4.... 3.411. 9% n .4: :1. {83. an . 3"»er m :39... . 1.5 @551. s «1 a ‘ g... Jun... .. .2... . w. .23: .- 1. .. 1.5 st 5. 3.!5. J.‘ 7‘ ‘ if ‘ (If G: In" LS {s 300; This is to certify that the dissertation entitled Molecular Control of Interfacial Properties: The Construction, Characterization and Application of Tailored Interfaces presented by Jaycoda Sandor Major has been accepted towards fulfillment of the requirements for Ph . D . degree in Chemistry flee ¥gf/ /M/ajor prof sor Date §/// 01/ 0A MS U is an Affirmative Action/Eq ual Opportunity Institution 0-12771 LIBVARY 0i Michigan mate University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cJCIFiC/DateDuopSS—ot 5 Molecular Control of Interfacial Properties: The Construction, Characterization and Application of Tailored Interfaces By J aycoda Sandor Major A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree DOCTOR OF PHILOSOPHY - Department of Chemistry 2002 ABSTRACT Molecular Control of Interfacial Properties: The Construction, Characterization and Application of Tailored Interfaces By J aycoda Sandor Major The work reported in this dissertation has focused primarily on the design of novel layered interfacial materials where there is discrete control over the resulting molecular properties of the material achieved as a function of design and assembly. In the first part of this work, we report on the design of a variety of polymer materials synthesized with various side groups that can be used in several covalent linking strategies. As a consequence of design, these polymers are capable of incorporating a variety of functional constituents within the matrix of the material. The ability to control the identity/composition of the reactive functionalities selectively within the matrix of the material affords the direct control over the resulting property of the material. We have also demonstrated our ability to control the effective loading density of the polymer adlayers through catalytic and/or dehydration mediation of the various covalent crosslinking reactions we report. Such control can undoubtedly be used in the design and improvement of sensors where the ability to increase the number of the sensing elements within each layer can lead to enhanced sensitivity. This characteristic also allows for the design of materials where the conductivity of the material can be controlled by the number elements incorporated within the matrix. The second part of this dissertation focuses on the assembly and characterization of these materials. Here, we employ an in-house assembled system for simultaneous w “Haunt-o9...“ quartz crystal microbalance (QCM) gravimetry and optical ellipsometry — a system that allows the simultaneous in situ acquisition of isotherm and thickness data. In this work, we have demonstrated our ability to control adsorption behavior of the interfaces as a function of the order and chemical identity of the interfacial adlayers. We have also shown that, by varying the functional groups within each layer, it is possible to control the adsorption response relative to the single functional group systems. The underlying implication of this work is that it is indeed possible to design systems that exhibit a chemical potential gradient where, in this particular case, the gradient is generated by varying the dipole moment of the individual organic side groups incorporated into the polymers used in making interfacial layers. The ability to design systems that exhibit a gradient in some property can find use in separations systems where the desired outcome is the continuous extraction of some analyte/solute leading to improved separations. Finally, we have demonstrated that through simple design strategies, we are able to tailor interfacial material properties with predictable results. We have designed a copolymer of 4-hydroxyphenymaleimide and a vinyl ether phosponate, where the phosphonate groups remain protected while the aromatic hydroxyl groups are used in the covalent crosslinking of interfacial multilayers. Subsequent to adlayer growth, we deprotect the phosphonate groups and use them for metal ion sequestration. This work illustrates the range of potential application for these materials. DEDICATION Dedicated to my parents, Theodore and Christina and to my other family members. To my nieces and nephews, this is the first; don’t let it be the last. -iv- ACKNOWLEDGEMENTS I would first like to express my sincere gratitude to my advisor Dr. Gary Blanchard who has been so instrumental in my scientific and professional development. You were always so understanding and helpful. I owe you so much, and I hope that at some point in life, I am able to be of some assistance to you. Thanks to Dr. Simon Garrett for always listening to me vent about my unpleasant experiences (at least the major one) at MSU. Thanks to both you and Gary for convincing me to stay at MSU. You have saved me a few years and a few brain cells. To Dr. Kathy Severin, you were also so helpful to me, always giving me my last five or ten QCMs. YOU ROCK KATHY!!! Thanks to all of my committee members for being so understanding. I would like to express my sincere thanks to Dr. Yvonne Gindt who convinced me, actually, insisted that I attend graduate school. Thanks for always assuring me that even the bad times “built character”. You were always willing to listen to me let off steam and you always had great advice. Thanks for checking in on me over the years, this level of commitment and concern has truly been appreciated and will never be forgotten. To my friends at church that made my stay in Lansing more enjoyable, especially Bazi, Mitchell, Lulama, Inga, Neliswa, Schvonne, Anne, Brent, Simone, Ruth and all of the other people whose names escape me at this late hour, thank you. You guys helped me through some rough times and helped me keep my faith in God intact. To all of my other friends that have made my stay enjoyable especially Charles and Karen Ngowe, Dan and Nicole Wampfler, Simona, Dalila, Erik, Fadi, Kam, Matt and Maggie, and everyone else that I don’t remember right now (that old timers disease is kicking in, as well as the lack of food and sleep) thank you. Julian and Letitia Newbold, Garnett Burgess, Carla Anderson, Antionette Moxey, Patrice King, Valderene Gardiner and Deon Stewart, thank you guys for being such great friends. Even though miles and years have separated us, you have never allowed these things to affect our relationships. Thanks for racking up the telephone bills or spending monies to fly all the way from the Bahamas, or wherever you are just to spend some time with me and to make sure that I was always okay. YOU ARE THE TRUE DEFINITION OF FRIENDS. Finally, to all the past and present members of the Blanchard group (most of you anyway) especially Steve, Michelle, (“yeah baby, yeah ”) Mark, Punit, Lee, Scott, Shawn and all of you other crazy people that made my transition to, and stay in the Blanchard lab such an UNFORGETABLE experience - thanks. If I can be of assistance in the future, you know how to contact me. Finally, to those individuals who had direct hands in making parts of my stay at MSU a less than pleasant ordeal — THANK YOU!!! It was those experiences that helped to strengthen my resolve to work harder and to be successful - it worked. I would also like to dedicate this work to you. God bless you and I wish you all the best. -Vl- Table of Contents Page List of Tables ......................................................................................... ix List of Figures ........................................................................................ x List of Schemes ...................................................................................... xv Chapter 1. Introduction ........................................................................ 1 1.1 Literature Cited .................................................................... 12 Chapter 2. Strategies for Covalent Multilayer Growth 1. Polymer Design and Characterization ................................................................... l6 2. 1 Introduction ............................................................... 17 2.2 Experimental .............................................................. 19 2.3 Results and Discussion .................................................. 26 2.4 Conclusions ............................................................... 43 2.5 Literature Cited ........................................................... 44 Chapter 3. Strategies for Covalent Multilayer Growth 2. Interlayer Linking Chemistry ........................................................................... 47 3. 1 Introduction ............................................................... 48 3.2 Experimental .............................................................. 49 3.3 Results and Discussion .................................................. 52 3.4 Conclusions ............................................................... 78 3.5 Literature Cited .......................................................... 79 Chapter 4. Acid-Enhanced Interfacial Polymer Layer Growth .......................... 82 4. 1 Introduction ............................................................... 83 4.2 Experimental ............................................................. 86 4.3 Results and Discussion ................................................. 89 4.4 Conclusions ............................................................. 104 4.5 Literature Cited ......................................................... 108 Chapter 5. Adsorption Behavior of Polymer Modified Interfaces ..................... 112 5.1 Introduction .............................................................. 1 13 5.2 Experimental ............................................................ 1 15 5.3 Results and Discussion ................................................ 118 5.4 Conclusions ............................................................. 135 5.5 Literature Cited ......................................................... 137 -vii- Chapter 6. 6.1 6.2 6.3 6.4 6.5 Chapter 7. 7.1 7.2 7.3 7.4 7.5 Chapter 8. 8.1 8.2 Molecular control of Adsorption Properties of Polymer Modified Interfaces ......................................................................... 139 Introduction ............................................................. 140 Experimental ............................................................ 142 Results and Discussion ................................................. 149 Conclusions ............................................................. 160 Literature Cited ......................................................... 161 Covalently-Bound Polymer Multilayers for Efficient Metal Ion Sorption ............................................................................ 163 Introduction ............................................................. 164 Experimental ............................................................ 166 Results and Discussion ................................................ 171 Conclusions ............................................................. 184 Literature Cited ......................................................... 186 Conclusions and Future Prognosis ............................................ 188 Conclusions ...................................................................... 188 Future Prognosis ................................................................. 190 - viii - wry-o”:- my Table 4.1 Table 4.2 Table 4.3 Table 5.1 Table 6.1 Table 6.2 LIST OF TABLES Slopes of the UV-Visible absorbance data for the addition of HCL and H230; as a function of concentration for the crosslinking of poly(NPM-VOB) with adipoyl chloride to form an ester linkage .............. 106 Dependence of ellipsometric thickness on acid addition for amide, ester, urea, and urethane crosslinked adlayers ................................................. 106 Dependence of absorbance data on acid addition for amide, ester, urea and urethane polymer adlayers ......................................................... 107 Layer thickness, monolayer adsorbate volume (le) and enthalpy of desorption (Adcs) extracted from experimental data and fits to BET isotherm .............................................................................. 136 Calculated dipole moments of selected polymer side groups. Calculations Performed using Hyperchem V. 6.0 with PM3 parameterization ............ 152 Results of fitting experimental adsorption isotherm data to equation 6.2 .......................................................................... 160 -ix- LIST OF FIGURES Figure 2.1 Polymer structure. (a) Poly(NPM-VOB), (b) Poly(NPM-APVE), (c) Poly(NPM-4-PC), (d) Poly(NAI) .......................................... 24 (e) Poly(3CPM-hex), (f) Poly(MP-APVE), (g) Poly(4-HPM-VEP) ...... 25 Figure 2.2 FTIR spectrum (KBr pellet) of poly(NPM-VOB) ........................... 28 Figure 2.3 1H-NMR spectrum of poly(NPM-VOB) ....................................... 30 Figure 2.4 FT IR spectrum (KBr pellet) of poly(NPM-APVE) .......................... 31 Figure 2.5 1H-NMR spectrum of poly(NPM-APVE) .................................... 32 Figure 2.6 (a) FTIR spectrum of poly(NPM-4-PC), (b) lH-N MR spectrum of poly(NPM-4-PC) ................................................................. 34 Figure 2.7 (a) FI'IR spectrum of poly(4NAI), (b) 1H-NMR of pol y(4NAI) .......... 36 Figure 2.8 FTIR spectrum of poly(3CPM-hex) ........................................... 37 Figure 2.9 1H-NMR spectrum of poly(3CPM—hex) ....................................... 38 Figure 2.10 (a) FI‘IR spectrum (KBr pellet) of poly(MP-APVE), (b) 1H-NMR spectrum of poly(MP-APVE) .................................................. 40 Figure 2.11 (a) FI‘IR spectrum (KBr pellet) of poly(4-HPM-VEP), (b) 1H—NMR spectrum of poly(4-HPM-VEP) ................................................ 42 Figure 3.1 (a)FTIR spectrum of poly(3CPM-APVE) with amide interlayer linkage (b) UV-visible spectra as a function of number of layers for poly(3CPM-APVE) .............................................................. 57 Figure 3.2 Dependence of ellipsometric thickness on number of layers for poly(3CPM-APVE) with amide interlayer linkage .......................... 58 Figure 3.3 (a) Dependence of ellipsometric thickness on number of layers for poly (NPM-VOB) with ester interlayer linkage. (b) UV-Visible spectra as a function of number of layers. Inset: Dependence of absorption maximum on number of layers .............................................................. 61 Figure 3.4 Poly(HPM-VEP) with ionic and covalent interlayer linkages. (a) Infrared spectrum of ionically-bound multilayers and (b) IR spectrum of covalently-bound multilayers. The spectra have been offset for clarity...62 Figure 3.5 Figure 3.6 Figure 3.7 Figure 3.8 Figure 3.9 Figure 3.10 Figure 3.11 Figure 3.12 Figure 3.13 Figure 4.1 UV-Visible spectra as a function of numerb of layers. Inset: Dependence Of absorption maximum on number of layers .............................. 64 (3) Dependence of ellipsometric thickness on number of layers for poly(NPM-VOB) with ether interlayer linkage. (b) UV-Visible spectrum for poly(NPM—VOB) with ether interlayer linkage. Inset: Dependence of absorption maximum on number of layers .................................. 66 IR spectrum for poly(NPM-VOB) with ether interlayer linkage ......... 67 Ether chemistry for modified substrate. (3) FTIR showing C-O-C stretch of epoxide prior to reaction, (b) FTIR of ring-opened system after reaction indicating ketone and acid functionalities .................................... 