.C., 1983. 2. Gungor, A.; Smith, C.; Wescott, J. M.; Srinivasan, S. and McGrath, J. E., Polym. Prep., Am. Chem. Soc. Div. Poly. Chem., 32(1) (1991): 172. 3. Gilbert, A. K. and Kantor, S. W. J. Polym. Sci., 40, 35 (1959). Sormani, P. M. and McGrath, J.E. “Kinetics and Mechanisms of the Anionic Ring Opening Polymerization of Cyclosiloxanes in the Presence of Bis(1,3-aminopropyl tetramethyldisiloxane)” in Ring Opening Polymerization: Kinetics, Mechanisms and Synthesis, J. E. McGrath, ed., ACS Symposium Ser. #286, (1985). 4. McGrath, J. E.; Sormani, P. M.; Elsbernd, C. S. and Kilic, S. Makromol. Chem. Macromol. Symp. 6 (1986): 67. 5. Senger, C. L., Synthetic Studies on Siloxane Hompolymers and Segmented Copolymers, Ph.D. dissertation, VPI&SU, 1988. 6. Summers, J. D.; Elsbernd, C. S.; Sormani, P, M; Brandt, P.J.A.; Arnold, C.A.; Yilgor, I.; Riffle, J.S.; Kilic, S.; McGrath, J.E. ACS Symposium Series 360, pp 180-198. 7. Arnold, C.A.; Summers J.D.; Chen Y.P.; Bott R.H.; Chen, D.; McGrath, J. E. Polymer 30(6) (1989): 989. 8. Arnold, C.A.; Summers J.D.; Bott R.H.; Taylor, L.T.; Ward, T.C.; McGrath, J. E. SAMPE Proceedings 32 (1987): 586. 9. Arnold, C.A.; Summers, J.D.; Bott, R.H.; Taylor, L.T.; Ward, T.C.; McGrath, J. E. Polym. Preprints, ACS 28(2) (1987): 217. 10. Arnold, C. A.; Summers, J. D.; Chen, Y.P.; Chen, D. H.; Graybeal J. D. and McGrath, J. E., SAMPE Proc 33 (1988): 960. 138

11. Arnold, C. A.; Summers J. D. and McGrath, J. E., Polym. Eng Sci 29(20), (1989): 1413. 12. Kreuz, J. A.; Endrey, A. L.; Gay, F. P. and Sroog, C. E. Journal of Polymer Science: Part A-1 4 (1966): 2607-2616. 13. Brekner, M.-J. and Feger, C. J., Polymer. Sci., Polym. Chem. Ed. 25 (1987): 2005. 14. Facinelli, J. F. The Preparation of High Performance Polymer for Composites and Blends: a) Thermally Stable Ion Containing Polymers b) Epoxy and Hydroxy Functional Polyolefin Macromers, Ph.D. dissertation, VPI&SU, 1996. 15. Facinelli, J. V.; Gardner, S. L.; Dong, L.; Sensenich, C. L.; Davis, R. M. and Riffle, J. S., Macromolecules, 29(23) (1996): 7343. 16. Fitzgerald, J. J.; Tunney, S. E. and Landry, M. R., Polymer 34(9) (1993): 17. Fox, T.G and Flory, P.J. J. Appl. Phys. 21 (1950): 581. 18. Polyimides, Wilson, D.; Stenzenberger, H. D.; Hergenrother, P. M., ed.; New York: Chapman and Hall, 1990. 19. Farr, I., Synthesis and Characterization of Novel Polyimide Gas Separation Membrane Material Systems, Ph. D. Dissertation, VPI&SU, 1999. 20. York, G.A. Structure Property Relationships of Multiphase Copolymers, Ph.D. Dissertation, VPI&SU, 1990. 21. York, G.A.; Waldbauer, R.O.; Arnold, C. A.; Rogers, M. E., Gungor, A., Rodrigues, D., Wilkes, G. L., and McGrath, J. E. “Investigation of the

Morphological Structure and Physical Behavior of High Tg Polyimide Thermoplastic Homopolymers and Siloxane Segmented Copolymers,” 35th International SAMPE Symposium (1990), pp. 579. 22. Noshay, A., McGrath, J. E. “Block Copolymers: Overview and Critical Survey,” New York: Academic Press, 1977. 23. Keohan, F.L.; Swint, S.A.; Buese, M. A., J. Polym Sci: Part A: Polym Chem 29 (1991): 303. 24. McGrath, J. E.; Dunson, D. L.; Mechum, S. J.; and Hedrick, J. L., “Synthesis and Characterization of Segmented Polyimide-Polyorganosiloxane Copolymers,” Adv. Polym. Sci., 140 (1999): 61. 139

CHAPTER 4. CONTROLLED MOLECULAR WEIGHT END-FUNCTIONAL POLYIMIDES FROM POLY(AMIC ACID) SALT PRECURSORS

4.1 SYNOPSIS

Previous work by John Facinelli1 established that controlled molecular weight BPADA/MPDA polyimides could be prepared from poly(amic acid) salts. Aqueous solutions of these salts were shown to be hydrolytically stable, whereas poly(amic acid)s are well known to degrade rapidly in water. This is thought to be a result of acid catalyzed hydrolysis. It is hypothesized that the stability of the poly(amic acid) salt is due to the absence of the acid proton. Synthesis of the poly(amic acid) precursors proceeded rapidly by a homogeneous method in THF.1,7 Interestingly, amidization was observed to proceed much more rapidly in THF relative to NMP, but the reasons for this remain unclear. This synthetic approach has now been extended to include polyimide oligomers with functional endgroups. The reactive endgroups of these tough, ductile thermoplastics allow them to be easily incorporated into various structures. In this chapter, the preparation of controlled molecular weight, hydroxy and amine terminated BPADA/MPDA polyimides via poly(amic acid) salt precursors will be discussed. The tripropylammonium salts formed stable aqueous solutions. These salts imidized rapidly in the melt to form stable polyimides with 5% weight losses well above 400°C. 140

4.2 EXPERIMENTAL

4.2.1 Materials and Their Purification

The utilization of pure reagents and solvents is vital for synthesis of controlled molecular weight polyimides. Therefore, the monomers and reagents were carefully purified and dried before use.

Tetrahydrofuran (THF) THF, (EM Sciences) was refluxed over sodium with a small amount of benzophenone as an indicator for dryness (which formed the purple sodium benzophenone ketyl). The solvent was distilled immediately prior to use. (b.p. 67°C) O

Toluene Toluene (Aldrich) was dried over calcium hydride overnight and distilled immediately prior to use. (b.p. 110°C)

CH3

Meta-phenylene diamine (MPDA) m-PDA, (Aldrich) was sublimed under vacuum using an 80°C oil bath and stored under nitrogen in the dark prior to use. (m.p. 64-66°C)

H2N NH2 141

2,2’-Bis[4-(3,4-dicarboxyphenoxy)phenyl]propane dianhydride (bisphenol- A dianhydride, BPADA)

BPADA (General Electric) was recrystallized from toluene (molar ratio 4:1 toluene:BPADA) and acetic anhydride (molar ratio 3:1 BPADA:acetic acid) under nitrogen. After filtering, the crystals were refluxed to dissolve them in toluene (molar ratio 4:1 toluene:BPADA) and acetic anhydride (molar ratio 15:1 BPADA:acetic anhydride). The second recrystallization process was repeated twice then the crystals were washed with fresh toluene and n-heptane. The purified BPADA was dried in a vacuum oven at 140°C overnight immediately prior to use. (m.p. 189.1-191.1°C) O O CH3 O C O O O CH3

O O

3-Aminophenol 3-Aminophenol (Aldrich), was recrystallized from dry toluene. The crystals were washed with fresh toluene and dried under vacuum. (m.p. 123.4-125.2°C)

HO NH2

4.2.2 Synthesis of BPADA/MPDA Amine Terminated Poly(amic acid)

An excess of MPDA (stoichiometrically determined by the Carother’s equation) was used to control the molecular weight and endcap the poly(amic acid)s. A representative procedure is discussed for an amine terminated poly(amic acid) with a Mn » 5,000g/mol. BPADA (20.000 g, 38.4 mmol) was 142 weighed into a dry 500 ml round bottom flask containing a magnetic stir bar. This flask was then sealed with a rubber septum and purged with nitrogen. Dry THF (300 ml) was added to the flask by cannula to dissolve the monomer. Then, m- PDA (4.288 g, 39.7 mmol) was added to a 100 ml round bottom flask. This flask was also sealed with a rubber septum and purged with nitrogen. Dry THF (50 ml) was added to the flask by cannula and the solution was stirred until dissolved. The solution was then added to the BPADA solution by cannula. The 100 ml flask was rinsed twice with 20 ml of dry THF which was also added to the reaction. The solution viscosity increased almost immediately. After 24 hours, the poly(amic acid) solution was poured into a teflon dish and placed inside a covered glass container equipped with a nitrogen sweep. The polymer solution was exposed to a brisk nitrogen flow at room temperature for several days to evaporate most of the solvent. The resulting amine terminated poly(amic acid) was then dried further for 48 hours under high vacuum at 20°C. A tough yellow glass resulted. Target molecular weights of 10,000 and 20,000 g/mol were prepared in the same manner. The amic acids were analyzed by 1H NMR to determine residual solvent content, and were then analyzed using size exclusion chromatography to determine the absolute molecular weight distributions.

4.2.3 Synthesis of Hydroxy Terminated BPADA/MPDA Poly(amic acid)

3-Aminophenol was used as an endcapping reagent to prepare the hydroxy terminated poly(amic acid)s. A representative procedure is discussed for a hydroxy terminated poly(amic acid) with a Mn » 5,000 g/mol. BPADA (18.7268g, 36.1 mmol) was weighed into a dry 500 ml round bottom flask containing a magnetic stir bar. This flask was then sealed with a rubber septum and purged with nitrogen. Dry THF (~200 ml) was added to the flask by cannula to dissolve the monomer. The m-PDA (3.439 g, 31.8 mmol) and the endcapping reagent 3-aminophenol (0.4954 g, 0.00454 mmol) were added to a dry 250 ml round bottom flask. This flask was also sealed with a rubber septum and purged 143 with nitrogen. Dry THF (100 ml) was added to the flask by cannula and the solution was stirred until dissolved. The solution was then added to the BPADA solution by cannula. The 250 ml flask was rinsed twice with 20 ml of dry THF. The washings were also added to the reaction flask in the same manner. After 24 hours, the poly(amic acid) solution was poured into a teflon dish and placed inside a covered glass container equipped with a nitrogen sweep. The polymer solution was exposed to a brisk nitrogen flow at room temperature for several days to evaporate most of the solvent. The resulting hydroxy terminated poly(amic acid) was then dried further for 48 hours under high vacuum at 20°C. A tough clear light yellow glass resulted. Target molecular weights of 10,000 and 20,000 g/mol were prepared in the same manner. The amic acids were analyzed by 1H NMR to determine residual solvent content and size exclusion chromatography (SEC) for absolute molecular weight distributions.

4.2.4 Determination of Weight % Solvent in Poly(amic acid)

To reduce the possibility of error in molecular weight determination and poly(amic acid) salt preparation, the weight % solvent in the poly(amic acid) samples was determined utilizing 1H NMR on a Varian 400 MHz instrument. The peak area due to the 4 THF protons at 3.6 ppm was compared to the peak area resulting from the 2 amide protons per poly(amic acid) repeat unit at 10.4 ppm. This method proved to be accurate in the determination of the residual THF in the poly(amic acid) samples. Approximately 19-22 weight % remained in the poly(amic acid) samples.

4.2.5 Preparation of Poly(amic acid) Salts

The tripropylamine base was titrated with 0.025 N HCl that was previously standardized against Na2CO3. The poly(amic acid) salt was prepared as follows: 144

To a 500 ml round bottom flask containing the poly(amic acid) (22.950 g, 36.5 mmol repeat units) was added a solution of methanol (350 ml) and tripropylamine (18.09 ml, 80.3 mmol). The volume of tripropylamine added was a 10 mole % excess relative to the carboxylic acid groups. The mixture was stirred until homogeneous, indicating complete conversion to the salt. The solvent was then evaporated under a nitrogen flow. The poly(amic acid) salts were further dried under vacuum at 20°C for 48 h. The poly(amic acid) salts were analyzed by thermogravimetric analysis (TGA).

4.2.6 Thermal Imidization of Poly(amic acid) Salts

The functionalized poly(amic acid) salts were imidized under nitrogen at 260°C for 10 min. Polyimides resulting from the imidization were analyzed by size exclusion chromatography (SEC), differential scanning calorimetry (DSC), and thermogravimetric analysis (TGA).

4.3 CHARACTERIZATION

4.3.1 Nuclear Magnetic Resonance

1H NMR analysis of the poly(amic acid)s and polyimides was performed on a 400 MHz Varian Unity NMR spectrometer using d-chloroform or d6- dimethylsulfoxide.

