fig .. . 3 IL; 1“. . . ma... 9% m: , e .541 .‘.{.vtl. ,1 w DRILL 1 3g.fiqv-§vuw ‘ . 2”: Nu. on". i . ’ mm .2 5443 :530 LIBRARY MiChigan State Th‘ ' t rt'fyth tth - - ISIS oce l a e Ul‘llVG rsrty dissertation entitled SELECTIVE, ULTRATHIN MEMBRANE SKINS PREPARED BY DEPOSITION OF NOVEL POLYMER FILMS ON POROUS ALUMINA SUPPORTS presented by Anagi Manjula Balachandra has been accepted towards fulfillment of the requirements for the Ph.D. degree in Chemistry fl ILL-u Lust/‘7 7 Major Professor’s We 1. / i r ' 5/ .1 o/ A on 3 ' I Date MSU is an Affirmative Action/Equal Opportunity Institution 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 OCT 2 II 2005 6/01 c:/CIRC/DateDue.p65-p. 15 SELECTIVE, ULTRATHIN MEMBRANE SKINS PREPARED BY DEPOSITION OF NOVEL POLYMER FILMS ON POROUS ALUMINA SUPPORTS By Anagi Manjula Balachandra A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHOLOSOPHY Department of Chemistry 2003 ABSTRACT SELECTIVE, ULTRATHIN MEMBRANE SKINS PREPARED BY DEPOSITION OF NOVEL POLYMER FILMS ON POROUS ALUMINA SUPPORTS By Anagi Manjula Balachandra Membrane-based separations are attractive in industrial processes because of their low energy costs and Simple operation. However, low permeabilities ofien make membrane processes uneconomical. Since flux is inversely proportional to membrane thickness, composite membranes consisting of ultrathin, selective skins on highly permeable supports are required to Simultaneously achieve high throughput and high selectivity. However, the synthesis of defect-free skins with thicknesses less than 50 nm is difficult, and thus flux is often limited. Layer-by-layer deposition of oppositely charged polyelectrolytes on porous supports is an attractive method to synthesize ultrathin ion-separation membranes with high flux and high selectivity. The ion-transport selectivity of multilayer polyelectrolyte membranes (MPMS) is primarily due to Donnan exclusion; therefore increase in fixed charge density should yield high selectivity. However, control over charge density in MPMS is difficult because charges on polycations are electrostatically compensated by charges on polyanions, and the net charge in the bulk of these films is small. To overcome this problem, we introduced a templating method to create ion-exchange sites in the bulk of the membrane. This strategy involves alternating deposition of a Cu”- poly(acrylic acid) complex and poly(allylamine hydrochloride) on a porous alumina support followed by removal of Cu” and deprotonation to yield free -COO' ion-exchange sites. Difffusion dialysis studies Showed that the Cl'/SO42' selectivity of Cu2+-templated membranes is 4-fold higher than that of membranes prepared in the absence of Cu”. Post-deposition cross-linking of these membranes by heat-induced amide bond formation further increased Cl‘/SO42' selectivity to values as high as 600. Room-temperature, surface-initiated atom transfer radical polymerization (ATRP) provides another convenient method for formation of utrathin polymer Skins. This process involves attachment of polymerization initiators to a porous alumina support and subsequent polymerization from these initiators. Because ATRP is a controlled polymerization technique, it yields well-defined polymer films with low polydispersity indices (narrow molecular weight distributions). Additionally, this method is attractive because film thickness can be easily controlled by adjusting polymerization time. Gas- permeability data showed that grafted poly(ethylene glycol dimethacrylate) membranes have a COz/CH4 selectivity of 20, whereas poly(Z-hydroxyethyl methacrylate) G’HEMA) films grown from a surface have negligible selectivity. However, derivatization of PHEMA with pentadecafluorooctanoyl chloride increases the solubility of C02 in the membrane and results in a COZ/CH4 selectivity of 9. Although composite PHEMA membranes have no significant gas-transport selectivity, diffusion dialysis studies with PHEMA membranes showed moderate ion- transport selectivities. Cross-linking of PHEMA membranes by reaction with succinyl chloride greatly enhanced anion-transport selectivities while maintaining reasonable flux. The selectivities of these systems demonstrate that alternating polyelectrolyte deposition and surface-initiated ATRP are indeed capable of forming ultrathin, defect-free membrane skins that can potentially be modified for specific separations. Copyright by ANAGI MANJULA BALACHANDRA 2003 To my dearest parents, Lalini and Henry Balachandra and my husband, Dhammika. Your endless support and encouragement made this dissertation and doctoral degree possible. ACKNOWLEDGMNETS I would like to extend my deep appreciation and gratitude to my adviser, Prof. Merlin Bruening, for his excellent guidance, support and encouragement. I was really lucky to be in your group and enjoyed working with you. Thank you for being such a wonderful adviser and a mentor. I would also thankful to Prof. Gregory Baker for his expertise advices during the polymerization project. I am also thankful for my guidance committee members; Profs, Pinnavaia, Watson and Broderick for their time and suggestions. I greatly acknowledge the financial support from the NSF Center for Sensor Materials at Michigan State University. I am very much thankful to Bruening group members past and present; Dan, Jeremy, Wenxi, Milind, Jinhua, Kangping, Yinda, Skanth, Sandra, Keith, Jacque, Bo Young, Matt, Brian. J in, Lei, Sri, Christin and Xiaoyun. Especially, to Dr. Wenxi Huang for teaching me many of the polymerization techniques. I also want to thank Baker group members I B, Ying and Bao, it was very nice working with you. Finally, I like to thank my family, especially, my parents for all their support, encouragement and love. I owe you very much for every Single achievement in life. Biggest thanks go to my husband, Dhammika for his constant support, understanding and love. It want be possible for me to finish my Ph.D. without you. vi TABLE OF CONTENTS List of Schemes List of Tables List of Figures List of Abbreviations Chapter 1: Introduction 1.1 Structure of the Dissertation 1.2 Formation and Development of Thin Polymer Films 1.3 Multilayer Polyelectrolyte Films 1.4 Ion Permeation through MPFS 1.5 MPFS as Ion-separation Membranes 1.6 Grafted Polymer Brushes 1.7 Atom Transfer Radical Polymerization (ATRP) 1.8 Formation of membranes by surface-initiated ATRP 1.9 Gas-Separation Membranes 1.10 Mechanism of Gas Transport through Polymer Membranes 1.11 Summary 1.12 References Chapter 2: Enhancing the Anion-Transport Selectivity of Multilayer Polyelectrolyte Membranes by Templating with Cu2+ 2.1 Introduction 2.2 Experimental 2.2.1 Chemicals and Solutions 2.2.2 Film Preparation 2.2.3 Film Characterization 2.2.4 Ion-Transport Studies 2.3 Results and Discussion 2.3.] Synthesis and Characterization of Cu2+-Templated PAA/PAH Films 2.3.2 Anion Transport through Cu2+-Templated PAA/PAH Membranes 2.3.3 Cross-linked Cu2+-Templated PAA/PAH Membranes 2.3.4 Changing the Surface Charge of Membranes 2.3.5 Anion-Transport through Partially Cross-linked, Cu2+-Templated Membranes Deposited at Different pH values 2.3.6 Anion-Transport through Partially Cross-linked, Cu +-Templated Membranes Deposited with Different Cu2+ Concentrations 2.3.7 Modeling of Anion-Transport through Cu2+-Templated PAA/PAH Membranes 2.4 Conclusions 2.5 References and Notes vii xi xii xvi 36 36 39 39 4O 41 42 44 44 47 50 53 54 56 59 68 69 Chapter 3: Preparation of Composite Membranes by Atom Transfer Radical Polymerization Initiated from a Porous Support 3.1 Introduction 3.2 Experimental 3.2.1 Chemicals and Solutions 3.2.2 Polymerization of EGDMA 3.2.3 Polymerization of HEMA and Subsequent Derivatization 3.2.4 Film Characterization on Gold Wafers 3.2.5 Film Characterization on Alumina Supports 3.2.6 Gas-Permeation Studies 3.3 Results and Discussion 3.3.1 Synthesis and Characterization of PEGDMA films 3.3.2 Synthesis and Characterization of PHEMA films 3.3.3 Gas Permeation through Polymer Membranes 3.3.4 Gas Permeation through Composite PEGDMA Membranes 3.3.5 Gas Permeation through PHEMA Membranes 3.4 Conclusions 3.5 References and Notes Chapter 4: Ion Transport through Grafted Poly (2-hydroxyethyl methacrylate) Membranes and their Derivatives 4.1 Introduction 4.2 Experimental 4.2.1 Chemicals and Solutions 4.2.2 Polymerization of HEMA from Gold-coated wafers and Porous Alumina 4.2.3 Derivatization of PHEMA with Crown ethers 4.2.4 Chemical Cross-linking of PHEMA 4.2.5 F ilrn Characterization 4.2.6 Ion-Transport Studies with PHEMA Membranes 4.3 Results and Discussion 4.3.1 Synthesis and Characterization of PHEMA Membranes 4.3.2 Ion-Transport Studies with Composite PHEMA Membranes 4.3.3 Chemically Cross-linked PHEMA Membranes and their Ion-Transport Properties 4.3.4 Crown Ether-Derivatized PHEMA Membranes and their Ion-Transport Properties 4.4 Conclusions 4.5 References viii 73 73 75 76 78 80 80 80 82 82 87 89 94 98 108 109 113 113 115 115 116 116 117 117 117 118 118 119 122 129 134 135 Chapter 5: Conclusions and Future Work 5. 1 Conclusions 5.2 Suggestions for Future Work 5.3 References ix 136 136 138 140 Scheme 1.1 Scheme 3.1 Scheme 3.2 Scheme 3.3 Scheme 4.1 Scheme 4.2 LIST OF SCHEMES Mechanism of ATRP. Growth of cross-linked PEGDMA from an anchored initiator. Growth of PHEMA brushes from a gold surface: (1) formation of a monolayer of initiator and (2) polymerization from the initiator-modified gold surface. Derivatization of PHEMA with perfluorooctanoyl chloride. Cross-linking of PHEMA by reaction with succinyl chloride. F unctionalization of PHEMA with 2-hydroxy methyl 18-crown-6. 17 86 91 102 124 132 LIST OF TABLES Table 2.1 Anion fluxes (moles cm'zs'l) through bare porous alumina and alumina coated with PAA/PAH and PAA-Cu/PAH films cross-linked at different temperatures. 51 Table 2.2 Thicknesses, fluxes (moles cm‘zs’l), selectivities and Cu2+ concentrations (M) in partially cross-linked, 10.5-bilayer PAA-Cu/PAH films deposited at different pH values. 55 Table 2.3 Anion Fluxes (moles cm’zs’l), selectivities, thicknesses, and estimated Cu2+ concentrations (M) of partially cross-linked (130 °C) 10.5-bilayer PAA-Cu/PAH membranes and films deposited using different Cu + concentrations. 58 Table 2.4 Diffusion coefficients obtained from modeling ion transport through 10.5-bilayer PAA-Cu/PAH membranes cross-linked at different temperatures. 67 Table 3.1 Gas permeability coefficients, estimated film thicknesses, and idea selectivities for PEGDMA membranes. 96 Table 3.2 Gas Permeability Coefficients, Estimated Film Thicknesses and Idea Selectivities for Underivatized PHEMA Membranes 100 Table 3.3 Estimated film thicknesses, gas permeability coefficients and ideal selectivities for fluorinated PHEMA membranes 107 Table 4.1 Film thicknesses, fluxes and selectivities for composite PHEMA membranes (porous alumina support). 121 Table 4.2 Hydrated radii and hydration energies of various ions. 123 Table 4.3 Fihn thicknesses, fluxes and selectivities for cross-linked PHEMA membranes in diffusion dialysis with single-salt solutions. 128 Table 4.4 Film thicknesses, normalized fluxes and selectivities for cross-linked PHEMA membranes in diffusion dialysis with mixed salt solutions. . 130 xi Figure 1.1 Figure 1.2 Figure 1.3 Figure 1.4 Figure 1.5 Figure 1.6 Figure 2.1 Figure 2.2: Figure 2.3 Figure 2.4 LIST OF FIGURES Deposition of a Langmuir-Blodget film from a floating Langmuir monolayer. Figure adapted from. Schematic diagram of the “Dip and Rinse” procedure for alternating deposition of oppositely charged polyelectrolytes on a charged substrate. In the actual film structure, polycations and polyanions are interdigitated. Synthesis of Cu2+-templated polyelectolyte films. Step 1: deposition of a polycation (PAH) in the presence of uncomplexed Cu2+. Step 2: deposition of partially Cu2+ -complexed PAA. Step 3: removal of Cu2+. Step 4: deprotonation of the free carboxylic acid groups of PAA. Repetition of steps 1 and 2 produes a multilayer film. Inertwining of layers is not shown for figure clarity. Two methods for grafting of polymer films onto solid surfaces. Cartoon showing the different stages of surface-initiated ATRP. Surface initiated atom transfer radical polymerization from (a) polyelectrolyte deposited on alumina and (b) Au-coated alumina. Preparation of Cu2+- templated polyelectrolyte films on porous alumina supports. Step l-adsorption of partially Cu2+-complexed PAA on porous alumina. Step 2-adsorption of a polycation (PAH). Step 3- removal of Cu“. Step 4-deprotonation of the free carboxylic acid groups of PAA. Repetition of steps 1 and 2 produces multilayer films. lntertwining of layers is not shown for figure clarity. Apparatus for ion-transport measurements. Cyclic voltammetry of a MPA-modified gold electrode coated with10 bilayers of PAH/PAA-Cu before (solid line) and after exposing to pH 3.5 water (dotted line). External reflectance FTIR spectra of (a) a 10-bilayer PAH/PAA-Cu film, (b) the same film after exposure to pH 3.5 water, (c) the fihn after exposure to pH 5-6 water, and (d) a 10-bi1ayer PAH/PAA film deposited without Cu2+. All films were deposited on a gold wafer coated with a monolayer of MPA. xii 13 15 19 21 37 43 46 48 Figure 2.5 Figure 2.6 Figure 2.7 Figure 2.8 Figure 2.9 Figure 3.1 Figure 3.2 Figure 3.3 Receiving phase concentration as a function of time for a bare porous alumina membrane (black), an alumina membrane coated with 10.5 bilayers of Cu2+-templated FAA/PAH (open) and an alumina membrane coated with 10.5 bilayers of FAA/PAH (grey). Different symbols represent different salts: circles-NaCl and squares-NaZSO4. The inset shows 8042' concentration vs. time for pure PAA/PAH (grey) and templated PAA/PAH (open). The membrane separates the receiving phase (initially deionized water) from a 0.1 M salt solution. External reflectance FTIR Spectra of IO-bilayer PAH/PAA-Cu films cross-linked at different temperatures: (a) unheated (b) heated to 160°C. Films were deposited on MPA-coated gold. Cyclic voltammetry of 10-bi1ayer films of PAH/PAA-Cu2+ deposited at different deposition pH values on MFA-modified gold surfaces: pH 5.5-dashed line, pH 6-solid line, and pH 6.6-dotted line. Areas of the reduction peaks were calculated by drawing the base line from the current value at a potential of 0.3V to the current value at -0.5V. Cyclic voltammetry of IO-bilayer PAH/PAA-Cu films deposited with different Cu2+ concentrations: 2.5 mM Cu2+ -dotted line, 5.0 mM Cu2+—solid line and 7.5 mM Cu2+-dashed line. The area of the reduction peak was calculated after drawing the base line from the current at a potential of 0.3V to that at -O.4V. Films were deposited at pH 6 on MPA-modified gold. Schematic representation of the model used to simulate ion transport through templated MPMS. The membrane consists of two charged layers: a surface layer and the membrane bulk. The line represents a hypothetical concentration profile for the excluded ion. me' and mez are the fixed charge concentrations of bulk of the film and surface layer respectively. Apparatus for gas-permeation measurements Schematic diagram showing polymerization of EGDMA from a polyelectrolyte surface modified with an initiator. External reflection FTIR spectra of PAH/PSS/PAH on a MPA-coated Au wafer before (a) and after reaction with an acid bromide initiator (b). Spectrum (c) is from a PEGDMA film grown for 20 hours from PAH/PSS/PAH modified with initiator. xiii 49 52 57 60 61 81 83 84 Figure 3.4 Figure 3.5 Figure 3.6 Figure 3.7 Figure 3.8 Figure 3.9 Figure 3.10 Figure 3.11 Figure 4.1 Figure 4.2 FESEM images of the filtrate side of porous alumina (0.02 pm surface pore diameter) (a) before depositing polyelectrolytes (b) after depositing one bilayer of PSS/PAH, and (c) after growth of PEGDMA from the surface for 20 hours. Image ((1) is a cross-section of the membrane shown in (c). 88 Reflectance F TIR spectra of (a) a monolayer of BrC(CH3)2COO(CH2)1 13); and (b) a grafied PHEMA layer. 90 Ellipsometric thicknesses of PHEMA films vs polymerization time. Thickness values are the average of three different films, and the error bars represent standard deviations. 92 Fluxes of different gases through a PEGDMA membrane (50 nm thick) as a function of transmembrane pressure drop. The outlet pressure was 1 atm and the measurements were performed at room temperature. PEGDMA was deposited on porous almnina. 95 Fluxes of different gases through a PHEMA membrane (60 nm thick) as a function of transmembrane pressure drop. The outlet pressure was 1 atm and the measurements were performed at room temperature. PHEMA was deposited on porous alumina. 99 Reflectance FTIR spectra of a PHEMA film grown from (BrC(CH3)2COO(CH2)1 IS); on gold before (a) and after reaction with pentadecafluorooctanoyl chloride (b). 103 Transmission FTIR spectra of a PHEMA film grown from (BrC(CH3)2COO(CH2)1 IS); on gold-coated porous alumina (a) before and (b) after reaction with pentadecafluorooctanoyl chloride. 104 Fluxes of different gases through a fluorinated-PHEMA membrane (100 nm thick) as a function of transmembrane pressure drop. The outlet pressure was 1 atm and the measurements were performed at room temperature. 106 Plot of receiving phase concentration as a function of time in diffusion dialysis when the source phase (0.1 M salt) was separated from the receiving phase (initially deionized water) by a porous alumina substrate capped with 28 nm of PHEMA. 120 Reflectance F TIR spectra of a 28-nm thick grafted PHEMA film on a gold-coated wafer before (a) and after reaction with succinyl chloride (b) and subsequent exposure to water (0). 125 xiv Figure 4.3 Figure 4.4 Plot of receiving phase concentration as a function of time in diffusion dialysis when the source phase (0.1 M salt) was separated from the receiving phase (initially deionized water) by a composite membrane of 35 nm cross-linked PHEMA film. The inset expands the concentration scale for 8042' and Fe(CN)63'. PHEMA was cross-linked by succinyl chloride. 127 Reflectance FTIR spectra of (a) a grafted PHEMA layer (28 nm) (b) CDI-derivatized PHEMA and (c) CDI-derivatized PHEMA after reaction with 2-hydroxymethyl-18-crown-6 (43 nm). 