68 (a) IR spectrum of poly(3CPM-APVE) with urea interlayer linkage. (b) UV-Visible spectra as a function of number of layers for poly(3CPM- APVE) with urea interlayer linkage. Inset: Dependence of absorption maximum on number of layers ................................................ 71 Dependence of ellipsometric thickness on number of layers for poly(3CPM-APVE) with urea interlayer linkage ............................ 72 Dependence of ellipsometric thickness on number of layers for poly(NPM-VOB) with urethane interlayer linkage ......................... 74 (a) IR spectrum for poly(NPM-VOB) with urethane interlayer linkage. (b) UV-Visible spectra as a function of number of layers for poly(NPM-VOB) with urethane interlayer linkae. Inset: Dependence of absorption maximum on number of layers ................................................ 75 Poly(NPM-4-PC). (a) Structures - with amide and ester linkers. (b) IR spectrum of a poly(NPM-4-PC) layer bound to the substrate by an ester linkage and with remaining acid chloride functionalities hydrolyzed by exposure to air ................................................................... 77 Figure 1. Data for amide-linked poly(3CPM-APVE). (3) Dependence of optical absorption at 245 nm on number of polymer adlayers. I are data points for acid-free growth, 0 are data for adlayer growth with 0.145 M HCl and A are data for adlayer growth with 0.216 M HZSO4. Inset: UV- Visible absorption spectrum for a four layer assembly of poly(3CPM- APVE). (b) Ellipsometric thickness of poly(3CPM-APVE) adlayers as a function number of layers deposited for growth without acid added (I) and with 0.216 M H2304 added (A ). (c) FT IR spectrum of a two-layer assembly of poly(3CPM-APVE) .............................................. 92 -xi- Figure 4.2. Figure 4.3. Figure 4.4. Figure 4.5. Figure 4.6. Figure 5.1 Figur 5.2 FI‘IR spectrum of substrate treated with adipoyl chloride showing essentially complete conversion of the terminal acid chlorides to carboxylic acids ................................................................... 93 Data for ester-linked poly(NPM-VOB). (a) Dependence of optical absorption at 231 nm on number of polymer adlayers. I are data points for acid-free growth, 0 are data for adlayer growth with 0.145 M HCl and A are data for adlayer growth with 0.216 M H2304. Inset: UV-Visible absorption spectrum for a four layer assembly of poly(NPM-VOB). (b) Ellipsometric thickness of poly(NPM-VOB) adlayers as a function number of layers deposited for growth without acid added (I), with 0.145 M HCl (0) and with 0.216 M H2804 added (A). (c) FTIR spectrum of a two- layer assembly of poly(NPM-VOB) ............................................ 95 Data for urea-linked poly(4BPM-APVE). (a) Dependence of optical growth, 0 are data for adlayer growth with 0.145 M HCl and A are data for adlayer growth with 0.216 M H2804. Inset: UV-Visible absorption spectrum for a four layer assembly of poly(4BPM-APVE). (b) Ellipsometric thickness of poly(4BPM-APVE) adlayers as a function number of layers deposited for growth without acid added (I), with 0.145 M HCl (0) and with 0.216 M H2804 added (A). (c) FTIR spectrum of a two-layer assembly of poly(4BPM-APVE) ................................... 98 Data for urethane crosslinked poly(NPM-VOB). (a) Dependence of optical absorption at 231 nm on number of polymer adlayers. I are data points for acid-free growth, 0 are data for adlayer growth with 0.145 M HCl and A are data for adlayer growth with 0.216 M H2304. Inset: UV- Visible absorption spectrum for a four layer assembly of poly(NPM- VOB). (b) Ellipsometric thickness of poly(NPM-VOB) adlayers as a function number of layers deposited for growth without acid added (I), with 0.145 M HCl (0) and with 0.216 M H2804 added (A). (c) FTIR spectrum of a two-layer assembly of poly(NPM-VOB) .................... 101 Dependence of ester-bound poly(NPM-VOB) monolayer absorbance at 231 nm on acid concentration for the initial polymer layer. (a) HCl added, (b) H2804 added. The different concentration dependencies for each acid indicates the importance of both protonation and dehydration in the ester formation reaction ............................................................... 103 Structure of molecules bound to QCM surfaces. (a) 6-mercapto-1- hexanol, (b) poly(4HPM-VEP), (c) poly(4MPA-VEP) .................... 119 (d) poly(4BPM-APVE), (e) poly(3CPM-APVE) ........................... 120 Adsorption isotherm data for uncoated gold QCM. (a) Data and best-fit line for adsorption of methanol and (b) data and best-fit line for adsorption -xii- fl 03”.“;fl‘fl Figure 5.3 Figure 5.4 Figure 5.5 Figure 5.6 Figure 6.1 Figure 6.2 Figure 6.3 Figure 6.4 Figure 6.5 Figure 6.6 of hexane. The data were fit t the BET model as given in Eqs. 2 and 3. Results of the fits are presented in Table 5.1 ................................ 124 Adsorption isotherm data for a QCM modified with 6-mercapto-l- hexanol. (a) Data and best-fit line for adsorption of methanol and (b) data and best-fit line for adsorption of hexane. The data were fitted to the BET model as given in Eqs. 2 and 3. Results of the fits are presented in Table 5.1 ................................................................................. 126 Adsorption isotherm data for a QCM modified with poly(4I-IPM-VEP). (a) Data and best-fit line for adsorption of methanol and (b) data and best- fit line for adsorption of hexane. The data were fitted to the BET model as given in Eqs. 2 and 3. Results of the fits are presented in Table 5.1 .................................................................................. 130 Absorption spectrum of poly(4MPA-APVE) on a quartz substrate. The band centered near 325 nm is the Sz<—So transition for the trans side groups and the band centered near 240 nm is the Szé—So transition for the cis side groups. (b) Data and best-fit line for adsorption of methanol (0) and hexane (o). The data were fited to the BET model as given in Eqs. 2 and 3. Results of the fits are presented n Table 5.1 ........................ 131 Methanol adsorption isotherm data and best-fit lines for a QCM modified with (a) poly(4BPM-APVE) and (b) poly(3CPM-APVE) with one polymer layer (0) and two polymer layers (O). The data were fitted to the BET model as given in Eqs. 2 and 3. Results of the fits are presented in Table 5.1 .......................................................................... 134 BET isotherm for the adsorption of methanol on S-BPM-NPM and S- NPM-BPM ........................................................................ 152 BET isotherm for he adsorption of hexane on S-BPM-NPM and S-NPM- BPM ............................................................................... 153 Isotherms for the adsorption of methanol on S-BPM-CPM and S-CPM- BPM ............................................................................... 155 Isotherms for the adsorption of hexane on S-BPM-CPM and S-CPM - BPM ............................................................................... 156 Isotherms for the adsorption of methanol on S-BPM-HPM and S-HPM- BPM ............................................................................... 157 Forward and backward isotherms for the exposure of S-BPM-CPM to methanol .......................................................................... 159 - xiii - Figure 7.1 Figure 7.2 Figure 7.3 Figure 7.4 Figure 7.5 Figure 7.6 Figure 7.7 Ellipsometric data for a four-layered MVE polymer system. Each layer consists of the MVE polymer and the adipoyl chloride underlayer. The slope of the best-fit line through the data is 16i1A/layer .................. 173 (a) FT IR absorbance data layers 1-4. The spectra are offset for clarity of presentation. (b) Carbonyl stretching band absorbance as a function of number of layers .................................................................. 174 UV-Visible absorbance spectra of MVE polymer layers as a function of number of layers. Inset: Dependence of the 231 nm absorption intensity as function of number of layers ................................................ 176 FT IR absorbance data for a four-layered system prior to deprotection of the phosphonate groups with BTMS (spectrum I), after deprotection with BTMS followed by hydrolysis (spectrum 11), and after exposure to ZrOClz (spectrum III). The spectra are offset for clarity of presentation. . . . . 177 Comparison of the four-layered system, after exposure to ZrOClz (spectrum HI), and after rinsing with absolute ethanol, then water (spectrum IV). These spectra are offset for clarity of presentation ...... 179 XPS survey scan of a surface with four layers of poly(4HPM-VEP) on a gold substrate ..................................................................... 181 QCM data showing rapid uptake of Zr4+ in a four-layer film .............. 183 -xiv- Scheme 2.1 Scheme 2.2 Scheme 3.1 Scheme 3.2 Scheme 3.3 Scheme 3.4 Scheme 3.5 Scheme 4.1 Scheme 6.1 LIST OF SCHEMES Reaction of maleic anhydride with an aromatic amine to produce the maleamic acid, followed by ring closure to the maleimide .................. 20 Synthetic route for the polymerization of maleimide and vinyl ether monomers (top), and for maleimide and vinyl monomers (bottom). Both reactions are initiated with AIBN ............................................... 21 Reaction schematic and idealized structures of amide-linked MVE layered assembly. (a) Amine-containing MVE polymer bound to acid chloride functionalized surface and (b) acid chloride containing MVE polymer bound to amine-functionalized surface ............................................................................... 55 Reaction schematic and idealized structures of ester-linked MVE layered assembly. (a) Alcohol-containing MVE polymer bound to acid chloride functionalized surface and (b) acid chloride containing MVE polymer bound to hydroxyl-functionalized surface ..................................... 59 Reaction schematic of a glicidyl ether with an alcohol (top) to form an ether and an idealized structure of an alcohol-containing MVE polymer bound to an expoxide functionalized substrate. Subsequent layer growth requires reaction of the alcohol-containing MVE polymer with a diglicidyl ether ................................................................................ 65 Reaction schematic of an isocyanate with an amine (top) to form a urea and an idealized structure of an amine-containing MVE polymer bound to an isocyanate-functionalized substrate. Subsequent layer growth requires reaction of the amine-containing MVE polymer with a diisocyanate. . . ..70 Reaction schematic of an isocyanate with an alcohol (top) to form a urethane and an idealized structure of an alcohol-containing MVE polymer bound to an isocyanate-functionalized substrate. Subsequent layer growth requires reaction of the alcohol-containing MVE polymer with a diisocyanate ............................................................... 73 Top panel, left to right: Structures of pol y(3CPM-APVE), poly(NPM- VOB) and poly(4-BPM-APVE). Bottom panel: Reaction schemes for the several crosslinking strategies employed in this work ...................... 88 Interface structures. (a) S-NPM-BPM, (b) S-BPM-NPM .................. 145 (c) S-BPM-CPM, (d) S-CPM-BPM ........................................... 146 -xv- Scheme 7.1 Scheme 7.2 (e) S-BPM-HPM, (f) S-HPM-BPM ........................................... 147 (a) Monomers used in the synthesis of he MVE alternating copolymer reported here. (b) Deprotection (activation) chemistry for the MVE copolymer. In this scheme, the isopropyl groups are hydrolyzed subsequent to layer assembly ................................................. 167 Idealized schematic of the MVE polymer system showing two layers connected by a diester linkage ................................................ 169 -xvi- Chapter 1 Introduction The past few decades have seen tremendous research effort in the areas of materials design and assembly where the ultimate goal is molecular scale control over the macroscopic properties exhibited by the material. Several methods have been developed for the design and application of thin films chemistry in a variety of areas 3-6 such as chemical separation,“2 biological and chemical sensors, electrode 7-10 passivation, optical information storage, nonlinear optical materials, molecular ll-l4 recognition and studies of energy transfer. Among the more extensively employed methods are Langmuir-Blodgett films, self-assembled monolayers,'5‘2' spin casting/coating, covalent interlayer linkage strategies and metal phosphonate ”‘37 A number of these methods employ a layer-by-layer strategy for chemistry. material assembly, where this approach extends the potential applicability of these materials into a variety of areas and allows for the molecular control of material property. This level of control is due to the fact that it is possible to introduce specific chemical functionalities into each individual layer. The ability to control the identity of the functional groups incorporated into each layer renders accessible the ability to selectively tune material properties in a relatively well-defined manner — the major focus of the work we report in this dissertation. As mentioned above, a variety of methods have been developed for materials design and assembly. There are however, shortcomings associated with each of these methods that necessitate the continued development of and improvement upon the current technologies. One such system that has gained significant attention over the years has been Langmuir-Blodgett films, based on the work of Irving Langmuir and 8.39 . - - 3 These materials have been used in numerous studies, Katharine Blodgett. however, they suffer significantly from lack of robustness, thereby limiting their application particularly in harsh environments. In this approach, surface monolayers of amphiphilic molecules preassembled on a liquid surface are adsorbed onto a substrate as it is withdrawn from or introduced to the liquid subphase. The adsorption is made possible by the affinity of the head- or tail-groups of the amphiphile to the substrate or to themselves. The source of the problems associated with this method is the association of the molecular layers through weak van der Waals interactions that are subject to disruption, often by rinsing with polar solvents. While this method of layer formation is typically easy to employ and can be used to prepare a variety of useful systems where the weak non-selective interaction does not destroy the functionality of the species being adsorbed, the fact that the intermolecular interactions are so weak renders L—B films limited in their utility. A second method that has been studied extensively is alkanethiol-gold self- 4044 In this strategy, surface modification is assembly pioneered by Allara and Nuzzo. possible due to the affinity of molecules bearing a terminal thiol for gold and/or silver substrates. While this method has received much attention, and has been employed extensively in the design of a variety of materials, it too has proven to be of limited utility due to the weak association of the thiol group with the surface (energies of 45 .46 interaction AGQ‘cs ~ 20 kJ/mol), consistent with weak chemisorption. A second limitation of this method is that surface modification is often limited to one molecular layer due to the selective interaction of the head group with the substrate. Multilayer assembly is, however, possible with this type of chemistry if the terminal functionality of the alkanethiol is capable of participating in additional reactions.”5 ‘ An additional limitation of this chemistry is that thiol groups dictate the type of substrate (typically Au or Ag) that can be used. Despite these shortcomings, this method has found extensive application in systems where specific surface terminal groups are desired. Once the surface has been modified, it is possible to perform further chemistry on the surface that would have not been possible. For instance, using the alkanethiols to assemble a variety of systems on gold allows electrochemical characterization of these systems where the gold surface behaves as a working electrode. The metal phosphonate cherrristry, based on the pioneering work of the 7,3 . . 