4.3.2 Molecular Weight Determination

The molecular weights of the hydroxy terminated polyimides were determined by SEC in N-methylpyrrolidone (NMP) (dried over phosphorus pentoxide) at 60oC. A Waters 150C instrument equipped with differential refractive index and viscosity (Viscotek Model 100 Differential Viscometer) 145 detectors in parallel was utilized. Absolute molecular weight averages were calculated by the “universal calibration” procedure that assumes the hydrodynamic volumes of solute molecules appearing at the same elution volume were equal. Since differential refractive indices were used to determine solute concentration, it was further assumed that the average refractive index of the block copolymer molecules appearing at a given elution volume was the same as that appearing at any other elution volume. The molecular weights of the amine terminated polyimides were determined by potentiometric titration (based on the amount of amine endgroups present) with 0.025 N HBr (standardized against KHP). Titrations were performed using a MCI GT-05 (COSA Instruments Corp.) automatic potentiometric titrator. Between 0.25 and 0.3 grams of polyimide were dissolved in 50 mL of chloroform (EM Science HPLC Grade) with stirring. After the samples were completely dissolved, 25 mL of glacial acetic acid (Fisher Scientific ACS Plus) was added with stirring. The samples were potentiometrically titrated with a solution of HBr in methanol (0.02581 N) standardized against potassium hydrogen phthalate. The endpoints of the titration were calculated by the auto- titrator as the inflection point in the potential vs. volume titrant plot. Three titrations were performed for each polyimide sample and the results were averaged. The molecular weight of the polyimides were calculated as follows:

Ws(g)*2 < Mn >= (1) Nt(mol / L)*Vt(L)

where Ws was the polyimide sample weight and Nt and Vt were the titrant normality and titration volume, respectively.

4.3.3 Thermogravimetric Analysis

Dynamic thermogravimetric analysis (TGA) was performed to determine the relative thermal stability of the polyimides using a Perkin Elmer TGA 7 146 thermogravimetric analyzer. Samples of the polyimides obtained via the amic acid route and those obtained via the amic acid salt route were placed in a platinum pan connected to an electric microbalance. The samples were heated at a rate of 10°C per minute in a flowing air atmosphere. Weight loss of the sample was measured as a function of time and temperature. The thermal stabilities of the polymers were reported as the 5 weight % loss temperatures. In addition, isothermal TGA was performed on poly(amic acid) salts to assess their relative imidization times.

4.3.4 Differential Scanning Calorimetry

Differential scanning calorimetry (DSC) was used to determine the glass transition temperatures of the polyimides using a Perkin-Elmer DSC 7. The samples were run using a heating rate of 10°C per minute from 25°C to 250°C.

The Tg values were obtained from the second heating cycle after rapid cooling from above the Tg of the polymer. The values were taken as the midpoint temperatures of the change in specific heat in the transition region.

4.4 RESULTS AND DISCUSSION

An excess of MPDA (stoichiometrically determined by the Carother’s equation2) was used to control the molecular weight and endcap the amine terminated poly(amic acid)s (Figure 4-1), and a calculated amount of 3- aminophenol was used to endcap the hydroxy terminated poly(amic acid)s (Figure 4-2). The molecular weights of the hydroxy terminated poly(amic acid)s were determined using SEC. The polymers exhibited unimodal distributions with polydispersities close to 2.0, which is expected for step growth polymerization polymers. 147

O O CH3 H2N NH2 O C O O O + CH3

O O (excess) Dry THF (room temperature)

O O CH3 H2N HN NH NH2 O C O

HO CH3 OH

O O n Figure 4-1. Synthetic route to amine terminated BPADA/MPDA poly(amic acid) 148

O O CH3 H2N NH2 O C O + O O CH3

O O

HO NH Dry THF 2 (room temperature)

O O O O CH CH HO HN 3 NH NH 3 NH OH O C O O C O

HO CH3 OH HO CH3 OH O O O O n

Figure 4-2. Synthetic route to phenol terminated BPADA/MPDA poly(amic acid) 149

The weight % THF remaining in the poly(amic acid)s was determined using 1H NMR according to the calculations shown in Figure 4-3. Approximately 19-22 weight % THF was calculated for each poly(amic acid). The theoretical amount of THF was approximately equal to 1 THF per 1 –COOH group. This value matched closely to that obtained by Facinelli3 of 19 weight % for analogous structures. BPADA/MPDA trialkylammonium salts have been previously shown to thermally imidize to yield polyimides with controlled molecular weights.3 However, the water solubility of the triethylammonium poly(amic acid) salt was low for high molecular weight poly(amic acid) precursors. It was proposed that the hydrated ionic diameter of the tripropylammonium ion would provide good water solubility while providing excellent molecular weight control.1,4,6 The hypothesis was that the steric bulk of this larger cation would not pack well against the poly(amic acid) backbone, and that it might allow for better water solvation. Thus, in this study, conversion of the poly(amic acid)s to aqueous or methanolic salt solutions was carried out using a 10 mole % excess of tripropylamine in methanol (Figure 4-4). As predicted, the resultant polymeric salts were all soluble and stable in water. Using size exclusion chromatography, the molecular weights were compared for a polyimide where the salt had been dissolved in water for 36 h and one that had not been exposed to water (Table 4-1). The results showed no significant difference in the polyimide molecular weights, thus suggesting these materials form stable aqueous salt solutions. Further evidence was demonstrated through time dependent viscosity measurements of the aqueous salt solutions, which did not change over a 24 h period. These salts were imidized in less than 10 minutes at 260°C under

1 nitrogen. H NMR analyses of the polyimides confirmed that imidization under these conditions was complete since there was no peak associated with the amide protons at 10.4 ppm. Further confirmation was demonstrated from thermogravimetric analyses which showed no weight losses for the imidized 150

Area due to 4 THF protons (3.6 ppm) = 21.24 21.24 72g = 5.31* = 382.32g 4 1mol Area due to 2 amide protons of the poly(amic acid) = 5.22 4.81 628g = 2.40* = 1510.34g 2 1mol 382.32 Weight % THF = *100 = 20.2% (3.82.32 + 1510.34)

Figure 4-3. Sample calculation for residual THF weight % remaining in poly(amic acid) sample 151

O O CH HN 3 OH O C O

HO CH3 NH O O :B (base, tripropylamine) Methanol (or water)

O O + CH :BH HN 3 O- O C O - O CH :BH+ 3 NH O O

Figure 4- 4. Formation of BPADA/MPDA poly(amic acid) salt from poly(amic acid) precursor. 152

Table 4-1. Molecular weight results for polyimides from stable poly(amic acid) salt solutions

(g/mol) (g/mol) molecular weight of imide 10,600 19,700 whose salt was not exposed to water molecular weight of imide 11,200 20,600 whose salt was exposed to water for 36 h 153 materials until above 400°C. TGA also indicated that imidization proceeded relatively rapidly at temperatures just above Tg. One of the poly(amic acid) salts was heated at 200°C/min to 260°C where it was held isothermally (Figure 4-5). Weight loss due to imidization was complete within about 3 minutes and occurred at approximately 250°C which was 40°C above the polyimide Tg (210°C). The salts imidized to give controlled molecular weight polyimides very close to the targets of 5, 10, and 20,000 g/mol. Likewise, the polyimides resulting from poly(amic acid) precursors (as well as the poly(amic acid) itself) had controlled molecular weights (Table 4-2). The molecular weights of the amine terminated polyimides were determined by potentiometric titration (based on the amount of amine endgroups present), and the molecular weights of the hydroxy terminated polyimides were determined by size exclusion chromatography.

It is well known that Tg’s of linear polyimides with molecular weights between 5,000 and 20,000 g/mol increase with . This trend was also seen with the end-functional polyimides with Tg’s ranging from 190°C to 200°C (Table 4-3). The amino-functional materials exhibited slightly reduced thermal stabilities when compared to the hydroxy-functional polyimides. This occurrence is suspected to be a result of the high temperature reactivity of the basic amine endgroups. Nevertheless, each of the polyimides showed a 5% weight loss well above 500°C indicating excellent thermal stability (Table 4-3).

The molecular weight dependence of the Tg can be examined through the ¥ free-volume theory relationship, Tg = Tg - c/ (plotted as Tg versus the

5 ¥ reciprocal of in Figure 4-6). Tg is the value of the glass transition temperature for a linear polymer with infinitely few chain ends, or infinitely high

. Therefore, both the hydroxy and amine terminated polyimides should in

¥ theory have the same Tg value. Extrapolation of the Tg vs. 1/Mn (for the polyimides from poly(amic acid) salt precursors) plots yielded approximately the same Tg¥ value, 224°C for the amino-functionalized material and 227°C for the 154

Temperature °C 45 100 150 200 260

100

80

60

40 Weight % Remaining 20

0 0 1 2 3 4 5 6 7 8 9 10 Time (min) Figure 4-5. Isothermal TGA analysis of an end-functional poly(amic acid) salt demonstrating complete imidization within about 2-3 min. 155

Table 4-2. Molecular weight results for end-functional polyimides from amic acid and amic acid salt precursors

Endgroup (poly(amic acid)) (polyimide (polyimide from (g/mol) from amic acid amic acid salt precursor) precursor) (g/mol) (g/mol) 4800 6400 5,900 Hydroxy 14,500 13,300 13,900 18,600 20,500 20,300

5830 6130 4,400 Amine 10,000 10,700 10,500 21,900 22,100 22,000 156

Table 4- 3. Thermal properties of end-functional BPADA/MPDA polyimides

Endgroup 5% Weight Loss Tg (°C) (°C) 5,900 535 189 Hydroxy 13,300 537 210 20,300 546 216

4,400 521 190 Amine 10,500 516 206 22,000 544 220 157

250

240

230

220

210 Amine Terminated y = -15.602x + 224.47 R2 = 0.9536 200 Tg (°C)

190

180 Hydroxy Terminated y = -22.406x + 226.95 R2 = 1 170

160

150 0.00 0.50 1.00 1.50 2.00 2.50 -4 1/Mn (x10 )

Figure 4-6. Glass transition temperature vs. 1/Mn for end-functional BPADA/MPDA polyimides 158

¥ hydroxy-functionalized material. These were slightly lower than the Tg value of ~237°C reported by Facinelli et al. 3

4.5 CONCLUSIONS

Controlled molecular weight end-functional UltemÔ type polyimides have been prepared from poly(amic acid) salt precursors. The salts of these materials formed stable aqueous solutions and imidized rapidly in the melt at temperatures slightly above their Tg. The resultant polyimides were thermally stable with 5 % weight loss temperatures above 500°C.

4.6 ACKNOWLEDGEMENTS

Appreciation is extended to the Corporate Research and Development Center of the General Electric Co. for their generous supply of BPADA.

4.7 REFERENCES

1. Facinelli, J. V., “The Preparation of High Performance Polymers for Composites and Blends: a) Thermally Stable Ion Containing Polymers b) Epoxy and Hydroxy Functional Polyolefin Macromoners,” Ph.D. Dissertation, VPI&SU, 1996. 2. Carothers, W. H., Trans. Faraday Soc. 32 (1936): 39. Odian, G., Principles of Polymerization; 3rd Ed., New York: John Wiley & Sons, 1991, pp. 110. 3. Facinelli, J. V.; Gardner, S. L.; Dong, L.; Sensenich, C. L.; Davis, R. M.; and Riffle, J. S., Macromolecules 29(23) (1996): 7342. 4. Kielland, J., Journal of American Chemical Society 59 (1937): 1675-78. 5. Fox, T. G. and Flory, P. J., J. Appl. Phys. 21 (1950): 581. 159

CHAPTER 5. CHARACTERIZATION OF POLY(SILOXANE IMIDE) BLOCK COPOLYMER SURFACES

5.1 SYNOPSIS

Microphase separation of poly(siloxane imide) block copolymers occurs even at relatively low block lengths due to the dissimilarity between the chemical structures of the polydimethylsiloxane (PDMS) and polyimide blocks.1 Surface modifications rendered by the incorporation of the PDMS component are of particular interest since small bulk concentrations of PDMS blocks can result in rather dramatic surface enrichment. In this chapter, the surface characterization of BPADA/MPDA and fluorinated poly(siloxane imide) block copolymers will be discussed in view of systematically varied block copolymer compositions. Since the surface free energy of PDMS is very low, the siloxane segments migrate to the more hydrophobic top (air) surface to form a silicon enriched layer. Angle-dependent x- ray photoelectron spectroscopy (XPS) proved to be an efficient method for analyzing the surface migration. Furthermore, water contact angle analyses indicated a more hydrophobic low energy surface. The topography of the microphase separated block copolymer surface was investigated using Atomic Force Microscopy (AFM). 160

5.2 EXPERIMENTAL

5.2.1 Materials

UltemÔ Type Poly(dimethylsiloxane imide) Block Copolymers The UltemÔ type poly(dimethylsiloxane imide) block copolymers used in this study (Table 6-1) were synthesized as described in Chapter 3.