133 XV ATRP bpy CV DMF EGDMA FESEM FTIR HEMA MPA MPF MPM PAA PAA-Cu PAH PEGDMA PHEMA PSS LIST OF ABBREVIATIONS atom transfer radical polymerization bipyridine cyclic voltammogram N,N-dimethylformamide ethylene glycol dimethacrylate field emission scanning electron microscopy fourier transform infrared spectroscopy 2-hydroxyethyl methacrylate mercaptopropionic acid multilayer polyelectrolyte film multilayer polyelectrolyte membrane poly(acrylic acid) poly(acrylic acid) copper complex poly(allylamine hydrochloride) poly(ethylene glycol dimethacrylate) poly(2-hydroxyethyl methacrylate) poly(styrene sulfonate) xvi CHAPTER 1 Introduction 1.1 Structure of the Dissertation Membrane-based separations are attractive because of low energy costs and simple Operation. However, low flux through selective membranes oflen limits their utility."2 The most common strategy for increasing flux is the use of asymmetric or composite membranes that consist of an ultrathin, selective skin on a highly porous support. The minimal thickness of the Skin allows high flux, while the support provides mechanical strength.3 The selective skin should be aS thin as possible to achieve the highest flux, but synthesis of defect-free skins with thicknesses less than 50 nm is an ongoing challenge.4 My research focused on the development of methods for formation of defect-free, ultrathin polymer Skins and the use of these Skins for separations. Specifically, this dissertation discusses the fabrication of ultrathin skins on porous alumina supports using two techniques: alternating polyelectrolyte deposition and grafting of polymers from a surface using atom transfer radical polymerization (ATRP). These techniques afford composite membranes that allow selective ion and/or gas transport. Chapter 1 of this dissertation describes previous research on thin film formation with special emphasis on alternating polyelectrolyte deposition and polymerization from a surface using atom transfer radical polymerization (ATRP). This chapter also contains background information about membranes for gas and ion separations and sets forth the motivation behind my research. Chapter 2 concerns templating of multilayer polyelectrolyte films (MPFS) to enhance their ion-transport selectivities. I first discuss some of the limitation of MPFS as ion-separation membranes and then show how these challenges can be overcome by using Cu2+ as a placeholder during polyelectrolyte deposition. Specifically, I utilized Cu2+-chelated poly(acrylic acid) (FAA) and post-deposition removal of Cu2+ to introduce ion-exchange Sites into the bulk of FAA/poly(allylamine hydrochloride) films. Diffusion dialysis studies showed 5-fold higher Cl’/SO42' selectivities with Cu2+-templated membranes than with similar membranes deposited without Cu”. Post-deposition cross- linking of Cu2+-templated membranes by heat-induced amide bond formation from carboxylate and ammonium groups further increased C1'/SO42‘ selectivity to values as high as 600. Ion-transport simulations suggest that both analyte size and analyte charge are important in effecting selective transport. The third chapter of the dissertation discusses formation of composite membranes by polymerization from porous alumina and examines gas transport through these systems. We synthesized cross-linked poly(ethylene glycol dimethacrylate) (PEGDMA) and non cross-linked poly(2-hydroxyethyl methacrylate) (PHEMA) films by ATRP fiom initiators immobilized on porous alumina. Gas-permeation studies with PEGDMA films showed a COZ/CH4 selectivity of 20, whereas PHEMA fihns exhibited a selectivity of 0.7. However, fluorination of PHEMA through reaction of hydroxyl groups with pentadecafluorooctanoyl chloride increased the COZ/CH4 selectivity to 9. Chapter 4 describes the performance of PHEMA and derivatized-PHEMA films as ion-separation membranes. Diffusion dialysis studies indicate that PHEMA membranes have a Cl’/SO42' selectivity of 15, a K+/Mg2+ selectivity of 47 and a Fe(CN)(,3' /Cl' selectivity of 160. Unlike polyelectrolyte membranes, PHEMA is a neutral polymer, and the observed selectivities are probably due to differences in size or hydration energies among the various ions. Although PHEMA has moderate selectivities, Cl' fluxes through these membranes were quite low. To achieve high selectivity while maintaining reasonable flux, we cross-linked very thin (28 nm) PHEMA films by reaction with a di-acid chloride. Diffusion-dialysis studies with these cross-linked PHEMA membranes showed a Cl’/SO42‘ selectivity of 300 and only a 50% decrease in C1' flux compared to that of bare alumina. Finally, chapter 5 contains conclusions and suggestions for future work. 1.2 Formation and Development of Thin Polymer Films Development of thin organic films has been a highly active area of research for nearly 100 years.5'7 Coating a solid substrate with a thin film plays a vital role in controlling surface properties for applications in optics,8 sensingf”10 corrosion protection,ll and separations,”14 and many of these applications require well-defined films with uniform morphologies.7 5 Langmuir-Blodgett (LB) films were the first examples of multilayer coatings prepared in a controlled layer-by-layer fashion.15 "6 The deposition of LB films begins by mechanically assembling an array of arnphiphilic molecules on a water surface. Once the molecules are compressed to the desired organization, the film can then be transferred to a solid support as shown in Figure 1.1. This results in deposition of a single layer of 71% 359K WRETEZWI <- Figure 1.1: Deposition of a Langmuir-Blodgett film from a floating Langmuir monolayer. Figure adapted from hgp://www.nes.coventry.ac.uk/research/cmbe/filmbal.htm. molecules, and subsequent immersions in the trough result in the deposition of more layers. LB fihns have been used to investigate phenomena ranging from optical properties of chromophores to ordered polymerizations on a substrate.7 However, these coatings have two limitations which greatly restrict their applications. First, LB films are fragile because the layers are linked by weak van der Waals forces.17 Second, fabrication of these materials requires pro-assembly at the air-water interface, so only arnpiphillic molecules form films. A more recent step in the development of ultrathin films was the discovery of self-assembled monolayers (SAMS).5 ’18'2' The principle behind formation of these monolayers is simple; a molecule containing a head group that adsorbs to a surface, e. g. thiols on gold, assembles on the substrate under the constraints of intermolecular forces and adsorption site geometry. Unlike LB films, formation of SAMS doesn’t require any pre-assembly, so their synthesis is simple and convenient. Additionally, because they are bound to a surface, SAMS are more robust than LB films.22 The most common family of SAMS is organothiols adsorbed on gold, and the first systematic study of these materials was done by Nuzzo and Allara in 1985.1823 Since then, organothiol monolayers have been the subject of thousands of investigations. By employing thiols with different tail groups, SAMS can be easily used to modify surface properties. In addition to Au, other substrates such as Al/Ale3, Si/SiOz and Cu have also been used to support SAMS.7’24 The main drawback to SAMS is the limited film thickness available from monolayer formation. Additionally, although self-assembled coatings are more convenient and stable than LB films, the stability of Au—thiol films is still an issue at high temperatures.25 More recently a number of techniques were developed to prepare multilayer fihns. 26'” The most convenient of these methods is probably alternating deposition of anionic and cationic polyelectrolytes on charged supports.3 "32 This strategy overcomes many of the limitations imposed by LB films and SAMS, although multilayered polyelectrolyte films are not well-ordered structures. Polyelectrolyte films are very stable because of the multiple electrostatic interactions between layers, and synthesis of these coatings requires only a simple dip and rinse procedure. The versatility and simplicity of alternating polyelectrolyte deposition are evident from the variety of charged substrates used for 33-36 33.37-44 deposition and the wide range of polyelectrolytes that can form these fihns. Because polyelectrolyte films form the basis of much of my work, I describe them in more detail below. 1.3 Multilayer Polyelectrolyte Films Using colloidal alumina and colloidal silica, Iler45 first demonstrated the fabrication of multilayers by sequential adsorption of oppositely charged materials. Later, Decher and Hong extended this idea and employed alternating deposition of oppositely charged polyelectrolytes (e. g. poly(allylamine hydrochloride) and sodium poly(styrenesulfonate)) on charged supports to form multilayer polyelectrolyte films (MPFS).31’32’46 Their technique allowed rapid formation of multilayer films with control over film thickness on the nm scale. More recently, alternating polyelectrolyte deposition has been demonstrated on a variety of charged substrates with polyelectrolytes ranging from poly(styrenesulfonate)37’46’47 to DNA47 and charged viruses.38 Deposition of MPFS occurs as shown in Figure 1.2.48 The procedure begins with immersion of a charged substrate into a solution containing an oppositely charged (with respect to the substrate) polyelectrolyte. A film forms due to electrostatic attraction between the substrate and the polyelectrolyte, but the thickness of this film is limited by electrostatic repulsion of incoming chains by adsorbed polymer. (This picture is oversimplified because the main driving force behind film formation is actually the increase in entropy that results when adsorption of a polyelectrolyte chain displaces many counterions from the substrate surface).49 Rinsing of the substrate with water and immersion in a second solution containing an oppositely charged polyelectrolyte then yields another layer on the surface, and repetition of this adsorption sequence results in a multilayer film. Charge overcompensation at each deposition step is the key to subsequent adsorption of an oppositely charged polyelectrolyte. Typically, film thickness increases linearly with the number of polyelectrolyte layers after deposition of the first 3 50-52 or 4 layers. Several features of alternating polyelectrolyte deposition make it a unique technique for thin film formation. First, this procedure is environmentally friendly because in most cases, the solvent is water. Second, the only stipulation on the substrate for film formation is that it should contain sufficient charge. Thus, substrates with a wide range of topologies can support film growth. The electrostatic interactions between polyelectrolytes and surface charge also allow good adhesion between the polymer and the substrate. Finally, the thickness of films can be easily controlled by varying a: .3 n: L H U) .0 3 (I) 1. Polyanion ((39990 3. Polycation O —— 2. Wash (9 Q (E) 4. Wash 03 ..: G) ®®®®®®®®®® Figure 1.2: Schematic diagram of the “Dip and Rinse” procedure for alternating deposition of oppositely charged polyelectrolytes on a charged substrate. In the actual film structure, polycations and polyanions are interdigitated. Figure adapted from reference 46. 50.53 deposition parameters such as the number of adsorption steps, the pH of deposition solutions,54'55 deposition time,37 and the supporting electrolyte concentration in deposition solutions.50’56'58 Since the introduction of MPF S, these materials have been subjected to numerous studies to understand their structure and the mechanism of film formation. Small-angle X-ray scattering38 and UV/visible spectroscopy33'35’40 provided evidence for layer-by- layer growth, but neutron reflectometry showed that multilayer films are not highly stratified. Strong interdigitation of polycations and polyanions occurs over several neighboring layerssg’6O A variety of other methods were also used to characterize these multilayers, including: electrochemical techniques,58 surface plasmon resonance,36’56’61 ellipsometry,58 quartz-crystal-microbalance gravimetr'y,36’62 X-ray photoelectron 57,63 14,34 . . 4 . . spectroscopy, atomic force mrcroscopy,3 and scannlng electron microscopy. Many recent studies showed that MPFS have potential applications in areas such 43,66-70 56,71,72 “‘65 non-linear optics, sensors, conductive as light-emitting devices, 76,77 13,14,78-85 73,74 separation membranes coatings, surface patterning,75 protective coatings, and nano-particle formation.86 Some of these applications require impermeable films (e. g., protective coatings), while others necessitate films with high or selective permeabilities (e.g., separation membranes). In the specific area of separation membranes, efficient separation requires an understanding of the factors that affect transport of molecules or ions through MPF S. Below, I discuss literature that relates specifically to ion transport through MPFS. 1.4 Ion Permeation through MPFS Several recent papers reported studies of ion permeation through MPFS. Schlenoff and co-workers used electrochemical methods to examine the transport of redox-active ions thorough a thin MPF on an electrode.87 Transport rates of redox-active ions decreased with increasing charge, suggesting that diffusion through MPFS occurs via hopping between ion-exchange sites. The need for more ion-exchange Sites to compensate the charge on multivalent ions should result in fewer hopping sites for these species, and hence slower transport. MOhwald and von Klitzing studied the transport of neutral quenchers through polyelectrolyte films by total internal reflection fluorescence spectroscopy.88 They Showed that the diffusion of molecules through the bulk of the film is much slower than through the outer layers. This probably occurs because layers in the bulk of the film pack more tightly than outer layers. In situ ellipsometric and cyclic voltammetric studies done by our group showed that the permeability of MPFS depends on the pH and ionic strength of polyelectrolyte deposition solutions, the number of layers in the film, and the nature of constituent polymers.58 More recently, MOhwald and co-workers investigated the release of fluorescein from multilayered polyelectrolyte capsules and showed that the rate of release is a function of the number of assembled polyelectrolyte layers.89’90 These studies provide important background for designing separation membranes containing MPFS. 1.5 MPFS as Ion-separation Membranes Several recent studies exploited the simplicity and versatility of layer-by—layer adsorption of polyanions and polycations to form ion-separation membranes. Krasemann 10 and Tieke examined ion permeation through 60 bilayers of poly(sodium styrene sulfonate) (PSS)/poly(allylamine hydrochloride) (PAH) deposited on a polymeric support. These membranes exhibited a Nai'lMg2+ selectivity of 112 and a Cl'/SO42‘ selectivity of 45.79 Selectivity increased with both the number of bilayers and the ionic strength of deposition solutions. Krasemann and Tieke suggested that the increase in selectivity with the number of layers is due to the multi-bipolar nature of the polyelectrolyte film. They thought that electrostatic repulsion would occur at each bilayer, resulting in higher monovalent/divalent ion selectivities for films with many layers. However, several studies suggest that polyelectrolyte bilayers are completely intertwined and that the bulk of these films have a net neutral charge.91 Thus, repulsion of ions should occur only at the film-solution interface. Perhaps large numbers of bilayers resulted in membranes with fewer defects. Our group Showed that membranes composed of 5 PSS/PAH bilayers on porous alumina have a Cl'/SO42' selectivity of 6 and that Cl°/SO42’ selectivity does not increase after deposition of an additional 5 bilayers.14 Selectivity is largely due to the electrostatic exclusion of multiply charged ions by surface charge, as selectivity depends on whether the membrane is terminated with a polycation or a polyanion. Stair et.al80 found that upon changing the surface of PAA/PAH-capped films from PAA to PAH, Cl’/SO42' selectivity decreased from 360 to 2. This observation further confirmed that outer layer charge plays a dominant role in determining selectivity. All of these previous studies suggest that electrostatic exclusion is a major factor behind ion-transport selectivity in multilayer polyelectrolyte membranes. Therefore, increasing the charge density in the bulk or at the surface of these membranes should ll enhance monovalent/divalent ion—transport selectivity. However, as mentioned, control over charge density in the bulk of MPFS is difficult because polycation charge is essentially completely compensated by polyarrion charge, giving little net fixed charge density in interior of the film.42'9| Insertion of net charge into MPF interiors likely requires a post-deposition reaction. We utilized Cu2+ as a template to create ion-exchange Sites (fixed charge) in the bulk of poly(acrylic acid) (PAA)/PAH membranes. This strategy involves alternating deposition of PAA-Cu (1 Cu2+ per 8 COO‘ groups) and PAH on porous alumina supports followed by removal of Cu2+ and deprotonation to yield free -COO' ion-exchange sites (Figure 1.3). Diffirsion dialysis studies showed that the selectivity of Cu2+-templated membranes is dramatically higher than that of membranes prepared in the absence of Cu”, presumably due to the higher concentration of fixed charge in the bulk of the film. Post-deposition cross-linking of these membranes by heat-induced amide bond formation further increased Cl'/SO42' selectivity to values as high as 600. These remarkable selectivities are achieved with no significant decrease in flux relative to pure FAA/PAH membranes. 1.6 Grafted Polymer Brushes Attachment of polymer chains to substrates provides another attractive way of controlling surface properties.92'94 Assemblies of polymers that are linked to a surface and yet highly extended into solution are often termed polymer brushes.95 These tethered polymer chains have potential applications in chemical separations, sensing, stabilization of colloidal suspensions, control of wetting and adhesion, corrosion resistance, 12 NH; [git-COO coo-4%; coo NH; 3‘)?! O = 2 COO‘ 000—ng— COO- NH; +H3 —coo- NH; +H3 0 —COO- NH3 H3 NH; —000H (2) O =§ coo-Cu2+-ooc 6 Figure 1.3: Synthesis of Cu2+-templated polyelectolyte films. Step 1: deposition 0 a polycation (PAH) in the presence of uncomplexed Cu2+. Step 2: deposition of IDar‘iially Cu2+-complexed PAA. Step 3: removal of Cu2+. Step 4: deprotonation of e fl‘ee carboxylic acid groups of PAA. Repetition of steps 1 and 2 produces a multilayer film. lntertwining of layers is not shown for figure clarity. microfluidics, fouling resistance and “chemical gating”.9("99 Methods for formation of polymer brushes include covalent attachment and physisorption of chains to a substrate. The latter method generally utilizes block copolyrners containing one block that strongly interacts with the surface and a second block that forms the brush layer.100 In this case, attachment to the surface occurs via van der Waals forces or hydrogen bonding, and thus these brushes can be desorbed by good solvents or displaced by other polymers. Covalent tethering of polymer brushes to a surface greatly enhances film stability. Covalent attachment of polymer brushes to a substrate can occur by either the '0‘ In the case of “grafting to” a “grafting to” or the “grafting from” method (Figure 1.4). surface, functional groups on a pre-synthesized polymer react with groups on the substrate to anchor the polymer.102 Although this technique provides strong adhesion between the surface and the polymer, thicknesses of films prepared in this way are generally limited to less than 5 nm. After formation of a relatively thin film, a diffusion barrier prohibits more polymer molecules from reaching reactive surface Sites, resulting in thin films with low grafting densities. The “grafting from” technique is attractive because of the potential for formation of long polymer brushes with a high grafting density.93 In this technique, initiators are firSt itIllnobilized on the surface, and polymer growth subsequently proceeds from these Sites- Growth occurs as small monomers diffuse to the surface and reach growing chain ends, and thus, in contrast to the “grafting to” technique, there is no large diffusion baxrier to polymer growth. Among the numerous polymerization techniques, free radical polymerization is 14 “Grafting to” a surface “Grafting from” a surface Z {jut—2 l Figure 1.4: Two methods for grafting of polymer films onto solid surfaces. F igure adapted from reference 101. the most widely used process because it is relatively easy to perform, and a wide range of monomers can be used.'03 However, free radical polymerization provides limited control over molecular weight and molecular weight distribution because of rapid radical-radical termination reactions. This limitation inspired the emergence of new controlled/“living” radical polymerization techniques. The concept of “living” polymerization was first introduced by Szwarc in 1956.104 The key feature of any “living” process is that chain growth proceeds without the occurrence of irreversible termination steps, i.e., chain transfer, radical coupling, and disproportionation, and molecular weight is a linear function of conversion. The first reported “living” procedure was the sequential anionic polymerization of two non-polar monomers to produce block copolyrners.105 Since then, numerous studies demonstrated contro lled/“living” polymerization techniques.103 In the mid 19908, two research groups report ed the discovery of a controlled/“living” radical polymerization method that employs transfer of a halide atom between a transition metal salt and a growing radical. This process was termed atom transfer radical polymerization (ATRP).1°6’107 Today, of the ma11y different “living” polymerization techniques (cationic, anionic, ring-opening, and ni‘Zl‘oxide-mediated polymerization;108 and reversible addition fragmentation chain transfer1 09'1”), ATRP is probably the most powerful, versatile and attractive technique. 1 '7 Atom Transfer Radical Polymerization (ATRP) ATRP, as the name implies, is the reversible formation of a radical by transfer of a - hallde atom from an alkyl halide to a transition metal of low oxidation state (Scheme 1 - l 1 ) ‘ 11 Upon transfer of the halide atom, the transition metal undergoes a one-electron l6 oxidation. After formation, radicals can either propagate via reaction with monomer or reform the dormant Species by abstraction of a halide atom from a metal-ion complex. (311(1) complexes are the most common ATRP catalysts, but other transition metal ions have also been used (Ru(u),‘°7 Fear),1 ‘2" ‘3 Ni(II),' '4" '5 Pd(II) and Rh(II)' '5). Monomers polymerized using ATRP include styrenes, (meth)acrylates, acrylonitriles, (meth)acrylamides, methacrylic acids and vinylpyridine.l '6 Compounds containing weak carbon-halogen or hetero-halogen bonds, e.g., a-bromocarbonyl groups or sulfonyl haI ides, serve as initiators. k ’R° 4' Cu"le Ligand R—x + CuIXILigand - ——§- Initiator Catalyst kd Q4) ."