3, 9. and Thompson3 6525 5 60 groups, M all 0le34.252729.32.52.55 K at223.28.30.35.37,56—58 provides a route to prepare materials that are both thermally and chemically robust. In this method, a substrate that initially presents surface hydroxyl or silanol groups is primed by the initial exposure of those groups to a functionalized silane (typically an amino- terminated silane). The silane-modified interfaces are then exposed to phosphorus oxychloride and subsequently exposed to aqueous solutions of metal ions such as Hf“ or Zr“. It should be noted that, even in the absence of silane surface priming, it is possible to react phosphorus oxychloride directly with the surface 54.55 silanol groups with no apparent compromise in reactivity or surface morphology. The phosphate/zirconium ion terminated substrate is then exposed to a solution of an 0t,u)-bisphosphonate or an asymmetric (ii-hydroxyphosphonate, where the ionic interaction between the positively charged metal ion with the phosphate or phosphonate head group leads to layer deposition. Subsequent to the deposition of the initial layer, the unreacted phosphonate head groups are exposed to an aqueous solution of the metal ion followed by exposure to a bisphosphonate. In the case of the (o-hydroxyphosphonate, the hydroxyl group is reacted with phosphorus oxychloride, followed by exposure to the aqueous metal ion solution, followed by reaction with either am-bisphosphonate or (n-hydroxyphosphonate. In this manner, it is possible to assemble robust multilayer structures. While metal bisphosphonate/phosphate chemistry has been shown to be robust (undoubtedly due to the strong, selective interactions) one of the shortcomings of this chemistry is the lack of stability of the layers parallel to the plane of the substrate. In work performed by Kohli and Blanchard, a slightly modified version of the zirconium phosphonate (ZP) chemistry was demonstrated, where an alternating maleimide functionalized copolymer presenting phosphonate side-groups was used in material design/assembly.56 In this scheme, the phosphonate functionalities are initially protected with isopropyl groups, which are then removed by reaction with bromotrimethylsilane (BTMS), allowing for controlled layer-by-layer assembly. One of the major advantages of this growth methodology is enhanced structural stability afforded by the polymer backbone in the plane parallel to the substrate. Additionally, the maleimide chemistry we employ allows for the introduction of a variety of functional groups into the polymer layer. One limitation of metal phosphonate chemistry was revealed in work performed by Home and Blanchard, where interlayer excitation transport in materials assembled using zirconium bisphosphonate interlayer linkages was shown to be inefficient.57 Those authors have offered a possible explanation where the polarizable ZP interlayer screens dipolar coupling between layers. To this end, the development of robust chemical routes to multilayer structures that do not involve ionic interlayer connections might be useful in removing screening effects that serve to minimize interlayer energy migration. The majority of the work reported in this dissertation involves the use of a suite of reactions that have been developed for the controlled layer-by-layer assembly of materials where the interlayer linkages are all covalent in nature. It should be noted that this work is not a study of energy migration; it is worth noting however, that the synthetic routes reported herein are very likely amenable to the design of systems that exhibit this phenomenon. There are three major thrusts of the work reported in this dissertation. First, the development of a suite of unique strategies for surface modification is reported. In these methods, the uniqueness iies not only in the linkage strategies, but also the demonstrated ability to introduce a variety of functional moieties into the matrix of the polymer materials that can be used to tune the properties of the material selectively. Depending on the chromophore incorporated into the layered material, it is also possible to probe the microenvironment formed within these adlayers. Secondly, the characterization of these materials is reported. In addition to traditional methods of characterization such as optical ellipsometry, UV-visible, FTIR and NMR spectroscopies, an in-house assembled apparatus that combines the techniques of ellipsometry and quartz crystal microbalance (QCM) gravimetry was employed to acquire adsorption isotherm data for these systems. By this method, the adsorption of vapor phase adsorbate molecules is followed in real time and plots of adsorbate volume as a function of adsorbate vapor pressure at constant temperature generate the isotherm. From the isotherm data, it is possible to access information about the energies of interaction between various adsorbates and the interfaces that are prepared for this study. It is also possible to determine the volume of the monolayer, i.e. the volume of the adsorbate that is required to cover the interface with a single molecular layer. The incorporation of an ellipsometer into the apparatus allows the determination of thickness changes associated with the adsorption of the adsorbate molecules as a function of vapor pressure. Ellipsometric thickness in concert with adsorbate volume can be used to acquire information about the specific surface area of the adsorbent. In the final phase of this project, the application of one of these interfaces is demonstrated. In that body of work, 4-hydroxyphenylmaleimide was copolymerized with a vinyl ether phosphonate, where the phosphonate groups were protected with isopropyl groups. After the polymer has been assembled into the desired number of layers by crosslinking between the hydroxyphenylmaleimide functionalities in adjacent layers, the phosphonate groups were deprotected and this assembly was used in selective metal ion sequestration. Building upon the maleimide vinyl ether chemistry used in the Blanchard group, a new system for multilayer assembly was demonstrated. Here, a suite of copolymers was prepared using a variety of functionalized maleimide monomers, all prepared in-house. The maleimides were copolymerized (radical polymerization) with a variety of allyl—, allyloxy-, vinyl— and vinyloxy— functionalized compounds. We have chosen to work with maleimides due to the fact that they (1) are easy to prepare, (2) are very stable, and (3) have been shown to form alternating copolymers, a feature that allows for greater uniformity/control over material property. Additionally, the comonomers were chosen due to the variety of functional side- groups that can be introduced into the polymer, allowing us to realize multilayer assembly using a variety of interlayer crosslinking strategies, where these crosslinks are either ionic or covalent in nature. We focus here primarily on covalent crosslinking of these materials due to the extensive body of literature dealing with ionic interlayer linkages. In Chapters two and three of this dissertation, we report the synthesis (Chapter 2) and multilayer assembly (Chapter 3) of the various polymer systems. The major focus of this project is the ability to prepare materials where we can introduce various functional moieties into the matrix of the bulk material that can be used to tune or modify a macroscopic material selectively. We approach this task from two fronts. First, we design polymer systems that present functionalities capable of participating in a variety of interlayer linking strategies, where the crosslinking chemistries are all covalent in nature. Second, by simple substitution of the functional moiety of the maleimide, it is possible to introduce a wide variety of functional groups into the matrix of these polymers. A variety of methods have been used in the characterization of these materials. FTIR (KBr pellet) and lH-NMR spectroscopies have been employed in the determination of the structures of the polymers, and GPC was used to determine molecular weights. Subsequent to layer assembly, optical null ellipsometry was used to determine the thickness of the layers as a function of deposition cycle for adlayers assembled on either gold-coated or silicon wafers. UV- visible and FTIR spectroscopies were employed to confirm material deposition and to verify the identity of the cross-linking chemistry. In Chapter 4 we report the acid enhanced deposition of polymer materials. In this work, we have demonstrated the ability to control the loading density of the chromophore - as evidenced by the absorption spectra and ellipsometry - by simple acid catalysis and/or dehydration. During layer deposition, concentrated hydrochloric or concentrated sulfuric acid is added to the reaction vessel and in both cases we observe significant enhancement in material deposition relative to the corresponding acid—free reactions. In all cases, we observed an increase in ellipsometric thickness for the layers as well as an increase in the intensity of the absorption band of the chromophore in the uv-visible spectra. We have explained these findings in the context of acid catalysis and/or dehydration. It should be noted that, in all cases reported, regardless of the chemistry being used for multilayer assembly, the addition of sulfuric acid led to a higher polymer adlayer loading density. We understand these findings based on the ability of sulfuric acid to function as both an acid and an efficient dehydrating agent. In the cases of amide and ester linkage formation, these reactions proceed with the elimination of water as a by-product. The addition of hydrochloric and sulfuric acid probably serves to protonate the reactive species, while sulfuric acid is simultaneously capable of sequestering product water, driving the reaction toward product formation according to Le Chatelier’s principle. In the cases of urea and urethane linkage reactions, the concentrated acid again serves to protonate the reactive species. It should be noted that the reactive intermediate in these reactions is an unstable carbamic acid that is susceptible to attack by adventitious water, leading to the degradation of this species. In the case where sulfuric acid is added, dehydration can thus play a significant role. Finally, we note the dependence of polymer adlayer density on acid concentration. We found that increasing the acid concentration eventually led to a decrease in the slopes of the absorbance versus concentration curves, suggesting that there is an optimal range over which these reactions are most efficient. We note, however, that increasing the concentration led to a linear increase in absorbance for the initial layer deposited. The ability to increase the loading density of the functional species with the material matrix could prove beneficial in sensor design and the design of conductive or resistive materials, for example. The main focus of Chapters 5 and 6 is the characterization of these materials with respect to their adsorption isotherm. From the adsorption isotherm data on the various materials, we have been able to demonstrate that, by simply varying the functional groups within the matrix of the material, we can control the adsorption energetics of the material. In Chapter 5, we focus on the deposition of single polymer layers with different functional groups and acquire their vapor phase isotherms. In Chapter 6, we have prepared the interfaces using multilayers of polymers, where we have varied the identity of the functional groups within each layer and report the observe changes in the isotherm response relative to the single layered systems of Chapter 5. In this work, we recover noticeable changes in the functional form of the isotherms as well as in the energetics of the systems when different groups are substituted into the matrix. Another very interesting observation is that when the order of polymer deposition is reversed, even in these two—layered systems, we recover substantially different adsorption isotherm data, indicating our ability to control the kinetic and thermodynamic properties of the system by synthetic means. We have used semi-empirical calculations to estimate the dipole moment of the various side-groups we employ in this study and have attempted to rationalize our findings in the context of adlayer polarity. The ability to control functional group identity within individual layers lends itself directly to the preparation of interfaces that exhibit a chemical potential gradient, where these interfaces can be used in separation schemes capable of continuous extraction. It should also be noted that, in all cases, we observed no thickness change for our polymer adlayers with adsorption, suggesting the porous nature of the interfaces. In previous work performed by Karpovich and Blanchard,58 it was shown that this system was sensitive enough to determine adlayer thickness changes as a function of partial pressure of the vapor phase adsorbate molecules. In that work, the adlayers were relatively well-ordered alkanethiol monolayers. Here, we use alternating copolymers, where these systems, by their very nature, are less well ordered and more porous. In Chapter 7 we describe a specific application of one particular polymer system we have prepared. In this work, we employed a hydroxyphenylmaleimide and copolymerized this group with a vinyl ether phosphonate. Multilayer assembly was made possible by generating ester linkages across the hydroxyphenylmaleimide and using the phosphonate groups for metal ion sequestration. It should be noted that during layer deposition, the phosphonate groups remain protected, precluding the -10- formation of phosphoesters. In a second step, these groups are deprotected using 59 We have employed a variety of analytical methods to bromotrimethylsilane. characterize these systems. X-ray photoelectron spectroscopy (XPS) data confirm the sequestration of various metal ions by these systems. Using this polymer and the various characterization methods, we have been able to demonstrate the selective and in some case irreversible sequestration of metal ions. There are three distinct advantages of the materials and assembly strategies reported in this dissertation. First, as mentioned above, the simplicity of the maleimide chemistry employed allows the introduction of a variety of chemical functional side-groups into the matrix of these materials. This ability will be exploited to develop materials that exhibit a chemical potential gradient. According to Giddings,60 the creation of a gradient in any particular property of a material can be used to create a chemical potential gradient. In this work, we generate a gradient in the dipole moments of the side-groups incorporated into each individual layer. Second, the side groups of the comonomers allow us to use very simple covalent interlayer linking strategies such as amide, ester, ether, urea and urethane formation to assemble these materials into robust multilayer structures. The third advantage is a direct result of the covalent cross-linking chemistry. We have been able to demonstrate that, through simple acid catalysis and/or dehydration of the cross— linking reactions, we are able to control material deposition. This feature can prove useful in a variety of applications such as sensor design, where increasing the number density of he sensing elements could lead to enhanced detection. -11- 1.2 Literature Cited 1. Vrancken, K. C.; VanderVoort, P.; Gillisdhamers, I.; VanSant, E. F.; Grobet, P.; J. Chem. Soc, Faraday Trans. 