b» D! l a .41. n . c u U I263.S~'.§flm :ipiw. th» in. v. .1 ¢,Q\‘....P£C 11 1. a. .munriu 9: Ar . .. . u»...<.w$w. .3. F. x... n ‘ It. 1: 2:! E 13:31:. : 55.2. v. n: 3.1%. I L 8 O. 4-00.31”! I 7” o. I! .r 4. a .\ Jad‘vduuh It" fits a." a. t « q- (E . .f.»L.fi.m :xr: on“. . .3»? 1ftn¢fi§§ . if e: .21... r. y .514»: .Ic 1:. . .~ 9‘ i 11...}: .1 If!!! 2,. . 2 . 317.... x 9.. v. «flarinqifieuvlat 1. x» . ‘ u ‘14 .unfilIKhtml; .3 .QI p.11 .3 300} This is to certify that the dissertation entitled Poly(ethylene oxide) based polymer brushes: Synthesis, properties and applications presented by Ying Zheng has been accepted towards fulfillment of the requirements for the PhD. degree in Chemistry flgé’ééa l/Major Professor’s Signature ZS 4q9¢,5'/‘ 2.006 Date MSU is an Affirmative Action/Equal Opportunity Institution '—_———— -——- 4%,, —— 7 LIBRARY Michigan State University PLACE IN RETURN Box to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/05 p2/CIRC/Date0ue.indd-p.1 POLY(ETHYLENE OXIDE) BASED POLYMER BRUSHES: SYNTHESIS, PROPERTIES AND APPLICATIONS By Ying Zheng A Dissertation Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2006 ABSTRACT POLY(ETHYLENE OXIDE) BASED POLYMER BRUSHES: SYNTHESIS, PROPERTIES AND APPLICATION By Ying Zheng Surface-initiated atom transfer radical polymerization (ATRP) of poly (ethylene glycol) methyl ether methacrylate (PEGMA) macromonomers from gold substrates leads to thin films of comb-like polymer brushes. Polymerization of these monomers in water resulted in essentially constant polymerization rates and yielded coatings with thicknesses as high as 400 nm. Polymers with side chains consisting of 22-23 ethylene oxide repeating units aligned to form crystalline polymers analogous to their non-brush analogs. However, polarized optical microscopy confirmed their crystallization as two- dimensional spherhulites, and the film morphologies, crystallization rates and lamellae orientations all depended on the film thickness. The crystallization rate decreased with the film thickness, as reflected by the film morphology changes. An Avrami analysis of growth rates indicated heterogeneous nucleation and one-dimensional growth. Both AFM images and reflectance FT-IR spectra of the films were consistent with crystalline lamellae preferentially oriented normal to the surface for brushes with thicknesses above 100 nm. As the film thickness decreased, the favored orientation shifted to crystalline lamellae parallel to the surface, with the PEO side-chains oriented perpendicular to the surface. The lamellae orientations are related to polymer-surface interactions as well as to the “comb-like” polymeric structure. Polymerizations were optimized to grow comb-like polymer brushes on flat surfaces and on silica nanoparticles. Kinetic studies of the growth of polymethacrylates with (PEG) polyethylene glycol segments of various lengths identified optimum polymerization condition for the growth of thick poly(PEGMA) films on planar substrates. These materials have potential utility as membranes, electrolytes for lithium ion batteries, and other applications. Crosslinked brushes were grown from gold substrates by copolymerization of a macromonomer, poly(ethylene glycol) methyl ether methacrylate (PEGMA), and a crosslinker, bisphenol A ethoxylate dimethacrylate (BisA-EDMA) using surface-initiated ATRP. The crosslink density of the film was estimated from FT-IR measurements and from the aqueous swelling of the films, as measured by in-situ ellipsometry. These films show increased dimensional stability with cross-link while keeping high swelling in water. These crosslinked films are attractive for their mechanical stability and their ability to resist protein adsorption in biological applications. Comb-like and crosslinked polymer brushes containing PEO segments were grafted on the surface of porous alumina membrane supports and the perrneabilities of C02 and H2 through the membranes were studied. These films are highly permeable to CO; and selectively permeate CO; with a C02 permeation coefficient as high as 300 Barrer and a COZ/Hz selectivity of 14 respectively for poly(PEGMA) film. The Cog/gas permeability decreased for shorter PEG segments and for highly crosslinked films. To My Famny -iv- a. ‘m’g 1.15 .. ACKNOWLEDGMENTS First, I would like to express my deep appreciation to my advisor, Professor Gregory Baker. Thank you for your guidance and constant encouragement, which built my confidence and enabled me to achieve my Ph.D. at Michigan State University. The polymer science knowledge, strict scientific attitude and initiative in research that I have learned from you will be invaluable in my future career. Professor Babak Borhan not only taught me good spectroscopic techniques and organic chemistry, but he also helped me in the first step of my Ph.D. studies in Baker’s group. I greatly appreciate him for his trust and help. I learned from Professor Merlin Bruening a sincere approach to science and critical insight. I also like to thank Professor Peter Wagner for being my committee member. I would like to give my thanks to all Baker group members: JB, Bao, Ping, Qin, Leslie, F eng, quei, Jon, DJ, Erin and Sampa for useful discussions and for help in my research and my life. I also learned good surface science from Dr. Wenxi, and I enjoyed working with Lei, Tiffany, Matt and Anagi. I would like to thank all my friends for making my life in MSU enjoyable and memorable. Special thanks to Lisa Dillingham for her help during my stay in chemistry department, which made my life easier. Finally, I want to thank my family. They have always been so supportive in every stage of my life. My life would not have been the same without their love and support. My dearest parents and my brother deserve my biggest thanks and love. TABLE OF CONTENTS Page List of Tables ....................................................................................... ix List of Figures ....................................................................................... x List of Schemes .................................................................................... xvii List of Abbreviations ............................................................................. xviii Chapter I. Introduction to polymer brushes ................................................. 1 Synthesis of polymer brushes by surface-initiated polymerization ........................ 4 Surface-initiated living anionic and living cationic polymerizations ............... 5 Surface-initiated ring opening metathesis polymerization (ROMP) ................ 9 Surface-initiated ring opening polymerization (ROP) .............................. 12 Surface-initiated controlled radical polymerization (CRP) ......................... l4 Atom Transfer Radical Polymerization (ATRP) ........................................... 17 ATRP kinetics ............................................................................ 19 ATRP components ....................................................................... 20 ATRP of PEGMA ........................................................................ 3O Polymer brushes formed by surface-initiated ATRP ....................................... 32 Methods for the controlled growth of polymers from surfaces ..................... 32 Controlled ATRP from surfaces of particles .......................................... 36 Physicochemical properties of polymer brushes ........................................... 38 Physical states of polymer brushes .................................................... 38 Stimuli-responsive polymer brushes ................................................... 40 Applications of polymer brushes ............................................................ 47 Nanotechnology—Patteming of polymer brushes ................................... 47 Biological application in protein resistance and cell adhesion ..................... 49 Polymer brushes grafied membranes ................................................... 51 References ....................................................................................... 53 Chapter II. Crystallization of Polymer Brushes ............................................ 61 Introduction ...................................................................................... 61 Experimental Section ........................................................................... 66 Materials .................................................................................. 66 Synthesis of initiator [Br-C(CH3)2-COO(CH2)118);] ............................... 66 Preparation of immobilized initiators on Au surfaces .............................. 67 Surface-initiated polymerization of PEGMA ........................................ 68 -vi- Solution ATRP of PEGMA ............................................................ 68 Characterization methods ............................................................... 69 Crystallization studies ................................................................... 69 Results and discussion ......................................................................... 70 Preparation of polymer brushes ........................................................ 70 Morphology of crystalline polymer brushes .......................................... 75 Spectroscopic characterization ......................................................... 82 Crystallization kinetics of polymer brushes .......................................... 90 Conclusions .................................................................................... 100 References ..................................................................................... 101 Chapter 111. Synthesis of Comb-like Polymer Brushes on Planar and Nanoparticulate Surfaces104 Introduction .................................................................................. 104 Experimental ................................................................................. 106 Results and Discussion ..................................................................... 112 Surface-initiated ATRP of PEGMA on Au ......................................... 112 Rapid grth of comb-like polymer brushes on the surface ..................... 119 Synthesis of PEG-functionalized nanoparticles .................................... 124 Growth of polymer brushes on silica wafers ....................................... 124 Synthesis of comb-like polymers on the surface of fumed silica ................ 127 Summary and Outlook ..................................................................... 132 References .................................................................................... l 33 Chapter IV. Crosslinked Polymer Brushes Containing PEO Segments .............. 135 Introduction .................................................................................. 135 Experimental ................................................................................. 136 Materials ................................................................................. l 36 Characterization methods .............................................................. 136 Synthesis of crosslinked polymer brushes .......................................... 137 Determination of cross-link density of polymer brushes .......................... 138 Results and Discussion ..................................................................... 140 Synthesis of crosslinked polymer brushes .......................................... 140 Characterization of crosslinked polymer fihns ..................................... 143 Swelling behavior of crosslinked copolymer films ................................ 148 Conclusion and Outlook ................................................................... 150 References ................................................................................... 1 5 1 -vii- Chapter V. Polymer Membranes for Gas Separation .................................... 152 Background .................................................................................. 1 52 Experimental ................................................................................. 158 Materials ................................................................................. 1 5 8 Characterization methods .............................................................. 1 5 8 Anchoring initiators on the surface of porous alumina membrane support. . 1 59 Surface-initiated ATRP of . poly(PEGMA) films using CuBr/CuBrszpy catalyst ................................................................................... 1 59 Polymerization of PEGMA using CuBr/Me4Cyclam/CuBr2(anbpy)2 catalyst complexes ..................................................................... 160 Growth of crosslinked copolymer film .............................................. 160 Gas permeation measurements ....................................................... 161 Results and Discussion ..................................................................... 163 Synthesis of and characterization of polymer fihns ................................ 163 Pure gas permeation of poly(PEGMA) brushes .................................... 165 Pure gas permeation through crosslinked polymer films [poly (PEGMA-co- Summary and Outlook ..................................................................... 177 References .................................................................................... 179 - viii - LIST OF TABLES Table Page Table 2.1. Crystalline PEO absorption band assignment ...................................... 84 Table 2.2. Avrami exponent n determined from 2-D spherulite growth monitored by in situ polarized optical microscopy ................................................................. 94 Table 3.1. Reaction conditions for polymerization of PEGMA(22-23) and PEGMA(4-5) from the surface of fumed silica particles ...................................................... 129 Table 3.2. Results of TGA control experiments ............................................... 130 Table 3.3. DSC results of poly(PEGMA) silica particle composites ........................ 131 Table 5.1 Physical properties of H2, C02, and N2 gases ..................................... 154 Table 5.2. Representative examples of pure gas C02 permeability and COz/Hz selectivity in PEG-containing polymer ..................................................................... 157 Table 5.3. Summary of gases permeation through poly(PEGMA) films .................. 171 Table 5.4. Gas permeability coefficients and average selectivities for crosslinked poly(PEGMA-co-BisA-EDMA) membrane ................................................... 176 -ix- LIST OF FIGURES Figure Page Figure 1.1. Polymer brushes on flat and curved surfaces ....................................... 1 Figure 1.2. “Grafting to” (a) and “grafting from” (b) methods for the generation of polymer brushes. Adapted with permission from J. Am. Chem. Soc. 1999, 121, 1016- 1022. Copyright ©1999 American Chemical Society ........................................... 3 Figure 1.3. Examples of polymer brushes grown by living anionic polymerization. . . . . ....7 Figure 1.4. Scheme showing the surface-initiated polymerization of 2—oxazolines for the preparation of the amphiphilic nanocomposites, with 2-ethyl—2-oxazoline and N,N-di-n- octadecylamine used as monomer and terminating agent, respectively. Adapted with permission from Macromolecules, 2001, 34, 1606-1611. Copyright © 2001 American Chemical Society ..................................................................................... 8 Figure 1.5. Polymer brushes grown by living cationic polymerization ........................ 9 Figure 1.6. Surface-initiated ROMP applied to the synthesis of norbornene-based polymer brushes ..................................................................................... 10 Figure 1.7. Graphic representation of ROMP initiated by DPN and a monomer coated AFM tip. Reprinted with permission from Angew. Chem, Int. Ed. 2003, 42, 4785-4789. Copyright © 2003 Wiley Intersciences ........................................................... 11 Figure 1.8. PCL and PLA brushes formed by surface-initiated ROP ........................ 12 Figure 1.9. Synthesis of poly(HEMA)—g-(polylactide) ......................................... 13 Figure 1.10. Synthetic strategy for the anionic ring-opening polymerization of ethylene oxide initiated from a silica surface modified by Al(OiPr)3/GPS ............................ 14 Figure 1.11. Synthesis of OEG containing polymer brushes on the silica surface. . . . . 17 Figure 1.12. Structure of the [CuI (bpy)2]+ cation in [CuI (bpy)2]+[ClO4]’ .................. 26 Figure 1.13. Structure ofthe [Cu’ (bpy) x2] (X=Br, c1) dimer .............................. 26 Figure 1.14. Molecular structure of [CuI(HI\/ITETA)]+[Cu‘C12]_ ............................. 27 Figure 1.15. Structure of the [Cu"(bpy)2Br]+ cation in [Cu"(bpy)2Br]+[Br]- ............... 27 Figure 1.16. Molecular structure of [Cu"(HMTETA)Br]+[Br]’ .............................. 28 Figure 1.17. General synthetic route to polymer brushes on gold and silica substrates... 33 Figure 1.18. Schematic illustration of linear, comb-like and crosslinked polymer brushes ................................................................................................ 35 Figure 1.19. Synthesis of mixed PMMA/PS brushes from an asymmetric di-functional initiator-terminated SAM (Y—SAM) by combining ATRP and NMRP techniques. Reprinted with permission from J. Am. Chem. Soc. 2004, 126, 6124-6134. Copyright © 2004 American Chemical Society ................................................................ 36 Figure 1.20. Poly(PEGMA) and PMEMA brushes grown from silica nanoparticles ...... 37 Figure1.21. Preparation of comb-coil polymer brushes on the surface of silica nanoparticles ......................................................................................... 38 Figure 1.22. Synthetic scheme of side chain LCP brushes by surface-initiated ATRP. . ..39 Figure 1.23. Speculative model for nanopattem formation in di-block copolymers ....... 41 Figure 1.24. Reversible response of triblock copolymer brushes to different solvents. . ..41 Figure 1.25. Tapping mode AFM height images of PDMS-b-PS-b-PDMSA brushes after the following treatments: (left) after immersion in toluene and drying with nitrogen, (right) after immersion in toluene, gradual addition of hexane and drying with nitrogen. Reprinted with permission from Macromol. Chem. Phys. 2004, 205, 411. Copyright © 2004 Wiley Intersciences .......................................................................... 43 Figure 1.26. 1H NMR spectra of PAA/PS particles dispersed in (a) CDCl3, (b) DMF-d7, and (c) CD3OD. A drop of DMF-d7 was added into the pa1ticles prior to CD0; and CD3OD to increase the concentration of the dispersed nanoparticles. Reprinted with permission from J. Am. Chem. Soc. 2005, 127, 6248-6256. Copyright © 2005 American Chemical Society ................................................................................... 44 Figure 1.27. Diagram showing reversible formation of intermolecular hydrogen bonding between PNIPAAM chains and water molecules (left) and intrarnolecular hydrogen bonding between the C=O and N-H groups in PNIPAAM chains (right) below and above the LCST. This mechanism is proposed to explain the thermally responsive wettability of a PNIPAAM thin film .............................................................................. 45 -xi- Figure 1.28. Surface-roughness-enhanced wettability of a PNIPAAM-modified surface. a) The relationships between groove spacing (D) of rough surfaces and the water contact angle at 25 ° C (A), and at 40 C (I). b) Water drop profile for thermally responsive switching between super-hydrophilicity and super-hydrophobicity of a PNIPAAM- modified rough surface with a groove spacing of about 6 pm, at 25 °C and 40 °C. The water contact angles are ~O ° and 149.3 °, respectively. 0) The temperature (T) dependence of water contact angles for PNIPAAM thin films on a rough substrate with groove spacing of about 6 pm (A) and on flat substrate (I). (1) Water contact angles at two different temperatures for a PNIPAAM-modified rough substrate with a groove spacing of 6 pm. Half cycles: 20 °C; and integral cycles: 50 °C. Reprinted with permission from Angew. Chem, Int. Ed. 2004, 43, 357-360. Copyright © 2004 Wiley Intersciences ........................................................................................ 46 Figure 1.29. (top) Preparation of surface-confined PNIPAAM polymer brush nanopattems by combining “nanoshaving” and surface-initiated ATRP using a surface- tethered thiol initiator. (Bottom) Contact mode AF M height images (20 um X 20 um) and corresponding typical height profiles of a PNIPAAM brush line nanopattem imaged at room temperature in (a) air, (b) MQ-grade water, and (c) a mixture of MeOH/water (1:1, v:v). Reprinted with permission from Nano Lett. 2004, 4, 373-376. Copyright © 2004 American Chemical Society ....................................................................... 48 Figure 1.30. A) a PEGMA polymerized by ATRP from a micropattemed gold surface. (B) NIH 3T3 fibroblast cells cultured on patterns of adsorbed fibronectin (20 um circles and 40 um stripes). Reprinted with permission from Biomacromolecules, 2005, 6, 2427-2448. Copyright © 2005 American Chemical Society ......................................... 50 Figure 1.31. Preparation of polymer brushes grafted to porous membranes and an illustration of the mechanism for regulating water and solute permeation through a polymer brush grafted membrane and a hydrogel membrane ................................. 51 Figure 2.1. Architecture of comb-like poly(PEGMA) brushes grown from a surface. . 63 Figure 2.2. Reflectance FTIR spectra of (a) an initiator monolayer, and (b) an amorphous 114 nm thick poly(PEGMA) film. Spectrum b was taken immediately after its synthesis. Trace (c) shows a poly(PEGMA) film of the same fihn thickness after crystallization at room temperature .................................................................................... 73 Figure 2.3. Evolution of the ellipsometric film thickness with time for the ATRP of PEGMA at room temperature from initiators anchored to gold surfaces. Conditions: [CuBr] = 1 mM, [CuBrz] = 0.3 mM, [bpy] = 3 mM, and [M]:[H20] = 2:1 (v/v) ........... 74 -xii- Figure 2.4. Optical micrographs of poly(PEGMA) brushes grown from gold surfaces and viewed through crossed polarizers: (a) and (b) show the crystallization of a 240 nm at 10 °C and at 5 °C, respectively; (c) and (d) show the crystallization of 148 nm and 76 nm, films, respectively at 5 °C. Images in this dissertation are presented in color .............. 