This is to certify that the dissertation entitled SURFACE INITIATED ATRP OF SUBSTITUTED STYRENES AND FUNCTIONAL MONOMERS ON FLAT SURFACES presented by Sampa Saha has been accepted towards fulfillment of the requirements for the PhD. degree in Chemistry Wavy Professor’s Signature (2 my 2.0/0 Date MSU is an Affirmative Action/Equal Opportunity Employer {\UB\ . may M'Ch‘ga" State l L University PLACE IN RETURN BOX to remove this Checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5108 KlProj/AcoSPres/ClRCIDateDue.indd SURFACE INITIATED ATRP OF SUBSTITUTED STYRENES AND FUNCTIONAL MONOMERS ON FLAT SURFACES By Sampa Saha A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Chemistry 2010 ABSTRACT SURFACE INITIATED ATRP OF SUBSTITUTED STYRENES AND FUNCTIONAL MONOMERS ON FLAT SURFACES Surface initiated atom transfer radical polymerization (ATRP) of substituted styrenes leads to rapid synthesis of uniform and thick substituted polystyrene brushes (>100 nm in 1 hour) from gold and silicon surfaces. High growth rates were observed for styrenes substituted with electron withdrawing groups in meta/para positions. The effects seen in surface and solution polymerizations are similar for styrenes with electron withdrawing groups, and for electron donors in ortho and para positions. However, electron donors at meta sites have surprisingly fast growth rates, which may be due to steric inhibition of termination. The overall surface polymerization rates for substituted styrenes was analyzed and found to follow the Hammett relation with p = 0.51. The ratio of kp to kt, is as an indicator of the likelihood that a reaction will reach high degrees of polymerization before termination. During surface initiated polymerization, thiols desorb from gold surfaces at low temperatures (< 60 C’C) and terminate growing polymers during surface initiated ATRP. Thiol desorption was prevented by forming a cross-linked poly(siloxane) primer layer on the gold surfaces prior to attaching the initiator layer. These modified surfaces provide polymer film thicknesses comparable to films grown from silicon surfaces. This strategy above 100 OC since the film delaminates from the substrate. Usually, difficult monomers, such as 2-vinyl pyridine, a polyelectrolyte precursor, cannot be grown from thio-initiators anchored on gold, but the cross-linked initiator enabled growth of thick polymer films. Polyacrylate brushes with pendent terthiophenes (PTTMM) were successfully grown from ITO and gold using surface initiated ATRP. Using cyclic voltammetry, the PTTMM brush was electrochemically cross-linked to form a conjugated polymer network. The conjugation lengths in the film were short, but were increased via heterocoupling. These uniformly grafted conducting polymer brushes may find use in photovoltaic devices. Click chemistry was used for the post—functionalization of hydrophilic polymer brushes. The polymer brushes were random copolymers of AZPMA (azidopropyl methacrylate), a functional monomer with a pendent azo group, and oli go ethylene glycol methyl ether methacrylate (PEGMA) with varying ethylene oxide chain lengths, which enabled control over the hydrophilicity and functional group density in the copolymer. Subsequent post-functionalization of homo and copolymer brushes was demonstrated by appending an alkyne—modified dye to AZPMA via click reactions. Kinetic studies showed that the modification of surface-grafted homo/copolymer brushes was fast (> 60% conversion in 1 min) irrespective of the copolymer composition. In the case of water soluble high molecular weight alkynes, surface-grafted copolymers with highest amount of hydrophilic monomer (PEGMA) gave the highest degree of immobilization, which indicates its potential application in bioconjugation. To My Beloved Family iv ACKNOWLEDGMENTS I would like to thank my advisor, Professor Gregory L. Baker for being great mentors during my graduate study at Michigan State University. He has been an excellent research advisor who really taught me how to think and come up with own ideas. The Openness to ask questions, propose solutions and the emphasis for critical thinking were important attributes that made the work environment both fun and challenging. I express my deep gratitude to all my committee members, Dr. Merlin Bruening, Dr. GK. Chang and Dr. William Wulff. Professor Merlin Bruening has a great science attitude and critical thoughts. He is exceptionally good at math. I found plenty of helpful suggestions from him to achieve my research goal. As my second reader, I want to thank Professor William Wulff for his great suggestions especially toward my seminar and a great party with lot of delicious food at his home. I would like to thank all past and present Baker group members: Bao, Ying, DJ, Erin, Qin, Gina, Hui, Heyi, Yiding, Wen and Quanxuan. I also enjoyed working with Bruening group members especially Fei, Weihan, David and Parul. I really enjoyed my life here with my husband and all of my friends. Special thanks to Debbie Roper for her kindness to me during my stay in Chemistry Department, especially in my last semester. She has been always a big help for me. Though last year I lost my father, I believe his blessings are always with me. Whatever I am today, it’s all because of my parent’s encouragement. Finally, I would like to thank my husband who was my second mentor during my graduate study. Last but not least I would like to thank my sister Dona and brother-in law Apurba for constant positive support. My beloved family deserve the greatest thank and love. vi TABLE OF CONTENTS Page List of Tables ................................................................................................................... x List of Figures .................................................................................................................. xi List of Schemes ............................................................................................................. xviii List of Abbreviations ..................................................................................................... xx Chapter 1. Introduction ............................................................................................... 1 1.1. Polymer brushes —- a brief definition .................................................................. 1 1.2. Structure of polymer brushes ............................................................................. 2 1.3. Synthesis of polymer brushes ~— recent advances ............................................... 5 1.3.1. Polymer Brushes by free radical polymerizations ..................................... 5 1.3.2. Surface-initiated controlled radical Polymerizations ................................. 12 1.3.2.1 Polymer brushes by Atom Transfer Radical Polymerization ............ 12 1.3.2.2 Polymer brushes by Nitroxide Mediated Polymerization .................. 18 1.3.2.3 Polymer brushes by Reversible Radical Addition-Fragmentation Chain Transfer (RAFT) .................................................................................. 20 1.3.3. Polymer brushes by Cationic Polymerizations .......................................... 22 1.3.4. Polymer brushes by Anionic Polymerizations ........................................... 23 1.3.5. Polymer brushes by Ring Opening Polymerizations ................................. 27 1.3.6. Polymer brushes by Ring Opening Metathesis Polymerizations ............... 29 1.4. Application of Surface Initiated Polymerizations ............................................... 31 1 .5. References ........................................................................................................... 40 Chapter 2. Substituent effect in surface AT RP of polystyrene brushes ................... 46 2.1 . Introduction ............................................................................................................... 46 2.2. Experimental Section ................................................................................................ 51 2.2.1. Material ............................................................................................................ 51 2.2.2. Characterization Methods ................................................................................ 51 2.2.3. Preparation of initiator immobilized Au Substrates ......................................... 52 2.2.4. Preparation of initiator immobilized Si Substrates .......................................... 52 2.2.5. Polymerization of substituted styrene from initiator immobilized Au surfaces for kinetic Study ...................................................................................... . ......... 53 2.3. Results and Discussions ............................................................................................ 54 2.3.1. Synthesis of substituted styrene monomers ..................................................... 54 2.3.2. Synthesis of substituted styrene brushes from Au surfaces ............................. 56 2.3.3. Kinetic study of polymerization of substituted styrene brushes ...................... 57 2.3.4. Substituent effects on polymerization rate ....................................................... 63 2.3.5. Synthesis of substituted styrene brushes from Si02 surfaces .......................... 71 vii 2.4. Conclusions ............................................................................................................... 73 2.5. References ................................................................................................................. 74 Chapter 3. Surface effects in surface ATRP ................................................................ 77 3.1. Introduction ............................................................................................................... 77 3.2. Experimental Section ................................................................................................ 78 3.2.1. Materials .......................................................................................................... 78 3.2.2. Characterization Methods ................................................................................ 79 3.2.3. Preparation of Initiator-Immobilized Au and SiOg substrates ......................... 80 3.2.4. Surface—Initiated Polymerizations .................................................................... 81 3.3. Results and Discussions ............................................................................................ 81 3.3.1. Surface-Initiated Polymerizations from Au and Si02 substrates .................... 81 3.3.2. Block 00polymer formation on Au and SiOz substrates ................................. 84 3.3.3. Formation of cross-linked initiator and grow polymer there from .................. 90 3.4. Conclusions ............................................................................................................... 103 3.5. References ................................................................................................................. 104 Chapter 4. Conducting Polymer tethered on flat surface via “grafting from” approach .................................................................................................... 107 4.1. Introduction ............................................................................................................... 107 4.2. Experimental Section ................................................................................................ 108 4.2.1. Materials .......................................................................................................... 108 4.2.2. Characterization Methods ................................................................................ 108 4.2.3. Synthesis of 3-methyl thienyl methacrylate ..................................................... 109 4.2.4. Synthesis of [2, 2’:5’, 2”-terthiophen]~3’-yl methyl methacrylate ................. 110 4.2.5. Synthesis of silane initiator .............................................................................. 112 4.2.6. Preparation of initiator immobilized flat substrates ......................................... 113 4.2.7. Surface initiated polymerization on gold and ITO substrates .......................... 113 4.2.8. Electrochemistry .............................................................................................. 114 4.3. Results and Discussions ............................................................................................ 115 4.3.1. Monomer synthesis .......................................................................................... 115 4.3.2. Silane Initiator Synthesis for ITO substrate ..................................................... 116 4.3.3. Deposition of SAM on ITO surface ................................................................. 117 4.3.4. Synthesis of polymer brush on gold and ITO surfaces .................................... 118 4.3.5. Electrochemical crosslinking ........................................................................... 122 4.3.6. Anodic copolymerization of the grafted polythiophene units and additional thiophene derivative in solution ~ Heterocoupling .......................................... 131 4.4. Conclusions ............................................................................................................... 135 4.5. References ................................................................................................................. l 36 viii Chapter 5. Polymerization of azidomethacrylates on gold surface and its elaboration via click chemistry ..................................................................................... 140 5. l . Introduction ............................................................................................................... 140 5.2. Experimental Section ................................................................................................ 141 5.2.1. Materials .......................................................................................................... 141 5.2.2. Home and Co-polymerization of Azpma, Egma and Pegma from initiators immobilized on Au and ITO substrates ........................................................................... 142 5.2.3. Click functionalization of the homo and co-polymer brushes ......................... 143 5.2.4. Characterization Methods ................................................................................ 144 5.3. Results and Discussions ............................................................................................ 145 5.3.1. Synthesis of uniform PAzpma brushes from gold surface ............................... 145 5.3.2. Copolymerization of Azpma with Egma and Pegma on gold surface ............. 152 5.3.3. Derivatization of the copolymer brush via click chemistry with a dye ........... 162 5.3.4. Derivatization of the copolymer brush via click chemistry with water soluble polymer ...................................................................................................................... 173 5.4. Conclusions ............................................................................................................... 179 5.5. References ................................................................................................................. 180 ix LIST OF TABLES Table Page Table 2.1. Propagation and termination rate constants for representative methacrylates ........................................................................................... 49 Table 2.2. Apparent Rate Coefficients in surface ATRP obtained from model and Absolute Propagation Rate Constants of Substituted Styrene .......... 66 Figure Figure 1.1. Figure 1.2. Figure 1.3. Figure 1.4. Figure 1.5. Figure 1.6. Figure 1.7. LIST OF FIGURES Graphical representation of a) the ‘grafting to’ technique, and b) the ‘ grafting from’ technique .............................................................. a) Schematic showing a thin film comprised of polymer chains in a random conformation. b) Schematic showing a thin film comprised of polymer chains in a brush conformation Examples of polymer systems comprising polymer brushes ................................................... Time-dependent properties of polymer chains grown by surface- initiated free radical polymerization of styrene: (a) molecular weight Mn, (b) grafting densities 5(PS), and (c) polydispersity of the covalently attached polymers. ........................................................... Examples of monomers that have been used to synthesize polymer Page ..... 2 ...2 .9 brushes via ATRP .................................................................................... 14 Optical micrographs of patterned surfaces: (left image) 10-um features in a continuous polymer brush showing regions of poly(tert—butyl acrylate) (dark) and poly(acrylic acid) (light) and (right image) interaction of a water droplet with ZOO-pm features showing an unusual wetting profile and preferential interaction with poly(acrylic acid) brush domains ..................................................... 33 Top image: PTPAA brushes and bottom image: Cartoon of inferred structure for CdSe nanocrystal infiltrated polymer brush photovoltaic device (From bottom to top: ITO-coated glass slide modified by surface attachment of a bromine end-caped trichlorosilane self-assembled-monolayer (SAM) (blue squares), polymer brushes grown from the SAM (red lines), CdSe nanocrystals infiltrated into the brush network exhibiting some degree of phase separation in the plane of the film (small black circles), and caped with an aluminum cathode. ...................................... Top image: PMETAC brushes and bottom image: Change in the wetting characteristics of PMETAC brushes (height, h ~20 nm) ' after exchanging the two contrasting counterions: TFSI (a) and polyphosphate (PP) (b). c) Representation of 8 AW as a function of counter ion (PP and TFSI). The plot depicts the reversible behavior of PMETAC brushes over repeated cycles of TFSI and PP counter xi .35 Figure 1.8. Figure 2.1. Figure 2.2. Figure 2.3. Figure 2.4. Figure 2.5. Figure 2.6. Figure 2.7. Figure 3.1a. ion exchange. On the right the chemical structures of both counter ions are represented .................................................................................. 37 Schematic description of the procedure. Five acetylene group- containing compounds: 1-hexyne, 5-hexyn-1—ol, 4—pentanoic acid, propargyl benzoate, and biotin-PEO-LC-N-pentynoate. ......................... 39 Evolution of the ellipsometric brush thickness with time for the polymerization of substituted styrenes from gold surface. ...................... 58 Representative examples of reflectance F TIR spectroscopy of gold surface coated with (a) an immobilized initiator (b) 70 nm polystyrene brush (0) 55 nm poly (3,5 di-tert butyl styrene) brush ((1) 5 nm poly (2,6 di-methoxy styrene) brush (e) 27 nm poly(a- methoxy styrene) brush (f) 130 nm poly(3,5 di trifluoromethyl styrene) brush ........................................................................................... 59 Evolution of the ellipsometric brush thickness with time for the polymerization of para substituted styrene from gold surface ................. 61 Representative examples of reflectance F TIR spectroscopy of para substituted styrene grown from gold surface (a) 70 nm polystyrene brush (b) 10 nm poly(4-methoxy styrene) brush (c) 200 nm poly (4-trif1uoromethy1 styrene) brush (d) 55 nm poly (4-methyl styrene) brush (e) 100 nm poly (4-bromo styrene) brush (f) 150nm poly (4- tert-butyl styrene) brush ........................................................................... 62 A representative example of the best fit curve to Eq 6. Thickness of poly (4-bromo styrene) film vs time for ATRP on gold wafers. ......... 67 Hammett plots for (kp/kt) in surface ATRP of substituted styrenes and for absolute kp in conventional polymerization of substituted styrenes .................................................................................................... 70 Evolution of the ellipsometric brush thickness with time for the polymerization of 3, 5 ditrifluoromethyl styrene, 4-tert-buty1 styrene, styrene, 4-methy1 styrene and 2-methoxy styrene grown from SiOz surface .................................................................................... 72 Evolution of the ellipsometric brush thickness with time for the polymerization of MMA (methyl methacrylate) from gold and silicon surfaces at 50 OC .......................................................................... 83 xii Figure 3.1b. Evolution of the ellipsometric brush thickness with time for the polymerization of styrene firom gold and silicon surfaces at 50 0C ........ 84 Figure 3.2. Reflectance FTIR spectra of gold substrates coated with (a) 97 nm PtBA brushes grown from the initiator layer; (b) PtBA (97 mm)- block-PMMA (210 um) copolymer brushes grown from the PtBA brush surface ............................................................................................ 86 Figure 3.3a. Surface polymerization study of the formation of PtBA-b-PMMA from gold surface ..................................................................................... 88 Figure 3.3b. Surface polymerization study of the formation of PtBA-b-PMMA from SiOz surface ................................................................... 89 Figure 3.4a. Surface polymerization study of the formation of PMMA brush from gold surface immobilized with crosslinked and non- crosslinked initiator and SiOz surface ..................................................... 93 Figure 3.4b. Surface polymerization study of the formation of PS brush from gold surface immobilized with crosslinked and non-crosslinked initiator and SiOz surface.................................. 94 Figure 3.5a. Reflectance FTIR spectra of gold substrates coated with (a) PMMA brushes grown from a non-crosslinked thio initiator in 1 hour. ......................................................................................................... 95 Figure 3.5b. Reflectance FTIR spectra of gold substrates coated with (0) PS brushes grown from a crosslinked thio initiator in 8 hours (<1) PS brushes grown from a crosslinked thio initiator in 8 hours ..................... 96 Figure 3.6a. Temperature dependence study of PMMA brush grown for 1 hour from various initiators. ............................................................................. 97 Figure 3.6b. Temperature dependence study of PS brush grown for 1 hour from various initiators ....................................................................................... 98 Figure 3.7. Optical micro graphs of PS brushes grown from gold surfaces immobilized with crosslinked initiator and viewed through crossed polarizers: (a) and (b) show a 70 nm thick film grown for 1 hour at 115°C ...................................................................................................... 99 xiii Figure 3.8. Surface polymerization study of the formation of PVP brush from various initiators immobilized on flat surface ......................................... 100 Figure 3.9. Reflectance FTIR (right) of PVP brush grown for 8 hours from gold surface immobilized with (a) crosslinked initiator (b) non- crosslinked initiator .................................................................................. 101 Figure 3.10. Possible reason of termination of surface-bound radicals on Au ............ 102 Figure 4.1. Evolution of the ellipsometric brush thickness with time for the polymerization of PMTM and PTTMM from gold and ITO surfaces .................................................................................................... 121 Figure 4.2. Representative examples of reflectance FTIR spectroscopy of gold surface grafted with (a) 50 nm PTTMM brush (b) 100 nm PMTM brush ......................................................................................................... 122 Figure 4.3. Cyclic voltammetry (20 cycles) of ~30 nm PTTMM brush coated on ITO surface from -25 to 1500 mV at 100 mV/s .................................. 124 Figure 4.4. Scan rate dependency study of PTTMM brushes coated on ITO surface at scan rates of 20-80 mV/s in a 0.1 M TEAP/CH3CN electrolyte solution ................................................................................... 125 Figure 4.5. Plot of peak current intensity versus scan rate at the maximum of the oxidation wave of PTTMM brush coated on ITO surface ................. 126 Figure 4.6. Representative examples of reflectance FTIR spectroscopy of gold surface coated with (a) 50 nm PTTMM brush (b) electrochemically crosslinked PTTMM brush ...................................................................... 128 Figure 4.7. UV-vis absorption spectra of PTTMM brush grown on ITO surface (blue line) and electrochemically cross linked PTTMM (EPTTMM) brush on ITO surface ............................................................................... 129 Figure 4.8. Topographical AFM images of ITO surface coated with PTTMM brush and electro polymerized PTTMM brush taken with tapping mode imaging ................................................................................. i .......... 130 Figure 4.9. Cyclic voltammetry (20 cycles) of ~30 nm PTTMM brush grafted on ITO surface in CH3CN added with TEAP (0.05 M) in the presence of EDOT in solution (first and 20th scan) at the scan rate of 20 mV/s ................................................................................................ 133 xiv Figure 4.10 UV-vis absorption spectra of PTTMM brush grown on ITO surface and electrochemically heterocoupled PTTMM brush with EDOT on ITO surface ......................................................................................... 134 Figure 5.1. Evolution of the ellipsometric brush thickness with time for the polymerization of Azpma from initiator monolayers on Au substrates using different catalyst systems .............................................. 148 Figure 5.2. Reflectance FTIR spectra of gold substrates coated with PAzpma brushes grown for (a) 0.5 h (b) 4 h (c) 6 h and (d) 8 h from the initiator layer ............................................................................................ 150 Figure 5.3. Topographical AFM image of gold surface coated with 250 nm PAzpma brush taken with tapping mode imaging ................................... 151 Figure 5.4. Comparison of the ellipsometric brush thickness with time for the homopolymerization of Azpma and Egma from the initiator monolayers on Au substrates ................................................................... 153 Figure 5.5. Evolution of the ellipsometric brush thickness with time for the co- polymerization of Azpma and Egma with various initial feed ratios from initiator immobilized gold substrates using CuCl / PMDETA / DMF system at 50 OC ............................................................................ 154 Figure 5.6. Reflectance FTIR spectra of gold substrates coated with copolymer brush of Azpma and Egma at different feed ratios .................................. 155 Figure 5.7. Results from FTIR analysis of the copolymer brushes mentoned in figure 5.6 .................................................................................................. 156 Figure 5.8. Evolution of the ellipsometric brush thickness with time for the co- polymerization of Azpma and Pegma with various initial feed ratios from initiator immobilized gold substrates using CuCl / PMDETA / DMF system at 50 °c .......................................................... 158 Figure 5.9. Reflectance FTIR spectra of gold substrates coated with copolymer brush of Azpma and Pegma at equimolar (50/50) feed ratio in different set of polymerization time ......................................................... 159 Figure 5.10. Reflectance FTIR spectra of gold substrates coated with copolymer brush of Azpma and Pegma at different feed ratios ................................. 160 XV Figure 5.11. Figure 5.12. Figure 5.13. Figure 5.14. Figure 5.15. Figure 5.16. Figure 5.17. Figure 5.18. Figure 5.19. Figure 5.20. Figure 5.21. Results from FTIR analysis of the copolymer brushes mentioned in Figure 5.10 ............................................................................................... 