1992, 88, 3197. 2. Pfleiderer, 13.; Albert, K.; Bayer, 13.; J. Chromatogr. 1990, 506, 343. 3. Demas, J. N.; Degraff, B. A.; Coleman, P. 8.; Anal. Chem, News & Features, 1999, 793A. 4. Yoon, H. C.; Kim, H-S. Anal. Chem. 2000, 72, 922. 5. Delamarche, E.; Sundarababu, G.; Biebuyck, H.; Michel, B.; Gerber, C.; Sirgrist, H.; Wolf, H.; Ringdorf, H.; Xanthopoulos, N.; Mathieu, H. J .; Langmuir, 1996, 12, 1997. 6. Brousseau, HI, L.C.; Aurentz, D. J.; Benesi, A. J .; Mallouk, T. E.; Anal. Chem, 1997, 69, 688. 7. Bakiamoh, S.; Blanchard, G. J.; Langmuir, 2001, 17(11), 3438. 8. Katz, H. E.; Scheller, G.; Putvinski, T. M.; Schilling, M. L.; Wilson, W. L.; Chidsey, C. E. D. Science, 1991, 254, 1485. 9. Li, D.; Ratner, M. A.; Marks, T. J.; Zhang, C. H.; Yang, J.; Wong, G. K.; J. Am. Chem. 506., 1990, 112, 7389. 10. Kanis, D. R.; Ratner, M. A.; Marks, T. J. Chem. Rev. 1994, 94, 195. 11. Zak, J .; Yuan, H.; Ho, M.; Woo, L. K.; Porter, M. D.; Langmuir 1993, 9, 2772. 12. Finklea, H. O.; Hanshew, D. D. J. Am. Chem. Soc. 1992, 114, 3173. 13. Sun, L.; Crooks, R. M. Langmuir 1993, 9, 1775. 14. Sun, L.; Kepley, L. J.; Crooks, R. M. Langmuir 1992, 8, 2101. 15. Bain, C. D.; Whitesides, G. M. J. Am. Chem. Soc. 1988, 110, 3665-3666. -12- 16. Bain, C. D.; Biebuyck, H. A.; Whitesides, G. M. Langmuir 1989, 5, 723. 17. Nuzzo, R. G.; Allara, D. L. J. Am. Chem. Soc. 1983, 105, 4481. 18. Nuzzo, R. G.; Dubois, L. H.; Allara, D. L. J. Am. Chem. Soc. 1990, 112, 558. 19. Nuzzo, R. G.; Fusco, F. A.; Allara, D. L. J. Am. Chem. Soc. 1987, 109, 2358. 20. Xia, Y.; Whitesides, G. M. Angew. Chem, Int. Ed. 1998, 37, 550. 21. Troughton, E. B.; Bain, C. D.; Whitesides, G. M.; Nuzzo, R. G.; Allra, D. L.; Porter, M. D. Langmuir 1988, 4, 365. 22. Hong, H.-G.; Sackett, D. D.; Mallouk, T. E. Chem. Mater. 1991, 3, 521. 23. Katz, H. E. Chem. Mater. 1994, 6, 2227. 24. Lee, H.; Kepley, L. J .; Hong, H.-G.; Akhter, S.; Mallouk, T. E. J. Phys. Chem. 1988, 92, 2597. 25. Lee, H.; Kepley, L. J .; Hong, H.-G.; Mallouk, T. E. J. Am. Chem. Soc. 1988, 110, 618. 26. Putvinski, T. M.; Schilling, M. L.; Katz, H. E.; Chidsey, C. E. D.; Mujsce, A. M.; Emerson, A. B. Langmuir 1990, 97, 237. 27. Rong, D.; Hong, H.-G.; Kim, Y.-I.; Krueger, J. S.; Mayer, J. E.; Mallouk, T. E. Coord. Chem. Rev. 1990, 97, 237. 28. Katz, H. E.; Schilling, M. L.; Chidsey, E. E. D.; Putvinski, T. M.; Hutton, R. S. Chem. Mater. 1991, 2, 699. 29. C30, G.; Rabenberg, L. K.; Nunn, C. M.; Mallouk, T. M. Chem. Mater. 1991, 3, 149. 30. Ungashe, S. B.; Wilson, W. L.; Katz, H. E.; Scheller, G. R.; Putvinski, T. M. J. Am. Chem. Soc, 1992, 114, 8717. -13- _ —-.' r i a 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. Vermeulen, L.; Thompson, M. E. Nature 1992, 358, 656. Yang, H. C.; Aoki, K.; Hong, H.-G.; Sackett, D. D.; Arendt, M. F.; Yau, S.-L.; Bell, C. M.; Mallouk, T. E. J. Am. Chem. Soc. 1993, 115, 11855. Frey, B. L.; Hanken, D. G.; Corn, R. M. Langmuir, 1993, 9, 1815. Yonemoto, E. H.; Saupe, G. B.; Schmehl, R. H.; Hubig, S. M.; Riley, R. L.; Iverson, B. L.; Mallouk, T. E. J. Am. Chem. Soc. 1994, 116, 4786. Katz, H. E.; Bent, S. E; Wilson, W. L.; Schilling, M. L.; Ungashe, S. B. J. Am. Chem. Soc. 1994, 116, 6631. Thompson, M. E. Chem. Mater. 1994, 6, 1168. Katz, H. E.; Wilson, W. L.; Scheller, G. J. Am. Chem. Soc. 1994, 116, 6636. Langmuir, 1.; J. Am. Chem. Soc. 1917, 39, 1848. Blodgett, K. B.; J. Am. Chem. Soc. 1935, 57, 1007. Allara, D. L.; Nuzzo, R. G. Langmuir 1985, I, 45. Allara, D. L.; Nuzzo, R. G. Langmuir 1985, 1, 52. Nuzzo, R. G.; Zegarski, B. R.; Dubois, L. H. J. Am. Chem. Soc 1987, 109, 733. Bain, C. D.; Whitesides, G. M. Science 1988, 240, 62. Strong, L; Whitesides, G. M. Langmuir 1988, 4, 546. Karpovich, D. S.; Blanchard, G. J. Langmuir 1994, 10, 3315. Schessler, H. M.; Karpovich, D. S.; Blanchard, G. J. J. Am. Chem. Soc. 1996, 118, 9645. Sabapathy, R. C.; Crooks, R. M. Langmuir 2000, 16, 1777. Chechik, V.; Crooks, R. M. Langmuir 1999, 15, 6364. Lackowsi, W. M.; Franchina, J. G.; Bergbrieiter, D. E.; Crook, R. M. Adv. Mater. -14- 1999, 11, 1368. 50. Bruening, M. L.; Zhou, Y.; Aguilar, G.; Agee, R.; Bergbreiter, D. E.; Crooks, R. M. Langmuir, 1997, 13, 770. 51. Zhou, Y.; Bruening, M. L.; Bergbreiter, D. E.; Crook, R. M.; Wells, M J. Am. Chem. Soc. 1996, 118, 3773. 52. Vermeulen, L. A.; Snover, J. L.; Sapochak, L. S.; Thompson, M. E. J. Am. Chem. Soc. 1993, 115, 11767. 53. Snover, J. L.; Byrd, H.; Suponeva, E. P.; Vicenzi, E.; thompson, M. E. Chem. Mater. 1996, 8, 1490. 54. Putvinski, T. M.; Schilling, M. L.; Katz, H. E.; Chidsey, C. E D.; Mujsce, A. M.; Emerson, A B. Langmuir, 1990, 6, 1567. 55. Bakiamoh, S. B.; Blanchard, G. J. Langmuir 1999, 15, 6379. 56. Kohli, P.; Blanchard, G. J. Langmuir, 1999, 15, 1418. 57. Home, J. C.; Blanchard, G. J. J. Am. Chem. Soc. 1999, 121, 4427. 58. Karpovich, D. S.; Blanchard, G. J. Langmuir, 1997, 59. Major, J. S.; Blanchard, G. J. Langmuir 2001, I7, 1163. 60. Giddings, C. J. Unified Separation Science; Wiley Interscience, 1991. -15- Chapter 2 Strategies for Covalent Multilayer Growth 1. Polymer Design and Characterization Abstract We describe the synthesis and characterization of several maleimide-vinyl ether and maleimide-primary alkene alternating co-polymers capable of being incorporated into covalently-bound multilayer assemblies. Synthetic control over the monomer substituents allows substantial versatility in the choice of layer formation chemistry and in polymer properties. We report on the co—polymerization of substituted allyl-, allyloxy- vinyl- and vinyloxy- monomers with a variety of N-substituted maleimides. The substituted maleimides were prepared by reaction of maleic anhydride with selected aromatic amines. Polymerization yields are typically in the range of 75 to 95% consumption of the monomers. lH-NMR, FT IR and UV-visible spectroscopies were used to characterize the polymers. The materials we report in this paper are used to create a family of covalently-bound multilayer structures, and we report these findings in chapter 3 of this dissertation. -l6- 2.1 Introduction A major effort in the materials community lies in the design, construction and characterization of novel interfacial thin films, with the goal being the ability to control the resultant macroscopic interface properties. Such materials have been designed and demonstrated for specific target applications, including chemical separations,"2 nonlinear optics,3'5 chemical and biological sensing,6'9 and heavy metal analysis and remediation.” ‘2 These interfacial materials have been applied to surfaces by spin coating, Langmuir- 13-16 Blodgett film formation, monolayer self-assembly (SAMs) and metal phosphonate ”’19 To extend the utility of interfacial chemistry into new areas and to chemistry. improve upon current technology, the development of a range of chemical approaches that are easy to implement and that result in chemically and physically robust materials will be required. We report here on the synthesis and characterization of several alternating copolymers that can be used for layered interface growth. The chemical identity of the side groups of these alternating copolymers renders them capable of discrete layer growth, where the interlayer linkages are covalent. In addition to having control over the interlayer linking chemistry, we can also demonstrate control over the resulting properties of the interface through the judicious choice of maleimide monomer functionality. The primary point of the work we present here is the demonstration of two distinct levels of synthetic and structural control we exercise over the properties of alternating copolymers. We design alternating copolymers with specific pendant side groups to gain explicit control over the way in which individual polymer layers can be -17- assembled or cross-linked into multilayers. The functionality incorporated into the polymers allows control over the macroscopic properties of the resulting interface. A specific aim of this work is to report the synthesis of polymers capable of covalent interlayer attachment. Earlier work has shown that metal phosphonate interlayer linkages limited excitation transport because of the polarizable nature of the inorganic “sheets” that connected the individual layer constituents.20 While this property may be advantageous for certain applications, it can be a liability for others. To circumvent this issue, we have devised a means of forming well-controlled multilayer assemblies where the interlayer linking chemistry is covalent in nature. Because of the number of different functionalities we use in achieving this goal, we report in this paper on the synthesis and characterization of the polymers that comprise the individual layers. This paper is not intended to be a comprehensive review of covalent multilayer assembly, nor is it a study 2' The work we present here is intended to of energy transfer in confined environments. demonstrate the structural control and versatility of selected alternating copolymer systems. We report the synthesis and characterization of a family of alternating copolymers that will be used for controlled multilayer assembly.22 The chemistry we use to synthesize the polymers is based on established maleimide-vinyl ether alternating copolymerization; we demonstrate here that vinyl and vinyl ether monomers can be used with equal success. These polymers are robust and afford substantial control over their properties through the identity of the monomer substituents. We use vinyl and vinyl ether monomers possessing terminal functionalities capable of participating in amide, ester, ether, urea and urethane formation chemistry. We report the assembly of these -13- polymers into discrete multilayer structures in the following chapter.22 A portion of this work is dedicated to the synthesis of substituted maleimide monomers because these moieties influence the properties of the bulk polymer significantly. 2.2 Experimental We have designed two categories of monomers and polymers that use and build upon the maleimide vinyl ether polymerization chemistry reported previously by our -2 group.23 5 Syntheses: Alternating copolymers were prepared by radical polymerization of the various monomer units as indicated in Scheme 2.1. In the first group of syntheses, substituted maleimide monomer units are copolymerized with a variety of vinyl and vinyl ether co-monomers possessing functionalities capable of participating in several different interlayer linkage schemes.22 In the second group of syntheses, the functional groups of the maleimide moiety are capable of participating in interlayer linkage chemistry. These substituted maleimides are copolymerized with the ethylvinylether-Z— diisopropylphosphonate (VEP) to form a polymer that is also capable of participating in ionic interlayer linking chemistry. For this type of multilayer growth, control over the extent of ionic bonding within these materials is determined by stoichiometric 26’” and can potentially be used to mediate (ionic) conductivity in deprotection chemistry these materials. We have shown that the presence of the VEP groups can be used to sequester metal ions, extending the utility of these materials further. Maleimide monomer preparation: The maleirrride monomers were prepared by reacting maleic anhydride with a substituted aniline. This reaction results in the -19- formation of the corresponding maleamic acid with typical yield of >95% for all monomers reported here (Scheme 2.1). Cyclization of the maleamic acid is accomplished by reaction with sodium acetate in acetic anhydride at 70° C for 2-3 hours (Scheme 2.1). Typical yield for the ring closing reaction is greater than 75%. After ring closure, the maleimide monomers are collected from ice-cold water with stirring followed by filtration and rinsed with cold water. Ring closure was confirmed by the absence of a carboxylic acid proton in the 1H-NMR spectrum. The resulting substituted maleimide is used in copolymer preparation. Substituted vinyl and vinyl ether monomer were available commercially (Aldrich), except ethyl vinyl ether diisopropylphosphonate which was prepared as described elsewhere.24 0 o o R R R ..+..@_. / @ ——-. me 011 o o 0 Scheme 2.1. Reaction of maleic anhydride with an aromatic amine to produce the maleamic acid, followed by ring closure to the maleimide. Copolymerization of the monomer units (Scheme 2.2) was performed according to the following procedures unless otherwise noted. The constituent monomers were dissolved in chloroform and 1-25 mol% of the radical initiator 2,2’-azobisisobutyronitrile (AIBN) was added to the solution. The reaction mixtures were refluxed with stirring at 65°C for 2 - 24 hours, as determined by observing a change in solution viscosity or, in the -20- case of the N-phenylmaleimide acrylic acid polymer, a white suspension/precipitate was obvious after about two hours and yield was ~95%. Reactions were carried out under an argon atmosphere. Upon completion of the reaction, excess solvent was removed by rotary evaporation and the resulting solution was diluted with ether or hexanes and stirred for ~30 minutes to separate residual monomer from the polymer product. It should be noted that collection of the product from hexanes gives significantly higher yields in all cases. The resulting polymer was collected by filtration, air-dried and characterized by UV-visible, FIIR, and 'H-NMR. R\ o fifiiif:i>htt +' Qi? -AUU1bL> n o N o o N o I \R. I R R R! A + \ AIBN. o o o o n i‘ i R R' R Scheme 2.2. Synthetic route for the polymerization of maleimide and vinyl ether monomers (top), and for maleimide and vinyl monomers (bottom). Both reactions are initiated with AIBN. -21- Purification of the polymers was achieved by re-dissolving the product in a small amount of chloroform and collection from ether or hexane once or twice as necessary. Poly(N-phenylmaliemide-I -vinyloxy-4-butanol): Poly(N-phenylmaliemide- 1 - vinyloxy-4-butanol), (poly(NPM-VOB), (Figure 2.1a) was prepared by reacting equimolar amounts of N-phenylmaleimide with 1-vinyloxy-4-butanol in the presence of the initiator. Typical yield for this polymerization is ~80%. Purification of the product was accomplished by dissolution in chloroform and collection from ether or hexane. Poly(N~phenylmaleimide-3-aminopropyl-1 -vinyl ether): Poly(N- phenylmaleimide-3-aminopropyl-l-vinyl ether), (poly(NPM-AVPE), (Figure 2.1b) was prepared by reacting equimolar amounts of N-phenylmaleimide and 3-aminopropyl-l- vinyl ether in chloroform in the presence of AIBN. Reaction procedures and conditions are the same as reported above. Typical yield for this polymerization is ~75%. Poly(N-phenylmaleimide-4-pentenoyl chloride): Poly(N—phenylmaleimide-4- pentenoyl chloride), (poly(NPM-4PC), (Figure 2.1c) was prepared by reacting equimolar amounts of N-phenylmaleimide and 4-pentenoyl chloride in chloroform in the presence of AIBN. Reaction procedures and conditions are the same as reported above. Typical yield for this polymerization is 50% with 10 mol% AIBN initiator and > 85% with 20 mol% AIBN. Here again, polymer purification was accomplished by re-dissolving the product in a small amount of chloroform and collecting from ether or hexane. The polymer possesses acid or acid chloride side groups and the spectroscopic data do not show evidence for hydrolysis of the acid chloride group even after air-drying the product for extended periods. -22- Poly(4-nitr0phenylma1eimide-allylisocyanate ): Poly(4-nitrophenylmaleimide- allylisocyanate), (poly(4NAI), (Figure 2.1d) was prepared by reacting equimolar amounts of 4-nitrophenylmaleimide and allylisocyanate monomers in chloroform in the presence of AIBN initiator. Reaction conditions and procedures are as reported above, except the reaction time required for this polymerization was 12-18 hours. Typical yield for this polymerization is ~80%. Poly(3-chlorophenylmaleimde and 5-hexe-2-0ne): Poly(3—chlorophenylmaleimde and 5-hexe-2-one), (poly(3CPM-hex), (Figure 2.1e) was prepared by reacting equimolar amounts of 3-chlorophenylmaleimde and 5—hexe-2-one in chloroform in the presence of AIBN (20 mol%). Reaction conditions were the same as those reported above. Typical yield for this polymerization is ~50%. Pol y( I -maleimid0pyrene and 3 -aminopr0pyl -1 -vinylether): Poly( 1 - maleimidopyrene and 3-aminopropyl—1-vinyl ether), (poly(MP-APVE), (Figure 2.11) was prepared by dissolving equimolar amounts of both monomers in chloroform using AIBN as initiator (10 mol%). Reaction conditions were the same as reported above. Typical yield for this polymerization is 90%. Poly(4-hydroxyphenylmaleimide-vinyl ether diisopropylphosphonate): Poly(4- hydroxyphenylmaleimide-vinyl ether diisopropylphosphonate), (poly(4—I-IPM-VEP), (Figure 2.1g) was prepared by reaction of equimolar amounts of 4- hydroxyphenylmaleimide and vinyl ether diisopropylphosphonate in chloroform, with the addition of AIBN initiator (10 mol%). Typical yield for this polymerization is 75%. -23- O 93 O D g:— /\/\O= c1 N02 Figure 2.1. Polymer structures. (a) Poly(NPM-VOB), (b) Poly(NPM-APVE), (c) Poly(NPM-4-PC), (d) Poly(NAI) -24- NHz OH - VE , Figure 2 1 cont’d. Polymer structures. (e) Poly(3CPM-hex), (f) Poly(MP AP ) (g) Poly(4-HPM-VEP) -25- The reactive functionalities introduced into these alternating copolymers makes the resulting materials amenable to a variety of deposition strategies where interlayer linking chemistry can be accomplished by the formation of amide, ester, ether, urea or urethane functionalities. The functionalized maleimide monomers employed in this work were prepared in-house as reported previously,28 except N-phenylmaleimide, which is available commercially, and was used after recrystallization from absolute ethanol. The reaction schemes for substrate preparation and multilayer assembly are presented and discussed more extensively in chapter 3.22 2.3 Results and Discussion The primary goal of this paper is to report the synthesis and characterization of several alternating copolymers that can be used as layer constituents in the construction of discretely layered interfacial structures. In this work, the monomer species we use are capable of forming covalent interlayer linkages and we report on that aspect of the work in the following chapter.22 We have synthesized and characterized an extensive library of these compounds and report on a representative selection of seven alternating copolymers here. The structural versatility of the polymers we report here allows us to achieve control over the identity of the reactive polymer side group used in covalent interlayer linkage formation. We also have control over polymer properties by maleimide substitution to incorporate a variety of functional groups that can be used to determine the properties of the polymer matrix. The alternating copolymer structural motif of these materials removes the -26- substantial structural variability associated with random block copolymers and thus affords control over the properties of the polymer and resulting interface.24 We discuss the characterization of each polymer next. For all of the polymers, lH-—NMR and FF IR show no evidence of unreacted C=C-H groups, indicating that separation of the monomer reactants from the polymer product was essentially complete. The IR data exhibit no hands for the vinyl ether C=C stretch in the 1660-1610 cm'1 region nor any =C-H wagging resonance in the 900 cm'1 to 1000 cm'l region. Previous GPC results for polymers prepared under similar conditions indicated molecular weights of approximately M, = 10,800. Poly(NPM—VOB): The infrared spectrum of poly(NPM-VOB) contains several characteristic bands that are useful to the structural characterization of this polymer (Figure 2.2). First are the characteristic aromatic methylene stretches associated with the N-phenyl ring between 3000 and 3100 cm'l. The aliphatic CH2 stretches between 2850 cm’1 and 2950 cm'1 are associated with the butyl group and the band centered around 3475 cm"1 is characteristic of hydroxyl stretches. The width of this band is indicative of intermolecular hydrogen bonding within this polymer.29 There are two carbonyl stretches, the first centered at 1775 cm‘l and the second at ~1710 cm". This spectral signature is characteristic of the presence of the maleimide groups and is evident in all of the polymers we report here. The band at 1775 cm'1 is associated with the symmetric C=O stretch and the band at 1710 cm"1 is associated with the asymmetric C=O stretch in the succinimide moiety. We also observe the characteristic C=C stretching modes of the phenyl ring at 1597 cm", 1500 cm], and1457 cm'1 and the out-of—plane ring bending at 692 cm".29 -27- absorbance (a.u.) WW I A a l 4000 3500 3000 2500 2000 1500 1000 500 frequency (cm'l) Figure 2.2. FTIR spectrum (KBr pellet) of poly(NPM-VOB). The band centered at 1387 cm‘1 has been assigned to the C-N bond of the tertiary aromatic amine (aniline).29 This band is centered at 1387 cm'1 rather than the expected 1360 cm'1 due to resonance stabilization of the amine group with the phenyl ring as well as the presence of the succinimide carbonyl functionalities. The 1189 cm'I band has been assigned to the maleimide C-N-C stretch.30 Because C—O-C stretches also occur in this region, it is possible that the intensity of this band is the result of overlap of the C-N-C -28- and C-O-C vibrations. This is likely the case based on the noticeably higher intensity of this band in the vinyloxy- copolymers. The bands at 1073 cm'1 and 1090 cm'1 are also associated with the C-O-C stretch. The bands at 755 and 734 cm'1 have been assigned to the C-N symmetric stretching modes. lH-NMR spectroscopy has also been useful in characterizing the polymers we report here. An important indicator of the polymerization process is the disappearance of C=C-H bands of both monomers (5 = 7.1 ppm for the maleimides and 5:62, 3.95 and 4.05 ppm for the protons adjacent to the ether oxygen and the two terminal protons of 1- vinyloxy-4-butanol, respectively), consistent with essentially complete separation of the polymer product from the monomers (Figure 2.3). We see no evidence of homopolymerization of any monomers. Extensive overlap and broadened peaks in some regions of the spectra makes unequivocal peak assignment impossible, particularly for aliphatic methylene groups. We observe the protons of the NPM aromatic ring at 8 = 7.2 - 7.6 ppm. The butyl protons adjacent to the hydroxyl oxygen are found at 5 = 4.2 - 4.6 ppm and both the aliphatic protons of the butyl group and the bridging protons between the maleimide and the vinyloxy- groups are seen as a broad resonance between 5 = 1.0 and 1.6 ppm. The succinimide protons are seen at 5 = 3.3 ppm and the multiplet feature at 3.6 ppm is assigned to the methine protons adjacent to the ether oxygen. The alcoholic proton is not seen in the spectrum because of facile proton exchange. -29- - MLLAC YTYTTYTTTfiII YfiT‘IfTrTYYYYTY'WYIYTYYIY‘TYYII I "IffiVTY‘ Y‘IV VTWIV I 11Tfi ”1511109 8 7 6 5 4 3 210 - - chemical shift (Dom) Figure 2.3. 1H-NMR spectrum of poly(NPM-VOB). -30.. Poly(NPM-APVE): The FTIR spectrum of this polymer reveals several characteristic bands in addition to those seen for all of the polymers we report here, as noted above (Figure 2.4). The band centered at 3408 cm“1 has been assigned to the associated NH; stretches of the amino group. We also see the NH bending mode resonance at 1648 cm". absorbance (a.u.) A 4000 3500 3000 2500 2000 1500 1000 500 frequency (cm'l) Figure 2.4. FTIR spectrum (KBr pellet) of poly(NPM-APVE). -31- The 1H-NMR spectrum of poly(NPM-AVPE) (Figure 2.5) contains the expected features such as the OH resonances associated with the phenyl ring between 5 = 7.2-7.6 ppm. The aliphatic methylene bands are contained in the broad resonance between 5 = 1.2-1.8 ppm, succinirrride protons are found at 5 = 3.2 ppm, and the propyl methylene protons adjacent to the ether oxygen are at 5 = 4.45 ppm. The proton exchange rate for the amino proton precludes observation of this resonance, in analogy to the hydroxyl proton resonance for poly(NPM-VOB). VV '1‘" V v' r—— r chemical shift (ppm) Figure 2.5. 1H-NMR spectrum of poly(NPM-APVE). -32- Poly(NPM-4-PC): Much of the vibrational and electronic spectral signature of poly(NPM-4-PC) is the same as that of the other copolymers we report here (Figure 2.6a). We also see the characteristic aliphatic methylene stretches associated with the pentenyl- portion of the polymer. There are several additional features associated with the acid chloride moiety in this polymer. For poly(NPM-4-PC), the 1189 cm'1 band is weaker than that seen for poly(NPM-AVPE) because there are only the C-N-C stretches for poly(NPM-4-PC). The absence of contributions from the vinyloxy C-O-C band accounts for the attenuation of this feature. For poly(NPM-4-PC) there is a resonance at ~3450 cm", suggesting intermolecular hydrogen bonding of the hydrolyzed acid chloride functionality. The carbonyl stretching region is characterized by bands associated with the succinimide, the acid chloride and some carboxylic acid (hydrolyzed acid chloride). The 1H-NMR spectrum of poly(NPM-4-PC) (Figure 2.6b) shows no evidence of hydrolysis of the acid chloride to the acid even after extensive periods of atmospheric exposure, and this finding is not consistent, at face value with the IR spectra. These two pieces of information suggest that the FI‘IR (KBr crystals) sample may have been wet. Aromatic, aliphatic and succinimide protons occur in the same region as seen for the other polymers. -33- absorbance (a.u.) W 4000 3500 3000 2500 2000 1500 41% 4 1000 500 frequency (cm'l) wait. Y ‘Tj f1 T Ti 1' 1' I 1 V V Y Y ‘ V V I I Y ‘ V V Y Y 12"‘1r“'m”"9” 8 7 6 5 4....3....,2...1 chemical shift Figure 2.6. (a) IR spectrum of poly(NPM-4-PC). (b) 1H-NMR spectrum of poly(NPM- 4-PC). -—4 -34- Poly(4-NAI): The IR band of the isocyanate group at 2265 cm'1 region is characteristic of this polymer (Figure 2.7a). We also observe the characteristic aromatic and aliphatic methylene stretches and the succinimide C=O stretches at 1775 and 1720 cm']. In addition, we observe the asymmetric ArNOz stretch at 1523 cm'1 and the symmetric ArNOz stretch at 1347 cm'l. We note the lower frequencies of the phenyl ring as well as the splitting of some of these bands into doublets due to ring substitution. We also see an intensity decrease in the C-N-C stretching region due to the absence of the C- O-C functionality in this polymer. The 854 cm'1 band is associated with the aromatic C- N stretch of the nitro group. We observe a symmetric splitting of the aromatic protons peak in the 1H-NMR spectrum of poly(4-NPM—AI) which is expected due to the presence of the N02 group in the 4—position. Para- substitution of the phenylmaleimide leads to equivalence of the two protons adjacent to the substituent leading to one peak for both protons (Figure 2.7b). The resonance for the methylene group protons adjacent to the isocyanate group is in the same region as the bridging methine and succinimide protons, 5 = 3.15 - 3.5 ppm. -35- absorbance (a.u.) l n l 1 l 4 l n 4000 3500 3000 2500 2000 1500 1000 500 frequency (cm'l) 10 9 8 7 4 3 2 1 chemical shift (ppm) Figure 2.7. (a) IR spectrum of poly(4NAI), (b) lH-NMR of poly(4NAl). —36- Poly(3CPM-hex): For this polymer, the IR band centered at 3475 cm'1 is likely an overtone of the carbonyl band (Figure 2.8). We make this assignment because of the absence of amino or hydroxyl functionality in this polymer. We observe characteristic aromatic and aliphatic methylene stretches between 2850 and 3100 cm‘I and maleirrride C=O stretches are seen at 1780 cm'1 and 1715 cm'l. The broadened carbonyl band at 1715 cm'1 is the result of spectral overlap between the maleimide C=O stretch and the ketone C=O stretch. Aromatic C=C stretches are at slightly lower frequencies than for N- phenylmaleimide due to the presence of the Cl— substitution at the 3- position. The C-N stretch of the tertiary amine is seen at 1382 cm'1 for poly(3CPM-hex). absorbance (a.u.) ti 1 l L l 4000 3500 3000 2500 2000 1500 1000 500 frequency (cm'l) Figure 2.8. IR spectrum of poly(3CPM-hex). -37- lH-NMR spectroscopy of poly(3CPM-hex) (Figure 2.9) shows the expected aromatic protons at 5 = 7.2 - 7 .6 ppm. The aliphatic methylene resonances are seen at 5 = 1.5 ppm with the methyl protons adjacent to the C20 functionality occurring at 5 = 2.05 ppm. The succinimide resonances are found at 5 = 3.3 ppm, consistent with the NMR spectra of the other polymers we report here. chemical shift (ppm) Figure 2.9. ‘H-NMR spectrum of poly(3CPM-hex). Poly(MP-APVE): The IR spectrum of this polymer contains several features unique to this polymer (Figure 2.10a). We observe the -NH2 stretch at 3277 cm]. The position of this band suggests that the NH; groups exist primarily in the associated form. -33- We note the presence of a high-frequency shoulder on the -NH2 band, indicating some free -NH2 groups. The aromatic C-H stretches are consistent with the pyrenyl ring structure. We also observe the characteristic maleimide C=O bands and a broad band at 1632 cm], which we assign to the N-H bending mode(s). This band could also be present due to some amount of maleirrride ring opening during polymerization. We observe a relatively strong band at 1194 cm'1 that we attribute to the presence of both the CDC and C-N-C functionalities in the polymer. The intense band at 844 cm'1 has been assigned to the out-of—plane C-H bending modes of the polynuclear aromatic moiety. Examination of the 1H-NMR spectrum (Figure 2.10b) reveals a complex spectral signature for the aromatic proton resonances at 5 = 8.0 - 8.8 ppm, consistent with the ring structure of the pyrene moiety. We observe the methylene and succinirrride protons at the same chemical shifts as seen for the other polymers we report here. There is a small resonance at 5 = 10.8 ppm, suggesting some succinimide ring opening occurring either as the result of incomplete ring closure of the monomer or as a result of the polymerization reaction. Given the integration of this band for 0.3 protons, it appears that ring opening, if it is proceeding, is not a dominant process. As a point of reference, the 1H NMR spectrum of the pyrene maleimide monomer shows no evidence of incomplete ring closure. -39_ a Of?