76 Figure 2.5. AFM images of a 108 nm crystalline polymer brush film. At left are the height images and at right are the corresponding 3-D views. In panel a, the scan distance is 4 x 4 pm, with a 2 distance of 25.44 nm. In panel b, the scan distance is 5 X 5 pm, and the 2 distance is 30.42 nm. Images in this dissertation are presented in color .............. 77 Figure 2.6. AFM images of crystalline polymer brushes. (3) shows data for a 310 nm thick film and (b) shows data for a the 122 nm film. Both image are 5 x 5 pm, and the heights of the 310 and 122 nm films are ~20 and 41 nm, respectively. Images in this dissertation are presented in color ................................................................ 79 Figure 2.7. (Left) AFM height images (5 x 5 pm) of crystalline polymer brushes with thickness of: (a) 92nm, 2 distance is 24.49nm; (b) 70nm, 2 distance is 31.46nm; (c) 35nm, 2 distance is 10.97nm. (Right) Phase images of crystalline polymer brushes with thickness of: (a) 92nm, 2 distance is 475mv; (b) 70nm, 2 distance is 919.13mV; (c) 35nm, 2 distance is 561mv. Images in this dissertation are presented in color ..................... 80 Figure 2.8. AFM height images of a crystalline film with thickness of 55 nm. Scan distance of the left image is 5 x 5 pm with a 2 distance of 15 nm. The scan distance of the right image is 1 x 1 pm with a z distance of 15 nm. Images in this dissertation are presented in color ................................................................................... 81 Figure 2.9. Crystalline structure of PEO with top view and side view ...................... 83 Figure 2.10. Reflectance FTIR spectra of poly(PEGMA) brushes of various thicknesses; a) 13 nm;b) 35 nm; e) 70m; (1) 92 nm; e) 177 nm;f)240nm .............................. 86 Figure 2.11. Lamellae orientations in the films of poly-PEGMA: (a) Chain axis of PEO side chains is oriented perpendicular to the substrate, lamellae parallel to the surface; (b) Chain axis of PEO side chains is parallel to the substrate, edge-on lamellae ............... 89 Figure 2.12. Optical micrographs showing the growth of spherulites in a 240 nm thick film at 23 i 2 °C. The samples are viewed using crossed polarizers. The scale bar represents 100 um. Images in this dissertation are presented in color ....................... 92 Figure 2.13 a) The evolution of spherulite radii with time for the crystallization of poly(PEGMA) brush films of various thickness at room temperature. b) Spherulite grth rates measured as a function of the poly(PEGMA) film thickness .................. 93 - xiii - Figure 2.14. Linear fit of the Avrami equation to the experimental data for crystalline polymer brushes with different film thickness .................................................. 95 Figure 2.15. ‘H-NMR spectrum of poly(PEGMA) ............................................. 97 Figure 2.16. Reflectance FT-IR spectra of 56 and 290 nm thick poly(PEGMA) films that were spin-coated on the gold surfaces ........................................................... 98 Figure 2.17.Crystal growth rates measured for spin-coated poly(PEGMA) and poly(PEGMA) brushes of comparable film thickness (~300 nm) ............................ 99 Figure 3.1 Evolution of the ellipsometric film thickness with time for the grth of poly(PEGMA)(22-23) from Au. Conditions: [M]:[H20] = 3:2 (v/v). (a) [CuBr] =10 mM, [CuBrz] = 5 mM, [bpy]= 30 mM. (b) [CuCl] =10 mM, [CuBrz] = 5 mM, [bpy]= 30 mM. The curves are a guide to the eye ................................................................ 113 Figure 3.2. Evolution of the ellipsometric film thickness with time for the growth of poly(PEGMA)(8-9) from Au. Conditions: (a) [CuBr] = 28.5 mM, [CuBrz] = 9.6 mM, [bpy] = 85.5mM; (b) [CuCl] = 28.5 mM, [CuBrz] = 9.6 mM, [bpy] = 85.5 mM. The curves are a guide to the eye ..................................................................... 114 Figure 3.3. Evolution of the ellipsometric film thickness with time for the grth of poly(PEGMA)(22-23) films from Au. [M]:[H20] = 3:2 (v/v). (a) [CuBr] = 0.7 mM, [CuBrz] = 0.21 mM, [bpy] = 2.1 mM; (b) [CuBr] = 2 mM, [CuBrz] = 0.6 mM, [bpy] = 6 mM; (c) [CuBr] = 5 mM, [CuBrz] = 1.5 mM, [bpy] = 15 mM; (d) [CuBr] = 10 mM, [CuBrz] = 3 mM, [bpy] = 30 mM ............................................................... 116 Figure 3.4. Evolution of the ellipsometric film thickness as a function of time for the polymerization of poly(PEGMA)(8-9). [M]:[[HZO] = 1:1.(a) [CuBr] = 28.5 mM, [CuBrz] = 9.6 mM, [pr] = 85.5 mM. (b) [CuBr] = 3 mM, [CuBrz] = 1 mM, [pr] = 9 mM. The curves are a guide to the eye ..................................................................... 117 Figure 3.5. Evolution of the ellipsometric film thickness with time for the polymerization of PEGMA(8-9) at different monomer concentrations. Conditions: [CuBr] = 28.5 mM, [CuBrz] = 9.6 mM, [bpy] = 85.5 mM. (a) [M]:[HzO]=2:1 and (b) [M]:[[HZO]=I:1. The curves are a guide to the eye ..................................................................... 118 Figure 3.6. Evolution of the ellipsometric fihn thickness with time for the polymerization of PEGMA(4-5) using Me4Cyclam as the ligand. Conditions: [M]:[HzO+DMF] = 1:1; [CuBr] = 2 mM, [CuBrz/(anbpy)2] = 0.6 mM, [Me4Cyclam] = 6 mM ................... 120 Figure 3.7. Evolution of the ellipsometric film thickness with time for the polymerization of PEGMA(4-5) using HMTETA as the ligand. Conditions: [M]/[HzO+DMF] = 1:1, -xiv- [CuBr] = 2 mM, [CuBrz] = 0.6 mM, [HMTETA] = 6 mM. The data points are the average of two independent polymerizations ............................................................ 121 Figure 3.8. Evolution of the ellipsometric film thickness with time for the polymerization of PEGMA(4-5) using HMTETA“) and Me4Cyclam (I) as ligands. The data are re— plotted from Figures 3.6 and 3.7 ................................................................. 123 Figure 3.9. Evolution of the ellipsometric film thickness with time for the polymerization of PEGMA(4-5). Conditions: [M]:[H20] = 2:1, (a) [CuBr] = 2 mM, [CuBrz] = 0.6 mM, [bpy] = 6 mM; (b) [CuBr] = 1 mM, [CuBrz] = 0.3 mM, [bpy] =3mM ..................... 126 Figure 3.10. FT-[R spectra of (a) initiator anchored fumed silica particles; (b) poly(PEGMA)(22-23) grown from fumed silica particles .................................... 128 Figure 3.11. TGA analysis of: (a) bare fumed silica; (b) initiator-anchored fumed silica; (c) poly-PEGMA (22-23) attached fumed silica .............................................. 130 Figure 3.12. TGA analysis of (a) bare fumed silica; (b) initiator-anchored fumed silica; c) poly(PEGMA)(4-5) attached fumed silica ...................................................... 132 Figure 4.1 Crosslinking and branching in films containing BisA-EDMA ................. 139 Figure 4.2 Reflectance FT-IR spectra of polymer films grown from an initiator anchored planar Au surface:(a) a poly(PEGMA) film and (b) a crosslinked film with 30 mol% crosslinker in the feed ............................................................................. 142 Figure 4.3. Evolution of film thickness with time for the copolymerization of PEGMA and BisA-EDMA. The overall concentration of monomer and crosslinker for all runs is 0.4 M. Catalyst: [CuBr] = 1mM; [CuBrz] = 0.3mM; [bpy] = 3mM. The inset shows the mole % composition of BisA-EDMA in the monomer feed solution ....................... 144 Figure 4.4. The relationship between the feed composition and the measured film. . 145 Figure 4.5. The fraction of crosslinkers in which both two double bonds polymerized. Each point is the average of at least three independent samples ............................ 146 Figure 4.6. Crosslink density vs. the mole percentage of crosslinker in feed ............ 147 Figure 4.7. a) (Top) thickness of dry and water-swollen films as a firnction of the mole percentages of crosslinker in the monomers feed solution; b) (bottom) the relationship between the normalized swelling of films and the fihn crosslink density ................. 149 Figure 5.1. Apparatus for gas permeation measurements .................................... 162 -xv- Figure 5.2. Synthesis of composite membranes with polymer skins ....................... 164 Figure 5.3. Transmittance FT-IR spectrum of poly(PEGMA)(22-23) and poly(PEGMA) (4-5) grafted to porous alumina membrane supports .......................................... 165 Figure 5.4. Single-gas fluxes of CO2 and H2 through fihns prepared by polymerization of PEGMA from porous alumina as a function of trans-membrane pressure drop. The outlet pressure was 1 atm and the measurements were performed at room temperature. . . . . ....166 Figure 5.5. FESEM image of the filtrate side of a porous alumina after grth of poly- PEGMA film from the surface .................................................................. 167 Figure 5.6. Single-gas fluxes of CO2 and H2 through poly-PEGMA grafted membrane as a function of trans-membrane pressure drop. The initially crystalline film was soaked in water, and then measurements were taken after the film was dried ........................ 168 Figure 5.7. Single-gas fluxes of CO2 and H2 through films prepared by polymerization of PEGMA with 8-9 ethylene oxide units from porous alumina as a function of trans- membrane pressure drop .......................................................................... 169 Figure 5.8. F ESEM image of the filtrate side of porous alumina after growth of (a) poly- PEGMA(8-9) fihn and (b) poly-PEGMA(4-5) film .......................................... 170 Figure 5.9. Single-gas fluxes of CO2 (I), H2 (9) and N2 (A) through cross-linked co- polymer films prepared by polymerization of PEGMA with a cross-linker (BisA-EDMA) from porous alumina as a function of trans-membrane pressure drop. For the film from 50mol% crosslinker, the flow rate of N2 is too slow to be measured ....................... 173 Figure 5.10. FESEM images (cross-sectional) of porous alumina (0.02 pm surface pore diameter) after grth of crosslinked films, which were polymerized from the monomer solution containing (a) 30mol% crosslinker; (b) 50 mol% crosslinker; (c) 80 mol% crosslinker; and (d) 100 mol% crosslinker (a poly(BisA-EDMA) film) ................... 174 -xvi- LIST OF SCHEMES Scheme Page Scheme 1.1. The general mechanism of controlled radical polymerization and some typical examples .................................................................................... 15 Scheme 1.2. The mechanism of ATRP of poly(ethylene glycol) methyl ether methacrylate (PEGMA) ............................................................................ 18 Scheme 1.3. Representative monomers used for ATRP ....................................... 21 Scheme 1.4. Representative examples of initiators used for ATRP .......................... 23 Scheme 1.5. Examples of ligands often used for copper-mediated ATRP .................. 25 Scheme 1.6. Synthetic approaches used to incorporate PEO segments in polymers 31 Scheme 2.1. Synthetic procedure for preparation of initiator on Au substrate ............. 68 Scheme 2.2. Growth of poly(PEGMA) films from gold ....................................... 72 Scheme 3.1. Surface-initiated ATRP of PEGMA from gold surfaces ...................... 105 Scheme 3.2. Synthetic route to (11-(2-bromo-2-methyl)propionyloxy) undecyldimethyl- chlorosilane ......................................................................................... 109 Scheme 3.3. Growth of comb-like PEO polymer brushes fi'om SiO2 surfaces ........... 125 Scheme 3.4. Modification of silica particles by surface-initiated ATRP of PEG- methacrylates ...................................................................................... 127 Scheme 4.1. Synthetic route to crosslinked copolymer films grown from gold substrates ........................................................................................... 141 Scheme 5.1. Structures of representative PEG (meth)acrylates and PEG di(meth)acrylates used in this study and reported in the literature for membrane separations ................ 156 - xvii - AFM AGET ATRP BisA-EDMA bpy uCP CRP DMAEMA DMF DP dNbpy anbpy DSC DPE DPN EDTA EGDMA FESEM FT -IR GMA GPC GPS HEA HEMA HMTETA kp kt LCP LCST LDPE Me4Cyclam MegTREN MA LIST OF ABBREVIATIONS atomic force microscopy activator generated by energy transfer atom transfer radical polymerization bisphenol A ethoxylate dimethacrylate (Mn =1 ,700g/mol) bipyridine (or 2,2’-dipyridyl) micro—contact printing controlled radical polymerization dimethyl amino ethyl methacrylate N, N-Dimethylformamide degree of polymerization 4,4’-di(5-nonyl)-2,2’-bipyridine 4,4’-di(n-nonyl)-2,2’-bipyridine differential scanning calorimetry diphenylethylene dip pen nanolithography ethylenediaminetetraacetic acid ethylene glycol dimethacrylate field emission scanning electron microscopy Fourier transform infrared spectroscopy glycidyl methacrylate gel permeation chromatography 3-glycidyloxypropy1 trimethoxysilane 2-hydroxy ethyl acrylate 2-hydroxy ethyl methacrylate 1 , 1 ,4,7,1 0, 10-hexamethyltriethylenetetramine propagation rate termination rate liquid crystalline polymer lower critical solution temperature low density polyethylene 1 ,4,8,1 1-tetramethyl-1 ,4, 8,1 l-tetraazacyclotetradecane tris[2-(dimethylamino)ethyl]amine methyl acrylate methyl methacrylate number average molecular weight weight average molecular weight nitroxide mediated radical polymerization nuclear magnetic resonance N-propyl (2-pyridyl) methanimine oligo(ethylene glycol) - xviii - PAA PCL PDI PDMS PDMSA PEO PEG PEGMA PEGDA PEGDMA PGMA PI PLA PMEMA PMDETA PMMA PNIPAAM PS PtBA Rp Rs RAFT ROP ROMP SAM TEMPO TGA THF poly(acrylic acid) poly(e-caprolactone) polydispersity index calculated as MW/Mn poly(dimethoxysiloxane) Poly(3-(dimethoxymethylsilyl)propyl acrylate) poly(ethylene oxide) poly(ethylene glycol) poly(ethylene glycol) methyl ether methacrylate, Mn= 300, 475, and 1100 g/mol poly(ethylene glycol) diacrylate poly(ethylene glycol) dimethacrylate poly(glycidyl methacrylate) polyisoprene poly(lactic acid) or poly(lactide) poly((2-N-morpholino)ethyl methacrylate) N, N, N’, N”, N”-pentamethyl diethylene triamine poly(methyl methacrylate) poly(N-isopropylacrylamide) polystyrene poly(tert-butyl acrylate) rate of polymerization radius of gyration reverse addition-fragmentation transfer ring opening polymerization ring opening metathesis polymerization self-assembled monolayer 2, 2, 6, 6-tetramethyl-1-piperidyloxy glass transition temperature thermal gravimetric analysis tetrahydrofuran melting point -xix- Chapter I Introduction to Polymer Brushes Thin polymer films are useful for modifying surface properties. Common methods for the deposition of polymers on surfaces include Spin and Spray coating, however, physically adsorbed films are easily displaced by good solvents. In contrast, the chemical stability and mechanical robustness of polymer brushes have led to their widespread use for surface modification. The term “polymer brush” generally refers to an ensemble of polymer chains with each chain bound by one end to a surface or interface. The high tethering density forces the chains to stretch away from the surface or interface, resulting in a brush height greater than the polymer radius of gyration (Rgfl’2 221122216 Figure 1.1. Polymer brushes on flat and curved surfaces. In general, polymer brushes are covalently bound to surfaces by reaction of functional groups on polymer chains with surface reactive sites, or by selective adsorption of block copolymer segments that strongly interact with the surface. Non- covalently attached polymer brushes also suffer from poor stability in good solvents. This discussion will emphasize covalently bound polymer brushes. -1- Strategies for anchoring polymer brushes onto surfaces can generally be categorized as “grafting to” and “graftingfrom” methods (Figure 1.2).3 The “grafting to” approach refers to tethering preformed polymer chains onto surfaces, which facilitates formation of polymer brushes with well-defined architectures. However, this method is limited in that it usually results in low polymer grafting densities and low film thickness due to the steric hindrance of previously attached polymer chains. In contrast, the “grafting fiom” method involves attaching initiators on the surface, and initiating polymerization of monomers from the surface. Since the only limit to the polymer chain propagation is the diffusion of monomers to the surface reactive sites, this method usually generates thicker polymer films with higher grafting densities than the “grafting to” approach.4 The “grafting fiom” method is also called surface-initiated polymerization. In order to achieve a high degree of control over the polymer brush architecture, the use of controlled polymerizationss’ 6 to provide well-defined polymers with low polydispersity (PDI) is highly desirable. Controlled surface-initiated polymerizations offer several advantages when forming well-defined dense and thick films including: 1) good control over the film thickness by controlling the polymerization time, 2) simplified separation and purification issues since the active radicals are confined on the surface and no polymers are formed in solution, eliminating the need for exhaustive extraction,7 3) tunable grafting densities by controlling the surface coverage of the initiator,8 4) the ability to simultaneously grow polymer chains from multifunctional cores or surfaces, and 5) the ability to control brush architectures, ranging from linear tethered polymers, graft (comb-like) copolymers, hyperbranched (star-like or dendritic) polymers, to cross- a) “Grafting to” b) “Grafting from” M M M M R. in i ............. l Propagating chain ends L Layer thickness I I H 2, , ”Ill _. SAM average distance between grafting points Figure 1.2. “Grafting to” (a) and “grafting from” (b) methods for the generation of polymer brushes. Adapted with permission from J. Am. Chem. Soc. 1999, 121, 1016- 1022. Copyright ©1999 American Chemical Society. linked polymer networks. Surface-initiated polymerization also provides good control over the “vertical” compositions of tethered linear polymers (normal to the surface) through the growth of block copolymers. Polymer brushes with varying compositions and dimensions have been prepared by their controlled growth from flat surfaces (gold, silica wafers, and glass slides), non—planar surfaces (gold, silica, latex particles, magnetic beads, polystyrene particles, carbon nanotubes), porous supports, chromatography media, and other materials."‘10 The physical properties of polymer brushes largely depend on the grafting densities, polymer chain lengths and flexibility, as well as the brush architecture. Polymer brushes have potential applications in many areas and they can greatly affect surface properties such as wettability and adhesion. Polymer brushes function as “smart” surfaces that are responsive to the environmental stimuli.”l3 barriers for chemical etching,”"6 layers resistant to adsorption of cells and biomolecules,l7 dielectric layers in electronic device,18 and other applications. Atom transfer radical polymerization (ATRP) is one of the most popular controlled polymerization methods due to its simplicity and functional group tolerance.” 2' ATRP also is a versatile technique for preparing well-defined (co)polymer films. In addition, ATRP of various monomers from preformed firnctional surfaces or colloidal initiators has provided polymer brushes with precise dimensions and functionalities. The nanoscale features or nano-objects in these films are typically characterized by various techniques, such as ellipsometry, contact angle measurements, and atomic force microscopy (AF M). Synthesis of polymer brushes by surface-initiated polymerization Surface-initiated polymerization generally includes anchoring initiator on the surface followed by the polymerization of monomer from the initiator. Initiators can be anchored to surfaces using well-known chemistry including the formation of self- assembled monolayers (SAM), such as thiols self-assembled on gold surfaces, silylation of silicon or glass surfaces, or by plasma treatment of substrates to directly form initiating species on polymer surfaces. Different types of surface-initiated polymerizations have been reported, including 22-27 28-30 2 anionic, cationic, conventional free radical?" 3 ring opening metathesis ”'36 and ring-opening polymerization (ROP).”39 Of these, polymerization (ROMP), controlled radical polymerization (CRP) is prominent due to its broad functional group tolerance compared to anionic and cationic polymerizations, and its ability to form polymers with high molecular weight and a PDI usually less than 1.5. Conventional -4- anionic polymerizations are the prototype living polymerization, but anionic polymerizations from surfaces have slow growth rates (usually days to generate a 100 nm film), and the polymerization system is sensitive to oxygen, moisture and impurities. Living cationic polymerization also can generate complex polymer architectures and polymers with low PDIs. However, the film growth rates for surface initiated cationic polymerizations are even slower than their anionic analogs, and also quite sensitive to moisture and other impurities. Much faster polymerization rates and films of predictable thicknesses are provided by traditional free radical polymerizations from surfaces, however free radical polymerization is incompatible with formation of certain architectures such as block copolymers. By maintaining a low concentration of active radicals in the system to suppress termination reactions, CRPs provide the advantages characteristic of living systems such as anionic and cationic polymerization. ATRP is tolerant of many functional groups and impurities, and because it is relatively easy to implement, it is the most popular CRP methodology for the controlled synthesis of polymer brushes. ROP has been used for preparing many commercially important polymers such as poly(s-caprolactone) (PCL) and polylactide (PLA). Strained cyclic monomers such as norbomenes are polymerized by ROMP. Some examples of surface- initiated polymerizations will be provided before the discussion of surface-initiated ATRP. Surface-initiated living anionic and living cationic polymerizations Due to its lack of chain transfer and termination reactions, anionic polymerizations provide the best possible control over polymer architectures. Surface- initiated anionic polymerization has been used to grow polymer brushes from various -5- small particles such as silica gel, carbon black, and silicon wafers. However, the challenge of using anionic polymerization is its stringent polymerization conditions, slow film growth rates, and the limited range of monomers that are compatible with anionic polymerizations. Jordan et al.24 used lithiated biphenyl SAMS and initiated the polymerization of styrene from gold substrates, obtaining uniform 18 nm thick films after a three-day reaction. Later, Advincula et al.22 activated diphenylethylene (DPE) SAMS with n-butyllithium to initiate anionic polymerization of styrene. A 5-day reaction gave a polystyrene (PS) film only 23 nm thick. However, the formation of polystyrene-block- polyisoprene (PS-b-PI) copolymer brushes demonstrated the living nature of the polymerization. Polymer brushes have also been grown from DPE derivatives anchored through quaternary ammonium tethers to clay surfaces. Zheng et a1.25 grafted polybutadiene brushes to the surface of PS nanoparticles, and reported that the grafting density and degree of polymerization greatly affect the dispersion of nanoparticles in matrices and the mechanical properties of nanocomposites. F oster26 used DPE-containing SAMS to generate polyisoprene-block-poly(ethylene oxide) (PI—b-PEO) block copolymer brushes from silicon surfaces, creating a hydrophilic surface by introducing the PEO block. Again, only thin films (24 nm) were obtained. Recently, Advincula et al.27 grew 14 nm thick PEO homopolymers and polyisoprene-block—poly(methyl methacrylate) (PI- b-PMMA) copolymer brushes on gold surfaces. From these examples, we see that the traditional monomers used for anionic polymerizations can be applied to surface-initiated polymerization, however, obtaining thin films after long reaction times restricts broader application of anionic polymerization to the synthesis of polymer brushes. Figure 1.3. Examples of polymer brushes grown by living anionic polymerization. H. .i- . .. .E? ””“l ”5 “ES; Si“ 7L WCHz-CH+ Adapted from Jordan et al. J Am. Chem. Soc. 1999 121 1016 1022 b) Adapted from Foster et al. Macromolecules 2002 35 9964 9974 c) AU S—CHZ n-BuLi S‘CHZ O phosphonium base Adapted from Advincula et al. J. Polym. Sci. A .' Polym. Chem 2006 44 769 Little work has been reported on the growth of polymer brushes using cationic polymerization. Ulman et al. reported surface-initiated cationic polymerization of 2- oxazolines from gold coated wafers28 and gold nanoparticle surfaces.29 Figure 1.4 shows the synthetic pathway used to obtain amphiphilic polymer brushes using this chemistry. i orf ‘ O O Et—(Oj N 6802CF3 Et Et 7? 