~.‘kt “‘ R-R kp Initiator - Alkyl halide M = Monomer Catalyst — Transition metal (Cu, Fe, Ru, Ni, Pd) complexed by one or more ligands X = Halogen atom Scheme 1.1: Mechanism of ATRP (adapted from reference 11 l). ATRP is a controlled or “living” process when the atom-transfer equilibrium Strongly favors the dormant species to give low radical concentrations. Because radical Coupling and disproportionation kinetics are second order with respect to radical oncentratron, termination In ATRP can be minimal compared to propagation, leading to t he f‘Ol‘mation of well-defined polymers with low polydispersity. To control Do 1 5"Itlerization, the transition metal/ligand ATRP catalyst is generally selected so that the 17 activation rate constant (k3) is much lower than the deactivation rate constant (kd). Advantages of ATRP over conventional radical polymerizations include: compatibility with a variety of functional monomers, tolerance to trace impurities (water, oxygen, and i nhibitor), control over molecular weights and molecular weight distributions, and possible block-copolymer formation by sequential activation of the dormant chain end in t he presence of different monomers. ATRP initiated from surfaces provides an attractive and convenient way to synthesize dense, uniform polymer brushes with controllable thickness (Figure 1.5). When initiators are covalently attached to a surface, atom-transfer results in the formation of radicals on the substrate but not in solution, limiting unwanted polymerization in ”“21 also helps to avoid solution. The ability to perform ATRP at room temperature autopolymerization in solution, and thus extensive extraction of adsorbed polymer after growth of a polymer brush is not necessary. Miminal solution polymerization is especi ally important when synthesizing cross-linked polymer films because cross-linked, physisorbed polymer is difficult to remove from a surface.1 '9 Recently, numerous reports described the use of ATRP to grow polymer brushes from a variety of substrates in a Well-defined manner.l '7‘120’122'125 The wide range of monomers that can be used in ATRP Shoald permit tailoring of the properties and composition of polymer brushes. 1 ‘8 Formation of membranes by surface-initiated ATRP Like alternating polyelectrolyte deposition, surface-initiated ATRP provides a ethod for formation of ultrathin membrane skins on porous alumina supports. This Qess Involves Immobilization of an Initiator on porous alumma and subsequent 18 O—C—Br 1 1 O—C - O—O—C- O—O—O—c- O—O—O—O—O—c- O—O—O—O—O—C—Br I l (dormant) O—O—O—O—O—O—C- O—O—O—O—O—O—O—C- Time CuBr/ligand CuBr2 lligand CuBr2 lligand CuBr2 lligand CuBr2 lligand CuBr lligand CuBr2 lligand CuBr2 lligand CuBr2 lligand Figure 1.5: Cartoon showing the different stages of surface-initiated ATRP. l9 polymerization from the anchored initiator sites (Figure 1.6). We used two strategies to anchor the initiator to alumina. The first method employed adsorption of a few layers of charged polyelectrolytes and subsequent attachment of the initiator to the MPF surface, while the second procedure involved gold sputtering followed by formation of a self- assembled monolayer of a disulfide initiator. Polymerization from these immobilized i ni tiators occurred using room-temperature ATRP. Using these procedures, we grew two kinds of polymer films, cross—linked poly(ethylene glycol dimethacrylate) (PEGDMA) and non cross-linked poly(2-hydroxyethyl methacrylate) (PHEMA). Cross-linked pol ytner films are very attractive in separations because of their mechanical stability and low free volumem'127 PHEMA films are also attractive membrane materials because their hydroxyl groups can be easily derivatized in high yields to tailor films for specific separreltions.118 The polymer grth from alumina was monitored by transmission-FT IR spectroscopy and scanning electron microscopy (SEM). Both top-down and cross- sectional SEM images of these polymer films showed that they effectively covered the surface pores of the alumina. Gas permeation studies with cross-linked PEGDMA membranes showed a gas- tr311513011 selectivity of 20 for CO; over CH4. In contrast, non cross-linked PHEMA membranes exhibited minimal gas transport selectivities that depended primarily on the “1013? masses of the permeating gases. However, after derivatization of the hydroxyl groups of PHEMA with pentadecafluorooctanoyl chloride, COz/CH4 selectivity increased to N 9 . A detailed description of gas-separation membranes prepared by ATRP from porous alumina supports will form chapter 3 of the dissertation. Below I give some ba cl(ground on gas-separation membranes to provide some context for my studies. 20 5 nm-Au Coating \ Porous Alumina Dnitiator Anchoring I l Initiator Anchoring—l 1* it. t t t} trim /////// Polymerization , l Grafted Polymer Film (5?; {5.9; eteariogmefie Figure 1.6: Surface-initiated atom transfer radical polymerization from (a) a Polyelectrolyte film deposited on alumina and (b) a self-assembled initiator mOnolayer on Au—coated alumina. 21 1.9 Gas-Separation Membranes Gas separation with polymer membranes was initially described over a century ago. Mitchell first reported that different gases permeate through natural rubber at different rates.108 In 1866, Graham demonstrated the enrichment of air with O; by permeation through a natural rubber membrane.128 He showed that a mixture of gases could be separated according to their molecular weights by permeation through a microporous membrane, and the proportionality of gas flux to the reciprocal of the square root of molecular weight later became the well-known Graham’s Law of Diffusion. Subsequently, there was little development of gas-separation membranes until the 1960’s when Loeb and Souirajan129 invented the first asymmetric membrane of industrial interest. This membrane, which was prepared from cellulose acetate by phase inversion, was originally made for desalination of water and later modified for gas separation. The asymmetric membrane contained a thin, dense, selective skin at the surface of a highly porous material, and this structure allowed much greater fluxes than thick, homogeneous membranes. The dense skin also exhibited higher selectivities than the porous structures that behaved according to Graham’s law, while the underlying porous material provided mechanical strength. The era of commercial gas separations began in the 1970’s. Monsanto initiated the first large-scale separation of gases in 1977 for the recovery of H; from industrial gas streams using membranes made of polysulfone. In the 19808, Perrnea introduced the prism membrane for separation and recovery of hydrogen from purge gas streams of ammonia plants. This was a polysulfone membrane coated with silicone rubber. In the 22 mid 1980s, Cyanara, Separex and GMS used dried cellulose acetate membranes for removal of CO2 from natural gas, and in 1982, Generon produced the first N2/Air separation membrane using poly(4-methyl-l-pentene). This system had an O2/N2 selectivity of ~ 4. In the mid 19903, Generon, Praxair and Medal produced a polyimide membrane for O2/N 2 separation with a selectivity of 6-8.3 Other recent developments in gas-separation membranes include the commercialization of composite membranes. 3 Composite membranes are made by depositing a thin layer of polymer on a highly porous substrate. The thin selective layer acts as a discriminating film to give selectivity, and the porous layer provides mechanical stability to the system. Composite membranes have the advantage that only a small amount of expensive skin material is needed, while the porous support can be made from an inexpensive polymer. Membrane geometry is also critical for practical separations, as surface area must be maximized. Asymmetric membranes can be packed as hollow fibers or in a spiral- wound configuration. A spiral-wound module consists of series of membrane envelopes, and each envelope consists of two membrane sheets, which are separated by a feed spacer. In a hollow fiber module, asymmetric hollow fibers are bundled together to achieve a very high surface area. Some commercial hollow fiber modules contain more than a million hollow fibers. 130’” 1 Current membrane-based gas separations include a wide range of applications such as recovery of H2 from synthesis gases and petrochemical process streams, removal of CO2 from mixtures of hydrocarbons and natural gases, N2 or 02 enrichment from air, S02 removal from smelter gas streams, H28 and water removal from natural gas and air - - - 107 streams, and NH3 removal from recycle streams In arrunonra synthesrs. These 23 processes use both composite and asymmetric membranes as well as several different membrane geometries including hollow fiber and Spiral wound systemsf”132 Of special relevance to this thesis, several studies demonstrated the possibility of using various poly(alkyl methacrylates) in gas separation membranes, e. g., poly(ethyl ‘33 and a styrene/methacrylate co- methacrylate),4 poly(tert-butyl methacrylate), polymer.134 Yoshikawa and co-workers Showed that the presence of amine moieties in poly(methacrylate) films greatly enhances CO2/N2 separation.135 The use of fluorinated poly(methacrylates) provides another way to enhance CO2/N2 or CO2/CH4 selectivity.‘36 Most previous studies with poly(alkyl methacrylates) employed cast membranes with large thicknesses, and decreasing the thickness of these films would greatly enhance flux. Although many successes have been achieved in gas separation membrane research, fabrication of utrathin (<50 nm) polymer skins is still a challenge, so the focus of this work was to develop methods for deposition of selective, ultrathin polymer skins. To develop either thin or thick membrane materials, one needs to understand the factors affecting gas separation and the mechanism of separation. Below I discuss the mechanism of gas transport through polymer membranes and the factors that determine selectivity for one gas over another. 1.10 Mechanism of Gas Transport through Polymer Membranes Gas-transport selectivity is usually based on one of three mechanisms: Knudsen '37 Knudsen diffusion dominates when diffusion, molecular sieving or solution/diffusion. membrane pores are larger than the gas molecules being separated but smaller than the mean free path of these gases. Permeation rates are inversely proportional to the square 24 root of the gas molecular weight of the gas so selectivity is the reciprocal of the square root of the ratio of the molecular weights of the gases being separated.‘38 Because most gases of interest have similar molecular weights, separations based on Knudsen—diffusion are not highly selective. Molecular sieving of common gases occurs when pore diameters in a membrane are smaller than 7 A. Selectivities between gases of different sizes can be nearly infinite ”9"40 This method, however, is limited by when one gas is incapable of entering a pore. relatively low fluxes and the difficulty of preparing defect-free membranes with uniform pore Sizes. Solution-diffusion is the most common mechanism that operates in practical gas separations. In this mechanism, transport of gases occurs in three steps: sorption of the penetrant into the polymer film at the high-pressure interface, diffusion of the penetrant through the polymer film, and finally, desorption at the permeate Side (low-pressure interface). Thus, gas flux depends on the diffusivity and solubility of the gas in the polymer as well as the transmembrane partial pressure gradient. F ick’s first law (equation 1.1) describes the transport of a species within a nonporous membrane. Flux, J, is proportional to the concentration gradient, dc/dx, and the diffusion coefficient, D, for the molecule in the membrane. dc =-D—- 1.1 J dx According to Henry’s law, the concentration, C, of a specific gas in the membrane at the high- or low-pressure interface is proportional to partial pressure, p, and the solubility coefficient, S (equation 1.2). 25 C=Sp L2 Using Henry’s law to determine concentrations at the two gas-membrane interfaces, and assuming steady-state flux and constant values of D and S allows transformation of equation 1.1 to equation 1.3. J = D S Ap/d 1.3 In this equation, Ap is the partial pressure difference between the feed and permeate, and 6 is the membrane thickness. The permeability coefficient, P, of a particular membrane for a specific gas is then defined by equation 1.4. P = D S = J d/Ap 1.4 Selectivity for one gas over another, (1,413, is given by the ratio of the permeability coefficients of two gases (equation 1.5). This ratio is a measure of the relative fluxes of the two gases at the same driving force. Since permeability coefficients depend on both solubility and diffusion coefficients, selectivity also contains diffusivity (DA/DB) and solubility (S A/SB) components. Diffusion selectivity generally favors small molecules, while solubility selectivity favors more condensable gases. _£-&§_A_ - 1.5 a“ P, DES, Both selectivity and permeability determine membrane performance, and thus the careful selection of polymer materials is vital for efficient separations. 26 1.1 1 Summary As discussed earlier, the main objective of this work is the fabrication of defect- free, ultrathin polymer skins on porous supports and the use of these composite membranes in ion and gas separation to achieve high selectivity along with high flux. In this chapter, I tried to Show the need for ultrathin polymer skins in separation membranes and the challenges in forming these polymer skins with controllable thicknesses. I also discussed background on the development of polyelectrolyte films and polymerization from a surface, the theory behind gas transport through polymer membranes and some relevant past research done on the development of gas and ion separation membranes. 27 1 .1 l (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) (17) (18) References Spillman, R. W. Chem. Eng. Prog. 1989, 85, 41-62. Henis, J. M. S.; T ripodi, M. K. Science 1983, 220, 11-17. Baker, R. W. Ind. Eng. Chem. Res. 2002, 41, 1393-1411. Chiou, J. S.; Paul, D. R. J. Membr. Sci. 1989, 45, 167-189. Ulman, A. Chem. Rev. 1996, 96, 1533-1554. Swalen, J. D.; Allara, D. L.; Andrade, J. D.; Chandross, E. A.; Garoff, S.; Israelachvili, J.; McCarthy, T. J .; Murray, R.; Pease, R. F.; et a1. Langmuir 1987, 3, 932-950. Ulman, A. Introduction to Ultrathin organic films: from Langmuir-Blodgett to self-assembly; Academic Press: Boston, 1991. Bubeck, C.; Neher, D.; Kaltbeitzel, A. Makromol. Chem, Macromol. Symp. 1990, 3 7, 239-245. Miller, L. S.; McRoberts, A. M.; Walton, D. J .; Peterson, 1. R.; Parry, D. A.; Sykesud, C. G. D.; Newton, A. L.; Powell, B. D.; Jasper, C. A. Thin Solid Films 1996, 284-285, 927-931. Petty, M. C. Biosensors Bioelectron. 1995, 10, 129-134. Whitesides, G. M.; Laibinis, P. E. Langmuir 1990, 6, 87-96. Stroeve, P.; Vasquez, V.; Coelho, M. A. N.; Rabolt, J. F. Thin Solid Films 1996, 284-285, 708-712. van Ackem, F.; Krasemann, L.; Tieke, B. Thin Solid Films 1998, 327-329, 762- 766. Harris, J. J.; Stair, J. L.; Bruening, M. L. Chem. Mater. 2000, 12, 1941-1946. Langmuir, I. J. Am. Chem. Soc. 1917, 39, 1848-1906. Blodgett, K. B. J. Am. Chem. Soc. 1935, 57, 1007-1022. Embs, F.; Funhoff, D.; Laschewsky, A.; Licht, U.; Ohst, H.; Prass, W.; Ringsdorf, H.; Wegner, G.; Wehrmann, R. Adv. Mater. 1991, 3, 25-31. Allara, D. L.; Nuzzo, R. G. Langmuir 1985, 1, 45-52. 28 (19) (20) (21) (22) (23) (24) (25) (26) (27) (28) (29) (30) (31) (32) (33) (34) (35) (36) Allara, D. L.; Fowkes, F. M.; Noolandi, J.; Rubloff, G. W.; Tirrell, M. V. Mater. Sci. Eng. 1986, 83, 213-226. Bishop, A. R.; Nuzzo, R. G. Curr. Opin. Coll. Int. Sci. 1996, I, 127-136. Nuzzo, R. G.; Fusco, F. A.; Allara, D. L. J. Am. Chem. Soc. 1987, 109, 2358- 2368. Knoll, W.; Matsuzawa, M.; Offenhausser, A.; Riihe, J. Isr. J. Chem. 1997, 36, 357-369. Allara, D. L.; Nuzzo, R. G. Langmuir 1985, 1, 52-66. Bain, C. D.; Troughton, E. B.; Tao, Y. T.; Evall, J .; Whitesides, G. M.; Nuzzo, R. G. J. Am. Chem. Soc. 1989, 111, 321-335. Bain, C. D.; Evall, J.; Whitesides, G. M. J. Am. Chem. Soc. 1989, 111, 7155- 7164. CaO, G.; Hong, H. G.; Mallouk, T. E. Acc. Chem. Res. 1992, 25, 420-427. Fang, M.; Kaschak, D. M.; Sutorik, A. C.; Mallouk, T. E. J. Am. Chem. Soc. 1997, 119,12184-12191. Kim, H.-N.; Keller, S. W.; Mallouk, T. E.; Schmitt, J.; Decher, G. Chem. Mater. 1997, 9,1414-1421. Kohli, P.; Blanchard, G. J. Langmuir 1999, 15, 1418-1422. Huc, V.; Bourgoin, J .-P.; Bureau, C.; Valin, F.; Zalczer, G.; Palacin, S. J. Phys. Chem. B 1999, 103, 10489-10495. Decher, G.; Hong, J. D. Makromol. Chem, Macromol. Symp. 1991, 46, 321-327. Decher, G.; Hong, J. D. Ber. Bunsenges. Phys. Chem. 1991, 95, 1430-1434. Lvov, Y.; Ariga, K.; Ichinose, 1.; Kunitake, T. J. Am. Chem. Soc. 1995, I I 7, 6117-6123. Caruso, F .; Furlong, D. N.; Ariga, K.; Ichinose, 1.; Kunitake, T. Langmuir 1998, 14, 4559-4565. F erreira, M.; Rubner, M. F. Macromolecules 1995, 28, 7107-7114. Caruso, F .; Niikura, K.; Furlong, D. N.; Okahata, Y. Langmuir 1997, I3, 3427- 3433. 29 (37) (38) (39) (40) (41) (42) (43) (44) (45) (46) (47) (48) (49) (50) (51) (52) (53) (54) Lvov, Y.; Decher, G.; MOehwald, H. Langmuir 1993, 9, 481-486. Lvov, Y.; Haas, H.; Decher, G.; MOehwald, H.; Mikhailov, A.; Mtchedlishvily, B.; Morgunova, E.; Vainshtein, B. Langmuir 1994, 10, 4232-4236. Lvov, Y. M.; Lu, Z.; Schenkman, J. B.; Zu, X.; Rusling, J. F. J. Am. Chem. Soc. 1998, 120, 4073-4080. Sukhorukov, G. B.; MOehwald, H.; Decher, G.; Lvov, Y. M. Thin Solid Films 1996, 284-285, 220-223. Schlenoff, J. B.; Laurent, D.; Ly, H.; Stepp, J. Adv. Mater. 1998, 10, 347-349. Laurent, D.; Schlenoff, J. B. Langmuir 1997, 13, 1552-1557. Lindsay, G. A.; Roberts, M. J .; Chafin, A. P.; Hollins, R. A.; Merwin, L. H.; Stenger—Smith, J. D.; Yee, R. Y.; Zarras, P.; Wynne, K. J. Chem. Mater. 1999, 11, 924-929. Lvov, Y.; Munge, B.; Giraldo, 0.; Ichinose, 1.; Suib, S. L.; Rusling, J. F. Langmuir 2000, 16, 8850-8857. Iler, R. K. J. Coll. Int. Sci 1966, 21, 569—594. Decher, G.; Hong, J. D; Schmitt, J. Thin Solid Films 1992, 210/211, 831-835. Lvov, Y.; Decher, G.; Sukhorukov, G. Macromolecules 1993, 26, 5396-5399. Decher, G. Science 1997, 277, 1232-1237. Bertrand, P.; Jonas, A.; Laschewsky, A.; Legras, R. Macromol. Rapid. Commun. 2000, 21, 319-348. Dubas, S. T.; Schlenoff, J. B. Macromolecules 1999, 32, 8153-8160. Laschewsky, A.; Mayer, B.; Wischerhoff, E.; Arys, X.; Jonas, A. Ber. Bunsenges. Phys. Chem. 1996, 100, 1033-1038. Arys, X.; Jonas, A. M.; Laguitton, B.; Laschewsky, A.; Legras, R.; Wischerhoff, E. Thin Solid Films 1998, 327-329, 734-738. Hoogeveen, N. G.; Stuart, M. A. C.; Fleer, G. J.; Boehmer, M. R. Langmuir 1996, 12, 3675-3681. Yoo, D.; Shiratori, S. S.; Rubner, M. F. Macromolecules 1998, 31, 4309-4318. 30 (55) (56) (57) (58) (59) (60) (61) (62) (63) (64) (65) (66) (67) (68) (69) (70) (71) Shiratori, S. S.; Rubner, M. F. Macromolecules 2000, 33, 4213-4219. Caruso, F.; Niikura, K.; Furlong, D. N.; Okahata, Y. Langmuir 1997, 13, 3422- 3426. Sukhorukov, G. B.; Schmitt, J .; Decher, G. Ber. Bunsenges. Phys. Chem. 1996, 100, 948-953. Harris, J. J .; Bruening, M. L. Langmuir 2000, 16, 2006-2013. Schmitt, J .; Griienewald, T.; Decher, G.; Pershan, P. S.; Kjaer, K.; LOesche, M. Macromolecules 1993, 26, 7058-7063. LOesche, M.; Schmitt, J.; Decher, G.; Bouwman, W. G.; Kjaer, K. Macromolecules 1998, 31 , 8893-8906. Kurth, D. G.; Osterhout, R. Langmuir 1999, 15, 4842-4846. Lvov, Y.; Ariga, K.; Onda, M.; Ichinose, 1.; Kunitake, T. Collid Surf A 1999, 146, 337-346. Chen, W.; McCarthy, T. J. Macromolecules 1997, 30, 78-86. Liu, Y.; Ma, H.; Jen, A. K. Y. Chem Commun. 1998, 2747-2748. Wu, A.; Yoo, D.; Lee, J. K.; Rubner, M. F. J. Am. Chem. Soc. 1999, 121, 4883- 4891. Koetse, M.; Laschewsky, A.; Mayer, B.; Rolland, 0.; Wischerhoff, E. Macromolecules 1998, 31 , 9316-9327. Laschewsky, A.; Wischerhoff, E.; Kauranen, M.; Persoons, A. Macromolecules 1997, 30, 8304-8309. Roberts, M. J .; Lindsay, G. A.; Herman, W. N.; Wynne, K. J. J. Am. Chem. Soc. 1998, 120, 11202-11203. Wang, X.; Balasubramanian, S.; Kumar, J .; Tripathy, S. K.; Li, L. Chem. Mater. 1998, 10, 1546-1553. Balasubramanian, S.; Wang, X.; Wang, H. C.; Yang, K.; Kumar, J .; Tripathy, S. K.; Li, L. Chem. Mater. 1998, 10, 1554-1560. Yang, X.; Johnson, S.; Shi, J .; Holesinger, T.; Swanson, B. Sens. Actuators, B 1997, B45, 87-92. 