- 013- d 015- d 011- Absorbance OJB— i 0.2 - fi OJ — 1 0.0 i 1 4000 3500 I I’ I I ' I - I ' I I I 3000 2500 2000 1500 1000 500 Frequency (cm") J J I A. 1 I V 1 V I V Y ,. I Y Yfi w I T 7 1 T ‘ V T V ‘ ‘ 7 I V I I Y I V I I V I ' 1 ‘ V Y Y I T Y r T fT V V I Y 12 11 10 9 8 7 6 5 4 chemical shift (ppm) “—4 N“. H Figure 2.10. (a) IR spectrum (KBr pellet) of poly(MP-APVE), (b) IH-NMR spectrum of poly(MP-APVE). -40- Poly(4-HPM-VEP): The IR spectrum of this polymer shows intermolecular hydrogen bonding based on the -OH stretching resonance of the phenol hydroxyl group at ~3430 cm'1 (Figure 2.11a). The strength and spectral width of this band obscures the aromatic methylene stretches. We note the presence of the maleimide C=O stretches. This polymer exhibits several IR bands that are characteristic of phosphate or phosphonate. The phosphoester P=O stretch is superimposed on the phenol ring C-O stretch at ~1200 cm". We also see the P-O-C stretch at 995 cm]. This band is important in structural assignment because it demonstrates the absence of the oxidation of the phosphoester groups to the corresponding phosphorus acid. We deliberately keep the phosphonate groups protected during multilayer assembly so they can be used for metal ion sequestration once the multilayer structure is formed.31 The 1H-NMR spectrum (Figure 2.11b) of this polymer reveals an aromatic resonance centered at approximately 7.0 ppm. Again we observe the splitting of this peak due to the presence of the para- hydroxyl group, where the splitting is a result of ring substitution at the 4- position. We also observe a peak centered at approximately 9.8 ppm which may be due to ring opening or, possibly, to hydrolysis of some of the phosphonate groups. However, either of these processes contribute negligibly due to the integration of the peak for ~ 0.3 relative to the other peaks. The spectral signature between 1.0 and 4.6 ppm has been assigned to the various contributions of the vinyl ether phosphonate resonances as reported previously.24 -41- 2? 3 Q) 8 .8 ‘5 i 8 (<3 4000 3500 3000 2500 2000 1500 1000 500 frequency (cm‘l) b chemical shift (ppm) Figure 2.11. (a) IR spectrum (KBr pellet) of poly(4-HPM-VEP), (b) 1H-NMR spectrum of poly(4-HPM~VEP). -42- 2.4 Conclusions We have reported on the synthesis and characterization of a family of alternating copolymers possessing a variety of functional side groups capable of forming covalent linkages for controlled multilayer assembly. Some of the polymers we have prepared are capable of participating in either ionic or covalent interlayer linking chemistry. We have demonstrated the ability to introduce a variety of functional groups into the matrix through substitution of the maleimide N—substituent. This substantial structural versatility is put to use in the following chapter and we anticipate using these polymers in the design of chemically selective interfaces. -43_ 25 Literature Cited 1. Vrancken, K. C.; VanderVoort, P.; Gillisdhamers, 1.; VanSant, E. F.; Grobet, P.; J. Chem. Soc., Faraday Trans. 1992, 88, 3197. 2. Pfleiderer, 8.; Albert, K.; Bayer, E.; J. Chromatogr. 1990, 506, 343. 3. Bakiamoh, S.; Blanchard, G. J.; Langmuir, 2001 17, 3438. 4. Katz, H. E.; Scheller, G.; Putvinski, T. M.: Schilling, M. L.; Wilson, W. L.; Chidsey, C. E. D.; Science, 1991, 254, 1485. 5. Li, D.; Ratner, M. A.; Marks, T. J.; Zhang, C. H.; Yang, J.; Wong, G. K.; J. Am. Chem. Soc., 1990, 112, 7389. 6. Demas, J. N.; Degraff, B. A.; Coleman, P. B.; Anal. Chem. News & Features, 1999, 793A. 7. Yoon, H. C.; Kim, H-S Anal. Chem. 2000, 72, 922. 8. Delamarche, E; Sundarababu, G.; Biebuyck, H.; Michel, B.; Gerber, C.; Sirgrist, H.; Wolf, H.; Ringsdorf, H.; Xanthopoulos, N.; Mathieu, H. J .; Langmuir 1996, 12, 1997. 9. Brousseau, 111, L. C.; Aurentz, D. J.; Benesi, A. J.; Mallouk, T. E.; Anal. Chem. 1997, 69,688. 10. Ng, S. C.; Zhou, X. C.; Chen, Z. K.; Miao, P.; Chan, H. S. 0.; Li, S. F. Y.; Fu, P.; Langmuir, 1998, I4, 1748. ll. Beatty, S. T.; Fischer, R. J .; Hagers, D. L.; Rosenberg, E.; Ind. Eng. Chem. Res. 1999, 38,4402. 12. Nooney, R. 1.; Kalyanaraman, M.; Kennedy, G.; Maginn, E. J.; Langmuir, 2001, 17, 528. -44- l3. 14. 15. l6. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. Ulman, A. Chem Rev. 1996, 96, 1533. Dubois, L. H.; Nuzzo, R. G.; Annu. Rev. Phys. Chem, 1992, 43, 437. Zhang, Z. J.; Hu, R. S.; Liu, Z. F.; Langmuir, 2000, 16, 1158. Skulason, H.; Frisbie, C. D.; Langmuir, 1998, 14, 5834. Lee, H.; Kepley,. L. J .; Hong, H. G.; Akhter, S.; Mallouk, T. E.; J. Phys. Chem. 1988, 92, 2597. Lee, H.; Kepley, L. J.; Hong, H. G.; Mallouk, T. E.; J. Am. Chem. Soc., 1988, 110, 618. Putvinski, T. M.; Schilling, M. L.; Katz, H. E.; Chidsey, C. E. D.; Mujsce, A. M.; Emerson, A. B.; Langmuir, 1990, 6, 1567. Home, J. C.; Blanchard, G. J.; J. Am. Chem. Soc. 1999, 121, 4427. Major, J. S.; Blanchard, G. J .; J. Phys. Chem, in preparation. Major, J. S.; Blanchard, G. J .; chapter 3. Kohli, P.; Scranton, A. B.; Blanchard, G. J.; Macromolecules, 1998, 31, 5681. Kohli, P.; Blanchard, G. J.; Langmuir, 1999, 15, 1418. Kohli, P.; Blanchard, G. J .; Langmuir, 2000, 16, 695. Jung, M. E.; Lyster, M. A.; J. Org. Chem. 1977, 42, 3761. Vickery, B. H.; Pahler, L. F.; Eisenbraun, E. J .; J. Org. Chem, 1979, 44, 4444. Kohli, P.; Blanchard, G. J. Langmuir, 2000, 16, 8518. Silverstein, Bassler and Morrill, Spectrometric Identification of Organic Compounds, 5th Ed., John Wiley & Sons, New York, 1991. -45- 30. Lin-Vien, D; Colthup, N. B.; Fateley, W. G.; Grasselli, J. G. The Handbook of Infrared and Raman Characteristic Frequencies of Organic Molecules; Academic Press, San Diego, 1991. 31. Major, J. S.; Blanchard, G. J. Langmuir 2001, 17, 1163. -46- Chapter 3 Strategies for Covalent Multilayer Growth. 2. Interlayer Linking Chemistry Abstract We report a strategy for the covalent assembly of polymer multilayers at interfaces, where growth is accomplished one layer at a time. The individual layer constituents are maleimide-vinyl ether alternating copolymers with side groups that possess reactive functionalities. The identity of the polymer layer side groups determines the chemistry we employ in interlayer linkage formation. We report on the selective creation of amide, ester, ether, urea and urethane interlayer linkages. We achieve controlled multilayer growth using each of these covalent bonding schemes. The resulting multilayer structures show linear growth in terms of thickness, measured ellipsometrically, and total mass loading, measured by UV-visible and FIIR spectroscopies. -47- vu—jr‘ '. 7"“T“"tl 3.1 Introduction The design of interfacial materials where there is explicit control over the identity and ordering of the constituent layers is an area of intense research effort. This chemical approach to surface modification can find use in sensor design,” chemical separationss'6 and nonlinear optics, 7'8 for example. To date, there have been several means devised to control layered material growth at interfaces, and among the most widely used are 9.10 11-14 11. Langmuir-Blodgett film growt thiol/gold monolayer self-assembly and metal bisphosphonate layer growth.15 '35 Each method has its advantages and shortcomings, so there is a continued need to identify novel layer growth chemistry where the resulting materials are relatively well organized and the chemistry binding the layers together is sufficiently robust to be useful. We describe here the assembly of several polymer multilayer systems, where the interlayer linking chemistry involves the formation of covalent bonds. The scheme we report here is, in a sense, the layer-by-layer cross- linking of a thin polymer film, where the structure of the polymers in a given layer (parallel to the substrate) can be controlled. The polymers we use in the creation of the multilayer structures have been published36 and described in detail in chapter 2. Careful choice of the polymer side groups allows layers to be added to the interface through 3738 allowing for the assembly 0f either ionic or covalent interlayer linking chemistry, materials that can function either as ionic conductors or as insulators. We are interested in covalently-linked polymer multilayers for several reasons. Earlier work on zirconium bisphosphonate (ZP) ionically-linked multilayer structures showed that interlayer excitation transport was inefficient by virtue of the polarizable zirconium bisphosphonate sheets that connect the layers in these materials.35 To control -43- interlayer excitation transport in layered interfaces, a necessary first step is the elimination of these polarizable layers; we replace the ZP functionality with covalent organic functionalities (vide infra). We have chosen to use polymers as the layer constituents in this work because of the combined ease of preparation and the degree of structural versatility available with maleimide-vinyl ether alternating copolymers.39 Polymers provide structural integrity within individual layers, allowing the opportunity to create relatively porous structures where there is a gradient in some property of the interface (e.g. dipole moment) along the surface normal axis.40 Such interface properties are likely to be of use in cherrrical separations and sensing. The polymer systems we reported in the preceding chapter allow control over the individual layers through synthetic design.33 The interlayer linking chemistry applied to these polymers that we report in this chapter allows for a second level of control; we can choose the means of interconnection between individual layers. In this work we will focus on an array of covalent interlayer linking chemistries that we have applied to layered material growth. The resulting materials are all robust and the strategies we discuss should be applicable to the growth of a variety of other materials in layered structural motifs. We anticipate that this level of interface structural control will find use in the design of materials intended for specific purposes, such as the creation of a chemical potential gradient normal to the substrate plane. 3.2 Experimental In the preceding chapter, we reported on the design and characterization of several maleimide-vinyl ether alternating copolymers. We use some of these polymers to -49- construct multilayer structures. We describe the procedures for substrate preparation and then present several strategies for covalent linking of polymer layers. We have used gold, quartz and silicon substrates for this work, with the choice of substrate being determined by the measurement to be performed. We used gold substrates for optical null ellipsometry and FTIR measurements, silica substrates for UV-visible spectroscopy and oxidized silicon for ellipsometry and Brewster’s angle transmission FTIR measurements. We have chosen specific chemical systems to demonstrate the chemistry of interest and have not attempted to report on all of the possible permutations of polymers and interlayer bonding schemes. Substrate preparation: Gold, quartz and silicon wafers were prepared by cleaning in piranha solution (3:1 H2804:H202) for ~15 minutes, followed by rinses with ethanol and water. The substrates were then dried under a stream of nitrogen gas. The quartz and silicon wafers were then exposed to a 5M HCl solution approximately 15 minutes followed by ethanol and water rinse and dried under nitrogen. After cleaning and preparation, the substrates were exposed to the appropriate solution to give the desired terminal functionality. To functionalize the gold surfaces with hydroxyl groups, the cleaned substrates were exposed to a 60:40 ethanolzwater 10 m1\_/l solution of 6-mercaptohexan-l-ol for ~30 minutes at ~40°C. For the quartz and oxidized silicon surfaces, the surface silanol groups were used directly and without prior treatment for further functionalization. To produce amine-terminated surfaces, the gold substrates were exposed to a 60:40 ethanolzwater 10 mM solution of cystamine dihydrochloride for 30 minutes at -50- ~40°C. Quartz and silicon wafers were exposed to ~5% solution of 3-aminopropyl- trimethoxysilane in toluene at room temperature overnight. To create a surface terminated with an acid chloride functionality, the substrate was reacted with an (0t,0))-diacid chloride. Hydroxyl-terminated gold substrates were exposed to a solution of 10 mL anhydrous acetonitrile, 0.3 mL 4-methylmorpholine and 0.3 mL adipoyl chloride under an inert atmosphere of argon at room temperature for 30 rrrinutes to produce a surface terminated with an acid chloride functionality. Quartz and silicon wafers were treated in the same way, where the surface silanol groups were reacted with adipoyl chloride. To create an epoxide-terminated surface, gold substrates that had been previously prepared to present a hydroxyl terminal functionality were exposed to a 1% (v/v) solution of 1,4-butanediglycidyl ether in anhydrous acetonitrile. Concentrated sulfuric acid (0.06 mL) was added drop-wise to the reaction vessel to mediate the epoxide ring opening. The reaction was allowed to proceed at room temperature for approximately 5 rrrinutes before rinsing. Silicon and quartz wafers were reacted directly with 1,4-butanediglycidyl ether in a manner similar to that for the gold substrate. For these substrates, the reaction is directly with the silanol groups present on the surface. For isocyanate-terminated substrates, the gold substrates treated with 6-mercapto- l—hexanol were immersed in a solution 10 mL anhydrous acetonitrile, 0.3 mL 4- methylmorpholine and 0.3 mL 1,6-diisocyanatohexane and heated to ~40°C for 30 minutes. As with the other reactions, the quartz and silicon substrates were treated in a similar manner and their native silanol groups were used as the reactive functionalities. -51- ...n i3 ’4“ ‘1... a.“ Optical Null Ellipsometry: Values of ellipsometric thickness of the deposited polymer multilayer assemblies were acquired using a Rudolph Auto-EL H optical null ellipsometer. This instrument is equipped with a HeNe laser operating at 632.8nm. The Software (Rudolph DAFIBM) used for data collection was acquired from the manufacturer. The refractive index was taken to be n = 1.54 + 0i for all films. Even though this value may differ by a small amount depending on the identity of the polymer film, changes over a physically realistic range affect the recovered thickness negligibly. U V— Visible Spectrophotometry. The absorption spectra of the multilayer assemblies were acquired using a Cary 300 UV-Visible spectrophotometer. The layers were synthesized on quartz and the data were acquired after the deposition of each individual layer. The scan rate for all measurements was 600 nm/min and the spectral resolution was 1 nm. The data were processed and plotted using Microcal Origin® version 6.1 software. In all cases, we recovered R2 values greater than 0.98 for linear fits. Infrared Spectroscopy. FTIR spectra of the multilayer assemblies on gold substrates were collected using a Nicolet Magna 750 FT IR spectrometer operating with Omnic® software. The spectral resolution was set to 4 cm'1 for all spectra collected. An external reflectance sample mount set at an incidence angle of 80° was used for data collection. Spectra of samples grown on silicon were acquired using a Nicolet Magna-IR 550 set up to acquire data in the transmission mode using a Brewster’s angle attachment. 