0(soz-cr=3)2 S H HOH O .40 >—~ n-1 —b —> —-b Au nanoparticle © N(CH2)1—7CH3 < (CHfi-CHa Hydrophobic 17 alkyl shell H_N(CH2)1—70H3 E>_N Terra—cm o ".1 E Hydrophilic 17 polymer shell Metal 3-D core SAM Figure 1.4. Scheme showing the surface-initiated polymerization of 2-oxazolines for the preparation of the amphiphilic nanocomposites, with 2-ethyl-2-oxazoline and N,N-di-n- octadecylamine used as monomer and terminating agent, respectively. Adapted with permission from Macromolecules, 2001, 34, 1606-1611. Copyright © 2001 American Chemical Society. Zhao et al.40 synthesized 30 nm thick PS brushes in one hour from cumyl methyl ether groups attached to SAMS on silicates. The reaction was carried out at -78 °C to suppress the chain transfer reactions characteristic of cationic polymerizations. 0\ W b 9H3 O’Si_©'C-CD3 = rMMQ—c CH —c1-r / . 1m ,+ 2 I 2;, CH3 Di-tert—butylpyridine Figure 1.5. Polymer brushes grown by living cationic polymerization. Surface-initiated ring opening metathesis polymerization (ROMP) Surface-initiated ROMP is well suited for the polymerization of functionalized norbomenes, and as shown in Figure 1.6, has been demonstrated on Au, SiO2 and Si surfaces using Grubbs-type catalysts (Cat-1 and Cat-2) linked to surfaces by various molecules (SAM-1, 2, 3). Whitesides et al.33 used a surface-immobilized Grubbs-type ROMP catalyst to grow functionalized polynorbomenes. They reported the rapid, controlled growth of 90 nm thick brushes polymer brushes in 30 min, and also formed diblock copolymer brushes by exposing the brushes to a second monomer solution. The same strategy was utilized by Moon et al.34 to form brushes from polymers with poly(p- phenylene ethynylene) Side chains for chemical sensing applications. Although very thin films (10 nm) were obtained, the brush fluorescence was brighter than comparable spin- cast films, which was attributed to reduced aggregation of polymer chains. Figure 1.6. Surface-initiated ROMP applied to the synthesis of norbomene-based polymer brushes /—_\ O Q“ “ Cln. R'u— . Tm / / Men PCY3 Cat-1 Cat-2 SAM-1 SAM-2 SAM-3 3) F1. _ / _ / SAM-1 r—Q-ISi Cat-1! . T0331 SI ”OH Si O -——> SI 0 / \ \ R F0H — R R - H 8' CB / - 9 I( )3 O bO‘filMi' 0M6 003H17 ______, Si O / n r _ . \ [R] — — C3H170 Adapted from Whitesides et al. Macromolecules 2000, 33, 2793-2795 and Moon et al. Macromolecules 2002, 35, 6086-6089. b) F 1. PCI5 M Cat-1 NRU Lb Si > St —* Si T 3‘ 2. XMQCH2CH=CH2 RU __ L. — Adapted from Grubbs, R. H. et al. Langmuir 2001, 17, 1321-1323. -10- Grubbs et al.35 directly attached initiators through a Si-C bond to a Si (111) surface, providing a link between the polymer and a Si wafer without the insulating SiO2 layer. By varying the monomer concentrations, they obtained brush thickness ranging from 0.9 nm to 5 pm. ROMP has been coupled with other techniques to form nanopattemed functionalized surfaces. Mirkin et al.36 combined dip-pen nanolithography (DPN) and ROMP to form polymer brush arrays on the nanometer length scale. The novelty of this approach is that combinatorial polymer brush arrays can be formed by controlling the tip and substrate reaction time and different monomers can be used for polymerization. r 2:23.21 41 _. edited Ru l v /?\\ /f\. Si & sro2 Constant tip-substrate Different monomer contact time . . Different tlp— Ru Ru substrate contact time Ru Ru R R R R1 Ph n Ph D I I Ph D Ph n l I /Sli\ /Sli\ - ' /&\ /&\ Figure 1.7. Graphic representation of ROMP initiated by DPN and a monomer coated AFM tip. Adapted with permission from Angew. Chem, Int. Ed. 2003, 42, 4785-4789. Copyright © 2003 Wiley Intersciences. -11- Surface-initiated ring opening polymerization (ROP) Surface-initiated ROP has been applied to the synthesis of poly(N- propionylethyleneimine), PCL, PLA, and polyglutamate brushes.3 Figure 1.8 shows examples of the polymerization of e-caprolactone and L-lactide initiated from SAMS terminated with ethylene glycol. o O O a OH I 0 ) PhCHZOH 0% t... EtaAl ”IOTO o o O : H b) XH ’ 0 ' Sn(Oct)2 Xi/‘H/ rm; X=O,NH I Figure 1.8. PCL and PLA brushes formed by surface-initiated ROP. Husseman et a1.37 grew PCL from gold using aluminum alkoxide-catalyzed ROP. Brushes up to 70 nm thick formed in a few hours using a diethylaluminum alkoxide based catalyst; the concurrent formation of polymers in solution required exhaustive washing of the substrates to remove adsorbed polymer. SAMS terminated with oligo(ethylene glycol) (OEG) was also used to initiate polymerization of lactide,38 however, the growth rate was slow, with a three day polymerization at 40 °C yielding a 12 nm thick film. The use of amine-terminated SAMS on SiO2 were also investigated, and -12- 70 nm thick PLA brushes were grown in three days at 80 °C. Kim et al. reported the polymerization of lactide from poly(2-hydoxyethy1 methacrylate) (poly(HEMA)) brushes at 95 °C using Sn(2-ethylhexanoate)2 as the catalyst.41 In this example, the terminal OH groups on the poly(HEMA) side chains initiated the ROP of lactide. Despite the use of a gold substrate, the film was stable at the polymerization temperature due to partial crosslinking of poly(HEMA) as a result of transesterification. Au R CHZ— 0 AU R CHz—C Br CO2CH2CH2CH n 140 °C) are required for RAFT of vinyl esters, and the residual thioester end groups can cause odor problems in low molecular weight species. Development of improved transferable groups for RAFT is desirable. ATRP has been applied to largest number of monomers and over a wide temperature range (-20 to 130 °C). Since the polymer end group is a halogen, it can be easily converted to other functional groups. Some tolerance to oxygen and compatibility with water makes ATRP probably the most popular CRP method. However, the biggest challenge for ATRP is to solve the challenge of removing/recycling the metal catalysts, which impedes its wide application in biotechnology and for industrial syntheses of polymers. A detailed discussion of ATRP appears in the following section. Andruzzi et al.44 reported the use of surface-initiated NMP to grow polymer brushes with pendant OEG segments from a silica surface. The monomers were styrene derivatives with pendent OEG units, and were polymerized at 125 °C for 48 h. The as- forrned polymer brushes were used to study protein adsorption and cell adhesion. ~16- 602 %(OCH20H2),,OCH3 3-a n= 3 34) n: 7 3 1 la) 2, styrene — ib) 2, 3-a, b yo—b " . I (OCH20HzanCHa 8102 -?i 3 (OCH2CHzinOCHa n = 3, 7 n = 3, 7 Figure 1.11. Synthesis of tethered OEG containing polymer brushes on silica surfaces. Atom Transfer Radical Polymerization (ATRP) The general mechanism of ATRP is Shown in Scheme 1.2. The terminal halogen atom of the initiator (R-X) is extracted by a metal catalyst, such as a CuI complex, during initiation to form a reactive radical. This process occurs with a rate constant of km and the reverse reaction occurs with a rate constant kdmt. The reactive radical will add to monomers to form a polymer just as in traditional free radical polymerization, with the propagation rate kp. The advantage of ATRP is the low concentration of reactive radicals in the polymerization system, which is maintained by the equilibrium between the reactive radical and dormant species (K,q = km /kdcact). Since kdcact > km, the equilibrium favors the dormant species. Termination (kt) also occurs in ATRP, mainly by bimolecular -17- coupling or disproportionation. However, if the radical concentration is sufficiently low, termination can be ignored. Therefore, a successful ATRP will yield polymers with high molecular weight and low PDI. k R-X + Cu(l)X/bpy .:i—-‘E‘__ R' + Cu(l|)lebpy kdeact \‘x‘ “\‘k‘ kp monomer “ termination R-CHyC—X am CH3 I —’ {3:0 i C“ )x’bpy kdeacr R-CH2-C::° + CU(||)X2/bpy (OCH2CH27n30CH3 9:0 (OCHZCHZ‘F‘OCHg kp ll monomer Polymer Scheme 1.2. The mechanism of ATRP of poly(ethylene glycol) methyl ether methacrylate (PEGMA). -13- A T RP kinetics For copper-mediated ATRP, the polymerization rate (RP) can be described as45 K..iMiIIi. i951— Eq. 1 R = k P [XCuII] P where ch = k ac. / kdcact. The PDI can be expressed as , k I M” =1+ °[ 1" u (3—1) Eq.2 M kdcact[xcu ] P n In Eq. 1, RP is first order with respect to the solution concentrations of monomer [M] and the Cu1 catalyst. The initial build up of the Cu” species eventually leads to control over the polymerization as the equilibrium between Cu1 and Cu" is established. Kcq depends on the monomer structure and can be adjusted by the choice of the metal and ligands in the catalysts. A high value for ch gives a faster rate of polymerization, but if the value of Keq is too high, ATRP simply becomes a conventional redox-initiated radical polymerization (de30. ~ 0) with substantial termination. If ch is too small, the polymerization is too slow to be useful. Eq. 2 indicates that the PDI (MW/Mn) decreases with conversion, p. For a given monomer, a decrease in the kp/kdmt ratio through fast deactivation of the catalyst will result in lower PDI. In addition, a higher concentration of deactivator (Cuu) will help control the polymerization (narrow molecular weight distribution), with the decrease of polymerization rate. Therefore, in some examples of Cu-catalyzed ATRP, small amounts of Cu" halides are added to the system in the initial stage of polymerization to control the polymerization. -19- A T RP components Monomers ATRP has successfully been employed on styrenes, acrylates, methacrylates, acrylonitriles, (meth)acrylamide, and other monomers. Scheme 1.3 lists representative monomers successfully polymerized using ATRP. Monomer structure is one of the determining factors of kp, since it defines the stability of the radicals formed after the addition of growing radicals. Each monomer also has its own equilibrium constant (Keq = kact/kdew), which defines its polymerization rate in ATRP. Therefore, each monomer has unique optimum polymerization conditions. The monomers used in this study are PEGMA macromonomers, methacrylates which contain pendent ethylene oxide segments. The novelty of these macromonomers is partly due to the incorporation of the biologically relevant PEO segments, and that they also directly afford comb-shaped polymers with regular and densely attached branches. Furthermore, crosslinked polymers can be prepared by homopolymerization of divinyl monomers or copolymerization of mono and divinyl groups.” 47 -20- Scheme 1.3. Representative monomers used for ATRP a) Styrenes R= CH3, tBu x=Br, c1, F b) Acrylates O \ 10 33:0 OH c) Methacrylates rk f) erofiofk MMA HEMA PEGMA __ o ==)=o ==>=o O O thi k— /> I n 2n CmF2m+1 BA 0 02 02 %7 NMe2 0 0:13== GMA DMAEMA EGDMA -21- Initiators The diversity of ATRP initiators is larger than for other controlled polymerizations. Scheme 1.4. lists representative examples of initiators. A good ATRP initiator should be sufficiently reactive to ensure that all chain ends start to grow polymers at the same time, and the equilibrium in the initiation step should favor the dormant species in order to reduce the chance of Side reactions. Substituent groups on initiators stabilize the radical in the order of CN > C(=O)R > C(=O)OR > Ph > Cl > Me, with or-haloesters the most common initiators used in surface-initiated ATRP. The general order in bond strength is R-Cl > R-Br > R-1, and while alkyl iodides are efficient initiatiors, bromo and chloro or-haloesters are the most often used Since alkyl iodides require special precautions to avoid side reactions in ATRP. Catalysts and ligands (a) Cu based ATRP A variety of transition metal complexes have been applied in ATRP, such as Mo,48 Re,49 Ru,50’ 5‘ Fe,52'54 Rh,55 Ni,56 57 Pd,58 and Cu 21’ 59 metal complexes. The following discussion will emphasize copper-based ATRP since Cu mediated ATRP was used in this study. Cu complexes are versatile ATRP catalysts, with styrenes, (meth)acrylates, amides and acrylonitriles all successfully polymerized by ATRP. -22- Scheme l.4. Representative examples of initiators used for ATRP a) Halogenated alkanes and benzylic halides R X CHC'z O O RCCI3 x c1 or R = H. Cl. Br. R = H, CH3, X = Br, CI Ph, CF3, 002CH3, X = Br. 0' CnHZn+1 b) 01- Haloesters and their derivatives 0 O O o O 0 Br '2 Br 0 O O o HO 8 B \ B \/\o ' Vokr ' V\o/“\r r wok“ The choice of ligands used in ATRP catalysts can be used to adjust the equilibrium between the dormant and reactive radicals, and to solubilize and stabilize the metal catalyst in the reaction media at different temperatures. For copper-based ATRP, nitrogen ligands work best, with sulfur, oxygen or phosphorous ligands less effective, possibly due to the inappropriate electronic effects or unfavorable binding constants. The electronic and steric effects of ligands greatly affect the catalyst reactivity. Significant steric hindrance at the metal center or the use of ligands with strong electron-withdrawing -23- groups decreases catalyst reactivity. Scheme l.5 shows examples of ligands used for copper-based ATRP. Bipyridine (bpy) has been the ligand of choice for copper-mediated ATRP of styrene and (meth)acrylic monomers. Cu/bpy complexes perform well in aqueous media, but their limited solubility in organic solvents reduces their control over the polymerization. To improve the solubility of the catalysts in organic solvents and maintain better control over the polymerization, bpy is often replaced by an alkyl- substituted analog 4,4’-di(n-nonyl)-2,2’-bipyridine (anbpy). The activity of nitrogen ligands usually increases with the number of coordinated sites (N4 > N3 > N2 >> N1), and decreases with the number C atoms that link nitrogen (C2 > C3 > C4). Multidentate ligands, such as tris[2-(dimethylamino)ethyl]amine (MegTREN) and 1,1,4,7,10,10- hexamethyltriethylenetetramine (HMTETA) have been successfully used for the ATRP of styrene and methacrylates. (b) Structures of CuI complexes Cationic CuI complexes prefer a tetrahedral or square planar configuration with a tetradentate ligand or with two bidentate ligands. Cul complexes with bidentate ligands such as bpy are generally represented as [CuI (bpy)2]+[Y]'. For Cul Br and CuI C1 salts, the counter ion [Y]' can also be the [Cu1 Br2 ]' and [CuI Cl2]' anions, respectively. In the solid state, the typical geometry of the [Cu1 (bpy)2]+ cation is a distorted tetrahedron, as shown in Figure 1.12.60 The average Cu-N bond length ranges from 1.985 to 2.057 A and is not greatly affected by the counter ion or substituents on the bpy ligand. The dihedral angle is in the range of 80-83° due to the restricted geometry of the bpy ligands. -24- Scheme 1.5. Examples of ligands often used for copper-mediated ATRP. a) Bidentate ligands (N2) CW -N \N/ -N \N/ -~ N-Pr bpy dNbpy anbpy b) Tridentate ligands (N3) R ' \ R, (\N/N ,R —N/——\N— i \ N R’N N‘R QT) ’ N z_ \ / R=H, DETA R=CH3, PMDETA c) Tetradentate Ligands (N4) R. /_\ R R m R N N’ F183 C: :3 f 3 ,_ N R’U R R7“ “TR ' R‘R R=H, Cyclam [2:1-1' TETA R=H, TREN R=CH3, M84 Cyclam R=CH3, HMTETA -25- NPI'PMI 'i i \ N | \ /N N / R=H, BPMA R=(CH2)7CH3 BPMOA /\ N \ ’N ri/ /N \l TPMA R=CH3,M86TREN “v“ Figure 1.12. Structure ofthe [Cu‘ (bpy)2]+ cation in [Cu' (bpy)2]+[ClO4]'. Bridging is observed in Cu'X (X = halide) (Figure 1.13). In [Cul (bpy) Br2], each CuI center has a distorted tetrahedral geometry and is coordinated to the two nitrogen atoms of a bpy ligand and two bridging bromides. \ /\/x\/N\ / Cu / \ / W \N/ \ Figure 1.13. Structure of the [Cu1 (bpy) X2] (X=Br, Cl) dimer. The structures of Cu1 /bpy complexes in solution also have been studied. In non- polar media, CuIX/bpy complexes were proposed to exist predominantly as the bridged complexes shown in Figure 1.13. However, in polar media, the predominant complexes are [Cu' (bpyirriYI. -26- Cu/HMTETA complexes are very reactive catalysts and have been used in my research. The X-ray structure of [Cu'(HMTETA)2]+[Cu'Cl2]' complex (Figure 1.14) indicates Cu1 in a distorted tetrahedral geometry. 9 9 Figure 1.14. Molecular structure of [Cu‘(HMTETA)]+[Cu'C12]‘. (c) Structures of Cun complexes It is generally accepted that the most common Cu"X2/bpy complexes are [Cu"(bpy)2X]+[X]-. As shown in Figure 1.15, [Cu"(bpy)2X]+ adopts a trigonal bipyramidal geometry. Figure 1.15. Structure of the [Cu"(bpy)2Br]+ cation in [Cu”(bpy)2Br]+[Br]_. -27- Cu(II)X2 complexes with tetradentate ligands such as HMTETA ([Cu"(HMTETA)X]+[X]-) involve penta-coordination. As shown in Figure 1.16, the Cu center coordinates with four nitrogen atoms of the ligand and the donor halogen atom, forming a square pyramidal geometry. Figure 1.16. Molecular structure of [Cu"(HMTETA)Br]+[Br]‘. Water as a polymerization solvent A variety of solvents have been used for ATRP, including toluene, anisole, diphenyl ether, ethyl acetate, acetone, DMF, alcohols, and water.20 Generally, the use of polar solvents improves the solubility of Cu" complexes and favors homogeneous, well- controlled ATRP. Water has been increasingly employed for ATRP of polar, water soluble monomers having hydroxyl, oxyethylene, amino, ammonium, and carboxylate functional groups. Water also can greatly accelerate the polymerization rate. The first aqueous copper-mediated ATRP was reported by Matyjaszewski et al.,61 the polymerization of 2-hydroxyethyl acrylate (HEA) in aqueous media (50:50 v:v) using a -23- CuBr/bpy catalyst at 90 °C. A 12 hour polymerization resulted in 87% conversion of monomer to polymer. Later, Armes et al.62’ 63 reported the remarkably fast ATRP of PEGMAS in water, 90% conversion in 30 min at 20 °C, while retaining control over the polymerization (PDI < 1.3). The fast polymerization rate was attributed to the formation of mononuclear CuI species in water. Huang et al. described the rapid growth of poly(HEMA) brushes in water using a catalyst prepared from 30 mol% CuCl, CuBr2, and bpy-64 Water used as a co-solvent accelerates polymerizations, but there are no reports of the catalyst structure in water. However, it has been proposed that [CuI (bpy)2]+[X]' is the predominant complex in polar media, and is believed to be the reactive catalyst. There is also the possibility of solvent and monomer coordination to the catalyst. Haddleton65 studied the effect of water on the polymerization of PEGMA(8-9) using a CuBr/N-(n- propyl)-2-pyridylmethanimine (NPrPMl) complex. They proposed the competitive coordination of water and ligand to the copper center, which results in an increase in polymerization rate. Later, Jonsson66 studied the ATRP of PEGMA(7-8) in aqueous media, and calculated the propagating radical concentration from the redox properties of the alkyl halide and Cu complex. The calculated radical concentration in water is 3 orders of magnitude higher than in acetonitrile, which contributes to the fast ATRP kinetics in water. -29- A TRP ofPEGMA There are two general strategies for incorporating PEO segments in polymer structures via ATRP. In the macroinitiator approach”72 shown in Scheme 1.6., an a)- hydroxy group of PEO is converted into an ATRP initiator by coupling with an or- bromoacyl bromide or or-bromocarboxylic acid. This scheme provides block copolymers. The macromonomer approach is the direct polymerization of a PEG-containing monomer such as a PEG methacrylate. This scheme provides densely grafted comb polymers with PEG teeth. The macromonomer approach is a convenient way to form macromolecular structures with PEO segments. Haddleton et al.73 investigated ATRP of PEGMA (Mn = 480 g/mol) in toluene mediated by Cu'Br/NPrPMI. Using an on-line lH-NMR analysis, they observed a fast polymerization rate, 90% conversion in 1 hour at 90 °C. These reaction conditions are harsh and incompatible with polymerizations from gold since the Au-S bond is unstable above 60 °C. Armes et al. 62’ 63 studied ATRP of PEGMAs with ~8 and 45 repeating units of ethylene oxide in water, and also found that polymerization rate was remarkably fast (90% conversion in 20 min at 20 °C), but still controlled (PDI < 1.3). Apparently, water greatly accelerated the polymerization rate, as discussed above. The structure of PEGMA plays an important role in accelerating polymerization since the sterically congested PEGMA inhibits radical-radical coupling, which reduces the termination rate (kt). Decreases in k with increased monomer Size also has been observed for alkyl-substituted methacrylate esters. For example, kt for undecenyl methacrylate is 0.6 x 1045 L/(mol ~ 5), 1/40 of the rate for methyl methacrylate (25.5 x 1045 L/mol - s). -30- Scheme l.6. Synthetic approaches used to incorporate PEO segments in polymers. (a) Macroinitiator approach 0 macroinitiator polymer .01-019 B r Br O ‘ ATRP /Of/\O)r: ' /O~(/\O)):‘\(Br (b) Macromonomer approach ll O macromonomer O 0 V0)?" R’OMO R’ Br > o O O CuX, Ligands P )- O m Ito et al.74 reported aqueous ATRP of two styrene-functionalized PEG macromonomers (Mn = 2300 g/mol). They reported that kp /k1 was ~0.25 in water, and proposed inter-micellar polymerization to explain the high molecular weight they