31 (72) (73) (74) (75) (76) (77) (78) (79) (80) (81) (82) (83) (84) (85) (86) (87) (88) (89) (90) (91) Lowy, D. A.; F inklea, H. O. Electrochim. Acta 1997, 42, 1325-1335. Fou, A. C.; Rubner, M. F. Macromolecules 1995, 28, 7115-7120. Lukkari, J .; Vinikanoja, A.; Salomaki, M.; Aaritalo, T.; Kankare, J. Synth. Met. 2001, 121, 1403-1404. Zheng, H.; Rubner, M. F .; Hammond, P. T. Langmuir 2002, 18, 4505-4510. Dai, J.; Sullivan, D. M.; Bruening, M. L. Ind. Eng. Chem. Res. 2000, 39, 3528— 3535. F arhat, T. R.; Schlenoff, J. B. Electrochem. Solid-State Lett. 2002, 5, 813-815. Krasemann, L.; Tieke, B. Mater. Sci. Eng. C 1999, C8-C9, 513-518. Krasemann, L.; Tieke, B. Langmuir 2000, 16, 287-290. Stair, J. L.; Harris, J. J.; Bruening, M. L. Chem. Mater. 2001, 13, 2641-2648. Stroeve, P.; van Os, M.; Kunz, R.; Rabolt, J. F. Thin Solid Films 1996, 284—285, 200-203. Krasemann, L.; Tieke, B. Chem. Eng. Tech. 2000, 23, 211-213. Krasemann, L.; Tieke, B. J. Membr. Sci. 1998, 150, 23-30. Meier-Haack, J .; Lenk, W.; Lehmann, D.; Lunkwitz, K. J. Membr. Sci. 2001, 184, 233-243. Levaesahni, J .-M.; McCarthy, T. J. Macromolecules 1997, 30, 1752-1757. Dai, J .; Bruening, M. L. Nano Letters 2002, 2, 497-501. Farhat, T. R.; Schlenoff, J. B. Langmuir 2001, 1 7, 1184-1192. Von Klitzing, R.; MOhwald, H. Thin Solid Films 1996, 284-285, 352-356. Qiu, X.; Donath, E.; MOhwald, H. Macromolecular Materials and Engineering 2001, 286, 591-597. Antipov, A. A.; Sukhorukov, G. B.; Donath, E.; MOhwald, H. J. Phys. Chem. B 2001, 105, 2281-2284. Schlenoff, J. B.; Ly, H.; Li, M. J. Am. Chem. Soc. 1998, 120, 7626-7634. 32 (92) (93) (94) (95) (96) (97) (98) (99) (100) (101) (102) (103) (104) (105) (106) (107) (108) (109) (110) Riihe, J .; Knoll, W. Supramolecular Polymers 2000, 565-613. Zhao, B.; Brittain, W. J. Prog. Polym, Sci. 2000, 25, 677-710. Uyama, Y.; Kato, K.; Ikada, Y. Adv. Polym. Sci. 1998, 13 7, 1-39. Milner, S. T. Science 1991, 251, 905-914. Ito, Y.; Nishi, S.; Park, Y. S.; Iminishi, Y. Macromolecules 1997, 30, 8096. Mayes, A. M.; Kumar, S. K. MRS Bull. 1997, 22, 43-47. Halperin, A.; Tirrell, M.; Lodge, T. P. Adv. Poly. Sci, 1991, 100, 31-71. Nagale, M.; Kim, B. Y.; Bruening, M. L. J. Am. Chem. Soc. 2000, 122, 11670- 11678. Belder, G. F .; ten Brinke, G.; Hadziioannou, G. Langmuir 1997, 13, 4102-4105. Jordan, R.; Ulman, A.; Kang, J. F.; Rafailovich, M. H.; Sokolov, J. J. Am. Chem. Soc. 1999, 121, 1016-1022. _ Mansky, P.; Liu, Y.; Huang, E.; Russell, T. P.; Hawker, C. Science 1997, 275, 1458-1460. Matyjaszewski, K. In ACS Symposium Series; Matyjaszewski, K., Ed.; American Chemical Society: Washington DC, 1998; Vol. 685, pp 2-30. Szwarc, M. Nature 1956, I 78, 1168-1169. Van Beylen, M.; Bywater, S.; Smets, G.; Szwarc, M.; Worsfold, D. J. Adv. Polym. Sci. 1988, 86, 87-143. Wang, J.-S.; Matyjaszewski, K. Macromolecules 1995, 28, 7901-7910. Kato, M.; Karnigaito, M.; Sawamoto, M.; Higashimura, T. Macromolecules 1995, 28, 1721-1723. Husseman, M.; Malmstrtim, 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. Baum, M.; Brittain, W. J. Macromolecules 2002, 35, 610-615. Tsujii, Y.; Ejaz, M.; Sato, K.; Goto, A.; Fukuda, T. Macromolecules 2001, 34, 8872-8878. 33 (111) (112) (113) (114) (115) (116) (117) (118) (119) (120) (121) (122) (123) (124) (125) (126) (127) (128) Matyjaszewski, K.; Xia, J. Chem. Rev. 2001, 101, 2921-2990. Wei, M.; Xia, J .; Matyjaszewski, K. Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 1997, 38, 233-234. Wei, M.; Xia, J .; McDermott, N. E.; Matyjaszewski, K. Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 1997, 38, 231-232. Granel, C.; Dubois, P.; Jerome, R.; Teyssie, P. Macromolecules 1996, 29, 8576- 8582. Uegaki, H.; Kotani, Y.; Kamigaito, M.; Sawamoto, M. Macromolecules 1997, 30, 2249-2253. Matyjaszewski, K.; Xia, J. Chem. Rev., 101, 2921-2990. Kim, J .-B.; Bruening, M. L.; Baker, G. L. J. Am. Chem. Soc. 2000, 122, 7616- 7617. Huang, W.; Kim, J .-B.; Bruening, M. L.; Baker, G. L. Macromolecules 2002, 35, 1 175-1 179. . Huang, W.; Baker, G. L.; Bruening, M. L. Angew. Chem. Int. Ed. 2001, 40, 1510- 1512. Matyjaszewski, K.; Miller, P. J .; Shukla, N.; Irnmarapom, B.; Gelman, A.; Luokala, B. B.; Siclovan, T. M.; Kickelbick, G.; Vallant, T.; Hoffrnann, H.; Pakula, T. Macromolecules 1999, 32, 8716-8724. Jones, D. M.; Huck, W. T. S. Adv. Mater. 2001, 13, 1256-1259. Shah, R. R.; Merreceyes, D.; Husemann, M.; Rees, 1.; Abbott, N. L.; Hawker, C. J .; Hedrick, J. L. Macromolecules 2000, 33, 597-605. Huang, X.; Wirth, M. J. Macromolecules 1999, 32, 1694-1696. Huang, X.; Doneski, L. J .; Wirth, M. J. Anal. Chem. 1998, 70, 4023-4029. Huang, X.; Doneski, L. J .; Wirth, M. J. Chemtech 1998, 28, 19-25. Jia, J.; Baker, G. L. J. Polym. Sci, Part B: Polym. Phys. 1998, 36, 959-968. Eisenbach, C. D. Polymer 1980, 21, 1175-1179. Graham, T. Philos. Mag. 1866, 32, 401-420. 34 (129) (130) (131) (132) (133) (134) (135) (136) (137) (138) (139) (140) Loeb, S.; Sourirajan, S. Adv. Chem. Ser. 1963, 38, 117-132. Henis, J. M. S.; Tripodi, M. K. J. Membr. Sci. 1981, 8, 233-246. Fritzsche, A. K. Polymer News 1988, 13, 266-274. Meyer, H. S.; Gamez, J. P. In Proceedings of the Laurance Reid Gas Conditioning Conference, 1995; Vol. 45th, pp 284-307. Wright, C. T.; Paul, D. R. Polymer 1997, 38, 1871-1878. Raymond, P. C.; Paul, D. R. J. Polym. Sci, Part B: Polym. Phys. 1990, 28, 2079- 2102. Yoshikawa, M.; Fujimoto, K.; Kinugawa, H.; Kitao, T.; Ogata, N. Chem. Lett. 1994, 243-246. Arnold, M. E.; Nagai, K.; Freeman, B. D.; Spontak, R. J .; Betts, D. E.; DeSimone, J. M.; Pinnau, I. Macromolecules 2001, 34, 5611-5619. Koros, W. J .; Fleming, G. K. J. Membr. Sci. 1993, 83, 1-80. Uhlhorn, R. J. R.; Keizer, K.; Burggraaf, A. J. J. Membr. Sci. 1989, 46, 225-241. Koresh, J. E.; Soffer, A. Sep. Sci. T echnol. 1987, 22, 973-982. Way, J. D.; Roberts, D. L. Sep. Sci. T echnol. 1992, 27, 29-41. 35 CHAPTER 2 Enhancing the Anion-Transport Selectivity of Multilayer Polyelectrolyte Membranes by Templating with Cu2+ 2.1 Introduction Alternating adsorption of polyanions and polycations is an attractive method for forming ultrathin separation membranes because of its versatility and simplicity."2 Previous studies of multilayer polyelectrolyte membranes (MPMS) deposited on porous supports showed selective separation of monovalent and divalent ions,3'5 modest gas separations3'f”9 and highly selective pervaporation.3’8’10’ll This study focuses on enhancing the anion-transport selectivities of MPMS by increasing their fixed negative charge density through templating with Cu”. Introduction of ion-exchange sites occurs due to partial Cu2+ complexation by the carboxylate groups of poly(acrylic acid) (PAA) during the deposition of FAA/poly(allylamine hydrochloride) (PAH) membranes. Removal of Cu2+ in acidic solution and subsequent-deprotonation of -CO0H groups yields fixed -C00’ ion-exchange sites as shown in Figure 2.1. Studies of ion transport through Cu2+-templated PAA (PAA-Cu)/PAH membranes show a 4-fold increase in CI' /S042' selectivities compared to pure PAA/PAH membranes deposited under Similar conditions. Cross-linking];l3 of templated films through heat-induced arnidation can yield Cl'/SO42' selectivities as high as 610, and these selectivities can be achieved without a diminution in flux relative to PAA/PAH membranes that are not templated or cross- linked. Such separations of ions of different valence are important in applications such as 14-16 17,18 removal of harmful ions from water, water softening, production of edible 36 O = Cu2*(COO')2 Figure 2.1: Preparation of Cu2 - templated2 polyelectrolyte films on porous alumina supports. Step 1-adsorption of partially Cu2 -complexed PAA ozn porous alumina. Step 2- adsorption of a polycation (PAH). Step 3-removal of Cu2 .Step 4- deprotonation of the free carboxylic acid groups of PAA. Repetition of steps 1 and 2 produces multilayer films. lntertwining of layers is not Shown for figure clarity. 37 salt from sea water,19 and prevention of fouling by cooling water.20 Most of these applications require high permselectivity among different ions as well as high flux. Previous research on ion-exchange membranes Showed that permselectivity can be ”’24 or surface controlled by alteration of hydrophilicity/hydrophobicity,2I’22 cross-linking charge density.25 To achieve maximum efficiency in ion-separation processes, a minimal membrane thickness is also vital for achieving high flux. Multilayer polyelectrolyte films (MPFS) are attractive for ion separation membranes in part because of their minimal thickness. Several studies on ion-exchange membranes already showed an increase in monovalent/ divalent ion selectivities after adsorption of one layer of polyelectrolyte to the surface of the membrane.”27 However, even better perrnselectivities might be obtained when membranes are exclusively composed of multilayer polyelectrolyte films (MPFS). MPFS are attractive as separation membranes because both their thickness and surface charge density can, in principle, be controlled by varying deposition conditions 30-34 such as pH,”29 salt concentration, and the number of adsorbed layers.35 Recent ”’13 that studies Show that MPMS exhibit monovalent/divalent ion-transport selectivities are at least in part due to Donnan exclusion at the charged surface layer of the polyelectrolyte films. Control over the charge density either at the surface or in the bulk of MPMS should thus yield control over membrane selectivity. Previous characterization‘1’6’37 of MPFS showed, however, that the bulk of the film is intrinsically charge compensated (i.e., polycations exactly neutralize the charge on polyanions), and total exchangeable charge resides only at the surface of the film. Enhancement of ion- 38 transport selectivity by introduction of charge into the bulk of MPFS will likely require post-deposition film modification. In this study, we use Cu2+-complexed PAA38'40 to control the charge density within PAA/PAH films. Several groups recently integrated metallosupramolecular polyelectrolytes into polyelectrolyte assembliesims For example, Kurth and coworkers used iron-coordinated terpyridine as a polycation to form MPFS with poly(styrene sulfonate).45 However, in the present case, we only partially complex PAA with Cu”, so that PAA is still deposited as a polyanion. This allows the introduction of cation- exchange sites after removal of Cu”, and formation of highly selective membranes. Diffirsion-dialysis studies and simple modeling of transport through these membranes suggest that selectivity is due to both Donnan exclusion and diffirsional selectivity. 2.2 Experimental 2.2.1 Chemicals and Solutions Poly(allylamine hydrochloride) (PAH) (MW = 70,000), poly(acrylic acid) (PAA) (MW = 2,000) and 3-mercaptopropionic acid (MPA) were used as received from Aldrich. We used a relatively low molecular weight PAA to increase its solubility when complexed with Cu”. NaCl, CuCl2'5H20, and Na2SO4 were used as received from Spectrum. AnodiscTM porous alumina membranes with 0.02 rim-diameter surface pores (Whatrnan Anodisc) were used as supports for deposition of polyelectrolyte films.“47 For cross-linked PAA/PAH membranes, the outer polypropylene support ring of the alumina membrane was burned off prior to film deposition by heating at 400 °C for 18-20 hours. This was done in order to prevent melting of the polymer ring into the pores of the 39 membrane during heat-induced cross-linking. Gold slides (200 nm of sputtered Au on 20 nm Cr on Si (100) wafers) were used as substrates for ellipsometry, external reflection FTIR spectroscopy and cyclic voltammetry. 2.2.2 Film Preparation Prior to film deposition, porous alumina substrates were cleaned in a UV/ozone cleaner (Boekel UV-Clean model 135500) for 15 min. Deposition ofpolyelectrolyte films began by dipping the substrate into a solution of PAA (0.04 M with respect to the repeating unit) containing CuCl2 (2.5mM to 7.5 mM) for 5 min followed by rinsing for 1 min with water (Milli-Q, 18 MQ-cm). The substrate was then immersed in a solution of PAH (0.04 M with respect to the repeating unit and containing the same CuCl2 concentration as PAA) for 5 min and rinsed with» water for l min. The above procedure was repeated until the desired number of bilayers was deposited. The pH values of the deposition solutions were adjusted to 5.5, 6, or 6.6 using dilute HCl and NaOH solutions. Both PAH and PAA solutions contained 0.5 M NaCl as a supporting electrolyte to increase film thickness, and PAH and PAA depositions were always done at the same pH. The porous alumina membrane is asymmetric such that the permeate side contains 0.2 rim-diameter pores, while pores on the filtrate side are 0.02 um in diameter. Deposition of polyelectrolytes was limited to the filtrate side by using an o-ring holder. After deposition of the desired number of layers, membranes were rinsed well with water and dried with N2. For cross-linking, membranes were placed in a flask that was subsequently purged with N2 for 30 min and then slowly heated to the desired temperature (~45 min ramping time). Heating continued at the desired temperature (100- 160 °C) for an additional 2 hours under N2 purging. 4O Similar to porous alumina supports, gold slides were UV/ozone cleaned prior to deposition. However, before polyelectrolyte adsorption, slides were immersed in an ethanolic solution of 2 mM MPA for 30 min and rinsed well with ethanol followed by deionized water (Milli-Q, 18 MQ-cm). This produces a carboxylic acid-containing monolayer on the surface that will be charged upon deprotonation. The polyelectrolyte deposition on gold was the same as for porous alumina except that depositions started with PAH rather than PAA. We used MPA rather than a long-chain alkanoic acid to avoid possible blocking of electron transfer at the gold surface.“49 2.2.3 Film Characterization Film thickness was determined with a rotating analyzer ellipsometer (J .A. Wollam model M-44), assuming a film refractive index of 1.5. The refractive index and absorption coefficient of the substrates were determined after deposition of the MPA layer. For each sample, thicknesses at three different spots were taken. External reflectance FTIR spectra were obtained with a Nicolet Magma-560 FTIR Spectrometer using a Pike grazing angle attachment (80° angle of incidence). The Spectrometer employs a MCT detector. Electrochemical measurements were performed with a CH- Instrument Electrochemical Analyzer (model 605) employing a standard three-electrode cell containing a Ag/AgCl (3M KCl) reference electrode and a Pt wire counter electrode. The working electrode was a gold slide in a plastic holder that exposed an area of 0.1 cm2. The supporting electrolyte in all electrochemical experiments was 0.1M Na2SO4. Before measuring cyclic voltarnmograms (scan rate of 0.1 V/s), solutions were purged with N2 for 20-30 min. 41 2.2.4 Ion-Transport Studies. Diffusion dialysis was performed using two glass half cells as shown in Figure 2.2. The membrane was clamped between the two half cells, with the film side of the membrane facing towards the feed cell, so as to expose 2 cm2 of membrane to the feed solution. The permeate cell contained deionized water (90 mL), and the feed cell contained 0.1 M solutions (90 mL) of NaCl or Na2SO4. After dialysis with a particular salt, the apparatus was washed well with water, and both cells were equilibrated with deionized water for 30 min before examining the next salt. Alternating NaCl (pH 5.3) and Na2S04 firH 5.6) transport experiments were performed until two successive chloride fluxes matched within 15%. Once the membrane achieved a steady Cl' flux value (this usually occurs after 4 to 5 NaCl and Na2S04 runs), it was removed from the permeability apparatus and dipped in pH 3.5 water (dilute HCl solution) for one hour to ensure that all the copper was removed from the membrane. Then the membrane was immersed in deionized water (adjusted to pH 5-6 with a dilute NaOH solution) for 1-2 hours to deprotonate the -COOH groups that were created upon removal of Cu“. Permeability experiments were then repeated, and Cl' fluxes differed by less than 5% when performed before and after a Na2SO4 permeability experiment. The C1‘/S042' selectivities and flux values reported here are calculated exclusively from permeability studies performed directly after immersing in pH 3.5 water followed by pH 5-6 water. (In fact, the initial conditioning runs are probably not necessary). 42 Conductivity meter Mow.m:rpr-lto.dwb Mechanical‘ Stirrer . fl \1 ‘1‘. \ J l l Receiving Phase Water (90 mL) 3:117 2/ / —L Source Phase 0.1 M Salt Solution Membrane (90 mL) Figure 2.2: Apparatus for ion-transport measurements. 43 Magnet The permeate-cell conductivity values were converted to concentration using a calibration plot of conductivity versus concentration for a particular salt. The flux (J) for each permeating ion was calculated using equation 2.1, and the selectivity (a) of one ion over the other was obtained from equation 2.2. AQK ,, At A ' *i 22 or J2 . In these equations, AC/At is the concentration change in the receiving cell with time obtained from the Slope of a plot of concentration versus time; V is the volume of the solution in the receiving cell after 90 min, A is the exposed surface area of the membrane; and subscripts 1 and 2 refers to the two different permeating ions. 2.3 Results and Discussion 2.3.1 Synthesis and Characterization of Cu2+-Templated PAA/PAH Films. Figure 2.1 shows schematically the preparation of Cu2+-templated PAA/PAH films on porous alumina supports. The procedure begins by preparing PAA complexed with Cu2+. To do this, we employ a PAA repeating unit to Cu2+ ratio of 8:1 so that ~25% of the -C00' groups of PAA will be complexed with Cu2+ (two -C00' groups should bind with one Cu2+ ion). The pH of the solution must be around 5.5 so that —C00' groups are mostly deprotonated and Cu(0H)2 does not precipitate. Alternating adsorption of the CUB-complexed PAA (uncomplexed -C00' groups allow PAA to act as a polyanion) and PAH molycation) results in a MPF. Although UV/visible Spectroscopy suggests that PAH does not form a complex with Cu2+ at pH 5.5,50 we use the same Cu2+ concentration in PAH deposition as for PAA to prevent leaching of Cu2+ fiom the deposited PAA-Cu layer during immersion in the PAH solution. After deposition of the desired number of layers, we expose films to an HCl solution (pH 3.5) to exchange protons for Cu2+ and create free -CO0H groups on PAA chains. Subsequent immersion in a pH 5.5 solution deprotonates these -CO0H groups (exchange of protons for Na+) and increases the fixed negative charge density in the bulk of the fihn. The Cu2+-templated PAA/PAH films differ from pure PAA/PAH films in that they contain -C00' groups that are electrically compensated by mobile cations (N a+) rather than neighboring ammonium groups of PAH. For cross-linked films, we use the same deposition procedure (Figure 2.1), except the removal of Cu2+ and the deprotonation of CO0H groups (Steps 3 and 4 in Figure 2.1) occur after heating the films for two hours. Heating results in the formation of amide cross-links from -C00'-NH3,+ pairs.12'5"52 Cyclic voltammetry (Figure 2.3) of PAH/PAA-Cu films deposited on gold wafers confirms the presence of Cu” in these films as well as its removal at low pH. The peaks due to Curr/Cu completely disappear after immersion of the electrode in water at pH 3.5 (pH adjusted with 0.1 M HCl). Integration of the reduction or oxidation peak allows estimation of the amount of Cu2+ in the film, and this proves usefirl in modeling of ion transport (vide infra). As a comparison, we also tried to put Cu2+ into a 10-bilayer PAH/PAA film (deposited under similar conditions) by immersing the film in a 0.1 M CuCl2 solution for 20 hours.53 Cyclic voltammetry of this film showed that the amount of adsorbed /absorbed Cu2+ is about 1/6 of that in a Cu2+-temp1ated film. 45 Current l I 0.6 0.4 0.2 0.0 -0.2 Potential (V vs. Ag/AgCl) Figure 2.3: Cyclic voltammetry of a MPA-modified gold electrode coated with 10 bilayers of PAH/PAA-Cu before (solid line) and after exposure to pH 3.5 water (dotted line). 46 Reflectance FTIR spectra also confirm templating of PAH/PAA-Cu films with Cu“. The spectrum of a PAH/PAA-Cu film (spectrum a, Figure 2.4) shows a broadening of the -C00' symmetric stretch compared with the Spectrum of a pure PAH/PAA film (spectrum (1, Figure 2.4). This broadening results from counter-ion-induced changes in the energy of the -C00' stretch.54 Upon exposure to pH 3.5 water and removal of Cu2+ (Spectrum b, Figure 2.4), the -C00' symmetric stretch looks like that of a pure PAH/PAA film. Further, a 50% increase in the acid carbonyl peak (1715 cm'l) after immersion in pH 3.5 water suggests that lowering of pH creates free -CO0H groups from the Cu“ complexes, as would be expected. Immersing the film in pH 5-6 water deprotonates -CO0H groups and results in a decrease in the acid carbonyl peak (spectrum 0, Figure 2.4). 