161 A representative example of kinetics of click reactions with a fluorescence dye monitored by Uv-Vis spectra of ITO substrates coated with ~ 180 nm copolymer brushes of Azpma and Pegma at equimolar (50/50) feed ratios ................................................................... 164 A representative example of kinetics of click reactions with a fluorescence dye monitored by reflectance FTIR spectra of gold substrates coated with ~ 200 nm copolymer brushes of Azpma and Egma at equimolar (50/50) feed ratios ..................................................... 165 Kinetic study of click reactions with a fluorescence dye monitored by reflectance FT IR spectra of gold substrates coated with ~200 i 30 nm copolymer brushes of Azpma and Egma as well as Azpma and Pegma at different feed ratios ........................................................... 166 Measurement of molar extinction coefficient of the alkyne modified fluorescein dye .......................................................................... 169 Click reactions for 5 min with a fluorescence dye monitored by Uv-Vis spectra of ITO substrates coated with ~ 200 :1: 30 nm copolymer brushes of Azpma and Pegma at different initial feed ratios ......................................................................................................... 170 Results obtained from Uv—Vis spectra of the copolymer brushes mentioned in Figure 5.16. ........................................................................ 171 The dye molecule binding capacity as a function of azide quantity in the 200 i 30 nm copolymer brush on ITO surface with different feed ratios. ................................................................................................ 172 A representative fluorescence microscopy image of the clicked surface and the reference prepared without copper catalyst under otherwise equivalent conditions ............................................................... 173 Click reactions for 12 hours with a water soluble polymer (alkynyl mPEG 5000) monitored by reflectance FTIR spectra of gold substrates coated with ~ 50 :1: 10 nm copolymer brushes of Azpma and Pegma at different initial feed ratios ................................................. 174 The disappearance of azide via click reaction with alkynyl mPEG 5000 for 12 hours at room temperature as a function of the azide quantity in the copolymer brush with different feed ratios ...................... 177 xvi Figure 5.22. Increase in thickness by the click reaction with alkynyl mPEG 5000 for 12 hours at room temperature as a function of azide content in the copolymer brushes ............................................................ 178 Images in this dissertation are presented in color. xvii Scheme Scheme 1.1. Scheme 1.2. Scheme 1.3. Scheme 1.4. Scheme 1.5. Scheme 1.6. Scheme 1.7. Scheme 1.8. Scheme 1.9. Scheme 1.10. Scheme 1.11. Scheme 1.12. Scheme 1.13. Scheme 2.1. Scheme 2.2. LIST OF SCHEMES Page Synthesis of polystyrene brushes on silica and cleavage of the polymers from the surface ....................................................................... 7 Mechanism of ATRP ............................................................................... 13 Synthesis of PMA-b-PMMA-b-PHEMA triblock copolymer brushes ..................................................................................................... 15 Polystyrene brushes grown by nitroxide-mediated polymerization ......... 18 Illustration of the intercalation of the quaternary ammonium alkoxyamine initiator into laponite by cation exchange and the subsequent formation of ionically bonded Polystyrene chains by surface-initiated NMP of styrene using a sacrificial alkoxyamine initiator ..................................................................................................... 19 Polymer brushes grown by RAFT polymerization of MMA ................... 20 General procedure for the preparation of atom transfer addition modified surfaces and subsequent diblock copolymer brush formation via RAFT ................................................................................. 21 Surface-initiated cationic polymerization of 2-oxazolines ...................... 23 Surface-initiated anionic polymerization of styrene on gold ................... 24 Immobilization of the DPE initiator followed by polymerization of the styrene homopolymer to form PS brushes ......................................... 25 Surface-Initiated Anionic Polymerization of Ethylene Oxide from multi-walled nanotubes (MWNTs) surface ............................................. 27 Surface-initiated ring opening polymerization of lactide ......................... 29 Schematic description of the formation of p(Nb-diMeOH)-b-p(Nb- COOMe) diblock copolymer brushes on a gold surface .......................... 31 Synthesis of tethered PtBA film on gold surface ..................................... 46 Possibilities of low kt due to steric constraints ........................................ 48 xviii Scheme 2.3. Scheme 2.4. Scheme 2.5. Scheme 2.6. Scheme 2.7. Scheme 3.1. Scheme 3.2. Scheme 3.3. Scheme 4.1. Scheme 4.2. Scheme 4.3. Scheme 4.4. Scheme 4.5. Scheme 4.6. Scheme 4.7. Scheme 5.1. General scheme of synthesis of substituted styrene via witti g reaction ..................................................................................................... 50 Synthesis of substituted styrenes ............................................................. 55 Surface initiated polymerization of styrene derivatives ........................... 56 The initiation, propagation and termination steps of ATRP .................... 64 Energy diagram for different substituents ................................................ 71 Surface initiated polymerization of MMA from silicon and gold surface ...................................................................................................... 82 Block copolymer formation on flat surface ............................................. 85 Formation of cross linked initiator on gold surface ................................. 91 Synthesis of 3-Methylthienyl Methacrylate (MTM) ................................ 1 15 Synthesis of [2, 2’ :5 ’, 2’ ’-terthiophen]-3’-ylmethyl methacrylate ......... 1 16 Synthesis of trichlorosilane initiator ........................................................ 117 Initiator SAM deposition on ITO surface ................................................ 118 Surface initiated polymerization of MTM from gold surface, TTMM from gold surface and TTMM from ITO surface ....................... 119 Electrochemical cross linking of PTTMM brush on ITO surface ........... 123 Heterocoupling of PTTMM brush with EDOT on ITO surface .............. 132 Synthesis of random copolymer brushes grafted on gold substrates and click functionalization ....................................................................... 147 xix LIST OF ABBREVIATIONS A AIBN AF M ATR ATRA ATRP Azpma 2-BPB bpy CL CM CV CYC LAM DMF dNbpy anbpy DP EDOT Absorbance 2,2’-Azobisisobutyronitrile Atomic force microscopy Attenuated total reflectance Atom transfer radical addition Atom transfer radical polymerization Average cross—sectional area of polymer chains Azidopropyl methacrylate 2—Bromopropionyl bromide 2,2 ’-Bipyn'dine 6—Caprolactone Chain transfer constant to monomer Cyclic Voltammetry 1,4,8,l 1-Tetraazacyclotetradecane N,N-dimethy1formamide 4,4'—Di(5-nonyl)—2,2'-bipyridine 4,4'-Di(n—nonyl)-2,2'-bipyridine Degree of polymerization Molar extinction coefficient 3,4 ethylenedioxy thiophene XX Egma EtOAc F TIR F GC GPC HEMA HMTETA ITO "p kpabs kpapp kt ktapp kact kdeact Lactide LB LCST Ethylene glycol methyl ether methacrylate Ethyl acetate free energy Fourier transform infrared Grafting density Gas chromatography Gel permeation chromatography 2-Hydroxyethyl methacrylate 1,1 ,4,7,10,10-Hexamethyltriethylenetetramine Indium tin oxide Polymerization rate constant Absolute propagation rate constant Apparent polymerization rate constant Termination rate constant Apparent termination rate constant Activation rate constant Deactivation rate constant 3,6-Dimethy1—l ,4-dioxane-2,5-dione Langmuir-Blodgett Lower Critical Solution Temperatures xxi MA Me4Cyclam Me6TREN iCP MMA mPEG MS MPS MTM MUD OTf PAA PBA Methyl acrylate Methyl 1 ,4,8,1 1 -tetramethyl-1,4,8, I l-tetraazacyclotetradecane trz's[2-(Dimethylamino)ethyl]amine Microcontact printing Methyl methacrylate Number average molecular weight methoxy Poly(ethylene glycol) Mass spectroscopy Mercaptopropyltrimethoxysilane 3-Methylthienyl Methacrylate Mercaptoundecanol Weight average molecular weight Multiwalled carbon nanotubes Avogadro’s number Nitroxide-mediated polymerization Nuclear magnetic resonance Norbomene Triflate Poly(acrylic acid) Poly(n-butyl acrylate) xxii PCL PDI PDMS PEI PEG Pegma PGMA Ph PHEMA phen PLA PLED PMA PMEMA PMMA PMDETA PNIPAAM POEGMA PPZ PS PtBA PVK PVP Poly(a-caprolactone) Polydispersity index calculated as Mw/Mn Polydimethylsiloxane N-propionylethylenimine Poly(ethylene glycol) Poly(ethylene glycol methyl ether methacrylate) Poly(glycidyl methacrylate) Phenyl Poly(2-hydroxyethyl methacrylate) 1,10-Phenanthroline Poly(lactide) Photo light emitting diode Poly(methacrylate) Poly(2-(N-morpholino)-ethyl methacrylate) Poly(methyl methacrylate) N ,N,N ',N ',N "-pentamethyldiethylenetriamine Poly(N-isopropylacrylamide) Poly(oligoethylene glycol methyl ether methacrylate) Poly(phosphazenes) Polystyrene Poly(tert-butyl acrylate) Poly(vinyl carbazole) Poly(4-vinyl pyridine) xxiii QR RAFT ROMP ROP SAM SEM SI tBA TEAP TEM TEMPO THF TMEDA TREN TTMM 4-VP Volume fraction Quenching and re-initiation Random copolymer Reversible addition-fragmentation chain transfer Ring-opening metathesis polymerization Ring-opening polymerization Rate of polymerization Radius of gyration Hammett constant Self-assembled monolayer Scanning electron microscopy Surface initiated tert-Butyl acrylate Tetraethyl ammonium perchlorate Transmission Electron Microscopy 2,2,6,6-Tetramethylpiperidinyloxy Tetrahydrofuran Tetramethylethylenediamine T ris[2-aminoethyl]amine [2, 2 ’:5 ’, 2 ’ ’-Terthiophen]-3 ’-ylmethyl methacrylate Ultraviolet 4-Viny1 pyridine xxiv XPS X-ray photoelectron spectroscopy p Reaction constant XXV Chapter 1 Introduction 1.1. Polymer Brushes — a brief definition The term “polymer brush” refers to an assembly of polymer chains tethered at one end to a surface and packed sufficiently dense that the polymer chains are forced into an extended conformation. Polymer brushes can be formed by using either a ‘ grafting to’ technique, where the polymers are absorbed onto a surface,1 or a ‘grafting from’ technique, where the polymers are grown directly from the surface (Figure 1.1).?"3’4 This study emphasizes the ‘grafting from’ technique due to its ease of synthesis, and the high density of polymer chains in the resulting film. A number of polymerization methods have been used to synthesize polymer brushes by the ‘grafting from’ technique, including free radical polymerization,5 as well as controlled polymerizations such as reversible addition-fragmentation polymerization (RAFT), 6 nitroxide mediated radical polymerization (NMP),7 ionic polymerization,8 ring opening metathesis polymerization (ROMP),9 and atom transfer radical polymerization (ATRP).10 Surface-initiated ATRP is one of the most important techniques for generating polymer brushes11 from surfaces 10 . . 12 such as gold or s111ca. a) b) polymer surface functionality monomer initiator monolayer l \ endgroup assemble functionality onto polymerize surface polymer Figure 1.1: Graphical representation of a) the ‘grafting to’ technique, and b) the ‘grafting from’ technique (Reprinted with permission from Gregory L. Whiting, Ph.D. thesis 2006, University of Cambridge). 1.2. Structure of Polymer Brushes polymer brush attachment site Figure 1.2: a) Schematic showing a thin film comprised of polymer chains in a random conformation. b) Schematic showing a thin film comprised of polymer chains in a brush conformation (Reprinted with permission from Gregory L. Whiting, Ph. D. thesis 2006, University of Cambridge). Due to steric interactions, the surface—attached polymer chains in a dense polymer brush film are expected to be more extended than randomly coiled polymers. Figure 1.2 compares the morphology of polymer chains in a random conformation (spin coated polymer) and in a brush conformation. A simple energy balance model can be constructed to compare the deformation of a polymer chain in a brush conformation and a randomly coiled polymer. Generally, maximization of entropy for polymer chains can be achieved by adopting a random coil conformation. However, this is not possible for polymer brushes due to the end—grafted chains and steric interactions from the overlap of other polymers. The Alexander model takes these two considerations into account for an end grafted polymer chain.13’14 This model, developed using the Flory approximation (from random walk experiments), shows that the free energy, f, for a polymer chain in a good solvent in the brush conformation is given by Equation 1.1.The model assumes that the concentration of repeating units is constant throughout the polymer brush film, so that (p = No/h (q): volume fraction of the grafted polymer chains) and that there is a fixed distance h, where all of the polymer chains terminate. Here, N is the number of repeating units of diameter a, o is the grafting density, T is temperature, k is the Boltzmann constant, and v is an excluded volume parameter. Minimizing this expression with respect to h (df/dh =0), provides Equation 1.2, which indicates that the thickness of the polymer brush film, in solution, varies linearly with the degree of polymerization. h 2+ thp ~kT f Na at 311 WIN2 df_ 1’3 ar*""(““~az“ ,2 > “(arwlwlwazl 1'2 This result implies that when the distance between the grafting sites is less than the radius of gyration, the chains stretch somewhat to minimize this steric energy. This is actually . . . . 1/2 . 1/2 observable when the unstretched charn dimensron 1ncreases as N (srnce Rg = N a), the brush thickness increases linearly with respect to N. So, it seems reasonable to accept that the polymer chains must be deformed somewhat from a completely random conformation, and the deformation will become more pronounced for higher degrees of polymerization (N). A similar strategy can be applied to a polymer brush film in a dry . . . .. . . . 2/3 state. In thrs case it can be shown that the equrlrbnum thrckness 1ncreases as N , implying that the polymer chains are also stretched and deformed from a random configuration in the dry state. This deformation leads to changes in physical properties of the brush film. Yamamoto, et al.15 showed that poly(methyl methacrylate) (PMMA) brushes consistently have higher glass transition temperatures (T g) than spin-coated PMMA films of similar thickness, and the corresponding brush films consistently have 8 OC higher T gs at higher thickness. Though the mechanism is obscure, the increased T g must be due to the stretched nature of the polymer brush chains.14 It should be pointed out that the authors formed spin-coated films from free polymer synthesized from the same polymerization as the surface-attached films. Therefore, the two polymer films (brush and spin-coated) are expected to have similar properties, and the Tg difference can not be due to differences in chain length, polydispersity or stereochemistry.16 Overall, these data suggest that polymer brushes exist in a stretched configuration different from that of a free chain in solution. 1.3. Synthesis of Polymer Brushes - recent advances There are two distinct pathways to achieve polymer brush architectures, the “grafting to” and “grafting from” approaches. “Grafting to” utilizes preformed polymers with an end functionality that interacts or anchors with the substrate surface. This generally involves a chemical or physical adsorption process. A limitation to this type of application is that the process is diffusion limited so that as more and more polymer chains are attached to the surface, the ability for a new polymer chain to diffuse to the surface of the substrate is greatly hindered (Figure 1.1a). On the other hand, the “grafting from” or surface initiated polymerization approach places the initiating groups directly on the surface, eliminating the diffusion problem to a great extent and providing control of the polymer chain length as well as the grafting density. Locating the initiator on the surface means that the growing polymer is attached to the surface and the monomer diffuses to the growing chain end throughout the reaction; 1.6., the growing chain eventually extends from the substrate (Figure 1.1b). This technique is amenable to a large . .. . . . . 17 number of surface 1n1t1ators as well as varrous polymenzatrons mechamsms as discussed below. 1.3.1 Polymer brushes by free radical polymerizations Surface initiated free radical polymerization was investigated in detail by Minko . 18 and coworkers using both theoretical and expenmental approaches. They attached azo and peroxide initiators to solid substrates by either physisorption or chemical immobilization, and grew polystyrene via a free radical pathway. The kinetics for polystyrene grown from a silica surface with attached azo—initiators was followed by in situ ellipsometric measurement of the film growth. The resulting kinetics showed a linear dependence of the polymerization rate on the concentration of the immobilized initiator, and an inverse square root dependence on the initiator concentration in bulk, which is in accord with conventional free radical polymerization. However, their method for anchoring initiators led to a low grafting density and side reactions. Later, Riihe et a]. developed a one-step initiator anchoring strategy and initiated free radical polymerization 5 . . . of styrene from surfaces. Scheme 1.1 shows the polymenzauon reactron and the method used to detach the polymer chains from silica gel. The initiator has three important functionalities: ( 1) an azo group that generates free radicals upon heating or UV irradiation, (2) a chlorosilane that allows the initiator to be anchored to the surface through reaction with silanol groups of the silica substrate, and (3) an ester that can be hydrolyzed to detach the polymer brushes from the surface. After polymerization of styrene, the ester bonds that connect the polystyrene to the surface are easily cleaved to determine the molecular weights of the polymers, which allow comparisons between the free radical polymerization in solution and from a surface. From the molecular weight and grafting density, they found that the average distance between tethered polystyrene chains was 2-3 nm, smaller than the radii of gyration of the corresponding polymer molecules. 5102 surface 1 O I + Q l g) > a: 2: Z \\ Z +5 m g o base 8102 Surface 5? (.0 E > o O 01 Z \\ r 3 m Styrene toluene, heat 4V2 8 g Me 0 Me CN .. '3 Mai—N W‘ b d a as non on e M 00-5 e O + polystyrene * n p—TsOH MeOH/Toluene 8 M Me CN e - N e ,0 1., E h-O—S.i_/\/OH Me N— N ill/1e + O .9 co ‘ "n Scheme 1.1: Synthesis of polystyrene brushes on silica and cleavage of the polymers from the surface Prucker and Rtihe also investigated the kinetics and mechanism of surface~ initiated free radical polymerization from a self assembled monolayer of azo initiators attached to the surface of silica particles.19 The rate of decomposition of the surface- immobilized initiator was monitored by DSC, as well as by quantifying the amount of nitrogen generated from the azo component during decomposition. After polymerization, the polymer chains were detached from the surface and the molecular weights of the de- grafted polymers and their distribution were studied as a fimction of the reaction parameters during polymerization. They also irradiated selected areas of surfaces with UV to pattern th1n polymer layers by surface-1n1t1ated free radrcal polymerization.2 Other substrates have been studied. For example, Velten et al. anchored cation-bearing peroxides to mica surfaces via ion exchange, and initiated styrene polymerization from . . . . 21 the 1n1t1ators to obtam surface-bound polymers. Wittmer and coworkers predicted strong differences between polymer brushes grown from surfaces and polymers generated in solution.22 They assumed that long chains are more efficient at adding monomers than short chains because they are more mobile and more accessible to monomers, and thus polymer brushes formed at the surface should have a higher polydispersity than the same reaction occurring in solution. However, the experimental data of Prucker and Riihe for polymers cleaved from surfaces indicate a small or no effect, with PDls ranging from 1.5 to 2, which is close to the PDI of free radical polymerizations in solution. Figure 1.3 shows the molecular weights and polydispersities of the detached polymer brushes. 175 ~ . ‘- 1 .:.—' 150 r - a _ 1 a, 125 ~ - ‘2: r l 37-; 100+ t 2 r 1 75 r 1:1 J 50 9 L A r 1 1 . l 1 1 . l . 1 0 2 4 6 8 1O 12 14 time [h] Figure 1.3: Time-dependent properties of polymer chains grown by surface- initiated free radical polymerization of styrene: molecular weight M“. (Reprinted with permission from Macromolecules 1998, 31, 602-613. Copyright 1998 American Chemical Society) 10 r 3" . as 8" c» i a) 6 P o. . 2 4 ~ 3 , 9:, 2 ’ to . 0 r- 0 2 4 6 8 10 12 14 time[h} Figure 1.3: Time—dependent properties of polymer chains grown by surface- initiated free radical polymerization of styrene: Grafting densities 5(PS) (PS=Polystyrene). (Reprinted with permission from Macromolecules 1998, 31 , 602-613. Copyright 1998 American Chemical Society) 10 2.0~ Cl- C " D .4 2" _ . E 1.9 2 b g 1.81— ~ 8 . 8 .22 1'7” 1 'O l. >~ '5 1.6~ « 0. i 1.5» - o 2 4 6 8 L10 time[h] 12114 Figure 1.3: Time-dependent properties of polymer chains grown by surface- initiated free radical polymerization of styrene: Polydispersity of the covalently attached polymers. (Reprinted with permission from Macromolecules 1998, 31, 602-613. Copyright 1998 American Chemical Society) Rtihe and coworkers expanded surface-initiated free radical polymerization to the preparation of block copolymer brushes, where one block was synthesized by ROP.23 A PCL (poly(caprolactone)) macroinitiator containing azo groups was physisorbed on a silicon oxide surface to initiate the radical polymerization of the other monomer to create 24 . . hydrophobic layers on hydrophilic surfaces. Zhao et al. later appl1edth1s strategy to the synthesis of amphiphilic Janus silica particles. Azo initiators were anchored to silica 11 particles, and the modified particles were suspended in a mixture of styrene and water. The particles were adsorbed at the liquid-liquid interface, with one hemisphere of each particle immersed in an aqueous phase, and the other in a styrene phase. After initiation at an elevated temperature (90 OC), poly(sodium methacrylate) grew from the hemisphere immersed in water, and polystyrene chains grew from the particle surface immersed in styrene. Thermogravimetric analysis and IR spectral data confirmed the grafting of polymer brushes on the surfaces. 1.3.2 Surface-initiated controlled radical polymerization Of the various polymerization methods used for polymer brush synthesis, controlled free radical polymerization (CRP) has attracted much attention due to its simplicity, wide functional group tolerance, and versatility compared to ionic processes. CRP provides control over the polymer chain length and length distribution, and also offers the possibility to design grafted polymer chains with controlled architectures. The most widely used CRP techniques are ATRP, RAFT, and NMP. All of these techniques have been used to build up highly dense polymer brushes from surfaces of various architectures (planar and nanoparticles). 1.3.2.1 Polymer brushes by Atom Transfer Radical Polymerization (ATRP) ATRP is a versatile technique for the polymerization of various vinyl monomers, primarily acrylates and methacrylates. Most aspects of ATRP have been 2526,27 reviewed. The mechanism for ATRP is shown schematically in Scheme 1.2. In ATRP systems, reversible termination is used to reduce the steady state concentration of growing radicals and suppress bimolecular radical termination reactions. Once a radical is generated from a dormant initiator, it can either add monomer or deactivate by reacting l2 with a metal complex (such as Cu(II)) to regenerate the dormant initiator and a metal in a lower oxidation state. The controlled nature of ATRP results from the equilibrium strongly favoring the dormant species. activation/ _ deactivation propagation termination k8 kt P——x a _______ p .___.___.. 1LT kd T' kp monomer Scheme 1.2: Mechanism of ATRP Polymers with low polydispersities are produced by ensuring that the initiation rate is faster than prOpagation with minimal termination. Since the dormant species is a growing polymer chain capped with a halogen atom, it is possible for a polymer chain to act as a macroinitiator, making ATRP a readily applicable technique for producing block copolymers. The kinetics of ATRP is shown in Equation 1.3.18 In this expression [M], [13.], [Cu(I)], [Cu(II)X], and [PX] are the concentrations of monomer, active polymer radicals, copper(l), copper(II) halide, and halide-capped polymer chain respectively. The rate constants kapp, kp, kact and kdeact refer to the apparent rate constant, the propagation rate constant, the activation rate constant and the deactivation rate constant respectively and finally, Kgq is the equilibrium constant. As per this expression, ATRP follows first order kinetics with respect to monomer concentration ([M]). 13 .. .. . .. [011'] RP - KapptMJ - KprP 11M] - kaquIII—mcuu X] [M] 1.3 Where» K _ kact __ [P'][Cu“X] eq kdeact [CU Il [PX] However, during surface initiated polymerization, the total amount of polymer synthesized on the surface is very small, and the concentration of monomer in solution remains essentially constant. This constant monomer concentration should lead to linear kinetic plots for surface-initiated polymer brush film thickness (proportional to the polymer molecular weight) with respect to time.28 1.3.2.1 Examples of Polymer Brushes via ATRP Surface—initiated ATRP has been the most widely employed methodology for the formation of polymer brushes, and Figure 1.4 shows examples of common monomers used to synthesize polymer brushes via ATRP. Figure 1.4: Examples of monomers that have been used to synthesize polymer brushes via ATRP. 14 Surface initiated ATRP can be applied to nonplanar surfaces such as nanoparticles. 29 In an early example, Huang, et al. used ATRP to grow polyacrylamide brushes from silica gel, and obtained a ~10 nm thick film after 40 hours at 130 C)C. Husseman et al.7 improved control over the growth of polymer brushes by adding a sacrificial initiator to decrease the concentration of active chains, and therefore reduced chain termination by coupling and disproportionation. However, the free initiators produced significant amount of polymer in solution that must be removed. An alternative method is to add a . . . 30 . .. . Cu(II) salt to the polymerrzatron solutron, and shift the equ111br1um more toward dormant chains. This is a ‘surface confined’ polymerization since no initiator is explicitly added to the solution, and therefore only small amounts of polymer is generated in solution. 15 o s~(CH2)-~o—ti-cH-ar 1‘ CH3 HEMA, CuBr/MeaTREN CuBr2/2anbpy, DMF/THF, 4o °c o cH3 1 Au S1-1-*(CH2) o— c-— CH CHz-c Br CH3 COZCHZCHZOH J \ofi Sn(Oct)2, Toluene. 95 00 V COZCHZCHZOH CH2 0 H 0 H O C‘C‘O‘C‘C‘O H CH3 CH3 n Scheme 1.3: Synthesis of PMA-b-PMMA-b-PHEMA triblock copolymer brushes Similarly, styrene and methyl acrylate were polymerized from an initiator layer of 2-bromoisobutyrate immobilized on silicon wafers. The polymerizations showed a linear . . . . . . 31 . increase 1n the polymer thrckness With reactron t1me. Controlled growth was achieved by the addition of a deactivating Cu(II) species. A combined study of the control of . . . . . . . 