obtained. However, no evidence for micellar polymerization was found in Arme’s system. Later, Jonsson et al.66 estimated kp to be 3.6 x 10 3 M‘l s'1 for PEGMA (Mn = 480 g/mol) in water by correlating the apparent polymerization rate with the redox potential (El/2) of copper complex. The high overall polymerization rate was ascribed to complexation of ethylene glycol groups to copper, displacing some of the diamine ligands to form a very reactive metal catalyst. -31- The above examples of PEGMA polymerization provided low molecular weights, with degree of polymerizations (DPS) ~10-33. A recent paper by Matyjaszewski et al.75 described aqueous AGET (activator generated by electron transfer) ATRP of PEGMA homopolymers and random copolymers with DPS >240 and PDIS <1.3. Given the results Of PEGMA polymerization in solution, we expect surface- initiated ATRP of these monomers to provide thick polymer films on surfaces. The surface-initiated ATRP of PEGMA is elaborated in Chapters 11 and III. Polymer brushes formed by surface-initiated ATRP An advantage of the ATRP system is the facile process by which substrates are converted into initiators. Functionalization of the target substrate uses commercially available or-haloesters or benzyl halides and their derivatives, rather than the multi-step syntheses required for alkoxyarnines and xanthates used in other controlled polymerizations. Not surprisingly, ATRP is the most commonly used controlled polymerization technique, and has been applied to the growth of well-defined polymer brushes on gold, Silicon wafers, inorganic particles/colloids, organic latexes, and a variety of other substrates. Methods for the controlled growth of polymers from surfaces. As described earlier, the preparation of polymer brushes requires attachment of an initiator on the surface followed by ATRP of vinyl monomers from the substrate. Figure 1.19 depicts the general scheme for polymer brush formation on flat surfaces. -32- 0 OH R‘ 1) Hs/‘H 2) BIJKF , (R'=H,CH3) m C 3 - Br S/HOW‘JQ m 0 Or I O Svl’toJKK ATRP m Br Polymer brushes OH R R1 m OH R*= benzyl halides, sulfonyl OH halides, ct-haloesters, etc. R= c1. OCH3, R,= c1. OCH3, CH, Figure 1.17. General synthetic route to polymer brushes on gold and silica substrates. The main challenge in surface-initiated ATRP from flat surfaces is the inherent low concentration of initiators in the system, and therefore, a low concentration of the Cull radical deactivator. At low CuII concentrations, ATRP may resemble a redox process, an uncontrolled polymerization dominated by termination. The addition of deactivator in the form of Cu11 can compensate for the low CuII concentration, or sacrificial initiator can be added. In the case of added sacrificial initiator, termination of radicals in solution will generate a sufficient concentration of deactivator to afford control over the polymerization. The addition of “free” initiator to the system is not only beneficial for controlled synthesis of polymer brushes, but also provides a way to characterize polymer brushes. Analysis of polymers formed in solution can be used to estimate the molecular weight of tethered polymer chains, since several experiments have shown that the polymer chains cleaved from surfaces have molar masses and PDIS -33- Similar to the polymers formed in solution. The DP of tethered polymer chains can be estimated from the concentrations of consumed monomers and sacrificial initiator, assuming all initiators were used. DP = A [M]/[initiator] However, a significant drawback of adding sacrificial initiator is the formation of polymer in solution, which can be very difficult to remove from surfaces, especially during the synthesis of crosslinked polymer brushes. The alternative, adding a Cu" deactivator, is particularly useful in controlling polymer brush growth. Using 30 mol% CuBr2 as a deactivator in HEMA polymerizations, Huang et al. obtained 400 nm thick films compared to 200 nm films without CuBr2.64 The results are sensitive to the CuBr2 concentration; the limiting film thickness using 5% CuBr2 (relative to the Cul concentration) was slightly >200 nm. A related study on the polymerization kinetics of methyl acrylate (MA) defined 0.1 x 10'3 M as the CuI concentration that yields the thickest films when the Cu11 concentration was fixed at 30 mol%.76 The Cu1 concentration was varied from 40 mM to 2.5 x 10‘4 M in this study. Diluting the catalyst concentration also helps control the growth of polymer brushes. Brush architectures generated by surface-initiated A T RP. The versatility of surface-initiated polymerizations makes it possible to form a variety of brush architectures on flat surfaces, including block copolymer, binary polymer 77, 78 79, 80 brushes, comb-like polymer brushes, hyperbranched polymers, and crosslinked polymer brushes,81 as shown in Figure 1.18. Huang et al. described the facile synthesis of crosslinked polymer brushes on gold by surface-initiated ATRP.81 Chapter III will -34- describe studies of comb-like polymer brushes with PEO segments grown on different substrates, and crosslinked polymer brushes will be described in Chapter IV. Homo or block Comb-like Crosslinked Figure 1.18. Schematic illustration of linear, comb-like and crosslinked polymer brushes. Polymers grown by ATRP retain a halogen at the terminus of the growing polymer chain. As described earlier, these sites can be reactivated to initiate the formation of block copolymers. As shown by Zhao et al.82 these Sites also can be permanently deactivated. Zhao designed novel binary polymer brushes by combining ATRP and NMP. As shown in Figure 1.19, a Y-shaped SAM was formed on the surface that incorporated both ATRP and NMRP initiators. ATRP of MMA was followed by removal of the terminal halogen using Bu3SnH. Styrene was then polymerized by NMP to form binary polymer brushes that Show lateral phase separation. -35- PMMA PS PMMA Br I ,. ,3 O O O)? N 0’N O/ O O O O o O O O O O O O o NMRP 1) ATRP ________, MMA, 75 °C Styrene R 115 °c 2) Buaan 7////// /////////// ///////// M'xed PMMA/PS Y-SAM PMMA Brushes Brlushes Figure 1.19. Synthesis of mixed PMMA/PS brushes from an asymmetric di-functional initiator-terminated SAM (Y-SAM) by combining ATRP and NMRP techniques. Reprinted with permission from J. Am. Chem. Soc. 2004, 126, 6124-6134. Copyright © 2004 American Chemical Society. Controlled A T RP from surfaces of particles. Surface-initiated ATRP is not confined to the formation of polymer brushes on planar substrates. Polymer/inorganic hybrid nanoparticles have also been prepared by surface-initiated ATRP of various monomers from colloidal initiators,6’ 80‘ 83'” aiming to improve their stability in different solvents. A variety of inorganic substrates can be modified by polymer brushes; the following discussion will emphasize polymer brushes -36- grown on the surfaces of silica nanoparticles. Using mild conditions (water, 20 °C), Armes et a1.89 grafted poly(PEGMA) and poly((Z-N-morpholino)ethyl methacrylate)) (PMEMA) brushes on the surface of 300 nm silica particles prepared by the Stober process. The PMEMA grafted silica particles Showed lower critical solution temperature (LCST) behavior, and began to aggregate at 34 °C. The aggregates completely re- dispersed on cooling below the LCST. Bromoisobutyrate -based siloxane Ethanol 20°C 300nm Figure 1.20. Poly(PEGMA) and PMEMA brushes grown from silica nanoparticles. Zhao et (11.90 reported the first synthesis of comb-coil polymer brushes on the surface of silica nanoparticles. A poly(HEMA) brush was grown from the nanoparticle by surface-initiated ATRP. Then, concurrent polymerization of tert-butyl acrylate (tBA) and lactide were initiated from the terminal halogen group by AGET ATRP, and from the OH groups of the poly(HEMA) side chains by ROP, respectively. -37- HO 0 Br 1. ,CH3 OH (c,H,O),SiCH,CH,NH2 BI‘C‘CCB' Br HO : CH3 : Br OH Br HO Br ATRP HEMA ( PLA LA B Br r L“ '\ PBA ° 0 e — 6 O O CuClzlbpy .1- 4—_____ Sn Oct Br 0 0 BA Br Figure 1.21. Preparation of comb-coil polymer brushes on the surface of silica nanoparticles. Physicochemical properties of polymer brushes Physical states of polymer brushes The physical nature of surfaces (“hard” or “soft”) largely depends on the physical states of polymer brushes (crystalline or amorphous). To date, most polymer brushes have been amorphous, including glassy polymers such as PS or rubbery polymers such as polydimethoxysilane (PDMS). Liquid crystalline polymer (LCP) brushes have attracted substantial interest due to the application of LC materials for displays and as semi-conducting materials. Peng et al.91 synthesized and studied side chain LCP brushes comprised of phenylbenzoates -33- tethered to a polymethacrylate with a spacer. These films, up to 200 nm thick and prepared by surface-initiated free-radical polymerization, were used as alignment layers for LC systems. Alignment layers orient LCs at surfaces and provide the large area domains needed for displays. As the brushes were heated or exposed to solvent, the LC phase underwent a reversible nematic to isotropic LC phase transition. Recently, Huck et al.92 reported the synthesis of side chain LCP brushes formed by surface-initiated ATRP of active acrylate mesogens on silica and glass substrates (Figure 1.22). Glass sides with polymer brushes were oriented face to face and separated by a spacer. A nematic LC, 4’- pentyl-4-biphenyl carbonitrile, was inj ected into the gap between the slides and spread as a result of capillary forces. Homeotropic alignment was observed, which was ascribed to the interaction of LC side chain with the LC phase. Alignment was also observed on nanopattemed substrates, particularly in regions where polymer brushes were present on both the top and bottom substrates. This suggests potential applications for brushes in displays and optoelectronics. =>_O o \—\_\—W o_©_\<°H : O O O O \ ./\/\ O/S' OJYB' CuCl, CuClz, CuBr2, bpy O DMF, acetone, 50 °C 0 O \S./\/\ W O /O C6H12 Hit—CW Figure 1.22. Synthesis of side chain LCP brushes by surface-initiated ATRP. -39- Although much progress has been made on amorphous polymer films and LCP brushes, there are few examples of studies on the crystalline state of polymer brushes. Studies of polymer brush crystallization should help explain polymer crystallization in constrained geometries. Chapter II will elaborate on the crystallization behavior of polymer brushes. Stimuli-responsive polymer brushes Stimuli-responsive polymer brushes are attractive for applications as “smart” or responsive surfaces. Such brushes would alter their conformations upon changes in temperature, solvents, pH, photo-illumination, or another external trigger, resulting in changes in the surface physical properties. Solvent induced polymer brush rearrangement. For simple A-B block copolymers, phase separation of the A and B blocks generally occurs when the heat of mixing is positive. Phase separation of A-B block copolymers also occurs on surfaces, but being tethered to the surface, the polymer chains are constrained to separate in nanometer-sized domains. Irnmersing block copolymer brushes such as PMMA-b-PS in different solvents results in solvent-induced structuring 1.93"94 studied the rearrangement of PMMA-PS block polymer of the surface. Brittain et a brushes in CH2C12 (a good solvent for both blocks) and cyclohexane (a good solvent for PS but not for PMMA). When the polymer brushes are in CH2Cl2, both blocks swell and stretch away from the substrate, yielding a smooth surface. When the polymer brushes -40- are in a mixture of CH2C12 and cyclohexane, the PMMA blocks collapse to form aggregates and avoid interaction with cyclohexane. addition of lmmersuon cyclohexane in CHZClz Figure 1.23. Speculative model for nanopattem formation in di-block copolymers. A-B-A triblock copolymers showed similar phenomena. For PMMA-b-PS-b- PMMA, the film showed extended conformations when exposed to the good solvent (CH2C12), and a folded brush conformation when exposed to cyclohexane, which is a good solvent only for the middle block.95 Good solvent for middle block —————> ) ‘— Good solvent for all blocks Figure 1.24. Reversible response of triblock copolymer brushes to different solvents. Polymer brushes can show different solvent induced changes in their surface morphologies provided they have sufficient conformational freedom. Pyun et al.96 synthesized an A-B-C triblock copolymer (PDMS-b-PS-b-Poly((3—(dimethoxy -41- methylsilyl)propyl acrylate (DMSA) using anionic ring opening polymerization and ATRP. The polymer was “grafted to” the silica surface resulting in polymer brushes with low grafting densities. The polymer brushes rearranged to form glassy or rubbery surfaces when immersed in different solvents (toluene and hexane) (Figure 1.25). Various polymer brush architectures were constructed on surfaces to study their properties. For example, Li et (11.97 grew mixed PtBA/PS brushes on silica particles by the sequential ATRP of tBA and NMP of styrene from Y-shaped-initiators anchored on silica particles. The block copolymer chains tethered to these particles reorganized in response to changes in solvent, forming stable suspensions in CHC13 and MeOH, which are good solvents for PS and poly (acrylic acid) (PAA, hydrolysis product of PtBA), respectively. -42- Swollen Brush in Toluene Collapsed Brush in Hexane PDMS p5 - PDMS excess hexane —’ PS - PDMSA excess toluene PDMSA Rubbery surface Ultralight tapping MA0 —> 1 5 Normal tapping AIAo = 0.9 500 Figure 1.25. Tapping mode AF M height images of PDMS-b-PS-b-PDMSA brushes after the following treatments: (left) after immersion in toluene and drying with nitrogen, (right) after immersion in toluene, gradual addition of hexane and drying with nitrogen. Reprinted with permission from Macromol. Chem. Phys. 2004, 205, 411. Copyright © 2004 Wiley Intersciences. -43- (b) ppm Figure 1.26. ‘H NMR spectra of PAA/PS particles dispersed in (a) CDCl3, (b) DMF-d7, and (c) CD30D. A drop of DMF-d7 was added into the particles prior to CDC13 and CD3OD to increase the concentration of the dispersed nanoparticles. Reprinted with permission from J. Am. Chem. Soc. 2005, 127, 6248-6256. Copyright © 2005 American Chemical Society. T emperature-responsive polymer brushes Poly(N-isopropylacrylamide) (PNIPAAM) films are widely studied as temperature responsive films since its LCST is ~32-33 °C, an easily accessible and biologically relevant temperature. Jiang et al. prepared PNIPAAM films on a well- controlled rough surface by surface-initiated ATRP.98 The films switch between super- hydrophilicity (contact angle ~O °) and super-hydrophobicity (contact angle ~150 °) within a 10° temperature range. At 25 °C (below the LCST of PNIPAAM), the polymer -44- adopts an extended brush architecture and the surface is hydrophilic due to H-bonding between the polymer chains and water. At 40 °C (above the LCST), the polymer structure is collapsed and the surface is hydrophobic due to H-bonding exclusively between polymer chains. Heafingup ——> 4_— Codmgdown >LCST Figure 1.27. Diagram showing reversible formation of intermolecular hydrogen bonding between PNIPAAM chains and water molecules (left) and intramolecular hydrogen bonding between the C=O and N-H groups in PNIPAAM chains (right) below and above the LCST. This mechanism is proposed to explain the thermally responsive wettability of a PNIPAAM thin film. -45- a) 160 b) about0° 1493° \40°C 120 . ‘I\‘\‘\ l 80 CAP” #/ Heating up 0, -Coo|ingdown ° °D§3n3° __. °° 25°C C) 160, 29°C 4.000 d) 160 .III'Iljislololllllrll'l . ‘ I ‘ l ' 120 HIiIlI‘IIIHI' li ill'il Ill i ii I II II J l I 30 [HIM ‘le l' ”I'IIHI' IIl'Il'iIi ' , ['IIIII IIIIIII “I III I CAI“ III'II IIIIII'H'I'U' IIlI'Il .. 0111...; Clll.l.l I 2b 30 4o 50 o 4 32$L 12 16 20 TI°C —- Cycle _. Figure 1.28. Surface-roughness-enhanced wettability of a PNIPAAM-modified surface. a) The relationships between groove spacing (D) of rough surfaces and the water contact angle at 25 °C (A), and at 40 °C (I). b) Water drop profile for thermally responsive switching between super-hydrophilicity and super-hydrophobicity of a PNIPAAM- modified rough surface with a groove spacing of about 6 pm, at 25 °C and 40 °C. The water contact angles are ~0° and l49.3°, respectively. c) The temperature (T) dependence of water contact angles for PNIPAAM thin films on a rough substrate with groove spacing of about 6 mm (A) and on flat substrate (I). (1) Water contact angles at two different temperatures for a PNIPAAM-modified rough substrate with a groove spacing of 6 pm. Half cycles: 20 °C; and integral cycles: 50 °C. Reprinted with permission from Angew. Chem, Int. Ed. 2004, 43, 357-360. Copyright © 2004 Wiley Intersciences. -46- Applications of polymer brushes Polymer brushes can be applied in many different areas. For example, LCP brushes, discussed previously, can be used as semiconducting polymers for LC displays. With the development of nanotechniques, switchable surfaces have been used for bioanalysis, microfluidic devices, and protein separations. Some representative examples of polymer brush applications are described below. Nanotechnology—Patterning of polymer brushes One of the most important applications of polymer brushes is that they are well- suited for the fabrication of nano- or micro-patterned arrays, which could be useful in cell growth control, biomimetic microfabrication, drug delivery and microelectronics. Polymer brushes can be patterned by traditional photolithography or chemical amplification of patterned SAMS. Photolithography was first explored by Ruhe et al., 99 who exposed polymers to UV light through a mask. A good example of the complex patterning that can be carried out on polymer brushes was described by Hawker and co- workers.100 A PtBA brush was formed on a silica substrate, followed by spin-casting on the surface a polystyrene film containing bis(tert-butylphenyl)iodonium triflate. After exposing the surface to UV irradiation through a mask, the PtBA brushes were converted to PAA brushes, resulting in a surface with distinct hydrophobic and hydrophilic areas. SAM patterning was obtained by micro-contact printing (uCP), which was first introduced by Whitesides et al.,101 and the pattern initiator monolayer was amplified from the by surface initiated polymerization. Caster and co-workers 102 used scanning probe lithography and surface-initiated ATRP to nanopattem PNIPAAM brushes. -47- 1. immobilization ofinitiator 2.ATRPof lePAAM w) m) 65 nm 20 20 20 1° 10 ”m 10 10 "m 55‘ 55‘1 2 3 4 5 55‘ A 1 2 3 4 5 1 2 3 4 5 E ll )1 :7 JJ A A __ A A ii A .1: 0 0 0 % z -55 -5s. . -55. . . 051015 0 51015 0 51015 Width (pm) Vlfidth (pm) width (um) Figure 1.29. (top) Preparation of surface-confined PNIPAAM polymer brush nanopattems by combining “nanoshaving” and surface-initiated ATRP using a surface- tethered thiol initiator. (Bottom) Contact mode AFM height images (20 um X 20 um) and corresponding typical height profiles of a PNIPAAM brush line nanopattem imaged at room temperature in (a) air, (b) MQ-grade water, and (c) a mixture of MeOH/water (1 :1, v:v). Reprinted with permission from Nano Lett. 2004, 4, 373-376. Copyright © 2004 American Chemical Society. -43- Biological application in protein resistance and cell adhesion PEO (or PEG) is very important in biological applications due to its ability to resist protein absorption and cell adhesion (non-specific adhesion resistance). PEO- containing surfaces used for cmrent nanotechnology application are generated by self- assembly of OEG alkane thiols on gold surfaces. PEO-containing polymer brushes are attractive due to their synthetic robustness, and high surface grafiing densities. As shown earlier (Figure 1.12), Andruzzi et al. compared the OEG tethered polystyrene brushes with OEG SAMS, and found that polymer brushes were superior in their ability to inhibit protein (less than 6%) and cell (less than 3%) adhesion when compared to surface assemblies with the same ethylene glycol length. Biocompatible and non-biofouling PEO containing polymer brushes patterned onto gold or silica by various lithography techniques can be applied to “molecular recognition”. The non-pattemed regions are backfilled with a molecule that has a specific interaction with cell or other biomolecules. For example, Figure 1.31 shows a nanopattemed gold surface modified with poly(PEGMA) brushes, '03 and a cell-adhesive protein was absorbed on the surface after the features were backfilled with SAM 2. Two different cells were patterned on the surface. -49- Figure 1.30. A) a PEGMA polymerized by ATRP from a micropattemed gold surface. (B) NIH 3T3 fibroblast cells cultured on patterns of adsorbed fibronectin (20 um circles and 40 um stripes). Reprinted with permission from Biomacromolecules, 2005, 6, 2427- 2448. Copyright © 2005 American Chemical Society. -50- Polymer brushes grafted membranes. Ito et al. reported pH sensitive,104 photosensitive,'°5 and oxidation/reduction sensitive106 polymer brushes covalently tethered to porous membranes, which can be used to regulate liquids flowing through porous membranes. .-’ solute solute (a) Hydrogel membrane or chemical Immoblllzatlon of ___- _ __ .— clonal censltlve PO'VMOI‘ Signal (b) Polymer brushes membrane Signal sensitive polymer Water so|ute Water SOlUte Porous membrane 1 U , Graft polymerization .1121); MM Signal n. Am I m; f g d Figure 1.31. Preparation of polymer brushes grafted to porous membranes and an illustration of the mechanism for regulating water and solute permeation through a polymer brush grafted membrane and a hydrogel membrane. Polymer brushes grafted on membranes afford selective gas transport. Balachandra et 01.107 grafted cross-linked poly(ethylene glycol dimethacrylate) (Poly(EGDMA)) and poly(HEMA) brushes on the surface of porous aluminum membrane supports. Gas permeation studies with poly(EGDMA) films showed a COz/CH4 selectivity of 20, and a C02 permeability coefficient of 20 Barrer. The COz/CH4 selectivity of poly(HEMA) films was only 0.9, however, fluorination of -51.. poly(HEMA) increased the COz/CH4 selectivity to 9. Described in Chapter V are the preparation of PEO-containing polymer brushes grafied from porous alumina membrane supports, and the permeability of gasses through the membranes. Multilayer Polyelectrolyte Film (MPF) 4/ BrCOCl-I(CH,)Br ~(c,H,),/TH1= o 'c 'c .3 1:5 , CuCl/ Cqu bpy c 9 . ‘ +— ;: .g .g- eooww water! DMF >2 S. 5'- RT 4)‘ (D‘ "- ‘3 z 5 MPF Cross-linked polymer — " V film Initiator studied to MPF Figure 1.32. Schematic diagram showing polymerization of EGDMA from a polyelectrolyte surface modified with an initiator. Reprinted with permission from J. Membr. Sci. 2003, 227, 1-14. -52- References: 1. 10. 11. 12. 13. 14. Zhao, B.; Brittain, W. J., Prog. Polym. Sci. 2000, 25, 677-710. Milner, S. T., Science 1991, 251, 905-914. Edmondson, S.; Osborne, V. L.; Huck, W. T. 3., Chem. Soc. Rev. 2004, 33, 14-22. Jennings, G. K.; Brantley, E. L., Adv. Mater. 2004, 16, 1983-1994. 