2.3.2 Anion Transport through Cu21-Templated PAA/PAH Membranes. Ion-transport studies Show that PAA-Cu/PAH membranes on porous alumina supports are significantly more selective than similar pure PAA/PAH membranes. Figure 2.5 shows a plot of receiving-phase concentration as a function of time for membranes sandwiched between deionized water (receiving phase) and 0.1 M NaCl or Na2SO4 (source phase). These plots Show that the Cl' flux through both Cu2+-templated and pure PAA/PAH membranes is about 40 % of that through bare porous alumina. However, 10.5-bilayer PAA-Cu/PAH membranes (the top layer in the film is PAA-Cu) show a 4- fold decrease in S042‘ flux relative to pure 10.5-bilayer PAA/PAH membranes as Shown in the inset of Figure 2.5. Overall, Cl'/S042' selectivity increases 4-fold due to templating of 10.5-bilayer films (Table 2.1). 47 Absorbance (b) COO-Cu2+ I l I l 1800 1700 1600 1500 1400 1300 Wavenumbers (cm4) Figure 2.4: External reflectance FTIR spectra of (a) a 10-bilayer PAI-I/PAA-Cu film, (b) the same film after exposure to pH 3.5 water, (c) the film after subsequent exposure to pH 5.5 water, and (d) a 10- bilayer PAH/PAA film deposited without Cu2+. All films were deposited on a gold wafer coated with a monolayer of MPA. 48 3.5 0.08 " 0.06 E 3.0 - v 0.04 8 2:: - 0.02 S 2.5 ‘g' 0.00 g 2.0 - O o 8 1.5 - m .C o. P’ 1 0 r ’ c . 0 Q’ (I) 0.5 ‘ “I m Q/ A/ /O/ .. f. m if, 0.0/L1 123 1! El 1931 1.1 H O 51015202530354045 Time (min) Figure 2.5: Receiving phase concentration as a function of time for a bare porous alumina membrane (black), an alumina membrane coated with 10.5 bilayers of Cu21-templated PAA/PAH (open) and an alumina membrane coated with 10.5 bilayers of PAA/PAH (grey). Different symbols represent different salts: circles-NaCl and squares-Na2S04. The inset Shows 8042’ concentration vs. time for pure PAA/PAH (grey) and templated PAA/PAH (open). The membrane separates the receiving phase (initially deionized water) from a 0.1 M salt solution. 49 2.3.3 Cross-linked Cu2+-Templated PAA/PAH Membranes. One possible limitation to Cl’/SO42' selectivity is that swelling in water may decrease charge density and reduce Donnan exclusion of S0421 In an effort to limit film swelling, we cross-linked PAA-Cu/PAH films by heating under N2 to form amide bonds through reaction of the ammonium groups of PAH and the carboxylate groups of PAA that are not complexed with Cu“. Reflectance FTIR spectroscopy confirms that the cross-linking reaction occurs.12 After heating at 160 °C, external reflectance FT IR spectra of 10-bilayer PAH/PAA-Cu films show a large reduction in the intensity of —- C00' peaks at 1570 cm'1 and 1400 cm’1 and the appearance of amide peaks at 1660 cm’1 and 1550 cm'1 (Figure 2.6). With lower heating temperatures, the amide peaks are more clearly visible after exposing the cross-linked films to low-pH solutions because peaks due to Cu21-C00‘ complexes also appear in this region of the spectrum. The degree of cross-linking depends greatly on heating temperature as indicated by amide peaks that increase with cross-linking temperature.12 Diffusion dialysis studies Show that as heating temperature (and hence the degree of cross-linking) increases, Cl'/S042' selectivity increases and then peaks at a heating temperature of 130 °C (Table 2.1, 10.5-bilayer PAA-Cu/PAH films). Partially cross- linked 10.5-bilayer PAA-Cu/PAH membranes (130 °C) show a 10-fold increase in C1” /S042’ selectivity relative to unheated, templated membranes, and this increase is achieved with only a 20% decrease in C1' flux. At higher cross-linking temperatures, 50 Table 2.1: Anion fluxesa (moles cm'zs'l) through bare porous alumina and alumina coated with PAA/PAH and PAA-Cu/PAH films cross-linked at different temperatures. Film Composition Cross-linking 108 x (31' 108 x S042' Cl'/S.0..;2'c T ( C) Flux Flux Selectivrty Bare - 4.211 3.3102 1.310.09 10 PAA/PAH - 1.0102 1.5102 0.710.03 10 PAA-Cu/PAHb - 2.3103 02710.01 911 10.5 PAA/PAH - 1.3104 01110.05 1313 10.5 PAA-Cu/PAHb - 1.6102 0.0310001 5513 10.5 PAA-Cu/PAHb 100 1.6105 00210004 80115 10.5 PAA-Cu/PAHb 120 1.4102 0.00610002 240180 10 PAA/PAH 130 0091003 002710005 310.4 10 PAA-Cu/PAHb 130 2.010.06 003210.005 62111 10.5 PAA/PAH 130 00710.01 00028100009 2617 10.5 PAA-Cu/PAH" 130 1.31005 00021100001 610120 10.5 PAA-Cu/PAH" 140 05110.2 00016100005 330170 10.5 PAA-Cu/PAH" 160 00871004 00003100001 2911.3 3‘ Flux values were calculated from the Slopes of plots of concentration in the receiving phase vs. time. Errors represent standard deviations of at least three measurements. Flux was measured after removal of Cu2+ from the membrane and deprotonation of newly formed —COOH groups. ° Calculated as the average of selectivity values for each membrane and not from average flux values. 51 0.01 C00- Absorbance fmide l Amide II (b) 1 800 1 700 1600 1 500 1400 Wavenumbers (cm-‘) Figure 2.6: External reflectance FTIR spectra of 10-bilayer PAH/PAA- Cu films (a) before and (b) after heating at 160°C. Films were deposited on MFA-coated gold. 52 Cl" flux drops rapidly, presumably due to a tighter membrane structure. Sulfate flux does not continue to drop significantly at higher cross-linking temperatures, and thus Cl'/S042' selectivity eventually decreases. Compared with pure PAA/PAH membranes, the selectivities and fluxes through partially cross—linked PAA-Cu/PAH membranes are remarkable. Table 2.1 shows that partially cross-linked (130 °C) 10.5-bilayer PAA- Cu/PAH membranes Show a 20-fold increase in Cl'/SO42’ selectivity relative to similar cross-linked pure PAA/PAH membranes. Additionally, the Cl' flux through these cross- linked Cu2+-templated membranes is 20-fold higher than the Cl' flux through pure PAA/PAH membranes cross-linked at the same temperature. This may be due to the formation of new transport pathways upon removal of Cu2+ or a lower degree of cross- linking in the templated film. 2.3.4 Changing the Surface Charge of Membranes. Our previous studiess’l3 of MPMS showed that much of the ion-transport selectivity in these systems is due to a high charge density at the membrane surface. However, Cu2+-templated membranes differ from previous MPMS in that they contain fixed charge throughout the membrane. In an effort to understand more about selectivities in PAA-Cu/PAH membranes, we changed the terminating layer of these films from PAA to positively charged PAH. If selectivity in these systems is largely due to charge at their surface, changing the outer layer from a polyanion to a polycation Should have a dramatic effect on ion transport. Changing the surface from PAA-Cu (10.5-bilayer films) to PAH (IO-bilayer films) in cross-linked (130 °C), templated films resulted in a 15-fold increase in 8042' flux and a 50% increase in CI‘ flux (Table 2.1). Thus Cl’/SO42' selectivity decreased from 53 610 to 60 on going from a 10.5-bilayer to a lO—bilayer cross-linked PAA-Cu/PAH film. In the case of unheated Cu2+-templated membranes, terminating with PAH rather than PAA-Cu yielded a decrease in C1'/SO42' selectivity from 55 to 9. These data clearly indicate that Donnan exclusion at the film surface plays a large role in determining selectivity. For unheated, pure PAA/PAH membranes, selectivity actually reverses (from 13 to 0.7) upon changing the top layer from PAA to PAH. However, with Cu“- templated fihns, we still see a significant Cl’/SO42’ selectivity when the surface of the membrane is positively charged because of fixed negative charge density in the bulk of the membrane. 2.3.5 Anion-Transport through Partially Cross-linked, Cu2+-Templated Membranes Deposited at Different pH values. Variation of the pH at which PAA-Cu/PAH films are deposited allows some control over the amount of Cu2+ in these films and may provide a means for controlling transport selectivity. Table 2.2 gives the Cl'/S042' selectivity values for partially cross- linked 10.5-bilayer PAA-Cu/PAH membranes deposited from solutions at three different pH values (5.5, 6, and 6.6). We also tried to deposit membranes at pH values <5.5, but under these conditions, polymer precipitates from deposition solutions. Selectivity is highest for films deposited at pH 5.5 and decreases at higher deposition pH values. The ellipsometric thicknesses (Table 2.2) of Similar films on gold wafers are independent of pH over this range of values,55 so selectivity differences are likely due to changes in charge density. The UV/visible Spectra of the PAH/Cu2+ solutions showed a shift in the 54 Table 2.2: Thicknesses, fluxes (moles cm'zs’1), selectivities and Cu2+ concentrations (M) in partially cross-linked, 10.5-bilayer PAA-Cu/PAH films deposited at different pH values. Deposition Thickness” cr Fluxc 8042' Flux° Cl‘/SO42‘ Cu2+ pHal (A) x 108 x 101 l Selectivityd Concentratione 5.5 170i3 l.3i0.05 2.1i0.1 610i20 1.032006 6 1703.6 1.1i0.2 2.6i0.4 430:90 0.9i0.04 6.6 170i10 1.63:0.3 170i42 11i4 0.3i0.1 aBoth PAA-Cu and PAH were deposited at this pH. bThiclmesses are for lO-bilayer PAH/PAA-Cu films deposited on gold wafers as described in the experimental section. CError values represents standard deviations. dCalculated as the average of selectivity values for each membrane and not from average flux values. eCu2+ concentration in the membrane was estimated from the area of the reduction peak in a cyclic voltammagram of a film on gold. This area was converted to number of moles of Cu21/cm2, and this value was divided by the ellipsometric film thickness to obtain the concentration. 55 Cu2+ absorption peak from 820 nm at pH 5.5 to 780 nm at pH 6 and to 710 nm at pH 6.6, suggesting that the amine groups of PAH begin to form complexes with Cu2+ at the higher pH values. In addition, at higher pH values Cu2+ can form hydroxide complexes. These competing reactions probably reduce the amount of Cu2+ deposited in the membrane as -C00'-Cu2+ complexes. To quantitatively investigate fixed-charge density in PAH/PAA-Cu membranes, we employed cyclic voltammetry to estimate Cu2+ concentrations in analogous films deposited on gold (Figure 2.7). By integrating the area of the reduction peak, Cu2+ concentrations could be estimated. In agreement with transport studies, Table 2.2 Shows that the maximum amount of copper is deposited at pH 5.5. Hence, after removal of Cu2+ from the film, higher charge densities should enhance Cl'/S042' selectivity for films deposited at the lower pH values. 2.3.6 Anion Transport through Partially Cross-linked, Cu21-Templated Membranes Deposited with Different Cu 1 Concentrations. Altering the amount of Cu2+ present during deposition should provide another means for controlling fixed charge and selectivity in membranes. To examine this possibility, we prepared cross-linked PAH/PAA-Cu membranes using Cu2+ concentrations of 2.5 mM, 5 mM or 7.5 mM in both PAA and PAH deposition solutions. Table 2.3 gives the Cl'/SO42' selectivities and flux values for these membranes. (We deposited PAA and PAH at pH 6 because at pH 5.5, higher Cu2+ concentrations resulted in precipitation). 56 Current L 0.6 0.4 0.2 0.0 -0.2 -0.4 -0.6 Potential (V vs. Ag/AgCl) Figure 2.7: Cyclic voltammetry of lO-bilayer PAH/PAA-Cu films deposited at different deposition pH values on MPA-modified gold surfaces: pH 5.5-dashed line, pH 6-solid line, and pH 6.6-dotted line. Areas of the reduction peaks were calculated by drawing the baseline from the value of current at a potential of 0.3 V to the value of current at -0.5 V. 57 $503738”. 8% 35308 832235 65680326 use weflezameo 80.39 90:88:? 0:93 momma eoEEHBoe £3 countenance $39. .5308 3368536 05 E eontomoe mm 158: N com 06 02 on weifiTmmBo coca Boom? Bow :0 ago :O-<<10‘9 1.4x10‘9 1.4x10‘3 1.1><10‘2 6 9 100 8.8><10‘9 1.1x10‘9 1.5><10‘3 1.2x10'2 8 10 120 7.4x10'9 3.6><10"° 1.6><10'3 1.3><10‘2 20.5 11.7 130 6.8><10‘9 1.5><10‘l0 2.0x10'3 1.6x10'2 45 13.6 140 2.4><10‘9 1.1x10"° 1.9><10'3 1.5><10‘2 22 15 160 4.1><10‘l0 2.2x10“° 2.1><10‘3 1.7><1o'2 1.7 17 a Determined using cyclic voltammetry and ellipsometry. b The diffusional selectivity is the ratio of diffusion coefficients obtained from the model. c Total Cl'/SO42' selectivity divided by diffusional selectivity. 67 These simple calculations suggest that the highest selectivities in cross-linked films are about equally due to diffusion and Donnan selectivities. We should note, however, that this model does not take into account activity coefficients or the effect of hydrophobicity on partitioning. Several previous studies showed that diffusion through charged membranes can be complicated due to electrostatic interactions between the membrane and the diffusing ions.“68 Additionally, charge distributions in our simulations are oversimplified and only approximate. However, the modeling studies do strongly indicate that selectivity is only partly due to Donnan exclusion. A full understanding of transport through MPMS will likely require measurement of diffusion and partition coefficients. 2.4 Conclusions Partial complexation of the -C00' groups of PAA with Cu2+ provides a convenient method to enhance fixed negative charge density in MPMS. Removal of Cu2+ leaves behind -COOH groups that behaves as ion-exchange sites. Diffusion-dialysis studies with Cu2+-templated membranes show that templating increases anion-transport selectivities, and post deposition cross-linking of these membranes further enhances Cl' /SO42' selectivities to values high as 610. Changing the surface layer from negatively charged PAA to positively charged PAH greatly reduces Cl'/SO42' selectivity, showing that selectivity is highly dependent on surface charge. Simulation of ion-transport data using a simple two-layer model of MPFS suggests that the observed Cl'/SO42' selectivities are due to both Donnan exclusion and differences in diffiisivities of ions. 68 2.5 References and Notes (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) (17) (18) (19) Decher, G.; Hong, J. D. Ber. Bunsenges. Phys. Chem. 1991, 95, 1430-1434. Decher, G. Science 1997, 277, 1232-1237. Krasemann, L.; Tieke, B. Mater. Sci. Eng, C 1999, 8—9, 513-518. Krasemann, L.; Tieke, B. Langmuir 2000, 16, 287-290. Harris, J. J .; Stair, J. L.; Bruening, M. L. Chem. Mater. 2000, 12, 1941-1947. Levasahni, J .-M.; McCarthy, T. J. Macromolecules 1997, 30, 1752-1757. Stroeve, P.; Vasquez, V.; Coelho, M. A. N.; Rabolt, J. F. Thin Solid Films 1996, 285, 708-712. van Ackem, F .; Krasemann, L.; Tieke, B. Thin Solid Films 1998, 329, 762-766. Kotov, N. A.; Magonov, S.; Tropsha, E. Chem. Mater. 1998, 10, 886-895. Krasemann, L.; Tieke, B. Chem. Eng. T echnol. 2000, 23, 211-213. Krasemann, L.; Tieke, B. J. Membr. Sci. 1998, 150, 23-30. Harris, J. J .; DeRose, P. M.; Bruening, M. L. J. Am. Chem. Soc. 1999, 121, 1978- 1979. Stair, J. L.; Harris, J. J .; Bruening, M. L. Chem. Mater. 2001, 13, 2641-2648. Hell, F.; Lahnsteiner, J .; Frischherz, H.; Baumgartner, G. Desalination 1998, 11 7, 173-180. Amor, Z.; Bariou, B.; Mameri, N.; Taky, M.; Nicolas, S.; Elmidaoui, A. Desalination 2001, 133, 215-223. Sata, T.; Yamaguchi, T.; Matsusaki, K. J. Chem. Soc., Chem. Commun. 1995, 11, 1 153-1 154. Mika, A. M.; Childs, R. F.; Dickson, J. M. Desalination 1999, 121, 149-158. Brett, S. W.; Gaterell, M. R.; Morse, G. K.; Lester, J. N. Environ. T echnol. 1999, 20, 1009-1018. Takata, K.; Yamamoto, Y.; Sata, T. Bull. Chem. Soc. Jpn. 1996, 69, 797-804. 69 (20) (21) (22) (23) (24) (25) (26) (27) (23) (29) (30) (31) (32) (33) (34) (35) (36) (37) (38) (39) (40) Bader, M. S. H. J. Hazard. Mater. 2000, B73, 269-283. Sata, T.; Mine, K.; Higa, M. J. Membr. Sci. 1998, 141, 137-144. Sata, T.; Mine, K.; Tagami, Y.; Higa, M.; Matsusaki, K. J. Chem. Soc., Faraday Trans. 1998, 94, 147-153. Sata, T.; Nojima, s. J. Polym. Sci. Polym. Phys. Ed. 1999,37, 1773-1785. Sata, T.; Emori, S. 1.; Matsusaki, K. J. Polym. Sci. Polym. Phys. Ed. 1999, 37, 793-804. Matsusaki, K.; Hashimoto, N.; Kuroki, M.; Sata, T. Anal. Sci. 1997, 13, 345-349. Tsuru, T.; Nakao, S.-I.; Kimura, S. J. Membr. Sci. 1995, 108, 269-278. Urairi, M.; Tsuru, T.; Nakao, 8.; Kimura, S. J. Membr. Sci. 1992, 70, 153-162. Yoo, D.; Shiratori, S. S.; Rubner, M. F. Macromolecules 1998, 31 , 4309-4318. Shiratori, S. S.; Rubner, M. F. Macromolecules 2000, 33, 4213-4219. Sukhorukov, G. B.; Schmitt, J .; Decher, G. Ber. Bunsenges. Phys. Chem. 1996, 100, 948-953. Sukhishvili, S. A.; Granick, S. J. Chem. Phys. 1998, 109, 6861-6868. Dubas, S. T.; Schlenoff, J. B. Macromolecules 1999, 32, 8153-8160. Ldsche, M.; Schmitt, J .; Decher, G.; Bouwman, W. G.; Kjaer, K. Macromolecules 1998, 31, 8893-8906. Lvov, Y.; Decher, G.; Mohwald, H. Langmuir 1993, 9, 481-486. Decher, G.; Hong, J. D.; Schmitt, J. Thin Solid Films 1992, 210/211, 831-835. Schlenoff, J. B.; Ly, H.; Li, M. J. Am. Chem. Soc. 1998, 120, 7626-7634. Laurent, D.; Schlenoff, J. B. Langmuir 1997, 13, 1552-1557. Gotoh, Y.; Igarashi, R.; Ohkoshi, Y.; Nagura, M.; Akamatsu, K.; Deki, S. J. Mater. Chem. 2000, 10, 2548-2552. Rivas, B. L.; Schiappacasse, N.; Basaez, L. A. Polym. Bull. 2000, 45, 259-265. Porasso, R. D.; Benegas, J. C.; Hoop, M. A. G. T. v. d. J. Phys. Chem. B 1999, 70 (41) (42) (43) (44) (45) (46) (47) (48) (49) (50) (51) (52) (53) (54) (55) (56) (57) 103, 2361-2365. Kurth, D. G.; Osterhout, R. Langmuir 1999, 15, 4842-4846. Caruso, F.; Schuler, C.; Kurth, D. G. Chem. Mater. 1999, II, 3394-3399. Xiong, H.; Cheng, M.; Zhou, Z.; Zhang, X.; Shen, J. Adv. Mater. 1998, 10, 529- 532. Zhang, X.; Shen, J. C. Adv. Mater. 1999, 11, 1139-1143. Schutte, M.; Kurth, D. G.; Linford, M. R.; Colfen, H.; Mohwald, H. Angew. Chem. Int. Ed. 1998, 3 7, 2891-2893. Chen, W.-J.; Aranda, P.; Martin, C. R. J. Membr. Sci. 1995, 107, 199-207. Chen, W.-J.; Martin, C. R. J. Membr. Sci. 1995, 104, 101-108. Dai, Z.; Ju, H. Phys. Chem. Chem. Phys. 2001, 3, 3769-3773. Bain, C. D.; Troughton, E. B.; Tao, Y. T.; Evall, J .; Whitesides, G. M.; Nuzzo, R. G. J. Am. Chem. Soc. 1989, 111, 321-335. At ph 5.5, the UV/vis spectrum of PAH/Cu2+ solution is identical to that of a CuClz solution containing the same Cu2+ concentration. Dai, J .; Sullivan, D. M.; Bruening, M. L. Ind. Eng. Chem. Res. 2000, 39, 3528- 3535. Dai, J .; Jensen, A. W.; Mohanty, D. K; Emdt, J .; Bruening, M. L. Langmuir 2001, 1 7, 931-937. Joly, 8.; Kane, R.; Radzilowski, L.; Wang, T.; Wu, A.; Cohen, R. E.; Thomas, E. L.; Rubner, M. F. Langmuir 2000, 16, 1354-1359. Deacon, G. 3.; Phillips, R. J. Coord. Chem. Rev. 1980, 33, 227-250. Rubner and coworkers reported that thickness of similar MPF's are dependent on deposition pH (reference 29). However, in our case we use a supporting electrolyte (NaCl) and Cu2+ in the deposition solution. Mendelsohn, J. D.; Barrett, C. J .; Chan, V. V.; Pal, A. J .; Mayes, A. M.; Rubner, M. F. Langmuir 2000, 16, 5017-5023. Fery, A.; Schoeler, B.; Cassagneau, T.; Caruso, F. Langmuir 2001, 1 7, 3779- 3783. 71 (5 8) (59) (60) (61) (62) (63) (64) (65) (66) (67) (68) Miyoshi, H. J. Membr. Sci. 1998, 141, 101-110. Manzanares, J. A.; Mafé, 8.; Pellicer, J. J. Phys. Chem 1991, 95, 5620-5624. Lebedev, K.; Ramirez, P.; Mafe, S.; Pellicer, J. Langmuir 2000, 16, 9941-9943. Lowack, K.; Helm, C. A. Macromolecules 1998, 31 , 823-833. To estimate the charge density in the surface layer, we first determined the concentration of Cu2+ in a 10-bilayer film using cyclic voltammetry and ellipsometry and then multiplied this value by a factor of 8 to get the -COO‘/ -COOH concentration in the film. Then we multiplied this value by a factor of 2 because the surface of the film is mostly PAA rather than PAH/PAA. We assumed that -COOH groups were completely deprotonated. Higa, M.; Kira, A. J. Phys. Chem. B 1995, 99, 5089-5093. Peeters, J. M. M.; Boom, J. P.; Mulder, M. H. V.; Strathmann, H. J. Membr. Sci. 1998, 145, 199-209. Sata, T. J. Membr. Sci. 1994, 93, 117-135. Sata, T.; Yamaguchi, T.; Matsusaki, K. J. Phys. Chem 1995, 99, 12875-12882. Beerlage, M. A. M.; Peeters, J. M. M.; Nolten, J. A. M.; Mulder, M. H. V.; Strathmann, H. J. Appl. Polym. Sci. 2000, 75, 1180-1193. Robertson, B. C.; Zydney, A. L. J. Membr. Sci. 1990, 49, 287-303. 