32 1n1t1ator densrty on gold and s111con substrates was carried out by Bao et. al. They prepared Au and 8102 substrates with various immobilized initiator densities using systems (size of initiator and diluent were matched) that should lead to a homogeneous l6 distribution of the initiator on the surface. Variations in the polymerization rate on both substrates were observed as a function of initiator density and it was found to be consistent with the decrease in bimolecular termination as well as a decrease in the number of chains on the surface when the initiator density drops below 10% on Au surfaces. A related effect of initiator density was reported by Wu et al.33 In their study, an initiator density gradient was formed on a silica surface, brushes were synthesized from the surface, and the thickness was found to vary systematically (as measured by ellipsometry) at different positions along the substrate. These results show that the polymer brush thickness decreases slightly with the reduction of the initiator density until a crossover point, where the polymer film thickness decreases markedly. At the crossover point the density of polymer is too low to force the polymer chains into a stretched conformation. As a result, the polymer chains do not interact with one another, and are present in the ‘mushroom’ regime as randomly coiled chains on the top of the surface. Since the polymer chains are already spaced far enough apart to not interact with each other in the mushroom regime, the thickness of the polymer chains does not correlate with initiator density. Though room temperature polymerizations are desirable to avoid thermal polymerization, ATRP has been carried out at elevated temperature to increase the rate of polymerization. However, thiol on gold SAMs are unstable at temperatures above 34 . . . . 60 CDC. One way to 1ncrease polymerization rates and decrease the react1on temperature, 17 . . . 10, 35 . . . . rs to use aqueous polymenzatron systems. It rs thought that the hrgh drelectrrc constant of the polar solvent increases the activity of the ATRP catalyst system. Another useful method for increasing the polymerization rate is to use ligands with high stability constants such as Me4Cyclam. Moreover, ATRP catalysts based on Me4Cyclam can grow 100 nm thick film of poly(tert-butyl acrylate) (PtBA) in 5 . 36 . . . . minutes. ATRP rs also a useful technrque for the format1on of block copolymers, srnce the polymer chains can act as macroinitiators for a new monomer. Several examples of block copolymer brushes via ATRP have been shown. Kim et al. used a simple quench and re-initiation (QR) approach to grow PMA-b-PMMA-b—PHEMA triblock copolymer brushes on Au (Scheme 1.3).37 Growing polymer brushes were quenched with a concentrated CuBrz/ligand solution, preserving the Br atoms at the chain ends for subsequent re-initiation of the next polymer block. The efficiency of the QR scheme was found to be superior over a simple solvent washing procedure, which resulted in a higher loss of active chains. In another example, ABA triblock copolymer brushes of polystyrene and poly(methyl methacrylate) were synthesized on silica surfaces. 38 Treatment with different solvents reversibly altered the surface morphology leading to applications in switchable and responsive surfaces. As the synthesis of polymer brushes via ATRP has become more controlled and an increasingly large number of monomers have been used, potential applications of this film morphology is now being explored. 18 1.3.2.2 Polymer brushes by Nitroxide Mediated Polymerization (NMP) The livingness of NMP depends on the reversible capping of the active chain-end radical with a nitroxide leaving group. Husseman er al. demonstrated that polystyrene (PS) chains could be grown from the surface by NMP (Scheme 1.4).7 They first attached alkoxyamine initiators onto the surface and then heated the system to 120 0C to initiate radical polymerization. During the initiating process, the stable nitroxide radical 2,2,6,6— tetramethylpiperidinyloxy (TEMPO) is cleaved and reversibly caps the chain-end radicals to control radical propagation. The addition of free alkoxyamine initiator provided more control over the molecular weight, but solution polymerization could not be avoided. CW - iii/glib Scheme 1.4 Polystyrene brushes grown by nitroxide-mediated polymerization Hawker and coworkers applied NMP to photolithography by patterning polymer brushes into well-defined hydrophobic and hydrophilic domains (Figure 1.5).39 PtBA brushes (hydrophobic) were synthesized by surface-initiated NMP and then hydrolyzed to form hydrophilic poly(acrylic acid) (PAA) brushes. Recently, Konn er al.40 grew polystyrene chains from the surface of synthetic Laponite clay platelets by nitroxide-mediated polymerization of styrene using N-tert-butyl—N-[l-diethylphosphono—(2,2- dimethylprOpyl)] (DEPN) as mediator (Scheme 1.5). A novel water-soluble quaternary ammonium alkoxyamine was synthesized and intercalated into the clay galleries by 19 cation exchange. Polystyrene (PS) chains with controlled molecular weight and narrow polydispersities were then grown from the organoclay which displayed a high stability and dispersibility in organic solvent as well as into polar and nonpolar monomers. Z— +\ DEPN Styrene, Toluene Styryl-DEPN, O DEPN N+ 110 C, 32h / I \ DEPN= ‘o-N Eto.) < EtO’ Ps0 Scheme 1.5: Illustration of the intercalation of the quaternary ammonium alkoxyamine initiator into laponite by cation exchange and the subsequent formation of ionically bonded PS chains by surface-initiated NMP of styrene using a sacrificial alkoxyamine initiator. (Reprinted with permission from Macromolecules 2007, 40, 7464. Copyright 2007 American Chemical Society.) 20 1.3.2.3 Polymer brushes by reversible radical addition-fragmentation chain transfer R~S 3 CN Ph CN COOMe a» 8 cu N=Nzl< 0 34 n Ph V‘oue 110°C 3 8Y8 ’ n Y3 Ph ©-// Scheme 1.6 Polymer brushes grown by RAFT polymerization of MMA ATRP is arguably the most widely used controlled radical polymerization technique for the preparation of surface initiated polymer brushes. However, RAFT offers potential benefits over ATRP in the polymerization of functional monomers. RAFT is a versatile technique based on a reversible degenerative chain transfer mechanism in which thiocarbonylthio compounds act as chain transfer agents providing 41 42,43,411 controlled growth of polymer chains (Scheme 1.6). Boyes et a1. modified surfaces with RAFT chain transfer agents and synthesized a series of homopolymer and diblock copolymer brushes (Scheme 17).“ They cleverly converted an immobilized ATRP surface initiator to the RAFT chain transfer agent surface by an atom transfer addition reaction. RAFT chain transfer agents were anchored to the surface of silicon wafers, and then homopolymer brushes of PMMA and PS and diblock copolymer brushes of PMMA-b-PS, PMMA-hPDMAEMA and PS-b-PMA were grown from the substrates. 21 RC“ 1)Piranha solution Me O 3' Me OH 0 Br O'“SI“ (CH2T‘O Me Me HMS Me 2) CI“ SMI" e(Cl“I2)‘1"10 Mee eToluene 56 °C, 4—6 hrs S 8.3 .. of .0 Me Me 3 Me "— o = o-sr- CH 0 s O—SeICHZi: Me CuBr, Cu(O) Me( 2Y1: JKQ N,N',N"-pentamethyldiethylenetriemine, 85 °C, 48 h Monomer 1 Chain transfer agent 2,2'-azobisisobutyronitrile Where R1= H, R2 = or II O)=o Me Me \ O—SNIe Me(CH2)—IOMSS Where R1 = CH3, R2 = n)K© >=O 0' 0:0 Monomer 2 o 0 Chain transfer agent 2,2'-azobisisobutyronitrile Me Me O—SMI: (CH2)"100 Scheme 1.7: General procedure for the preparation of atom transfer addition modified surfaces and subsequent diblock copolymer brush formation via Reversible Addition-Fragmentation Chain Transfer Polymerization. 22 1.3.3 Polymer brushes by Cationic Polymerizations Substantially less work has been reported on the application of cationic polymerization for the synthesis of polymer brushes from surfaces. Jordan and Ulman46 initiated cationic ring opening polymerization of 2-substituted 2-oxazolines from gold surfaces, resulting in a 10 nm poly(N-acylethylenimine) brush after 7 days of polymerization under reflux in chloroform (Scheme 1.8). They also showed that the polydispersity of the resulting linear polymers were narrow because of the highly living character of the ring—opening polymerization. This method allows block-copolymers synthesis by consecutive addition of different 2-substituted monomers. Furthermore, terminal functionalities are easily introduced by terminating the polymerization with nucleophiles. They extended this strategy to gold nanoparticles to the preparation of . . 47 . . 48 . functional composrtes at the nanometer scale. Zhao and Brittam deposrted SAMs terminated with cumyl methyl ether moieties on silicon wafer surfaces. Activation with TiCl4 in the presence of styrene and di-tert-butylpyridine, a proton scavenger, led to 30 nm thick polystyrene brushes in less than an hour. The brush growth was carried out at — 78 0C to suppress chain transfer reactions. The livingness of the polymerization was confirmed by re-initiating the polymer chains to grow additional polystyrene. Recently, Patton et (11.49 developed a new class of polymer brushes based on the hybrid inorganic- organic polymer backbone of poly(phosphazenes) (PPZs). PPZ brushes were synthesized using surface-bound phosphoranimines as active sites for living cationic surface-initiated polymerization of chlorophosphoranimines. 23 OH O IOZCFS 5‘ Et-{gj sozcr—‘3 o A o/ N/ o e N + Au ._ \SOZCF3 \__/ OTf EIOH; (1.; 16 h 1 day; 0 0C -—> reflux c .-= 1on1nol/l vapor phase R R HS S alk l alk l csz _ => 3 E 7 deals. N 1 day: N po'ymer CHCl3; reflux O ”-1 CHCl3; rt. 0 n-1 SAM R R Scheme 1.8: Surface-initiated cationic polymerization of 2-oxazolines. 1.3.4 Polymer brushes by Anionic Polymerizations The living nature of anionic polymerization has made it an attractive choice for the synthesis of well defined polymer brushes via the ‘grafting from’ approach. Early examples involving “grafting to” approaches such as the reaction of lithiated polystyrene . . . . 50 51 . . chains With surfaces were first described in the 19705. ’ This usually involved the reaction of carbanions of living polymers of polystyrene or polyvinylpyridine with surface bound chlorosilane groups to bind the macromolecule covalently to the silica surface. Later, Jordan et al.52 initiated anionic polymerization of styrene from gold substrates. Initially, a bromobiphenyl SAM group was converted to the corresponding lithium species by metal-halogen exchange with sec-butyllithium, and after addition of 24 styrene, uniform 18 nm thick films grew in 3 days. The initiating efficiency and grafting density were calculated to be 8% and 3.2-3.6 nmz/chain respectively. Ingall et al. . . . . . . . 53 . . polymerized acrylonltnle from 8102 usmg a Similar strategy. A bromine terminated SAM, formed from 3-bromopropyltrichlorosilane, was lithiated with lithium di—tert- butylbiphenyl, and subsequent addition of monomer resulted in a 245 nm thick film after ooomooowooo Scheme 1.9: Surface-initiated anionic polymerization of styrene on gold Advincula et al. used n-butyllithium to activate diphenyethylene-terminated SAMs for the anionic polymerization of styrene54 from gold and silicon substrates. The polymerization was slow, and reaction times of several days produced thin films (up to 16 nm). The results for silicon surfaces showed large variations in film thickness for similar reaction conditions. The addition of tetramethylethylenediamine (TMEDA) produced a thicker film (26 nm). To demonstrate the living nature of the polymerizations, polystyrene-block—polyisoprene and polybutadiene-block-polystyrene were synthesized . . . 55 56 57 . . by the sequential addition of monomers. Later, they ’ ' extended anionic surface- initiated polymerization to silica and clay nanoparticles. (Scheme 1.10). They confirmed 25 the “living” nature of the polymerization by demonstrating a linear relationship between monomer concentration and Mn. CH2 CH “0' ~ CI-SlMez-(CHZ 11 CH2 0 o H‘O-SlMez-(CH2h-1 Bu \ O G O . ? benzene .— 'S'Mez-(CH2111 ' Umml. ©/§ > growing polymer chain Scheme 1.10: Immobilization of the DPE initiator followed by polymerization of the styrene homopolymer to form PS brushes by SIP. Baskaran et al.58 used living surface-initiated anionic polymerization (Scheme 1.11) to synthesize poly(ethylene oxide) and polystyrene brushes from the surface of MWNTs. Using a “grafting from” strategy, MWNTs were covalently functionalized with 4-hydroxyethy1 benzocyclobutene (BCB-EO) and l—benzocylcobutene—l'-phenylethylene (BCB-PE) through [4 + 2] cycloaddition. Alkoxy anions and alkyllithium anions were and MWNTs-g-(BCB-PE),, using generated from MWNTs-g-(BCB-EO),, 26 triphenylmethane and sec-butyllithium for the polymerization of ethylene oxide and styrene in THF and benzene, respectively. The initiation of ethylene oxide and styrene from the surface alkoxy and alkyl anions respectively was found to be slow due to heterogeneous nature of the reaction. But with increasing reaction time, the polymerization of ethylene oxide from the surface of MWNTs~g-(BCB-EO),, exhibited linear growth as indicated by the monomer conversion and weight percent of PEO on the MWNTs. Moreover, this process provided surface grown polymers in high conversion. Mueller and coworkerssg utilized a similar strategy for the grafting of MWNTs with the such as poly(e-caprolactone) and biodegradable and biocompatible polymers poly(ethylene oxide). In their “grafting from” methodology, an initiator was covalently attached to the MWNT through a [4 + 2] Diels-Alder cycloaddition reaction. The resulting polymer—CNT hybrid materials contained a high percentage of polymers (as high as 98%), as revealed by TGA. 27 Scheme 1.11: Surface-initiated anionic polymerization of ethylene oxide from MWNTs-g-(BCB-EO),, surfaces. Although living anionic polymerization is useful for the synthesis of well-defined brushes with low polydispersity, it has several disadvantages such as the extreme sensitivity of anionic polymerization to impurities, which necessitates the use of specialized glassware and rigorous purification and drying of reagents. Restricted monomer functionality, long reaction times and low values for final thickness of the polymer films also hinder the use of this technique for polymer brush growth. 1.3.5 Polymer brushes via surface-initiated Ring-opening Polymerization Radical surface initiated polymerization is limited to vinyl monomers. However there is a need for biocompatible and biodegradable polymers such as polylactides and polylactones for medical applications60. These polyesters are usually synthesized by the ring opening polymerization of cyclic monomers initiated by organometallic reagents in a living/controlled fashion, which is the prerequisite for the controlled grafting of polymers from surfaces. Surface initiated ring opening polymerization is generally performed after the formation of a hydroxyl- or amine terminated self-assembled monolayer which initiates the polymerization (Scheme 1.12). The most comprehensive article on polyester grafting by ring opening polymerization of lactones (s-caprolactam) was reported by Carrot e1 (11.62 Covalent grafting of polyester chains was achieved by functionalizing silica surfaces with a trimethoxysilane agent containing. an amine functionality. Initiation occurred selectively from the surface by activation of amine groups by an aluminum alkoxide, and the controlled polymerization proceeded via a 28 coordination-insertion mechanism involving the coordination of the carbonyl group of the a-caprolactam by triethyl aluminum followed by the nucleophilic addition of the amine. In related work, Choi and Langer formed an oligo(ethylene glycol) terminated SAM on gold substrate and used tin(II) (2-ethylhexanoate)2 (Sn(Oct)2) to catalyze the ROP of L— lactide from Au and silicon substrates 63 (Scheme 1.12). Polymerization for 3 days provided 12 nm thick PLA brushes on Au surfaces, and 70 nm thick films on silicon surfaces. The PLA brushes were shown to be chiral and crystallized on the surface. Later, used tin(II) octoate to grow biodegradable, aliphatic 9 Choi and coworkers, poly(ether-esters) such as poly(p—dioxane) (PPDX) and poly(l,5-dioxepan-2-one) (PDXO) from an oligo(ethylene glycol)-terminated SAM by the ring opening polymerization. However the polymerization was not “living” in nature as suggested from the imperfect . 63 . . . shapes of the resultant particles. In another study, the Silanol groups of the Silica particles were used to initiate polymerization. A 140 nm thick polylactide brush was obtained, which was consistent with thermogravimetric analysis (TGA). 29 ”’1. O O O E i T 0 t )“ XH o o x \n/\o n ,. o Sn(Oct)2 SAMs, X = O or NH Au recto/V0“ was 9 3 9 3 ‘5. ‘é- r Sl/SIOZ O’IIW \/\NH2 (MeOI3SI\/\/N\/\NH2 \ Scheme 1.12: Surface—initiated ring opening polymerization of lactide. 1.3.6 Polymer brushes by Ring-opening metathesis polymerization (ROMP) ROMP is the transition metal catalyzed ring opening polymerization of strained cyclic olefins. Via ROMP, Whitesides and co-workers grew a variety of norbomene- . . . 66 . . derived polymer brushes from Silicon surfaces. The surface-bound catalytic Sites were produced by forming a trichlorosilane—derived SAM containing norbomene groups, and then exposing the SAM to a solution of a Grubbs-type ROMP catalyst. The polymerization was controlled with rapid initiation, producing 90 nm thick brushes in 30 min. The formation of block copolymer brushes and the use of microprinting to produce patterned surfaces also were described. In related work, Mingotaud er al. immobilized a ruthenium catalyst on 200 nm Silica nanoparticles67 by reacting a methathesis catalyst bearing a hydroxyl group with silica particles modified with acyl chloride functional groups, and then performing ROMP of norbomene. The authors estimated that 30% of 30 the catalyst initiated polymerization. TEM characterization showed a core—Shell morphology which suggested the presence of the catalyst on the silica surface. Using a recently invented ruthenium catalyst, [(HgIMes)(3-Brpy)2(C1)2Ru=CHPh], Choi and coworkers demonstrated that surface-initiated ring opening metathesis polymerization (SI-ROMP) can be utilized for the formation of diblock copolymer brushes from surfaces (Scheme 1.13).68 Taking advantage of the highly improved activity of the ruthenium catalyst and the rapid initiation step of SI-ROMP, they successfully formed thin films of well-defined diblock copolymers with 5-norbomene-2-endo,3—endo-dimethanol (Nb- diMeOH) and an endo/exo isomeric mixture (44:56) of norbomene carboxylic acid methyl ester (Nb-COOMe). 31 / \ .+~C' N-R1_ — /IuflPh B Q N r / / l \ OTBS s/‘Mfi 8' $ sflMeru] OTBS +~ Au / 1,, .0 1 TBS TB TBSO/ ~ “ \ores ) (300m> O O S n 3 2) Ethyl Vinyl ether COOMe p(Nb-leTBS) p(Nb-diOTBS)- b-p(Nb-COOM0) “\OH TBAF __ W0 THF 7 n 3 n v COOMe p(Nb-diMeOH)—b—p(Nb—COOMe) Scheme 1.13: Schematic description of the formation of p(Nb-diMeOH)-b—p(Nb- COOMe) diblock copolymer brushes on a gold surface. 1.4 Applications of Surface-initiated Polymerization A significant advantage of polymer brushes compared to other surface modification methods is their mechanical and chemical stability, accompanied by a high level of synthetic flexibility towards the introduction of functional groups. This is in contrast to the physisorbed, non-bound polymer films where chemical modification by using wet chemistry is difficult to conduct. Additionally, it is now possible to grow brushes on virtually every surface (flat surfaces, particles or macromolecules), to any thickness, of every composition, incorporating a multitude of functional groups and containing series of blocks. More recent applications of polymer brushes include 32 69 . . 70 . . . nanopattemed surfaces, photochemical deViceS, new adheSive materials,71 protein- . . 72 . . 73 . 74 1 reSistant biosurfaces, chromatographic deVices, lubricants, polymer surfactants, . .. l polymer compatibilizers and many more. One of the most attractive applications of surface-initiated polymerizations is the formation of nano-patterned surfaces by soft lithography techniques that combine microcontact printing (uCP) and graft polymerization. An elegant example is that of Hawker et al. who combined photolithography with nitroxide-mediated “living” free radical polymerization to obtain patterned polymer brushes with well-defined . . 4 . hydrophobic and hydrophilic domains (Figure 1.5). 9 They extended this concept to 75 synthesize patterned polymer layers by aqueous ATRP. 33 200 um features Water droplet \ Manual . 1/2000 PAA brush PTBA brush Figure 1.5: Optical micrographs of patterned surfaces: (left image) IO-pm features in a continuous polymer brush Showing regions of poly(tert-butyl acrylate) (dark) and poly(acrylic acid) (light) and (right image) interaction of a water droplet with ZOO-um features showing an unusual wetting profile and preferential interaction with poly(acrylic acid) brush domains.(Reprinted with permission from J. Am. Chem. Soc. 2000, 122, 1844-1845. Copyright 2000 American Chemical Society.) 34 68 Recently, Huck and coworkers have shown that charge-transporting polymer brushes (polytriphenyl amine acrylate) can be synthesized from a variety of surfaces (ITO, Si02 and conducting polymer) relevant for organic electronic device fabrication. These polymer brush films contain a greater level of ordering at the molecular level and display higher charge mobility than Spin-coated films of the same polymer, which was attributed to the controlled polymer brush architecture and morphology. They also demonstrated substantial uptake of CdSe nanocrystals (with diameter in the range 2.5-2.8 run) into the polymer brush layers (Figure 1.6) and a photovoltaic quantum efficiencies of up to 50%. 76 In another report, Advincula and coworkers 77 successfully grafted hole- transporting PVK (poly(vinyl carbazole)) brushes on transparent ITO electrodes. Using cyclic voltammetry, the PVK brush was electrochemically crosslinked, to form a conjugated polymer network film. Covalent linkage of PVK led to a direct electroluminescent PLED device, in which the electroluminecent polymer layer can be simply solution-cast onto the modified ITO. 35 —0 Br u\\\\\\i\\\\\\\\\\ O \ I / 9.) O anode (ITO) Figure 1.6: Top image: PTPAA brushes and bottom image: Cartoon of inferred structure for CdSe nanocrystal infiltrated polymer brush photovoltaic device (From bottom to top) ITO-coated glass Slide modified by surface attachment of a bromine end- caped trichlorosilane self-assembled-monolayer (SAM) (blue squares), polymer brushes grown from the SAM (red lines), CdSe nanocrystals infiltrated into the brush network exhibiting some degree of phase separation in the plane of the film (small black circles), and caped with an aluminum cathode. (Reprinted with permission from Nano Lett. 2005, 5, 1653. Copyright 2005 American Chemical Society.) 36 A more ambitious challenge in surface science is the design of smart surfaces . . . 78 . . With dynamrcally controllable properties. Such surfaces have characteristics that can be changed or tuned in an accurate and predictable manner by using an external stimulus. Recently, Huck and coworkers have shown that wetting properties of surfaces modified with cationic polyelectrolyte brushes strongly depend on the nature of the counter ion. Coordination of polyelectrolyte brushes bearing quaternary ammonium groups (QA+) With sulfate anions resulted in highly hydrophilic surfaces, 9 whereas, coordination of Similar brushes with C104 rendered the surface hydrophobic.80 37 e = Counterion °e° G ’ooeoec 069599000 n O 3'“ 0 e O 0 e v o e o o eo :9 e e o e o o G 0 0 G o o o e o o 0 o o 0 e 0 e e o If + L- 2...; io—o'um-«s--...K...‘..-.....r.'.';.-...;_u_.., 5‘44; _._ 4y“... ., euurs —N— ' - ”L“ \ (A) (B) PP o m "’ 90 ¢m TFSI 105 TFSI (I; g CF 90 i- ‘ _’F3C\ — I’ 3 4" ’/ ‘ ’ \\ 75 _ \ [fit ,1 [I f _ 0 N 0 - 1 l I \ l 60 " l I \ -/ \ l \ I . _ O O— . I . ' I l 9A 45 - \ I t ,‘ l I \ l . 0‘15 ,5/0 _ l I i l I \l _ C.” \O’ \90 30 ' U (I i ' 0:,P P’zo 15 (C) a» I J Y .' 00, ,0\ ,0 PP —0’l:l 5’90 Counterion Exchange Figure 1.7: Top image: PMETAC brushes and bottom image: Change in the wetting characteristics of PMETAC brushes (height, 11 ~20 nm) after exchanging the two contrasting counterions: TFSI (a) and polyphosphate (PP) (b). c) Representation of 0 A as a function of counter ion (PP and TFSI). The plot depicts the reversible behavior of PMETAC brushes over repeated cycles of TFSI and PP counter ion exchange. On the 38 right the chemical structures of both counter ions are represented. (Reprinted with permission from Angew. Chem. Int. Ed. 2005, 44, 4578. Copyright 2005 Wiley-VCH.) Recent research has focused on the Cu(I)-catalyzed, highly specific and efficient formation of 1,2,3-triazoles via the 1,3-dipolar cycloaddition of azides and terminal . ,, . 81 . _ alkynes (“click chemistry). This methodology has been used to modify surfaces of solid metals and cells, because the reaction provides high yields, stereospecificity and . . . 82 83 . . proceeds under mild conditions. ’ Click chemistry also has been used for . . . . . 84 85 . . . . functionalizmg polymers in solution. ’ Research in nanobiotechnology and biomedical sciences often involves the manipulation of interfaces between manmade surfaces and biomolecules (and cells), which generally requires the construction of surfaces that present chemically active functional groups from non-biofouling supporting materials. . 86 . . . . Chor and coworkers used “click” chemistry to couple aZide groups at the termmal of the non-biofouling polymeric film of poly(oligoethylene glycol methacrylate) with incoming molecules of interest containing terminal acetylenes (Figure 1.8). As a model for bioconjugation, biotin was immobilized onto the poly(oligoethylene glycol methacrylate) film via click chemistry, and biospecific recognition of streptavidin was demonstrated. 39 n n n Br \/\OM6 O\/\OMe O\/\OMe O O O O O O O O O CuBr/Bipy NaN3 Click ———> ———> OEGMA Alkyne 1 S .1. J. ft HN NH Alkyne 1 H H MN$0NO$ONNM O 0 Figure 1.8: Schematic description of the attachment of biotin to polymer brush via click chemistry [Acetylene group—containing biotin compound: biotin-PEO-LC-N- pentynoate (1)]. 4O 1.5 References —_ 1 Milner, S. T., Science 1991, 251, 905. 2 Zhao, B.; Brittain, w. J. Prog. Polym. Sci. 2000,25, 677. 5 Prucker, 0.; Riihe, J. Langmuir 1998, 14, 6893. 6 Baum, M.; Brittain, W. J. Macromolecules 2002, 35, 610. 7 Husseman, M.; Malmstrom, E. E.; McNamara, M.; Mate, M.; Mecerreyes, D.; Benoit, D. G.; Hedrick, J. L.; Mansky, P.; Huang, E.; Russell, T. P.; Hawker, C. J. Macromolecules 1999, 32, 1424. 8 Jordan, R.; Ulman, R. A.; Kang, J. F.; Rafailovich, M. H.; Sokolov, J. J. Am. Chem. Soc. 1999, 121, 1016. 9 Juang, A.; Scherman, O. A.; Grubbs, R. H.; Lewis, N. S. Langmuir 2001, 17, 1321. 10 Jones, D. M.; Huck, W. T. S. Adv. Mater. 2001, 13, 1256. ll Patten, T. E.; Matyjaszewski, K. Adv. Mater. 1998, 10, 901. 12 Ramakrishnan, A. ; Dhamodharan, R.; Rfihe, J. Macromol. Rap. Comm. 2002, 23, 612. 13 Alexander, S., J. Phys. 1977, 38, 977. 14 De Gennes, P. G., Macromolecules 1980, 13, 1069. 15 Yamamoto, S.; Tsujii, Y.; Fukuda, T. Macromolecules 2002, 35, 6077. 16 Yamamoto, S.; Ejaz, M.; Tsujii, Y.; Matsumoto, M.; Fukuda, T. Macromolecules 2000, 33, 5602. 17 Advincula, R.; Riihe, J.; Brittain, W.; Caster, K. E. Polymer Brushes: 0n the Way to T ailor-Made Surfaces; lst ed.; Wiley-VCH: New York, 2004. 18 (a) Minko, S.; Gafijchuk, G.; Sidorenko, A.; Voronov, S. Macromolecules 1999, 32, 4525. (b) Minko, S.; Sidorenko, A.; Stamm, M.; Gafijchuk, G.; Senkoysky, V.; Voronov, S Macromolecules 1999, 32, 4532. (c) Sidorenko, A.; Minko, S.; Gafijchuk, G.; Voronov, 3- Macromolecules 1999, 32, 4539. (d) Luzinov, I.; Minko, 3.; Senkovsky, V-§ VOTOUOV, A; Hild, S.; Marti, 0.; Wilke, W. MaCromolecules 1998, 31, 3945. (e) Mlnko, S. S.; 41 22 Wittmer, J. P.; Cates, M. E.; Johner, A.; Turner, M. S. Europhys. Lett. 1996, 33, 397. 23 Stohr, T.; Ruhe, J. Macromolecules 2000, 33, 4501. 24 Zhang, J .; J in, J .; Zhao, H. Langmuir 2009, 25, 6431. 25 Patten, T. E.; Matyjaszewski, K. Adv. Mater. 1998, 10, 901. 26 Tunca, U.; Hizal, G.; Acar, M. H.; Tasdelen, M. A.; Yagci, Y.; Mishra, M. K. Plast. Eng. (Boca Raton, FL, United States) 2009, 73, 231. 27 Pintauer, T.; Matyjaszewski, K. Chem. Soc. Rev. 2008, 37, 1087. 28 Edmondson, S.; Huck, W. T. S. J. Mater. Chem. 2004, 14, 730. 29 Huang, X. Y.; Wirth, M. J. Anal. Chem. 1997, 69, 4577. 30 Huang, W. X.; Kim, J. B.; Bruening, M. L.; Baker, G. L. Macromolecules 2002, 35, 1175. B.; Siclovan, T. M.; Kickelbick, G.; Vallant, T.; Hoffmann, H.; Pakula, T. Macromolecules 1999, 32, 8716. 32 Bao, Z.; Baker, G. L.; Bruening, M. L. Macromolecules 2006, 39, 5251. 33 Wu, T.; Efimenko, K.; Genzer, J. J. Am. Chem. Soc. 2002, 124, 9394. 34 Huang, W. X.; Skanth, G.; Baker, G. L.; Bruening, M.L. Langmuir 2001, 17, 1731. 