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, 1424-1431. Huang, X.; Wirth, M. J ., Macromolecules 1999, 32, 1694-1696. Kim, J. B.; Bruening, M. L.; Baker, G. L., J. Am. Chem. Soc. 2000, 122, 7616- 7617. Jones, D. M.; Brown, A. A.; Huck, W. T. S., Langmuir 2002, 18, 1265-1269. Pyun, J .; Kowalewski, T.; Matyjaszewski, K., Macromol. Rapid Commun. 2003, 24, 1043-1059. Pyun, J .; Matyjaszewski, K., Chem. Mater. 2001, 13, 3436-3448. Ionov, L.; Minko, S.; Stamm, M.; Gohy, J. F.; Jerome, K.; Scholl, A., J. Am. Chem. Soc. 2003, 125, 8302-8306. Currie, E. P. K.; Norde, W.; Stuart, M. A. C., Adv. Colloid. Interface Sci. 2003, 100, 205-265. Luzinov, I.; Minko, S.; Tsukruk, V. V., Prog. Polym. Sci. 2004, 29, 635-698. Huck, W. T. 8.; Yan, L.; Stroock, A.; Haag, R.; Whitesides, G. M., Langmuir 1999, 15, 6862-6867. -53- 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. Zhou, F.; Jiang, L.; Liu, W. M.; Xue, Q. J ., Macromol. Rapid. Commun. 2004, 25, 1979-1983. Zhou, F.; Liu, W. M.; Hao, J. C.; Xu, T.; Chen, M.; Xue, Q. J ., Adv. Funct. Mater. 2003, 13, 938-942. Senaratne, W.; Andruzzi, L.; Ober, C. K., Biomacromolecules 2005, 6, 2427- 2448. Rutenberg, I. M.; Scherman, O. A.; Grubbs, R. H.; Jiang, W. R.; Garfunkel, E.; Bao, Z., J. Am. Chem. Soc. 2004, 126, 4062-4063. Kamigaito, M.; Ando, T.; Sawarnoto, M., 2001, 101, 3689-3745. Matyjaszewski, K.; Xia, J. H., Chem. Rev. 2001, 101, 2921-2990. Patten, T. E.; Matyjaszewski, K., Adv. Mater. 1998, 10, 901. Advincula, R.; Zhou, Q. G.; Park, M.; Wang, S. G.; Mays, J .; Sakellariou, G.; Pispas, S.; Hadjichristidis, N., Langmuir 2002, 18, 8672-8684. Advincula, R. C., J. Dispersion Sci. Tech. 2003, 24, 343-361. Jordan, R.; Ulman, A.; Kang, J. F .; Rafailovich, M. H.; Sokolov, J ., J. Am. Chem. Soc. 1999, 121, 1016-1022. Zheng, L.; Xie, A. F .; Lean, J. T., Macromolecules 2004, 3 7, 9954-9962. Quirk, R. P.; Mathers, R. T.; Cregger, T.; Foster, M. D., Macromolecules 2002, 35, 9964-9974. Sakellariou, G.; Park, M.; Advincula, R.; Mays, J. W.; Hadjichristidis, N., J. Polym. Sci. A: Polym. Chem. 2006, 44, 769-782. Jordan, R.; Ulman, A., J. Am. Chem. Soc. 1998, 120, 243-247. -54- 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. Jordan, R.; West, N.; Ulman, A.; Chou, Y. M.; Nuyken, 0., Macromolecules 2001, 34, 1606-1611. Zhao, B.; Brittain, W. J., J. Am. Chem. Soc. 1999, 121, 3557-3558. Prucker, 0.; Ruhe, J ., Macromolecules 1998, 31, 592-601. Prucker, 0.; Ruhe, J ., Macromolecules 1998, 31 , 602-613. 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 , 2793-2795. Moon, J. H.; Swager, T. M., Macromolecules 2002, 35, 6086-6089. Juang, A.; Scherman, 0. A.; Grubbs, R. H.; Lewis, N. S., Langmuir 2001, 17, 1321-1323. Liu, X. G.; Guo, S. W.; Mirkin, C. A., Angew. Chem, Int. Ed. 2003, 42, 4785- 4789. Husemann, M.; Mecerreyes, D.; Hawker, C. J .; Hedrick, J. L.; Shah, R.; Abbott, N. L., Angew. Chem, Int. Ed. 1999, 38, 647-649. Choi, I. S.; Langer, R., Macromolecules 2001, 34, 5361-5363. Joubert, M.; Delaite, C.; Bourgeat-Lami, E.; Dumas, P., Macromol. Rapid Commun. 2005, 26, 602-607. Zhao, B.; Brittain, W. J ., Macromolecules 2000, 33 , 342-348. Kim, J. B.; Huang, W.; Wang, C.; Bruening, M. L.; Baker, G. L., Bottle Brush Brushes: Ring-Opening Polymerization of Lactide from Poly(hydroxy methacrylate) Surfaces. Wiley-VCH VerlagGmbH&Co. KGaA: Weinheim, 2004. Baum, M.; Brittain, W. J ., Macromolecules 2002, 35 , 610-615. -55- 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. Matyjaszewski, K., Controlled Radical Polymerization. Washington, DC, 1998. Andruzzi, L.; Senaratne, W.; Hexemer, A.; Sheets, E. D.; Ilic, B.; Kramer, E. J .; Baird, B.; Ober, C. K., Langmuir 2005, 21, 2495-2504. Matyjaszewski, K., Macromol. Symp. 2002, 183, 71-81. Connal, L. A.; Gurr, P. A.; Qiao, G. G.; Solomon, D. H., J. Mater. Chem. 2005, 15, 1286-1292. Wang, A. R.; Zhu, S. P., Macromolecules 2002, 35, 9926-9933. Brandts, J. A. M.; van de Geijn, P.; van Faassen, E. E.; Boersma, J .; van Koten, G., J Organometallic Chem. 1999, 584, 246-253. Kotani, Y.; Kamigaito, M.; Sawamoto, M., Macromolecules 1999, 32, 2420-2424. Kato, M.; Kamigaito, M.; Sawamoto, M.; Higashimura, T., Macromolecules 1995, 28, 1721-1723. Bielawski, C. W.; Louie, J .; Grubbs, R. H., J. Am. Chem. Soc. 2000, 122, 12872- 12873. Wang, J. S.; Matyjaszewski, K., J. Am. Chem. Soc. 1995, 117, 5614-5615. Matyjaszewski, K.; Wei, M. L.; Xia, J. H.; McDermott, N. E., Macromolecules 1997, 30, 8161-8164. Muhlebach, A.; Gaynor, S. G.; Matyjaszewski, K., Macromolecules 1998, 31 , 6046-6052. Percec, V.; Barboiu, B.; Neumann, A.; Ronda, J. C.; Zhao, M. Y., Macromolecules 1996, 29, 3665-3668. Uegaki, H.; Kotani, Y.; Kamigaito, M.; Sawamoto, M., Macromolecules 1998, 31 , 6756-6761. -56- 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. Moineau, G.; Minet, M.; Dubois, P.; Teyssie, P.; Senninger, T.; Jerome, R., Macromolecules 1999, 32, 27-35. Lecomte, P.; Drapier, 1.; Dubois, P.; Teyssie, P.; Jerome, R., Macromolecules 1997, 30, 7631-7633. Patten, T. E.; Matyjaszewski, K., Acc. Chem. Res. 1999, 32, 895-903. Pintauer, T.; Matyjaszewski, K., Coordin. Chem. Rev. 2005, 249, 1155-1184. Coca, S.; Jasieczek, C. B.; Beers, K. L.; Matyjaszewski, K., J. Polym. Sci. Polym. Chem. 1998, 36, 1417-1424. Wang, X. S.; Lascelles, S. F .; Jackson, R. A.; Armes, S. P., Chem. Commun. 1999,1817-1818. Wang, X. S.; Armes, S. P., Macromolecules 2000, 33, 6640-6647. Huang, W. X.; Kim, J. B.; Bruening, M. L.; Baker, G. L., Macromolecules 2002, 35, 1175-1179. Perrier, S.; Haddleton, D. M., Macromol. Symp. 2002, 261-272. Coullerez, G.; Carlmark, A.; Malmstrom, E.; Jonsson, M., J. Phys. Chem. A 2004, 108, 7129-7131. Lee, S. B.; Russell, A. J .; Matyjaszewski, K., Biomacromolecules 2003, 4, 1386- 1393. Ali, M. M.; Stover, H. D. H., Macromolecules 2004, 37, 5219-5227. Liu, S. Y.; Weaver, J. V. M.; Tang, Y. Q.; Billingham, N. C.; Armes, S. P.; Tribe, K., Macromolecules 2002, 35, 6121-6131. Cai, Y. L.; Armes, S. P., Macromolecules 2004, 3 7, 7116-7122. -57- 71. 72. Fujii, S.; Cai, Y. L.; Weaver, J. V. M.; Armes, S. P., J, Am. Chem. Soc. 2005, 127, 7304-7305. Dufresne, M. H.; Gauthier, M. A.; Leroux, J. C ., Bioconjugate Chem. 2005, 16, 1027-1033. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. Haddleton, D. M.; Perrier, S.; Bon, S. A. F ., Macromolecules 2000, 33, 8246- 8251. Nomura, E.; Ito, K.; Kajiwara, A.; Kamachi, M., Macromolecules 1997, 30, 2811- 2817. 0h, J. K.; Min, K.; Matyjaszewski, K., 2006, 39, 3161-3167. Kim, J. B.; Huang, W. X.; Miller, M. D.; Baker, G. L.; Bruening, M. L., J. Polym. Sci. A: Polym. Chem. 2003, 41, 386-394. Lemieux, M.; Usov, D.; Minko, S.; Stamm, M.; Shulha, H.; Tsukruk, V. V., Macromolecules 2003, 3 6, 7244-7255. Minko, S.; Patil, S.; Datsyuk, V.; Simon, F.; Eichhom, K. J .; Motomov, M.; Usov, D.; Tokarev, 1.; Stamm, M., Langmuir 2002, 18, 289-296. Mori, H.; Boker, A.; Krausch, G.; Muller, A. H. E., Macromolecules 2001, 34, 6871-6882. Mori, H.; Seng, D. C.; Zhang, M. F.; Muller, A. H. E., Langmuir 2002, 18, 3682- 3693. Huang, W. X.; Baker, G. L.; Bruening, M. L., Angew. Chem, Int. Ed. 2001, 40, 1510. Zhao, B.; Haasch, R. T.; MacLaren, S., J. Am. Chem. Soc. 2004, 126, 6124-6134. Pyun, J .; Jia, S. J .; Kowalewski, T.; Patterson, G. D.; Matyjaszewski, K., Macromolecules 2003, 36, 5094-5104. -53- 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. von Weme, T.; Patten, T. E., J. Am. Chem. Soc. 2001, 123, 7497-7505. Bontempo, D.; Tirelli, N.; Masci, G.; Crescenzi, V.; Hubbell, J. A., Macromol. Rapid Commun. 2002, 23, 418-422. von Weme, T.; Patten, T. E., J. Am. Chem. Soc. 1999, 121, 7409-7410. Mandal, T. K.; Fleming, M. 8.; Walt, D. R., Chem. Mater. 2000, 12, 3481-3487. Xiao, D. Q.; Wirth, M. J ., Macromolecules 2002, 35, 2919-2925. Perruchot, C.; Khan, M. A.; Kamitsi, A.; Armes, S. F.; von Weme, T.; Patten, T. E., Langmuir 2001, 1 7, 4479-4481. Zhao, H. Y.; Kang, X. L.; Liu, L., Macromolecules 2005, 38, 10619-10622. Peng, B.; Johannsmann, D.; Ruhe, J ., Macromolecules 1999, 32, 6759-6766. Hamelinck, P. J.; Huck, W. T. S., J. Mater. Chem. 2005, 15, 381-385. Zhao, B.; Brittain, W. J .; Zhou, W. S.; Cheng, S. Z. D., J. Am. Chem. Soc. 2000, 122, 2407-2408. Zhao, B.; Brittain, W. J ., Macromolecules 2000, 33, 8813-8820. Boyes, S. G.; Brittain, W. J .; Weng, X.; Cheng, S. Z. D., Macromolecules 2002, 35, 4960-4967. Pyun, J .; Jia, S. J .; Kowalewski, T.; Matyjaszewski, K., Macromol. Chem. Phys. 2004, 205, 411-417. Li, D. J .; Sheng, X.; Zhao, B., J. Am. Chem. Soc. 2005, 12 7, 6248-6256. Sun, T. L.; Wang, G. J.; Feng, L.; Liu, B. Q.; Ma, Y. M.; Jiang, L.; Zhu, D. B., Angew. Chem, Int. Ed. 2004, 43, 357-360. -59- 99. 100. 101. 102. 103. 104. 105. 106. 107. Prucker, 0.; Schimmel, M.; Tovar, G.; Knoll, W.; Ruhe, J., Adv. Mater. 1998, 10, 1073. Husemann, M.; Morrison, M.; Benoit, D.; Frommer, K. J .; Mate, C. M.; Hinsberg, W. D.; Hedrick, J. L.; Hawker, C. J., J. Am. Chem. Soc. 2000, 122, 1844-1845. Xia, Y. N.; Whitesides, G. M., Angew. Chem, Int. Ed. 1998, 37, 551-575. Kaholek, M.; Lee, W. K.; LaMattina, B.; Caster, K. C.; Zauscher, S., Nano Lett. 2004, 4, 373-376. Ma, H. W.; Hyun, J. H.; Stiller, P.; Chilkoti, A., Adv. Mater. 2004, 16, 338-+. Ito, Y.; Park, Y. S.; Imanishi, Y., J. Am. Chem. Soc. 1997, 119, 2739-2740. Ito, Y.; Ochiai, Y.; Park, Y. S.; Imanishi, Y., J. Am. Chem. Soc. 1997, 119, 1619- 1623. Ito, Y.; Nishi, S.; Park, Y. S.; Imanishi, Y., Macromolecules 1997, 30, 5856-5859. Balachandra, A. M.; Baker, G. L.; Bruening, M. L., J. Membr. Sci. 2003, 227, 1- 14. -60.- Chapter II Crystallization of Polymer Brushes This chapter describes the crystallization of comb-like polymer brushes formed by surface-initiated atom transfer radical polymerization (ATRP) of PEG-methacrylates (PEGMA) from gold surfaces in aqueous media. Optical microscopy and atomic force microscopy (AF M) show that the ethylene oxide side chains of poly(PEGMA) crystallize to form two-dimensional spherulites. The crystal growth rates are film-thickness dependent. AFM and reflectance FT-IR measurements on < 100 nm fihns are consistent with formation of crystalline lamellae predominantly oriented parallel to the surface, but thicker films are mainly comprised of perpendicularly-oriented lamellae. The crystallization of polymer brushes occurs by heterogeneous nucleation and one- dimensional growth. Introduction Thin and ultra-thin films are attractive for tailoring surface properties because of their wide range of physicochemical properties. Studies on polymer films, primarily physically absorbed films deposited on substrates by spin-coating, show that the interactions between the substrate and polymer layers usually lead to physical properties that are different than those of the bulk polymer. For example, the glass transition temperature (T 8) may be enhanced or depressed depending on the film thickness and interfacial interactions,1 and restriction of chain mobility greatly affects the 10.11 Crystallization process. Polymer morphology, 2'9 degree of crystallinity, preferred -61- 2, 7,12 3,10,11,13-18 polymer chain orientation, and crystal growth rates may vary with the film thickness. Although physical adsorption of films on substrates is a convenient and versatile method for surface modification, the stability of physisorbed coatings is limited in the presence of good solvents. For this reason, films covalently bonded to substrates, such as polymer brushes, are being rapidly adopted for the study of thin films. Some properties of polymer brushes are unique to their architecture and have not been explored in detail. For example, the effect that the polymer brush architecture may have on crystallization is unclear since the extended chains of densely grafted brushes may pre-organize and favor crystallization, while chain entanglements and the restricted mobility of the polymer chains in these brushes may retard the crystallization process. Luzinov et al. 19’ 20 studied the surface morphology of PEG brushes formed by reacting the end groups of activated PEGs of different molecular masses with a surface primed with poly(glycidyl methacrylate). After PEG grafting, the surface was either partially or totally covered by finger-like crystalline domains, depending on the molecular mass of the constituent PEGs. Since the thicknesses of the films were similar, the variations in crystalline morphology must be related to the grafting density and the molecular masses of the PEGs. To further understand the effects of brush architecture on crystallization, this work examines the crystallization of comb-like polymer brushes obtained fiom ATRP of poly(ethylene glycol) methyl ether methacrylate (PEGMA) from initiators immobilized on Au surfaces. In the case of PEGMA macromonomers, polymerization from surfaces yields methacrylate comb polymer brushes with hydrophilic poly(ethylene oxide) (PEO) side chains as shown in Figure 2.1. The side chains of poly(PEGMA) and other comb -62- polymers crystallize when they are sufficiently long, irrespective of the tacticity of the polymer backbone.”23 9H2 H3C-C-ICI-O(CH2CH20>CH3 + 0 22-23 Figure 2.1. Architecture of comb-like poly(PEGMA) brushes grown from a surface. The melting temperature (T,,.) of the PEO segment in comb polymers depends on the length of side chains and ranges from below room temperature to ~65 °C, the T m of PEO. This range is compatible with thermally unstable substrates such as self-assembled monolayers on Au, enabling detailed IR analysis of the structure and orientation of the crystalline PEO side chains. In addition to their relevance for understanding crystallization of thin films, poly(PEGMA) brushes have important potential applications as gas separation membranes and materials for limiting non-specific protein adsorption on surfaces. Studies of chain crystallization in poly(PEGMA) brushes complement extensive research on crystallization of bulk PEO,” 25 PEO films on surfaces,“ 3‘ ‘4’ ‘5 PEO segments in block copolymers,” 27 as well as films of PMMA/PEO blends.7' 28' 29 Crystalline PEO comb-polymers are usually arranged in a “sandwich” structure, with layers of amorphous backbones alternating with lamellae of crystalline side chains. For example, Inomata et al. suggested a crystalline lamellae packing for PEO side chains of poly(MMA-stat-PEGMA), with amorphous backbone connecting lamellae.23 -63- Neugebauer et al. proposed a model of alternating disordered and crystalline domains for polymethacrylates with PEG and alkyl side chains. 2‘ Although the chain conformations in adsorbed films and brushes are different, polymer crystallization studies of physisorbed films provide a useful starting point for understanding the crystallization of brushes. As film thickness decreases to nm dimensions in the physisorbed polymer films, the spherulitic morphology observed in 6’7’ 30 or finger-like patterns for bulk polymers evolves into sheaf-like aggregates, monolayer films.9’ 3" 32 AFM makes it possible to observe individual lamellae in spherulites. For most ultra-thin polymer films studied to date, crystalline lamellae preferentially orient parallel to the substrate, with the polymer chain axis oriented normal to the surface. The orientation of crystalline lamellae may change with the fihn thickness, and lamellae oriented parallel to the surface may give way to perpendicular lamellae as the film thickness increases. For example, Schonherr et al.2 studied PEO films and reported crystalline lamellae oriented parallel to the substrate when the films were < 300 nm thick, and perpendicular lamellae for film thickness > 1 am. This trend in lamellae orientation was also found in crystalline low density polyethylene (LDPE) films7 and poly(di-n-hexylsilane) films.12 The accepted explanation for these thickness effects is that perpendicular orientation minimizes the energy for the primary nucleation step, and therefore, the same orientation is favored for secondary nucleation and crystal growth, resulting in the favored perpendicular lamellae orientation. In ultra-thin films, lamellae lying flat on the surface reduce the surface energy more effectively and are favored over perpendicular lamellae. -64- Frank et al.11 reported that the crystallization rate and crystallinity of poly(di-n- hexylsilane) films decreased significantly when the fihn thickness was < 50 nm. Since then, several research groups confirmed a reduction in crystal grth rates as film thickness decreased, but conflicting results remain regarding the correlation between film thickness and crystal grth rate. Taguchi et al.16 studied the grth rate of isotactic polystyrene films and claimed that the fihn growth rate (G) can be related to film thickness (d) using a simple equation: G(d) = C(00) (1- a/d), where C(00) refers to the growth rate of bulk polymer, and the constant a is ~ 6 nm. In contrast to the above results, Schonherr et al.2 studied isothermal crystallization of PEO films on silicon substrates, and concluded from an Avrami analysis of FT-IR data that the crystallite growth rate and film thickness have an exponential relationship, i.e, G(d) = C(00) exp (1- constant/d). l. '4’ 18 studied isothermal crystallization of PEO films ranging from 13 Dalnoki-Veress et a nm to 2 pm, and reported a non-monotonic decrease in crystal grth rate for film thickness > 40 nm, which they related to morphology changes. This chapter describes the synthesis and characterization of crystalline polymer brushes with an emphasis on the effect of film thickness on morphology, chain orientation in crystallites and the crystallization rate. The crystallization kinetics for brushes is interpreted in the context of the Avrami-Evans theory. -65- Experimental Section Materials. Poly(ethylene glycol) methyl ether methacrylate (PEGMA, Mn=.1,100 g/mol), Cu(I)Br (99.999%), Cu(II)Br2 (Aldrich, 99.999%), and ll-mercapto-undecanol [HS(CH2)I 10H] were used as received. 2,2’-Bipyridine (bpy) was recrystallized from hot hexane and sublimed under vacuum at 60 °C. The disulfide initiator ((CH3)2CBrC00(CH2)IIS)2) was synthesized using a slightly modified version of a literature procedure.33 Deionized water (Milli-Q, 18 MQcm) was used as the polymerization solvent. Synthesis of initiator [Br-C(CH3)2-C00(CHz)uS)2] 33 CH2C12 (150 mL), 10% aqueous potassium bicarbonate (20 mL) and ll-mercaptoundecanol (4.0 g) were added to a 250 mL Schlenk flask. Bromine (0.51 mL) was slowly syringed into the well-stirred mixture, and the initial orange-red color of bromine quickly disappeared. After continued stirring for 2 h, the organic and water layers were separated, and the water layer was extracted with 50 mL of CH2C12. The combined organic layers were dried over MgSO4, filtered, and evaporated to dryness to yield 2.95 g (75%) of H0(CH2)1 IS-S(CH2)110H (1) as a white powder. 1H-NMR (300 MHz, CDC13): 5 1.23 (br, 32H, CH2), 1.52 (m, 4H, CH2), 1.62 (m, 4H, CH2), 2.62 (t, 4H, SCHz), 3.61(t, 4H, 0CH2). Under a blanket of Ar, 2-bromoisobutyryl bromide (1.46 mL) was added drop-wise to a 0 °C solution of disulfide [1] (2.0 g, 5 mmol) and triethyl amine (3.4 mL, 0.07 mol) in 120 mL CH2C12. After stirring at 0 °C for 1 h, and at 25 °C for 2 h, the solution was filtered to remove triethylamine hydrobromide and then extracted with 2N sodium carbonate solution saturated with aqueous ammonium chloride (2 x 50 mL). The dichloromethane was removed by rotary evaporation and the crude product was purified by column -66- chromatography using 10:1 hexane/ethyl acetate as the eluent. Removal of the solvent provided 1.5 g of the initiator (43%) as a pale yellow viscous liquid. 1H-NMR (300 MHz, CDC13): 5 1.31 (br, 28H), 1.7 (br, 8H), 1.9 (s, 12H), 2.7(t, 4H, J = 7.3), 4.18 (t, 4H, J = 6.6). 13C-NMR: 5 172 (C=0), 66.4 (0CH2), 56.2 (C), 39.4 (SCHz), 31.0(CH3), 29.7(CH2), 29.5(CH2), 29.4(CH2), 28.7(CH2), 28.6(CH2), 26.0(CH2). Anal. Calcd. for C30H56Br204S2: C, 51.13; H, 8.01. Found C, 51.27; H, 7.69. Scheme 2.1. Synthetic procedure for preparation of initiator 2 2 113-6011217011 l Br2 1 S+CH2)-OH 11 s—(CH2)—0H 11 0 E13N I Br/U\