72 CHAPTER 3 Preparation of Composite Membranes by Atom Transfer Radical Polymerization Initiated from a Porous Support 3.1 Introduction Synthesis of practical separation membranes requires methods for creating thin, selective skins at the surface of highly permeable supports."5 Composite membranes prepared by depositing an expensive skin on a relatively inexpensive support are especially attractive in this regard because only a small amount of the selective skin material is needed.6’7 The most common methods for formation of composite membranes include interfacial polymerizationf"9 casting,'0plasma polymerization,11 and solution coating.12 Even with the successes of these methods, synthesis of selective membrane skins with minimal (<50 nm) thicknesses is still difficult.3'13 This chapter describes our initial investigations into the possibility of using polymerization from a surface to create ultrathin membrane skins with unique structures. Many recent studies demonstrate the use of controlled polymerization techniques to grow polymer chains from surfaces in a well-defined way.”23 These procedures generally involve attachment of polymerization initiators to a surface and subsequent polymerization from these initiators. Of the many types of possible polymerization l,24 cationic,21 anionic,15 ring-opening,25 and ring-opening strategies (e.g., radica metathesis“), atom transfer radical polymerization (ATRP) is especially attractive because it yields polymers of low polydispersity and is compatible with a variety of functional monomers. Since the initial discovery of ATRP,”28 we and others have 73 adapted this technique for surface-initiated polymerization.”"19’29’30 The recent discovery of transition metal complexes that catalyze ATRP from a surface at room temperature is particularly important because low-temperature polymerization from a substrate can occur with minimal simultaneous polymerization in solution.'9’30'33 This helps to avoid physisorption of unbound polymer chains and allows synthesis of cross-linked polymer films. Additionally, the controlled nature of ATRP affords control over skin thickness by variation of polymerization time. This work demonstrates the versatility of surface-initiated ATRP for forming ultrathin membrane skins on a porous support and examines gas permeation through these membranes. We utilized ATRP to synthesize two kinds of membrane skins: cross- linked poly(ethylene glycol dimethacrylate) (PEGDMA)3 l and poly(2-hydroxyethyl methacrylate) (PHEMA).19 The synthesis involves covalent attachment of an ATRP initiator32 to a modified porous alumina support followed by room-temperature polymerization with a suitable monomer.. Cross-linked polymer membranes such as PEGDMA are potentially attractive for gas separations because they should be able to function in high levels of plasticizing and condensable vapors that ofien degrade membrane performance.”38 Recent methods for forming cross-linked membranes include UV-irradiation of benzophenone-containing 7.4 . 3 0 and chemlcal polymers,34’3"r”39 heating of polyimides that contain diacetylene groups, cross-linking of polyimides with diamino compounds.41 Koros and coworkers demonstrated that chemical cross-linking of carboxylic acid-containing polyimides with ethylene glycol greatly increases the C02 plasticization pressure and also increases COz/CH4 selectivity.10 Preparation of cross-linked membranes can also occur by casting 74 a solution containing cross-linkable monomer and subsequently polymerizing the film. Although this method does not result in ultrathin skins, Hirayama and co—workers showed that cross-linked polymer films containing polyethylene oxide chains have a COz/Nz selectivity of 65.42’43 This work demonstrates that ATRP from a surface allows controlled synthesis of ultrathin, cross-linked and derivatizable membrane skins. Gas-permeation studies with PEGDMA films grown on porous alumina supports show that these membranes are free of defects and have a COz/CH4 selectivity of ~20. In comparison, PHEMA brushes show selectivity values typical of Knudsen diffusion. One advantage of the PHEMA membranes, however, is that they can be readily derivatized with a variety of functional groups. Esterification of PHEMA with pentadecafluorooctanoyl chloride increases the CO; permeability of these membranes, but still yields a COz/CH4 selectivity of only 8. Future work aims at exploiting the versatility of ATRP for creating membranes for specialty applications. 3.2 Experimental 3.2.1 Chemicals and Solutions Poly(allylamine hydrochloride) (PAH) (Mw= 70,000), sodium poly(styrenesulfonate) (PSS) (MW: 70,000), 3-mercaptopropionic acid (MPA), pentadecafluorooctanoyl chloride (97%), pyridine, dimethylforrnamide (DMF, anhydrous, 99.8%), tetrahydrofuran (THF, anhydrous, inhibitor free, 99.8 %), methanol (anhydrous, 99.8%), 2-bromopropionylbromide (2-BPB), CuCl (99.999%), CuBr (99.999%), CuBrz (99%) and 2,2'-bipyridine (bpy, 99%) were used as received from 75 Aldrich. MnClz (Acros) and NaBr (Spectrum) were also used as received. Triethylamine (Spectrum, 98%) was vacuum distilled over CaHz. 2-Hydroxyethyl methacrylate (HEMA, Aldrich, 98%, inhibited with 300 ppm hydroquinone monomethyl ether (MEHQ)) and ethylene glycol dimethacrylate (EGDMA, Aldrich, 98%, inhibited with 100 ppm MEHQ) were purified by passing them through a column of activated basic alumina (Spectrum). Deionized water (Milli-Q, 18.2 M0 cm) was used for preparation of solutions and rinsing. The disulfide initiator, (BrC(CH3)2COO(CH2)118)2, was synthesized according to a literature procedure.16 AnodiscTM porous alumina membranes (Fisher) with 0.02 um-diameter surface pores were used as supports for membrane formation. Gold slides (200 nm of sputtered Au on 20 nm Cr on a Si (100) wafer) were used as substrates for ellipsometry and Fourier transform infi'ared (FTIR) external reflection spectroscopy. 3.2.2 Polymerization of EGDMA The initial step in the polymerization procedure is the attachment of initiating groups to the substrate. In some cases, we first deposited a multilayer polyelectrolyte film on the substrate and subsequently anchored initiators to these modified surfaces. The deposition of polyelectrolytes occurred as follows. Au-coated wafers were cleaned in a UV/ozone chamber (Boekel UV-Clean model 135500) for 15 minutes and subsequently immersed in a 1 mM ethanolic solution of MPA for 30 minutes, rinsed with ethanol and deionized H20, and dried with N2. This procedure yields a carboxylic acid- terminated surface. Substrates were then immersed in a polycation solution (0.02 M PAH, 0.5 M NaBr, pH 2.3) for 5 minutes and rinsed with deionized water. (Molarities of polymers are given with respect to the repeating unit.) Subsequent immersion in a 76 polyanion solution (0.02 M PSS, 0.5 M MnCl2, pH 2.1) for 2 minutes and rinsing with deionized water yielded a polyelectrolyte bilayer on the surface. A second layer of PAH was then adsorbed on top of the PAH/PSS layer to provide amine functional groups for initiator anchoring. Initiator was attached to PAH/PSS films via reaction with 2-BPB in the presence of triethylamine.32 The gold slide was first immersed in a solution of triethylamine (0.242 g in 10 mL THF), and then the initiator solution (0.432 g of 2-BPB in 10 mL THF) was added drop-wise while stirring. Because the reaction is exothermic, both solutions were cooled to 0 °C prior to the reaction. The reaction was stopped after 2 minutes by transferring the slide to a THF solution. Initiator attachment was performed in a glove box because the acid bromide is moisture sensitive. Further rinsing with ethyl acetate, ethanol, and deionized water followed by drying with N2 was done outside the glove box. Polymerization of EGDMA occurred by immersing the initiator-modified surfaces in a solution containing EGDMA (monomer), DMF, deionized H20 and the Cu catalyst system.31 In this procedure, the monomer mixture, 42 mL of solution containing EGDMA, H20 and DMF (323:8, vzvzv) was first degassed in a three necked flask by three freeze-pump-thaw cycles. Then, CuCl (180 mg, 1.8 mmol), CuBr2 (120 mg, 0.54 mmol), and bpy (731 mg, 4.68 mmol) were quickly added to the degassed mixture under a nitrogen atmosphere. This mixture was immediately degassed using another two freeze- pump-thaw cycles and then warmed to room temperature with continuous stirring until the solution became a homogeneous dark brown color. The sealed vessel containing the monomer/catalyst solution was next transferred into a glove bag that was subsequently 77 purged with nitrogen for at least an hour. The polymerization solution was then transferred to vials containing polyelectrolyte-coated gold wafers modified with initiators, and polymerizations were carried out in the glove bag at room temperature for different times. After polymerization, substrates were removed from the vessels, rinsed with DMF, sonicated (1 minute) in DMF, rinsed with THF followed by ethanol, and dried under a flow of N2. To grow PEGDMA on porous alumina, one bilayer of PSS/PAH was deposited directly on UV/ozone-cleaned alumina, and the initiator attachment and polymerization occurred as described above. Polyelectrolyte depositions were limited to the filtrate side of the alumina membrane by using a holder, and the initiator attachment and polymerization were done without the holder. 3.2.3 Polymerization of HEMA and Subsequent Derivatization For the polymerization of HEMA from gold wafers, substrates were first cleaned in a UV/ozone chamber for 15 minutes, immersed in an ethanol/water (50:50, v:v) solution for 10 minutes, rinsed with water, and dried with nitrogen. These Au-coated supports were then immersed in a 1 mM ethanolic solution of the disulfide initiator, (BrC(CH3)2COO(CH2)11S)2, for 12 hours to form a monolayer of initiator. After monolayer formation, the substrates were rinsed with ethanol and dried with N2. Polymerization of HEMA occurred by immersion of the initiator monolayer- coated substrates in a methanolic solution containing HEMA and the Cu catalyst system. To prepare this solution, 42 mL of HEMA and methanol (1:1, v:v) were first degassed in a three-necked flask by three freeze-pump-thaw cycles. Then, 552 mg (3.84 mmol) of CuBr, 86 mg (0.39 mmol) of CuBr2, and 1329 mg (8.52 mmol) of bpy were quickly 78 added to the HEMA/methanol while flowing N2 over the solution. This mixture was immediately subjected to another two freeze-pump-thaw cycles and subsequently stirred until a homogeneous dark brown solution formed. The sealed vessel containing the polymerization solution was then transferred to a glove bag, which was purged with N2 for ~l hour. The polymerization solution was finally transferred into vessels containing substrates modified with initiators, and polymerizations were carried out for different times. Afler polymerization, substrates were removed from the vessels, rinsed with methanol, sonicated (1 minute) in DMF, rinsed with THF followed by ethanol, and dried under a flow of N2. For the polymerization of HEMA from porous alumina, substrates were first coated with gold, and the initiator was attached as a self-assembled monolayer as described above. Prior to gold coating, substrates were immersed in boiling methanol for 10 minutes and subsequently cleaned in a UV/ozone chamber for 10 minutes. Substrates were then sputter-coated (filtrate side only) with 5 nm of gold and again UV/ozone cleaned. Initiator anchoring was done, as described above, by immersion in an ethanolic solution of disulfide initiator (this immersion occurred in an air-tight vessel that was initially purged with N2 gas). After monolayer formation, PHEMA was polymerized from the initiator surface as described above. To derivatize PHEMA coatings, films were immersed in 7 mL of anhydrous DMF containing pentadecafluorooctanoyl chloride (0.08 M) and pyridine (0.1 M). After 15 minutes, films were removed from the solution, rinsed with DMF followed by ethanol, and dried with a flow of nitrogen. The fluorination was monitored by FTIR spectroscopy 79 of films on gold wafers and alumina membranes. PHEMA membranes were fluorinated after initial gas transport measurements. 3.2.4 Film Characterization on Gold Wafers Ellipsometric thickness measurements were obtained using a rotating analyzer ellipsometer (model M-44; J .A. Woollam), assuming a film refractive index of 1.5. For each polymer film, thicknesses were measured at three different spots and averaged. At least three samples of each film were examined. External reflection FTIR spectroscopy was performed with a Nicolet Magna 560 FTIR using a Pike grazing angle (80°) accessory. 3.2.5 Film Characterization on Alumina Supports Film growth was monitored by transmission FT IR spectroscopy (Mattson Instruments, Infinity Gold) and F ield-Emission Scanning Electron Microscopy (FESEM, Hitachi S-470011, acceleration voltage of 15 kV). Membranes were coated with 5 nm of gold for imaging purposes. In the case of cross-sectional images, membranes were freeze-fractured under liquid N2 prior to sputter coating with gold. 3.2.6 Gas-Permeation Studies Gas-penneation studies were performed using a permeation cell with a pressure relief valve, and permeate flux was measured as a fimction of inlet pressure (5-45 psig) using a soap-bubble flow meter (Figure 3.1). 02, N2, H2, He, CH4, and CO2 were used for permeation studies, and measurements were performed for each gas separately in the above order. After examining all gases, O2 permeability was remeasured to check the 80 h c.1- '- cml.- .11 iifi!i!i!i!i£i£i!i!ifi£i§!fii£i§l1ifi!fihi ' ' ' ' 'WIHiITifili'jli’flifififililifif flimfliJ ' ,' $232555: L— 93 <1) E E 2 u. E E» D 5: E, 032 1% En , 2‘5 r em E 3 ‘1 t 002 He N2 Gas Controller 81 Apparatus for gas-permeation measurements. Figure 3.1 stability of the membrane, and the O2 flux changed by <10%. For each gas, the permeation cell was purged several times with the gas of interest over a 20-minute period to obtain a stable flux value. Gas permeation studies were done for each polymer (PEGDMA or PHEMA) at several different thicknesses, and for each thickness, three different membranes were tested. The area of the membrane exposed to gas was 2.0 cm2. The selectivity of one gas over another was obtained from the ratio of the respective permeability coefficients at 45 psig. 3.3 Results and Discussion 3.3.1 Synthesis and Characterization of PEGDMA films The first step in growing polymer films from a substrate is attachment of an initiator to the surface. We chose to attach initiators to adsorbed multilayer polyelectrolyte fihns because electrostatic adsorption provides a convenient way to introduce functional groups on a surface. Deposition of PAH/PSS/PAH films results in a surface rich in amine groups, and attachment of initiators to this surface via amide linkages occurs easily (Figure 3.2). Initially, we grew films on gold-coated Si wafers because this substrate facilitates film characterization by ellipsometry and reflectance FTIR spectroscopy. The ellipsometric thickness of PAH/PSS/PAH films on MPA-coated gold was 4.8 i 0.4 nm, and the reflectance FT IR spectra (Spectrum a, Figure 3.3) of these films had strong sulfonate peaks at 1219 and 1177 cm'1 as well as a number of peaks due to aromatic and —NH3+ modes.“ After reaction of the film with the acid bromide initiator, the reflectance FT IR spectrum looked similar to that of PAH/PSS/PAH, but there was a small increase in the peak intensity in the amide region (1650-1560 cm'l, Spectrum b, Figure 3.3), 82 080:3 :0 4:3 005.008 000080 050500.038 0 8on gnaw .Ho :owfitofibom wagon? 8003? 030835 "Nae. 953% “ES. 2 0050030 .2055 EE Ll 00:58 025.0020 “ES. “=20 Ea; 2.280 0 I + + v L - >3 >08 58 h (anteater/timea?» o. o a...» Efuovz 5118:0005 0:37.330 05982028 Cass EE 03.282028 00053.2 83 0.05 Absorbance .<_°.>__,/\ J WHO M KNOW I l l l l 3500 3000 2500 2000 1500 1000 Wavenumbers (cm'1) Figure 3.3: External reflection FTIR spectra of a PAH/PSS/PAH film on a MPA-modified Au surface before (a) and after reaction with an acid bromide initiator (b). Spectrum (c) is from a PEGDMA film grown for 20 hours from PAH/PSS/PAH modified with initiator. 84 indicating initiator attachment. Film thickness increased only slightly ( 023000—00 0w000>0 05 080 .00805808 8000.030 00:: 0o.“ 000030—00 000B Ami; m0 00v 00333020? 00805808 8000.006 008: 08 08065000 325008000 («0 0w000>0 08 000 0020> 00003 A 850000 was: 0003200 0003 08068000 0923008005“0 .00w08m Simmm 88.0 00008300 0003 000005—0500 8380 0 ~88 \80 3.09 M80 2-3 x“ n 005m _ mfiom :3 036; N02 fion”: 0.30.0 0600.0 No3.— modfimg ow om 000m mi: v.oH~.~ 000m foam; mdfiwd Nofimw Nomad fiofiam om om 0.0.3.0 0.0030 Noam; Sham N0 w 3”? @303 73 30. em N4 £000 £0000 ~23 60 .00 N: 0: N2 N0 .005 Awfiwmv 0 z: 00 0 A8080 v m :22 0o 0 0 _ 008000 00 000 :0 0 000MH030 8200: u 0. .0 _ m m 00 . b U 5.0 00 _ _ U _.m 008.com 00805808 0m0u0>0 05 0:0 080.5808 80:00.00 00:5 00.: 00003000 0:03 3me 3 00V 00_:>000_0m a 00:05808 8000020 00:5 00.: 08205000 053008000 .00 0w000>0 05 0:0 0020> 00:05 A 500:3 800 0000—0200 0003 08065000 002500809” 0 .m0w08_ EMmmm 80.0 00008000 0003 000005050 0 ad 0.0 ad 700 mduuwxz VHMN gum— m.ofl¢.0 mdflNS 00 v ad ed m6 700 ONfi: WHMN MH©~ 700 700. mm N 0.6 0.6 ad 3% NflN~ mflaN mflom Nfla 70w w~ g £000 £0000 N200 N00 £0 a: 0: N2 N0 .005 0320 30¢on DEC. _ - 033000—0m $00809 008005000 323008000 000 00003000 87m :0008 008.com 00:05:08 0.0/Ema 0000030008: :0.“ 000300200 #002 0:0 000005—030 EE 00008000 08003000 055008.800 000 "Nd 030,—. 100 on permeability. The permeability of C02 obtained in our study is similar to the reported C02 permeability for PEMA, but the observed low selectivity of PHEMA compared with the selectivity reported for PEMA suggests that the hydroxyl groups in PHEMA dramatically alter polymer packing. To improve the COz/CH4 or 02/N2 selectivity of PHEMA membranes, we explored derivatization of PHEMA with fluorinated compounds. One attractive feature of PHEMA is that its hydroxyl groups can be easily derivatized with various acid E chlorides or carbonyldiimidazole to introduce different functional groups.19 We reacted ’ PHEMA with pentadecafluorooctanoyl chloride in the presence of a base to obtain r PHEMA with perfluorinated side chains (scheme 3.3). Disappearance of the alcohol % peak (3500-3300 cm") in the reflectance FTIR spectrum of PHEMA (Spectrum b, Figure 3.9) indicates conversion of the hydroxyl groups to fluorinated esters. The appearance of a fluorinated ester peak at 1800 cm" and CFx peaks at 1250 cm”1 also confirm the esterification of PHEMA. Based on the density of HEMA and poly(1,l ’- dihydroperfluorooctyl methacrylate, we would expect to see a more than 100 % increase in thickness upon fluorination. Nevertheless, the ellipsometric thicknesses of PHEMA films increased only by ~70 % after reaction with pentadecafluorooctanoyl chloride, suggesting <100 percent derivatization of hydroxyl groups or very dense films. However, the disappearance of the alcohol peak in the IR spectrum of PHEMA points to virtually quantitative derivatization. In addition to examining PHEMA films on gold wafers, we also characterized the fluorinated films on alumina with transmission FTIR (Figure 3.10). Appearance of a carbonyl peak around 1800 cm‘1 suggests the incorporation of perfluorinated groups. The relatively small increase in film thickness 101 S? W 1*“ Au ——S—(CH2)11—O—C—(I3 CHz—C Br CH3 o:\/ n PHEMA O— CH2CH2 OH S0 45A \ (I? i kg CF3{CF2%C——Cl 5: DMFR T fir Fl uon' nated-PHEMA 9 0 9% AU —S—(CH2)11—O—C—IC CHz—C Br CH3 0:< n O— CHzCHzO‘nzf Csz’eCF3 0 Scheme 3.3: Derivatization of PHEMA with perfluorooctanoyl chloride. 102 0. 1 8 C N g < (a) M N I 1 1 l l 1 3500 3000 2500 2000 1500 1000 Figure 3.9: Reflectance FTIR spectra of a PHEMA film grown from (BrC(CH3)2COO(CH2)1 IS)2 on gold (a) before and (b) after Wavenumbers (cm'l) reaction with pentadecafluorooctanoyl chloride. 103 (b) Absorbance (a) l L l l l 4000 3 500 3000 2500 2000 l 500 1000 Wavenumbers (cm") Figure 3.10: Transmission FTIR spectra of a PHEMA film grown from (BrC(CH3)2COO(CH2)1 18); on gold-coated porous alumina (a) before and (b) afier reaction with pentadecafluorooctanoyl chloride. 104 . ' Q.