3.3 Results and Discussion In this work, we report the multilayer assembly of polymers that have been designed with various functional side groups that allow covalent interlayer linkages. The goal of this work is to demonstrate that careful choice of the functional groups -52- incorporated into the copolymers provides accessibility to simple reactions that form arrride, ester, ether, urea and urethane functionalities, allowing the regular growth of multilayer systems. In many cases, there is more than one way to generate a particular covalent interlayer linkage, with the order of the chemical functionalities used being determined by the native functionality or initial modification of the substrate. This work is intended to be more than a catalog of covalent interlayer linking chemistry. It is our intent to demonstrate the facile nature of our approach to the controlled growth of multilayer interface structures. In all of the examples we present here, the strategy is essentially the same. We use alternating copolymerization chemistry where the polymerization reaction is templated from the substrate initially. At each step of the growth process, the reactive interface cannot grow another layer until it is presented with the requisite monomer. By presenting the interface with only one type of monomer at a time, controlled growth of individual, discrete molecular layers is possible. In cases where the same multilayer is available through multiple routes (save for orientation of the interlayer linking functionality), we report the reaction(s) that we have demonstrated experimentally. Depending on the monomer species used in the MVE polymer synthesis, either the maleimide or vinyl ether side groups can be used for interlayer attachment. Amide chemistry. Amide interlayer linkages are created by reaction of an acid chloride with an amine (scheme 3.1) and we report three viable routes. (3) An amine- terrninated substrate is exposed to a solution of an MVE alternating copolymer with acid chloride side group functionalities, resulting in the direct formation of an amide bond between the polymer side groups and the substrate. Subsequent layers are added by -53- ". fimzl‘fla’ '1" exposure of the acid chloride containing polymer (topmost) layer to an (0t,c0)-diamine, followed by reaction of the aminated surface with the MVE alternating copolymer. (b) An acid chloride-terminated substrate is exposed to a solution of an MVE alternating copolymer with amine-terminated side groups. Controlled multilayer growth is achieved by reacting the polymer amine side group with an (0t,0))-diacid chloride (adipoyl chloride) followed by exposure to the MVE polymer. This layer growth scheme is identical to the first save for the orientation of the resulting amide interlayer linkage. (c) An amine- terrninated substrate is exposed to a MVE alternating copolymer containing acid chloride side groups, followed by exposure to a MVE alternating copolymer with amine- terminated side groups. In this scheme, no interlayer linker is necessary. The choice of initial polymer layer in this scheme is determined by the reactive functionality presented by the substrate. -54- ° )L /U\ + R'—NH2 ——4» /R Scheme 3.1. Reaction schematic and idealized structures of amide-linked MVE layered assembly. (a) Amine-containing MVE polymer bound to acid chloride functionalized surface and (b) acid chloride containing MVE polymer bound to amine-functionalized surface. Co-polymerization of 3-chlorophenylmaleimide and 3-aminopropyl-l-vinyl ether produces poly(3-CPM-APVE),36 a polymer that is capable of participating in amide interlayer linkage formation. Reacting poly(3-CPM-APVE) with adipoyl chloride leads to the formation of amide interlayer linkages. The IR spectrum of the amide-linked -55- multilayer assembly (Figure 3.1a) reveal amide I, H and HI bands, where the amide I band corresponds to the carbonyl stretching mode centered at approximately 1655 cm". Amide H and III bands correspond to N-H bending vibrations and C-N stretching modes occurring at approximately 1595 cm'l and 1421 cm", respectively. In addition, the band centered at ~3330 cm'1 corresponds to the stretching modes of the free -NH groups. UV—visible spectroscopy is used to monitor layer deposition, by observing the aromatic absorbance band centered at 248 nm (Figure 3.1b - inset). Errors were calculated based on the noise level and found to be less than 2% in all cases reported. The observed spectral red shift relative to the N-phenylmaleimide copolymer is the result of the chloro functionality at the 3- position. We observe a loading density for the amide- linked assembly that is constant with number of MVE polymer layers (Figure 3.1b). From the ellipsometric data (Figure 3.2) we recover a slope of 19 A/Iayer for the amide- linked multilayer, consistent with that seen for several other interlayer linking strategies. -56- ().()()(i — 31‘ t3 Eg’ ().()()4- — 5 t i O . jg ().()Cl2. — E 0 . O O 0 ‘ l ‘ ' ‘ I’IV - l g 1 z1000 3500 3000 2000 1500 1000 hequency (cm‘fi 0.12 — - (r08- 1 E . ff 0'10 _ £0.06» ,. V _ 0 04 Q) 0 .08 g ’ /+ g ' .§ 0.02- i e 0.06 — 0 00> . . . 8 ' i 0 1 2 3 4 "D — layer number {1 0.04 0.02 - 0 .00 ‘ ‘ ’f-iAvxjx‘fcm bibm-I 200 250 300 350 Wavelength (nm) Figure 3.1. (a) FTIR spectrum of poly(3CPM-APVE) with amide interlayer linkage (b) UV-Visible spectra as a function of number of layers for poly(3CPM-APVE). -57- 90 80 70 I 1 T T I—l 40; 30- 201 103- 0’ . . . . 0 r 2 3 4 Layer number Thickness (A) Figure 3.2. Dependence of ellipsometric thickness on number of layers for poly(3CPM-APVE) with amide interlayer linkage. Ester chemistry. Ester interlayer linkages are achieved by the reaction of an acid chloride with an alcohol (scheme 3.2). There are three routes to multilayer formation using this chemistry. (a) A MVE alternating copolymer with hydroxyl-terminated side groups is reacted with an acid chloride terminated substrate. Multilayer assembly is achieved by reacting the polymer hydroxyl side groups with a diacid chloride followed by exposure to the MVE polymer. (b) The complement of the first route can be used, where the MVE polymer contains acid chloride terminated side groups and the interlayer linking moiety is an (0t,0))-diol. -58- muar‘rfi-‘fl 0 AL A + R'—OH —> /R' 1 Scheme 3.2. Reaction schematic and idealized structures of ester-linked MVE layered assembly. (a) alcohol-containing MVE polymer bound to acid chloride functionalized surface and (b) acid chloride containing MVE polymer bound to hydroxyl-functionalized surface. To realize this chemistry, we used the alternating copolymer of N- phenylmaleirrride (NPM) and l-vinyloxy-4-butanol (VOB), poly(NPM-VOB); The reaction of poly(NPM-VOB) with adipoyl chloride results in an ester—linked multilayer assembly. Linear growth was seen in the ellipsometric data (Figure 3.3a), with a slope of 21 A/Iayer. As discussed in earlier work, it is not a straightforward matter to evaluate the thickness of a layer relative to a molecular structure calculation (e.g. molecular -59- wmm:m1l mechanics) because the loading density of the interface is not known quantitatively. The thickness) we recover experimentally is consistent with other multilayer structures we report here (vide infra) and have reported elsewhere. The absorbance data indicate a linear growth as well, (Figure 3.3b - inset), although it is not possible to determine the absolute loading density from these data because the extinction coefficient at 220 nm for poly(NPM-VOB) is not known. -60- mica . «YEA-til ii 100 I- 90 : ", 80 h ,”’ A 70 " ell’ °§ 60 - I,” 8 I- ’1 50 1" [’1 .4 . - 5 40 — I,’ 30 "' ’1” I- I.’ 20 -— ,z 10 — 0 p l l I l 0 l 2 3 4 Layer number _ 0.08~ g 0.07; ' 0.15 - § 0.06L ' A v > 3' g 0.05- . “i i E 0.04. G) "“ 0.03- 0 I— . g 0.10 , .8 r 0 I 2 3 4 o layernumber 3 <1 0.05 — 0.00 " 1 A I . Wm .3 . .2” ”44.14 200 250 300 350 Figure 3.3. (a Wavelength (nm) ) Dependence of ellipsometric thickness on number of layers for poly(NPM-VOB) with ester interlayer linkage. (b) UV-Visible spectra as a function of number of layers. Inset: Dependence of absorption maximum on number of layers. -61- -___.,._-m A slight modification of this reaction scheme has been reported previously.“ In that work, the reactive polymer side group is attached to the maleirrride monomer and the vinyl ether monomer is terminated with a diisopropylphosphonate functionality, the interlayer linking chemistry can proceed through the maleimide monomers. In this reaction scheme, the monomers are 4-hydroxyphenylmaleimide (HPM) and ethyl vinyl ether diisopropyl phosphonate (VEP), and the interlayer linking moiety is adipoyl chloride. We note that phosphoester formation is precluded in this layer growth scheme due to presence of the diisopropyl protecting groups on the phosphonate functionality.“ The MVE polymer is capable of participating in ionic and/or covalent interlayer linkages. .O O p—I O l a g 0.008 - e g 0.006L g . .0 § .8 b 4000 l 3500 ' 3000 2000 ' 1500 . 1000 ‘ frequency (cm-1) Figure 3.4. Poly(HPM-VEP) with ionic and covalent interlayer linkages. (a) Infrared spectrum of ionically-bound multilayers and (b) IR spectrum of covalently-bound multilayers. The spectra have been offset for clarity. -62- For ionic multilayer formation, the phosphonate groups are deprotected by reaction with bromotrimethylsilanef’l resulting in the efficient hydrolysis of these groups to the corresponding phosphonic acid. Sequential exposure of the interface to M4+ (M = Zr, Hf) and the MVE polymer, with deprotection chemistry used as required, results in ionic multilayer formation. Figure 3.4a shows the infrared spectrum of a multilayer of poly(HPM-VEP), linked by the formation of zirconium bisphosphonate moieties. For covalent interlayer linkage, the terminal hydroxyl group of the poly(I-IPM-VEP) polymer is capable of participating in ester, ether or urethane interlayer linkages. Figure 3.4b shows the IR spectrum of a 4 layer stack of poly(HPM-VEP) assembled using ester linking chemistry. The ester carbonyl stretching is centered at ~1720 cm”l (superimposed on the maleimide asymmetric C=O stretch). The broad band centered around 3200 cm'] is an indication of residual water or hydrogen-bonded phosphonic acid functionalities within the multilayer. The ellipsometric and UV-visible spectroscopic data for this multilayer indicate linear growth. We recover a slope of 16A/layer for the ellipsometric data (Figure 3.5). Regression of the UV-visible absorbance data was made possible by monitoring the aromatic absorbance band centered at 231 nm (Figure 3.5 - inset). -63- .r-Kiufi .0 o [\J o absorbance (231 um) i i 3 4 layer number Absorbance (a.u.) g: 8 200 A 250 I 300 350 Wavelength (nm) Figure 3.5. UV-Visible spectra as a function of number of layers. Inset: Dependence of absorption maximum on number of layers. Ether chemistry. We have demonstrated ether connections between MVE polymer layers using a ring-opening reaction of an epoxide with a hydroxyl group (scheme 3.3). The reaction of an MVE polymer with hydroxyl-terminated side groups with an epoxide-functionalized substrate in the presence of a strong acid leads to ether- linked multilayer assemblies. Subsequent exposure of the MVE polymer surface to an (01,0))-diepoxide yields an epoxide-terminated surface ready for further reaction. The reaction of an adlayer of poly(NPM-VOB) with 1,4-butane diglycidyl ether leads to ether linkages. Multilayer assembly was monitored using ellipsometry, and FI‘IR and UV-visible spectroscopies. A linear response was observed for the regression -64- 1s citrate-19:1. of the ellipsometric data (Figure 3.6a), and we recover a slope of 29 A/Iayer. This thickness is substantially greater than that which we observe for the other multilayer assemblies suggesting either greater loading density or greater conformational disorder within the MVE polymer layer. We note that the formation of ether interlayer linkages appears to be limited to the deposition of two or three layers before terminating. There are several interesting features contained in the data for this system. j"; R—Q) + R'—OH ——> R O‘R' H* WT.— HO Scheme 3.3. Reaction schematic of a glicidyl ether with an alcohol (top) to form an ether and an idealized structure of an alcohol-containing MVE polymer bound to an epoxide functionalized substrate. Subsequent layer growth requires reaction of the alcohol-containing MVE polymer with a diglicidyl ether. -65- 0.08 0.06 0.04 absorbance (a.u.) 0.02 0.00 100 Z 90 80 : 70 _ 60 _ 50 40 30 20 10 Thickness (A) r I 0.16 ' _ A. ’ a g 0.12~ E? g 0.08» . " E 0.04» i 0'000 I 2 3 4 _ layer number - J A l g l A I 200 250 300 350 wavelength (nm) b - .I’ I I, ... I” l’ I— I, 1’. [I r f” . L 1 1 1 0 1 2 3 4 Layer number Figure 3.6. (a) Dependence of ellipsometric thickness on number of layers for poly(NPM-VOB) with ether interlayer linkage. (b) UV-Visible spectrum for polym(NPM-VOB) with ether interlayer linkage. Inset: Dependence of absorption maximum on number of layers. First, the UV-visible spectroscopy suggests that there is a much higher loading density of the chromophore when compared to the deposition of the same polymer -66- system using either ester or urethane interlayer linkage (Figure 3.6b), and this finding is consistent with the ellipsometric data. The anomalously high layer density may be the result of the way we performed the reaction. We used concentrated sulfuric acid to facilitate the epoxide ring-opening reaction and to promote ether formation. There is no analogous reaction step in the formation of ester or urethane interlayer bonds. Repeating the epoxide ring opening reaction with HCl instead of H2804 yielded lower density layers. We observe a similar pattern when the reactions for the ester-, amide- and urethane-linked system are performed in the presence of H2S04 versus HCl, suggesting the importance of dehydration by the H2804 in these reactions. We will explore this issue more fully in a subsequent study. We also note that there is an unexpectedly large 0.02 r- O O p—s l absorbance (a.u.) 0.00 Maj! 1 . 1 4000 3500 3000 2000 1500 1000 frequency (cm'l) Figure 3.7. IR spectrum for poly(NPM-VOB) with ether interlayer linkage. -67- IR resonance near 1700 cm'l for the ether-linked assembly (Figure 3.7). This is not an expected result because the ring opening reaction should yield a secondary alcohol and an ether (scheme 3.3). One possible explanation for this observation is the oxidation of the alcohol to a ketone, and this result is consistent with the absence of a strong OH stretching resonance (Figure 3.7). To test for the possibility of this reaction, we reacted an oxidized silicon substrate with 1,4-butane diglycidyl ether. The IR spectrum of the ether-modified substrate is dominated by resonances associated with the asymmetric C- O-C stretch centered about 1100 cm'1 (Figure 3.8a). l A l L l 1 IV I l n l n 4000 3500 3000 2500 1500 1000 frequency (cm'l) Figure 3.8. Ether chemistry for modified substrate. (a) FI‘IR showing C-O-C stretch of epoxide prior to reaction, (b) FIIR of ring—opened system after reaction indicating ketone and acid functionalities. -6g- The substrate was then immersed in anhydrous acetonitrile and a few drops of concentrated sulfuric acid were added to facilitate the ring opening hydrolysis reaction. The substrate was then heated to ~40° for 30 minutes. The IR spectrum of the reacted substrate showed a marked decrease in the C-O-C stretch, consistent with the ring- opening reaction, and the presence of C=O stretching bands of almost equal intensity (Figure 3.8b). The carbonyl resonance appears as a doublet, with one peak centered at approximately 1726 cm'1 and the other at about 1745 cm]. These two resonances are consistent with the formation of an aldehyde and a carboxylic acid, respectively.42 We observe that this chemistry is difficult to control beyond the deposition of about two to three molecular layers, at which time, we observe no further growth. If the epoxide groups are being opened and subsequently oxidized, as the IR data suggest, then they are rendered capable of participating in interlayer bonding. It is also possible that the epoxide reaction proceeds in an intralayer fashion, which would also preclude the formation of interlayer ether linkages. It may be possible to minimize such side reactions by running these reactions at room temperature, using strictly anhydrous solvents, and minimizing atmospheric exposure. Urea chemistry. Urea linkages are made by the reaction of an amine with an isocyanate (scheme 3.4). An isocyanate-functionalized substrate is exposed to an MVE polymer with amine-terminated side groups. The surface layer of polymer is reacted with a ((1,0))-diisocyanate, then with another layer of the same MVE polymer. The complementary chemistry, where an amine terminated surface is reacted with a MVE polymer containing isocyanate side groups is also possible. -69- R—N=c=o + R'-—NH2 ——> R\ /U\ /R' Cl HzN 0% O O O N ‘1‘... HN l Scheme 3.4. Reaction schematic of an isocyanate with an amine (top) to form a urea and an idealized structure of an amine-containing MVE polymer bound to an isocyanate-functionalized substrate. Subsequent layer growth requires reaction of the amine-containing MVE polymer with a diisocyanate. We use poly(3-CPM-APVE) to grow urea-linked multilayer structures. Reacting poly(3-CPM-APVE) with 1,6-diisocyanatohexane leads to urea formation. The IR band assignment for the urea-linked multilayer structure is virtually identical to that for the amide-linked multilayer, save for differences in band intensities, as expected. The IR spectrum of the amide-linked multilayer assembly (Figure 3.9a) reveal Amide I, H and HI bands, where the amide I band corresponds to the carbonyl stretching mode centered at approximately 1655 cm]. Amide II and III bands correspond to N-H bending vibrations -70- and C-N stretching modes occurring at approximately 1595 cm'1 and 1421 cm'I respectively. In addition, the band centered at approximately 3330 cm'I corresponds to 0.005 — a 2;: 0.004 - 3 8 0.003 - 5 .0 ‘55 0.002 r .8 cs: 0.001 - 4000 3500 3000 2000 1500 ‘ 1000 -1 frequency (cm ) b l’ A 0.04. .,. 