99 % conversion of PEGMA (see Figure 2.15). M11 = 60,000 g/mol. Characterization methods. Reflectance FT-IR was performed using a Nicolet Magma-IR 560 spectrometer with a PIKE grazing angle of 80 °C. Reflectance FT-IR experiments used p-polarized light. Since only transition moment components perpendicular to the surface provide absorption peaks in the spectra, the molecular conformation of samples were inferred from IR absorption spectra using the previously reported PEG assignments. Film thicknesses were measured by a rotating analyzer ellipsometer (M-44; J .A. Woollarn) at an incident angle of 75°. The data were analyzed using WVASE32 software, and the thickness and refractive index determinations were performed on at least three spots on each substrate. The refractive index was fitted with the film thickness, and the measured refractive index was around 1.5. AF M images were measured using the close-contact mode of a nano-RTM instrument, and both height and phase images were recorded simultaneously. The silicon tip was used with a spring constant of 36 N/m, with tip curvature of 10-20 nm, and a resonance frequency of 286- 339 KHz. The scan rate was set to 0.85 Hz and the set-point was adjusted to optimize image quality. Crystallization studies. The morphology of crystalline polymer brushes were imaged using Nikon 0ptiphot2-POL polarizing optical microscope equipped with a Mettler FP 82 hot-stage and a CCD camera. Samples for isothermal crystallization experiments were heated at 50 °C for 8 h in a vacuum oven to erase the thermal history of -69- the film. The films were immediately transferred to the hot stage, and held at 23 i 2 °C under a flow of N2. In non-isothermal crystallization experiments, the samples were loaded into the hot stage and heated to 45 °C (above the melting temperature), and then cooled at 2 °C /min to ambient temperature, and further cooled at 1 °C /min to ~5 °C by purging the hot-stage with the boil off from liquid N2. For other experiments, the slides were cooled to ambient temperature in a desiccator under reduced pressure or under N2. Polymer solutions for spin-coating were prepared as 1-10 wt % solutions in solvent CH2Cl2 and filtered through 0.02 pm PTFE filters. All films were deposited at 3000 rpm, with the film thickness controlled by varying the concentration of the polymer solution. After spin-coating, the films were dried under vacuum and annealed at 30 °C. The films were characterized by external reflectance FT-[R and by ellipsometry. Results and discussion Preparation of polymer brushes. Scheme 2.1 shows the synthetic route used to grow poly(PEGMA) from gold. Methacrylates containing 4-5, 8-9, and 22-23 ethylene oxide repeat units were polymerized, but only poly(PEGMA) with 22-23 PEO units crystallized at room temperature. Research on bulk poly(PEGMA) shows that the atactic backbone of poly(PEGMA) and a portion of side chains adjacent to the backbone constitute an amorphous phase, and the ethylene oxide chain must exceed a minimum chain length, typically 9-10 ethylene oxide units to crystallize. In the remainder of this Chapter, PEGMA refers to the methacrylate with 22-23 PEO units, unless otherwise Specified. Polymerization studies of other PEG methacrylates appear in Chapter III. -70- Immobilization of the initiator was confirmed by the appearance of a carbonyl peak in the reflectance FT-IR spectrum (Figure 2.2a) as well as formation of a 17 i 2 A- thick film, as measured by ellipsometry. The PEGMA macromonomer was polymerized from the surface using aqueous solution containing the CuBr/CuBr2/bpy ATRP catalyst. During the first 8 h of the polymerization, the film thickness increased linearly with polymerization time, consistent with a controlled polymerization (see Figure 2.3), but at longer times, the film growth rate decreased, which may reflect reaction of active chain ends with adventitious oxygen in the glovebag. IR spectra of the films (Figure 2.2b) show characteristic bands for poly(PEGMA): C-H stretching (2950-2800 cm'l), C=0 stretching (1740 cm”), CH2 bending (1460 cm"), CH2 wagging (1350 cm"), overlapping CH2 twisting (~1300-1240 cm'l), and a broad C- 0-C stretching band (1150 cm'l). The width and position of the IR peaks are consistent with an amorphous polymer brush film. As described below, the evolution of the IR bands indicates that these films crystallize at room temperature. -71- o CH s(c1+2+o— -C+Br Initiator-anchored A CH3 substrate 0 CH3 s-(0H2—)-o— —c+ar CH3 0 CH3 11 CH3 0 CH3 +sfcnfi—o— -CI:+Br 0 01-13 +sfc112—)-o- c—éer CH3 CH3 H3C 22-23 io/VOTEF) Macromonomer PEGMA 0 22-23 CuBr, Cu(||)Br2.bpy / \ “N H20, r.t. sfcrufi—o-c— c—(—CH2—c’C—O)—Br Scheme 2.2. Growth of poly(PEGMA) films from gold. -72- \ / 0.04 1L c: Crystalline polymer film JL_ b: Amorphous polymer film “Ju- - m;- a: Initiator x 30 Absorbance W 3600 3200 2800 2400 2000 1600 1200 800 Wavenumbers (cm'l) Figure 2.2. Reflectance FTIR spectra of (a) an initiator monolayer, and (b) an amorphous 114 nm thick poly(PEGMA) film. Spectrum b was taken immediately after its synthesis. Trace (c) shows a poly(PEGMA) film of the same fihn thickness after crystallization at room temperature. -73- 500 400 - D E - 300 4 ' '3 in . U) .' 2 f, 200 - F .E l- 100 - P 13' 111 o I I I I I o 5 1o 15 20 25 30 Time (h) Figure 2.3. Evolution of the ellipsometric fihn thickness with time for the ATRP of PEGMA at room temperature from initiators anchored to gold surfaces. Conditions: [CuBr] = 1 mM, [CuBr2] = 0.3 mM, [bpy] = 3 mM, and [M]:[H20] = 2:1 (v/v). -74- Morphology of crystalline polymer brushes. The optical microscopy experiments described here provide a qualitative picture of the crystallization behavior as a function of temperature and film thickness. While stored at room temperature, the films crystallized and the mirror-like finish of as-prepared fihns developed surface features that were visible to the naked eye. When observed between crossed polarizers, the as-prepared polymer films initially appeared dark (amorphous), but eventually the Maltese cross patterns characteristic of spherulites filled the field of view. The images shown in Figure 2.4 are from samples annealed at 45 °C and then cooled to the desired temperature. Since the Tm of poly(PEGMA) films are ~35-40 °C, the super-cooling (AT = T c-Tm) at ~5 °C supports a sufficiently high crystallization rate to enable crystallites to be observed in a short period of time. Panels a and b in Figure 2.4 show the evolution in polymer morphology for a 240 nm-thick fihn. At 10 °C, large and well-defined “spherulites” cover most of the field of view. The domains have similar radii, consistent with nucleation of each domain at comparable times. At 5 °C, the crystalline domains grow to cover the entire surface, as shown in panel b. These disc-like domains are large, ~100 nm in diameter but only 240 nm thick, resulting in an aspect ratio comparable to digital video discs (DVDs). Given their symmetry and disc-like geometry, the domains may be viewed as a 2-dimensional slice through the center of a spherulite. Crystallization under isothermal conditions at room temperature under N2 also produced films with spherulitic morphologies. Panels b, c and (1 show the effects of film thickness on crystallization. All three samples were imaged at 5 °C under comparable conditions. While the 240 nm film was fully crystallized, some areas of the 148 and 76 nm films remained amorphous. The 76 -75- run film appeared as sheaf-like aggregates, which may be typical of the early stages of spherulite development. These observations show that the crystallization behavior poly(PEGMA) films is similar to thin PEO films, and that both the crystalline morphology and crystallization rate depend on the poly(PEGMA) film thickness. Figure 2.4. Optical micrographs of poly(PEGMA) brushes grown from gold surfaces and viewed through crossed polarizers: (a) and (b) show the crystallization of a 240 nm at 10 °C and at 5 °C, respectively; (c) and (d) show the crystallization of 148 nm and 76 nm, films, respectively at 5 °C. Images in this dissertation are presented in color. -76- AFM measurements provided a more detailed view of the crystalline polymer brush morphology. Figure 2.5 shows flat and 3-D height images obtained for a 108 nm film. Panel a shows a clearly resolved boundary (1 ) between two spherulitic domains and well-defined perpendicular lamellae (2) that emanate from the center of the “spherulite”. Panel b shows the apparent overlap of two spherulitic domains, possibly at the site of a film defect, resulting in twisted branches and a more irregular boundary between the two domains. The evolution of film morphology with thickness was examined using 20-300 nm films that were crystallized under identical conditions (isothermal crystallization). Figure 2.6 shows AFM images for 122 and 310 nm films. The image of the 310 nm film shows thick, closely packed perpendicular lamellae. The observed branching of lamellae in the 122 nm film is the expected consequence of the radial growth of “spherulites”. The morphologies of films < 100 nm are more complicated and are thickness- dependent. Panel a of Figure 2.7 Compared to thicker films, the lamellar morphology seen in the height image is less resolved, but more clearly shown in the phase image. The morphology of the 70 nm film is similar to that of the 92 nm film. A switch from perpendicularly oriented lamellae to lamellae presumably lying parallel to the surface came from the image of a 55 nm film, which clearly shows a worm-like film morphology that resembles aggregates of lamellae parallel to the surface (Figure 2.8). As the film thickness decreased to 35 nm, however, the AFM image of the film appeared homogeneous and nearly featureless, suggesting either an amorphous film, or a morphology dominated by parallel lamellae. -77- (a) 25 nm Onm (b) 30nm iOnm 0 pm 2.5 pm 5 pm Figure 2.5. AFM images of a 108 nm crystalline polymer brush film. At left are the height images and at right are the corresponding 3-D views. In panel a, the scan distance is 4 x 4 pm, with a 2 distance of 25 nm. In panel b, the scan distance is 5 x 5 pm, and the 2 distance is 30 nm. Images in this dissertation are presented in color. -73- (a) 310nm 20 nm 0 nm 0 pm 2.5 pm 5 pm (b) 122nm . 40 nm 0 nm Figure 2.6. AFM images of crystalline polymer brushes. (a) shows data for a 310 nm thick film and (b) shows data for a the 122 nm film. Both image are 5 x 5 um, and the heights of the 310 and 122 nm films are 20 and 40 nm, respectively. Images in this dissertation are presented in color. -79- (a) 92 nm 24 nm f. - 475 mV i’ ' { ' 2., it >- . ’25:?" y”:- '-.' #2- r: {3,15}? ' . gut-c; ./ 3 In?“ ' '1 14:13:- e‘ - '1?- “‘1‘: "'- " ': - Onm --».p; _ OmV 0pm 25pm 5pm 0pm 25pm 5pm (b) 70 nm 31 "m 919 mV Onm V Ont 0 pm 2.5 pm 5 pm 0 11m 2-5 pm 5 pm (c) 35 nm 11 nm 561 mV Onm OmV 0pm 25pm 5pm 0pm 25pm 5pm -30- Figure 2.7. (Left) AFM height images (5 x 5 pm) of crystalline polymer brushes with thickness of: (a) 92 nm, 2 distance is 24 nm; (b) 70 nm, 2 distance is 31 nm; (e) 35 nm, 2 distance is 10 nm. (Right) Phase images of crystalline polymer brushes with thickness of: (a) 92 nm, 2 distance is 475 mV; (b) 70nm, 2 distance is 919 mV; (c) 35 nm, 2 distance is 561 mV. Images in this dissertation are presented in color. 15 nm Figure 2.8. AFM height images of a crystalline film with thickness of 55 nm. Scan distance of the left image is 5 x 5 pm with a 2 distance of 15 nm. The scan distance in the right image is l x 1 pm with a 2 distance of 15 nm. Images in this dissertation are presented in color. ~81- Spectroscopic characterization Reflectance FT-IR measurements can provide spectroscopic confirmation of crystallization and information on the orientation of the poly(PEGMA) side chains relative to the surface. As seen in Figure 2.2c, crystallization causes several poly(PEGMA) IR bands to sharpen. Notable changes consistent with crystallization include the C-O-C band at 1150 cm'1 splitting into strong peaks at 1120 and 1148 cm], and the CH2 wagging peak at ~1350 cm'l evolving into a sharp peak at 1360 cm'1 with a small shoulder at 1345 cm'l. Both indicate a high degree of crystallinity for the PEO chain. PEO usually crystallizes as a 72 helix, which contains seven CH2CH2O repeat units and two turns of the helix in the 19.3 A between equivalent sites along the helix. The PEO chain in the helix is a succession of nearly trans (CCOC), trans (COCC), and gauche (OCCO) conformations along the chain.” 25 This structure, shown in Figure 2.9, has been confirmed by X-ray measurement as well as extensive spectroscopic studies. PEO adopts a planar zig-zag conformation under tension,34 but when the stress is removed, the zig-zag conformation disappears rapidly. The symmetry of the IR bands for the helical PEO have been assigned to two species, one with transition moment parallel to the helical axis, and the other with transition moment perpendicular to the helical axis (See Table 2.1). -32- Top view Side view Figure 2.9. Crystalline structure of PEO with top view and side view. Polymer brushes with different film thicknesses were characterized by reflectance FT -IR spectroscopy (Figure 2.10). All films were annealed at 50 °C to erase their thermal history, and then held at room temperature until crystallization was complete. Trace a is the IR spectrum of a 13 nm-thick film, and the width and position of the peaks indicate that this film was amorphous. All fihns of comparable thickness (< 20 nm) proved to be amorphous, possibly because the restricted conformations of ultrathin films prevent efficient packing of the PEO side chains into crystalline lamellae. Another possibility is that films < 20 nm thick have melting points at or below room temperature. However, the the latter seems unlikely, since a 300 nm fihn melts at 38 i 1 °C, and the melting point only decreases to 34 :l: 1 °C for a 80 nm film. -83- Table 2.1. Absorption band assignments for crystalline PEO Ban (1 assignment 3,1, Transition Wavenumber Wavenumber (cm'l) moments ° (cm") a (this study) 5(CH2) J- 1473,1466 1467 6((CH2) 11 1461,1453 d w (CH2) 4 1358 1360 w (CH2) 11 1342 1344 1(an ) J- 1278 1280 1(an ) 11 1240 1242 v(C0) J- 1147 1148 0(C0) -L 1116 1120 0(CO) 11 1103 e v(C0)+r(CH2)+u(CC) J- 1060 1062 r (CH2)+ 1) (CH2 ) 11 958 964 r (CH2) + 0(CC) J- 947 947 r(CH2)+u(CO) J- 844 843 3 Ref 24,29 b 5 = scissor; w = wag; t = twist; u = stretch; r = rock. c J- and II represent polarization direction perpendicular and parallel to the helical axis of PEO, d overlapped. e not observed. -84- The IR spectrum of a 35 nm thick fihn (spectrum b) shows only four peaks between 1400 and 900 em‘1 (1344, 1242, 1120, 964 cm"). The CH2 wagging (1360, 1344 cm"), CH2 rocking (964, 947 cm'l) and combination bands (843 cm") are sensitive to the conformation order.24 Absorption peaks at 1344, 1242, and 964 cm"1 in trace b correspond to vibrations with transition moments parallel to the PEO helical axis. Moreover, since the reflectance FT-IR measurements used p-polarized light with the active transition moment normal to the surface, the helical axis of ethylene oxide units was oriented perpendicular to the substrate. The FT-IR spectra evolved as the film thickness increased. Spectrum c shows a small band at 1148 cm], which appears as a shoulder of the 1120 cm'1 band. Splitting of this C-O-C absorption confirms crystallization of the ethylene oxide segments. In addition to the 1344 and 1242 cm'1 bands seen in thinner films, there are new absorptions at 1360 and 1280 cm}. These bands are more obvious in trace (1. The 1360, 1280, 1148, 947 and 843 cm'1 bands correspond to vibrations with transition moments perpendicular to the helical axis. Thus, there are two populations of helical PEO in 92 nm thick film, one with its axis perpendicular to the surface and the other parallel to the surface. In thicker films (traces e and f), the 1360, 1280, 947 and 843 cm‘1 bands dominate indicating an increasing proportion of PEO segments oriented parallel to the surface. The FT-IR data confirm the previous interpretations of the AFM images. The morphology of films >100 nm thick appears to be dominated by perpendicularly oriented lamellae, since polarized FT-[R measurements identify the dominant helix orientation in these films as parallel to the surface, and hence consistent with lamellae oriented normal to the surface. -85- 1148 1120 1467 1360 134412801242 I ||l| 843 I 0.1 964 947 1062 f) 240 nm e) 177 nm (1) 92 nm '1 A j \ c) 70 nm b) 35 nm a)13 nm (intensity x 5) A - A Absorbance 1 600 1500 1400 1 300 1200 1 100 1000 900 800 Wavenumber (cm'1) Figure 2.10. Reflectance FTIR spectra of poly(PEGMA) brushes of various thicknesses; a) 13 nm; b) 35 nm; c) 70 nm; (1) 92 nm; e) 177 nm; 1) 240 nm. The brushes were initially annealed at 50 °C and then allowed to crystallize at room temperature in N2. -86- The morphologies of films <100 nm thick are more complicated, with AFM showing a progressive loss of perpendicular lamellae as films become thinner. For very thin fihns such as the 35 nm thick film shown in Figure 2.7, the AFM image was featureless, but FT-IR detected relatively sharp bands at 1342 and 1242, 1120, and 964 cm'1 indicating some order, and by inference, crystallinity. In contrast, the 55 nm fihn of Figure 2.8 shows signs of structure in the AFM image, different than seen in thicker fihns, suggesting lamellae oriented parallel to the surface. The FT-IR spectra of thin films are consistent with such a conclusion. IR measurements indicate that the PEO chains in the 92 and 70 nm-thick fihns are orientated both parallel and perpendicular to the surface. AFM only showed irregular perpendicular lamellae, possibly due to the overgrth of perpendicular lamellae on the top of parallel lamellae. In addition, AF M usually detects the top layer morphology of the film, while vibrational bands shown in the IR spectra reflect mean orientations of polymer chains. The two orientations for crystalline lamellae in films are shown in Figure 2.11. Crystallization of thin polymer brushes in lamellae parallel to the surface would reduce the polymer and surface interaction, which is favorable. Rafailovich et al.7 claimed that for the crystallization of thick LDPE films, the energy of the primary nucleation step is minimal when the chains are oriented parallel to the surface. However, as described in later sections of this chapter, crystallization of poly(PEGMA) occurs by heterogeneous nucleation. Formation of perpendicular lamellae also may be driven by comb-like structure of poly(PEGMA), which favors effective side chain interactions when the side chains are oriented parallel to the surface. Bulk comb polymer samples typically -37- crystallize in layers with the crystalline side chains oriented roughly perpendicular to the backbone. Densely packed comb polymer brushes favor an extended polymer backbone when the surface density is high enough, with the polymer backbones roughly oriented perpendicular to the surface and the side chains parallel to the surface. Crystal grth follows the same orientation, leading to the perpendicular lamellae. A similar side chain orientation was described recently by Gabriel et al.35 In addition to identifying chain orientation, IR spectroscopy suggests that films must be >20 nm to exhibit detectable crystallinity. Several groups studied spin-coated PEO films and reported different results regarding the critical film thickness for crystallization. Schdnherr et al.2 did not detect crystallinity in fihns < 15 nm. However, Reiter et al. 9' 32 reported crystallization in monolayer-thick films (~ 5 nm). The poly(PEGMA) case is structurally different from spin-coated PEO. First, the backbones of comb-like structures usually interfere with crystallization side chain segments adjacent to the polymer backbone, which should lead to a higher critical thickness for crystallization than linear PEO films. (We note that the melting point of the poly(PEGMA) brushes are significantly lower than linear PEO, which suggests a lower thermodynamic stability). In addition, polymer brushes with high graft densities may be viewed as pre-organized for crystallization, but more restricted in terms of available chain conformations. More detailed studies of the crystallization rates will illuminate the importance of these issues. -33- Figure 2.11. Lamellae orientations in the films of poly(PEGMA): (a) Chain axis of PEO side chains is oriented perpendicular to the substrate, lamellae parallel to the surface; (b) Chain axis of PEO side chains is parallel to the substrate, perpendicular lamellae. -89- Crystallization kinetics of polymer brushes. The classical method used to interpret the grth of polymer spherulites is the application of the Avrami-Evans theory.3"‘”39 The Avrami equation is v/v =1 - exp(-Kt ") Eq.]. where v/v is the volume percent crystallinity, t is the crystallization time, K is a crystallization rate constant, and n is the Avrami exponent, a parameter that reflects the dimensionality of crystal growth. For atherrnal nucleation (homogeneous), all nucleation events occur at t = 0, and n can have values of 1, 2, and 3, corresponding to one, two, and three dimensional growth, respectively. For thermal nucleation (heterogeneous), nucleation occurs as a function of time), and n can have values of 2, 3, and 4, which correspond to one, two and three dimensional growth.” 41 The Avrami exponent is obtained by measuring crystallinity as a function of time and analyzing the experimental data using equation 2, a logarithmic version of equation 2.1. log [- In (1- v/v)] = n logt + log K Eq. 2. where n and K are extracted from the slope of a plot of log [- In (1- v/v)] vs. log t and the intercept, respectively. The crystallization of thin polymer brush films is conveniently monitored by optical microscopy, which allows real-time imaging of the growth of individual spherhulites. The volume percent crystallinity in thin films can be approximated as: vc/v~A,/A ~fc7rr2/7rR2 where A, and A are the crystalline area and the total area observed, respectively. For the case of a single spherulite, the radius r is tracked over time and related to R, the radius of the firlly grown spherulite with a degree of crystallinityfi. _90- The crystallization rates for a series of poly(PEGMA) thin films of various thickness were measured using polarized optical microscopy. Prior to crystallization, the films were annealed at 50 °C for 12 h to erase their thermal history. The fihns were immediately inserted into a hot stage purged with N2, quenched to 23 i 2 °C, and the evolution of the film morphology was captured by a video camera. Figure 2.12 shows the growth of spherulites in a 240 nm thick poly(PEGMA) film. The use of the term “spherulite” is technically correct only at the initial stages of crystallization, when the radius of the spherulite is less than the film thickness. This stage lasts only a few seconds and the radii of the spherulites quickly exceed the film thickness. With radii > 50 um, these disk-like objects correspond to a slice through the center of a “spherulite”, a cylindrical crystallite. The “spherulite” growth rates were measured for films of different thicknesses and plotted as a function of time (Figure 2.13a). For each sample, a constant growth rate (G: dR/dt) of 0.02 to 0.07 um/s was observed, which increased with fihn thickness. The average spherulite grth rate is plotted in Figure 2.13b. The growth rate increases with film thickness (regime I), but then appears to level off for fihns >300 nm (regime II). The increase in crystal grth rate is likely due to an increase in the average chain mobility of longer chains, enabling side chains interact with each other and incorporate into the lamellar structure more easily. At film thickness > 300 nm, the effect of the surface- anchoring on chain mobility has a minor influence on crystal growth, resulting in grth rates independent of fihn thickness. -91- Figure 2.12. Optical micrographs showing the growth of spherulites in a 240 nm thick film at 23 t 2 °C. The samples are viewed using crossed polarizers. The scale bar represents 100 um. Images in this dissertation are presented in color. -92- 500 450 _ O 370nm I 310nm 400 - A 250nm E I 210nm 3 35° ‘ o 122nm A 0 g 300 . O 113nm E A 80nm g 250 - in '6 m 200 - .2 E 150 - 100 - 50 - 0 . . . 0 50 100 150 200 Time (min) 8 7 _ Regime | {1‘ I. 161 8-4 1 l + Regime ll Growth rate (x10'2um/s) # o I l l l O 100 200 300 400 500 Thickness (nm) Figure 2.13 a) The evolution of spherulite radii with time for the crystallization of poly(PEGMA) brush films of various thickness at room temperature. b) Spherulite growth rates measured as a function of the poly(PEGMA) film thickness. -93- The linear fit of the Avrami equation to the experimental data is shown in Figure 2.14. For all data sets, the Avrami exponent extracted from the slopes was ~2, which is consistent with either homogeneous nucleation (atherrnal) and two-dimensional growth or heterogeneous nucleation (thermal), and one-dimensional growth (see Table 2.2.). For homogeneous nucleation, the distribution of nuclei in the film should be random, and therefore, the nucleation pattern in a given film should be different each time it is crystallized. For heterogeneous nucleation, secondary nucleation may occur at defect sites, and assuming the sites are not lost, the nucleation pattern should be conserved in a subsequent crystallization. Table 2.2. Avrami exponent n determined from 2-D spherulite grth monitored by in situ polarized optical microscopy. Film thickness (nm) Growth rate (><10'2 um/s) Avrami Exponent n 76 1.7 :t 0.3 2.1 113 2.4 :t 0.2 2.1 210 4 at 1.4 2.2 260 4.7 i 0.7 2.4 290 6.7 :t 0.3 — 310 6.2 i 0.2 2.2 370 6.6 i 0.2 2 -94- To identify the nucleation mode, we heated crystalline films above their melting temperature to erase their thermal history and let the samples cool to room temperature and crystallize. Alter crystallization was complete, we repeated the experiment three times, and followed the crystallization process using polarized optical microscopy. By comparing the images for each melt/crystallization cycle, we found that the distribution of nucleation sites were the same for each cycle, indicating heterogeneous nucleation and 1-D crystallization (radial growth). 