__' J‘ after fluorination may indication a high film density (low free volume) afier derivatization. This is consistent with the fact that fluorinated PHEMA is less permeable than poly(1,1’—dihydroperfluorooctyl methacrylate (vide infra), and the brush-like structure of these films might account for this high density. . Gas-permeation studies with PHEMA were repeated after fluorination, and Figure 3.11 shows the fluxes of several gases as a function of transmembrane pressure drop. 3‘ After fluorination, the fluxes of various gases decrease in the order C02>He>H2>02>N2, l CH4 and are no longer dependent solely on the molar masses of the gases. Because ’ fluorination enhances C02 flux relative to other gases, C02/CH4 selectivity increases ~10-fold compared to non-fluorinated films. Similar to PEGDMA, fluorinated PHEMA 1% also shows an increase in C02/CH4 selectivity with increasing film thickness. The highest selectivity was obtained for a lOO-nm thick fluorinated film (Table 3.3). At fihn thicknesses higher than 100 nm, flux values for some gases were lower than the detection limits of our flow meter. The increase in C02/CH4 selectivity and C02 permeability after fluorination probably occurs due to an increase in C02 solubility in the polymer matrix. Compared with the other gases we tested, C02 has a high polarizability and quadrupole moment48 that should allow it to interact with the polar fluorinated side chains (the C-F dipole moment is 1.39 D”). Thus C02 solubility is higher after fluorination compared to other gases hence we observe higher C02/CH4 selectivity. Arnold and coworkers reported the gas permeation properties of membranes prepared from poly(1,1’-dihydroperfluorooctyl acrylate) (PF 0A) and poly(1,1’- dihydroperfluoro methacrylate) (PFOMA).60 These membranes are similar to fluorinated 105 6 0 CD2 0 D He 5 " a H2 o 0 02 o a," I N S 4' c; ‘73 Tm V 4 0 D i F; El E’l 3 - o n D , E g 8 o D ‘ ‘2? 2 L D A 3 ° C. + A u. A O . D A <> 0 1 - A O D ‘ 0 <7 0 ‘3 3 o v V V e 6 V V O OH ‘ ‘ ‘ O 10 20 30 40 50 Pressure (psig) Figure 3.11: Fluxes of different gases through a fluorinated- PHEMA membrane (100 nm thick) as a function of transmembrane pressure drop. The outlet pressure was 1 atm and the measurements were performed at room temperature. 106 008000: 0:0 m00:0> 0m00>0 050:0 a00:0:0808 800.80 00:5 :8 00:23:00 003 80:: 00 :00 000300009 00:0:0808 800.80 00:5 :8 08000000 30500800 00 00890 05 0:0 m00:0> 000:: .: :000000 0:00 0808800 00>» 08000000 35000800.. .0008: $5.0m": 800 00:08:00 003 0000:0050.w 0000.0 00000 0.000.: 00mm 00000 NON: 00m: 0000. m :00 00: :.:0\r.0 0000.0 :.00w.: 00: 0 0000.0 ~00: m0:~ N.00m.0 0.0000 00 0000.0 0000.0 :.000.: m0m~ N00 0.00: ~00: :00 :00 00 ~Z\~OU £0000 NZ\~O Nov 0::0 mm: 0:: N7: N0 aA85 00508:. o3:>::00:0m @0803 0080000000 30500800 000 0808800 8:“: 005808 0.0/man: 00:08:03: :8 0003:0000 :000: 0:0 08000000 3050080: 000 00000585 8:: 00:08:00 "fin 0:00P 107 PHEMA and have similar C02/CH4 and 02/N 2 selectivities. However, the permeability coefficients of C02 in PF 0A and PF OMA are, respectively,18-fold and 5-fold higher than that of fluorinated-PHEMA.. The extra ester group in fluorinated PHEMA probably reduces the chain mobility and lowers the permeability, and as mentioned above, derivatized films may have a high density.. Other studies also showed that fluorinated side chains greatly enhance C02 permeability, and hence C02/CH4 selectivity.6"62 3.4 Conclusions Surface ATRP provides a convenient way of synthesizing ultrathin, cross-linked films and linear polymer brushes on porous supports. FTIR and F ESEM confirm film formation on these surfaces. Gas permeation studies with these films indicate that PEGDMA has a C02/CH4 selectivity of ~20 and a C02 permeability coefficient of 20 Barrers. Unlike cross-linked PEGDMA, linear PHEMA films show only Knudsen- diffusion based selectivity. However, esterification of the hydroxyl groups of PHEMA with pentadecafluorooctanoyl chloride increases the C02/CH4 selectivity to ~8 and the C02 permeability coefficient to ~20 Barrers. The derivatizability of PHEMA may make it a suitable candidate for specialty separations. 108 .305 (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) (17) (18) References and Notes Ferjani, E.; Lajimi, R. H.; Deratani, A.; Roudesli, M. S. Desalination 2002, 146, 325-330. Dutta, B. K.; Sikdar, S. K. Environ. Sci. & Tech. 1999, 33, 1709-1716. Li, X.-G.; Huang, M.-R. J. Appl. Polym. Sci. 1997, 66, 2139-2147. Karnada, K.; Kamo, J .; Motonaga, A.; Iwasaki, T.; Hosokawa, H. Polym. J. 1994, 26, 833-839. Marek, M., Jr.; Brynda, E.; Houska, M.; Schauer, J .; Hynek, V.; Sipek, M. Polymer 1996, 37, 2577-2579. Baker, R. W. Ind. Eng. Chem. Res. 2002, 41, 1393-1411. Pinnau, 1.; Freeman, B. D. In Polymer Membranes for Gas and Vapor Separation; Pinnau, 1., Freeman, B. D., Eds.; American Chemical Society: Washington, DC, 2000; Vol. 744, pp 1-22. Liu, C.; Martin, C. R. Nature 1991, 352.50-52. Petersen, R. J. J. Membr. Sci. 1993, 83, 81-150. Staudt-Bickel, C.; Koros, W. J. J. Membr. Sci. 1999, 155, 145-154. Hayakawa, Y.; Terasawa, N.; Hayashi, E.; Abe, T. J. Appl. Polym. Sci. 1996, 62, 951-954. Yanagishita, H.; Kitamoto, D.; Haraya, K.; Nakane, T.; Okada, T.; Matsuda, H.; Idemoto, Y.; Koura, N. J. Membr. Sci. 2001, 188, 165-172. Pinnau, 1.; Koros, W. J. Ind. Eng. Chem. Res. 1991, 30, 1837-1840. Jordan, R.; Uhnan, A. J. Am. Chem. Soc. 1998, 120, 243-247. Jordan, R.; Ulman, A.; Kang, J. F.; Rafailovich, M. H.; Sokolov, J. J. Am. Chem. Soc. 1999, 121, 1016-1022. Shah, R. R.; Merreceyes, D.; Husemann, M.; Rees, 1.; Abbott, N. L.; Hawker, C. J .; Hedrick, J. L. Macromolecules 2000, 33, 597-605. Huang, X.; Wirth, M. J. Macromolecules 1999, 32, 1694-1696. Huang, X.; Doneski, L. J .; Wirth, M. J. Chemtech 1998, 28, 19-25. 109 (19) (20) (21) (22) (23) (24) (25) (26) (27) (28) (29) (30) (31) (32) (33) (34) (35) Huang, W.; Kim, J .-B.; Bruening, M. L.; Baker, G. L. Macromolecules 2002, 35, 1 175-1 179. Ejaz, M.; Yamamoto, S.; 0hno, K.; Tsujii, Y.; Fukuda, T. Macromolecules 1998, 31, 5934-5936. Zhao, B.; Brittain, W. J. J. Am. Chem. Soc. 1999, 121, 3557-3558. Huang, W.; Skanth, 0.; Baker, 0. L.; Bruening, M. L. Langmuir 2001, 17, 1731- 1736. Zhao, B.; Brittain, W. J .; Zhou, W.; Cheng, 8. Z. D. J. Am. Chem. Soc. 2000, 122, 2407-2408. Prucker, 0.; Riihe, J. Macromolecules 1998, 31 , 602-613. Husemann, M.; Mecerreyes, D.; Hawker, C. J .; Hedrick, J. L.; Shah, R.; Abbott, N. L. Angew. Chem. Int. Ed. 1999, 38, 647-649. Jeon, N. L.; Choi, I. S.; Whitesides, G. M.; Kim, N. Y.; Laibinis, P. E.; Harada, Y.; Finnie, K. R.; Girolarni, G. S.; Nuzzo, R. G. Appl. Phy. Lett. 1999, 75, 4201- 4203. Wang, J .-S.; Matyjaszewski, K. Macromolecules 1995, 28, 7901-7910. Kato, M.; Kamigaito, M.; Sawamoto, M.; Higashimura, T. Macromolecules 1995, 28, 1721-1723. Huang, X.; Wirth, M. J. Anal. Chem. 1997, 69, 4577-4580. Matyjaszewski, K.; Miller, P. J .; Shukla, N.; Immaraporn, B.; Gelrnan, A.; Luokala, B. B.; Siclovan, T. M.; Kickelbick, G.; Vallant, T.; Hoffmann, H.; Pakula, T. Macromolecules 1999, 32, 8716-8724. Huang, W.; Baker, G. L.; Bruening, M. L. Angew. Chem. Int. Ed. 2001, 40, 1510- 1512. Kim, J.-B.; Bruening, M. L.; Baker, G. L. .1. Am. Chem. Soc. 2000, 122, 7616- 7617. Jones, D. M.; Huck, W. T. S. Adv. Mater. 2001, 13, 1256-1259. Kita, H.; Inada, T.; Tanaka, K.; Okamoto, K. J. Membr. Sci. 1994, 87, 139-147. Bos, A.; Pont, I. G. M.; Wessling, M.; Strathmann, H. J. Polym. Sci., Part B: Polym. Phys. 1998, 36, 1547-1556. 110 (36) (37) (38) (39) (40) (41) (42) (43) (44) (45) (46) (47) (48) (49) (50) (51) (52) (53) (54) McCaig, M. 8.; Paul, D. R. Polymer 1999, 41 , 629-637. Rezac, M. E.; Schoberl, B. J. Membr. Sci. 1999, 156, 211-222. Bos, A.; Punt, I. G. M.; Wessling, M.; Strathmann, H. Sep. Purif. Tech. 1998, 14, 27-39. Wright, C. T.; Paul, D. R. J. Membr. Sci. 1997, 129, 47-53. Rezac, M. E.; Sorensen, E. T.; Beckham, H. W. J. Membr. Sci. 1997, 136, 249- 259. Liu, Y.; Wang, R.; Chung, T. S. J. Membr. Sci. 2001, 189, 231-239. Hirayama, Y.; Tanihara, N.; Kusuki, Y.; Kase, Y.; Haraya, K.; Okamoto, K. i. J. Membr. Sci. 1999, 163, 373-381. Hirayama, Y.; Kase, Y.; Tanihara, N.; Sumiyarna, Y.; Kusuki, Y.; Haraya, K. J. Membr. Sci. 1999, 160, 87-99. Stair, J. L.; Harris, J. J.; Bruening, M. L. Chem. Mater. 2001, 13, 2641-2648. Parks, 0. A. Chem. Rev. 1965, 65, 177-198. Bluhm, E. A.; Bauer, E.; Chamberlin, R. M.; Abney, K. D.; Young, J. S.; Jarvinen, G. D. Langmuir 1999, 15, 8668-8672. Actually, PHEMA is partially branched due to intermolecular transesterification and in some cases is lightly cross-linked. Ghosal, K.; Freeman, B. D. Polym. Adv. Tech. 1994, 5, 673-697. Stern, S. A. J. Membr. Sci. 1994, 94, 1-65. Koros, W. J. J. Polym. Sci., Part B: Polym. Phys. 1985, 23, 1611-1628. Ghosal, K.; Chem, R. T.; Freeman, B. D.; Savariar, R. J. Polym. Sci, Part B: Polym. Phys. 1995, 33, 657-666. Hoover, J. M.; Smith, S. D.; DeSimone, J. M.; Ward, T. C.; McGrath, J. E. Polymer Preprints 1987, 28, 390-392. Robeson, L. M. J. Membr. Sci. 1991, 62, 165-185. Chiou, J. 8.; Paul, D. R. J. Membr. Sci. 1989, 45, 167-189. 111 (55) (56) (57) (58) (59) (60) (61) (62) Raymond, P. C.; Paul, D. R. J. Polym. Sci, Part B: Polym. Phys. 1990, 28, 2079- 2102. Wright, C. T.; Paul, D. R. Polymer 1997, 38, 1871-1878. Chiou, J. S.; Barlow, J. W.; Paul, D. R. J. Appl. Polym. Sci. 1985, 30, 1173-1186. Andrews, R. J .; Grulke, E. A. In Polymer handbook; Brandrup, J ., Immergut, E. H., Grulke, E. A., Eds; John Wiley and Sons: New York, 1999; pp VI 193- VI 277. Yampolskii, Y. P.; Alentiev, A. Y.; Shishatskii, S. M.; Shantarovich, V. P.; Freeman, B. D.; Bondar, V. I. Polymer Preprints 1998, 39, 884-885. Arnold, M. E.; Nagai, K.; Freeman, B. D.; Spontak, R. J .; Betts, D. E.; DeSimone, J. M.; Pinnau, I. Macromolecules 2001, 34, 5611-5619. Nagale, M.; Kim, B. Y.; Bruening, M. L. J. Am. Chem. Soc. 2000, 2000, 11670- 1 1678. Xiao, K. P.; Ham's, J. J .; Park, A.; Martin, C. M.; Pradeep, V.; Bruening, M. L. Langmuir 2001, 1 7, 8236-8241. 112 VE‘: '. '.. V‘) v 5'- CHAPTER 4 Ion Transport through Grafted Poly(2-hydroxyethyl methacrylate) Membranes and their Derivatives 4.1 Introduction Chapter 3 demonstrated the fabrication of ultrathin skins on porous supports using surface-initiated atom transfer radical polymerization. This chapter investigates the ion permeability of poly(2-hydroxyethyl methacrylate) (PHEMA) membranes that were also prepared by ATRP from a surface. Although such membranes did not prove highly selective in gas separations, they do allow selective transport of monovalent ions. Derivatization of PHEMA with a cross-linking agent, succinyl chloride, results in C1’ /SO42' selectivities as high as 300. Such selectivities may be important in nanofiltration (NF), which has become an important area of research in membrane-based separations. This technique is widely used in applications such as water softening and purification, removal of heavy metals from water streams, waste-water reclamation, and separation of organic solutes."2 NF is sometimes preferable to reverse osmosis because it occurs at lower pressures and hence, has lower energy costs. The separation characteristics in NF fall between reverse osmosis and ultrafiltration, and separation is based on sieving and electrostatic effects?”4 Typical NF membranes are synthesized from polymers such as polysufones, polyamides, modified aromatic polyamides, and derivatives of polyvinyl alcohols.S In 1965, Baddour and coworkers demonstrated that hydrogels made from PHEMA are capable of desalinating brackish or sea water.6’7 They showed that membranes synthesized by copolymerization of HEMA and ethylene glycol 113 dimethacrylate have NaCl rejections up to 87.6%.(”7 Haldon and Lee examined the permeability of PHEMA membranes that were prepared by copolymerization with different cross-linkers (ethylene glycol dimethacrylate, trimethylolpropane trimethacrylate (TPT) or pentaerythritol tetramethacrylate) and showed that the water- permeability of the membrane depends on the cross-linking density.8 A similar study by J adwin and coworkers found that water permeability through cross-linked PHEMA membranes decreases from 6X104’ m3-m/m2-day to 6><10’9 m3-m/m2-day, and salt rejection increases from 78% to 94% as the amount of cross-linking with TPT increases from 0 to 11 mole percent.7 However, such membranes were prepared by solution casting followed by photoirradiation, and thus, the thicknesses of these materials were relatively high (100-500 pm). This, of course,_results in unacceptably low fluxes. The main objective of this work was to determine whether ultrathin, PHEMA skins prepared by ATRP from a surface can successfully separate different ions. Ion- transport studies with composite PHEMA membranes showed moderate selectivities (Cl' /S042'selectivity of 15, K+/Mg2+ selectivity of 47 and Cl'/Fe(CN)63' selectivity of 164), but Cl' fluxes in diffusion dialysis were quite low (15% of that through the bare alumina support). (Higher selectivities were obtained with even thicker films, but flux was even lower.) The relatively low fluxes occurred in spite of the fact that fihn thickness was less than 100 nm, suggesting that films are relatively dense. One advantage of PHEMA is that it can be easily tailored for specific applications through derivatization of its hydroxyl groups.9 In chapter 3, I showed that derivatization of PHEMA with pentadecafluorooctanoyl chloride enhanced gas-transport selectivities. This chapter shows that reaction of PHEMA with succinyl chloride dramatically increases ion- 114 transport selectivities, while still allowing reasonable flux values with 28 nm-thick films (Cl‘ flux was 50% of that through bare alumina). 4.2 Experimental 4.2.1 Chemicals and Solutions Poly(allylamine hydrochloride) (PAH) (MW: 70,000), sodium poly(styrenesulfonate) (PSS) (MW: 70,000), 3-mercaptopropionic acid (MPA), pyridine, dimethylformamide (DMF, anhydrous, 99.8%), tetrahydrofuran (THF, anhydrous, inhibitor free, 99.8%), ethyl acetate, ethanol, 2-bromopropionylbromide (2-BPB), CuCl (99.999%), CuBr (99.999%), CuBr2 (99%), 2,2'-bipyridine (bpy, 99%), 1,1'- carbonyldiimidazole (CD1, 98%), 2-hydroxymethyl-18-crown-6 (90%) and succinyl chloride (90%) were used as received from Aldrich. MnCl2 (Acros) and NaBr (Spectrum) were also used as received. Triethylamine (Spectrum, 98%) was vacuum distilled over CaH2. 2-Hydroxyethyl methacrylate (HEMA, Aldrich, 98%, inhibited with 300 ppm hydroquinone monomethyl ether (MEHQ)) was purified by passing it through a column of activated basic alumina (Spectrum). Deionized water (Milli-Q, 18.2 M0 cm) was used for preparation of solutions and rinsing. AnodiscTM porous alumina membranes (Fisher) with 0.02 um-diameter surface pores were used as supports for membrane formation. Gold-coated slides (200 nm of sputtered Au on 20 nm Cr on a Si (100) wafer) were used as substrates for ellipsometry and Fourier transform infrared (FTIR) external reflection spectroscopy. 115 4.2.2 Polymerization of HEMA from Gold-coated wafers and Porous Alumina For p olymerization o f H EMA from gold w afers, 2 -BPB w as immobilized o n a multilayer polyelectolyte film (PAH/PSS/PAH) using the procedure described in Chapter 3. In the case of porous alumina, the initiator was attached to a PSS/PAH bilayer. Polymerization of HEMA occurred by immersion of the initiator-coated substrates in an aqueous solution containing HEMA and a Cu catalyst system.9 To prepare the catalyst solution, 42 mL of HEMA and deionized water (1:1, v:v) were first degassed in a three- necked flask by three freeze-pump-thaw cycles. Then, 55 mg (0.55 mmol) of CuCl, 36 mg (0.16 mmol) of CuBr2, and 244 mg (1.56 mmol) of bpy were quickly added to the HEMA/water solution under a flow of nitrogen. The mixture was immediately subjected to another two freeze-pump-thaw cycles and subsequently stirred until a homogeneous dark brown solution formed. The sealed vessel containing the polymerization solution was then transferred to a glove bag, which was purged with N2 gas for ~1 hour. The polymerization solution was finally transferred into vessels containing substrates modified with initiators, and polymerizations were carried out for different times. After polymerization, substrates were removed from the vessels, rinsed with deionized water, sonicated (1 minute) in DMF, rinsed with THF followed by ethanol, and dried under a flOW Osz. 4.2.3 Derivatization of PHEMA with Crown ethers To couple crown-ethers to PHEMA, a film-coated substrate was immersed into a 0.2 M solution of CD1 in DMF for 12 h and subsequently rinsed with DMF. CDI- functionalized PHEMA substrates were then immersed in a 10-mL DMF solution containing 2-hydroxymethyl-18-crown-6 (0.1 M) and triethylamine (0.1 M). The 116 reaction was carried out at 70 °C for 4.5 days, after which the substrate was rinsed with DMF, followed by ethanol and dried with a flow of nitrogen. Derivatization was performed after ion-permeability studies with the underivatized PHEMA membrane. 4.2.4 Chemical Cross-linking of PHEMA PHEMA-coated substrates were immersed in 10 mL of DMF containing succinyl chloride (0.1 M) and pyridine (0.1 M). After 10 minutes, substrates were removed from the reaction solution, rinsed with DMF, and dried with N2. Derivatization was performed after ion-permeability studies. 4.2.5 Film Characterization Ellipsometric measurements were obtained with a multiwavelength, rotating analyzer ellipsometer (model M-44; J .A. Woollam) using WVASE32 software at an incident angle of 75°. The refractive index of the films at all wavelengths was assumed to be 1.5. For each substrate, thicknesses were measured at three different spots and averaged. Reflectance FTIR spectroscopy was performed using a Nicolet Magna-IR 560 spectrometer containing a PIKE grazing angle (80°) attachment. The spectra were collected with 256 scans using a MCT detector. 4.2.6 Ion-Transport Studies with PHEMA Membranes Diffusion-dialysis studies with PHEMA membranes were performed using two glass half cells (Figure 2.2) as described in Chapter 2 (section 2.2). The permeate cell contained deionized water (90 mL), and the feed cell contained 0.1 M salt solutions (90 mL) of KCl, K2S04, K3Fe(CN)6 and MgCl2. After dialysis with each salt, the entire apparatus was rinsed well with deionized water and subsequently filled with water for 30 117 minutes to remove any adsorbed ions. Salts were examined in the same order as given above, and after MgCl2 dialysis, a second KCl dialysis was performed to check the integrity of the membrane. (The difference between KCl fluxes in the first and second dialyses was <10%). Conductivity (Orion model 115) of the receiving side was recorded at every 5 minutes for a period of 45 minutes, and conductivity values were converted to concentration using a calibration curve of conductivity vs. concentration for each salt. Fluxes were calculated from the slopes of concentration vs. time plots using equation 2.1. Selectivity of monovalent over divalent ions was calculated from the ratio of respective flux values (equation 2.2). For chemically cross-linked PHEMA membranes, I also examined dialysis with solutions containing 1000 ppm Cl' and 1000 ppm SO42' or 500 ppm Cl' and 2500 ppm 8042' (solutions were prepared with KCl and K2SO4.). Dialysis was carried out for 90 min, and 2 mL samples were withdrawn from both sides at 10-min intervals. These samples were analyzed with an ion chromatograph (Dionex 600) using an ASl4A anion column and an 8 mM Na2CO3/1 mM NaHCO; eluent. Normalized fluxes were calculated from the respective slopes of normalized concentration vs time plots, and the anion- selectivities were determined from the ratio of normalized fluxes. Normalization was performed by dividing by the source-phase concentration. 4.3 Results and Discussion 4.3.] Synthesis and Characterization of PHEMA Membranes Initiator anchoring and polymerization were initially performed on gold—coated wafers to facilitate film characterization. First, 2-BPB was covalently attached to the 118 . ‘ 4 '., It. . . 7v“ polyelectrolytes through amide linkages. As previously discussed (section 3.3.1), we did not observe a prominent change in the reflectance FTIR spectrum of polyelectrolytes after initiator anchoring, but there was a small absorbance increase in the amide region (1650-1560), confirming initiator attachment. Polymerization was carried out by immersing the initiator-modified substrates in HEMA/water solutions containing CuCl/CuBr2/bpy. Unlike gas-permeability studies, the HEMA polymerization was carried out in water, rather than methanol, because aqueous polymerization is a more controlled process that yields a‘relatively linear relationship between film thickness and polymerization time. Several previous studies showed that aqueous conditions also 9‘” The appearance of a strong carbonyl peak accelerate ATRP of hydrophilic monomers. at 1730 cm"1 in the reflectance FTIR spectrum of films after polymerization indicated formation of PHEMA on the surface. 4.3.2 Ion-Transport Studies with Composite PHEMA Membranes Figure 4.1 shows results of diffusion dialysis through a composite membrane containing a 28 nm-thick PHEMA skin. The linear relationship between receiving-phase concentration and time shows that flux is constant and confirms that ion concentrations in the receiving phase were negligible compared to those in the source phase. Table 4.1 gives the cumulative thicknesses of PHEMA fihns, ion-transport selectivities and ion fluxes through PHEMA membranes. As the PHEMA film thickness increases, selectivity increases, presumably because the film more completely covers the substrate. Since PHEMA is neutral, the observed selectivities among ions are probably due to differences in hydration energies and hydrated radii. Both hydration energy and 119 .v ‘ u I" 4 ~,:'_.;; Receiving Phase Concentration (mM) 2.5 O KCI I KZSO4 . ' MgCl2 2.0- A K3E'-'e(CN)6 . o 1.5- I o 1.0- . I I I I 0.5“ . . V V I ' v A ‘ ‘ I . ‘ 0.0? 