35 Jones, D. M.; Brown, A. A.; Huck, W. T. S. Langmuir 2002, 18, 1265. 36 B30, 2.; Baker, G. L.; Bruening, M. L. J. Am. Chem. Soc. 2006, 128, 9056. 37 Kim, J. 8.; Huang, W. X.; Bruening, M. L.; Baker, G. L. Macromolecules 2002, 35, 5410. 38 Boyes, S. G.; Brittain, W. J .; Weng, X.; Cheng, S. Z. D. Macromolecules 2002, 35, 4960. 42 41 Moad, G.; Solomon, D. H. The Chemistry of Radical Polymerization, 2nd fully revised ed.; Elsevier Ltd.: Amsterdam, 2006. 42 Le, T. P.; Moad, G.; Rizzardo, E.; Thang, S. H. SH PCT Int. Appl. W0 98 01478 A1 980115, 1998. 43 Moad, G.; Chiefari, J.; Chong, Y. K.; Krstina, J.; Mayadunne, R. T. A.; Postma, A.; Rizzardo, E.; Thang, S. H. Polym. Int. 2000, 49, 993. 44 Perrier, S.; Takolpuckdee, P. J. Polym. Sci, Part A: Polym. Chem.2005, 43, 5347. 45 Rowe, M. D.; Boyes, S. G. Macromolecules 2007, 40, 879. 46 Jordan, R.; Ulman, A. J. Am. Chem. Soc. 1998, 120, 243. 47 Jordan, R.; Nuki, W.; Yen-Ming, C.; Oskar, N.; Ulman, A. Macromolecules 2001, 34, 1606. 48 Zhao, B.; Brittain, W. J. Macromolecules 2000, 33, 342. 49 Jun, L.; Arthur, L.; Joel, H.; Derek, P. L. Abstracts of Papers, 237th ACS National Meeting, Salt Lake City, UT, United States, March 22-26, 2009, POLY-228. 50 Horn, J .; Hoene, R.; Hamann, K. Makromol. Chem. 1975, I, 329. 51 Donnet J .B.; Papirer, E. C olloq. Int. CNRS. 1975, 231, 117. 52 Jordan, R.; Ulman, A.; Kang, J. F.; Rafailovich, M. H.; Sokolov, J., J. Am. Chem. Soc. 1999, 121,1016. 53 Ingall, M. D. K.; Honeyman, C. H.; Mercure, J. V.; Bianconi, P. A.; Kunz, R. R., J. Am. Chem. Soc. 1999, 121, 3607. 54 Advincula, R.; Zhou, Q. (3.; Park, M.; Wang, S. G.; Mays, J.; Sakellan'ou, (3.; Pispas, 3.; Hadjichristidis, N. Langmuir 2002, 18, 8672. 55 Zhou, Q.; Fan, X.; Xia, C.; Mays, J.; Advincula, R. Chem. Mater. 200], 13, 2465. 56 Zhou, Q.; Wang, 8.; Fan, X.; Xia, C.; Mays, J.; Advincula, R. C. Langmuir 2002, 18, 3324. 43 57 Fan, X.; Zhou, Q.; Xia, C; Cristofoli, W.; Mays, J.; Advincula, R. C. Langmuir 2002, 18, 4511. 58 Sakellariou, G.; Ji, H.; Mays, J. W.; Baskaran, D. Chem. Mat. 2008, 20, 6217. 59 Prifiis, D.; Sakellariou, G.; Hadjichn'stidis, N.; Penott, E. K.; Lorenzo, A. T.; Mueller, A. J. J. Polym. Sci, Part A: Polym. Chem. 2009, 47, 4379. 60 Nouvel, C.; Ydens, 1.; Degee, P.; Dubois, P.; Dellacherie, E.; Six, J. L. Macromol. Symp. 2001, 175, 33. 61 Mecerreyes, D.; Jerome, R.; Dubois, P. Adv. Polym. Sci, 1999, 147, 1. 62 Carrot, G.; Rutot, D.; Pottier, A.; Degée, P.; Hilbom, J.; Dubois, P. Macromolecules 2002, 35, 8400. 63 Choi, I. S.; Langer, R., Macromolecules 2001, 34, 5361. 64 Yoon, K. R.; Koh, Y. J.; Choi, I. S. Macromol. Rap. Comm. 2003, 24, 207. 65 Yoon, K. R.; Lee, Y. J.; Lee, J. K.; Choi, I. S. Macromol. Rap. Comm. 2004, 25, 1510. 66 Kim, N. Y.; Jeon, N. L.; Choi, I. S.; Takami, S.; Harada, Y.; Finnie, K. R.; Girolami, G. S.; Nuzzo, R. G.; Whitesides, G. M.; Laibinis, P. E. Macromolecules 2000, 33, 2793. 67 Mingotaud, F.; Reculusa, S.; Mingotaud, C.; Keller, P.; Sykes, C.; Duguet, E.; Ravaine, S. J. Mater. Chem. 2003, 13, 1920. 68 Kong, 8.; Lee, J. K.; Choi, I. S. Langmuir 2007, 23, 6761. 69 Shah, R. R.; Merreceyes, D.; Husemann, M.; Rees, 1.; Abbott, N. L.; Hawker, C. J.; Hedrick, J. L. Macromolecules 2000, 33, 597. 70 Whiting, G. L.; Snaith, H. J.; Khodabakhsh, S.; Andreasen, J. W.; Breiby, D. W.; Nielsen, M. M.; Greenham, N. C.; Friend, R. H.; Huck, W. T. S. Nano Lett. 2006, 6, 573. 71 Raphael, E.; De Gennes, P. G. J. Phys. Chem. 1992, 96, 4002. 72 Amiji, M.; Park, K. J Biomat. Sci. Polym. Ed. 1993, 4, 217. 73 Vanzanten, J. H. Macromolecules 1994, 27, 6797. 74 Joanny, J. F. Langmuir 1992, 8, 989. 75 Vidal, A.; Guyot, A.; Kennedy, J. P. Polym. Bull. 1980, 2, 315. 44 76 Snaith, H. J.; Whiting, G. L.; Sun, B.; Greenham, N. C.; Huck, W. T. S.; Friend, R. H. Nano Lett. 2005, 5, 1653. 77 Fulghum, T. M.; Prasad, T.; Advincula, R. C. Macromolecules 2008, 41, 5681. 78 Lahann, J.; Mitragotri, S.; Tran, T. N.; Kaido, H.; Sundaram, J.; Choi, I. S.; Hoffer, S.; Somorjai, G. A.; Langer, R. Science 2003, 299, 371. 79 Moya, S. E.; Azzaroni, 0.; Farhan, T.; Osborne, V. L.; Huck, W. T. S. Angew. Chem, Int. Ed. 2005, 44, 4578. 80 Azzaroni, 0.; Moya, S. E.; Farhan, T.; Brown, A. A.; Huck, W. T. S. Macromolecules 2005, 38, 10192. 81 Feldman, A. K.; Colasson, B.; Fokin, V. V. Org. Lett. 2004, 6, 3897. 82 (a) Rostovtsev, V.; Green, L. G.; Fokin, V. V.; Sharpless, K. B. Angew. Chem, Int. Ed. 2002, 41, 2596. (b) Tomoe, C. W.; Christensen, C.; Meldal, M. J Org. Chem. 2002, 67, 3057. 83 Lewis, W. G.; Green, L. G.; Grynszpan, F.; Radic’, Z.; Carlier, P. R.; Taylor, R; Finn, M. G.; Sharpless, K. B. Angew. Chem, Int. Ed. 2002, 4], 1053. 84 Sumerlin, B. S.; Tsarevsky, N. V.; Louche, G.; Lee, R. Y.; Matyjaszewski, K Macromolecules 2005, 38, 7540. 85 Gao, H.; Louche, G.; Sumerlin, B. S.; Jahed, N.; Golas, P.; Matyjaszewski, K Macromolecules 2005, 38, 8979. 86 Lee, B. S.; Lee, J. K.; Kim, W. J.; Jung, Y. H.; Sim, S. J.; Lee, J.; Choi, 1. S. Biomacromolecules 2007, 8, 744. 45 Chapter 2 Substituent effects in surface ATRP of polystyrene brushes 2.]. Introduction The development of controlled radical polymerizations such as Atom Transfer . . 1, . .. . Radical Polymerization (ATRP), 2 Reversrble-Addrtion-Fragmentation Transfer Polymerization (RAFT)3 and Nitroxide-Mediated Polymerization (NMP),4’5provide powerful methods for the growth of polymer brushes from surfaces. These polymerization methods limit radical concentrations during polymerization, minimize bimolecular termination reactions and provide control over Mn and the polydispersity (PDI). However, compared to traditional radical polymerization, control comes at the expense of a substantial reduction in the polymerization rate. Recently, we reported the remarkably rapid growth of well—defined poly(tert—butyl acrylate) brushes under mild conditions (50 0C) using ATRP. The key aspect of the polymerization was the use of a highly active ATRP catalyst [Cu(1)1,4,8,11-tetramethy1-1,4,8,1 1-tetraazacyclotetradecane (Cu(l)-Me4Cyclam)] which provided 100 nm thick poly(tert-butyl acrylate) (PtBA) brushes on a flat Au surfaces in just 5 minutes (Scheme 2.1).6 Such polymerization rates are several orders of magnitude greater than typical controlled polymerizations from surfaces. The Cu(1)—Me4Cyc1am) system also enabled rapid growth of thick brush layers from other vinyl monomers such as styrene, vinyl pyridine and methacrylates. Rapid 46 growth of brushes could expand the scope and applications for polymer brushes by greatly reducing the time required for their synthesis. 1.. O o O Anisole/DMF _ Au —-S-(CH2)--O Br 7 AU —S*(CH2);1-O lCHz—CH‘j-Br 11 CuBr/ Me4Cyclam m 0 O CuBr2(anbpy)2.5O °C Me4Cyclam = E j anbpy 7' Scheme 2.1: Synthesis of tethered PtBA film on gold surface 6 The unusually rapid growth of PtBA films compared to other vinyl monomers can be rationalized as a combination of tBA’s fast propagation rate and reduced bimolecular coupling due to the steric bulk of the monomer. As shown in Scheme 2.2, sterically demanding monomers show a high propensity towards head to tail placement during polymer growth, and the same steric interactions would be expected to hinder termination by coupling (analogous to forming a head to head linkage). The generality of this phenomena can be seen in kp (propagation rate constant) and kt (termination rate constant) . 7,8 . values for the methacrylate monomers shown in Table 2.1. The first five entries show data for structurally related methacrylates obtained under similar experimental conditions. There are two important trends in the data; kt decreases as the size of the ester increases because of the steric crowding depicted in Scheme 2.2, but a more surprising result is that in some cases, kp also increases with the size of the ester group. Gilbert suggested that the 47 increases in kp may be due to the increased momentum associated with collisions , . . 9 . between higher molecular weight specres. Monomers With both high kp and low kt values, i.e. a high kp / k, ratio, should be the prime candidates for rapid polymerizations from surfaces. The last two entries show that there are limits to this approach. For example, kt for trityl methacrylate is <1><10"6 L mol'lsec'l, but the extreme size of the ester group apparently hinders any enhancement in kp. Ethylene is the least demanding vinyl monomer and the reported value for kt, 540x10.6 L mol-lsec’ , indicates facile termination by coupling and disproportionation. When combined with a kp ~242 Lmol' sec-l, the kp / kt ratio for ethylene is 0.45X10’6, which predicts ethylene as a poor candidate for rapid growth from surfaces by radical polymerization. The above arguments suggest that hp and more importantly the hp / kt ratio could be used to identify promising monomers for the rapid growth of polymer brushes from surfaces. 48 Scheme 2.2. Steric constraints favor head to tail addition and low kt values. The polymerization results for PtBA prompted us to examine other bulky monomers to see if they also would polymerize rapidly and provide thick films. Thick polystyrene brushes are difficult to grow via surface initiated ATRP (40 nm thick films in . . 6 1 hour compared to 100 and 200 nm films in 1 hour for MMA or tBA, respectively), 8 . . consistent with its small kp / kt ratio (4.3 at 60 OC) . We studied substituted styrene derivatives (Scheme 2.3) to evaluate the effects of steric congestion and electronic effects on the polymerization rate under the same conditions used for tBA system. These . . . . . 10,11,12,l3 monomers efficiently polymeriZe via free radical polymerization in solution, 49 but there are no reports of their polymerization via surface-initiated ATRP or other controlled radical processes. Table 2.1. Propagation and termination rate constants for representative methacrylates8 CH -6 6 H2O 2 kp k,(x10 ) kp/k,(>< 10 ) C) -1 -1 1 1 OR (L mol 560 ) (L mol' sec- ) --CH3 649 25.5 25 “CH2CH3 723 7.35 98 [CH3 ——C\H 740 4.52 163 CH3 “(CH2)3CH3 794 10 79 “(CH2)1 1CH3 1011 0.6 1670 9H3 —C'3-CH3 350 14 25 CH3 *C‘@>3 26 0.30 86 From data reported from other types of radical polymeriZations in 14,15,16 solution, we expected that substituents on the styrene ring may affect the polymerization rate, and enable fast growth of polystyrene brushes from surfaces. 50 Furthermore, we expected that such a study would provide a better understanding of the mechanism of surface initiated polymerizations (especially the correlation between surface polymerization rate and monomer structure) as there has been no systematic investigation of substituent effects in substituted styrenes in surface initiated polymerizations. Electron withdrawing groups Electron donating groups a a s s a 2% \ \ 6 45k. ES 6 Scheme 2.3. Substituted styrene monomers polymerized in the study Herein we report the rapid surface initiated polymerization of substituted styrenes, and the effect of various substituents on the polymerization kinetics studies. The data are discussed with regard to the Hammett equation: log (kx/kH) = po 51 2.2. Experimental Section 2.2.1. Materials Unless otherwise noted, all chemicals were obtained from Aldrich. ll—Mercapto- l-undecanol (MUD, 97%), 2-bromopropionyl bromide (2-BPB) (97%), anisole (99.7%), N,N—dimethylformamide (DMF, 99.8%), Cu(I)Br (999999.), Cu(II)Br2 (99.999%), Me4Cyclam (99%), and 4,4’—dinonyl-2,2’—bipyridine (anbpy. 97%) were used as received. Triethylamine was vacuum-distilled from calcium hydride. tert~Buty1 acrylate (tBA) (98%), MMA (99%), styrene (99%), 2-vinylanisole (95%), 4-bromostyrene, pentafluorostyrene, 4—methylstyrene and 4-tert-butyl styrene were passed through a 10 cm long column of basic alumina followed by distillation to remove inhibitors. 3,5- Bis(trifluoromethyl)styrene (97%) and 3,5-dimethylstyrene were purified by passing through a 10 cm long basic alumina column. After purification, the monomers and solvents were transferred to Schlerik flasks, degassed using three freeze-pump-thaw cycles, and then transferred into a drybox. Other monomers such as 3,5-di-tert-butyl styrene, 3,5-dimethoxystyrene, 2,6-dimethoxystyrene, 4-vinylanisole and 4- (trifluoromethyl)styrene were prepared as previously reported mostly from their aldehyde . . . . ll precursors vra the Wittig reaction. 2.2.2. Characterization Methods Film thicknesses were measured using a rotating analyzer ellipsometer (model M- 44, J. A. Woollam) at an incident angle of 75°. The data were analyzed using WVASE32 software, and thickness and refractive index determinations were performed on at least three spots on each substrate. The refractive index of the films was assumed to be 1.5 and 52 then fitted with the film thickness. Reflectance FTIR spectroscopy was performed using a Nicolet Magna-IR 560 spectrometer containing a PIKE grazing angle (80°) attachment. 2.2.3. Preparation of Initiator Immobilized Gold Substrates Gold-coated Si wafers (200 nm of Au sputtered on 20 nm of Cr on Si(100) wafers) were UV/O3-cleaned for 15 min before use and then transferred into a Nz-filled glovebag. Hydroxy-terminated SAMs were formed by immersing the Au-coated substrates in a vial containing a 1 mM ethanolic solution of MUD for 24 h. After removing the vial from the glovebag, the substrates were rinsed sequentially with ethanol and water and dried under a stream of N2. The ellipsometric thickness of the MUD layer was 10-15 A. MUD-coated substrates were transferred to a drybox filled with N2 and were dipped in a 10 mL solution of 0.12 M triethylamine in anhydrous THF at 0°C. After 1 min, 10 mL of a solution of 2—BPB in anhydrous THF (0.1 M) was added dropwise to the solution to form the immobilized initiator layer. The reaction time was limited to 2-3 min, since thiol— terminated SAMs could be unstable in the presence of the acyl bromide. After rinsing with THF, the Au substrates were removed from the drybox, and then rinsed sequentially with ethyl acetate, ethanol and deionized water (Milli-Q) and dried under a stream of N2. 2.2.4. Preparation of Immobilized Initiators on Si Substrates UV/O3 cleaned Si wafers with an ellipsometrically determined oxide thickness of 20 A were transferred to a glove bag filled with N2 and immersed in a dry toluene solution (20 mL) containing 30 11L of [11~(2-bromo-2-methyl)propionyloxy]undecyl trichlorosilane. After 24 h without stirring, the samples were removed from the solution, 53 placed in fresh toluene and sonicated for 1 min. Following additional rinsing with toluene, acetone, and ethanol, the substrates were dried under a stream of N2. The ellipsometric thickness of the initiator layer was ~ 20 A. 2.2.5. Polymerization of Substituted Styrenes from Initiators Immobilized on Au Substrates - Kinetic Experiments Note: For best results, solvents, initiators and monomers must be scrupulously purified and deoxygenated. In a Nz-filled drybox, CuBr (5.7 mg, 0.04 mmol) CuBrz, (4.5 mg, 0.02 mmol), Me4Cyclam (10.3 mg, 0.04 mmol) and anbpy (16.4 mg, 0.04 mmol) were added to a round—bottom flask containing 20 mL of a degassed solution of monomer in DMF/anisole (DMF/anisole ~1:1 (vzv), [styrene] ~ 4 M). The mixture was well-stirred and heated with an oil bath to 50 0C until a transparent, light green solution formed. The prepared solution was then transferred into a small vial containing an initiator-modified Au substrate to start the surface-initiated polymerization. After a predetermined reaction time at 50 0C, the substrate was removed from the vial, washed with ethyl acetate and THF sequentially, and then dried under a flow of N2 in the drybox. The same conditions were used for the polymerization of other monomers. 2.3. Results and Discussion 54 2.3.1. Synthesis of Substituted Styrenes The substituted styrenes used in this investigation are listed in Scheme 2.3. Most were available from commercial sources, however, 3,5-di-tert-butylstyrene, 3,5- dimethoxystyrene, 2,6-dimethoxystyrene, 4-vinylanisole and 4-(trifluoromethyl)styrene . . . 10-13,17 . were synthesrzed according to literature procedures. The synthetic protocols for 3,5-dimethoxystyrene, 4-vinylanisole and 4-(trifluoromethyl)styrene are outlined below (Scheme 2.4.). 55 / G) 6 Base, CH3PPh3Br \ ,z X \x x x = substituent / CH Li, CH P( Ph Br /O\©/O\ n—BuLi, TMEDA 3 3 )3 DMF,THF 1". 55 °C 0-10°C,1.5h 45% 72% tBuCl / A1013 CHB‘2 .40 °C to -100 °C NBS/benzoyl peroxide 2.5h CCI4, reflux,15h 86% 70% HMTA, water / methanol reflux, 4 h 68% / /0 CH31. PPh3, NaH THF, 50 °C, 6h 60% Scheme 2.4: The synthesis of substituted styrenes. (top) 3,5-Di-tert-butylstyrene, 3,5-dimethoxystyrene, 2,6-dimethoxystyrene, 4-vinylanisole and 4-(trifluoromethyl) styrene were synthesized via the Wittig reaction from commercially available aldehydes. (middle) Formylation of 1,3-dimethoxystyrene followed by Wittig olefination provided 2,6-dimethoxystyrene in moderate yield. Synthesis of 3,5-di-tert-butylstyrene (bottom) was prepared via a four step procedure. 2.3.2. Synthesis of Substituted Styrene Brushes via Surface Initiated Polymerization Scheme 2.5. outlines the synthesis of substituted styrene polymer brushes grown from gold substrates. The experimental procedure is similar to that described by Bao et al. for the rapid growth of tert-butyl acrylate from gold surfaces.6 A self assembled monolayer was prepared on the gold surface using MUD and converted into an initiator monolayer by reacting its terminal hydroxy group with 2-bromopropionyl bromide, as described previously. Initiator immobilization was apparent from the appearance of a carbonyl peak at 1743 cm'1 in the reflectance FTIR spectrum (Figure 2.2a). The purified substrates were transferred into a dry box, and the polymerization was performed inside the drybox to avoid contamination from oxygen. The catalyst system was a mixture of Cu(I)Br/Me4Cyclam and Cu(II)Br2/(anbpy)2 in 1:1 (v/v) solutions of DMF and anisole. The Cu(II) complex ensures deactivation of active radicals and provided some control over the rapid polymerization. Polymerizations of substituted styrenes were run at 55 0C for 8 hours under identical conditions. The substrates were sonicated, and then washed with ethyl acetate and THF. The IR reflectance spectra of the dried films confirmed formation of polymer brushes by the appearance of characteristic peaks that correspond to the substituents on the styrene ring (Figure 2.2). / \ l \, ’X 0 . o . H H ls—(CH2)-o—c-cH-Br an'sme’DME: I'S"(CH2);10 [CH2~CiBr 11 CH3 CuBr l Me4Cyclam \ n CuBr2(anbpy)2 x= substituent l ,\’ X Scheme 2.5.: Surface initiated polymerization of styrene derivatives 57 2.3.3. Kinetic study of surface initiated polymerization of substituted styrene from gold surface. Figures 2.1 and 2.3 show the evolution of film thickness with time for the polymerization of substituted styrenes. Figure 2.1 shows data for the surface polymerization of ortho and meta-substituted monomers while Figure 2.3 shows comparable data for para-substituted styrenes. The polymerization rates at 55 C>C were unusually high for the ATRP of 3,5-disubstituted styrenic monomers, providing 100-350 nm thick layers in 4 hours.18 In the early stages of polymerization, the nonlinear relationship between the film thickness and time suggests a relatively high concentration of radicals leading to some termination as well as a high polymerization rate. However, after >2 hours, growth in film thickness with time is roughly linear, indicating that some level of control was established during the polymerization. Polymerizations of ortho- substituted monomers were sluggish, likely due to steric effects, and none provided 100 nm thick films in 8 hours. 58 300 I 250 - o E 200 - ' 5 A 8 u g 150 ~ 0 .C I A '— o 100 ~ ' a o D Q a Cl 50 . g + + + + A 0 A 4 i 1 , 0 2 4 6 8 10 Time (hour) Figure 2.1: Evolution of the ellipsometric brush thickness with time for the polymerization of substituted styrenes from gold substrates. 3,5-bis- (trifluoromethyl)styrene (I), 3,5-dimethoxystyrene (O), 3,5-dimethylstyrene (A), 3,5-di— tert-butylstyrene (X), styrene (O), 2-(trifluoromethyl)styrene (D), 2-methoxystyrene (+), 2,6-dimethoxystyrene (A). 59 t x10 e x 100 x50 A C X100 1 b x100 a x 50 LE iii a? 9 j l 3000 2500 2000 1 500 1 000 Wavenumbers (cm") Figure 2.2.: Representative reflectance FTIR spectra of gold surfaces coated with (a) the immobilized initiator (b) 70 nm of polystyrene (c) 55 nm of poly(3,5-di-tert butylstyrene) (d) 5 nm of poly(2,6-dimethoxystyrene) (e) 27 nm of poly(2- methoxystyrene) and (f) 130 nm of poly(3,5-bis-(trifluoromethyl)styrene). 60 The IR spectra in Figure 2.2 confirm the growth of polymers from Au surfaces. Furthermore, the spectral intensities agree well with the kinetic data. Spectrum b shows the four peaks (overtones and combinations) from 2000-1700 ch characteristic of mono-substituted benzene rings (styrene). Strong methyl C-H stretching bands at 3000- 2850 cm.1 and the characteristic doublet for tert-butyl groups, a strong peak at 1370 cm.1 and a weaker signal at 1390 cm-1 confirms the growth of 3,5-di tert—butyl styrene in spectrum 0. The strong, broad, asymmetric (1280-1220 cm'l) and symmetric (1050-1000 cm'l) stretching modes for aryl ethers confirm the attachment of dimethoxy and monomethoxy styrene (spectra d and e). In spectrum f, a sharp strong peak at 1350 cm”1 (C-F stretching) identifies the CF3 group of poly(3,5-bis-(trifluoromethyl)styrene). In Figure 2.4, strong peaks between 1280-1220 cm"1 in spectrum b (C-O-C stretching of an alkyl aryl ether) confirm the growth of 4amethoxystyrene. A sharp strong peak at 1350 cm'1 (C-F stretching) identifies the CF3 group in poly(4-trifluoromethyl styrene) (spectrum c). The spectrum of poly(4-methylstyrene) (d) is similar to that of styrene (a), except for the 2 peaks between 2000-1700 cm'1 (overtones and combinations) for a 1,4-disubstitutedbenzene ring. The strong, narrow peak seen at ~1000 cm" in spectrum e is characteristic of aromatic C-Br stretching expected for poly(4- bromostyrene). In spectrum 0, the characteristic doublet for tert-butyl groups, a strong peak at 1370 cm" and a weaker signal at 1390 cm'1 confirms the growth of 4—tert-buty1 styrene. 61 250 I 200 ~ 75‘ 5 150 ~ I D In (D a) .5 '9- 100 P D o "3 o '- O I a Z A 50 - 0 8 A A 0 A T i T 0 2 4 6 8 10 Time (hour) Figure 2.3: Evolution of the ellipsometric brush thickness with time for the polymerization of 4-substituted styrenes from gold substrates. 4-(trifluoromethyl)styrene (I), 4-tert-butylstyrene (O), 4-bromostyrene (E1), styrene (O), 4-methylstyrene (A), 4- methoxystyrene (A). 62 0.2 l 1 x20 8 x20 5 M .0 I- 0 x50 (n .0 <1 x2 XW L l 3000 2500 2000 1500 1000 Wavenumbers (cm") Figure 2.4: Representative examples of reflectance FTIR spectra of 4- substituted styrenes grown from gold substrates (a) 70 nm polystyrene (b) 10 nm poly(4- methoxystyrene) (c) 200 nm poly(4-(trifluoromethyl)styrene) (d) 55 nm poly(4- methylstyrene) (e) 100 nm poly(4-bromostyrene) (t) 150 nm poly (4-tert-butylstyrene). 63 2.3.4. Substituent effects on polymerization rate. The data in Figures 2.1 and 2.3 show that monomers with electron withdrawing groups polymerize faster than monomers bearing electron donating groups, irrespective of the substitution pattern on the aromatic ring, (excluding ortho-substituted monomers where steric effects restrict propagation and termination. We also failed to grow a-methyl styrene from gold surface (data not shown) due to steric reasons.) Monomers with bulky electron donating groups (tert~butyl) at meta or para position polymerized faster than styrene and ortho-substituted monomers. Surprisingly, monomers with electron donating groups at meta sites polymerized faster than styrene or their ortho/para substituted analogues, irrespective of their steric demand. The observed film growth rates varied in the order of m-di—CFg, p-CF3> m-di methoxy> m-dimethyl, m-di-tert—butyl > p-tert- butyl > p-Br > H > p-methyl > o—methoxy > o-dimethoxy, p-methoxy. The results are consistent with solution polymerization rate data obtained from conventional radical polymerizations as well as controlled radical polymerizations such as ATRPI6 and living radical polymerizations initiated by TEMPO/EPOM’IS To analyze the surface initiated polymerization rate data, we constructed a . . . . . 16,19 Hammett plot Similar to the method used to analyze solution polymerizations. A Hammett plot requires rate constants for the surface polymerization of para-substituted styrene because, only para-substituted styrene Hammett constants (c) are known. The propagation and termination rate constants for surface polymerization can be estimated . . 20 by USlng a model proposed by the Wirth group. 64 The initiation and propagation steps of living radical polymerization have been well-documented and adopted for ATRP systems as shown in Scheme 2.6. Initiation: ka ~Rn-X + (Cu(|)Iigand)X # ~Rn' + (Cu(||)ligand)X2 (1) d Propagation: kp ~Rn' + M ~———-—> ~Rri+1 (2) Termination: kt ~ ~Rn° “Ra “‘"* in (3) "Rm Scheme 2.6: The initiation, propagation and termination steps of ATRP According to this mechanism, the disappearance of monomer only occurs during propagation (step 2 of Scheme 2.6). -9155]- ..- 1 dt .. kp[R][M] () If [R-] is constant, the monomer concentration can be reduced according to the first-order kinetics. m W2] = kp[R-]t (2) However, [R-] usually is not constant due to termination, especially in case of surface polymerization which is clearly visible in Figures 2.1 and 2.3. Assuming only termination by coupling for surface-initiated ATRP, and ignoring possible contributions from surface 65 fouling or side reactions such as radical transfer to solvent, the rate of termination can be 18 expressed as: d R is] ._. _ kt[R']2 (3) [13'] R = 0 <4) Recognizing that monomer-to-initiator ratio is so high that monomer concentration changes are negligible simplifies the solution of Equation 2. [M] (5) — = - k R' 1 [M10 1 pl 1 Substituting Equation 4 into Equation 5 reduces conversion to simple nonlinear time dependence: [Mic-[M] _ kle'lo‘ (6) [M10 1 +1R.10ktt Realizing that the ellipsometric thickness is proportional to [M]0 - [M], the data of Figure 2.3 were fit to Equation 6. The fits to the data, as shown in Figure 2.5 for 4- bromostyrene, is reasonable and allows extraction of the apparent rate constants for propagation (kp app = [M]0 hp [R-]0) and termination (kt app = [R°]0 kt). The data obtained for different para-substituted styrenes are tabulated in Table 2.2. 66 Table 2.2. Apparent rate coefficients in surface ATRP obtained from the Wirth . . 8 model and absolute propagation rate constants for substituted styrenes subsumem “21 kp app (moi-1161) kt app (5'1) kp abs (mol-1Ls-1)8 4-CN 0.66 — - 219 4-CF3 0.54 649 7.23 - 4-Br 0.23 111 1.35 186 4-Cl 0.23 - - 150 4-H 0.00 106 1.82 110 4-Me -0. 17 76 2.03 84 4_CMe3 -0.20 288 6.81 - 4-OMe -0.27 9.0 0.59 71 67 100 75 i- A E r 5 2 m R = 0.99513 8 50 C .. x y - 110.53t l (1 + 1.351) .2 fi 25 O J 1 r L 1 I r L O 2 4 6 8 Time (hours) Figure 2.5: A representative example of the polymerization data for 4-bromo styrene fit to Equation 6. The polymer was grown from Au substrates at 55 CC. Using the rate constants from fitting the data to Equation 6, we constructed a Hammett plot (Figure 2.6) relating the ratio of the apparent rate coefficients (kp app/kt app) (similar to analyzing product ratios, as applied by several authors to obtain Hammett 22,23,24 . . . and the Hammett constant, c, for different para substituents. By usmg the plots) ratio of the rate constants, we eliminate the radical concentration [R-]0, and in principle, kp app / kt app should reflect the ratio of absolute rate coefficients. Thus, 68 kp app / kt app = [Mlokp 1R'10/ 1R'10 kt: [M10 kp / kt and 108 (kx/ hr) = 108 [([M10 kp / kt)X / ([Mlo kp / kt)H1 =108 1(kp / kt)x / (kp / kt)H1 where X and H refer to the substituted styrene and styrene respectively. For conventional radical polymerization of para-substituted styrenes, there is a linear correlation between log(kp app / kt app) and the Hammett o constants for different substituents. The value of p for our surface ATRP data (p = 0.51) was similar to that of conventional radical systems run in solution (p = 0.55, Figure 2.6),25 but much smaller than the p value reported by Matyjaszewski for the solution ATRP (p = 1.5).16 The difference between the solution and surface ATRP results is the use of different ordinates. In solution ATRP, the correlation was established with regard to the apparent propagation rate constant which comprises [.R'], rather than the ratio of rate coefficients used for surface ATRP. Thus, the p values for surface ATRP and conventional (solution) radical polymerizations which also uses absolute rate coefficients as ordinate, are similar. Based on the qualitative success of the simple Hammett equation similar to the . . . . . 