O O O O CH3 CH3 R R CH2 Si O Si CH2 N N N N 3 n 3 CH3 CH3 O O O O x y

CH3 R = O O

CH3

Table 5-1. Compositions of BPADA/MPDA block copolymers used in the surface characterization study

PDMS (g/mol) Weight % PDMS 1000 10 1000 20 2000 10 2000 70 161

6FDA/MPDA Poly(siloxane imide) Block Copolymers The 6FDA/MPDA poly(siloxane imide) block copolymers used in this study (Table 5-2) were synthesized using 6FDA and m-phenylene diamine according to the procedure discussed in Chapter 3.

O O O O CH3 CH3 R R CH2 Si O Si CH2 N N N N 3 n 3 CH3 CH3 O O O O x y

CF3 R = C

CF3

Table 5- 2. Compositions of fluorinated block copolymers used in the surface characterization study

PDMS (g/mol) PDMS Weight % 1000 30 2000 10 2000 20 2000 30 5000 10 5000 30 11000 10 11000 50

Amine Terminated BPADA/MPDA Polyimide

The amine terminated BPADA/MPDA polyimide (» 20,000 g/mol) used in this study were synthesized as described in Chapter 4. 162

O O CH3 O C O H2N N N NH2 CH3 O O n

Fluorinated Polyimide

The fluorinated polyimide ( » 27,000 g/mol) used in this study was synthesized using 6FDA and meta-phenylene diamine according to the procedure outlined in Chapter 4. The Tg was found to be approximately 280°C.

O O CF3 C N N CF3

O O n

5.2.2 Preparation of Polymer Films for Surface Analysis

The block copolymers were soxhlet extracted with isopropanol for 2 days and dried at 100°C under vacuum for 2 hours to remove any residual unreacted PDMS. Dilute chloroform solutions (approximately 2-5 weight % solids) were prepared and vacuum filtered. Some of the solvent was removed via rotavap to yield a more viscous casting solution. Films of the copolymers were cast directly onto glass substrates for AFM and water contact angle analyses. Likewise, copolymer films were cast directly onto the sample probes for XPS. The cast films were dried with the aid of a heat lamp overnight under a nitrogen environment. Then they were heated under vacuum to just below the upper glass transition temperature (150-200°C) and held there for 2 hours before cooling to room temperature. 163

5.2.3 Copolymer Composition

The composition of the extracted and unextracted poly(siloxane imide) block copolymers were determined using NMR on a Varian 400 MHz instrument. 1H NMR was used for the BPADA/MPDA block copolymers. The peak area due to the 6 silicon methyl protons per soft segment repeat unit at 0 ppm was compared to the peak area resulting from the 6 isopropylidine methyl protons per hard segment repeat unit at 1.7 ppm. 13C NMR was used to analyze the block copolymers prepared with 6FDA. The peak area due to the 2 silicon methyl carbons per soft segment repeat unit at 0 ppm was compared to the peak area resulting from the tertiary carbon of the fluoro isopropylidine group per hard segment repeat unit at 62 ppm. This method proved to be effective for the determination of the relative block compositions. The results obtained were in agreement with the theoretical compositions. A sample calculation is shown in Figure 5-1 for a BPADA/MPDA block copolymer containing 80 weight % imide (PDMS = 2000 g/mol).

Area due to 6 methyl protons of the hard block (1.7 ppm) = 22.71 22.71 628g = 3.785mol * = 2376.98g 6 1mol Area due to 142 methyl protons of the soft block (0 ppm) = 41.54 41.54 2000g = 0.2925mol * = 585.07g 142 1mol 2376.98 Weight % Hard Block = *100 = 80.2% (2376.98 + 585.07)

Figure 5-1. Sample calculations for the determination of block copolymer compositions. 164

5.3 CHARACTERIZATION

5.3.1 X-Ray Photoelectron Spectroscopy (XPS)

Surface analysis of the poly(dimethylsiloxane-imide) block copolymer films were obtained on a Perkin Elmer X-ray Photoelectron Spectrometer 5400 Series using a Mg anode operating at 300 watts (14kV). The area analyzed by the input lens of the analyzer was 1mm x 1mm and the pressure of the system was below 5x10-6 Pa. Spectra were obtained using angle dependent XPS at 15º, 30º, and 45º take-off angles. The sampling depth (d) is related to the take-off angle (q) according to the following relationship: d = 3lsinq, where l is the inelastic mean free path of emitted photoelectrons.2 All binding energies were referenced to the carbon C-H photopeak at 285.0 eV. Atomic percent calculations were obtained and used in the estimation of surface compositions. Based upon the structures of the repeat units the following relationships were derived for the (1) BPADA/MPDA block copolymers and (2) fluorinated block copolymers.

C/Si = [2X + 41(1-X)] / X (1)

C/Si = [2X + 25(1-X)] / X (2) where C/Si is the atomic ratio of carbon to silicon obtained by XPS measurements and X is the bulk PDMS molar fraction of the samples. Using X the PDMS weight percentage was calculated according to the formula weights of the two repeat units.

Wu = 74X / [74X + 640(1-X)] * 100 (3)

Wf = 74X / [74X + 516(1-X)] * 100 (4) 165

where Wu and Wf are the PDMS weight percentages for BPADA/MPDA and fluorinated block copolymers, respectively. The surface composition calculated from XPS measurements is a complex average over the entire span of a surface region up to the sampling depth. Since the XPS photoemission intensity is attenuated exponentially with depth, the materials in the upper surface region make more contribution to the measurement than the inner materials of the same volume. 3 All of the XPS data presented are the complex average values without deconvolution.

5.3.2 Water Contact Angle Analysis

The water contact angle of the poly(siloxane-imide) films were measured using a Dage 550 Series Goniometer. Five microliter deionized water droplets were placed on the films and the contact angles (q) were measured at room temperature within 2 minutes (Figure 5-1). Five measurements were taken for each sample and averaged.

q

Poly(siloxane- imide) Film Figure 5-2. Setup for water contact angle analyses.

5.3.3 Atomic Force Microscopy

The block copolymer surfaces were analyzed using Atomic Force Microscopy (AFM) on a Digital Instruments’ Dimension 3000 scanning probe 166 microscope using the Nanoscopeâ IIIa controller. Measurements were performed at room temperature in air using the Tapping ModeÔ. The etched silicon tips had spring constants of ~20-1000 Nm-1 and resonance frequencies around 300 kHz. The images were collected with a free air amplitude of ~2.6 V using a set point of ~1.75 V (Ao/Asp » 0.67). The phase and height images are shown as raw data with minor flatness and tilt corrections. The mean surface roughness, which is the mean value of the surface relative to the center plane, was calculated using Nanoscopeâ IIIa software from the following equation

Ly Lx Ra = (1/ (Lx * Ly)) f (x, y) dxdy (5) ò0 ò0 where f (x,y) is a function of the surface relative to the center plane, and Lx and Ly are the dimensions of the surface.4 All of the mean roughness values were averaged from approximately three 1x1 mm images of a 5x5 mm scan.

5.4 RESULTS AND DISCUSSION

5.4.1 Copolymer Composition

The poly(siloxane imide) block copolymers were extracted with isopropanol to remove any unincorporated siloxane prior to surface analyses. After extraction, the copolymer compositions were assessed using 1H NMR (Figure 5-3) for the BPADA/MPDA block copolymers and 13C NMR (Figure 5-4) was used for the fluorinated block copolymers. The extracted results are compared to the values obtained for the unextracted in Table 3. The data suggest that no significant level of unincorporated PDMS oligomer was present in any of the copolymers since there was little difference between the predicted and measured compositions following extraction. 167

Table 5-3. Copolymer composition before and after extraction

Copolymer PDMS Weight % Weight % Weight %

PDMS PDMS PDMS (g/mole) (theoretical) (before) (after) 1000 20 17 16 1000 10 9.9 9.7 BPA-DA/m-PD/PDMS 2000 70 66 65 2000 50 50 50 2000 10 8 8

1000 30 32 32 1000 10 12.3 13 2000 30 32 33 6F-DA/m-PD/PDMS 2000 20 20 21 2000 10 13 13 5000 30 28 29 5000 10 9.3 10 11000 50 50.5 51 11000 10 11.5 11 168

a f g d O O O CH 3 O CH 3 CH 3 C C C O C O C NH N R1 N CH2 Si O Si N N H 3 x a CH CH C 3 3 i CH 3 e,c,b O y O O z h j k

f, g, i d

h

k

e c b

10 8 6 4 2 0 ppm

Figure 5-3. 1H NMR spectrum of a BPADA/MPDA poly(siloxane imide) block copolymer. 169

O O O CF3 O a CF3 l d CH3 CH3 n e i b,c,d C C CH Si O Si CH N N N N 2 3 2 3 CF n 3 m k CF3 CH3 CH3 g j O O O O x f y

c

m,n i,j d b e a l g k h f

150 100 50 PPM

Figure 5-4. 13C NMR spectra of a 6FDA/MPDA poly(siloxane imide) block copolymer 170

5.4.2 XPS Surface Analysis of Block Copolymer Films

The compositions of the block copolymer film surfaces were investigated using angle dependent XPS. Figure 5-5 shows XPS wide scans for selected BPADA/MPDA and 6FDA/MPDA block copolymers and their polyimide controls (0% PDMS). As apparent from the spectra no detectable amount of silicon was present in the polyimide control samples. Narrow scans were conducted on the elements obtained in the wide scans and their binding energies5 (BE) (Table 5-4) were used in establishing the surface chemical compositions. Table 5-5 contains the BEs observed in the XPS analysis of the polyimide controls and block copolymers presented in Figure 5-5. The experimental and theoretical values match closely, thus verifying the surface composition. Structurally the block copolymers at different compositions are identical, thus the data presented is representative of each series. Table 5-6 presents the surface compositions determined by XPS and the calculated bulk compositions. The following observations can be made from the results:

1) The BPADA/MPDA polyimide exhibited excellent agreement between the calculated bulk and experimental elemental compositions. 2) The experimental BPADA/MPDA block copolymers atomic compositions showed a higher Si atomic % and lower C atomic % than the bulk composition; however, the O atomic % and N atomic % compositions for the bulk and experimental were in agreement. 3) The 6FDA/MPDA polyimide showed a higher experimental F atomic % and lower C atomic % than the bulk atomic composition; however, the O atomic % and N atomic % bulk and experimental values were in agreement. 4) The 6FDA/MPDA block copolymers exhibited lower C , N, and F atomic percentages and higher O and Si atomic percentages than the bulk compositions. 171

(a) 10

8 -C 1s

6 -C KLL

-O 1s

-O KVV N(E)/E 4

-N 1s

2

-O 2s 0 1000 800 600 400 200 0 Binding Energy, eV

(b) 10

-C 1s 8

-C KLL 6

-O KVV

2p

-N 1s

N(E)/E 4

2s

Si

-

Si

- 2

-O 2s

O 1s-

0 1000 800 600 400 200 0 Binding Energy, eV 172

(c) 10

-O 1s

-C 1s 8 -C KLL

-O KVV

6 N(E)/E 4 -N 1s

F 1s- 2

F KLL-

-O 2s 0 1000 800 600 400 200 0 Binding Energy, eV

(d) 10

-O 1s

-C 1s 8

-O KVV

6

C KLL-

-N 1s

N(E)/E 4

F 1s- 2s

2p

Si

Si

- 2 -

F KLL-

-O 2s 0 1000 800 600 400 200 0 Binding Energy, eV

Figure 5-5. XPS wide scans of (a) BPADA/MPDA polyimide, (b)BPADA/MPDA block copolymer (20% PDMS, » 2000 g/mol), (c) 6FDA/MDA polyimide, and (d) 6FDA/MPDA block copolymer (10% PDMS » 2000 g/mol) 173

Table 5-4. XPS binding energies (BE) and chemical states5

Peak BE (eV) Chemical State 284.8-285.2 C-C, C-H

C1s 286.0-286.4 C-N 287.0-286.4 C=O (C-O) 287.0-290.2 C-F

O1s 530.5-532.0 C=O 532.6-532.7 O-Si

N1s 399.7-400.8 N-C

F1s 688.6-688.8 C-F

Si2p 101.9-100.5 Si-C, Si-O 174

Table 5-5. XPS Binding Energies (eV) for the polyimide homo- and copolymers referenced to C1s (285.0 eV) for the XPS wide scans in Figure 5-5.*

Element Ultem™ Ultem™ Fluorinated Fluorinated Polyimide Block Polyimide Block Copolymer Copolymer

O1s 532.5 532.6 532.5 532.4

N1s 400.8 400.6 400.4 400.6

Si2p ----- 102.4 ----- 102.4

F1s ------688.6 688.8 *The BPADA/MPDA block copolymer contains 20% PDMS of » 2000 g/mol and the fluorinated block copolymer contains 10% PDMS with » 2000 g/mol. 175