0.5 -l- 370nm 0 _ <> 310nm + 0 A I o 250nm +0 ~ ' I 240nm ++° -o 5 ~ 0 38 : 1:1 210nm + . g > '12, A \o X 12201“ + ? .1 - 0 I - .2, c o 80nm + 0 E. -1.5 ~ . X C) O 0 ° I, " . Cl 2 + -2.5 ~ 0 '3 1 r I 2.5 3 3.5 4 4.5 log t Figure 2.14. Linear fit of the Avrami equation to the-experimental data for crystalline polymer brushes with different film thickness. -95- F rank et al. used Avrami analysis to interpret the dimensionality of crystal growth in poly(di-n-hexylsilane)ll and PEO films.2 For the poly(di-n-hexylsilane) film, polymer chains are oriented parallel to the surface, and high T c and low film thickness favor one dimensional growth. For PEO films, the polymer chains were oriented perpendicular to the surface, and the crystal growth was 3-dimensional. The dimensionality of crystal grth is related to the polymer and interface interactions. Several studies have reported the crystallization rates of PEO thin films.2’ 3’ 14’ 18 The results are dependent on the crystallization conditions (T c, supercooling), the substrate, and the molecular weight of PEO. However, linear PEO is a poor model for the crystallization of polymer brushes because of the different architectures of linear and comb polymers as well as the processes involved in main-chain and side-chain crystallization. Therefore, we studied the crystallization behavior of spin-coated poly(PEGMA) thin films for comparison. Poly(PEGMA) was prepared by ATRP of the PEGMA macromonomer in water using a CuBr/CuBr2/bpy catalyst. After a 2 h polymerization at room temperature, the polymerization was exposed to air to terminate the reaction. Purification provided poly(PEGMA) as a white solid, with M" = 60,000g/mol 1H-NMR and IR spectra (see below) confirmed the chemical structure. Thin films were spin-coated onto gold coated silicon wafer substrate using CH2Cl2 as the solvent. After drying under vacuum and annealing at 30 °C, the samples were allowed to crystallize at room temperature. Optical microscopy of the spin-coated film showed a spherulitic morphology with extinction of banded rings, which is possibly due to the twisting and dislocation of lamellae. Like the poly(PEGMA) brush, the orientation of the crystalline lamellae in -96- spin-coated fihns also depended on the film thickness. For a 60 nm thick film, the preferred lamellae orientation is parallel to the surface with the PEO side-chains oriented perpendicular to the surface. FT-IR data from the thick film (290 nm) is consistent with the PEO side-chains parallel to the surface, suggesting an perpendicular lamellae orientation. d+e+f Figure 2.15.1H-NMR spectrum of poly(PEGMA). -97- DA A_ Absorbance 3500 3000 2500 2000 1500 1000 500 Wavenun'ber (cm'1) Figure 2.16. Reflectance FT-IR spectra of 56 and 290 nm thick poly(PEGMA) films that were spin-coated on the gold-coated silicon wafer. -93- The “spherulite” growth rate measured for a spin-coated 300 nm thick poly(PEGMA) film is shown in Figure 2.17. The spin-coated film crystallized twice as fast (0.12 urn/s) as a polymer brush (0.06 prn/s) of comparable thickness. The slower crystalline growth rate for polymer brushes indicates that the decreased mobility imposed by the brush architecture trumps any benefits of chain alignment. Alignment effects may prove to be important only when the density of polymer chains on the surface approaches full coverage. 700 o Spin-coated film -300 nm 600 ~ I Polymer brushes-300 nm 500 ~ 400 - G = 0.12 Irm/s 300 - Radius (um) 200 - 100 i 0 1000 2000 3000 4000 5000 6000 Time (s) Figure 2.17. Crystal grth rates measured for spin-coated poly(PEGMA) and poly(PEGMA) brushes of comparable film thickness (~300 nm). -99- Conclusions. Surface-initiated ATRP of PEGMA macromonomer from gold substrates lead to the nanometer-thck films of comb-like polymer brushes. Polymers with side chains of 22- 23 ethylene oxide units align and crystallize. The film morphology, orientation of crystalline lamellae, and crystal grth rates all depend on the fihn thickness. Polarized optical microscopy and AFM show that the polymer brush films crystallized into spherulitic domains. In addition, both AFM and reflectance FT-IR spectroscopy are consistent with crystalline lamellae preferentially oriented perpendicular to the surface when the film thickness exceeds 100 nm. For thinner films, the favored orientation shifted to crystalline lamellae parallel to the surface, with the PEO side-chains oriented perpendicular to the surface. Crystal growth rates decreased with fihn thickness, and an Avrami analysis of the data indicated heterogeneous nucleation and one-dimensional crystal growth. Polymer brushes crystallized slower than physically absorbed films, likely due to mobility restrictions imposed by the brush architecture. -100- References: 10. 11. 12. 13. 14. Prucker, 0.; Christian, S.; Bock, H.; Ruhe, J .; Frank, C. W.; Knoll, W., Glass transition in ultrathin polymer films. In Organic Thin Films, 1998; Vol. 695, pp 233-249. Schonherr, H.; Frank, C. W., Macromolecules 2003, 36, 1188-1198. Schonherr, H.; Frank, C. W., Macromolecules 2003, 36, 1199-1208. Pearce, R.; Vancso, G J ., Macromolecules 1997, 30, 5843-5848. Pearce, R.; Vancso, G J ., Polymer 1998, 39, 1237-1242. Mellbring, 0.; 0iseth, S. K.; Krozer, A.; Lausmaa, J .; Hjertberg, T., Macromolecules, 2001, 34, 7496-7503. Wang, Y.; Ge, S.; Rafailovich, M.; Sokolov, J .; Zou, Y.; Ade, H.; Luning, J .; Lustiger, A.; Maron, G, Macromolecules 2004, 37, 3319-3327. Sakai, Y.; Imai, M.; Kaji, K.; Tsuji, M., Macromolecules 1996, 29, 8830-8834. Reiter, G; Sommer, J. U., J. Chem. Phys. 2000, 112, 4376-4383. Despotopoulou, M. M.; Frank, C. W.; Miller, R. D.; Rabolt, J. F., Macromolecules 1995, 28, 6687-6688. Despotopoulou, M. M.; Frank, C. W.; Miller, R. D.; Rabolt, J. F., Macromolecules 1996, 29, 5797-5804. Hu, Z. J .; Huang, H. Y.; Zhang, F. J .; Du, B. Y.; He, T. B., Langmuir 2004, 20, 3271-3277. Mareau, V. H.; Prud'homme, R. E., Macromolecules 2005, 38, 398-408. Dalnoki-Veress, K.; Forrest, J. A.; Massa, M. V.; Pratt, A.; Williams, A., J. Polym. Sci. B: Polym. Phys. 2001, 39, 2615-2621. -101- 15. l6. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. Schultz, J. M.; Miles, M. J ., J. Polym. Sci. B; Polym. Phys. 1998, 36, 2311-2325. Taguchi, K.; Miyaji, H.; Izumi, K.; Hoshino, A.; Miyamoto, Y.; Kokawa, R., Polymer 2001, 42, 7443-7447. Marentette, J. M.; Brown, G R., Polymer 1998, 39, 1405-1414. Massa, M. V.; Dalnoki-Veress, K.; Forrest, J. A., Eur. Phys. J. E 2003, 11, 191-198. Zdyrko, B.; Klep, V.; Luzinov, 1., Langmuir 2003, 19, 10179-10187. Zdyrko, B.; Varshney, S. K.; Luzinov, 1., Langmuir 2004, 20, 6727-6735. Neugebauer, D.; Theis, M.; Pakula, T.; Wegner, G; Matyjaszewski, K., Macronolecules 2006, 39, 584-593. Xu, J. T.; J i, J ., Polymer 2003, 44, 6379-6385. Inomata, K.; Nakanishi, E.; Sakane, Y.; Koike, M.; Nose, T., J. Polym. Sci. B: Polym. Phys. 2005, 43, 79-86. Yoshihara, T.; Murahashi, S.; Tadokoro, H., J. Chem. Phys. 1964, 41 , 2902. Miyazawa, T.; Ideguchi, Y.; Fukushima, K., J. Chem. Phys. 1962, 37, 2764-&. Sun, X.; Zhang, H.; Zhang, L.; Wang, X.; Zhou, Q.-F., Polym. J. 2005, 3 7, 102- 108. Richardson, P. H.; Richards, R. W.; Blundell, D. J .; MacDonald, W. A.; Mills, P., Polymer 1995, 36, 3059-69. Wang, M. T.; Braun, H. G; Meyer, E., Polymer 2003, 44, 5015-5021. Hoffmann, C. L.; Rabolt, J. F., Macromolecules 1996, 29, 2543-2547. Bartczak, Z.; Argon, A. S.; Cohen, R. E.; Kowalewski, T., Polymer 1999, 40, 2367-2380. -102- 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. Reiter, G; Castelein, G; Hoemer, P.; Riess, G; Sommer, J. U.; Floudas, G, Eur. Phys. J. E. 2000, 2, 319-334. Reiter, G; Sommer, J. U., Phys. Rev. Lett. 1998, 80, 3771-3774. Shah, R. R.; Merreceyes, D.; Husemann, M.; Rees, 1.; Abbott, N. L.; Hawker, C. J .; Hedrick, J. L., Macromolecules 2000, 33, 597-605. Takahash.Y; Sumita, 1.; Tadokoro, H., J. Polym. Sci. B; Polym. Phys. 1973, 11, 2113-2122. Gabriel, S.; Dubruel, P.; Schacht, E.; Jonas, A. M.; Gilbert, B.; Jerome, R.; Jerome, C., Angew. Chem. Int. Ed. 2005, 44, 5505-5509. Avrami, M., J. Chem. Phys. 1939, 7, 1103. Avrami, M., J. Chem. Phys. 1940, 8, 212. Avrami, M., J. Chem. Phys. 1941, 9, 177. Evans, U. R., Trans. Faraday. Soc. 1945, 41, 365. Gedde, U., Polymer Physics, 1 st Edition. Kluwer Academic Publishers: Dordrecht, 1995. Wunderlich, B., Macromolecular Physics. Academic Press: New York, 1973, 1976; Vol. 1, 2. -103- Chapter 111 Synthesis of Comb-like Polymer Brushes on Planar and Nanoparticulate Surfaces This chapter describes the synthesis and characterization of comb-like polymer brushes on planar and non-planar surfaces using surface-initiated ATRP. Comb-like polymer brushes were formed on gold and silica surfaces by polymerization of PEGMA macromonomers in aqueous media. Polymerization of macromonomers is challenging because their high molecular weight results in low monomer concentrations. Studies of the kinetics of surface-initiated ATRP on planar substrates were used to identify optimum conditions for the growth of PEGMA films. In the second part of this chapter, these polymerizations were used to grow comb-like polymer brushes on the surface of silica nanoparticles. Introduction Poly(ethylene glycol) (PEG) coated surfaces are important in biological and medical applications due to PEG’s ability to resist non-specific protein adsorption and cell adhesion.1 A variety of chemical methods have been developed to deposit strongly adhering PEG coatings on surfaces including self-assembly of PEG-containing thiols on gold,2 coupling reactive groups on PEG with silicon substrates} 4 or the use of priming layers such as poly(glycidyl methacrylate) (PGMA). However, these “grafting to” methods usually result in a low PEG grafting density and thin films. Alternatively, PEG brushes can be grafted from initiators anchored on surfaces through the polymerization of —104- monomers that incorporate PEG segments. The advantage of using comb-like polymer brushes is that dense arrays of PEG side chains are formed in a one-step polymerization. A preference for well-defined brush structures implies their synthesis by a controlled polymerization technique such as ATRP. Zhu et al. 5 used ATRP to grow poly(PEGMA(4-5)) (the numbers refer to the average number of oxyethylene repeat units in the poly(ethylene glycol) methyl ether chain) from silica by ATRP and obtained films < 30 nm thick. Recently, Ma et al. 6 used ATRP to synthesize poly(PEGMA(8-9)) from a gold coated surface, however, the limiting film thickness was only 100 nm after a 16 h polymerization. Obviously, faster polymerizations of such monomers are needed to provide ready access to materials designed to exploit PEG-functionalized surfaces. In the previous chapter, we described the polymerization of PEGMA from gold and studied the crystallization behavior of polymer brushes. In this chapter, we describe optimization of the polymerization conditions for the synthesis of polymer brushes from PEG-substituted methacrylates (Scheme 3.1), with an emphasis on generating thick films on surfaces using short polymerization times. ,CH3 H20=C\ ,c=o (Lt/\oit’nCHa PEGMA 0 CH3 1 m=4-5 ll ' stcrlanIo-c—Qer 2. m=8-9 i? 9H3 [CH3 CH3 3. "1:22.23 siCHzio-C-g—PCHz—C—i-Br (1? CH3 c e c IIB L1 (13' H CH3 >=0n s(CH2-)-o—c+8r U r‘ ”( ) '2' “a" O 11 CH3 H20,l‘.i. ? o +- 3 H3O Scheme 3.1. Surface-initiated ATRP of PEGMA from gold surfaces. -105- Experimental Materials. Unless otherwise noted, all chemicals were obtained from Aldrich. PEGMA(4-5), (Mn = 300 g/mol) and PEGMA(8-9), (Mn = 875 g/mol) were eluted through a basic alumina column to remove inhibitor. PEGMA(22-23) (Mn = 1,100 g/mol), Cu'Br (99.999%), Cu'Cl (9999904101013:2 (99.999%), 1,1,4,7,10,10-hexamethyltriethyl enetetramine (HMTETA), tetramethyl-1,4,8,1l-tetraazacyclotetradecane (Me4Cyclam), 4,4’-di(n-nonyl)-2,2’-bipyridine (anbpy) 10-undecen-1-ol, 2-bromoisobutyryl bromide, pyridine, chlorodimethylsilane (97%) and H2PtC16 were all used as received. Milli-Q water (18 M9) and N,N-dimethylformamide (DMF) (HPLC grade, inhibitdr free) were used as polymerization solvents. Fumed silica (FS) (Aerosil A200), with a primary particle size approximately 12 nm and an average surface area of 200 mZ/g was kindly supplied by Degussa. Characterization methods. Film thicknesses were measured by a rotating analyzer spectroscopy ellipsometer (M-44; J .A. Woollam) at a 75° angle of incidence. The refractive index and film thickness were calculated simultaneously from the values of A and (1). Thermal gravimetric analyses (TGA) of fumed silicas were obtained in dry air using a Perkin Elmer TGA 7 instrument at a heating rate of 10 °C/min. Each sample was held at 110 °C for 30 min to remove adsorbed water prior to initiating the run. Differential scanning calorimetry (DSC) analyses of polymers were obtained using a Perkin Elmer DSC 7. Samples were run in sealed aluminum pans under an N2 atmosphere. The coolant for the system was liquid nitrogen. Each sample was heated from room temperature to 120 °C at a rate of 10 °C/min, held at 120 °C for 30 min, cooled -106- to -90 0C at a rate of -50 °C /min, and then heated to 120 °C at a rate of 10 °C/min. The melting point was taken as the maximum of the melting endotherrn. Polymerization using the CuBr/CuBr/HM TE T A catalyst complex. A catalyst solution was prepared in an oxygen-free drybox by adding CuBr (0.057 g, 20 mmol), CuBr2 (0.045 g, 6 mmol) and HMTETA (0.13 g, 60 mmol) to 20 mL of DMF, and then transferred to a N2-filled glove bag. A second Schlenk flask was loaded with PEGMA(4- 5) (30 g, 0.09 mol) and H20 (15 mL). The flask was connected to a vacuum line and the solution was degassed using three freeze-pump-thaw cycles, back-filled with N2, and then transferred to the N2-filled glove bag. Catalyst solution (5 mL) was added to the second flask, and the resulting blue-green solution was stirred until homogeneous. The final catalyst concentrations were [CuBr] = 2 mM, [CuBr2] = 1 mM and [HMTETA] = 6 mM. The solution was added to a series of vials, each containing one initiator-anchored slide. Slides were removed from vials at predetermined times, ranging from 5 min to 1 h. (Polymerizations longer than 1 h generally became viscous with some gel formation.) The polymer films were rinsed sequentially with water, DMF, and anhydrous ethanol, and finally dried under a stream of N2. Polymerization using CuBrMe4Cyclam/CuBr2(anbpy) 2 as the catalyst. In a drybox, CuBr2 and anbpy (1:2 molzmol) were mixed and the resulting green complex was stored for future use. A catalyst solution was prepared in an oxygen-free drybox by adding CuBr (0.057 g, 20 mmol), CuBr2(anbpy)2 (0.21 g, 10 mmol) and Me4Cyclam (0.1 g, 20 mmol) to 10 mL of DMF. The solution was stirred until homogeneous (~ 20 min). A second Schlenk flask was loaded with PEGMA(4-5) (18 g, 0.05 mol) and H20 (9 mL). The flask was connected to a vacuum line and the solution was degassed using three -107- freeze-pump-thaw cycles, and finally back-filled with N2. In a N2-filled glovebag, a portion of the catalyst solution (3 mL) was added into the flask. The final concentrations were [CuBr] = 2 mM, [CuBr2 (anbpy)2] = 1 mM, and [Me4Cyclam] = 2 mM. The catalyst monomer solution solutions were combined, causing the formation of a brown solution and deposition of a precipitate. The solution was immediately transferred into the vials, each containing an initiator-anchored gold slide. The polymer fihns were taken out from the solution at predetermined times, ranging from 5 min to 1 h, rinsed sequentially with water, DMF, and anhydrous ethanol, and finally dried under a stream of N2. Synthesis of undecen-I-yl, 2-bromo-2-methyl propionate (I).7 A 500 mL round bottom flask was loaded with 10-undecene-l-ol (42.78 g, 0.25 mol), anhydrous THF (250 mL), and 25 mL of pyridine. The flask was cooled to 0 0C, and 2-bromoisobutyryl bromide (24.8 mL) was added drop-wise to the flask. After stirring the reaction mixture at room temperature for 12 h, the solution was filtered to remove pyridinium hydrobromide, and THF was removed under reduced pressure. Hexane (100 mL) was added to the mixture, and the solution was washed with 2N HCl and water (2x). The organic phase was collected and dried over MgS04, filtered, and the solvent was removed by rotary evaporation. Purification by column chromatography on silica using 10/1 (v/v) hexane/ethyl acetate as the eluent provided 1 (Rf = 0.6) as a transparent liquid in 82.4% yield. 1H-NMR: 5 l.2-1.45 (br, 12H), 1.64 (m, 2H), 1.94 (s, 6H), 2.05 (m, 2H), 4.17 (t, 2H), 4.95 (d, 2H), 5.8 (m,1H) ppm. 13C-NMR: 8 171.6 (C=0), 139.1 (CH), 114.1 (CH2), 66.2 (OCH2), 56.01 (C), 32.97 (SCH2), 30.81 (CH3), 29.4 (CH2), 29.3 (CH2), 29.24 (CH2) 29.18 (CH2), 28.4 (CH2), 25.8 (CH2). -108- Synthesis of (11-(2-bromo-2-methyl)propionyloxy)undecyldimethylchlorosilane (2). H2PtC16 (5 mg) and undecen-l-yl, 2-bromo-2-methyl propionate (l) (3.00 g, 9.43 mmol) were added to a dry flask. Chlorodimethylsilane (10.43 mL, 94.3 mmol) was syringed into the mixture and the reaction mixture was stirred at room temperature under N2. The reaction was monitored by 1H-NMR, and alter the reaction was complete, the yellowish solution was diluted with anhydrous toluene, and quickly passed through a column of activated carbon and silica. Removal of the solvent provided (1 l-(2-bromo-2- methyl)propionyloxy)undecyldimethylchlorosilane (2) as a colorless liquid in 60% yield. lH-NMR: 5 0.4 (s, 6H), 0.9 (t, 2H) 1.22-1.42 (br, 16H), 1.8 (m, 2H), 1.93 (s, 6H), 4.16 (t, 2H). 13C-NMR: 5 171.7 (C=0), 66.2 (OCH2), 56.01 (C), 32.97 (SCH2), 30.81 (CH3), 29.4 (CH2), 29.3 (CH2), 29.24 (CH2), 29.18 (CH2), 28.4 (CH2), 25.8 (CH2), 22.1 (CH2), 19.0 (CH2), 1.69 (CH3). 0 _ ‘(CHaie-OH + 131/kg, Pyridine THF 0 1 ‘(CH2)9-o—5-
100 nm/h. Our goal is the synthesis of polymer membranes in minutes using surface-initiated polymerization. Dense films are preferable since they may enhance membrane selectivity. Our earlier results suggest that a more rapid polymerization system requires a low concentration of a highly active catalyst in a concentrated monomer solution. Copper based ATRP catalysts with tetradentate ligands such as HMTETA are more reactive than catalysts with bpy and other bidentate ligands.9 Recently, Bao et al. reported the ultrafast AT RP of tert-butyl acrylate (20 nm/min) from Au and silica surfaces using a CuBr/Me4Cyclam/CuBr2/(anbpy)2 catalyst in DMF/anisole. The same catalyst system might also be useful for the polymerization of PEGMA. Since the CuBr/Me4Cyclam/CuBr2/(anbpy)2 catalyst is sparingly soluble in water, we used DMF as a co-solvent. Despite the formation of a precipitate when a DMF solution of the catalyst was added to the monomer/water mixture, the PEGMA polymerization rate was -119- 300 250 - 0 200 - O 150 i 9 Thickness (nm) 100 — 50‘ 0 r 1 1 1 1 r 0 10 20 30 40 50 60 70 Time (min) Figure 3.6. Evolution of the ellipsometric film thickness with time for the polymerization of PEGMA(4-5) using Me4Cyclam as the ligand. Conditions: [M]:[H2O+DMF] = 1:1; [CuBr] = 2 mM, [CuBr2/(anbpy)2] = 0.6 mM, [MeaCyclam] = 6 mM. -120- 500 450 - 0 400 - . 350 - E 300 — . 3 250 i Q) E l g 200 .C l- 150 ~ 5 100 - 50 ~ ° 0 1 T 0 10 20 30 40 50 60 70 Time (min) Figure 3.7. Evolution of the ellipsometric film thickness with time for the polymerization of PEGMA(4-5) using HMTETA as the ligand. Conditions: [M]/[H2O+DMF] = 1:1, [CuBr] = 2 mM, [CuBr2] = 0.6 mM, [HMTETA] = 6 mM. The data points are the average of two independent polymerizations. -121- very fast, with 100 nm thick films formed in 5 min (Figure 3.6). Using a related ligand, HMTETA, a 5 min polymerization of PEGMA(4-5) using water as a solvent yielded a 60 nm thick film. Remarkably, the linear growth in film thickness with time implies that the polymerization is well-controlled. Figure 3.8. compares PEGMA polymerizations using the HMTETA and MeaCyclam ligands. Both systems rapidly polymerize PEGMA(4-5) (~10 nm/min), however, each system has limitations. Polymerizations using Me4Cyclam were much less controlled, and the rate of polymerization visibly decreased alter 15 min. Polymerization using HMTETA became viscous after 1 h polymerization, which suggests substantial chain transfer. Significant chain transfer has been observed in free-radical10 and photoiniferter mediated11 polymerization of PEGMA. The coefficient for chain transfer to PEG 12 is an order of magnitude higher than MMA (1.2 x 10'5 ),13 and is believed to be dominated by of hydrogen abstraction from the ethylene glycol units. Although increases in solution viscosity, and presumably chain transfer, were occasionally seen in the CuBr/bpy system, it is unclear why chain transfer is favored for the HMTETA catalyst. A complication of the Me4Cyclam system is the marginal solubility of the catalyst. The reaction solution is inhomogeneous and a portion of the CuBr catalyst complex precipitates. When pure DMF is used as the solvent, the HMTETA system fails to generate thick films. Another problem of heterogeneous polymerization is that they provide less uniform films. -122- 500 450 7 9 400 — 350 - 300 2 250 1 200 ~ Thickness (nm) 150- I Q 100— ' 50- 0 1 1 l 1 1 1 0 10 20 30 40 50 60 70 Time (min) Figure 3.8. Evolution of the ellipsometric film thickness with time for the polymerization of PEGMA(4-5) using HMTETA(O) and Me4Cyclam (I) as ligands. The data are re-plotted from Figures 3.6 and 3.7. -123- Synthesis of PE G-functionalized nanoparticles. Particles with PEO-modified surfaces have applications ranging from stationary phases for chromatographic separations to polymer electrolytes in lithium ion batteries.14 Initially, comb-like polymer brushes were grown on the surface of silicon wafers since this substrate facilitates fihn characterization by ellipsometry. Later, PEG-functionalized nanoparticles were prepared by the polymerization of PEGMA from the surface of fumed silica in aqueous media at room temperature. Growth of polymer brushes on silica wafers. Scheme 3.3 shows the synthetic steps used to grow polymer brushes from silicon wafer surfaces. Traditionally, people use trichlorosilanes to anchor initiators and other organic species on SiOz surfaces, but polymerization of trichlorosilanes competes with surface anchoring, complicating quantification of the anchoring step. Therefore, initiators were attached to silicon wafers and silica nanoparticles as monochlorosilanes. To generate initiator layers on $02, clean silicon wafers were placed in an initiator/toluene solution for 24 h. Ellipsometric measurements showed an increase in fihn thickness (20 i 3A) confirming successful attachment of the initiator. Et3N was occasionally used to accelerate the initiator anchoring reaction. Surface-initiated ATRP was catalyzed by a mixture of CuBr/CuBrZ/bpy in aqueous media at room temperature. Figure 3.9 shows the polymerization of PEGMA(4- 5). Thick films (>300 nm) were obtained in 8 h by using a relatively low catalyst concentration ([CuBr] = 2 mM, [CuBr2] = 0.6 mM, [bpy] = 6 mM). The film thickness- time relationship is linear, indicating a controlled polymerization, and a lower catalyst concentration yielded thinner films. Both are consistent with minimal termination. -124- \ i? Cl—S/i—(CH2)11-O—C—