26 . solution polymerization V18. free radical pathway, we conclude that these reaction constants are roughly correlated with Hammett parameters for the para substituents in surface polymerization with p = 0.51. Such substituent effects are mainly due to an increase in kp / kt for monomers substituted with electron-withdrawing substituents in para positions (Table 2.2). However, it does not exclude the possibility of a larger Keq (in addition to larger kp) for electron withdrawing monomers due to the decrease in the bond 69 dissociation energy (BDE) of the C-X bond by electron-withdrawing substituents and hence causes the additional enhancement of apparent polymerization rate coefficient with . . . . . 16 . . 6 (Similar to the solution polymerization) . The energy graph in Scheme 2.7 provrdes a qualitative picture of the substituent effects on polymerization rates in surface ATRP of substituted styrenes. Of the monomers studied, only 4—methoxystyrene failed to grow a significant surface film. Similar observations were made by Matyzasjewski et al. 16 in their study of the solution ATRP of 4-methoxystyrene, where they failed to obtain high molecular weight polymer. They suggested that the active growing species has a cationic nature due to the presence of strong electron-donating substituent. For this reason, we excluded 4- methoxystyrene from the Hammett plot. 70 ..... 10g (kx abs/[CH abs) 10g 1(kp app/kt app)X/(kp app/kt app)H1 ____ 0.5 04 _ ----y-_-o.55x-o.003 // ' R2=0.91 / 0.3 ' -01 . —y=0.51x-0.04 1?.2 =0.88 -O.2 '/ -0.3 ' ' m ' -0.4 -0.1 0.2 0.5 0.8 0 Figure 2.6: Hammett plots of (kp/kt) in surface-initiated ATRP of substituted styrenes (O) and .for absolute kp values from conventional polymerizations (Cl) of . 8 substituted styrenes. 71 4-0.423}, l : —CH2-CH \ X(H) 1 \ I X X X(donor) “—7)“— X(acceptor) —.‘r—'"’ “ s Energy BDE BDE BDE ’ " H x L—" —CH2-C-Bl’ . ’ H x”, —CH2'C-Bl' I \ __iL_ ,\' 1 \ X(acceptor) —CH2-C-Br /\,’( (H) l \ \, X(donor) smaller Keq larger Keq Scheme 2.7: Energy diagram for different substituents16 2.3.5 Synthesis of substituted styrene polymer brushes from SiOz substrates This polymerization system was broadened to silica substrates as well. As expected, similar polymerization trends were observed (i.e. monomers with electron withdrawing substituents polymerize faster than those bearing electron donating substituents). One interesting observation was that polymerizations initiated from gold substrates were always slower than those from silica substrates, irrespective of monomer . 6,27 . studied (Figure 2.7). This observation, prevrously reported by Bao et al. Will be explained in detail in Chapter 3. 72 500 I 400 ~ A I E 5 300 - til o I E g 200 E O i- I O O 100 - 0 A o 8 A + g t O l l J l 0 2 4 6 8 10 Time (hour) Figure 2.7: Evolution of the ellipsometric brush thickness with time for the polymerization of 3,5-bis(trifluoromethyl)styrene (I), 4-tert-butylstyrene (O), styrene (O), 4-methylstyrene (A),and 2-methoxystyrene (+) from silica surfaces. 73 2.4. Conclusions Surface initiated atom transfer radical polymerization of substituted styrenes provides uniform and thick brushes (>100 nm in 1 hour) from gold and silicon surfaces. Styrenes substituted with electron withdrawing groups in meta and para positions exhibited high growth rates. The substituent effects seen in surface and solution polymerizations are similar for styrenes with electron withdrawing groups, and for electron donors in ortho and para positions, but donors at meta sites have surprisingly fast growth rates, possibly due to steric inhibition of termination. We also showed that the surface polymerization rate varies with different substituents follow the Hammett relation with p = 0.51. The ratio of kp to kt, acts as an indicator of the likelihood that a reaction will reach high degrees of polymerization before termination. 74 2.5. References 3 Chiefari, J.; Chong, Y. K.; Ercole, F.; Krstina, J.; Jeffery, J.; Le, T. P. T.; Mayadunne, R. T. A.; Meijs, G. F.; Moad, C. L.; Moad, G.; Rizzardo, E.; Thang, S. H. Macromolecules 1998, 31, 5559. 4 Couvreur, L.; Lefay, C.; Belleney, J.; Charleux, B.; Guerret, 0.; Magnet, S. Macromolecules 2003, 36, 8260. 5 Grande, D.; Guerrero, R.; Gnanou, Y. J. Polym. Sci. Polym. Chem. 2005, 43, 519. 6 Bao, Z.; Bruening, M. L; Baker, G. L. J Am. Chem. Soc. 2006, 128, 9056. 7 Odian, G. Principles of Polymerization. 3rd ed.; John Wiley & Sons: New York, 1991. 8 Brandup, J .; Irnmmgut, E. H.; Grulke, E. A. Polymer Handbook. 4th ed.; John Wiley & Sons: New York, 1999. 10 Pillow, J. N. G.; Halim, M; Lupton, J. M.; Burn, P. L.; Risch, N.; Meyer-Roscher, B.; Langhals, M. Macromolecules 1999, 32, 5985. ll Risch, N.; Meyer-Roscher, B.; Langhals, M. Z. Naturforsch., B: Chem. Sci. 1994, 49, 141. 12 Overberger, C. G.; Burns, C. M. J. Polym. Sc, Polym. Chem. 1969, 7, 333. 13 Kawamura, T.; Uryu, T; Matsuzaki, K. Makromol. Chem. 1982, 183, 125. 14 Imoto, M.; Kinoshita, M.; Nishigaki, M. Makromol. Chem. 1965, 86, 217. 15 Kazmaier, P. M.; Daimon, K.; Georges, M. K.; Hamer, G. K. Am. Chem. Soc, Polym. Prepr. 1996, 3 7, 485. 16 Qiu, Jian ; Matyjaszewski, K. Macromolecules 1997, 30, 5643. 17 Okuma, K.; Sakai, 0.; Shioji, K. Bull. Chem. Soc. Jpn. 2003, 76, 1675. 18 Samadi, A.; Husson, S. M.; Liu, Y.; Luzinov, I.; Kilbey, M. Macromol. Rap. Comm. 2005, 26, 1829. 75 19 Coote, M. L.; Davis, T. P. Macromolecules 1999, 32, 4290. 20 Xiao, D. Q.; Wirth, M. J. Macromolecules 2002, 35, 2919. 21 Hansch, C.; Leo, A.; Taft, R. W. Chem. Rev. 1991, 91, 165. 22 Watkins, A. L.; Hashiguchi, B. G.; Landis, C. R. Org. Lett. 2008, 10, 4553. 23 Yajima, Tomoko; Okada, Kyoko; Nagano, H. Tetrahedron, 2004, 60, 5683. 24 Galardon, E.; Le Maux, P.; Simonneaux, G. Tetrahedron, 2000, 56, 615. 25 Walling, C.; Briggs, E. R.; Wolfstim, K. 8.; Mayo, F. R. J. Am. Chem. Soc. 1948, 70, 1537. 26 Coote, M. L.; Davis, T. P. Macromolecules 1999, 32, 4290. 27 Bao, Z.; Baker, G. L.; Bruening, M. Macromolecules 2006, 128, 9056. 76 Chapter 3 Surface effects in surface-initiated ATRP 3.1. Introduction Surface initiated polymerization is a powerful technique for preparing chemically l,2,3,4,5,6 modified surfaces. Studies aimed at understanding the fundamentals of polymer growth often use self-assembled monolayers (SAMs) as the initiator layer, because SAMs provide a closely packed, well-ordered and stable configuration on the surface. Most SAMs are generated on metal surfaces such as Au, Ag and Cu using thiol chemistry while alkyl siloxanes anchored to hydroxyl-terminated surfaces, such as $0; A1203 and other oxides via alkoxy and chlorosilanes. Surface initiated radical polymerization can be carried out on a broad range of substrates, but most studies have focused on silicon and gold substrates. Gold surfaces are chemically homogeneous, virtually free of contamination, easy to clean, applicable to a wide variety of analytical techniques for thin film characterization. A disadvantage of Au surfaces is that the Au-S bond that links the 9 9 . . . . . 78 . . initiator layer to the surface 18 labile above 60 OC and hence polymerizations are typically limited to <60 OC. However, at such temperatures, the low propagation rates for radical polymerization of important aromatic monomers such as styrene and 4-vinyl pyridine limit their growth. In addition, Bao et al. observed that film growth rates from $102 substrates were higher than those grown from Au substrates, regardless of 10 monomer even at temperatures < 60 0C. 77 9a . . . . . . . . Huang et al. investigated free-radical polymerization of styrene from azo-1n1t1ators immobilized on flat Au substrates, and showed that free-radical polymerization from Au is limited by the instability of alkanethiol monolayers. They reported that soluble free radicals accelerated thiol desorption from Au, and the desorbed alkanethiols appeared to be efficient chain—transfer reagents that terminate brush growth. To overcome monolayer instability, they utilized a simple cross-linking procedure to enhance the stability of SAMs, and made thermal radical polymerization from Au surfaces facile. Later Stephanie et al.9b restricted thiol desorption at elevated temperatures by coating gold electrodes with a layer of carboxylic acid functionalized polypyrrole, and in their work on one pot thermal polymerization from gold nanoparticles, Matyjaszewski et al.“C used a crosslinked polymer shell to prevent dissociation of linear brushes from gold nanoparticles at elevated temperatures. There are other factors that may lead to different polymerization rates on Au and silica. One difference between Au and SiOz surfaces is the number and density of sites available on the surface for immobilizing initiator. Studies show that silica surfaces have a limiting area/chain of 0.6 nmz,“ roughly 40% of the density for a SAM on Au, while . . 2 . . typical values for silica range from 0.65 nm2 to 1.54 nmzlcharn. Bao et all investigated the effect of initiator density on polymerization rates for Au and silica surfaces, and observed that decreasing the initiator density on silica by 50% decreased the polymerization rate, but for gold, the rate was nearly unchanged. Additionally, gold is a 78 3,14,15 transition metal capable of quenching radicals,1 which could reduce the number of active chains, resulting in thin films. This chapter describes a systematic evaluation of three factors that may be responsible for the lower film grth rates from gold: 1) differences in the initiator densities for gold and silica substrates, 2) termination resulting from desorption of thiols from SAMs on gold surfaces, which is absent in silicon surfaces 3) and radical quenching ’ 1 ,1 . by the gold surface. 5 Based on the work of Bao et al., we suspected that desorption of thiols from SAMs on gold surfaces at ambient temperatures, followed by chain transfer of growing polymer chains to thiols was the likely cause of thinner film growth on Au compared to silica surfaces. We tested this hypothesis by examining polymerizations from a crosslinked initiator anchored on gold substrates, and by analysis of block copolymer growth from Au and silica surfaces. 3.2. Experimental Section 3.2.]. Materials Unless otherwise noted, all chemicals were obtained from Aldrich and stored under nitrogen. ll-Mercapto-l-undecanol (MUD, 97%), 2-bromo—2-methylpropionyl bromide (2-BIB, 97%), 3-mercaptoprOpyltrimethoxysilane (MPS), anisole (99.7%), N,N- dimethylformamide (DMF, 99.8%), Cu(I)Br (99.999%), Cu(II)Br2 (99.999%), Me4Cyclam (99%), and 4,4’-dinonyl-2,2’-bipyridine (anbe. 97%) were used as received. (3-Aminopropyl)trimethoxysilane was distilled under vacuum prior to each use. Triethylamine was distilled from calcium hydride under a nitrogen atmosphere, and stored under nitrogen. tert-Butyl acrylate (tBA) (98%), MMA (99%) and styrene (99%) 79 were passed through a 10 cm column of activated basic alumina and then distilled from calcium hydride to remove inhibitors. The disulfide initiator (ll-[(2-bromo-2- methyl)propionyloxy]undecyldisulfide)7 [(Br-C(CH3)2-COO-(CH2)1[S)2], the trichloro- silane initiator (11-[(2-bromo-2-methyl)propionyloxylundecyl trichlorosilane” and the trimethoxysilane initiator [2—bromo—2-methyl-N—(3-trimethoxysilylpropyl)propionamide] were synthesized using slightly modified versions of literature proceduresl6 Silane compounds were used and stored under nitrogen. Toluene was distilled under nitrogen in the presence of sodium/potassium alloy using benzophenone as an indicator. After purification, the monomers, liquid chemicals and solvents were transferred to Schlenk flasks, degassed by three freeze-pump-thaw cycles, and then transferred into a drybox. Silicon wafers and Au-coated wafers (electron-beam evaporation of 200 nm of Au on 20 nm of Ti on Si (100) wafers, or 200 nm of Au sputter-coated on 20 nm of Cr on Si (100) wafers) were cleaned in a UV/O3 chamber for 30 min prior use. 3.2.2. Characterization Methods Ellipsometric measurements were obtained with rotating analyzer ellipsometer (model M-44, J. A. Woollam) using WVASE32 software. The angle of incidence was 750 for all experiments. For the calculation of film thickness, a refractive index of 1.50 was used. Reflectance Fourier Transform Infrared (reflectance FTIR) spectroscopy was performed using a Nicolet Magna—560 FTIR spectrometer containing a PIKE grazing angle (80°) attachment. Spectra were typically collected with 128 scans using a MCT detector. Changes in surface morphology of the polymer film during thermal degradation 80 - 1111111 on gold surface were observed using a Keyence Digital Microscope equipped with a video camera. 3.2.3. Preparation of immobilized initiators on gold and silicon substrates Gold—coated wafers were UV/O3-cleaned for 30 min, washed with water and ethanol, and after transferred into a N2 glovebag, immersed in a 1 mM ethanolic solution of the disulfide initiator [(Br-C(CH3)2-COO(CH2)11S)2] for 24 h to form the self- assembled initiator monolayer. The films were then washed with ethanol and dried under a stream of N2. A crosslinked initiator monolayer was formed by immersing Au-coated substrates in a vial containing a 2 mM methanolic solution of MPS for 12 h in a glovebag at room temperature. After deposition, the substrate was rinsed three times with 2 mL of methanol and dried with nitrogen. The attached silane monolayer was then hydrolyzed at room temperature with 0.1 M HCl for 15 h to afford a hydroxylated surface. The modified Au substrate was then treated with a 10 mM solution of trimethoxysilane initiator [2—bromo-2—methyl-N—(3-trimethoxysilylpropyl)propionamide] in toluene at 55 0C for 12 h under nitrogen. Following the deposition, the substrates were rinsed repeatedly with toluene and iSOpropanol, and then dried in a stream of nitrogen. The ellipsometric thickness of the crosslinked initiator layer on gold was 10-15 A. The trimethoxysilane initiator was immobilized on UV/O3-cleaned silicon wafers by immersing the substrate in a 10 mM solution of [2-bromo-2-‘methyl-N-(3- trimethoxysilylpropyl)propionamide] in toluene at 55 0C under nitrogen. After 12 h, the wafer was rinsed repeatedly with toluene and isopropanol, sonicated in toluene for 1 min 81 and dried in a stream of nitrogen. The ellipsometric thickness of the crosslinked initiator layer on silicon surface was 10-15 A. 3.2.4. Surface initiated polymerization In a NZ-filled dry box, CuBr (5.7 mg, 0.04 mmol), CuBr; (4.5 mg, 0.02 mmol), Me4Cyclam (10.3 mg, 0.04 mmol) and anbpy (16.4 mg, 0.04 mmol) were added to a round-bottom flask containing 20 mL of a degassed solution of monomer in DMF/anisole (monomer/DMF/anisole ) 2:1 :1 v:v:v, [monomer]~4.0 M). The mixture was well-stirred and heated with an oil bath to 50 0C until a transparent, light green solution formed. The prepared solution was then transferred into a small vial containing an initiator-modified Au / SiOz substrate to start the surface-initiated polymerization. After a predetermined reaction time at 50 CC, the substrate was removed from the vial, washed with ethyl acetate and THF sequentially, and then dried under a stream of N2 in the dry box. 3.3. Results and discussion 3.3.1. Surface initiated polymerization from gold and S102 substrates Bao et al. had reported that “identical” polymerizations of MMA from Au and Si02 surfaces gave different polymer film thicknesses, with polymer films grown from - . . 12 S102 systematically thicker than those grown from Au substrates. To confirm Bao’s observation, we polymerized MMA and styrene under identical conditions from initiators anchored on Au and SiOz surfaces (Scheme 3.1), and compared the evolution of the film thickness with time. Figures 3.1a and 3.1b show the growth of PMMA and polystyrene 82 brushes from flat Au and S102 surfaces. Irrespective of the monomer, Si02 surfaces yield thicker films and higher apparent polymerization rates than those grown from Au surfaces under similar conditions. However, the film thickness vs. time profiles are similar, suggesting that the primary difference between growth from gold and growth from silica is fewer active chains when brushes are grown from gold surfaces. These data confirm the observations of Huang et al. and Bao et al., that polymer films grown on Au substrates are systematically thinner than those polymerize on S102. Au Si b—OH P‘-S'-(C1"'12)""O Bf 11 MMA anisole / DMF CuBr / Me4Cyclam CuBr2(anbpy)2 —-———————-> O 55 °C O MeQ ”0H MeO-lSi-(Csz—N‘ Br MeO 3 H Toluene Overnight , 55 °C 31 o o L"O"Si“((:1‘12)"‘i‘{ i C I 3 H O \ / O 0. 9H3 ”‘O'Si‘WCHZl’N. CH2-C 1—Br \ 3 H n 0 CH3 Au -—s-—- CH -—0 -' ( 2) CH2 C Br 11 A " O OMe Si -O-’Si~(CH2)-N\ Br 0 3 H \ MMA, anisole / DMF CuBrl Me4Cyclam CuBr2(anbpy)2 55 °C A 0 OMe CH3 HzréT’Br A n 0 OMe Scheme 3.1: Surface initiated polymerization of MMA from silicon and gold surfaces 83 300 250 ~ 1 i g 2..- i (D 8 1 5 150 ~ .2 i ‘5 r 100 — f ' i 50 ~ D O F I L l 1 0 20 40 60 80 100 Time (min) Figure 3.13: Evolution of the ellipsometric brush thickness with time for the polymerization of MMA (methyl methacrylate) from 0 gold and I silicon surfaces at 50 OC. The points are the average of data from two independent runs, and the limits of the error bars are the measured film thicknesses from the two runs. 84 200 150 - 75‘ i 5 i 8 1:100 - i x .2 i f if 5 i I 50 - § 5 o I l 1 l 0 2 4 6 8 10 Time (hour) Figure 3.1b: Evolution of the ellipsometric brush thickness with time for the polymerization of styrene from 0 gold and I silicon surfaces at 50 oC. The points are the average of data from two independent runs, and the limits of the error bars are the measured film thicknesses from the two runs. 3.3.2. Block copolymer formation on Au and SiOz substrates It would be reasonable to expect that the different behaviors seen for films grown from Au and SiOz substrates would diminish when the growing polymer chains are distant from the substrate. This hypothesis can be tested by placing the initiator at increasing distances from the Au surface. One way to form an extended initiator is to use ATRP to grow polymers from the surface, and then quench the growing chain by adding 85 a large excess of Cu(II)Br2 as shown in Scheme 3.2. The resulting Br-terminated chains monomer, it was dried under a stream of N2, removed from the dry box, and characterized by ellipsometry and F TIR spectroscopy. We then returned the substrates to the drybox and initiated polymerization of MMA for 1 hour from the dormant PtBA chains. The same process was carried out on a Si02 substrate, which served as a control. Net growth of PMMA in 1 hour Br Br Br Br 3" BrBr tBA/98131361 MMA leatalyst .. .. SO'Ut'O” lg solution , 1 hour Initiator attached on gold PtBA brush on gold Block copolymer on surface or Silicon surface or Silicon surface (PtBA = Poly(tertbutyl acrylate)) Scheme 3.2: Block copolymer formation on Au and Si02 surfaces Figure 3.2 shows F TIR spectra of a 97 nm PtBA film and the PtBA-block—PMMA copolymer after growing 210 nm of PMMA from the initial PtBA film. The spectrum for the block copolymer is as expected; the intensity of the ester carbonyl increased, and the relative intensity of the t-butyl doublet decreased compared to the C-H rocking bands, (135045006114). 86 0.05 Absorbance ’3‘ 3500 3000 2500 2000 1500 1000 Wavenumbers (cm'1) Figure 3.2: Reflectance FTIR spectra of gold substrates coated with (a) 100 nm PtBA brushes grown from the initiator layer; (b) a PtBA-block-PMMA copolymer brush, synthesized by growing 210 nm of PMMA from a 100 nm PtBA film. Figures 3.3a and 3.3b show that PMMA blocks grown for 1 hr from PtBA brushes were ~200 nm thick, 2-3 times thicker than PMMA films grown directly from gold surfaces, but comparable to PMMA blocks grown directly from SiOz. These results suggest that a substantial fraction of the PtBA chains were active after polymerization of the initial PtBA block, and directly confirm the role of the Au surface in reducing the 87 film thickness when polymers are grown from Au surfaces. For Au substrates, the thickness of the PMMA layer grown from PtBA films in 1 hour increased with the PtBA thickness, saturated at ~ 200 nm, and then slowly declined. The decline likely reflects fewer dormant PtBA chains available to form the PtBA-block-PMMA copolymer. Since the same effect is not seen for SiOz substrates, the decline can be identified with higher termination rates for the growth of PtBA on Au, presumably by radical quenching by electron transfer, or by the loss of growing chains via desorption of thiols from Au surfaces. Desorbed thiol may act as a chain transfer agent - for polymer brush syntheses, chain transfer is equivalent to termination. An alternative desorption scenario is that polymers anchored through Au-S bonds may be mechanically unstable, especially for highly solvated polymers of high molecular weight. 88 700 600 500 .p. O 0 Thickness (nm) (to O O 200 100 A A A A .-9-._. A?" ""0. e, C1 C1 ."'9 1:1 . o E] 0 Lo . O 20 40 60 Time(min) 80 Figure 3.3a: Surface polymerization study of the formation of PtBA-block-PMMA films on gold surfaces. 1:] PtBA brushes grown from initiators anchored on gold for various times; A Total film thickness after 1 hour growth of PMMA from PtBA brushes grown on gold; 0 Net growth of PMMA from PtBA (total film thickness — PtBA brush thickness) after 1 hour of polymerization (The broken line drawn as a guide to the eye) 0 PMMA grown from initiators anchored on gold for various times, shown for comparison. 89 700 ‘ A 600 - ‘ ’E‘ 500 - 5 E] El fl) 3 400 ~ ‘ E] C {a T: 300 ~ I- .-""o'-._ 2001:” “-0. ....... 2 C] O 100 - O O E 1 I l o 20 . 40 60 80 Time(min) Figure 3.3b: Surface polymerization study of the formation of PtBA-block-PMMA films on SiOz surfaces. D PtBA brushes grown from initiators anchored on SiOz surfaces for various times; A Total film thickness after 1 hour growth of PMMA from PtBA brushes grown on $0; 0 Net growth of PMMA from PtBA (total film thickness — PtBA brush thickness) after 1 hour of polymerization (The broken line drawn as a guide to the eye) 9 a PMMA brush grown from initiators anchored on Si02 for various times, shown for comparison. 90 3.3.3. Formation and polymerization from a cross linked initiator The previous data do not distinguish between radical quenching and thiol desorption as the primary factor leading to thinner films growing on Au surfaces compared to 8102. However, preventing desorption of thiols from the Au surface should test the validity of the thiol desorption mechanism. To that end, we formed cross—linked initiators on gold surfaces, similar to those of Holzinger et al., 17 and compared polymerizations from cross-linked initiator layers to substrates with initiators based on standard thiols. Scheme 3.3 shows the formation of a cross—linked initiator layer on a gold surface. The thicknesses of the cross-linked and standard thiol initiator layers were comparable, and therefore, radical quenching by the gold surface should be comparable for both initiator layers. Immersion of Au-coated slides in a 2 mM solution of MPS in methanol formed an MPS monolayer. During hydrolysis, the trimethoxysilane group at the MPS terminus condensed to form a dense poly(siloxane) network. The siloxane layer provides lateral stabilization through inter-chain cross-linking and generated a hydroxylated surface for the subsequent attachment of the trimethoxysilane-ATRP initiator. 91 Au \ i ' OCH Q /\/\OFH3 SMSi-08H3SM31'OH OCH3 AU OCH33 _____3___.. Au E) OCH overni ht, rt Methanol flSMSi- OC3H3 g SMSl'OH overnight,rt J OCH3 /O V O MeQ M MeO- $i- (CH2)N Br MeO 3H \ / O Toluene O O M . o ~S Si-O-Si— CH -N Br overnIght. 55 C \/\/ ( 2) . \ 3 H Au 0 o o “SWSj-O-Si-(CH2)-N Br 0 3 H / \ MMA. anisole / DMF 55 Cl CuBr/Me4Cyclam CuBr2(anbpy)2 \ / O O Q —S\/\/S\i0 Si (CHz)3N.H CH2- (ST—Br P'S\/\.Si O- Si- (CH) 23“ CH C Br ‘0 o 2 0:3:— ’ \ Aoue Scheme 3.3: Formation of cross linked initiators on gold surface We used FTIR spectroscopy and ellipsometry to follow the attachment of the ATRP initiator to MPS. The reflectance FTIR spectrum of the MPS layer on Au (not shown here) shows vibrational bands characteristic of MPS (2938 cm.1 for overlapping CH3 and CH2 bands, 2846 cm-1 for the CH2 symmetric stretch, and 1114 cm.1 for the Si-O-C stretch). After hydrolysis, the methyl peaks disappeared and the peak at 1114 -1 . . cm greatly decreased, indicating nearly complete hydrolysrs of the MPS layer. The ellipsometric thickness of the hydrolyzed MPS was 1.0 nm. The above results are in good agreement with literature data. 2 c . . “’18’19’ 0 After attachment of the ATRP Initiator, the film 92 thickness increased to 2.2 nm and amide peaks (1652 and 1548 cm-1) appeared in the reflectance F TIR spectrum confirming successful attachment of the ATRP-initiator. Figures 3.4 and 3.5 show data for the surface polymerization of MMA and styrene from gold and silicon surfaces with the crosslinked initiator. For both MMA and styrene, films grown from the crosslinked initiator on gold had thicknesses comparable to those grown on SiOz surfaces. Additionally, the data in Figure 3.5 show increases in the characteristic IR peaks of the respective polymers consistent with the ellipsometry data. Therefore, stabilizing the initiator layer via cross-linking efficiently restricts thiol desorption, and there is no evidence for radical quenching by gold. To explore the stability of the cross-linked thiol initiator, we ran polymerizations from room temperature to 115 0C using standard initiators immobilized on Au and Si02 surfaces. The data (Figure 3.6) show that thinner films are formed on Au substrates, which suggests significant thiol desorption even at low temperatures. In contrast, the polymerization data for the cross-linked initiator track the data for films grown on SiOz. Furthermore, the crosslinked initiator enables polymerization of MMA and styrene from Au surfaces up to ~100 oC, providing 300 nm and 120 nm films grown in 1hr, respectively. The apparent temperature limit for growth from Au surfaces is ~110 OC. Inspection of a polystyrene film grown at 115 OC showed large-scale delamination of the polymer films (Figure 3.7), which is absent in films grown at 90 0C. Kim et al. made similar observations for poly(hydroxyethyl methacrylate) brushes grown on Au.21 93 350 300 - A250 a f % E 5 3 200 r 3 i '5 150 - E I " I 100 - § § 50 » I! ¢ 0 I L _L l O 20 40 60 80 100 Time (min) Figure 3.43: Surface initiated polymerization of MMA from gold and Si02 surfaces. 0 PMMA grown from a standard (non-crosslinked) initiator on gold; 9 PMMA grown from a crosslinked thio initiator; Cl PMMA grown from Si02_. The points are the average of data from three independent runs, and the limits of the error bars are the measured film thicknesses from the three runs. 94 200 150 - E 5 m ¥ 8100 ~ .5 i 2 i r l- ? § 50 ~ Q 2 0 1 J l l o 2 4 6 8 10 Time (hour) Figure 3.4b: Surface initiated polymerization of styrene from gold and SiOz surfaces. 