Table 5-6. Elemental surface compositions compared with the bulk stoichiometries

(Obtained From XPS Peak Areas) Polymer C % O % N % F % Si %

Calculated 45° Calculated 45° Calculated 45° Calculated 45° Calculated 45° BPADA/ 82.3 81.9 13.3 13.7 4.4 4.6 ------MPDA (0% Sx) BPADA/ 49.5 59.7 23.9 22.2 1.8 1.6 ------24.8 16.5 MPDA (20% Sx) 6FDA/ 67.6 65.3 10.8 10.1 5.4 5.1 16.2 19.5 ------MPDA (0% Sx) 6FDA/ 63.4 56.0 14.0 22.0 4.6 1.8 12.3 2.6 5.9 16.2 MPDA (10% Sx) *The BPADA/MPDA block copolymer contains 20% PDMS of » 2000 g/mol and the fluorinated block copolymer contains 10% PDMS with » 2000 g/mol. 176

The results reflected surface segregation of the polysiloxane components both block copolymer series, as expected. In addition, surface segregation of fluorine was observed for the polyimide homopolymer. Buchwalter6 reported similar results for fluorinated polyimides with CF3 pendant groups. This phenomenon was not observed for the fluorinated block copolymers since the

7 8 critical surface tension of PDMS is » 20.9 mN/m while it is »6 mN/m for –CF3 based materials. A plausible explanation is at room temperature the imide segments are below their Tg, thus these components are too immobile to migrate to the surface area. On the other hand, at room temperature the siloxane segments are well above Tg. This enables the mobile segments to migrate to the hydrophobic surface area. Observation of emitted electron intensities at the three take-off angles (15, 30, and 45°) permitted estimation of block copolymer composition as a function of depth. PDMS weight % in and near the surface was estimated from the atomic percentages and compared with the average bulk compositions. The results revealed that the PDMS weight % was much higher at all three sampling depths than its average bulk compositions in every case (Figure 5-6). Examining this phenomenon using block copolymers with even lower PDMS bulk compositions than 10 weight %, as presented in Figure 5-6, revealed similar results (Figure 5-7). Block copolymers containing 5 weight % PDMS (5,000 g/mol oligomer) and 4 weight % (11,000 g/mol oligomer) showed results similar to those block copolymers containing higher PDMS bulk compositions. The surface concentrations were 85% and 73% respectively. This demonstrates that at surfactant level bulk concentrations the PDMS segments are effective in forming a silicon enriched layer. This behavior is much like that observed for PDMS containing polymeric blends. Chen and Gardella9 investigated PDMS containing polycarbonate block copolymers. The authors reported that when the PDMS containing block copolymers were incorporated into blends at 6 weight % or less, the surface PDMS concentrations of the blends could be as high as 95%. When comparing similar bulk concentrations of PDMS, it became evident that the longer PDMS segments were more effective in increasing the surface 177

(a)

100

90

80

70

60

50

40

30

20 Weight % PDMS in Surface Region

10

0 10%(1K) 20%(1K) 10%(2K) 70%(2K) Bulk PDMS Concentration (Oligomer Molecular Weight) Ultem Type Block Copolymers

(b)

120

100

80

60

40

Weight % PDMS in Surface Region 20

0 10%(1K) 30%(1K) 10%(2K) 20%(2K) 30%(2K) 10%(5K) 30%(5K) 10%(11K) 50%(11K) Bulk PDMS Concentration (Oligomer Molecular Weight) Fluorinated Block Compolymers

15° 30° 45°

Figure 5-6. Angle dependent XPS data depicting PDMS surface concentrations in block copolymer films as a function of PDMS bulk weight % (a) UltemÔ type and (b) Fluorinated block copolymers at take-off angles of 15°, 30°, and 45° respectively. 178

100 90 80 70 60 50 40 30 20 PDMS Weight % in Surface Region 10 0 5%(5K) 4%(11k) Bulk PDMS Concentration (Oligomer Molecular Weight) Fluorinated Block Compolymers

15° 30° 45°

Figure 5-7. Angle dependent XPS data for fluorinated block copolymers containing 5 weight % (5000 g/mol) and 4 weight % (11000 g/mol) PDMS at take- off angles of 15°, 30°, and 45°. 179

100

90

80

70

60

50

40

30

20 Weight % PDMS in Surface Region 10

0 0 2 4 6 8 10 12 -3 PDMS Oligomer x 10

15 deg 30 deg 45 deg

Figure 5-8. Influence of PDMS oligomer molecular weight (at 10 weight % bulk concentration) on the surface composition of fluorinated block copolymers. 180

concentrations. For example, the fluorinated block copolymer prepared with a 1,000 g/mol PDMS oligomer (10 weight %) had approximately 76 weight % PDMS on the surface, whereas the block copolymer prepared with a 2000 g/mol oligomer (10 weight %) had approximately 86 weight % PDMS on the surface. This is likely related to better microphase separation as the length of the PDMS oligomer was increased. Figure 5-8 shows the influence of the PDMS oligomer size on the surface composition. It is noted that increasing the PDMS oligomer molecular weight above 2000 g/mol has little further effect on the surface composition. Dwight and coworkers10 reported analogous findings for perfectly alternating copolymers with bis-A sulfone, aromatic ester, urea and imide (BTDA/ODA) structures. The authors reported that the degree of phase mixing depended on the polysiloxane % and block lengths. Furthermore, the authors reported that a siloxane molecular weight between 6,800 and 12,800 was required to form a complete siloxane monolayer.

5.4.3 Water Contact Analysis of Block Copolymer Films

Further evidence in support of the siloxane surface enhancement was demonstrated using water contact angle analyses. High water contact angles are indicative of a hydrophobic low energy surface monolayer. As shown in Figure 5-9, the lowest polysiloxane composition investigated was 5 weight %

PDMS ( = 5000 g/mol) in the bulk. This significantly changed the surface free energy with a 25% increase in the water contact angle whereas increasing the weight % PDMS in the bulk further had little effect. Once the monolayer forms there is no further energetic driving force for surface segregation. Others have demonstrated that low PDMS bulk weight percentages yield similar results.11 PDMS has a reported water contact angle of approximately 101°.7 Phase separated PDMS containing block copolymer films show water contact angle values very close to that of pure PDMS. This suggests that the block copolymers 181

110.0

100.0

90.0 F-PSI(1K) F-PSI(2K) F-PSI(11K) 80.0 PSI(1K) PSI(2K) F-PSI(5K) 70.0 Water Contact Angle (degrees) 60.0

50.0 0 10 20 30 40 50 60 70 80 Bulk PDMS Weight %

Figure 5-9. Water contact angles of copolymer films as a function of PDMS content for 6FDA/MPDA (FPSI) and BPADA/MPDA (PSI) poly(siloxane imide) block copolymers with 1000, 2000, 5000 and 11000 PDMS oligomers. 182 containing 5000 g/mol and 11000 g/mol PDMS oligomer have better phase separation than those containing the smaller oligomers.

5.4.4 AFM Analysis of Block Copolymer Surfaces

The surface morphologies of the solvent cast fluorinated and UltemÔ type block copolymer films were characterized with Tapping ModeÔ AFM. Using moderate to hard tapping forces, the softer component (lower modulus) appears slightly darker than the harder component (higher modulus). Thus, the siloxane containing phases appear darker than the imide containing phases in the AFM phase image micrographs. Figure 5-10 contains representative AFM phase images for fluorinated homo and block copolymers containing 0 – 30 weight % PDMS ( » 2000 g/mol). The phase images show phase separated morphologies for the block copolymers containing 10 and 20 weight % and a relatively featureless morphology for the polyimide homopolymer. At 30 weight % the PDMS monolayer is believed to be thicker than at 10 and 20 weight percentages. Using the same cantilever force the AFM does not detect a significant difference in the modulus of the sample; therefore, the surface appears featureless. Although not pictured, similar results were observed for the BPADA/MPDA block copolymers. Not only can the block copolymer phase separation be observed as differences in the phase images, it can also be seen topographically. The surface roughness of the block copolymer films has been related to the degree of phase separation between the two blocks12 Therefore, the surface roughness should increase with increasing amount and molecular weight of the PDMS segments. This type of morphological difference is also influenced by the difference in the miscibility or compatibility between the PDMS and polyimide segments.12 Furukawa and coworkers12 investigated the surface topography of microphase separated DSDA/BAPP and BTDA/BAPSM based poly(siloxane imide) block copolymers using AFM (contact mode). The authors reported an 183

0 1mm 0 1mm 0 1mm 0% 10% 20% 30%

Figure 5-10. AFM phase images for fluorinated polymers with varying amounts of incorporated PDMS ( » 2000 g/mol) 184

(a) (b)

(c) (d)

Figure 5-11. AFM surface topographic images of BPADA/MPDA homo and copolymers with PDMS weight percentages of (a)0%, (b)10% ( »1000 g/mol), (c) 10% (Mn »2000 g/mol), and (d) 20% (Mn »2000 g/mol). (z-axis = 10 nm) 185

(a) (b)

(c) (d)

(e) (f)

Figure 5-12. AFM surface topographic images of fluorinated homo and copolymers with PDMS weight percentages of (a)0%, (b)10% ( »2000 g/mol), (c) 20% (Mn »1000 g/mol), (d) 20% (Mn »2000 g/mol), (e) 30% (Mn »1000 g/mol), and (f) 30% (Mn »2000 g/mol). (z-axis = 10 nm) 186 increase in surface roughness with 10 weight % PDMS bulk concentrations. Figures 5-11 and 5-12 contain AFM surface images of the homo and copolymers for the BPADA/MPDA and 6FDA/MPDA materials. The images pictorially demonstrate the influence of varying PDMS content and molecular weight. The PDMS domains form “islands” in the “sea-island” structure. With the addition of more PDMS or larger segments the domains stand up on the surface to produce steep hills and valleys. This behavior is similar to that in the formation of monolayer films, which by increasing the composition the layer thickness is increased (Figure 5-13). The observation of surface roughness can be expressed quantitatively using a roughness factor (Ra). This auto correlation function estimates the film surface roughness using the AFM data and expresses it in terms of nanometers. Table 5-7 contains the roughness calculation results for the copolymer films. As the surface images indicated, the roughness value increases with increasing PDMS content and in some cases molecular weight. This is related to better phase separation as observed with the XPS results. The relationship between the block copolymer composition and surface roughness can be pictured graphically as a plot of mean roughness versus PDMS weight percent (Figure 5- 14). The fluorinated block copolymers prepared with PDMS ( » 2000 g/mol) showed a linear relationship between the block copolymer composition and surface roughness.

5.5 CONCLUSIONS

Both XPS and water contact angle analyses demonstrated the surface enrichment of the copolymers with the low energy dimethylsiloxane component. The results showed that at bulk PDMS concentrations as low as 5 weight % there was a 25% increase in the water contact angle on the block copolymer film, thus indicating a more hydrophobic, low energy surface. The surface topography was 187 influenced by the degree of phase separation and was characterized using AFM. This was quantitatively represented with the roughness factor, Ra. It was found

PDMS Domains

Figure 5- 13. Diagram relating the block copolymer surface arrangement to the formation of monolayer films. Increasing the PDMS concentration in the polyimide-polysiloxane “sea-island” structure increases the PDMS domains as with the deposition of monolayers. 188

Table 5-7. Mean Surface Roughness of Block Copolymer Films

Type PDMS PDMS Weight % Mean Roughness (nm) (g/mol) 0 (control) 0 (control) 0.341 +/- 0.006 BPADA/MPDA/ 1000 10 0.880 +/- 0.059 PDMS 2000 10 0.887 +/- 0.113 2000 20 2.353 +/- 0.333

0 (control) 0 (control) 0.400 +/- 0.044 1000 20 0.507 +/- 0.039 6FDA/MPDA/ 1000 30 0.694 +/- 0.076 PDMS 2000 10 0.841 +/- 0.044 2000 20 1.654 +/- 0.199 2000 30 2.620 +/- 0.166 189

3.0

2.5

2.0

1.5

y = 0.0747x + 0.2578 R2 = 0.9751 1.0 Mean Roughness (Ra (nm))

0.5

0.0 0 10 20 30 40 Bulk PDMS Weight%

Figure 5-14. Graph of bulk PDMS weight % plotted against mean roughness (Ra) for the fluorinated block copolymers prepared with PDMS » 2,000 g/mol. The graph shows a linear relationship between the composition and surface roughness. 190 that the surface roughness increased with increasing PDMS oligomer content and molecular weight.

5.6 RECOMMENDATIONS FOR FUTURE WORK

One recommended area of investigation is to explore methods for tailoring the surface morphologies of the PDMS containing block copolymers by casting from selective solvents and exposing the materials to specified environments (eg. water). AFM can be used to characterize these surfaces in water. Another area would be to extend the contact angle study to include determination of the surface tension. This would yield more information about the acid/base properties of the surfaces. Young’s Equation, gs,g = gl,gcos(q) + gl,g, (where q is the contact angle, gs,g , gl,g, gl,g are the interfacial tensions of the solid/gas, liquid/gas, and solid/liquid interfaces) can be used to calculate the surface tensions. Drop contact angles on surfaces are influenced by surface roughness. Perhaps, correlation between surface roughness and the calculated surface tensions could be established. Additional recommended areas of investigation include an evaluation of the effects of film aging on the surface compositions.