0 10 20 30 40 Time (min) Figure 4.1: Plot of receiving phase concentration as a function of time in diffusion dialysis when the source phase (0.1 M salt) was separated from the receiving phase (initially deionized water) by a porous alumina substrate capped with 28 nm of PHEMA. 120 :_ no? \ .' I' .«‘-'—n_.'.. :00 505 00:88:00 00>» 80880000 :05 00 02.: -:U 00:00: 05 00 0800 05 003 :0: 8:00:00 .:0 00:00 .m0=:0> 02.: 00000 800 :0: 0:0 0:0:0808 0000 :8 029 3:300:00 .:0 000:0>0 05 00 00:23:00 .00::0> 03:: .:0 02:02:00: .:0 8:00:o 0808080008 00.5 :000: :0 .:0 0:000:00 0:00:00 800080: 80::m: 0:08 08:: .0> 000:: 083000: 8 85880800 .:0 000:0 05 808 0:080:00 0:03 0020:: 00:00. 0:05 8: 0.0 003 00:03 .80.: :0:0:::::\0:30::00:0300 05 0020:: 000005082. :00: :05 000: 3: 00::0> 00:03 0808080008 03: .:0 00000 05 800080: 0030:» 0:80? 0:00 :0 08:0 Emma: 00 080800820 3: 0088000 0:03 0000:0083. a 000mm: 0:. :0000 0:000 00000000 :00000000 00000000 0000.0 :0 00: m :000 2000: 00m : 00.00:N0 :000000 :0000 :00 00 00: :00 000 0000.0 0000.0 :00 000: 0000 00 00 £025: 00:20:00-0 0.000 .002 00:20:»: -000 :0 :50 :58: 25: «0005:0000 85000839: U3350000 00 : xAm~80\m0:080 A:02": 8:0: 0:00:00 08820 30:03 005808 0.0/Ema: 0:009:00 :8 000050000 0:0 0002.: 00000585 80...: 3.0 030.0 121 hydrated radii decrease in the order Mg2+ > SO42' > C1' > K+ (See Table 4.2).l2 Since hydration energy and hydrated radius are higher for Mg2+ than SO42", one would expect to observe higher selectivity values for KVMg2+ than for Cl'/SO42'. This was indeed the case for all three PHEMA thicknesses. The Cl'/Fe(CN)63' selectivity is higher than monovalent/divalent selectivities probably because of the large hydration radius and hydration energy of Fe(CN)63'. Although we were successful in obtaining reasonable selectivities with PHEMA in: membranes, Cl' flux values were relatively low for thicker films. For 28 nm-thick i i PHEMA films, Cl- flux was 80% of that through bare alumina (Cl' flux through bare alumina was 5.2><10’8 moles/cmzs'), but flux decreased dramatically with increasing 9 PHEMA thickness (Table 4.1). We thought that cross-linking of thin PHEMA membranes might allow both high flux and high selectivity. Cross-linking should reduce swelling and may allow thin films to fully cover a substrate. 4.3.3 Chemically Cross-linked PHEMA Membranes and their Ion-Transport Properties We cross-linked PHEMA by derivatizing films with a di-acid chloride that would react with hydroxyl groups in adjacent PHEMA chains. Scheme 4.1 outlines the reaction of grafied PHEMA layer with succinyl chloride. Reflectance F TIR spectra of PHEMA films after derivatization showed the disappearance of the hydroxyl peak and a ~2-fold increase in the intensity of ester carbonyl peak, suggesting quantitative conversion of hydroxyl groups to esters (Spectrum b, Figure 4.2). However, the appearance of a small shoulder at 1819 cm'1 suggested the presence of unreacted acid 122 Table 4.2: Hydrated radii and hydration energies of various ions.‘2 Ion Hydrated Hydration Energy Radius (pm) (kJ/mol) K+ 212 330 CI' 224 365 S047“ 278 h 1035 Mg2+ 299 1945 Fe(CN)63' 396 123 9 9H3 9H3 ° . WWNH—C—Cl: CHz—C CKBF) + (”W O=§ H n O DMF, RT it (H) CH3 (EH3 CH3 WW‘NH--C—C|3 CHz—C CHz—C C|(Br) H 0:< n O \ n-1 0 g0 30 O O O —O o 0 Cl OCHZCHZO— Scheme 4.1: Cross-linking of PHEMA by reaction with succinyl chloride. 124 __ s - ,:-_ A I 0.02 . X 0.8 (c) : x 0.8 (b) ' M (a) Absorbance W‘r l l l l l l 3500 3000 2500 2000 1500 1000 Wavenumbers (cm") Figure 4.2: Reflectance FT IR spectra of a 28-nm thick grafted PHEMA film on a gold-coated wafer before (a) and after reaction with succinyl chloride (b) and subsequent exposure to water (c). 125 chlorides, implying less than 100% cross-linking. Disappearance of the 1819 cm'l peak upon exposure to water confirmed that this peak is probablydue to unreacted acid chloride (spectrum c, Figure 4.2). Nevertheless, the small size of the acid chloride peak suggests that cross-linking is >50%. Figure 4.3 shows results of diffusion dialysis through a composite membrane containing a 35 nm-thick cross-linked PHEMA skin. (The PHEMA film was ~28 nm- thick prior to cross-linking, so these results can be compared to those in Figure 4.1.) Table 4.3 summarizes the fluxes and selectivities of chemically cross-linked PHEMA membranes. For the 35 nm-thick film (initially 28 nm of PHEMA), reaction with succinyl chloride increased Cl’lSO42' selectivity by a factor of 100 and Cl’/Fe(CN)63' selectivity by a factor of 90. The large increase in monovalent/multivalent anion selectivities suggests that in addition to cross-linking, which reduces membrane swelling and increases size-based selectivity, we also introduced fixed negative charges into the PHEMA film. As reflectance FTIR spectra indicate, these negative charges probably result from hydrolysis of unreacted acid chlorides. When the membrane is negatively charged, Donnan potentials cause substantial exclusion of multiply charged anions such as S042” and Fe(CN)63' and hence, higher monovalent! multivalent anion selectivities. In contrast to anion transport, K+/Mg2+ selectivity increases only 4-fold after cross-linking. Negative charges in the membrane should reduce monovalent/divalent cation selectivity, and thus we see only a small increase in Ki'fMg2+ selectivity after cross-linking. The 4- fold increase in selectivity is likely due to a reduction in film swelling. Consistent with decreased film swelling, large increases in anion-transport selectivities with thin PHEMA 126 Receiving Phase Concentration (mM) 1.6 o 0.06 ~ . " ' o 0.04 r . I 1.2 ~ 0.02 ~ . ' ' - A 0 ‘* i T ‘ f ‘ f ‘ T 0 0 10 20 30 40 o 0.8 - 0 O C KCI MgClz 0.4 - ' . K2304 . A K3Fe(CN)6 v V V o o T x x I X X X n n n 0 10 20 3O 40 Time (min) Figure 4.3: Plot of receiving phase concentration as a function of time in diffusion dialysis when the source phase (0.1 M salt) was separated from the receiving phase (initially deionized water) by a composite membrane of 35 nm cross-linked PHEMA film. The inset expands the concentration scale for SO42' and F e(CN)63 ’. PHEMA was cross-linked by succinyl chloride. 127 - . .....i..1,..p. 000:? 08:: 00 8:00”v .:0::000 0808:0000 0:: 8 000::0000 00 :0:: 08:: .0> 000:: 088000: :0 50880800 :0 00:80 0:: 80:: 00:83:00 0:03 00::0> 03:”? 00::000: 0:0 0000058581.: 0803800 800000 0808.898 8 0080000 00 0:0003 0:00 :0 08:0 080:0:00800 :0 00000538: 080800050 0080008 80:: 00:08:00 00 00000585. 0 128 NN how 00m , N: mod mod mm mm N 00 mm 0: de 0.: 0m :N .0025: 0220086 00020 :02 022000: -000 :0 :55 «00088:? $838200 o0:x 00880208: 003:: 8:0: 0:88:00 ::00-0:080 :83 80380 :0:0:.:::0 8 00:0:0808 <:>:m:::n: 0088:0080 :0: 0088:0200 0:0 0003:: 00000808: 8:”: "0.0 030,—. films are accompanied by a 40% decrease in C1' flux (See Tables 4.1 and 4.2). However, flux is still about 50% of that through bare alumina. The ion-transport studies described above were performed with source-phase solutions that contained single salts. Actual ion separations occur from mixed salts, however, so we briefly examined ion-transport with both KCl and K2S04 in the source phase. Table 4.4 contains preliminary selectivities and flux values obtained with cross- linked PHEMA membranes. The Cl'/SO42' selectivity obtained with solutions of mixed salts is a factor of 1.6 to 4.5 higher (depending on the ratio of C1' to 8042') than the selectivity obtained from single-salt experiments. The increase in selectivity with mixed solutions is presumably due the fact that the diffusion potential is lower in mixed KCI/K2S04 solutions than it is when only K2S04 is present. In K2SO4 solutions, diffusion of K” is faster than diffusion of S042} so a diffusion potential develops to resist current flow. The diffiision potential decreases K+ transport and increases SO42' transport. When KCl is present, a smaller diffusion potential develops because Cl' has a higher mobility than SO42" The no-current condition can be reached with a smaller potential that increases Cl’ flux much more than SO42' flux. 4.3.4 Crown Ether-Derivatized PHEMA Membranes and their Ion-Transport Properties We hoped to improve the selectivity of PHEMA membranes by derivatization with a complexing agent that can selectively bind a particular ion. Selective hopping of ions between crown-ether binding sites could facilitate the transport of one ion over another. Crown ethers are well known for both selective binding and fast release of alkali metal 129 Table 4.4: Film thicknesses, normalized fluxes and selectivities for cross-linked PHEMA membranes in diffusion dialysis with mixed salt solutions. Normalized F luxd X109 Selectivity° Film Thickness“l 2 (mm) or so4 ‘ cr/soi' 21b 450 4 106 3 5c 440 0.9 490 aThicknesses are estimated from measured ellipsometric thicknesses of corresponding films on gold wafers as described in experimental section. Thicknesses after derivatization are reported. bDiffusion dialysis was done using a 1000 ppm Cl' and 1000 ppm S0421 °Diffusion dialysis was done using 500 ppm Cl' and 2500 ppm S0421 dNormalized fluxes are reported. °Ratio of noramalized fluxes. 130 v cations.l3 We derivatized PHEMA with 2-hydroxymethyl-18-crown-6 because the size of the crown ether cavity matches with the ionic radius of K+.M’15 In the derivatization of PHEMA, films were first reacted with CD1 for 12 hours (Step 1, Scheme 4.2).9 After coupling of PHEMA with CD1, the reflectance FTIR spectrum of the film showed complete disappearance of the hydroxyl peak (3500 cm"- 33000m'1) and the appearance of a strong carbonyl peak at 1771 cm"1 due to the imidazole carboxylic acid intermediate, suggesting quantitative reaction (Spectrum b, Figure 4.4).9 Since imidazole groups can be displaced with nucleophiles such as amines and alcohols, CDT-derivatized PHEMA is attractive for the immobilization of a variety of functional groups. In this case imidazole-derivatized PHEMA films were reacted with 2- hydroxymethyl-l 8-crown-6 to introduce crown ether functionalities into the film (Step 2, Scheme 4.2). Since the reaction was very slow, the solution was heated to 70 °C for 4.5 days. Disappearance of the carbonyl peak at 1771 cm'1 and the appearance of peaks at 1265 cm'1 and 1155 cm'1 (C-O-C stretches of the crown ether) in the reflectance FTIR spectrum of derivatized films confirmed the reaction with the crown ether (spectrum c, Figure 4.3). Additionally, the ellipsometric thickness increased by 55 %, indicating incorporation of crown ether moieties. However, diffusion dialysis yielded a K+ flux (2.3 x 10''0 molescm'zs'l) through crown ether derivatized membranes that was smaller than Na+ flux (2.9 x 10'10 molescm' 2s"). The lack of K‘i/Na+ selectivity could be due to crown ether moieties that are not in the right conformation for K+ ion binding, or perhaps hopping between binding sites is slower than diffusion. A better way of performing this derivatization would be to use a K+-salt of the crown ether and subsequently remove the K+ ions afier derivatization. 131 (H) CH3 / \ wwNH— C— C %Hz— Cl(Br) + (\N—C—N/W N C”) \éN OCHchZOH con Step 1 DMF RT, 12h CH3 0... 'c' (30%05' +00.) __ :<—o CHZCHZOC— 51:: /CH20H Step 2 % fl DMF, 70°C 0 CH3 WWNH— C— C—ECHOz—C :+CI(Br) CHzCHZOCOCHz Scheme 4.2: Functionalization of PHEMA with 2-hydroxy methyl 18- crown-6. 132 0.02 N Absorbance . 0.8 (b) J (a) M l 1 J l l l 3500 3000 2500 2000 1500 1000 Wavenumbers (cm") Figure 4.4: Reflectance FTIR spectra of (a) a grafted PHEMA layer (28 nm) (b) CDI-derivatized PHEMA and (c) CDI-derivatized PHEMA after reaction with 2-hydroxymethyl-l 8-crown-6 (43 nm). 133 4.4 Conclusions Surface-initiated ATRP is an attractive and versatile method of forming ultrathin polymer skins on porous supports. Ion—permeability studies with composite PHEMA membranes showed moderate selectivities for monovalent over divalent and trivalent ions. However, reaction with succinyl chloride enhanced Cl'/SO42' and Cl'/ Fe(CN)63' selectivities by 2 orders of magnitude, presumably due to the introduction of cross- linking and negative charges into the membrane. Using mixed-salt solutions, even better selectivities were obtained because of a reduction in diffusion potentials. 134 4.5 (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) References Xu, X.; Spencer, G. Desalination 1997, 113, 85-93. Wang, X.-L.; Zhang, C.; Ouyang, P. J. Membr. Sci. 2002, 204, 271-281. Bowen, W. R.; Mohammad, A. W.; Hilal, N. J. Membr. Sci. 1997, 126, 91-105. Bowen, W. R.; Mukhtar, H. J. Membr. Sci. 1996, 112, 263-274. Haraya, K.; Li, S.; Mizoguchi, K. Springer Series in Materials Science 1999, 35, 95-124. Baddour, R. F .; Graves, D. J .; Vieth, W. R. J. Colloid Sci. 1965, 20, 1057-1069. Jadwin, T. A.; Hoffman, A. S.; Vieth, W. R. J. Appl. Polym. Sci. 1970, 14, 1339- 1359. Halden, R. A.; Lee, B. E. Br. Polym. J. 1972, 4, 491-501. Huang, W.; Kim, J .-B.; Bruening, M. L.; Baker, G. L. Macromolecules 2002, 35, 1 175-1 179. Wang, X. S.; Lascelles, S. F.; Jackson, R. A.; Armes, S. P. Chem. Commun. 1999, 1817-1818. Robinson, K. L.; Khan, M. A.; de Banez, M. V.; Wang, X. S.; Annes, S. P. Macromolecules 2001, 34, 3155-3158. Marcus, Y. Biophys. Chem. 1994, 51, 111-127. Noble, R. D. J. Chem. Soc. Faraday Trans. 1991, 8 7, 2089-2092. Thunhorst, K. L.; Noble, R. D.; Bowman, C. N. J. Membr. Sci. 1999, 156, 293- 302. Reusch, C. F.; Cussler, E. L. AIChE Journal 1973, 19, 736-741. 135 CHAPTER 5 Conclusions and Future Work 5.1 Conclusions The work reported in this dissertation demonstrated the versatility of multilayer polyelectrolyte deposition and surface-initiated atom transfer radical f polymerization (ATRP) in the formation of defect-free, ultrathin membrane skins. These skins are quite selective in both ion and gas separations. Chapter 2 described a method of introducing fixed negative charges into the bulk of multilayer polyelectrolyte membranes (MPMS) to enhance anion-transport selectivities. This method relies on complexation of Cu2+ by the —C00' groups of poly(acrylic acid) (PAA) during the deposition of PAA/ poly(allylamine hydrochloride) (PAH) films on porous alumina supports. Subsequent removal of Cu2+ and deprotonation results in ion-exchange sites in the bulk of the membrane skin. Diffusion dialysis studies with Cu2+-templated PAA/PAH membranes showed a 4-fold increase in Cl'/SO42' selectivity compared to pure PAA/PAH membranes deposited under similar conditions. Post deposition cross-linking of these membranes further increased Cl' /SO42' selectivity to values as high as 600. These remarkable selectivities are presumably due to increased fixed negative charge density in the bulk of the membrane, which increases the Donnan potential to give greater exclusion of divalent than monovalent anions. However, modeling of ion-transport data suggested that selectivity is due to both Donnan exclusion and diffusivity differences among ions. 136 The second method I utilized in thin film formation was room-temperature, surface-initiated ATRP. Chapter 3 showed the versatility of this technique in the synthesis of cross-linked polymer fihns from a modified porous alumina support. Cross-linked poly(ethylene glycol dimethacrylate) (PEGDMA) membranes exhibited a COz/CH4 selectivity of 20. In comparison non cross-linked poly(2-hydroxyethyl methacrylate) (PHEMA) grown form a modified porous alumina support showed minimal gas-transport selectivities. However, derivatization of the hydroxyl groups of PHEMA with fluorinated acid chloride yielded moderate gas-transport selectivity. ":1 ' VA -._.— Polymer growth was monitored by scanning electron microscopy (SEM) and transmission FTIR spectroscopy. Both top-down and cross-sectional SEM images showed that these polymer films effectivelycover the surface pores of the alumina support. Chapter 4 discussed the promise of PHEMA membranes in ion separations. PHEMA by itself showed moderate ion-transport selectivities. However, cross- linking PHEMA via reaction with succinyl chloride increased Cl'/SO42' and Cl' /Fe(CN)63' selectivity by IOO-fold. This large increase in anion-selectivities is probably due to reduced film swelling and introduction of negative charge by hydrolysis of unreacted acid chlorides. Overall, these studies demonstrate novel methods for the formation of ultrathin membranes. Although new procedures for film deposition may not be practical on a large scale, they should allow development of membrane for specific small-scale (i.e., analytical) separations. Work with both multilayer polyelectrolyte and grafted polymer films shows that the minimal thickness of these systems allows 137 high flux. Additionally, the wide variety of functional groups that can be included in these membranes allows tailoring of transport properties. 5.2 Suggestions for Future Work My success in growth of a cross-linked film from a porous substrate should now allow investigation of new types of membrane systems such as imprinted polymers. Cross-linked polymers are attractive materials for separations because of their low free volume (which will reduce film swelling) and ability to withstand drastic separation conditions. An interesting area of research will be the examination of PEGDMA membranes in ion separations. However, ion-transport through PEGDMA itself is quite low, and thus, incorporation of non cross-linkable monomers into PEGDMA films will likely be necessary to achieve a desirable flux. Introduction of charged functionalities into these films could be achieved by co-polyrnerizing ethylene glycol dimethacrylate (EGDMA) with t-butyl methacrylate and subsequently hydrolyzing t-butyl groups to —COOH groups. Deprotonation of these —COOH groups will yield ion-exchange sites in the membrane. Such membranes have the potential to be extremely selective in anion separations. In chapter 4, I showed the potential of derivatized poly(2-hydroxyethyl methacrylate) (PHEMA) films in anion separations. Similarly PHEMA could be derivatized for cation separation by reaction with a di-arnine (e. g. ethylene diamine). This d erivatization c ould b e e asily p erformed b y first coupling P HEMA with 1 ,1'- carbonyldiimidazole (CD1) and then reacting CDI-derivatized PHEMA with the di- amine.‘ Similar to reaction of PHEMA with succinyl chloride, derivatization with a 138 diamine should result in cross—linking along with residual free amine groups that can give fixed positive charge to the membrane. It will also be interesting to test PHEMA and its derivatives in neutral-molecule separations. This will provide valuable information in understanding the separation mechanism in these systems. As previous studies demonstrated2 that thick PHEMA membranes can be used in nanofiltration, it will definitely be interesting to test ultrathin PHEMA and cross- linked PHEMA (HEMA copolymerized with EGDMA) films in nanofiltration. Another very interesting and challenging area of research will be the synthesis of ultrathin, imprinted polymer membranes. In imprinting, a template, a monomer and a cross-linker are polymerized together, and the subsequent release of the template results in a polymer that contains cavities that are selective for the template. 3‘5 Imprinted polymers could be prepared by copolymerizing EGDMA (cross-linker) with HEMA that was hydrogen bonded with a template molecule. For templates, we could use any molecule of interest (possibly a high molecular weight species) that can interact with HEMA through hydrogen bonding. Removal of the template should leave behind recognition sites in the membrane through which transport could take place. These imprinted membranes may be useful in selective separation of the template molecule. In summary, the development of new techniques for forming ultrathin membrane skins has the potential to yield a wide variety of separation membranes ranging from imprinted to ion-exchange systems. 139 I; 1'1 -'.v.’.::i'flT.a‘ “km 5.3 (1) (2) (3) (4) (5) References Huang, W.; Kim, J.-B.; Bruening, M. L.; Baker, G. L. Macromolecules 2002, 35, 1175-1179. Jadwin, T. A.; Hoffman, A. S.; Vieth, W. R. J. Appl. Polym. Sci. 1970, 14, 1339-1359. Shea, K. J. Trends Polym. Sci. 1994, 2, 166-173. Wulff, G. Angew. Chem. Int. Ed. 1995, 34, 1812-1832. Kriz, D.; Rarnstrom, 0.; Mosbach, K. Anal. Chem. 1997, 69, 345A-349A. 140 Wthvruiuihflm.