0 Polystyrene grown from a non-crosslinked thio-initiator on gold 0 Polystyrene grown from a crosslinked thio initiator D Polystyrene grown from Si02_The points are the average of data from three independent runs, and the limits of the error bars are the measured film thicknesses from the three runs. 95 0.05 (a) 23 8 3 (b) .J’k J 3500 3000 2500 2000 1 500 1000 Wavenumbers (cm") Figure 3.53: Reflectance FTIR spectra of gold substrates with (a) PMMA brushes grown from a non-crosslinked thio initiator in 1 hour (b) PMMA brushes grown from a crosslinked thio-initiator in 1 hour 96 0.05 Absorbance l l l I 3500 3000 2500 2000 1500 1000 Wavenumbers (cm") Figure 3.5b: Reflectance FTIR spectra of gold substrates with (c) PS grown from a crosslinked thioinitiator in 8 hours ((1) PS grown from a crosslinked thioinitiator in 8 hours 97 400 350 - II I Ezso- ¥ 3 «woo-é i C X 0 3150- l- 100 - § i i r i r r 50~ O l l 30 55 80 105 Temperature (°C) Figure 3.63: Temperature-dependent surface-initiated polymerization of MMA from various initiators; <> PIVHVIA grown from standard thioinitiators anchored to An surfaces; 0 PMMA grown from the crosslinked thioinitiator shown in Scheme 3.3; and Cl PMMA grown from SiOz as shown in Scheme 3.3. Each data point indicates the film thickness after 1 hour of growth. The points are the average of data from two independent runs, and the limits of the error bars are the measured film thicknesses from the two runs. All polymerizations were run for 1 hour. 98 160 140 2 , I —L O O r FOB-i PB" Thickness (nm) CD 0 60 ' 6 § 9 Q 5 § 40 - 8 9 o 20 - ° 0 l l l 30 55 80 105 130 Temperature (°C) Figure 3.6b: Temperature-dependent surface-initiated polymerization of styrene from various initiators; O polystyrene grown from standard thioinitiators anchored to Au surfaces; 9 polystyrene grown from the crosslinked thioinitiator shown in Scheme 3.3; and Cl polystyrene grown from SiOz as shown in Scheme 3.3. Each data point indicates the film thickness after 1 hour of growth. The points are the average of data from two independent runs, and the limits of the error bars are the measured film thicknesses from the two runs. All polymerizations were run for 1 hour. 99 (a) (b) Delaminated polymer 1mm Figure 3.7: Optical micrographs of polystyrene brushes grown from crosslinked initiators immobilized on gold surfaces: (a) and (b) show a 70 nm thick film grown for 1 hour at 115 oC. The use of cross-linked initiators to improve polymer film grth fi'om Au surfaces can be generalized to other monomers such as poly(vinyl pyridine) (PVP). Despite its potential utility as polyelectrolyte brush, we are unaware of examples of thick 12b . PVP brushes grown from Au surfaces (20 nm in 2 hours by Bao et al. and 6 nm 1n 5 hours. by Husson et al.22). However, Riihe 23 grew 430 nm thick PVP films from surface-anchored azo initiators on SiOz in 14 hours. Growing PVP from cross-linked initiators should provide films on Au with thicknesses that approach those of Riihe. Polymerization of vinyl pyridine for 8 hours at 50 oC provided 200 nm films from both Au and Si02 surfaces (Figure 3.8), approximately 10X thicker than previous examples. The increase in IR intensities, shown in Figure 3.9, is consistent with the ellipsometric data. 100 250 O 200 - C] 5 U E 5150 ~ (0 (D o I: x .2 100 — i5 6 50 - <> 0 ° 0 O l l l l O 2 4 6 8 10 Time (hours) Figure 3.8: Surface-initiated polymerization of vinyl pyridine from various initiators at 50 C)C; O poly(vinyl pyridine) grown from a non-crosslinked thio-initiator on gold 9 poly(vinyl pyridine) grown from a crosslinked thio initiator on gold Cl poly(vinyl pyridine) grown from a SiOz surface 101 0.05 2 2 (a) 2 ML (b) l P I l 3500 3000 2500 2000 1 500 1000 Wavenumbers (cm") Figure 3.9: Reflectance FTIR spectra of poly(vinyl pyridine) brushes grown for 8 hours from gold surfaces (a) using a crosslinked thioinitiator and (b) a standard thioinitiator. 102 growing chain terminated chain / non- growing 5 / \) chain S E’) ' S .__.._____+ S I -1” Thermal desorption l Radical induced | termination Radical induced desorption Effect of free thiol in solution - an analogy Mo _.—_._.__.___.> Chain Transfer Polymer growth [:ASC] :35: Figure 3.10. Proposed pathways for terminating surface-bound radicals on Au: (top) thermal or radical-induced desorption of surface thiol radicals terminate growing chains, (bottom left) radical termination by reaction with a thiol (analogous to chain- transfer), which competes with propagation (bottom right). Our experiments using cross-linked initiators indicate thiol desorption as the main limitation to growing thick polymer brushes on Au. However. the desorption mechanisms are unknown. Figure 3.10 depicts some pathways that lead to termination of surface- bound radicals on Au surfaces. The thermal instability of thiol SAMs on Au may not be important for low temperature polymerizations (< 60 0C). but likely more important at high temperatures. Radical—induced desorption of thiol SAMs from growing chains can 103 be important over a broad temperature range. The copper catalyst also may contribute to thiol desorption. We have observed chain desorption (reduction of the carbonyl peak height in the FTIR spectra of PMMA brushes) at high copper concentrations. Though the copper concentration used here is small, the effect may not be negligible since partial desorption of initiator-containing monolayers from the Au surface would result in a decreased surface initiator concentration. 3.4. Conclusions Polymerizations initiated from Au surfaces generally provide thinner polymer brushes than comparable polymerizations initiated from SiOz surfaces from room temperature to 100 0C. This disparity can be eliminated by forming a crosslinked poly(siloxane) primer layer on gold surfaces, which prevents thiol desorption. Using the crosslinked initiator, polymerizations fi'om Au provide film thicknesses comparable to polymerizations initiated from SiOz. These results implicate thiols in terminating growing polymer brushes, and polymer chains with low molecular weights and low film thickness. These results also exclude radical quenching by gold or the difference in initiator density between gold and silicon as significant factors. Crosslinked initiators extend the temperature range for polymerizations initiated from Au surfaces to 100 C)C, enabling rapid polymerization of monomer less active than methacrylates, such as styrene and vinyl pyridine. Above 100 0C, we observed macroscopic delamination of polymer films grown on Au surfaces. 104 References: l Milner, S. T. Science 1991, 251, 905. 2 Huang, X.; Doneski, L. J .; Wirth, M. J. Anal. Chem. 1998, 70, 4023. 3 Bruening, M. L.; Zhou, Y.; Aguilar, G.; Agee, R.; Bergbreiter, D. E.; Crooks, R. M. Langmuir 1997, 13, 770. 4 Sidorenko, A.; Minko, S.; Schenk-Meuser, K.; Duschner, H.; Stamm, M. Langmuir 1999, 15, 8349. 5 Prucker, 0.; Naumann, C. A.; Riihe, J.; Knoll, W.; Frank, C. W. J. Am. Chem. Soc. 1999, 121, 8766. 6 Edmondson, S.; Osborne, V. L.; Huck, W. T. S. Chem. Soc. Rev. 2004, 33, 14. 7 Ulman, A. Chem. Rev. 1996, 96, 1533. 8 Schlenoff, J. 8.; Li, M.; Ly, H. J. Am. Chem. Soc. 1995, 117, 12528. 9 (a) Huang, W.; G, Skanth; Baker, G. L.; Bruening, M. Langmuir 2001, 17, 1731. (b) Fan, X.; Xia, C.; Fulghum, T.; Park, M.-K.; Locklin, J.; Advincula, R. C. Langmuir 2003, 19, 916. (c) Dong, H.; Zhu, M.; Yoon, J. A.; Gao, H.; Jin, R.; Matyjaszewski, K. J. Am. Chem. Soc. 2008, 130, 12852. (d) Shan, J .; Nuopponen, M.; Jiang, Hua; K., Esko; T., Heikki Macromolecules 2003, 36, 4526. 10 Bao, Z. “Surface-Initiated Polymerizations on Initiator Anchored Substrates: Synthesis and Characterization of Nanometer Thick Functional Polymer Films.” Ph.D., Michigan State University, East Lansing, MI, 2006. 11(a) Fan, X.; Xia, C.; Fulghum, T.; Park, M.-K.; Locklin, J .; Advincula, R. C. Langmuir 2003, I9, 916. (b) Kawai, T.; Saito, K.; Lee, W. J. Chromatogr. B 2003, 790, 131. 12 (a) Bao, Z.; Bruening, M. L.; Baker, G. L. Macromolecules 2006, 39, 5251. (b) Bao, Z.; Bruening, M. L; Baker, G. L. J. Am. Chem. Soc. 2006, 128, 9056. 13 Henglein, A.; Lilie, J. J. Am. Chem. Soc. 1981, 103, 1059. 14 Sakata, Y.; Imahori, H. J. Phys. Chem. B 2000, 104, 1253. 15 Ganesh, V.; Lakshminarayanan, V. J. Phys. Chem. B 2005, 109, 16372. 16 (a)Yoon, K. R.; Chi, Y. S.; Lee, K.; Lee, Jungkyu K.; Kim, D. J.; Koh, Y.; Joo, S.; Yun, W. s.; Choi, r. s. J. Mater. Chem. 2003, 13, 2910. (b) Mulvihill, M. J.; Rupert, B, 105 L.; He, R.; Hochbaum, A.; Arnold, J .; Yang, P. J. Am. Chem. Soc. 2005, 127, 16040. 17 Holzinger, D.; Liz-Marzan, L. M.; Kickelbick, G. J. Nanosc. Nanotech. 2006, 6, 445. 18 Yan, Y.; Bein, T. J. Phys. Chem. 1992, 96, 9387. 19 Mihailova, B.; Engstrom, V.; Hedlund, J .; Holmgren, A.; Sterte, J. J. Mater. Chem. 1999, 9, 1507. 20 Mintova, S.; Schoeman, B.; Valtchev, V.; Sterte, J .; Mo, S. Y.; Bein, T. Adv. Mater. 1997, 9, 585. 21 Kim, J. 8.; Huang, W. X.; Wang, C. L.; Bruening, M. L.; Baker, G. L. “Bottle Brush Brushes: Ring-Opening Polymerization of Lactide from Poly(hya’roxyethyl methacrylate) Surfaces” in Polymer Brushes, Advincula, R. C.; Brittain, W. J .; Caster, K.C.; Riihe, J ., Eds, Wiley-VCH, Weinheim, Germany, pp. 105-117 (2004). 22 Li, X.; Wei, X. L.; Husson, S. M. Biomacromolecules 2004, 5, 869. 23 Biesalski, M.; Riihe, J. Langmuir 2000, 16, 1943. 106 Chapter 4 Surface Tethered Conducting Polymers via a Grafting-from Approach 4.1. Introduction Over the past several decades, n-conjugated organic polymer materials have been . . . . . la . 1b extensrvely investigated as the active element In sensors, optoelectronic and . . . 2 . . semiconducting devrces, and as electroluminescent, photoconductrng, electron- . . . . 3 . transporting, hole-transporting, and ron-dopable materials. One of the most wrdely used materials is poly(thiophene), which has attractive properties such as high electrical . . . . 4 . conductrvrty, electrochromrsm and electrolumrnescence. Poly(thlophene) and related polymers are synthesized by two general methods, electro-polymerization directly from electrode surfaces, or step-growth methods such as chemical coupling of di-bromo monomers, or chemical oxidation by FeCl3 and other oxidants. Electropolymerization conveniently deposits electroactive polymers directly onto sensor substrates, and permits control over the coating thickness, since the polymerization proceeds by application of a precisely controlled potential. In spite of these advantages, poor interfacial adhesion between the conducting polymer and the 5,6,7,8,9, . . 10 . . . . electrode rs a major problem. One strategy for Improvmg adhesron rs to form conducting polymers from polymerizable precursors preadsorbed on a metallic substrate . 5-7 . . (gold, nickel, or platinum) via thiol functional groups. An altematrve approach rs to graft conducting polymers onto metal oxide surfaces, such as indium tin oxide (ITO),9 107 where the key step is-the chemisorption of a preformed conducting polymer or its precursor. Recently, conjugated polymer network films were formed on conducting surfaces by the synthesis of precursor polymers from monomers with pendant electroactive units and then elaborating the pendant unit by electropolymerization or chemical oxidation. ll,12,13,14,15,16,17 . . . . . The resulting polymer films are rntrrnsrc conducting polymer networks with having both inter- and intramolecular cross-links between the pendant monomer units. Surface-initiated ATRP is attractive for modifying surfaces since surface properties are easily modified by varying the composition of the polymer brush, grafting density, the degree of polymerization18 and most importantly, delamination of polymer layer from surface can be eliminated. Recently, Carter et al. reported the successful grafting of disubstituted polyacetylene brushes grown from modified silicon and quartz surfaces . . . . . . 19 . usrng a transition-metal—catalyzed polymerization technrque. Huck and Friend used surface-initiated ATRP to produce tethered poly(triarylamine acrylate) hole transport materials for use in photovoltaic devices.20 Compared to devices prepared by solution- casting methods, devices prepared from tethered polymers showed enhanced conductivity, which was attributed to control over polymer architecture and morphology. Advincula et al. grew poly(vinyl carbazole) brushes from ITO by surface initiated free radical polymerization, controlling the graft density and brush length. In a second step, they cross-linked the poly(vinyl carbazole) brush electrochemically.21 108 Herein, we report the growth of poly(terthiophene methacrylate) from gold and ITO substrates using ATRP. Electropolymerization of the pendant terthiophene groups forms conducting cross-linked polymer network films of oligo— and polythiophene. 4.2. Experimental section 4.2.1. Materials Unless otherwise noted, all chemicals were obtained from Aldrich. ll-Mercapto- l-undecanol (MUD, 97%), 2-bromopropionyl bromide (2-BPB) (97%), 3- thiophenemethanol, thiophene-3-carboxaldehyde, tetrakis(triphenylphosphine) palladium, thiophene-2-boronic acid, methacryloyl chloride (98%), 3,4-ethy1enedioxythiophene (EDOT, 97%), NJV-dimethylformamide (DMF, 99.8%), dimethoxyethane (98%), NaBH4, Cu(I)C1 (99.999%), Cu(I)Br (99.999%) and hexmethyltriethylenetetraamine (HMTETA, 99%) were used as received. Triethylamine and acetonitrile were distilled from calcium hydride under reduced pressure and under an inert atmosphere. Tetraethylammonium perchlorate (TEAP) was dried under vacuum at 80 0C for 12 hours and stored in glove bag. ITO (indium tin oxide) coated glass slides were obtained from Delta technologies. 4.2.2. Characterization Methods 1 l3 . . H and C NMR analyses were camed out at room temperature on a Varran UnityPlus-SOO spectrometer at 500 and 300 MHz, respectively, with the chemical shifts reported in ppm and referenced to signals from residual protons in the solvent. Film thicknesses were measured using a rotating analyzer ellipsometer (model M-44, J. A. Woollam) at an incident angle of 75°. The data were analyzed using WVASE32 software, and thickness and refractive index determinations were performed on at least three spots 109 on each substrate. The refractive index of the films was assumed to be 1.5 and then fitted with the film thickness. Reflectance FTIR spectroscopy was performed using a Nicolet Magna-IR 560 spectrometer containing a PIKE grazing angle (80°) attachment. UV—vis measurements were taken on an Agilent technologies 8453 spectrometer. Atomic force microscopy (AFM) images were obtained in tapping mode with Multimode AFM and NanoScope IV software (Digital Instruments, Santa Barbara, CA) at room temperature. A tapping mode probe (NSClS) with a nominal frequency of 300 kHz was used for all experiments. 4.2.3. Synthesis of 3—methylthienyl methacrylate (MTM) 3-Thiophenemethanol (5.7 g, 50 mmol), dry triethylamine (7.3 g, 71 mmol), and CuCl 25 mg were dissolved in 35 mL of dry diethyl ether. A solution of methacryloyl chloride (5.35 g, 51 mmol) in 35 mL of dry diethyl ether was added slowly at O 0C. The mixture was stirred for 2 h, and then filtered through a silica gel column to remove triethylammonium chloride. After solvent evaporation, the residue was stirred overnight in a 1:1 mixture of methylene chloride and 2 M NaOH. The organic layer was separated, washed twice with water, and dried over CaClz. After solvent evaporation, the residue was distilled under vacuum (Vigreux column). bp (1 mmHg): 90 OC. Yield: 70%. 1H NMR (CDC13): 8 1.9 (s, CH3), 5.1 (s, 0CH2), 5.5 (5, vinyl H), 6.1 (s, vinyl H), 7.06 (m, ring H), 7.2 (m, ring H). 110 4.2.4. Synthesis of [2, 2':5', 2"-terthiophen]-3’-ylmethyl methacrylate (TTMM) 4.2.4a. Synthesis of 2,5-dibromoformyl-3-thiophene (1): A solution of Brz (0.562 mL, 20 mmol) in anhydrous CHC13 was added drop wise and under vigorous stirring, to a solution of thiophene-3—carboxaldehyde (0.500 mL, 5.5 mmol) in anhydrous CHC13 (1.5 mL) in a 10 mL flask, at 60 CC under inert atmosphere. The reaction mixture was refluxed for 5 h, and brought to room temperature and then poured into ~10 mL of water at 0 oC. The organic phase was neutralized with a NazCO3 saturated solution, dried with MgSO4, filtered, and evaporated to dryness. The residue was chromatographed on flash silica gel 60 using 90/ 10 hexanes/ethyl acetate as eluant to give a yellowish orange solid. Crystallization from heptanes affords pale yellowish needlelike crystal in 40% yield. 1H NMR (CDC13): 5 = 9.80 (s, 1H, CHO), 7.34 (s, 1H, H4). 4.2.4b. Synthesis of 3'-formyl-2,2':5',2"-terthiophene (2): A 50 mL 3-neck flask, equipped with condenser, magnetic stirrer and N2 inlet, was charged with 1, (0.2086 g, 0.77 mmol), tetrakis(triphenylphosphine) palladium(0) (0.0531g, 0.046 mmol) and 8 mL of l,2-dimethoxyethane. After stirring for 10 min at room temperature, thiophene-Z-boronic acid (0.2362 g, 1.84 mmol) was added, and followed immediately by 5 mL of aqueous 1M NaHCO3 solution. The reaction mixture was refluxed for 4 h under nitrogen. After cooling to room temperature, the mixture was filtered and the organic solvent was evaporated under reduced pressure. After removal of the solvent, 10 mL of water were added to the residue, and the mixture was extracted with diethyl ether (3X50 mL). The combined organic phases were washed with water and with saturated 111 NaCl solution, and dried over MgSO4. After filtration and solvent evaporation, the crude product was chromatographed on flash silica gel 60, using 90/10 hexanes/ethyl acetate as the eluant. Removal of the solvent yielded greenish-yellow powdery solid. Yield 75%, 1H NMR (CDC13): 8 10.09 (s, 1H, CHO), 7.57 (s, 1H, H4'), 7.51 (dd, 1H,J== 5.1 Hz,J= 1.2 Hz, H5), 7.32 (dd,1H, J= 3.6 Hz, J=1.2 Hz, H3), 7.30 (dd, 1H, J: 5.1Hz, J=1.2 Hz, H5" ), 7.23 (dd, 1H, J= 3.6 Hz, J: 1.2 Hz,H3"),7.17 (dd,1H,J= 5.1 Hz, J = 3.6 Hz, H4), 7.05 (dd, 1H, J= 5.1 Hz, J== 3.6 Hz, H4”). 4.2.4c. Synthesis of 3'-hydroxymethyl—2,2'5',2"-terthiophene (3): Compound 3 was synthesized by the procedure of Zanardi et al.23in 95% yield. 1H NMR(CDC13): 8 7.33 (dd, 1H, J= 5.1 Hz, J: 1.0 Hz, H5), 7.23 (s, 1H, H4' ), 7.22 (dd, 1H, J= 5.1 Hz, J =1.0 Hz, H5H ), 7.20 (dd, 1H, J: 3.6 Hz, J== 1.0 Hz, H3), 7.17 (dd, 1H, J: 3.6 Hz, J= 1.0 Hz, H3H ), 7.07 (dd, 1H, J= 3.6 Hz, J: 5.1 Hz, H4), 7.02 (dd, 1H, J= 3.6 Hz, J== 5.1 Hz, H4”), 4.74 (s, 2H,CH2), 1.78 (5 (br), 1H, 0H). 4.2.4d. Synthesis of [2,2':5',2"-terthiophen]-3’-y1methyl methacrylate (TTMM): Compound 3 (4.5 g), 3.15 mL of dry triethylamine, and CuCl (25 mg) were dissolved in 50 mL of dry diethyl ether. A solution of freshly distilled methacryloyl chloride (1.68 mL) in 50 mL of dry diethyl ether was added slowly at O C)C., and the mixture was stirred for 2 h. The solution was filtered a silica gel column to remove triethylammonium chloride, and the solvent was evaporated to dryness. The residue was stirred overnight in a 1:1 mixture of methylene chloride and 2 M NaOH. The organic 112 layer was separated, washed twice with water, and dried over CaClz. After solvent evaporation, the residue was re-crystallized from diethyl ether to provide TTMM as green crystals. Yield: 50%. 1H NMR (CDC13): 8 7.37 (dd, 1H, J: 5.5 Hz, J: 1.0 Hz, H5), 7.25 (dd, 1H, J: 5.5 Hz , J= 0.5 Hz, H4’), 7.22 (m, 2H, H5”), 7.20 (dd, 1H, J: 3.5 Hz, J= 1.0 Hz, H3), 7.11 (dd, 1H, J: 4.0 Hz, J: 5.0 Hz, H3”), 7.04 (dd, 1H, J: 4.0 Hz, J= 5.5 Hz, H4), 6.18 (5, vinyl H), 5.62 (5, vinyl H), 5.24 (s, OCHZ), 2.0 (s, CH3). 13c {1H} NMR (CDC13): 167.41, 136.84, 136.38, 136.15, 134.65, 134.52, 133.32, 128.14, 128.12, 127.0, 126.79, 126.76, 126.29, 125.09, 124.29, 60.49, and 18.61. High resolution MS: m/z calc. for C17H1402S3+z 346.0156; Found: 346.0150. 4.2.5. Synthesis of the silane initiator: The trichlorosilane initiator was . . . . 18b . . synthesrzed followrng prevrously publlshed procedures. 2-Bromorsobutyryl bromide (1.85 mL, 15.0 mmol), was added dropwise to a stirring solution of allyl alcohol (1.02 mL, 15.0 mmol) and triethylamine (2.51 mi. 18.0 mmol) in dichloromethane (10 mL) at 0 C>C, under nitrogen atmosphere. After 1 hour of stirring the temperature was raised to room temperature and the reaction mixture was stirred for a further 3 hours. The precipitate was then removed under reduced pressure and the organic layer was washed with an aqueous saturated NH4C1 solution, followed by a wash with water. The organic layer was then dried over anhydrous MgSO4 and the solvent removed in vacuo. The product was purified by column chromatography (silica) using 9:1 hexane: ethyl acetate as the eluent. The solvent was then removed under reduced pressure to yield the clear, 113 liquid product prop-2-enyl—2-bromo-2-methyl propionate (1.72 g, 8.31 mmol, 55.4% yield). 1H NMR (CDC13, 500 MHz): 5 = 1.94 (6H, s), 4.66 (2H, d), 5.27 (1H, d), 5.38 (1H, d), 5.93 (1H, m). A solution of hexachloroplatinic acid (21 mg, 51 umol) in 1:1 (v/v) ethanol: 1,2- dimethoxyethane (3.75 mL) was added dropwise to a solution of pr0p-2-enyl-2-bromo-2- methyl propionate (0.97 g, 4.7 mmol) in trichlorosilane (15 mL, 0.15 mol) under nitrogen atmosphere. The reaction was stirred (in the dark) for 18 hours. Toluene (5.0 mL) was then added and unreacted trichlorosilane was removed under reduced pressure. Dichloromethane (20 mL) was added and then removed in vacuo to remove all remaining trichlorosilane. The resulting product, 2-bromo-2-methyl-propionic acid 3- trichlorosilanyl—propyl ester was used without further purificatioan NMR (CDCl3, 300 MHz): 8 = 1.50 (2H, m), 1.93 (8H, m), 4.20 (2H, m). 4.2.6. Preparation of initiator immobilized flat substrates: ITO slides (~1 cmz) were first sonicated in acetone (10 min), followed by sonication in IPA (10 min). All substrates were cleaned using UV/O3 chamber for 1 hour. ITO substrates were then placed in a ~l mM solution of the trichlorosilane initiator, containing ~10 mM triethylamine in toluene. Substrates were left, covered, in the solution, at room temperature, for 18 hours under N2 atmosphere. The substrates were then removed and, successively: washed with toluene, sonicated for 1 minute in toluene, washed with acetone, washed with ethanol, and dried under a stream of nitrogen. The surfaces were stored under nitrogen until further use. 114 4.2.7. Surface initiated polymerization on gold and ITO substrates: In a N2- filled drybox, 0.7 mg of CuBr and 5.55 mg of HMTETA were added to a round-bottom flask containing 1 mL of a degassed solution of TTMM monomer (500 mg) in DMF (DMF : monomer ~1 :1 (vzv). The mixture was well-stirred and heated with an oil bath to 50 0C until a transparent, green solution formed. The prepared solution was then transferred into a small vial containing an initiator-modified Au substrate to start the surface-initiated polymerization. After a set reaction time at 50 0C, the substrate was removed from the vial, washed with THF and isopropanol sequentially, and then was dried under a flow of N2. The similar polymerization condition. has been applied for growing MTM from surfaces. 4.2.8. Electrochemistry: All electrochemical modifications were performed using a CHI650a computerized potentiostat (CH Instruments, Inc., Austin, TX).The electrochemical experiments, i.e., CV (cyclic voltammetry) were carried out in acetonitrile containing tetraethyl ammonium perchlorate (TEAP~5><102 M) as a conducting salt. For heterocoupling experiments, the solution contained 10‘2 M of EDOT along with TEAP. All experiments were carried out in a glovebox under an inert and dry atmosphere at room temperature. In a three electrode system, the working electrode (gold or ITO-coated glass substrate) was placed between the reference electrode (Pt wire) and the counter-electrode (graphite). All potentials were measured in this work with respect to a Pt wire used as a quasi-reference electrode. For this reason, the potentials may not be directly compared. 115 4.3. Results and Discussion: 4.3.1. Monomer synthesis: As shown in Schemes 4.1 and 4.2, the synthesis of 3- methylthienyl methacrylate (MTM) and the precursors to 3’-methylterthienyl methacrylate (TTMM), i.e. compounds 1, 2 and 3, were synthesized by modified versions of the procedures of Yagci22 and Seeber23et al. TTMM was obtained in high purity by the reaction of 3 with methacryloyl chloride in dry diethyl ether, similar to the synthesis of monomer MTM. Methacrylate monomers were chosen for their facile polymerization via ATRP, while the polymerizable thiophene and terthiophene groups have shown good . . . 24,25,26 conducting pr0pert1es In related polymers. HO O Et3N, Dry Ether o / \ + O > 8 Cl 0 cc. 2 h / \ 70% 3 Scheme 4.1: Synthesis of 3-methylthienyl methacrylate (MTM) 116 KP! Biz, CHCI3 KP! / \ j» Br /\ Br S RT,24h S 70% O H Pd PPh ,NaHCO ¢( 314 3 U + / \ 600/0 8 B(OH)2 Br 8 Br 85‘V NaBH4, RT 0 45 min Et3N, Dry Ether + O F (3' RT, 24 h 78% Scheme 4.2: Synthesis of [2,2':5',2"-terthiophen]-3'-ylmethyl methacrylate (TTMM) 4.3.2. Synthesis of silane initiators for ITO substrates Typically, we use a ll-carbon long alkyl chain in trichlorosilane initiators to ensure a well-defined SAM on ITO. However, we used a short alkyl chain since the longer chain may act as a resistive element and inhibit the polymerization of the pendent thiophene methacrylares. The initiator was synthesized following the method of Huck et al.20 (Scheme 4.3), esterification of 2-bromoisobutyryl bromide with allyl alcohol, followed by hydrosilation of the intermediate ester to give the trichlorosilane initiator. The trichlorosilane functionality was chosen because trifunctional organosilanes are more . . . 27 reactive towards surfaces than thelr monosubstltuted analogues. These SAMs are 117 expected to be stable due to formation of polysiloxane networks bound to the substrate 28,29 surface For gold surfaces, we used a 11-carbon long ATRP initiator as shorter chain initiators failed to form uniform self-assembled monolayers on gold surfaces. Br 0 El N 2 h WOH 4' Br/lksr S730? WOW 0 O l H Br 01‘ Cl 0 Scheme 4.3: Synthesis of the trichlorosilane initiator Cl Pt(0)l C'xs'i/Cl 100% 4.3.3. Deposition of SAMs on ITO surface The formation of trichlorosilane-based SAMs on silica surface is well known.28 However, generating a well defined SAM on an ITO surface was expected to be more difficult because of the high surface roughness of ITO and the low coverage of hydroxyl groups.30 Nevertheless, some methods have produced SAMs on ITO. Examples include microcontact printing,31 a one-hour soak in a 1 mM solution of the chlorosilane at room temperature,32 and refluxing a solution of trimethoxysilane in toluene over ITO for 7 days.30 For the latter case, Markovich, et al. achieved about 90% coverage of surface sites. Considering the higher reactivity of a trichlorosilane compared with a . . . ' 20 trimethoxysilane, the reaction was carried out srmllar to the method of Huck et al., at room temperature for 18 hours to avoid polymerization (Scheme 4.4). In addition, triethylamine was added18b to drive the reaction to completion. While this SAM 118 deposition method may not provide complete surface coverage, estimates of initiator efficiency suggest that ~10% of surface bound initiating molecules initiate a polymer brush,33 and therefore having less than full surface coverage should not be a significant factor in the polymer brush synthesis. Typically, a full ITO wafer was completely . . . . 2 . . denvatlzed, stored under nrtrogen, and fractured lnto (~1 X 2 cm ) pleces Just before use. The surfaces were stable and polymerizations carried out with l-year old initiator monolayers gave similar film thicknesses when used in similar polymerization conditions. Br Cl\ ,OH O -OH Cl—Si—(CH2)—-—-O *O-Si—(CH2)—-O Br 01’ 3 o \ 3 lTO ;> ITO O Toluene, RT,18 h / O —OH "O‘Sl—(CH2)"‘O \OH 3 Br Scheme 4.4: Initiator SAM deposition on ITO surfaces 4.3.4. Synthesis of polymer brushes from gold and ITO surfaces The syntheses of polyMTM and polyTTMM brushes from gold and ITO surfaces are shown in Scheme 4.5. Films on gold surfaces were characterized using ellipsometry, but for ITO surfaces, tapping mode AFM was used to measure PolyTTMM film thicknesses since the film absorbs light, which makes ellipsometric measurement ambiguous. 119 ITO O r--O 0 Au L—S—(CH2)1—1-O Br OH Br AU F's—(CH2)1T'O O —o-$i-(CH2)-o 3 Br 0 =2 CuBr I PMDETA DMF , 55 °C ‘2. O / \ \ / Wig —> CuBr/ PMDETA DMF , 55 °C CuBrl HMTETA DMF, 90 °C ll Au H—o lTo—o-jSI—(CH )o ( ——-o 28 0 9H3 I_...S_ __ _ (CH21110 (Cszzgr O O [l \i S O Scheme 4.5: Surface initiated polymerization of MTM from gold surface (top), TTMM from gold surface (middle) and TTMM from ITO surface (bottom). 120 PMTM brushes were grown from gold surfaces at 55 0C in anhydrous DMF using a CuBr/HMTETA catalyst. This system was previously shown to be compatible with the synthesis of methacrylate monomers via ATRP,34 and also proved successful for the synthesis of polyMTM brushes. The polyMTM brush was typically synthesized from a concentrated polymerization solution (1 g/mL), which allowed access to a greater range of film thicknesses. Similar solutions were used for the synthesis of polyTMM brushes. The kinetic plots for the synthesis of both brushes (Figure 4.1) show the characteristics of significant termination during ATRP;35the polymer chains initially grow rapidly, and as termination consumes growing chain ends, the plots plateau. Complete termination occurs more quickly for the polyTTMM systems than than for polyMTM. High radical concentrations result in increased probability of radical combination and termination of growing polymer chains. Usually, increasing the reaction temperature provides thicker films, but when the polymerization temperature was increased from 55 0C to 90 0C, the growth rate for polyTTMM brushes grown on ITO surfaces did not increase. No CuBr2 was explicitly added to the polymerizations, but Cu(II) should be present as a consequence of of CuBr with the initiator. For polyTTMM brushes, early termination may be related to the stronger complexing ability of terthiophene units with Cu(II) than for monothiophenes, as evidenced by Rajesh et al.36 Thus, Cu(II) generated during initiating of polymerization is sequestered by the pendent terthiophene of TTMM, leading to a high concentration of radicals and almost no deactivation. 