100% 0 l I I T o 1 2 3 4 5 Time (h) Figure 4.3. Evolution of film thickness with time for the copolymerization of PEGMA and BisA-EDMA. The overall concentration of monomer and crosslinker for all runs is 0.4 M. Catalyst: [CuBr] = 1mM; [CuBr2] = 0.3mM; [bpy] = 3mM. The inset shows the mole % composition of BisA-EDMA in the monomer feed solution. -144- 1 _ 0.9- .. ,x’ o 1- , g 0.8 ~ 9. o 0.7 s 0 ,/ 9 0 ,’ .5 0'6 7 j x// .ii. 2 "6 0.5 — f / g g .3 0.4 ‘ i ,’/ “E 0.3 ~ . '3 i x E 02 — ,/ i: 0.1 - ,/ o I I I U I o 02 0.4 0.6 0.8 1 f2, Mole fraction of M2 in comonomer feed Figure 4.4. The relationship between the feed composition and the measured film composition. Each point is the average of at least three independent samples. We are interested in quantifying the residual vinyl groups in the film since they may be used for further elaboration of the membranes. As detailed in the experimental section, comparison of the vinyl C=C and the aromatic C=C stretching bands provide information on the fraction of BisA-EDMA double bonds that were not polymerized The data show that the fraction of BisA-EDMA vinyl bonds polymerized is approximately 80%, and independent of the BisA-EDMA composition in the feed. There was, however, significant run to run variation in films prepared with high crosslinker feed ratios. -145- 1.2 2 .0 3 8 1.0- i .. .. g8 o.a~ ii i T . a i 0 w E S. 0,- 7) 3 ' JL in O. 2 in ' u U *5 g 0.4 — r c 9 . O '3 g 0.2 ~ 9 § 0.0 I I I l o 0.2 0.4 0.6 0.8 1 1‘. f2, Mole fraction M2 in comonomer feed Figure 4.5. The fraction of crosslinkers in which both double bonds participate in the polymerization. Each point is the average of at least three independent samples. The crosslink density can be defined from the data above: M(x - link) . M(BisA — EDMA) M(BisA — EDMA) M(BisA — EDMA) + M(PEGMA) Density(X — link) = M (x - link) Density(X — link) = M (BisA — EDMA) + M (PE GMA) (4.5) The near linear relationship between the feed and measured compositions, and the nearly constant crosslink efficiency (~80%) imply that the crosslink density should be linearly related to the mole fraction of the crosslinker in the feed, as shown in Figure 4.6. Except for the 80 mol% point, the general trend is a linear relationship, which may reflect some variability in membrane synthesis or the limitations of the IR data. -146- 100 l 80 r 70 ~ 60 r 50 — 40 r 30 a 20. ,‘s i 10- Crosslinking density (%) 0 20 40 60 80 100 Feed Crossiinker mo|e% Figure 4.6. Crosslink density vs. the mole percentage of crosslinker in the feed. -147- In addition to estimating the crosslink density, FT-IR spectra also provide information on the morphology of the films. The IR data show that films copolymerized from feed solutions with less than <15 mol% BisA-EDMA are crystalline, but all BisA- EDMA mole fractions >15 mol% produced amorphous films. Swelling behavior of crosslinked copolymer films. Figure 4.7. shows dry and water swollen thicknesses for polymer films prepared at different crosslinker feed ratios. The dry films ranged from 20-30 mm and correspond to compositions of 0% to 100% poly(BisA-EDMA). In Figure 4.7b, the normalized swelling is plotted as a function of crosslinking density to better illustrate the relationship between film structure and swelling behavior. The swelling ratio is defined as d(Wet) ‘ d(d’y) X 100% (4.6) d (dry) Swelling ratio = Where d is the film thickness. As shown in Figure 4.7b, the swelling ratio decreased with cross-link density and pl ateaued at >40% crosslink density. The former indicates increasing dimensional Stability with crosslinking, which is characteristic of crosslinked films, while 160% SVVelling may represent the inherent swelling due to the long PEO segments in the two I11¢Z>nomers. Qiao et al. 10 observed the increased swelling of PEG dimethacrylate network as the length of PEG segments increased. -148- 2000 I in water 1600 0 dry state 1200 * 800 1 I Film thickness (A) 400 3 0 20 40 60 80 100 Crosslinker feed percentage ( mo|°/o) 300 , 280 l 260 - 24o - 220 - 200 - 130 - 160 - , 14o - 120 - 1 00 T . . . 0 20 4O 60 8O 1 00 Crosslinking density (%) Swelling percentage (%) 99 o e Figure 4.7. a) (Top) thickness of dry and water-swollen films as a function of the mole percentages of crosslinker in the monomers feed solution; b) (bottom) the 1‘EIationship between the normalized swelling of films and the film crosslink density. -149- Conclusion and Outlook: Crosslinked polymer brushes containing PEO segments were grown from the surface of gold substrates using surface-initiated ATRP. The addition of the cross-linker to PEGMA slowed the polymerization rate. Using the aromatic ring of BisA-EDMA as a reference, the film structure and crosslink density were measured by an analysis of the IR spectra of crosslinked films. Aqueous swelling of copolymer films in water decreases with increasing crosslink density, and is nearly constant for crosslink densities >40%. The gas permeation characteristics of crosslinked-polymer membranes grafted on porous alumina supports are described in Chapter V. -150- References: Nolan, C. M.; Reyes, C. D.; Debord, J. D.; Garcia, A. J .; Lyon, L. A., Biomacromolecules 2005, 6, 2032-2039. Groll, J .; Amirgoulova, E. V.; Ameringer, T.; Heyes, C. D.; Roecker, C.; Nienhaus, G. U.; Moeller, M., J. Am. Chem. Soc. 2004, 126, 4234-4239. Lin, H.; Wagner, E. V.; Freeman, B. D.; Toy, L. G.; Gupta, R., Science 2006, 311, 639-642. Heggli, M.; Tirelli, N.; Zisch, A.; Hubbell, J. A., Bioconjugate Chem. 2003, 14, 967-973. Matsumoto, A.; Ohashi, T.; 0e, 11.; Aota, H.; Ikeda, J. I., J. Polym. Sci. A: Polym. Chem. 2000, 38, 4396-4402. Baker, J. P.; Hong, L. H.; Blanch, H. W.; Prausnitz, J. M., Macromolecules 1994, 27, 1446-1454. Elliott, J. E.; Macdonald, M.; Nie, J .; Bowman, C. N., Polymer 2004, 45, 1503- 1510. Kizilay, M. Y.; Okay, O., Macromolecules 2003, 36, 6856-6862. Odian, G., Principles of Polymerization. Wiley-Interscience: New York, 2004. Qiao, C.; Jiang, S.; Dong, D.; Ji, X.; An, L.; Jiang, B., Macromol. Rapid Commun. 2004, 25, 659-663. -151- Chapter V Polymer Membranes for Gas Separation This chapter describes the development of polymeric membrane materials and studies of gas permeation through membranes, emphasizing the separation of CO2 from light gases such as H2 and N2. Comb and crosslinked polymer brushes containing PEO units were grafted to the surface of porous alumina membrane supports by surface- initiated ATRP. Due to the interactions between CO2 and the ether oxygens of the ethylene oxide units, the modified membranes are highly permeable to CO2 and selectively permeate CO2. For poly(PEGMA) films, the measured permeability of CO2 is ~300 Barrer with a C02 /H2 selectivity of 14. CO2 permeability and selectivity decreased as the PEO side chain of the comb polymer was shortened. CO2 and H2 permeability, and the CO2/H2 permselectivity decreased as more crosslinker was incorporated in the film. Background During the last decade, gas separation by polymeric membranes has been widely studied as an alternative to traditional gas separation technology. The emphasis of this research has focused on the recovery of H2 from petrochemical process streams, removal of CO2 from hydrocarbon mixtures, and removal of acidic gases from natural gases. 1 The steady state permeability of gas A, through a film with thickness of l, is defined as:2 -152- where FA is the steady state flux of gas through the film (cm3 (STP)/cm2 s), l is the film thickness and Ap is trans-membrane pressure (cmHg). Permeability coefficients are commonly expressed in units of Barrers, where 1 Barrer = 10'10 cm3 (STP) cm/(cm2 s cmHg). When the upstream pressure (p) is much higher than the downstream pressure (p2), Ap can be approximated as pi. The permeability of gas A through a polymer matrix can be expressed as the product of the average effective diffusivity, DA, and the apparent solubility of gas A in the polymer, S A? PA: DA X 3A (2) According to this diffusion/solubility model, a penetrant gas dissolves in the high pressure side of the film, diffuses through the film, and then desorbs at the low pressure side of the film. The ideal membrane selectivity for gas A over gas B is expressed as: on (NR) = PA/PB = (DA/DB) x (SA/ SB) = a r) x or s (3) where or D is the diffusivity selectivity, and a s is the solubility selectivity. The gas diffusivity increases with decreasing penetrant (gas) size (lower critical volume or kinetic diameter of the penetrant), increasing polymer chain flexibility (increased free volume), and decreasing polymer chain and penetrant interactions. The penetrant solubility increases with increasing condensability (higher boiling point or critical temperature) and stronger gas-polymer interactions.3’ 4 As shown in Table 5.1, the critical volume and kinetic diameter (dk) are used as a metric for penetrant size. Considering only molecular size, H2 is smaller and has a higher diffusion coefficient than CO2. However, CO2 has a higher boiling point and critical temperature than H2, and therefore, it is more condensable than H2. -153- Many commercially available membranes are based on glassy polymers, and are designed to maximize the gas diffusivity selectivity. Those membranes are more permeable to H2 than to CO2, since an favors H2 much more strongly than as favors CO2. One way to improve the permeability of glassy membranes is to increase chain mobility by adding a plasticizer. However, plasticization of glassy polymers would be detrimental to membrane performance. When the polymer absorbs a high concentration of a condensable gas (CO2), the polymer swells and softens. The diffusion coefficient of large molecule (CO2) increases more rapidly than small molecule (H2), resulting in the loss of diffusion selectivity. Table 5.1. Physical properties of H2, C02, and N25, 6 Size Condensibility S°lub1my paramenter Critical Boilin oint Critical Volume (1 1, (nm) (13p temperature 8 (MPa 0'5) (cm3/mol) (K) H2 65.1 0.289 20.4 33.2 6.6 N2 89.8 0.364 77.4 126.2 5.3 CO2 93.9 0.330 195 304.2 12.2 Since the diffusivity ratio OLD [=D(CO2)/D(H2)] is unfavorable, a common strategy for separating CO2 and H2 is to design rubbery membranes that have a high solubility ratio as [= (S(CO2)/ S(H2)]. A high solubility for CO2 relative to light gases also is very useful for CO2/N 2 separation, since their sizes and critical volumes are similar (see Table 5.1.). The solubility of a gas depends on the physical properties of the gas, such as boiling point and critical temperature, but more importantly, solubility is related to polymer/gas -154- interactions. PEO is perhaps the best material for the separation of CO2 due to favorable interactions between the ether oxygens and CO2. Membranes more permeable to CO2 than H2 are called “reverse selective” membranes. Plasticization, which is detrimental to the glassy polymers, actually benefits PEO-containing reverse selective membranes, since it increases D(CO2)/D(H2), yielding a higher CO2/H2 selectivity. High molecular weight PEO is semi-crystalline. The formation of a crystalline phase would reduce the CO2 gas permeability because only ethylene oxide segments in the amorphous phase could interact with CO2 and solubilize CO2. There are several strategies for incorporating PEO segments into polymers while avoiding PEO crystallization including the use of low molecular weight PEO, block copolymers with short PEO blocks, graft and comb polymers with short PEO segments as side chains, and cross-linked PEO networks. PEO-containing monomers commonly used to synthesize such structures are shown in Scheme 5.1. Table 5.2 lists CO2 permeabilities and CO2/H2 selectivities for representative membrane materials. More CO2/N 2 separation data appear in reference 10. -155- _ ,CH3 Hzc—q lc=o CH Oi~/\0’)j1 3 1. n=4-5 PEGMA 2.n=88 3. n = 22-23 Wofiok: 6. n=4 7. n=14 trio-ick- 4. n=4 PEGDMA 5. n = 14 O TYVO ”O O 0 30949: 8. BisA-EDMA Scheme 5.1. Structures of representative PEG (meth)acrylates and PEG di(meth)acrylates used in this study and reported in the literature for membrane separations. -156- Table 5.2. Representative examples of pure gas CO2 permeability and CO2/H2 selectivity in PEO-containing polymers. Permeability and Selectivity P(CO2) 01(C02/H2) ° Reference b Polymer (Barrer) Semi-crystalline PEO 13 (35 °C) 6.7 (1 atm) Lin7 Liquid PEG blend with solid PDMS N/A 7 .5(25 °C) Lin 8 PEG soft block and polyimide hard block 66 (35 °C) 7.8 (35 °C) Bonderg' ‘0 120 (35 °C) 9.8 (35 °C) Crosslinked 50 wt% 2/ 50 wt% 7 a 260 (35 °C) 10 Freeman 8 PEO films 70 wt% 2 / 30 wt% 7 320 (35 °C) 11 (35 °C) 40 (-20 °C) 99 wt% 2 570 (35 °C) 12 Crosslinked PEG di-acrylate (7) 70 (23 °C) 9.5 Patel ”"3 ol ethers p y 50/50 w/w 100 5 PEG di-acrylate/PPG diacrylate a The number refers to the structures in Scheme 5.1. b Permeability coefficient: lBarrer =10-10 cm3 (STP) cm/(cm2 s cmHg). ° Selectivity 0i (CO2/H2) = P (CO2)/ P(H2). -157- Our strategy is to develop composite polymer membranes in which a thin polymer skin is deposited on the surface of a porous aluminum membrane support. Solution casting,14 and solution coating,15 plasma polymerization 16‘ ‘7 and UV-induced grafting,'8’ 19 are the most common ways to prepare composite membranes. The principal challenge in these methods is the difficulty in controlling film thickness. Surface-initiated ATRP allows controlled growth of polymer brushes on the surface, and it may provide access to membranes with well-defined polymer brush skins. In this chapter, we describe the grafting of PEO-containing combs and crosslinked polymer brushes from the surface of porous alumina membrane supports, and examine the CO2, H2 and N2 gases permeability through these membranes. Experimental Materials. Poly(ethylene glycol) methyl ether methacrylate (PEGMA, M,, = 300 g/mol and 1,100 g/mol), Cu(I)Br (99.999%), Cu(II)Br2 (99.999%), tetramethyl-1,4,8,l 1- tetraazacyclotetradecane (Me4Cyclam), 4,4’-di-(n-nonyl)-2, 2’-bipyridine (anbpy) were used as received from Aldrich. Bisphenol A ethoxylate dimethacrylate (BisA-EDMA, M,, = 1,700 g/mol) was passed through a column packed with a commercial inhibitor removal agent (Aldrich) to remove hydroquinone inhibitor. Milli-Q water (18 M0), N,N- dimethylformamide (DMF) (HPLC grade, inhibitor free) were used as polymerization solvents. Characterization methods. Field-emission scanning electron microscopy (FESEM, Hitachi s-47OOII, acceleration voltage of 15 kV) was used to characterize the film grth on the porous aluminum membrane support. Membranes were freeze- -158- fractured using liquid N2 prior to sputter coating with 5 nm gold for cross-sectional image analysis. Anchoring initiators on the surface of porous alumina membrane support. Porous alumina membranes (0.02 pm) were cleaned in an UV/O3 chamber for 15 min, washed with anhydrous ethanol and dried with N2. The membranes were sputter-coated with 5 nm gold, and then were cleaned in the UV/O3 chamber for an additional 10 min, washed with ethanol and dried under a stream of N2. A self-assembled initiator monolayer was formed by immersing the gold-coated membrane in a 1 mM ethanolic solution of [Br-C(CH3)2-COO(CH2)ioS]2 for 24 h. The films were washed with a large amount of ethanol, sonicated in ethanol for 1 min, washed again with ethanol, and finally dried under a stream of N2. To directly attach initiators to the alumina membrane, the membrane was immersed in a DMF solution of the trichlorosilane initiator (see chapter 111) for 24 h, and then the membrane was washed with DMF and anhydrous ethanol, and dried under a stream of N2 stream. The synthetic scheme is illustrated in Figure 5.2. Surface-initiated A T RP of poly(PE GMA ) films using CuBr/CuBr/bpy catalyst. A catalyst solution was prepared in an oxygen-free drybox by adding CuBr (0.057 g, 20 mM), CuBr2 (0.045 g, 6 mM) and bpy (0.18 g, 60 mM) to 20 mL of DMF, and then transferring the solution to a N2-filled glove bag. PEGMA (15 g) and Milli Q water (14 mL) were added to a Schlenk flask and stirred until homogeneous. The flask was connected to a vacuum line and the solution was degassed using three freeze-pump-thaw cycles, back-filled with N2, and then transferred to the N2-filled glove bag. A portion of the catalyst solution (1.5 mL) was added to the flask, resulting in catalyst concentrations of [CuBr] = 1 mM, [CuBr2] = 0.3 mM, and [bpy] = 3 mM. The solution was transferred -159- into a container holding an initiator-anchored membrane. After 12 h, the membrane was removed from polymerization solution, rinsed with large amount of water and DMF, cleaned in an ultrasonic bath with DMF, washed with anhydrous ethanol, and finally dried under a stream of N2. Polymerization of PE GMA using a C uBr/lile4C yclam/CuBr2(anbpy) 2 catalyst. In a drybox, CuBr2 and anbpy (1:2, molzmol) were mixed together to form a green complex and stored for future use. A catalyst solution was prepared in an oxygen-free drybox by adding CuBr (0.057 g, 20 mM), CuBr2(anbpy)2 (0.21 g, 10 mM) and Me4Cyclam (0.1 g, 20 mM) to 10 mL of DMF. The solution was stirred until homogeneous (~ 20 min). A second Schlenk flask was loaded with PEGMA(4-5) (18 g, 0.05 mol) and H20 (9 mL). The flask was connected to a vacuum line and the solution was degassed using three freeze-pump-thaw cycles, and finally back-filled with N2. In a N2-filled glovebag, a portion of the catalyst solution (3 mL) was added into the flask. The final catalysts concentrations were: [CuBr] = 2 mM, [CuBr2] = [(anbpy)2] = 1 mM, and [Me4Cyclam] = 2 mM. The catalyst solution was added to the monomer solution, causing it to turn brown with the formation of a precipitate. The solution was immediately transferred into a vial containing the initiator-anchored membrane. The membrane was taken out at a predetermined times, ranging from 3 min to 10 min, to control the film thickness. The membrane was rinsed with a large amount of water and DMF, cleaned in an ultrasonic bath with DMF, washed with anhydrous ethanol, and finally dried under a stream of N2. Growth of crosslinked copolymerfilm. A stock catalyst solution was prepared in a drybox by adding CuBr (0.043 g, 10 mM), CuBr2 (0.02 g, 3 mM), and bpy (0.14 g, 30 -l60- mM) to 30 mL of DMF. The solution was sealed in a flask and transferred to a N2 filled glovebag. Monomer (PEGMA), crosslinker (BisA-EDMA) and water were added to a Schlenk flask and stirred until homogeneous. The total concentration of monomer and crosslinker was fixed at 0.4 M, and the monomer to crosslinker ratio was varied to obtain the desired polymer brush composition. For a polymerization with 30 mol% crosslinker, 15.4 g of PEGMA (0.014 mol), 10.8 g of BisA-EDMA (6 mmol), and 19 mL water were added to a Schlenk flask. The solution was degassed via three freeze-pump-thaw cycles, and back-filled with N2. A portion of the stock catalyst solution (5 mL) was added to the mixture under N2, resulting in final catalyst concentrations of [CuBr] = 1 mM, [CuBr2] = 0.3 mM, and [bpy] = 3 mM. The solution was stirred until homogeneous and then was transferred into vials containing the initiator-anchored membranes. After a 12 h polymerization, membranes were removed from the solution, rinsed with a copious amount of water, ultrasonically cleaned in a DMF bath, washed with ethanol, and finally dried under a stream of N2. Gas permeation measurements. The permeation of gases (CO2, H2, and N2) through the modified membranes was measured using a permeation cell with a pressure relief valve. The permeate flux was measured as a function of gas inlet pressure (5-50 psig) using a soap-bubble flow meter. Measurements were performed for each gas separately in the order H2, N2, and CO2. For each gas, the permeation cell was purged with the gas several times to remove traces of other gases and to establish a stable gas flux. The flowmeter readout was converted to the flux using Eq 4: F=v/A (4) -l61- where F is the flux of the gas (cm3 (STP)/cm2 s), v is the flow rate of the gas (lemin), and A is the area of the membrane (cmz). Permeability coefficients were calculated according to Eq. 1. N, Q @ Erliii co2 H2 Gas controller 12 . 34 Pressure gauge permeation cell '— i—1 L Pressure 7" *4 ‘4 relief valve Digital Flow Meter Polymer membrane Figure 5.1. Apparatus for gas permeation measurements. -162- Results and Discussion Synthesis of and characterization of polymer films Polymer brushes were grafted from the surface of alumina membrane supports (0.02 pm pore size) using two methods. In the first, the membrane was sputter coated with 5 nm of gold, and then the thiol terminated initiator was attached by forming a self- assembled monolayer. Alternatively, trichlorosilane initiator precursors were directly coupled with OH groups to form Si-O-Si bonds. Using surface-initiated ATRP, PEGMA macromonomers were polymerized from the initiator layers to form comb polymer brushes with PEO side chains. Including a cross-linker (BisA-EDMA, 30-100 mol%) in the monomer pool allowed the growth of cross-linked films with various crosslink densities on the surface of porous alumina membranes. Figure 5.2 shows transmission FT-IR spectra of polymethacrylates grafted on the surface of a porous alumina membrane support. Characteristic peaks observed in the spectrum, include C-H stretching at 2885 cm"1 and ester C=O stretching at 1729 cm". The lengths of the PEO side chains in the two polymers are different, hence the different intensity ratios seen for C-H stretching and ester C=O bands. The ethylene oxide C-O-C stretching band is obscured by the Al-O-Al stretching band, located at ~1110 cm". Because of the high pore density in the alumina substrates, ellipsometric measurements Were not possible on these surfaces, and thus film thicknesses were estimated from the Cross-sectional FESEM images. To demonstrate the potential of these materials in gas Separations, we examined CO2, N2 and H2 fluxes through these films grown on the Surface of porous alumina membranes. -l63- Porous alumina U U U U U U membrane support Cl\ 9 CI—Si—(CH2)11—O-C—
40%. Freeman and co- workers have pointed out that changes in gas diffusion and permeability in crosslinked polymer systems are more related to changes in the glass transition temperature (T S) than crosslink density changes.20 In the BisA-EDMA/PEGMA system, introduction of the crosslinker increased the polymer T 3 due to the rigid benzene rings of BisA-EDMA, which likely reduced the free volume of the polymer, and hence affected the gas diffusion through the membrane. -175- agape? A<2Qm-