5.7 ACKNOWLEDGEMENTS

Appreciation is extended to Dr. Kenneth Wynne and Steve Bullock of the Naval Research Lab for their informative discussion on characterization with Atomic Force Microscopy. Appreciation is also extended to Dr. Malcolm Robertson and Maggie Bump for help in AFM characterization. 191

5.8 REFERENCES

1. Chen, X., Lee, H. F.; and Gardella, Jr. J. A., Macromolecules 26 (1993): 4601. 2. Ratner, B. D. and Horbett, T. A., J. Coll. Interf. Sci. 83 (1981): 630. 3. Flory, P. J. and Erman, B., J. Polym. Sci.: Polym. Phys. Ed., 22 (1984): 49. 4. Digital Instruments Nanoscope™ Command Reference Manual, update version 4.10, Aug 1995. 5. Wagner, C. D.; Riggs, W. M. and Davis, L.E., “Handbook of Photoelectron Spectroscopy,” Mondla, J. F. and Muilengerg, G E., eds., Eden Prairie, MN: Perkin-Elmer Corporation Physical Electronics Division, 1979. 6. Buchwalter, L. P. J. Adh. Sci. Technol. 10(11) (1996): 1141. 7. Clarson, S. J. and Semlyen, J. A., eds., Siloxane Polymers, New Jersey: PTR Prentice Hall, 1993. 8. Zisman, W. A. in “Contact Angle, Wettability and Adhesion,” Adv. in Chemistry, ACS, vol. 43, Washington, DC, 1964, p. 1. 9. Chen, X. and Gardella Jr., J. A. Macromolecules 27 (1994): 12. 10. Dwight D., McGrath, J.E.; Lawson, G.; Patel, N.; York, G., “Surface and bulk microphase sepration in siloxane-containing block copolymers and their blends: The roles of composition and kinetics from multiphase macromolecular systems,” Contemp. Topics in Polym. Sci. 6 (1989): 265. 11. Tezuka, Y.; Ono, T and Imai, K, J. Coll. Interf. Sci. 136(2) (1990): 408. 12. Furkawa, N.; Yamada, Y.; Furukawa, M.; Yuasa, M; and Kimura, Y.; J. Polym. Sci. Part A: Polym. Chem 35(11) (1997): 2239. 192

CHAPTER 6. FUNCTIONAL POLYIMIDES AND POLY(SILOXANE IMIDE) BLOCK COPOLYMERS AS INTERPHASE MATERIALS

6.1 SYNOPSIS

Fluorinated and BPADA/MPDA polyimides and poly(siloxane imide) block copolymers have been prepared. Research has been devoted to the application of these materials to composites and electronic devices to serve as interphase materials. The first section of this chapter discusses the efforts toward characterizing water soluble UltemÔ type poly(amic acid) salts and polydimethylsiloxane (PDMS) containing poly(amic acid) salts as interphase agents for composites.1-5 Figure 6-1 contains the general synthetic approach used in the preparation of the composite interphase materials. Research efforts have been focused on the potential use of thermoplastic polymer fiber coatings as interphase agents to toughen fiber reinforced composites. The optimum design for composite fiber matrix interfaces has not been established. The hypothesis is that the interfacial material should meet the following requirements: ductile, applicable from water, and either miscible with, or reactive with the matrix to facilitate sizing-matrix resin interdiffusion and adhesion. Water solubility is considered an important processing advantage. This allows the materials to be applied as fiber “sizings” from aqueous solutions or to be incorporated into waterborne resins as thermoplastic toughening agents6- 8. The second section of this chapter discusses efforts toward characterizing the fluorinated materials for their use as interlayer dielectrics or interphase agents for electronic applications. The hypothesis was that low energy, low Tg blocks in block copolymers selectively segregate toward substrate interfaces in multilayer electronic structures, and that this might alleviate interfacial stress caused by mismatches in thermal expansion coefficients between silicon wafers and polyimide films. The approach for studying this phenomenon was to introduce polydimethylsiloxane blocks into a fluorine-containing polyimide 193

O O HN OH CH3 O O HO NH CH3 O O n BPADA/MPDA Poly(amic

:B, (Base)

O O :BH+ HN O - CH3 O O -O NH + CH3 :BH O O BPADA/MPDA Poly(amic acid) n

Heat

O O

CH3 N N O O CH O 3 O BPADA/MPDA Polyimide n

Figure 6-1. Preparation of interphase material 194

backbone. The interfacial properties of the poly(siloxane imide)-substrate systems were investigated.

6.2 INTERPHASES FOR COMPOSITE APPLICATIONS

6.2.1 EXPERIMENTAL

6.2.1.1 Materials

DER 331 Epoxy Resin DER 331 was kindly donated by DOW Chemical Company and used as received. (Mn~380g/mole)

O CH3 OH CH3 O

CH2 O C O CH2 CH CH2 O C O CH2

CH3 CH3 n

4-Aminophenyl Sulfone (DDS) DDS was obtained from Aldrich and used as received. (m.p. 175°C-177°C)

O

H2N S NH2 O 195

Amine Terminated UltemÔ Type Polyimides

The amine terminated UltemÔ type polyimides (Mn» 5000, 10000, and 20000 g/mol) used in this study were synthesized as described in Chapter 4.

O O CH3 O C O H2N N N NH2 CH3 O O n

Amine Terminated UltemÔ Type Poly(amic acid) Salt

The amine terminated UltemÔ type poly(amic acid) salt (Mn» 20, 000 g/mol) used in this study was synthesized as described in Section 4.3.5.

O O CH3 H2N N N NH2 H O O H

CH3 O- O- + :BH+ :BH O O n

UltemÔ Type Poly(dimethylsiloxane imide) Block Copolymers The UltemÔ type poly(dimethylsiloxane imide) block copolymers used in this study (Table 6-1) were synthesized as described in Chapter 3. 196

O O O O CH3 CH3 R R CH2 Si O Si CH2 N N N N 3 n 3 CH3 CH3 O O O O x y

CH3 R = O O

CH3

Table 6-1. BPADA/MPDA poly(siloxane imide) block copolymer compositions used in the composite interphase study

PDMS Molecular Weight % Hard Weight (g/mol) Segment 1000 38 1000 50 1000 80 2000 38 2000 58

6.2.1.2 Preparation of Bilayer Films as Interphase Models

Bilayer films were prepared containing the amine terminated UltemÔ type polyimide and poly(dimethylsiloxane imide) block copolymers as interphase models. The polyimide (or block copolymer) films were solvent cast into a silicone mold from an NMP solution. The solvent was allowed to evaporate with the aid of a heat lamp for 3 days. The films were subsequently heated (120°C) under vacuum for several hours to remove most of the remaining solvent. DDS (2.6251g, 0.0106 mol) was added to a round bottom flask (equipped with a mechanical stirrer) containing DER 331 epoxy resin (8.0351g, 197

0.02114 mol). The mixture was stirred and degassed at 130°C under vacuum to obtain a clear solution. The hot epoxy resin mixture was poured into the mold containing the polyimide (or poly(siloxane imide)) film. The mold was covered, and the samples were cured at 180°C for 1 h, then 220°C for 1 h. Cross-sections of the resulting bilayer films were analyzed using AFM.

6.2.1.3 Preparation of Epoxy - Polyimide Networks

The amine terminated polyimide ( » 5, 10, and 22,000 g/mol) was incorporated into the epoxy resin in systematically varied weight percentages. The amine terminated polyimide modifiers and the crosslinking reagent (meta or para DDS) were reacted stoichiometrically with the epoxy resin to produce epoxy-polyimide networks. A typical procedure is outlined for an epoxy- polyimide network containing 17 weight % of amine terminated polyimide ( » 22,000 g/mol). DER 331 epoxy resin (26.5720 g, 0.0699 mol) and the amine terminated polyimide (5.3144 g, 0.242 mmol) were added to a 100 ml, 2 neck round bottom flask equipped with a mechanical stirrer and a vacuum adapter. The mixture was heated to 200°C and stirred under vacuum until homogeneous, then cooled. A stoichiometric amount of p-DDS (8.618 g, 0.0347 mol) was added to the reaction mixture. This solution was heated to 135°C and stirred under vacuum until homogeneous. The viscous solution was then poured into a silicone mold and placed in a preheated, programmable oven. The samples were heated at 180°C for 1 hour, then 220°C for 1 hour. Analogous procedures were conducted with 9 and 27 weight % amine terminated polyimide ( »22,000 g/mol).

6.2.1.4 Carbon Fiber Sizing For Composite Fabrication

AS4-12K carbon fiber tow (Hercules) was sized from an aqueous 2 weight % amine terminated poly(amic acid) salt solution using a lab scale sizing line.9 198

The sizing conditions were as follows: line speed (1.50 meters/min), tension (3 lbs), and fiber temperature in the drying oven (270°C). The sized carbon fibers were cut and aligned unidirectionally for stitching. The presence of the polyimide on the surface of the fibers was confirmed using X-ray photoelectron spectroscopy.

6.2.1.5 Measurement of Sizing Level and Uniformity

The quantitative level of sizing on the carbon fibers was determined by pyrolsis in a nitrogen atmosphere. The method used was a modified version of test method # SRM 14 specified by SACMA.10 The fiber was dried at 105°C in air for approximately 12 hours prior to the pyrolysis. The fiber was then weighed and placed in a high temperature furnace with a nitrogen purge at a flow rate of approximately 40 cm3/sec. This led to a purge gas residence time of 30 seconds. Typically 15 minutes were required for heating the chamber from room temperature to the pyrolysis temperature of 500°C. Thus, the pyrolysis chamber typically was purged with gas for about 30 residence times prior to reaching the pyrolysis temperature. This ensured that there was no unwanted oxidation of the carbon fibers. Pyrolysis then proceeded for 30 minutes at 500°C. The sizing level was then determined from the weight of the pyrolyzed fiber. The pyrolysis technique was verified with fiber samples containing a known amount of polyimide determined independently by a gravimetric technique. 199

6.2.2 CHARACTERIZATION

6.2.2.1 Transmission Electron Microscopy (TEM) and Atomic Force Microscopy (AFM)

Those samples with lower glass transitions temperatures above 20°C were microtomed at room temperature using a Reichert-Jung Ultracut-E ultramicrotome. The resulting sections were approximately 500-700 Å thick. The sections were collected on 300 mesh copper grids and analyzed using TEM. The samples with glass transition temperatures below room temperature were cryo-microtomed at –120°C using the FC-4D cryo attachment. These sections were collected on 600 mesh copper grids. The TEM was performed using a Philips 420T TEM at 100 kV. A Digital Instruments Dimension 3000 instrument using the Nanoscope IIIa controller was used to obtain AFM images. The images were collected in the Tapping Mode™ using etched silicon tapping tips with spring constants in the 20-100 N/m range. The samples were microtomed smooth before analysis.

6.2.2.2 Fracture Toughness Analysis

Three point bend tests were utilized to characterize the toughness of the polyimide/epoxy networks in terms of the critical-stress intensity factor K1c, according to ASTM Standard D 5045-91. The test specimens had dimensions of thickness B (3.12 mm), width w (6.28 mm), notch depth (2.2 mm), and notch width (0.8 mm) (Figure 6-2). The single-edge notch bending (SENB) method was used. First, a sharp notch was prepared in the specimen by sawing. A natural crack was initiated by inserting a fresh razor blade (cooled by immersion in liquid nitrogen) and tapping. The pre-cracked notched specimen was loaded crack down into a three-point bend fixture and tested using an Instron model 200

4204. The SENB rig has moving rollers to avoid excess plastic indention. The three-point bend fixture was set up so that the line of action of applied load passed midway between the support roll centers within 1% of the distance between these centers. The crosshead speed was 10 mm/min, and the testing was done at room temperature.

W a B = W/2 2.2W 2.2W Figure 6-2. Schematic of three point bend specimen, where W, a, and B are width, crack length, and thickness, respectively.

6.2.2.3 X-Ray Photoelectron Spectroscopy (XPS)

Surface analysis of the sized carbon fibers was obtained on a Perkin Elmer X-ray Photoelectron Spectrometer 5400 Series using a Mg anode operating at 300 watts. Spectra were obtained at a 45º exit angle.