121 The polyTTMM film thicknesses were intentionally kept low to ensure their crosslinking by electrochemical methods. In thicker films, electrochemical-based cross- linking may be hindered by low permeation rates of anions or monomer through the polyTTMM matrix. PolyMTM and polyTTMM brushes grown from gold surfaces were characterized by FTIR. The spectra for polyMTM and polyTTMM show the expected bands for the carbonyl peak at 1733 cm-1 and sp2 C-H stretching at 3100 cm.1 expected for thiOphene rings (Figure 4.2). 250 O 8 E O 5150 r to (o “c’ 3:, 100 2). .c l- O 50 1- Q [j I: A gm 0 I _L A P _1 0 5 10 15 20 25 30 Time (hours) Figure 4.1: Evolution of the ellipsometric brush thickness with time for the polymerization of polyMTM and polyTTMM from gold and ITO surfaces; a polyMTM brush grown on gold surface (9), a polyMTM brush grown on ITO surface (0), a polyTTMM brush grown on gold surface ([3), a polyTTMM brush grown on ITO surface (A). 122 0.02 (a) Absorbance (b) I I I l 3500 3000 2500 2000 1500 1000 Wavenumbers (cm'l) Figure 4.2: Representative examples of reflectance FTIR spectroscopy of gold surface grafted with (a) 50 nm polyTTMM brush (b) 100 nm polyMTM brush. 4.3.5. Electrochemical crosslinking PolyTTMM films (~30 nm) on ITO and gold substrates were placed in a 0.1 M tetraethylammonium perchlorate (TEAP) solution in acetonitrile under an inert atmosphere to avoid overoxidation, and then crosslinked electrochemically (Scheme 4.6). The substrates served as the working electrode, a graphite electrode as the counter electrode and a Pt wire as the reference electrode. The substrate was scanned at 100 mV/s 123 O Electrochemical l—O . . CH ._ . Polymerlzatlon 0. 9H3 . 3 lTO—O-SI- CH O C B ._ L. - -_ o’ ( 2):3 (CH2 imr _ fl ITO 00,81 (CH2)3O CHzC—tCHzCigBr O O O O Scheme 4.6: Electrochemical cross linking of polyTTMM brush on ITO surfaces from -25 to +1500 mV for up to 20 cycles in order to cross-link the polyTTMM film (Figure 4.3). The direct electrochemical polymerization of the polyMTM brush was not achievable under our conditions, but consistent with previous reports of the low reactivity of thiophenes substituted with esters in the 3-position.26 PolyTTMM’s higher reactivity . . 26 . stems from the extended n-structure of the terthienyl morety which lowers the oxidation potential of polyTTMM brush relative to polyMTM. In addition, the terthienyl unit is less sterically challenged with respect to coupling in the 2- and 5”-positions, compared to the 2- and 5-positions of a single thienyl unit. A well-defined oxidation peak starting at ~1.1 V/Pt (Figure 4.3, first scan) was observed for the polyTTMM brush by cyclic voltammetry. Because films of grafted poly (alkylacrylate), e.g., PEA (poly(ethyl acrylate)), do not react upon anodic polarization, this peak can only be attributed to the oxidation of the aromatic thiophene rings. The reduction peak of the accordingly formed polyTTMM two-component films is observed at ~ 0.63 V/Pt during the reverse scan. These observations confirm that the thienyl rings attached to the conducting substrate (ITO / gold) as a result of the surface initiated polymerization via ATRP of the parent acrylates (TTMM), remain available to electrochemical polymerization and form polymers with a more extended 7: electron 124 2.50E-04 2.00E-04 )- 2: 1505-04 ~ 5 S. 5- 1005-04 - 2 5' 5.00E—05 - 0.00E+00 - -5.00E-05 e . 1 . -0.5 0 05 1 1.5 2 'E(v) Figure 4.3: Cyclic voltammetry (20 cycles) of ~30 nm polyTTMM brush coated on ITO surface from -25 to 1500 mV at 100 mV/s. conjugation. The voltammetric scans were repeated (20 times) (Figure 4.3, 20th scan) or less if the polymerization peak was no longer observable. On the last scan (20th scan), the oxidation peak appears at a lower anodic potential (E~0.75 V/Pt), and corresponds to a 37,38,39,40,4l more extended conjugated system than the monomeric terthiophenes. The de-doping peak remains visible. These redox potentials are higher than values reported . . . . . . 42 . for terthiophene reductlon under the srmrlar condrtrons but, they are comparable to the data reported by Advincula et al.21 for copolymers of carbazole and terthiophene + + containing methacrylate (EC = 0.6 V/Ag/Ag and Ea = 1.08 V/Ag/Ag ). The reduced 125 mobility of a thiophene attached to polymethacrylate chains might explain the formation of polythiophenes with shorter conjugation lengths. Scan rate dependence studies (Figure 4.4) of the cross-linked polyTTMM at scan rates of 20-80 mV/s in a 0.1 M TEAP/CH3CN electrolyte solution (potentials are reported relative to Pt as a quasi reference electrode) revealed linear behavior (Figure 4.5) similar to data of Sotzing et al.39 for electropolymerized poly(terthiophene) films. This indicates a surface-confined reaction as the sole phenomenon. 5.00E-05 Scan rate decreasing 4.00E-05 2 3.00E-05 - 2.00E-05 - 1.00E-05 r Current (Amp) 0.00E+OO r -1.00E-05 - -2.00E-05 1 ' ' ‘ .05 0 0.5 1 1.5 2 E (v) Figure 4.4. Scan rate dependency study of polyTTMM brushes coated on ITO surface at scan rates of 20-80 mV/s in a 0.1 M TEAP/CH3CN electrolyte solution. 126 3.5 - ’5. E a - Z‘ 17, 2.5 - C .9 .E 2 - 72'. t 1.5 - 3 0 2 1 - Q) a 0.5 - 0 r r r 10 3O 50 7O 90 Sean rate (mv/sec) Figure 4.5: Plot of peak current intensity versus scan rate at the maximum of the oxidation wave of polyTTMM brush coated on ITO surface. Examination of the films by UV-vis before and after the electro-polymerization showed a small spectral difference arising from the formation of conjugated oligothiophene species (Figure 4.7). The Amax for the polyTTMM brush before . . 39 . crosslinking was ~ 360 nm (characteristic vibronrc pattern of terthiophene), and shifted to ~ 370 nm after electro polymerization. This pattern is comparable to the sexithiophene . 43 . based structure investigated prevrously. IR spectroscopy also suppports terthiophene coupling. The spectrum of polyTTMM (Figure 4.6) shows bands at 3100, 1450 and 790 cm'] due to OH stretching, wagging and C-H out-of-plane deformation modes from the thiophene rings respectively. After crosslinking, these bands decreased dramatically. A 127 -l . . new band at 1620 cm , due to the conjugated throphene, appears in the spectrum (Figure 4b). The persistence of the carbonyl stretching mode at 1710 cm.1 in both spectra confirms that the polymer backbone is apparently unchanged. These results are consistent with a network of conducting polythiophene chains of short conjugation length, perhaps dimers, grown from pristine polyTTMM. The morphology of the pregrafted polyTTMM films may change upon oxidation of the terthiophene units. Films before and after oxidation were investigated by tapping mode AFM. The surface shows an irregular morphology (Figure 4.8), compared to the smooth films before oxidation. We suspect that anodic oxidation rigidified the polythiophene segments as well as the polyacrylate chains, which resulted in a more heterogeneous surface dominated by grains of various sizes. However, the surface roughness did not change significantly after oxidation. 128 0.01 (b) d) 0 C N D L- O (I) .0 < (a) 3250 2750 2250 1750 1250 750 Wavenumbers (cm") Figure 4.6: Representative examples of reflectance FTIR spectroscopy of gold surface coated with (a) 50 nm polyTTMM brush (b) electrochemically crosslinked polyTTMM brush. 129 0.05 PTTMM brush on ITO surface or ‘/ o c In a o ElectrOpolymerized PTTMM brush on ITO 3 surface 42 300 400 500 600 700 800 Wavelength(nm) Figure 4.7: UV-vis absorption spectra of polyTTMM brush grown on ITO surface (blue line) and electrochemically cross linked polyTTMM (EPTTMM) brush on ITO surface (red line). 130 5.00 100.0 nm 50.0 nm 0.0 2.50 nm ' 0.0 0.0 2.50 5.00 pm 5'00 50.0 nm 25.0 nm 0.0 nm 2.50 , 0.0 0.0 2.50 5.00 mm Figure 4.8: Topographical AF M images of ITO surface coated with polyTTMM brush (upper image) and electro polymerized polyTTMM brush (bottom image) taken with tapping mode imaging. Upper image is a 5 X 5 pm survey scan with high surface coverage and an rrns roughness of 3.8 nm, as calculated with Nanoscope IV software. The bottom image is a 5 X 5 am scan with an rms value of 2.9 nm. 131 4.3.6. Anodic copolymerization of the grafted polythiophene units and additional thiophene derivative in solution - heterocoupling The data for the electrochemical oxidation of polyTTMM suggest that terthiophene cross-linking was inefficient. To increase the conjugation length of polyTTMM films, the potentiostatic oxidation of polyTTMM can be performed in the presence of thiophene derivatives such as 3,4 ethylenedioxythiophene (EDOT) (Scheme 4.7). 44 ’46 According to Zotti and coworkers,42 EDOT is the best choice for heterocoupling terthiophene monolayers as well as polyterthiophene films. Electropolymerization involves the formation of carbon-carbon bonds by reaction . . 45,47 . . . . . between two radical cations, and the prrmary reqursrte for heterocoupling IS the generation of radical cations on the electrode surface (for the polyTTMM brush) and in the solution close to the electrode (for the monomer). Thus, the heterocoupling is favored by a small difference in oxidation potential between the surface and the solution components. The second factor for the coupling process is the reactivity of radical cations toward coupling. Zotti et al.42 claim the second factor is predominant as the oxidation potential of polyterthiophene films (0.5 V/ Ag/Ag+) is much lower than the oxidation potential of EDOT (1.16 V/ Ag/Ag+). They surmised that the strong positive shift of polyterthiophene oxidation moves it close to EDOT oxidation which should kinetically favor the copolymerization. Figure 4.9 shows the voltarnmogram recorded for the oxidation of a polyTTMM film in the presence of EDOT. The polymerization peak starts at ~l.3 V/Pt, nearly the 132 same potential as the oxidation of the polyTTMM film in the absence of EDOT. Repeated cycling (20 times) of the polyTTMM film between -0.5 V and 1.5 V, while immersed in a dilute solution of EDOT, resulted in the deposition of a polymer film and a concomitant growth of a redox wave in the cyclic voltammogram (Figure 4.9, 20th scan). With each scan, more material deposited onto the electrode, resulting in an increase in the anodic and cathodic peak currents associated with the polymer. The polymerization peak (~1.1V) of the as-prepared conducting polymer occurs at lower cathodic potentials compared to polyTTMM in the absence of EDOT, which is consistent with the formation of chains with longer conjugated length. The UV-vis spectra of a polyTTMM brush before and after heterocoupling with EDOT (Figure 4.10) shows broadening of the absorption band and a shift of oscillator strength to longer wavelengths, consistent with increased conjugation. However, we cannot exclude the possibility of a small amount of EDOT homopolymer adsorbed on the ITO surface after extensive washing. Electrochemical O 0 CH3 crosslinking H—0 0 CH L _‘ -_ . . H , 3 |T0__OO.SI (CH2)30 (CHzeigr ——> lroeo-si—(cnzgo (CHZClFBr r—\ 00 To 00 Scheme 4.7: Heterocoupling of polyTTMM brush with EDOT on an ITO surface 133 5.00E—03 4.00E-03 r 3.00E-03 r 2.00E-03 )- 1.00E-03 *- 0.00E+00 - E \ -1.00E-03 th 20 scan Current (amp) -2.00E-03 H 4 ‘ ‘ ' -1 05 0 0.5 1 1.5 2 E (v) Figure 4.9: Cyclic voltammetry (20 cycles) of ~30 nm polyTTMM brush grafted on ITO surface in C H3CN added with TEAP (0.05 M) in the presence of EDOT in solution (first (blue line) and 20th scan (red line)) at the scan rate of 20 mV/s 134 0.05 PTTMM brush on ITO surface a: u : en E O Electrochemically heterocoupled PTTMM brush 3 on ITO surface < 300 400 500 600 700 800 Wavelength(nm) Figure 4.10: UV-vis absorption spectra of polyTTMM brush grown on ITO surface (blue line) and electrochemically heterocoupled polyTTMM brush with EDOT on ITO surface (red line). 135 4.4. Conclusions Poly(terthiophene methacrylate) brushes were successfully grown from ITO and gold electrodes using surface initiated ATRP. The resulting films were homogeneous with smooth domains and features. The polyTTMM brush were electrochemically oxidized by cyclic voltammetry, to form a crosslinked polymer network with short segments of conjugated poly(terthiophene) segements. Including EDOT during oxidation resulted in heterocoupled EDOT and polyTTMM with increased conjugation lengths. Uniformly grafted conducting polymer brush may be useful in photovoltaic devices. 136 4.5. References: 1 (8) Wang, X.; Kim, Y.-G.; Drew, C.; Ku, B.-C.; Kumar, J.; Samuelson, L. A. Nano Lett. 2004, 4, 331. (b) Tong, H.; Wang, L.; Jing, X.; Wang, F. Macromolecules 2003, 36, 2584. (c) Murphy, C. B.; Zhang, Y.; Troxler, T.; Ferry, V.; Martin, J. J .; Jones, W. E., Jr. J. Phys. Chem. B 2004, 108, 1537. 2 Dodabalapur, A.; Torsi, L.; Katz, H. E. Science 1995, 268, 270. 3 Skotheim, T. A.; Reynolds, J. R. Handbook of Conducting Polymers, 3rd ed.; CRC Press: Boca Raton, FL, 2007. 4 (a) Barbarella, G.; Melucci, M.; Sotgiu, G. Adv. Mater. 2005, 17, 1581. (b) Kirchmeyer, S.; Reuter, K. J. Mater. Chem. 2005, 15, 2077. (c) Nalwa, S. H. Handbook of Organic Conductive Molecules and Polymers, Wiley—VCH, New York, 1997. (d) Fichou, D. Handbook ofOligo— and Polythiophenes, Wiley—VCH, New York, 1998. 5 McCarley, R. L.; Willicut, R. J. J. Am. Chem. Soc. 1998, 120, 9296. 6 Smela, E. Langmuir 1998, 14, 2996. 7 Sayre, C. N.; Collard, D. M. Langmuir 1995, 11, 302. 8 Lang, P.; Mekhalif, Z.; Garnier, F. J. Chim. Phys. 1992, 89, 1063. 9 Ntlesch, F .; Si-Ahmed, L.; Franciiois, B.; Zuppiroli, L. Adv. Mater. 1997, 9, 222. 10 Simon, R.; Ricco, A.; Wrighton, M. J. Am. Chem. Soc. 1982, 104, 2031. 11 Sebastian, R. M.; Caminade, A. M.; Majoral, J. P.; Levillain, E.; Huchet, L.; Roncali, J. Chem. Comm. 2000, 507. 12 Taranekar, P.; Baba, A.; Fulghum, T. M.; Advincula, R. Macromolecules 2005, 38, 3679. (b) Taranekar, P.; Fulghum, T.; Baba, A.; Patton, D.; Advincula, R. Langmuir 2007, 23, 908. 13 Deng, S. X.; Advincula, R. C. Chem. Mater. 2002, 14, 4073. 14 Taranekar, P.; Fan, X.; Advincula, R. Langmuir 2002, 18, 7943. 15 Inaoka, S.; Roitman, D. 13.; Advincula, R. C. Chem. Mater. 2005, 17, 6781. 16 Park, M. K.; Xia, C.; Advincula, R. C.; Schutz, P.; Caruso, F. Langmuir 2001, 17, 7670. 137 17 Jang, S. Y.; Sotzing, G. A.; Marquez, M. Macromolecules 2002, 35, 7293. 18 (a) Ejaz, M.; Yamamoto, S.; Ohno, K.; Tsujii, Y.; Fukuda, T. Macromolecules 1998, 31, 5934. (b) Husseman, M.; Malmstrom, E. E.; McNamara, M.; Mate, M.; Mecerreyes, D.; Benoit, D. G.; Hedrick, J. L.; Mansky, P.; Huang, E.; Russell, T. P.; Hawker, C. J. Macromolecules 1999, 32, 1424. (c) Kamigaito, M.; Ando, T.; Sawamoto, M. Chem. Rev. 2001, 101, 3689. 19 Jhaveri, 8.; Carter, K. Langmuir 2007, 23, 8288. 20 (a) Snaith, H. J.; Whiting, G. L.; Sun, B.; Greenham, N. C.; Huck, W. T. S.; Friend, R. H. Nano Lett. 2005, 5, 1653. (b) Whiting, G. L.; Snaith, H. J.; Khodabakhsh, S.; Andreasen, J. W.; Breiby, D. W.; Nielsen, M. M.; Greenham, N. C.; Friend, R. H.; Huck, W. T. S. Nano Lett. 2006, 6, 573. 21 Fulghum, T.; Taranekar, P.; Advincula, R. C. Macromolecules 2008, 41, 5681. 22 Cirpan, A.; Alkan, S.; Toppare, L.; Hepuzer, Y.; Yagci, Y. J. Polym. Sci, Part A: Polym. Chem. 2002, 40, 4131. 23 Zanardi, C.; Scanu, R.; Pigani, L.; Pilo, M.I.; Sanna, G.; Seeber, R.; Spano, N.; Terzi, F.; Zucca, A. Electrochim. Acta 2006, 51, 4859. 24 Jang, S. Y.; Sotzing, G. A. Macromolecules 2002, 35, 7293. 25 Jang, S. Y.; Sotzing, G. A. Macromolecules 2004, 3 7, 4351. 26 Watson, K. J .; Wolfe, P. S.; Nguyen, S. T.; Zhu, J.; Mirki'n, C. A. Macromolecules 2000, 33, 4628. 27 Fadeev, A. Y.; McCarthy, T. J. Langmuir 2000, 16, 7268. 28 Ulman, A. Chem. Rev. 1996, 96, 1533. 29 Zhuang, Y. X.; Hansen, 0.; Knieling, T.; Wang, C.; Romback, P.; Lang, W.; Benecke, W.; Kehlenbeck, M.; Koblitz, J. J. Micromech. Microeng. 2006, 16, 2259. 30 Markovich, 1.; Mandler, D. J. Electroanal. Chem. 2001, 500, 453. 31 Koide, Y.; Such, M. W.; Basu, R.; Evmenenko, G.; Cui, J.; Dutta, P.; Hersam, M. C.; Marks, T. J. Langmuir 2003, 19, 86. 32 Lee, J.; Jung, B. J.; Lee, J. 1.; Chu, H. Y.; Do, L. M.; Shim, H. K. J. Mater. Chem. 2002, 12, 3494. 138 33 Jones, D. M.; Brown, A. A.; Huck, W. T. S. Langmuir 2002, 18, 1265. 34 Patten, T. F.; Matyjaszewski, K. Adv. Mater. 1998, 10, 901. 35 Huang, W. X.; Kim, J. B.; Bruening, M. L.; Baker, G. L. Macromolecules 2002, 35, 1175. 36 Mahajan, R. K.; Sood, P. Int. J. Electrochem. Sci, 2007, 2, 832. 37 Skompska, M.; Peter, L. M. J. Electroanal. Chem. 1995, 383, 43. 38 (a) Taranekar, P.; Baba, A.; Fulghum, T. M.; Advincula, R. Macromolecules 2005, 38, 3679.. (b) Taranekar, P.; Fulghum, T.; Baba, A.; Patton, D.; Advincula, R. C. Langmuir 2007, 23, 908. 39 Jang, S. Y.; Sotzing, G. A.; Marquez, M. Macromolecules 2002, 35, 7293. 40 Fulghum, T.; Karim, S. M. A.; Baba, A.; Taranekar, P.; Nakai, T.; Masuda, T.; Advincula, R. C. Macromolecules 2006, 39, 1467. 41 Taranekar, P.; Fulghum, T.; Baba, A.; Patton, D.; Advincula, R. Langmuir 2007, 23, 908. 42 Zotti, G.; Zecchin, S.; Vercelli, B.; Berlin, A.; Grimoldi, S.; Groenendaal, L.; Bertoncello, R.; Natali, M. Chem. Mat. 2005, 1 7, 3681. 43 F ichou, D.; Horowitz, G.; Xu, 2.; Gamier, F. Synth. Met. 1992, 48, 167. 44 Labaye, D. E.; Jerome, C.; Geskin, V. M.; Louette, P.; Lazzaroni, R.; Martinot, L.; Jerome, R. Langmuir 2002, 18, 5222-5230. 45 (a) Waltman, R. J.; Bargon, J. Tetrahedron 1984, 40, 3963. (b) Waltman, R. J.; Bargon, J. Can. J. Chem. 1986, 64, 76. (c) Andrieux, C. P.; Audebert, P.; Hapiot, P.; Saveant, J. M. J. Phys. Chem. 1991, 95, 10158. (d) Audebert, P.; Catel, J. M.; LeCoustumer, J .; Duchenet, V.; Hapiot, P. J. Phys. Chem. 1995, 99, 11923. (e) Guyard, L.; Hapiot, P.; Neta, P. J. Phys. Chem. 1997, 101, 5698. (f) Audebert, P.; Catel, J.M.; Duchenet, V.; Guyard, L.; Hapiot, P.; LeCoustumer, J. Synth. Met. 1999, 101, 642. 139 Chapter 5 Polymerization of azidomethacrylates on gold surface and its elaboration via click chemistry 5.1. Introduction Atom Transfer Radical Polymerization (ATRP), one of the most widely deveIOped controlled radical polymerization (CRP) techniques, is extensively used to prepare polymer materials with predetermined molecular weights, narrow polydispersities, and advanced architectures.l’2’3’4 ATRP is performed under mild conditions, and is applicable to a wide range of monomers and solvents. Recently, the copper-catalyzed Huisgen 1,3-dipolar cycloaddition, one of several chemistries termed as “click reactions”, has drawn attention due to their high efficiency, tolerance to various . . . . . . . . 5,6,7 . . . functional groups, technical srmplrcrty and high specrficrty. Combrnlng click reactions with ATRP is a versatile strategy for the synthesis of highly functionalized polymers.8 Generally, polymer functionalization strategies are limited to end-modified and side-modified polymers. In the first approach, an ATRP initiator containing an azide or alkyne moiety is prepared and used to mediate the polymerization of various monomers. The resulting polymers contain terminal alkynyl or azido functionalities, which can be used in click reactions with functional azides or alkynes, respectively. Using this approach, Tsarevsky and Haddleton et al. separately synthesized functional telechelic polymersg’10 In another approach, a polymer with pendant alkyne or azido groups are first synthesized by ATRP, and subsequently functionalized via click reactions. Recently, Matyjaszewski et al. reported the modification of poly(3-azid0pr0pyl 140 1 . . . . . . l methacrylate) 1 (Side-chain functionalization) and linear polystyrene 2 (end functionalization ) by click chemistry. Surface modification with synthetic polymers, especially polymer brushes, is of . . . . . . . . . 13 great interest for 1ndustr1al and biological applications as well as in academic research. Strategies have been developed for the functionalization of surfaces with polymer 14,15,16,17 brushes to immobilize biomolecules, and to realize “smart surfaces” with . . . 8,1 responsrve and adaptive pr0pert1es.1 9 The recent development of ATRP has proved to be a versatile tool for modifying surfaces with a variety of functional polymers with grafted polymer chains ranging from low to high graft densities. Surface-based click 20,21,22,23 chemistry is attracting increasing attention, especially for attaching biomolecules. Bein and coworkers presented a novel approach for the covalent modification of mesoporous silica surfaces (modified with azide monolayer) with intact enzymes (derivatized with an alkyne) via click chemistry.24 They reported a high density of covalently bound enzyme while retaining enzyme activity and the absence of leaching. Recently, Choi and coworkers demonstrated that the click chemistry could be used to modify the termini of polymeric nanobrushes (synthesized via ATRP) by replacing the bromine at the end of the polymer chain with azide groups. To expand the versatility of surface-initiated ATRP we report post polymerization modification of polymer brushes by combining surface-initiated ATRP and click reactions. Azidopropyl methacrylate (AZPMA) was polymerized from initiators anchored on gold and indium tin oxide substrates, and via post-polymerization functionalization, dyes and water soluble polymers were attached to the pendent azides via click chemistry. 141 The AZPMA homopolymer has a high density of azides, and modification of the homopolymer may create a sterically congested system that reduces the efficiency of the click reaction. More importantly, to be suitable for click chemistry in the aqueous environments required for the surface attachment of water soluble molecules (e.g. biological molecules) AZPMA must be copolymerized with hydrophilic monomers. Random copolymer brushes are especially attractive since they should exhibit a uniform distribution of functional groups as well uniform swellingzs’26 Moreover, if both of the comonomers are functional, then polymer brushes can be created with orthogonal functionalities. In this chapter, we describe the ATRP of AZPMA (hydrophobic functional monomer) with ethylene glycol methyl ether methacrylate (EGMA, hydrophilic) and polyethylene glycol methyl ether methacrylate (PEGMA). The lengths of the ethylene oxide chain in EGMA and PEGMA are different, which permits tuning of the hydrophilicity of the AZPMA copolymers. Click chemistry was used to modify poly(AZPMA) and AZPMA copolymers with an alkynylated organic dye and a water soluble polymer, polyethylene glycol monomethyl ether. 5.2. Experimental Section 5.2.1. Materials Unless otherwise noted, all chemicals were obtained from Aldrich. Fluorescein, polyethylene glycol monomethyl ether (mPEG, Mn = 5000 g/mol), sodium hydride, propargyl bromide (80% solution in toluene), ll-mercapto-l-undecanol (MUD, 97%), 2- bromopropionyl bromide (2-BPB, 97%), anisole (99.7%), N,N-dimethylformamide (DMF, 99.8%), Cu(I)Br (99.999%), Cu(II)Br2 (99.999%), Me4Cyclam (99%) 4,4’- 142 dinonyl-2,2’-bipyridyl (anbpy, 97%) and pentarnethyldiethylene triamine (PMDETA, 97%) were used as received. 2,2’-Bipyridine (bpy, 99%) was recrystallized from hexanes and sublimed prior to use. Triethylamine was distilled from calcium hydride under an argon atmosphere at reduced pressure. EGMA (ethylene glycol methyl ether methacrylate, 99%) and PEGMA (poly(oligoethylene glycol methyl ether methacrylate), MW ~300, 98%) were passed through a 10 cm column of basic alumina to remove inhibitors. Azidopropyl methacrylate (AZPMA) was synthesized using the two step procedure reported in the literature.11 After purification, monomers and solvents were transferred to Schlenk flasks, de-gassed using three freeze-pump-thaw cycles and then transferred into a drybox. Alkynylated mPEG (Mw ~5000) was synthesis by a published procedure.” The alkynylated fluorescein methyl ester was synthesized over two steps starting from . . . 28 . . . . . . . fluorescein usrng the method of HVilsted et al. The process of immobilizmg initiators on gold and ITO substrates was described in Chapter 4. 5.2.2. Homo and Copolymerization of AZPMA, EGMA and PEGMA from initiators immobilized on An and ITO substrates The procedure for the polymerization of AZPMA is described. In a Nz-filled drybox, CuBr (6 mg, 0.04 mmol), CuBrz (5 mg, 0.02 mmol), Me4Cyclam (10 mg, 0.04 mmol), and anbpy (16 mg, 0.04 mmol) were added to a round bottom flask containing a 20 mL solution of monomer in DMF/anisole (AZPMA/DMF/anisole = 2:1:1 v:v:v, [AZPMA] = 4.0 M). The well-stirred mixture was heated in an oil bath at 50 0C until a 143 transparent light green solution formed. The solution was then transferred into small vials containing initiator-modified Au or ITO substrates to start the surface-initiated ol erization. After a set reaction time at 50 0C, the substrate was removed from the P ym vial, washed sequentially with ethyl acetate and THF, and then dried under a flow of N2 in a drybox. The same conditions were used for homo polymerization of EGMA ([EGMA] = 4 M). The same procedure was used for homo and copolymerizations of the monomers using copper catalysts with PMDETA and bpy. In all cases, the ratio of [monomer]:[Cu(I)]:[Cu(II)]:[ligand] was 300:1:0.l:1.1 and the monomer:solvent was 1:1 (v/v). For copolyrnerizations, the total monomer concentration was 4 M, and the mole ratio of the co monomers was varied. 5.2.3. Click functionalization of homo and copolymer brushes Click chemistry was performed in a drybox by placing gold (or ITO) substrates coated with AZPMA homopolymer or copolymer brushes in a DMF solution (5 mL) of alkynylated fluorescein (97 mg, 0.25 mmol), CuBr (9 mg, 0.0625 mmol) and PMDETA (l3 uL, 63 umol) at 50 c)C. After a set reaction time, the substrate was thoroughly washed with DMF and THF, sonicated for l min in THF, and again rinsed with THF to remove unreacted dye. The rinsed films were then dried in a stream of nitrogen. For modification of polymer brushes by aqueous click chemistry, substrates were transferred inside a glove bag filled with N2 and placed in degassed solutions (5 mL) of alkynylated mPEG (0.252 g), CuBr (1.8 mg), and bpy (4 mg) in Milli-Q water (18.2 M!) cm). The reaction was run for 12 hours at room temperature, and then the substrates were rinsed with deionized water and dried in a stream of N2. 144 5.2.4. Characterization Methods 13 . . 1H and C NMR analyses were carried out at room temperature on a Varian UnityPlus—SOO spectrometer at 500 and 300 MHz, respectively, with the chemical shifts reported in ppm and referenced to signals from residual protons in the solvent. Film thicknesses were measured using a rotating analyzer ellipsometer (model M-44; J. A. Woollam) at an incident angle of 75°. The data were analyzed using WVASE32 software, and thickness and refractive index determinations were performed on at least three spots on each substrate. The refractive index of the films was assumed to be 1.5 and then fitted with the film thickness. Reflectance FTIR spectroscopy was performed using a Nicolet Magma—IR 560 spectrometer containing a PIKE grazing angle (80°) attachment. UV-vis measurements were taken on Perkin Elmer lambda 400 spectrometer. Atomic force microscopy (AFM) images were obtained in tapping mode with Multimode AFM and NanoScope IV software (Digital Instruments, Santa Barbara, CA) at room temperature. A tapping mode probe (NSCIS) with a nominal frequency of 300 kHz was used for all experiments. The confocal system was the Olympus Flroiew FVlOOO Laser Scanning Confocal Microscope (Tokyo, Japan). The images were collected using the 20>< UPlanFLN (NAO.5) objectives with a 3x optical zoom. The fluorescence was excited using the 488 nm laser line of the Ar gas laser. The emission was collected using a 535- 565 nm band-pass filter. 145 5.3. Results and Discussion 5.3.1 Synthesis of uniform ponAZPMA brushes from gold surfaces Scheme 5.1 shows the synthetic route to polymethacrylate copolymer brushes with pendent alkynes and their modification by click chemistry. Initially, we examined the kinetics for the surface initiated homopolymerization of azidopropyl methacrylate (AZPMA) using several catalyst systems (Figure 5.1). Usually, well-controlled ATRPs maintain a low concentration of active radicals to minimize termination, indicated by a linear relationship between film thickness and time, and provide control over the molecular mass and polydispersities. Typically, the rate of ATRP is low, but the use of Me4Cyclam/anbpy as ligands for Cu catalyst systems yields unusually rapid film growth and high film thicknesses. However, Figure 5.1 shows a decline in film growth rate with time for this system, and suggests significant termination as the consequence of a relatively high radical concentration that also leads to rapid polymerization. Compared to the Me4Cyclam/anbpy catalyst systems, using PMDET A as the Cu ligand provides more controlled polymerizations, as evidenced by a nearly linear increase in thickness with time for the first 4 hours of polymerization (after an rapid rate following initiation). At longer times, the polymerization rate slows and the film thickness reaches 260-300 nm after 12 hours (data not shown here). In prior research, solution polymerization of AZPMA using PMDETA/CuCl and bpy/CuBr provided significant control over the molecular weight distribution and retention of the azide functionality.1 1’29 Liu and coworkers found that the PMDETA/CuCl system gave higher monomer conversion and lower polydispersities than catalysts with 2,2'-bipyridyl or 146 Me6TREN as the ligands.29 It has been well-established that the combination of a CuCl catalyst and a bromo-capped initiator results in halogen exchange and increases the relative rate of initiation compared to propagation, and results in a controlled radical ’ polymerization with a high initiating efficiency. 147 0 Au F__S(_’CHZ);?JJVKBr O OCHZCHZCH2N3 CuCl, PMDETA AZPMA DMF, 55 °c O(CH2CH20)BCHZCHZOCH3 p = 0: EGMA p ~ 4: PEGMA V o H39 H39 (— CHQCHCHz-C‘>Br Au r-S CH 0 1-m {hi 0 0 m Ofl 0 o 0 O A A O A I I I l O 1 2 3 4 5 Time (hours) Figure 5.]. Evolution of the ellipsometric brush thickness with time for the polymerization of AZPMA from initiator monolayers on Au substrates using CuBr/Me4Cyclam/anbpy)/DMF at 50 °C (+), CuCl/PMDETA/DMF at 50 °C (0), CuCl/PMDETA/CuBrz/DMF at 50 0C (O), CuCl/bpy/CuBrz/isopropanol at 50 0C (A) and CuCl/bpy/CuBrz/isopropanol at RT (A). Here, each point represents a different film. 149 However, surface polymerizations of AZPMA using bpy as the catalyst ligand yielded ~lOO nm thick polyAZPMA films after 4 hrs of polymerization at 50 C’C, compared to ~200 nm thick brushes obtained under similar conditions using PMDETA as the ligand. As proposed by Matyjaszewski, 32 multidentate ligands such as Me4Cyclam and PMDETA may complex Cu(II) species more efficiently than bpy, shifting the equilibrium toward the Cu(II) complex and providing a higher radical concentration and faster polymerizations than those catalyzed by Cu-bpy complexes. Previous studies demonstrated that Me4Cyclam and PMDETA Cu complexes are . . . 33,34 . . highly active catalysts for solution ATRP. To control the polymerizations, we added 10% CuBr; to the CuCl/PMDETA catalyst system and observed sluggish, but controlled growth of thick polyAZPMA brushes. As shown in Figure 5.1, the CuCU/PMDETA catalyst system was superior to the other catalyst systems that we tested, and it provides reasonable growth rates with minimal termination. This system was used exclusively in subsequent experiments. The successful growth of polyAZPMA brushes was apparent in the reflectance FTIR spectra of the film by the appearance of strong carbonyl and azide bands at 1740 and 2150 cm-1, respectively (Figure 5.2). The growth in film thickness matches the changes in the intensity of the IR bands. In addition, a topographical AF M (height) image of a 250 nm thick polyAZPMA brush was smooth and uniform with an nns roughness < 2 nm (< 1% of the film thickness). 150 0.1 Absorbance :5, l l l I 3500 3000 2500 2000 1500 1000 Wavenumbers (cm'1) Figure 5.2. Reflectance FTIR spectra of gold substrates coated with polyAZPMA brushes grown for (a) 0.5 h (b) 4 h (c) 6 h and (d) 8 h from the initiator layer. A UV/O3 cleaned gold slide was used as a background. 151 Figure 5.3: Topographical AFM image (tapping mode) of a gold surface coated with a 250 nm polyAZPMA brush. The image is a 5 X 5 um survey scan with high surface coverage and an rms (root mean square) roughness of 1.9 nm, as calculated by the Nanoscope IV software. 152 5.3.2 Copolymerization of AZPMA with EGMA and PEGMA from gold surfaces The conditions used for the homopolymerization of AZPMA were applied to the homopolymerization of EGMA from gold surfaces. Kinetic studies showed that the CuCl/PMDETA system also provided good control over EGMA polymerizations; the CuBr/Me4Cyclam/anpr system gave higher growth rates but with less control (Figure 5.4). With good control and comparable polymerization rates for AZPMA and EGMA, the CuBr/Me4Cyclam/anpr system was used for the copolymerization studies. Copolyrnerizations were run at various comonomer feed ratios and characterized by ellipsometry and reflectance FTIR (Figures 5 .5 and 5 .6 respectively). The copolymer composition was calculated from the reflectance FTIR data by integrating the characteristic azide and carbonyl peaks from AZPMA and the carbonyl peak from EGMA. The copolymer composition correlates well with the initial feed ratios of the comonomers (Figure 5.7), consistent with a random copolymer brush. Moreover, copolymers and homopolymer films synthesized at various feed ratios and characterized at identical polymerization times had similar ellipsometric thicknesses, suggesting that the reactivities of AZPMA and EGMA are comparable and consistent with a random . . 25,26 copolymerization. 153 350 300 r- i] A 250 r l E i 5 a, 200 - * ¥ 0) 3 g 150 L D '3 r? i I 100 t“ { E i 50 § i O I I I J¥ 0 1 2 3 4 5 Time (hours) Figure 5.4. The evolution of the ellipsometric brush thickness with time for the homopolymerization of AZPMA and EGMA from initiator monolayers on Au substrates using CuBr/Me4Cyclam/anbpy/DMF as the catalyst at 50 oC (IAZPMA, DEGMA) and CuC /PMDETA/DMF at 50 °C (OAZPMA, OEGMA). Here, each point represents the average of three independent runs, and the error bars correspond to the standard deviation. 154 250 8 200 *- A 0 I A ’5‘ D 5150 ~ ' a : _g 100 - B .c I- o o o 50 E o 0 l I I i O 1 2 3 4 5 Time(hours) Figure 5.5. Evolution of the ellipsometric brush thickness with time for polyAZPMA and poly(AZPMA-co-EGMA) at various initial feed ratios from initiator monolayers immobilized on gold substrates using CuCl/PMDETA/DMF as the catalyst system at 50 °C. OAZPMA/EGMA (100/0), AAZPMA/EGMA (75/25), IAZPMA/EGMA (SO/50), DAZPMA/EGMA (25/75), OAZPMA/EGMA (0/100). Here, each point represents a different film. 155 >6) 3500 3000 2500 2000 1500 1000 Wavenumbers (cm'1) Figure 5.6. Reflectance FTIR spectra of polyAZPMA and poly(AZPMA-co-EGMA) brushes grown from gold substrates at different feed ratios. (a) AZPMA/EGMA (0/ 100) (b) AZPMA/EGMA (75/25) (c) AZPMA/EGMA (SO/50) (d) AZPMA/EGMA (25/75) (e) AZPMA/EGMA (100/0). 156 y = 0.97x + 0.03 R2 = 0.99 9.0.0.0999 w‘hU'ICDVGDCD—k .09 6N L _J I L 0 0.2 0.4 0.6 0.8 1 Mole fraction of Azpma in copolymer 0 Mole fraction of Azpma in feed Figure 5.7. FTIR analysis of the copolymer data in Figure 5.6. The ratio of the comonomers in the copolymer was determined by integrating the azide and carbonyl peaks and then normalized by the carbonyl peak area. 157 We also copolymerized AZPMA and PEGMA. The ethylene oxide side chain of PEGMA (average molecular weight ~ 300) is significantly longer compared to EGMA, and copolymer brushes with PEGMA are more hydrophilic. Because of the large difference in side chain length in AZPMA and PEGMA, we carried out a similar copolymerization study to assess the randomness of copolymerizations. We first used reflectance FTIR to analyze several gold surfaces coated with poly(AZPMA-co-PEGMA) ([AZPMA] = [PEGMA] in the feed) corresponding to different polymerization times from 5 min to 4 h. Figure 5.8 shows that the mole fraction of AZPMA, determined from the peak ratios of the azide and carbonyl bands, was 0.50 i 0.02 for samples from 90 nm to 250 nm. We then synthesized poly(AZPMA-co-PEGMA) brushes at different feed ratios and characterized them by ellipsometry and FTIR (Figures 5.9 and 5.10). Figure 5.9 shows that the polymerization profiles of PEGMA and AZPMA are similar, but copolymerizations were faster than homopolyrnerizations. This behavior suggests that copolymerization favors alternation, i.e. the growing chains preferentially add the comonomer. However, analysis of the FTIR data for the copolymer brushes (Figure 5.11) indicates that the copolymer compositions match the feed ratios, as expected for a random copolymerization. Returning to the copolymerization data for poly(EGMA-co- AZPMA) in Figure 5.5, there may be similar rate enhancements for copolymerizations, but not as significant as poly(PEGMA-co-AZPMA). However, the IR data indicate that irrespective of their side chain length, EGMA and PEGMA form random copolymer brushes with AZPMA. 158 0.2 Absorbance > E L I I _l 3500 3000 2500 2000 1500 1000 Wavenumbers (cm'1) Figure 5.8. Reflectance FTIR spectra of gold substrates coated with poly(AZPMA-co- PEGMA) brushes at 50 0C using an equimolar feed ratio at different polymerization times (a) 5 min (b) 30 min (0) 1 h (d) 2 h and (e) 4 h. The polymerization system was CuCI/PMDETA/DMF. 159 300 250 e 9 A A I ’5‘ 200 e A x I: V X 3’» o g 150 - D g A O O I— 100 a 0 Cl 9 50 2) O O O L 4 _L 4 o 1 2 3 4 5 Time (hours) Figure 5.9. Evolution of the ellipsometric thickness with time at 50 0C for polyAZPMA and poly(AZPMA-co-PEGMA) brushes with time grown from initiators immobilized on gold substrates at various initial feed ratios, using the CuCl/PMDETA/DMF system. OAZPMA/PEGMA (100/0), AAZPMA/PEGMA (75/25), xAZPMA/PEGMA (SO/50), DAZPMA/PEGMA (25/75), OAZPMA/PEGMA (0/100). Each point represents a different film. 160 0.2 Absorbance > 9’: _L 4 l A 3500 3000 2500 2000 1500 1000 Wavenumbers (cm'1) Figure 5.10. Reflectance FTIR spectra of gold substrates coated with polyAZPMA, poly(AZPMA-co-PEGMA) and polyPEGMA brushes grown at different feed ratios. (a) AZPMA/PEGMA (lOO/O) (b) AZPMA/PEGMA (75/25) (c) AZPMA/PEGMA (SO/50) (d) AZPMA/PEGMA (25/75) (e) AZPMA/PEGMA (0/1 00). The polymerization system was CuCl/PMDETA/DMF at 50 °C. 161 0-8 F y = 0.94x + 0.03 R2 = 0.99 Mole fraction of Azpma in copolymer O ‘7 I i I _I_ O 0.2 0.4 0.6 0.8 1 Mole fraction of Azpma in feed Figure 5.11. FTIR analysis of the copolymer brush data of Figure 5.10. The copolymer composition was determined by integration of the azide and carbonyl peaks and normalized by the carbonyl peak area. 162 5.3.3. Derivatization of copolymer brushes with dyes via click chemistry Surface grafted polyAZPMA and AZPMA copolymers are excellent substrates for the synthesis of modified polymer brushes via click chemistry. The number of reactive azide groups can be controlled by the c0polymer composition, which is simplified by the apparent random incorporation of PEGMA and EGMA comonomers into the brush. C0polymers are particularly useful since they reduce the high local concentration of azides within the polymer film which may reduce the efficiency of click functionalization compared to reactions in dilute solutions of polyAZPMA. 11 To improve click efficiency, increase polymer hydrophilicity, and reduce the functional group density, we synthesized copolymer brushes with the hydrophilic comonomers, PEGMA and EGMA. To assess the click reaction rate for grafted copolymer brushes, we appended a fluorescent dye, modified with a terminal alkyne, to AZPMA homopolymers and c0polymers in DMF using CuBr/PMDETA as the catalyst at 50 OC. Both gold and ITO surfaces were used as substrates; gold for analysis by reflectance FTIR, and ITO for UV-vis spectrosc0py. Figure 5.12 shows representative UV-vis spectra of ~200 nm thick poly(AZPMA-co- PEGMA) (SO/50) brushes grown from ITO surfaces for different reaction times. The click reaction was almost complete within 5 min, as the absorbance marginally increased from 5 min to 1 hour. To quantify the click reaction kinetics, we used reflectance FTIR to monitor the disappearance of azide peak at ~ 2100 cm-1 and the appearance of aromatic C=C . -l . . . stretching at ~16OO cm during the first hour span of the reaction, where desorption observed was negligible. The kinetic study (Figure 5.13 and Figure 5.14) confirmed that 163 the click reaction is fast, with > 60% of the azides reacted within 1 min. The rate was nearly independent of the brush architecture, with copolymers having fewer azides reacting slightly faster than polyAZPMA. There are several reasons for the rapid click reaction: (1) the local concentration of azide groups in a brush is much higher than soluble polyAZPMA, as observed by Li et al. in the case of azide grafted on silica nanoparticles;35 (2) while the azide is immobilized on a polymer brush, the dye molecule is small and readily diffuses through the brush; (3) the azide group forms triazoles, which bind and stabilize Cu(I)36’37and hence enhances the local concentration of copper catalyst; and (4) CuBr/PMDETA is a highly efficient click catalyst.11 Interestingly, the click reaction never went to completion when the substrate was polyAZPMA brushes or copolymer brushes with >50% AZPMA. This may reflect reduced diffusion of the alkynes through the brush architecture, especially near the substrate surface. The effect seems to be limited to azides near the surface, since the reaction rates are similar for homopolymers, and copolymers with varying ethylene oxide chain lengths. The rapid click reaction may be assisted by the DMF solvent, since DMF at 50 0C is a good solvent for homopolymer and copolymer brushes, enabling the dye molecule and catalysts to diffuse through the brush. An important observation is that prolonged soaking the brush-coated substrate in the click solution results in brush desorption. After 24 h of reaction, the dye absorbance decreased instead of increasing. A similar phenomenon was observed via reflectance FTIR spectroscopy where the intensities of the characteristics peaks of thick films (>150 nm) decreased over time. 164 Absorbance 0.2 ~ 0.1 L O J L L I 300 400 500 600 700 800 Wavelength (nm) Figure 5.12. UV-vis spectra of ITO substrates coated with ~l80 rim of poly(AZPMA- co-PEGMA) (50/50) showing the evolution of the click reaction between the copolymer and a fluorescent dye . Blue line: 5 min, pink line: 1 h, red line: 4 h and green line: 24 h. 165 0.05 I i [I l Absorbance l l g E E g l I _l L I 3500 3000 2500 20001 1 500 1 000 Wavenumbers (cm‘ ) Figure 5.13. Reflectance FTIR spectra of gold substrates coated with ~200 nm copolymer brushes of poly(AZPMA-co-PEGMA) (SO/50) showing the evolution of the click reaction between the copolymer and a fluorescent dye. (a) 1 h (b) 30 min (0) 15 min (d) 5 min (e) l min (f) 0 min. 166 100 CI 3} [:1 Di * * ‘ O X 80 L 2 8 g ifi R A A X 0 23 50 - C .9 2 °>’ O U 20 r O 1 I_I I I l i l I 1 I l I I O 15 30 45 60 75 Time (min) Figure 5.14. Kinetics of the click reaction between a fluorescent dye and ~200 d: 30 nm films of poly(AZPMA), poly(AZPMA-co-EGMA) and poly(AZPMA-co-PEGMA) grown from gold substrates. The data were extracted from reflectance FTIR spectra analogous to those shown in Figure 5.13. <>AZPMA/EGMA (lOO/O), — AZPMA/EGMA (75/25), OAZPMA/EGMA (SO/50), DAZPMA/EGMA (25/75), xAZPMA/PEGMA (75/25), AAZPMA/PEGMA (SO/50), +AZPMA/PEGMA (25/75). Here, each point represents a different film. 167 We used the absorbance values at Amax to quantify the covalently bound dye in the copolymer brushes. Alternative methods such as measuring the change in molecular weight or using ellipsometry to monitor changes in film thickness after the click reaction were less precise. UV-vis spectra of the dye were obtained in air and in various solvents, and since we found no changes in the spectra, we assumed that the molar extinction coefficients for the dye in solution and on surfaces is identical. Rearranging the Lambert- Beer’s law, A = a b c; where A = absorbance, a = molar extinction coefficient, b = path length, and c = concentration of the solution to A = (c/a) a b c; where, a = area of the solution exposed to the light, provides A = s F; where F = the grafting density, (moles/area). To calculate the grafiing density (F), we need to know the surface absorbance at Amax and molar extinction coefficient of the dye. The molar extinction coefficient of the dye (a = 52227 cmZ/mol) was calculated from the absorbance values at kmax in THF at 4 different concentrations (Figure 5.15). We estimated the dye molecule grafting density (F) from c and the data from Figures 5.16, 5.17 and 5.18. The dye absorbance tracks the azide content in poly(AZPMA-co- PEGMA), until it reaches its maxima for the homopolymer (100% azide). Figure 5.19 shows fluorescence micrOSCOpy images from ITO coated with ~180 nm of poly(AZPMA-co-EGMA) (1:1) after a 5 min click reaction with the alkynylated dye, and a control surface, treated identically but without the copper catalyst. The fluorescence of the clicked surface confirms that the dye has been bound to the surface, while the lack of fluorescence for the control surface confirms that physical adsorption of . . 28 the fluorescent reactant can be disregarded, as observed by Hvrlsted and coworkers. 168 9.0.0 7‘? A 03 oo _. m 4:- Absorbance at peak maximum O to 0 0.0E+00 5.0E-06 1.0E—05 1.5E-05 2.0E-05 2.5E-05 Concentration (mol/L) Figure 5.15. Measurement of the molar extinction coefficient of the alkyne modified fluorescein dye. 169 0.1 f Absorbance I 4 4 J I l 1 I l 300 400 500 600 700 800 Wavelength (nm) Figure 5.16. UV-vis spectra of ~ 200 i 30 nm thick polyAZPMA and poly(AZMPA-co- PEGMA), and polyPEGMA brushes grown from ITO substrates after a 5 min click reaction with a alkynylated fluorescent dye. Green line: before clicking to AZPMA (100/0), blue line: after clicking to AZPMA (25/75), magenta line: after clicking to AZPMA/PEGMA (SO/50), black line: after clicking to AZPMA/ PEGMA (75/25), red line: after clicking to AZPMA/ PEGMA (100/0). 170 .0 .0 .0 h 0) 00 l I l O Absorbance at peak maxima O to f I I I 25 50 75 100 O % of Azpma in the copolymer brush Figure 5.17. Absorbance at kmax (463 nm) obtained from the UV-vis spectra in Figure 5.15. 171 Bound dye molecule (pg/cmz) O L I I I O 25 50 75 100 °/o of Azpma in the copolymer brush Figure 5.18. Dye molecule binding as a function of the per cent AZPMA in a 200 i 30 nm poly(AZMPA-co-PEGMA) brushes on ITO surface. The dye molecule binding was determined by dividing the UV-vis spectral peak maxima of the different copolymer brushes (Figure 5.17) with the molar extinction coefficient obtained from Figure 5 .15. 172 20 um 20 um Figure 5.19. Representative fluorescence microscopy images of the clicked surface (left image) and the control (right), prepared without the copper catalyst but under otherwise equivalent conditions. The images were recorded using equal lighting and camera settings. 5.3.4 Derivatization of copolymer brushes via click chemistry with a water soluble polymer The previous data show that dye molecules in DMF show comparable reactivity in click reactions with homopolymer and copolymer substrates. However, tethering biomolecules to brushes requires an aqueous environment, and therefore we synthesized a water soluble copolymer brush to test whether the click reaction would be successful in aqueous environments. For this experiment. alkynylated mPEG 5000, a water soluble polymer, was used for the click reactions. Figure 5.20 shows reflectance F TIR spectra for ~50 nm thick films of polyAZPMA and poly(AZPMA-co-PEGMA) grown from gold surfaces, before and after click reactions with an alkynylated mPEG 5000 in water at room temperature for 12 hours. Compared to click reactions with the alkynylated dye, the reaction of the alkynylated mPEG polymer was slow. likely due to slower diffusion of the higher molecular weight of the polymer (5000 g/mol). 173 Absorbance LED, I I L I 3500 3000 2500 2000 1500 1000 Wavenumbers (cm'1) Figure 5.20. Click reactions for 12 hours with a water soluble polymer (alkynylated mPEG 5000) monitored by reflectance FTIR spectra of gold substrates coated with ~ 50 i 10 nm of poly(AZPMA) and poly(AZPMA-co-PEGMA) brushes (a) after clicking to AZPMA/PEGMA (100/0), (b) before clicking to AZPMA/PEGMA (100/0), (c) after 174 clicking to AZPMA/PEGMA (75/25), (d) before clicking to AZPMA/PEGMA (75/25), (c) after clicking to AZPMA/PEGMA (50/50), (f) before clicking to AZPMA/PEGMA. (SO/50), (g) after clicking to AZPMA/PEGMA (25/75), (h) before clicking to AZPMA/PEGMA (25/75). Other reasons for the slow click reaction may include (1) the hydrophobicity of azide-containing polymer which could collapse and restrict access to the azide, (2) the lower reaction temperature (ambient temperature vs. 50 0C, and (3) the water soluble catalyst, CuBr/bpy, may be less efficient than the CuBr/PMDETA complex.38 Integrating the azide peaks at ~2100 cm-1 in FTIR spectra before and after the click reactions (Figure 5 .20) reveals that click conversion increases as the PEGMA mole fraction in the polymer brush increases and becomes more hydrophilic (Figure 5.21). In addition, the changes in the intensities of the C-0 stretching bands at 1100 and 1350, and OH stretching at 2900 cm.1 are consistent with the azide peak analysis. Moreover, the thickness of the most hydrophilic brush, poly(AZPMA-co-PEGMA) (AZPMA/PEGMA = 25/75) nearly doubled, from 47 nm to 92 nm, after the click reaction. The thickness of the poly(AZPMA) brush was almost unchanged afier the click reaction (< 10% change) under identical conditions (Figure 5.22). The expansion of clicked poly(AZPMA-co-PEGMA) (25/75) film, and the IR analyses confirm that the click reaction occurs in the bulk film, while poly(AZPMA) films react primarily on the surface. Figures 21 and 22 also show that the click conversion and copolymer compositions, inferred from the change in film thickness, (Figure 5.22), are consistent for all polymers tested. Also, the more hydrophobic 175 poly(AZPMA—co-EGMA) films (shorter ethylene oxide chain, p=0, see Scheme 1) was marginally better than poly(AZPMA) films in terms the water-based click reactions (data not shown here). Therefore, the longer ethylene oxide chains in poly(AZPMA-co- PEGMA) films (p ~4, see Scheme 1) make the polymers sufficiently hydrophilic to effect immobilization of water soluble polymers by the click reaction. 176 70 6C 4 (1‘ A0\0v coumhw>coo 70 O 60 r- A5O - o\° o .540 ‘ 2 o “3’30 r O O 20 r 10 L o O L l l l I 0 25 50 75 100 125 °/o of azide in copolymer Figure 5.21. The disappearance of azide the via click reaction with alkynylated mPEG 5000 for 12 hours at room temperature as a function of the molar per cent azide in the COPOIymer brush for different feed ratios (see Figure 5.20). The conversion was calculated from FTIR spectra by integrating the azide peak before and after the click reaction. 177 100 90 . 8O . " 70 - 60 ~ . 50 - % increase in thickness 10L I O i L I l l 0 25 50 75 100 125 % of azide in copolymer Figure 5.22. Increase in thickness after the click reaction with alkynylated mPEG 5000 for 12 hours at room temperature as a function of the molar per cent azide in the copolymer brush. 178 Interestingly, FTIR spectroscopy detected significant chain desorption while immobilizing alkynylated mPEG onto thick c0polymer brushes (> 150 nm), and the click reaction was reasonably fast (significant conversion within 30 min). However, no desorption (no change of carbonyl peak height) occurred for thin films (~ 50 nm) even after 12 h of reaction. The severe desorption for poly(AZPMA—co-PEGMA) brushes, may be due to the long water-soluble side chains39 and extreme steric crowding of the thick film. However, this needs to be investigated in detail. 5.4. Conclusions Azidopropyl methacrylate (AZPMA) a functional monomer with a pendent clickable moiety, was grown from gold and ITO surfaces using surface-initiated ATRP. AZPMA enables post-polymerization modification of polymer brushes via the 1,3-dipolar cycloaddition of azides with acetylenes, and copolymerization with PEGMA and EGMA, provides control over the hydrophilicity and fimctional density of AZPMA polymers. Post-polymerization modification of homo and copolymer brushes was demonstrated by reacting with an alkyne-modified dye via click reactions. Kinetic studies showed that click reactions on surface-grafted homo/copolymer brushes were fast when run in DMF at 50 0C (> 60% conversion in 1 min) irrespective of the copolymer composition. The outcome of aqueous click reactions depends on the composition of the polymer brush. Using a alkynylated polyethylene glycol methyl ether (mPEG, Mn ~5000), the most hydrophilic c0polymers, i.e. those with the highest fraction of PEGMA, immobilized the most mPEG, as evidence by the increase of the dry film thickness. Aqueous click reactions offer an alternative to ester or amide formation for the covalent immobilization of biomacromolecules. 179 References: 1 Wang, J. S.; Matyjaszewski, K. J. Am. Chem. Soc. 1995, 117, 5614. 2 Xia, J.; Gaynor, S. G.; Matyjaszewski, K. Macromolecules 1998, 31 , 5958. 3 Matyjaszewski, K.; Xia, J. Chem. Rev. 2001, 101, 2921. 4 Kamigaito, M.; Ando, T.; Sawamoto, M. Chem. Rev. 2001, 101, 3689. 5 Rostovtsev, V. V.; Green, L. G.; Fokin, V. V.; Sharpless, K. B. Angew. Chem. Int. Ed. 2002, 41, 2596. 6 Tomoe, C. W.; Christensen, C.; Meldal, M. J. Org. Chem. 2002, 67, 3057. 7 Kolb, H. C.; Finn, M. G.; Sharpless, K. B. Angew. Chem, Int. Ed. 2001, 40, 2004. 8 Golas,P. L.; Matyjaszewski, K. QSAR Comb. Sci. 2007, 26, 1116. 9 Tsarevsky, N. V.; Sumerlin, B. S.; Matyjaszewski, K. Macromolecules 2005, 38, 3558. 10 Mantovani, G.; Ladmiral, V.; Tao, L.; Haddleton, D. M. Chem. Commun. 2005, 2089. ll Sumerlin, B. S.; Tsarevsky, N. V.; Louche, G.; Lee, R. Y.; Matyjaszewski, K. Macromolecules 2005, 38, 7540. 12 Gao, H.; Louche, G.; Sumerlin, B. S.; Jahed, N.; Golas, P.; Matyjaszewski, K. Macromolecules 2005, 38, 8979. 13 Advincula, R. C.; Brittain, W. J.; Caster, K. C.; Ruhe, J. Eds.; Polymer Brushes; Wiley-VCH Verlag: Weinheim, 2004. 14 Cullen, S. P.; Liu, X.; Mandel, I. C.; Himpsel, F. J.; Gopalan, P. Langmuir 2008, 24, 913. 15 Dai, J.; Bao, 2.; Sun, L.; Hong, S. U.; Baker, G. L.; Bruening, M. L. Langmuir 2006, 22, 4274. 16 Tsarevsky, N. V.; Braunecker, W. A.; Vacca, A.; Gans, P.; Matyjaszewski, K. Macromol. Symp. 2007, 248, 60. 17 Jain, P.; Dai, J .; Baker, G. L.; Bruening, M. L. Macromolecules 2008, 41 , 8413. 18 Riihe, J.; Ballauff, M.; Biesalski, M.; Dziezok, P.; Grdhn, F.; Johannsmann, D.; Houbenov, N.; Hugenberg, N.; Konradi, R.; Minko, S.; Motomov, M.; Netz, R.; Schmidt, 180 M.; Seidel, C.; Stamm, M.; Stephan, T.; Usov, 0.; Zhang, H. Adv. Polym. Sci. 2004, 165, 79, and references therein. 19 Minko, S. Ed.; Responsive Polymer Materials: Design and Applications; Blackwell Publishing: Boston, 2006. 20 Lummerstorfer, T.; Hoffmann, H. J. Phys. Chem. B 2004, 108, 3963. 21 Gallant, N. D.; Lavery, K. A.; Amis, E. J .; Becker, M. L. Adv. Mater. 2007, 19, 965. 22 Kacprzak, K. M.; Maier, N. M.; Lindner, W. Tetrahedron Lett. 2006, 47, 8721. 23 Prakash, 3.; Long, T. M.; Selby, J. C.; Moore, J. 8.; Shannon, M. A. Anal. Chem. 2007, 79, 1661. 24 Schlossbauer, A.; Schaffert, D.; Kecht, J.; Wagner, E.; Bein, T. J. Am. Chem. Soc. 2008, 130, 12558. 25 Sanjuan, S.; Tran, Y. J. Polym. Sci, Polym. Chem. 2008, 46, 4305. 26 Sanjuan, S.; Tran, Y. Macromolecules 2008, 41, 8721. 27 Schoffelen, 8.; Mark, H.; Lambermon, L.; Mark, B.; Eldijk, V.; Jan, C. M.; Hest, V. Biocon. Chem. 2008, 19, 1127. 28 Daugaard, A. E.; Hvilsted, S.; Hansen, T. S.; Larsen, N. B. Macromolecules 2008, 41, 4321. 29 Zhang, J .; Zhou, Y.; Zhu, 2.; Ge, 2.; Liu, S. Macromolecules 2008, 41, 1444. 30 Karanam, S.; Goossens, H.; Klumperrnan, 13.; Lemstra, P. Macromolecules 2003, 36, 3051. 31 Shipp, D. A.; Wang, J. L.; Matyjaszewski, K. Macromolecules 1998, 31, 8005. 32 Matyjaszewski, K. Curr. Org. Chem. 2002, 6, 67. 33 Pintauer, T.; Matyjaszewski, K. Coord. Chem. Rev. 2005, 249, 1155. 34 Tsarevsky, N. V.; Matyjaszewski, K. J. Polym. Sci, Polym. Chem.2006, 44, 5098. 35 Li, Y.; Benicewicz, B. C. Macromolecules 2008, 41 , 7986. 36 Chan, T. R.; Hilgraf, R.; Sharpless, K. B.; Fokin, V. V. Org. Lett. 2004, 6, 2853. 181 37 Sumerlin, B. S.; Tsarevsky, N. V.; Louche, G.; Lee, R. Y.; Matyjaszewski, K. Macromolecules 2005, 38, 7540. 38 Golas, P. L.; Tsarevsky, N. V.; Sumerlin, B. S.; Matyjaszewski, K. Macromolecules 2006, 39, 6451. 39 Lebedeva, N. V.; Sun, F. C.; Lee, H.; Matyjaszewski, K.; Sheiko, S. S. J. Am. Chem. Soc. 2008, 130, 4228. 182 MICHIGAN STATE UNIVERSITY LIBRARIES lll ll '1 ll ll 3 03063 8658 "3"129