6.2.3 RESULTS AND DISCUSSION

The aim of this research has been to toughen composites for infrastructure by toughening the fiber-matrix interface. Previous work with thermoplastic modifiers has shown that toughness increases significantly with inversion of morphology, the point at which the continuous phase is changed from thermoset to thermoplastic.6-8 The hypothesis is that by placing a small 201 amount of tough ductile material at the interface, toughening of the composite will be achieved due to the inversion of morphology at the interface. Epoxy-Polyimide Networks. In collaboration with C. S. Tyberg, epoxy- polyimide networks were prepared with DER 331 epoxy resin and the reactive amine terminated polyimides at various compositions (Figure 6-3). The epoxy resin was stoichiometrically cured using meta or para-DDS. TEM of the networks containing the lower molecular weight reactive polyimides did not exhibit phase separation, whereas those incorporating the higher molecular weight polyimide (22,000 g/mol) showed microphase separation. Figure 6-4 contains micrographs of epoxy-polyimide networks prepared with 9, 12, and 27 weight % of the » 22,000 g/mole polyimide. Due to higher electron density relative to the epoxy, the polyimide enriched areas appear slightly darker than the epoxy enriched areas. In all three micrographs the polyimide is the included phase and the epoxy is the continuous phase. Fracture toughness measurements of the epoxy-polyimide networks (Table 6-2) were determined using three point bend specimens (cf. Figure 6-2). These values indicated that the networks containing 9 weight % reactive polyimides are only slightly toughened. This may be attributed to the fact that the ductile polyimide is the included phase. These results are consistent with the work reported on polyethersulfone modifiers.16-17 In that system, the authors proposed that at low weight percentages of modifier the mechanism of toughening was ductile fracture of polyethersulfones. By contrast, at higher modifier amounts, after phase inversion, the mechanism of toughening was ductile yielding of the continuous phase. Therefore, it was expected that the polyimide thermoplastic would only cause a large increase in toughness if it were the continuous phase. Due to the high viscosities associated with higher levels of polyimide incorporation, fracture toughness samples with large amounts of polyimide could not be measured by this method. 202

O O C C H2N CH3 N N NH2 O O O O + C C DER 331 O CH3 O n Epoxy Resin Amine Terminated Polyimide R 190 °C

O OH OH O DER 331 DER 331 N R N DER 331 DER 331 O OH OH O

O

H2 N S NH2 125 °C O p-DDS OH OH

OH OH R N N R DER 331 DER 331 OH O OH N S N OH OH O DER 331 DER 331 N N R R OH OH

OH OH EPOXY-POLYIMIDE NETWORK

Figure 6-3. Schematic of epoxy-polyimide network synthesis 203

(a) 9 weight % (b) 17 weight % (c) 27 weight %

Figure 6-4. Transmission electron micrographs of epoxy-polyimide networks with 9, 17, and 27 weight % of the = 22,000 g/mole polyimide, respectively. (The white marker represents 1mm). 204

Bilayer Films as Model Composites. The interphase region was modeled using bilayer films (Figure 6-5). These films were prepared by first solvent casting the polyimide (or poly(siloxane imide) block copolymer) followed by addition of the epoxy resin. After curing the samples, the morphologies of the epoxy-polyimide bilayers were characterized using AFM. Bilayers were prepared with an end-functional (amine terminated) and non-functional (commercially available UltemÔ) polyimide. Figure 6-6 contains AFM cross-sectional images of the bilayers prepared with end-functional and non-functional polyimides. The micrographs clearly show that diffusion occurred in the interphase regions. The included phase in the polyimide layer (left side) is epoxy, and the included phase in the epoxy layer (right side) is polyimide. It appears that the epoxy resin diffused into the polyimide region to a greater degree than the polyimide into the epoxy layer. This was true irrespective of whether the temperature during the first stage of the cure reaction was slightly below or above the Tg of the polyimide (220°C). As discussed previously, this phase inversion was expected to lead to increased toughness at the interface. Although diffusion clearly occurred in the interphase, a clear demarcation remained between the two layers. This suggested that the two materials interdiffused at high temperatures, but phase separated upon cooling. Figure 6-7 shows an AFM cross-sectional image of the entire interphase region spanning over approximately 60 mm. A gradient in composition is clear. 205

Table 6-2. Fracture toughness of epoxy-polyimide networks prepared with DER 331 epoxy resin

Curing Agent Amount Amine K1c (MPa*m0.5) Terminated Polyimide (weight %) p-DDS 0 0.62 +/- 0.02 m-DDS 0 0.58 +/- 0.04 m-DDS 9 0.73 +/- 0.06 m-DDS 9 0.75 +/- 0.08 206

Epoxy Resin (DER 331 + DDS)

Interphase Region Polyimide or Poly(siloxane imide)

Figure 6-5. Epoxy-polyimide (or poly(siloxane imide) block copolymer) bilayer films were investigated as model interphase materials. 207

(a) (b)

Polyimide Epoxy Polyimide Epoxy Polyimi

Figure 6-6. AFM images of (a) non-functional (Ultem™) and (b) end-functional polyimide/epoxy bilayer films showing interdiffusion at the interface. 208

Polyimide Epoxy Side Side

20.0 40.0 60.0 Figure 6-7. AFM image showing entire interphase region of ~60 mm (end- functional polyimide). 209

In contrast, images of the epoxy-poly(siloxane imide) bilayers prepared with copolymers having 50 weight % imide or less (in the copolymers) showed negligible interdiffusion (PDMS »1000 g/mol), although the copolymers appeared miscible with the epoxy resin at the curing temperature. Similar results were obtained for the copolymers prepared with 2,000 g/mol PDMS. It is believed that the hydrophobic PDMS component in the block copolymer film segregates toward the air interface during bilayer preparation, and may therefore prevent diffusion. However, bilayers prepared with poly(siloxane imide) containing an imide continuous phase showed diffusion of the epoxy resin into the poly(siloxane imide) side of the bilayer. Figure 6-8 shows AFM cross- sectional images of epoxy-poly(siloxane imide) bilayers prepared with 30, 50, and 80 weight % imide block. Diffusion between the layers occurred for the sample with 80 weight % imide in the copolymer. Sized Carbon Fibers for Composite Fabrication. Carbon fibers were sized using a lab scale sizing line (Figure 6-9). The carbon fibers proceeded through a series of rollers before reaching the sizing bath containing a 2 weight % aqueous poly(amic acid) salt solution. The poly(amic acid) salt was imidized directly on the fibers in the drying tower of the sizing line. The coated fibers were removed from the drumwinder, aligned unidirectionally, and cut into 6 x 6 inch panels. The sized fibers were analyzed to establish the presence and amount of coating. Surface analysis by XPS confirmed the presence of the polyimide on the fiber surface. The atomic weight percentages on the surface of the sized fibers matched the atomic composition derived from a film of the polyimide sizing material, and also matched the theoretical values based on bulk composition (Table 6-3). SEM photo-micrographs suggest smooth fiber coatings (Figure 6- 10). Burn-off measurements indicated 1.1 weight % sizing was on the fibers. 210

(a) (b) (c)

Epoxy Epoxy Epoxy

PSI PSI PS I

Figure 6-8. AFM images of poly(siloxane imide) (PSI) bilayers prepared with (a) 30 weight % imide, (b) 50 weight % imide, and (c) 80 weight % imide (showing epoxy resin diffusion at the interface). 211

Table 6-3. Surface composition in atomic % of the BPADA/MPDA polyimide sized carbon fibers (determined by XPS at a 45° take-off angle)

Element Theoretical Film Sized Carbon Fiber (atomic %) (atomic %) (%)

C 82.3 81.9 81.8 O 13.3 13.7 13.9 N 4.4 4.6 4.3 212

Process Variables IR Forced Convection Dryer (Fiber temperature is equal • Line Tension 0-30 lbs to imidization temperature) • Fiber Spreading up to 6 inches • Line Speed 0.1-12.0 ft/min • Oven Temperature

Load cell Winder/Drum Winder Sizing Bath (contains aqueous salt solution) Motorized Nip Rollers

Figure 6-9. Depiction of lab scale sizing line 213

Figure 6- 10. SEM micrographs of polyimide sized carbon fibers at different magnifications 214

6.2.4 CONCLUSIONS

Networks prepared with 22,000 g/mol amine terminated BPADA/MPDA polyimide tougheners yielded microphase separated structures. By contrast, lower molecular weight amine terminated BPADA/MPDA polyimides incorporated into these networks did not phase separate upon curing. The phase separated networks with 9 weight % imide as the included phase showed slight increases in fracture toughness. The interphase was modeled using matrix-sizing bilayer films. Interphase regions prepared with epoxy matrices and polyimide sizings contained gradient compositions spanning over approximately 60mm widths. Diffusion at the interface was observed for the epoxy-polyimide bilayer films regardless of the presence of functional endgroups. Interdiffusion was not observed for epoxy- poly(siloxane imide) bilayers where the copolymers had siloxane continuous phases. This is thought to be a result of the insolubility of the siloxane enriched surface in the epoxy resin. By contrast, bilayers prepared with poly(siloxane imide)s where the copolymers had an imide continuous phase showed interdiffusion. It has been shown that significant toughness is achieved at the point at which the continuous phase is changed from thermoses to thermoplastic. The hypothesis is that by placing a small amount of tough ductile material at the interface, composite toughening is achieved due to the morphology inversion at the interface.

6.2.4 RECOMMENDATIONS FOR FUTURE WORK

In view of finding that interdiffusion occurs at the interface for the epoxy- polyimide and poly(siloxane imide) (with an imide continuous phase) bilayers, a study should be undertaken involving the preparation of single fiber composites of these materials. TEM and AFM analysis of these materials would shed light 215 on the actual fiber matrix interphase region. Other important areas of future research include evaluation of the adhesion and composite properties of poly(amic acid) salt (homopolymer and copolymer) sized carbon fibers with epoxy resin. The fracture toughness of epoxy-poly(siloxane imide) networks should also be evaluated.

6.2.5 REFERENCES

1. Sensenich, C. L.; Facinelli, J. V.; Dong, L.; Gardner, S. L.; Davis, R. M. and Riffle, J. S., Polym. Prep. 37(2) (1996): 400. 2. Bowens, A. D.; Sensenich, C. L.; Venkatesan, V.; Robertson, M. A.; McCartney, S. R.; Lesko, J. J. and Riffle, J. S., Polym. Prep. 38(2) (1997): 632. 3. Sensenich, C. L.; Bowens, A. D.; Ji, Q.; McCartney, S. R.; Robertson, M. A. F. and Riffle, J. S. “Waterborne Polyimides and Poly(Imide-Siloxane)s,” Proc. ICCI ’98, Vol 1, pp. 4. 4. Bowens, A. D.; Robertson, M. A.; and Riffle, J. S., Polymer Preprints, 39(1) (1998): 569. 5. Robertson, M. A.; Bump, M. B.; Verghese, S. R. McCartney, Lesko, J. J.; Riffle, J. S.; Kim, I.-C.; and Yoon, T.-H., Journal of Adh. Sci., in press. 6. Chang, J.; Bell, and Shkolnik, S., J. Appl. Polym. Sci., 34 (1987): 2105. 7. Dagli, G. and Sung, N. H.; Proc. ACS, Polym. Materials: Sci. Eng. 56 (1987): 410. 8. Crasto, A. S.; Own, S. H.; and Suramanian, R. V., in Composite Intrfaces, Ishida,H. and Koening, J. L., ed., New York: North-Holland, 1986. 9. Broyles, N. S.; Chan, R.; Davis, R. M.; Lesko, J. J. and Riffle, J. S, Polymer 39(12) (1998): 2607. 10. Test Methods of Suppliers of Advanced Composite Materials Association, SACMA, Test Number SRM 14. 216

6.3 INTERPHASES FOR MICROELECTRONIC APPLICATIONS

6.3.1 EXPERIMENTAL

6.3.1.1 Materials

Fluorinated Poly(dimethylsiloxane imide) Block Copolymers The fluorinated poly(dimethylsiloxane imide) block copolymers used in this study (Table 6-4) were synthesized using 6FDA according to the procedure discussed in CHAPTER 3.

O O O O CH3 CH3 R R CH2 Si O Si CH2 N N N N 3 n 3 CH3 CH3 O O O O x y

CF3 R = C

CF3

Table 6- 4. 6FDA/MPDA block copolymer compositions used in interphase study for electronic applications

PDMS (g/mol) PDMS Weight % 1000 30 2000 10 2000 20 2000 30 5000 10 5000 30 11000 50 217

Fluorinated Polyimide

The fluorinated polyimide ( » 27,000 g/mol) used in this study was synthesized using 6FDA and meta-phenylene diamine according to the procedure outlined in Chapter 4. (Tg » 280°C)

O O CF3 C N N CF3

O O n

Aminopropyl Terminated Polydimethylsiloxane (PDMS) Oligomer

The PDMS oligomer ( » 2000 g/mol) used in this study was synthesized as described in Section 3.2.2. (Tg » -125°C)

CH3 CH3

H2N (CH2)3 Si O Si (CH2)3 NH2 n CH3 CH3

N,N-Dimethylacetamide (DMAc) DMAc (Fisher) was used as received. (b.p. 163-165°C)

O CH3 CH3 C N CH3

Silicon Wafers Polished 2 inch silicon wafers were kindly donated by IBM. The wafer surfaces 218 were cleaned with methylene chloride and hexane and dried at 100°C for 1 h before cooling to room temperature immediately prior to use.

Metal Substrates The tantalum foil (Aldrich, 0.025 mm thick) and titanium (Aldrich, 0.025 mm thick) used in this study were obtained from with a purity of 99.99%. The films were cut into 1.5x1.5 in squares. The surfaces were cleaned with methylene chloride and hexane and dried at 100°C for 1 h before cooling to room temperature immediately prior to use.

6.3.1.2 Block Polymer/Metal Interphase Study

An NMP solution of the block copolymer containing 20 weight % PDMS ( » 2000 g/mol) was prepared with 20 weight % solids. Films of the block copolymer were cast directly onto the tantalum and titanium substrates. The films were dried with the aid of a heat lamp in a covered glass container equipped with a nitrogen sweep. After 24 h, the films were heated to 200°C under vacuum where they were held for 2 h before cooling to room temperature. The dried films were removed from the metal substrates by submerging the samples in liquid nitrogen for 10 seconds and peeling them off with the aid of a razor blade. The films were quickly transferred into the XPS chamber (polymer/metal interface side up) where they were analyzed at 15°, 30°, and 45° take off angles.

6.3.1.3 PDMS/Silicon Wafer Interface Study

The contact angle of PDMS ( » 2000g/mol) on silicon wafers was measured in DMAc using a Dage 550 Series Goniometer. DMAc was added to a small rectangular shaped glass container containing the silicon wafer and a metal support (Figure 6-11). Two types of silicon wafers were investigated: UV 219 treated (to remove organics) and an untreated wafer. Approximately 5-10 mL PDMS droplets were placed on the submerged silicon wafer and the contact angle was measured at room temperature.

q Camera Stage Light Source

Figure 6- 11. Goniometer setup used for polymer/silicon wafer interface study.

6.3.1.4 Residual Stress Analyses

NMP solutions of the fluorinated polyimide and block copolymers containing 5 and 10 weight % PDMS ( » 5000 g/mol) were prepared with 20 weight % solids. The solutions were vacuum filtered. Films of the block copolymers were cast directly onto the titanium substrates. The films were dried with the aid of a heat lamp in a covered glass container equipped with a nitrogen sweep. After 24 h, the film coated metal samples were heated to 300°C under vacuum where they were held for 2 h before cooling to room temperature. The dried film coated metal samples were cut into strips. Kapton tape (0.8 mm in width) was placed on the metal substrate edges before film casting to provide a polymer free area for attachment (Figure 6-12) to the DMA clamps. The residual stress tests were performed on a TA DMA 2980 in the Compression/Penetration Mode using the procedure described by Dillard and Park et al.1 The samples were heated from 25°-350°C at 1°C/min for approximately 6 cycles and the displacements were measured. 220

Polymer Coating

Metal Substrate

Increasing Temperature

Figure 6-12. Single cantilever beam setup for residual stress analyses

The residual stress of the block copolymer films was estimated using the following equation4:

E t 2 æ 1 1 ö s s ç ÷ s f = ç - ÷ (1) 6t f (1- ns ) è Rf R¥ ø

where R is the radius of curvature, E is Young’s modulus, t is thickness, n is Poisson’s ratio, and the subscripts f and s represent the polymer film and substrate respectively. R¥ is the radius of curvature of the substrate without the polymer film. Assuming that the ends of a strip of length l are resting on a flat frictionless surface, R may be determined from the following relationship: 221

l2 + h2 R = (2) 2 cosq 6.3.2 CHARACTERIZATION

6.3.2.1 Refractive Index Analysis

The refractive indices of the 6F/MPDA block copolymer films were determined at room temperature using a Metricon Spectrometer Model 2010.

6.3.2.2 X-Ray Photoelectron Analysis

Surface analyses of the poly(dimethylsiloxane-imide) copolymer interphases and metal substrates were obtained on a Perkin Elmer X-ray Photoelectron Spectrometer PHI 5400 Series using a Mg anode operating at 300 watts and 14 kV. Typical operating pressures were below 5 x 10-7 torr. Spectra were obtained at 15º, 30º, and 45º exit angles. All binding energies were referenced to the carbon C-H photopeak at 285.0 eV. Atomic percent calculations were obtained and used in the estimation of surface compositions as described in Chapter 5.

6.3.3 RESULTS AND DISCUSSION

Refractive Indices of Block Copolymer Films. A low dielectric constant is an important requirement for microelectronic applications. The Iower dielectric constants provided by polymers yield lower signal propagation delays and the possibility of higher wiring densities for interchip connections over the inorganic alternatives. Common flexible chain polyimides typically have dielectric constants of approximately 3.5 at 20°C.2 Incorporation of PDMS and fluorinated moieties 222 into the polyimide backbone offers the possibility of even lower dielectric constants. Table 6-5 shows the refractive index and calculated approximate dielectric constant values of the fluorinated block copolymer films. The refractive index describes a material’s ability to bend light and can be used in the estimation of the dielectric constant according to equation 3.3

2 D = n¥ (3) where D is the dielectric constant and n¥ is the refractive index of a material for light of long wavelength. This equation is valid for nonpolar materials with no permanent dipoles. As is evident from the tabulated values the refractive index decreases with increasing PDMS content and oligomer size within a given series. However, it appears that the amount of PDMS has a greater effect than the oligomer size. The dielectric constant of each of the block copolymers is predicted to be below 3 which is desirable for microelectronic applications. Arnold et al.9 investigated the influence of molecular structure on the dielectric behavior of poly(siloxane imide) segmented copolymers. The results showed that the fluorinated structures yielded dielectric constants <3. A decrease was observed with the addition of just 10 weight % PDMS when compared to the polyimide homopolymer. Block Copolymer /Metal Interphase Study. The hypothesis is that the PDMS component of a poly(siloxane-imide) block copolymer film will alleviate the residual stress in electronic devices through stress relaxation by migrating to the interface.4 Lin measured the interfacial PDMS content of poly(siloxane imide) block copolymer films prepared with BTDA (3,3’,4,4’- benzophenonetetracarboxylic dianhydride) and DDS (3,3’-diaminopiphenyl sulfone) cast on aluminum, zinc, and titanium. The study showed the presence and amount of PDMS at the interface between the coated film and substrate was affected by the substrate used.5 223

Table 6-5. Refractive indices and calculated* dielectric constant for 6FDA/MPDA block copolymers polymer films

PDMS Weight % PDMS n¥ D (g/mol) (experimental) (predicted) 0 0 1.5837 2.5081 1000 30 1.5278 2.3342 2000 10 1.5700 2.4649 2000 20 1.5500 2.4025 2000 30 1.5256 2.3274 5000 10 1.5732 2.4750 5000 30 1.5417 2.3768 11000 50 1.5210 2.3134 * equation (3) in section 6.3.3 224

Silicon wafers used in electronic applications can be coated with any number of layers including, SiN, SiOx, Ta, TaN, Ti, TiN, etc. The outermost surfaces of these inorganic coatings consists of an oxide layer containing hydroxyl groups (Figure 6-13). The acid-base properties of the oxide layers have been found to affect polymeric adhesion as well as influence the interfacial properties.7,8 McCafferty and Wightman reported surface hydroxyl group concentrations of 15, 13, 11, 8, and 6 OH nm-2 for oxide covered aluminum, chromium, titanium, silicon and tantalum films, respectively.7 The same authors also reported isoelectric points of 5.2-5.3, 9.5, and –0.7 for oxide-covered chromium, aluminum, and tantalum films, respectively.8 The metal films used in the polymer/metal interphase study were characterized using XPS. The binding energies of the metal and metal oxides are reported in Table 6-4.10 Figure 6-14 shows the wide scan and photopeak positions for the Ta 4f7/2,Ta 4f5/2 doublet confirming the presence of Ta2O5 on the surface. Likewise, the titanium oxide (TiO2) layer was confirmed with the positions of the Ti 2p3/2, Ti 2p1/2 doublet photopeaks (Figure 6-15). Since the surface free energy of PDMS is very low, the siloxane segments are expected to migrate to the more hydrophobic top (air) surface to form silicon enriched layers.11-13 Figure 6-16 shows the results for the block copolymer containing 20 weight % PDMS. As expected, the polymer/air surface side of the film showed considerable PDMS surface enhancement over the bulk concentration, with ~85 weight % PDMS in the surface region observed at a 15° take-off angle. 225

Organic Carbon Overlayer C,O

Chemisorbed Water H2O

Hydroxylated Region OH, O-2, M+n

O-2, M+n

METAL SUBSTRATE M0

Figure 6-13. A schematic diagram for the model of an oxide-covered metal showing the various layers and their compositions.8 226

Table 6-6. XPS binding energies for the metal and metal oxide layers10

Metal Substrate Binding Energy (eV)

Ta2O5 (Ta4f7/2, 4f5/2) 26.8, 28.6

Ta (Ta4f7/2, 4f5/2) 21.8, 23.6

TiO2 (Ti2p3/2, 2p1/2) 458.9, 464.6

Ti2O3 (Ti2p3/2, 2p1/2) 457.2, 462.9

TiO (Ti2p3/2, 2p1/2) 455.2, 461.0

Ti (Ti2p3/2, 2p1/2) 453.6, 460.0 227

Element Conc.%

4f

-O 1s

-C KLL C 47.78

Ta

-

-O KVV

-C 1s

4p

4p

-O 2s O 39.86

Ta

Ta

- - Ta 12.37 N(E)/E

1000 800 600 400 200 0 Binding Energy, eV Ta+5 N(E)/E Tao

38 35 32 29 26 23 20 Binding Energy, eV

Figure 6-14. XPS wide scan and Ta 4f7/2,Ta 4f5/2 photopeaks for tantalum film 228

KLL

2p3 Element Conc.%

C

- O 1s

LMM Ti

-

Ti

LMM C 48.01

-

2s

- O KVV

Ti

2p1

-

Ti O 41.47

- Ti

- Ti 10.52

- C 1s N(E)/E

3p

Ti

-

- O 2s

1000 800 600 400 200 0 Binding Energy, eV Ti+4

Ti+4 N(E)/E Ti+2

475 470 465 460 455 Binding Energy, eV

Figure 6-15. XPS wide scan and the Ti 2p3/2, Ti 2p1/2 photopeaks for titanium substrate 229

Samples for the polymer/metal interphase study were prepared by solvent casting the block copolymer films onto the titanium and tantalum substrates. Once removed from the substrates, the polymer interfaces were characterized using angle dependent XPS at 15º, 30°, and 45° exit angles (Figure 6-16). The interfacial PDMS composition of the titanium/polymer system was significantly higher than the bulk composition. The PDMS concentration at the tantalum/polymer interface was approximately equal to the bulk concentration. The enriched siloxane presence at the titanium/polymer interface was consistent with a higher concentration of surface hydroxyl groups on the metal substrate. This suggests that the TiOH surface groups hydrogen bond with the siloxane structure (Figure 6-17). Table 6-7 contains the surface atomic concentrations of the metal substrates before coating and after film removal. Although the metal surfaces visually appeared clean after the polymer films were removed, XPS analysis of the metals indicated the presence of residual polymer. Polymer remaining on the metal side could affect the results obtained for the interfacial compositions. However, the amounts obtained followed the trends observed in Figure 6-17 for the titanium and tantalum interfaces.

PDMS/Silicon Wafer Interface. Ideally, for electronic applications interest lies with the polymer/silicon wafer interphase. However, due to the brittle nature of silicon wafers it was impossible to conduct this study on such a system. Thus, the polymer/silicon wafer interphase was investigated using PDMS contact angle analyses. Figure 6-18 contains the XPS wide scan and photopeak for the silicon wafer. The peak position of the Si 2p3/2 doublet confirmed the presence of the

SiO2 layer. The Si photopeak has a binding energy of 99.1 eV and the SiO2 has a binding energy of 103.4 eV.14 230

100 Polymer/Air Surface 90 Polymer/Ti Surface 80 Polymer/Ta Surface

70 S

M 60 D P

%

50 t h g i 40 e W 30

20 BULK

10

0 15deg 30deg 45deg Take-off Angle

Figure 6-16. Interfacial PDMS surface compositions for the polymer/metal interphase study (dashed line denotes bulk PDMS weight percentage). 231

CH3 Si O n CH3 CH 3 Si O n CH3 H H O O

Titanium

Figure 6-17. Schematic showing the interaction of the PDMS component of the block copolymer with the metal substrate suggesting that the TiOH surface groups hydrogen bond with the siloxane structure. 232

Table 6-7. Surface compositions of metal substrates before coating and after film removal in atomic %.

Metal Atomic Concentration (%) Substrates C O Si N F Metal Titanium 1 48.01 41.47 0 0 0 10.52 2 64.30 24.28 4.38 1.99 1.25 3.80 Tantalum 1 47.78 39.86 0 0 0 12.37 2 56.08 22.31 1.82 7.25 5.38 7.18 (1 - before coating; 2 - after film removal) 233

The polymer/silicon wafer interphase study was conducted in an “imide- like” environment. A PDMS droplet was placed on the silicon wafer submerged in DMAc and the contact angle was measured. This was done on an ultraviolet light treated silicon wafer (to remove surface organics) and on an untreated wafer. The PDMS droplet completely wetted the surface of the treated and untreated wafers, thus yielding a 0° contact angle. The results suggest that PDMS should spread at the imide/silicon wafer interface. These findings are in agreement with the argument presented earlier that the substrate oxide layer hydrogen bonds with the siloxane structure. Preliminary Results for Residual Stress Analyses in Fluorinated Block Copolymer Thin Films. The block copolymers were coated onto titanium metal substrates for residual stress analyses and the displacements were measured as the temperature was increased. Investigations by Dillard and Park et al.1 demonstrated that the point at which the coated substrate became straight (indicating no further displacement) corresponded to the stress free temperature, Tsf, and was typically located in the vicinity of the glass transition temperature of the material. An ideal curve is shown in Figure 6-19. The initial heat cycle is due to differences in thermal history whereas the subsequent heat/cooling cycles indicate reproducibility. Dillard and Park et al.1 also reported that the hysteresis loop is associated with the differential thermal lag. Figures 6- 20 to 6-22 show the results for the 6FDA/MDA polyimide control and poly(siloxane imide) block copolymers containing 20 weight % (PDMS =

2000 g/mol) 30 weight % (PDMS = 5000 g/mol) cast on titanium. The curves exhibited some shifting during subsequent heating/cooling cycles. It is not clear if this is a result of substrate yielding or a non-uniform film thickness.

The approximate value of the Tsf occurred slightly above the Tg of the material. (See section 3.4.4.2 for the thermal properties of these materials). 234

Element Conc. % C 10.21 O 46.37 N 0.25

O KVV

Si2p

-

-

Si2s F 0.76

5

-

C1s Si 42.29

F1s

-

Sn3d N1s

-

- -

_ O2s Sn 0.12

-

O1s 1000 800 600 400 200 0 Binding Energy, eV Si

SiO2

N(E)/E

108 106 104 102 100 98 96 Binding Energy, eV

Figure 6-18. XPS wide scan and Si 2p3/2 photopeaks for the silicon wafer

235 Travel distance (h) distance Travel Probe Position

Temperature T sf

Figure 6-19. Ideal deflection curve for a polymer coated strip for several temperature cycles. 236

Tsf

Figure 6-20. Deflection curve for a 6FDA/MPDA polyimide cast on titanium. 237

Tsf

Figure 6-21. Deflection curve for a 6FDA/MPDA poly(siloxane imide) block copolymer containing 30 weight % PDMS ( = 5000 g/mol) cast on titanium. 238

Tsf

Figure 6-22. Deflection curve for a 6FDA/MPDA poly(siloxane imide) block copolymer containing 20 weight % PDMS ( = 2000 g/mol) cast on titanium. 239

The sample with the less travel distance (Figure 6-19) or the shallower curve is considered to be “less stress generating.” Due to differences in the film thickness this observation could not be adequately carried out (Table 6-8). The radius of curvature (Figure 2-23) was estimated from the data according to equation (2) presented in section 6.3.1.4. This value was then used to calculate the residual stress according to equation (1) in section 6.3.1.4. The preliminary test results indicated that the polyimide film has a lower residual stress than the block copolymer films (Table 6-9). At this point it is not clear if the higher stress observed for the block copolymers are a result of differences in the CTE between the block copolymer and the metal substrate or a result of film thickness. 240

Table 6-8. Probe travel distance of cast block copolymer films

Substrate/Polymer Thin Film Thickness Travel Distance (mm) (mm) Ti / Polyimide Control 20 34

Ti / F-PSI30(5K) 40 150 (30 weight % PDMS, Mn 5000 g/mol)

Ti / F-PSI20(2K) 30 300 (20 weight % PDMS, Mn 2000 g/mol) 241

l

q h

R

Figure 6-23. Illustration of the radius of curvature, R, determined using the single cantilever beam setup. 242

Table 6-9. Preliminary results for curvature measurements of thin films

SAMPLE/ PDMS Weight Radius of Residual SUSTRATE % (Oligomer Curvature Stress , g/mol) (mm) (MPa) F-PI / Ti 0 6174 1.963 F-PSI70(5k) / Ti 30(5K) 1438 6.637 F-PSI80(2k) / Ti 20(2K) 714 17.823

Parameters Used:

Ti Modulus: 105 Gpa Ti CTE: 0.34 ppm 243

6.3.4 CONCLUSIONS

Fluorinated poly(siloxane imide) block copolymers have been evaluated for their use as interphase materials in electronic applications. The refractive indices of the fluorinated block copolymer films suggest these materials will have low dielectric constants, < 3.0. In combination with their known high glass transition temperatures and superior thermal and mechanical properties, these materials are excellent candidates as interlayer dielectrics. The hydroxyl groups of the metal oxide layer promote hydrogen bonding between the substrate and siloxane segments. Thus, the presence of PDMS at the interface increases with an increase in surface hydroxyl concentration. Contact angle studies carried out in DMAc suggested that PDMS will spread at the polymer/silicon wafer interface, which may aid in the reduction of residual stress. Preliminary results from the residual stress analysis of thin films showed that the method employed was promising. However, additional work is needed to investigate substrates and film thickness to better understand the residual stress analysis.

6.3.5 RECOMMENDATIONS FOR FUTURE WORK

In future studies it would be interesting to investigate the PDMS/silicon wafer interface using solvents having structures that might better mimic the polyimide structure. The dielectric properties and water absorption of the block copolymer films should be determined. A research effort should be undertaken involving the preparation of more rigid poly(siloxane imide) block copolymers or polyimide-siloxane networks for electronic applications. These materials would have better chemical resistance, thermal stability, and higher glass transition temperatures than the block copolymers used in this study. Other areas of potential research include the fabrication of a multilayer structure containing these materials as interlayer dielectrics. 244

6.3.6 ACKNOWLEDGEMENTS

Special thanks are extended to IBM for supplying the silicon wafers.

6.3.7 REFERENCES

1. Dillard, D. A.; Park, T. G.; Zhaung, H.; Chen, B., Adh Society, Proceedings of the 22nd Annual Meeting, (1999): 336. 2. Elsner, G., J. Appl. Polymer Sci. 34 (1987): 815. 3. Glorstone, S., Textbook of Physical Chemistry, 2nd edition, 1948, p 536. 4. Hedrick, J. L.; Brown, H. R.; Volksen, W.; Sanchez, M; Plummer, C. J. G.; and Hilborn, J. G., Polymer 38(3) (1997): 605. 5. Lin, T., Surface and Interphase Studies of the Adhesion of a Siloxane- Modified-Polyimide Coating on Metals, M.S. Thesis, VPI&SI, 1990. 6. Metson, J. B.; Hyland, M. M.; Gillespie, A.; and Hemmingesen-Jensen, M., Colloid Surf. A 93 (1994): 173. 7. Wightman, J. P.; Lin. T.D.; and Webster, H. F., Intl. J. Adhes. 12 (1992): 133. 8. McCafferty, E. and Wightman, J. P., Surf. Interface Anal. 26 (1998): 549. 9. Arnold, C. A.; Summers, J. D.; Chen, Y. P.; Bott, R. H.; Chen, D. and McGrath, J. E., Polymer 30 (1989): 986. 10. Hofman, S. and Sanz, J. M., J. Trace Micro. Tech. 1 (1982/83): 213. 11. Chen, X. and Gardella Jr., J. A., Macromolecules 27 (1994): 12. 12. Zhuang, H.; Gardella, Jr.; Incavo, J. A.; Rojstacze, R. S.; Rosenfeld, J. C., J. Adh., 63(1) (1997): 199. 13. Dwight, D., McGrath, J. E.; Lawson, G.; Patel, N.; York, G., “Surface and bulk microphase separation in siloxane-containing block copolymers and their blends: The roles of composition and kinetics from multiphase macromolecular systems,” Contemp. Topics in Polym, Sci. 6 (1989): 265. 245

14. Wagner, C. D., Riggs, W. M.; Davis, L. E.; Moulder, J. F.; Muilengerg, G. E. (eds), Handbook of X-ray Photoelectron Scpectroscopy, , Eden Prairie, MN: Perkin-Elmer Corporation Physical Electronics Division (1979). 246

SUMMARY/CONCLUSIONS

Synthetic aspects, thermal, morphological, surface, and interfacial properties were explored in efforts to obtain a fundamental understanding for the application of polyimide and poly(siloxane imide) block copolymers as interphases in composite and microelectronic components. The two-step poly(amic acid) synthesis in THF proved to be an efficient method of preparation for the systematically varied poly(siloxane imide) block copolymers containing BPADA/MPDA and 6FDA/MDA structures. The BPADA/MPDA poly(siloxane imide) block copolymers were prepared from water soluble precursors. The poly(siloxane amic acid) salts can be imidized quickly (<10 min. at 260°C), thus suggesting that the potential for online processing in composite manufacturing is promising. Utilizing the same methodology hydroxy and amino-functionalized BPADA/MPDA poly(amic acid) salts were prepared. The salts of these materials are water soluble and their aqueous solutions are stable in air. Block copolymers containing 6FDA/MPDA structures were prepared via melt imidization of their amic acid precursors. Both block copolymer series produced microphase separated materials with thermal stabilities well above 400°C. It was demonstrated that the thermal stability and char formation of the block copolymers was influenced by the amount of PDMS incorporated. The reduction in the upper Tg of the copolymers is thought to be due to the lowering of the imide phase molecular weight. This is not completely ignoring the possibility of phase mixing to some extent, especially with those copolymers prepared with 1000g/mol PDMS oligomers. Both XPS and water contact angle analyses demonstrated the surface enrichment of the copolymers with the low energy PDMS component. The results showed that at bulk PDMS concentrations as low as 5 weight % there was a 25% increase in the water contact angle on the block copolymer film, thus indicating a more hydrophobic, low energy surface. The surface topography was influenced by the degree of phase separation and was characterized using AFM. This was quantitatively represented with the roughness factor, Ra. It was found 247 that the surface roughness increased with increasing PDMS oligomer content and molecular weight. Fundamental aspects were evaluated for composite and microelectronic applications. Networks prepared with 22,000 g/mol amine terminated BPADA/MPDA polyimide tougheners yielded microphase separated structures. By contrast, lower molecular weight amine terminated BPADA/MPDA polyimides incorporated into these networks did not phase separate upon curing. The phase separated networks with 9 weight % imide as the included phase showed slight increases in fracture toughness. The interphase was modeled using matrix-sizing bilayer films. Interphase regions prepared with epoxy matrices and polyimide sizings contained gradient compositions spanning over approximately 60mm widths. Diffusion at the interface was observed for the epoxy-polyimide bilayer films regardless of the presence of functional endgroups. Interdiffusion was not observed for epoxy- poly(siloxane imide) bilayers where the copolymers had siloxane continuous phases. This is thought to be a result of the insolubility of the siloxane enriched surface in the epoxy resin. By contrast, bilayers prepared with poly(siloxane imide)s where the copolymers had an imide continuous phase showed interdiffusion. Fluorinated poly(siloxane imide) block copolymers were evaluated for their use as interphase materials in electronic applications. The refractive indices of the fluorinated block copolymer films suggest these materials will have low dielectric constants, <3.0. In combination with their known high glass transition temperatures and superior thermal and mechanical properties, these materials are excellent candidates as interlayer dielectrics. The hydroxyl groups of the metal oxide layer promote hydrogen bonding between the substrate and siloxane segments. Thus, the presence of PDMS at the interface increases with an increase in surface hydroxyl concentration. Contact angle studies carried out in DMAc suggested that PDMS will spread at the polymer/silicon wafer interface, which may aid in the reduction of residual stress. 248

Preliminary results from the residual stress analysis of thin films showed that the method employed was promising. However, additional work is needed to investigate substrate and film thicknesses to better understand the residual stress analysis. 249

Vita

The Andrea Demetrius Bowens was born in Florence, Alabama on September 23, 1972 the older of two children to Lillian Francine Poole and John Luster Bowens. After graduating from Bradshaw High School in 1990 she accepted a 5 year B.S. / M. S. Office of Naval Research Scholarship to attend Clark Atlanta University in Atlanta, Georgia. While at Clark Atlanta University she studied organic synthesis and spent two summers interning at 3M Company in St. Paul, MN. Her master’s thesis entitled Benzofulvenes for the Incorporation into Polymers with Potential Nonlinear Optical Behavior was completed under the advisement of Dr. Eric Mintz in 1995. Following graduation in 1995, she entered the Ph.D. program at Virginia Polytechnic Institute and State University under the advisement of Dr. Judy S. Riffle. During her tenure at Virginia Tech she spent one summer working at Colgate-Palmolive Company, presented papers at several national meetings, and was an active member of the Division of Polymer Chemistry (ACS) Executive Board as a student representative. Upon completion of her Ph.D. requirements in December 1999, the author will begin work at Procter & Gamble Company in Cincinnati, OH as a scientist in Beauty Care Technology.