0.10 "' g r ." 9, 0.03- ,, f? " {3’ . x ,. :3 a) C6 0.08 — Q 0.02“ "’x. V 5 . ”x 8 E 0.01 - " 5 0.06 e 000. g ' 0 1 2 3 4 ,8 0.04 _ layer number < 0.02 \ “N. 0.00 I J A ‘t¥.>sAfe;'.r.. .. ,.,... 200 250 300 350 Wavelength (nm) Figure 3.9 (a) IR spectrum of poly(3CPM-APVE with urea interlayer linkage. (b) UV-Visible spectra as a function of number of layers for poly(3CPM-APVE) with urea interlayer linkage. Inset: Dependence of absorption maximum on number of layers. -71- the stretching modes of the free NH groups. This band is significantly larger in the urea- linked multilayer than in the amide-linked layers, as expected because each urea moiety possesses two free NH groups. We also observe larger amide II and HI bands for the same reason. As with the amide-linked multilayers, we monitor the 248 nm absorbance band - which corresponds to the 3-chlorophenyl maleimide moiety - using UV-visible spectroscopy (Figure 3%). We observe a higher loading density of the polymer for the amide-linked assembly than for the urea-linked multilayer. One explanation for this finding is that urea formation may be less efficient than amide formation under our reaction conditions (reaction conditions were identical for both deposition schemes). The ellipsometric data for the urea-linked assembly show a slope of IDA/layer (Figure 3.10). This result is smaller by a factor of ~2 compared to the amide-linked system and is consistent with the uv-visible absorbance data. 70 .- 60 l 50 l 3-x 40 ; ,4" 30 L i» 20 - ,”§/’ 10 1 O l J l 0 1 2 3 4 Layer number Thickness (A) :- Figure 3.10. Dependence of ellipsometric thickness on number of layers for poly(3CPM-APVE) with urea interlayer linkage. -72- Urethane chemistry. The formation of urethane linkages can be accomplished by the reaction of a hydroxyl group with an isocyanate group (scheme 3.5). We report two such schemes here. (a) A hydroxyl-terminated substrate is reacted with an ((1,0))- R—NZCZO + R'—OH ———’ N O H Scheme 3.5. Reaction schematic of an isocyanate with an alcohol (top) to form a urethane and an idealized structure of an alcohol-containing MVE polymer bound to an isocyanate-functionalized substrate. Subsequent layer growth requires reaction of the alcohol-containing MVE polymer with a diisocyanate. diisocyanate, followed by exposure to an MVE polymer with hydroxyl terminated side groups. Multilayers are assembled by alternate exposure of the substrate to the diisocyanate and the MVE polymer. (b) An MVE polymer with isocyanate-terminated side groups is exposed to a hydroxyl-terminated substrate. This reaction results in the -73- formation of a urethane linkage. Multilayer assembly is achieved through alternate exposure to a diol and the MVE polymer. We use poly(NPM-VOB) and 1,6-diisocynatohexane to form urethane-linked multilayer assemblies. The ellipsometric data (Figure 3.11) reveal linear growth with a slope of 18 Allayer, consistent with that seen for the amide-linked layers and larger by a factor of ~2 than for the urea-linked multilayers. This finding suggests that the reaction of isocyanates with alcohols is more efficient than with amines. Growth of the urethane- linked layers was monitored using FI‘ IR (Figure 3.12a) and UV-visible (Figure 3.12b) spectroscopies, and regular growth was observed with these methods as well. 100 90;- . 80 7 28:: i 50 ~ 40 '- 30 '- {r 20 "- ’ 10 - O p 1 1 I 0 1 2 3 4 Layer number I' ‘fi 0 Thickness (A) Figure 3.11. Dependence of ellipsometric thickness on number of layers for poly(NPM-VOB) with urethane interlayer linkage. In addition to the alternating copolymerization of maleimides and vinyl ethers, we can also use alkenes as reactive monomers. The interlayer linking chemistry we describe above is equally applicable to such polymers, and in all cases is determined by the -74- m*“”““‘1 I 0.016 I 0.012 I 0.008 absorbance (a.u.) 0.004 ~ 0.000 AMA 4000 3500 3000 2000 1500 A 1000 frequency (cm'l) E b o . 2 0.25 r- 3 0.101,“. “9 ’ o 0.081 . 2: 0-20- s - g . ,2 0.06 t g 0.15 — :9: 0.04: x”. ’ 2 0.10 _ < . layer number 0.05 L O . O O " l . I ‘ ~ v‘fi'i’.“t “avg/«,3, .’ 200 250 300 350 W avelength (nm) Figure 3.12. (a) IR spectrum for poly(NPM-VOB) with urethane interlayer linkage (b) UV-Visible spectra as a function of number of layers for poly(NPM-VOB) with urethane interlayer linkage. Inset: Dependence of absorption maximum on number of layers. -75- terminal functionalities of the resulting polymer side groups. The co-polymerization of NPM and 4-pentenoyl chloride, poly(NPM-4PC), is capable of participating in both ester and amide interlayer linkages as shown in Figure 3.13a. This reaction scheme appears to be less successful than the same interlayer linking chemistry used for the MVE alternating copolymers. For poly(NPM-4PC) we see layer growth that is limited to one or two molecular layers before terminating. One explanation for this behavior is that the oxidation of the acid chloride groups to the acid followed by intermolecular hydrogen bonding between the acid functionalities to render these groups less accessible for subsequent layer deposition. It is also possible for the putative carboxylic acid groups to hydrogen bond with solvent or adventitious water during rinsing and measurement. Examination of the IR data (Figure 3.13b) is consistent with these explanations. The broad —OH stretch centered at ~3100 cm'1 is characteristic of intermolecular hydrogen bonding. A second piece of evidence in support of this information is the position of the C=O band in these polymer adlayers. Dimerization of carboxylic acids shifts the carbonyl stretch to a lower frequency than that of the free carboxylic acid. For a poly(NPM-4PC) adlayer the C=O stretch is centered at ~1710 cm'l, whereas for an acid chloride, the C=C stretch is expected to be near 1775 cm". -76- ] 1 0.0010 0.0008 0.0006 0.0004 L 0.0002 P 0.0000 _ -00002 - . - . - . - . . . 4000 3500 3000 2500 2000 1500 I I V absorbance (a.u.) A 1000 frequency (cm'l) Figure 3.13. Poly(NPM-4-PC). (a) Structures — with amide and ester linkers. (b) IR spectrum of a poly(NPM-4PC) layer bound to the substrate by an ester linkage and with remaining acid chloride functionalities hydrolyzed by expose to air. -77- 3.4 Conclusions We have demonstrated our ability to design and assemble polymer multilayer assemblies with layer-by-layer control of the layer identity and interlayer linking chemistry. We have presented the design and synthetic schemes for the preparation of these polymers in the preceding chapter. In this paper we have presented our strategies for assembling these materials into layered structures where the identity of the polymer side groups allows simple chemical reactions to be used to link the individual layers. The functionalities formed by this simple chemistry include amide, ester, ether, urea and urethane. The design and assembly schemes we present here have the distinct advantage of ease of preparation of the monomers and polymers, ease of assembly of these materials into multilayers and relatively short deposition times. We have also shown that, depending on the functional groups of the copolymer, it is sometimes possible to assemble these materials where prior surface preparation is not necessary, as is the case for reactions on silicon and quartz wafers. In these cases, the silanol groups serve as the reactive moiety to which our polymers can be tethered. -73- 3 .5. Literature Cited 1. Kepley, L. J.; Crooks, R. M.; Ricco, A. Anal. Chem. 1992, 64, 3191. 2 Demas, J. N.; Degraff, B. A.; Coleman, P. B. Analytical Chemistry News & Features, 1999, 723A. 3. Yoon, H. C.; Kim, H-S. Anal. Chem. 2000, 72, 922. 4. Delamarache, E.; Sundarababu, G.; Biebuyck, H.; Michel, B.; Gerber, C.; Sirgrist, H.; Wolf, H.; Ringsdorf, H.; Xanthopoulos, N.; Mathieu, H. J. Langmuir, 1996, 12, 1997. 5. Vrancken, K. C.; Van Der Voort, P.; Gillis-D’Hammers, I.; Vansant, E. F.; Grobet, P.; J. Chem. Soc., Faraday Trans. 1992, 88, 3197. 6. Pfleiderer, B.; Albert, K.; Bayer, E. J. Chromatogr. 1990, 506, 343. 7. Katz, H. E.; Scheller, G.; Putvinski, T. M.; Schilling, M. L.; Wilson, W. L.; Chidsey, C. E. D. Science 1991, 254, 1485. 8. Li, D.; Ratner, M. A.; Marks, T. J.; Zhang, C. H.; Yang, J.; Wong, G. K. J. Am. Chem. Soc. 1990, 112, 7389. 9. Langmuir, 1.; J. Am. Chem. Soc., 1917, 39, 1848. 10. Blodgett, K. B.; J. Am. Chem. Soc., 1935, 57, 1007. ll. Ulman, A.; Chem. Rev., 1996, 96, 1533. 12. Nuzzo, R. G.; Allara, D. L. J. Am. Chem. Soc., 1983, 105, 4481. 13. Ulman, A. An Introduction to Ultrthin Films: From Langmuir-Blodgett to Self- Assembly; Academic Press, Inc.: New York, 1991. 14. Nuzzo, R. G.; Dubois, L. H.; Allara, D. L. J. Am. Chem. Soc., 1990, 112, 558. -79- ‘l 2.“ IA‘ -.11 ‘r-I-‘V ‘ 15. l6. 17. 18. 19. 20. 21. 22. 23. Lee, H.; Kepley, L. J.; Hong, H.-G.; Mallouk, T. E. J. Am. Chem. Soc. 1988, 110, 618. Lee, H.; Kepley, L. J .; Hong, H.-G.; Akhter, S.; Mallouk, T. E. J. Phys. Chem. 1988, 92, 2597. Yonemoto, E. H.; Saupe, G. B.; Schmehl, R. H.; Hubig, S. M.; Riley, R. L.; Iverson, B. L.; Mallouk, T. E. J. Am. Chem. Soc. 1994, 116, 4786. Yang, H. C.; Aoki, K.; Hong, H. — G.; Sackett, D. D.; Arendt, M. F.; Yau, S.-L.; Bell, C. M.; Mallouk, T. E. J. Am. Chem. Soc. 1993, 115, 11855. Cao, G.; Rabenberg, L. K.; Nunn, C. M.; Mallouk, T. E. Chem. Mater. 1991, 3, 149. Rong, D.; Hong, H. — G.; Kim, Y. - I.; Krueger, J. S.; Mayer, J. E.; Mallouk, T. E. Coord. Chem. Rev. 1990, 97, 237. Thompson, M. E. Chem. Mater. 1994, 6, 1168. Vermeulen, L.; Thompson, M. E. Nature 1992, 358, 656. Vermeulen, L. A.; Snover, J. L.; Sapochak, L. S.; Thompson, M. E. J. Am. Chem. Soc.1993,115,11767. 24. Snover, J. L.; Byrd, H.; Suponeva, E. P.; Vicenzi, E.; Thompson, M. E. Chem. Mater. 1996, 8, 1490. 25. 26. Katz, H. B.; Wilson, W. L.; Scheller, G. J. Am. Chem. Soc. 1994, 116, 6636. Katz, H. E.; Bent, S. P.; Wilson, W. L.; Schilling, M. L.; Ungashe, S. B. J. Am. Chem. Soc. 1994, 116, 6631. 27. Ungashe, S. B.; Wilson, W. L.; Katz, H. E.; Scheller, G. R.; Putvinski, T. M. J. Am. Chem. Soc. 1992, 114, 8717. -80- 28. Katz, H. E.; Schilling, M. L.; Chidsey, C. E. D.; Putvinski, T. M.; Hutton, R. S. Chem. Mater. 1991, 3, 699. 29. Putvinski, T. M.; Schilling, M. L.; Katz, H. B.; Chidsey, C. E. D.; Mujsce, A. M.; Emerson, A. B. Langmuir 1990, 6, 1567. 30. Katz, H. E. Chem. Mater. 1994, 6, 2227. 31. Byrd, H.; Whipps, S.; Pike, J. K.; Ma, J.; Nagler, S. E.; Talham, D. R.; J. Am. Chem. Soc., 1994, 116, 295. Alli "v. 32. Byrd, H.; Pike, J. K.; Talham, D. R.; Chem. Mater., 1993, 5, 709. 33. Kohli, P.; Blanchard, G. J. Langmuir 1999, 15, 1418. m;— 34. Home, J. C.; Huang, Y.; Liu, G.—Y.; Blanchard, G. J. J. Am. Chem. Soc. 1999, 121, 4419. 35. Home, J. C.; Blanchard, G. J. J. Am. Chem. Soc. 1999, 121 , 4427. 36. Major, J. S.; Blanchard, G. J., Chem. Mater. in review. 37. Kohli, P.; Blanchard, G. J .; Langmuir, 2000, 16, 4655. 38. Kohli, P.; Blanchard, G. J .; Langmuir, 2000, 16, 8518. 39. Kohli, P.; Scranton, A. B.; Blanchard, G. J. Macromolecules 1998, 31 , 5681. 40. Major, J. S.; Blanchard, G. J ., Langmuir, in preparation. 41. Major, J. S.; Blanchard, G. J., Langmuir, 2001, 17, 1163. 42. Silverstein, R. M.; Bassler, G. C.; Mon‘ill, T. C. Spectrometric Identification of Organic Compounds, 5ed., New York, John Wiley & Sons, Inc., 1991. -81- Chapter 4 Acid-Enhanced Interfacial Polymer Layer Growth Abstract We report on the growth of interfacial multilayer structures formed from maleimide-vinyl ether alternating copolymers. We have demonstrated that the thickness and density of these polymer layers can be controlled by adding acid to the interlayer cross-linking reaction. We have demonstrated this control for several different interlayer crosslinking strategies, where amide, ester, urea and urethane interlayer covalent bonds are formed. For all reactions, the addition of concentrated acid during polymer layer deposition resulted in a two- to four-fold increase in the loading density of the polymer relative to the acid-free reaction, depending on the acid used and its concentration. These findings are consistent with acid catalysis (HCl) and/or dehydration (H2804). -32- w? l..t_i OYAT‘. j 4.1 Introduction Designing and constructing polymeric materials where there is explicit control over the identity of each molecular layer in conjunction with covalent interlayer linking chemistry is an emerging effort in interface science. The field of layered growth started with the work of Langmuir and Blodgett,"2 resulting in physisorbed L-B films, advanced 3-13 through the discovery of self-assembling alkanethiol/gold monolayer structures to 14-17 18-34 covalent, ionic and polyelectrolyte multilayers.”46 Such materials have found 47-51 22.52.53 use in chemical sensing, surface second harmonic generation, electronics and electro-optics.54 A common concern in all of these application areas is the extent of order and the unifomiity of coverage of the layered assembly. We have developed a means to deposit maleimide-vinyl ether (MVE) alternating copolymer55 multilayers, one layer at a 5957 and are interested in achieving time, with covalent interlayer linking chemistry, control over the density of individual polymer layers. We observe an increase in layer thickness and, in some cases, density when HCl or H2S04 are added to the crosslinking reactions used in the polymer layer growth process. This work demonstrates that we have three distinct types of control over the construction of these materials. First, we can control the identity of the functional constituents incorporated into the polymer adlayer by substitution of the polymer side group(s).57 Second, we have control over the covalent chemistry used to form the interlayer linkage for assembly of multilayer structures.56 Third, the work we report here demonstrates that we can control the loading density of the polymer at each layer through the addition of HCl or H2SO4 to the crosslinking reaction. -83- 1:7“ "h" Tu: . 'J; _ The ability to design well-behaved thin film systems where there is some level of control over the density of specific functionalities within the adlayer matrix could prove beneficial in areas such as the design of sensor materials, for example, where each functional group behaves as an individual sensing element. Controlling the density of functional constituents within a layer could also provide a means to mediate conductivity or diffusion. Our interest in the design of covalently bound polymer multilayer assemblies is based on the ability of these materials to mediate interfacial adsorption and desorption phenomena. We have reported recently on design schemes and deposition strategies for a variety of alternating copolymers using covalent interlayer attachment chemistry.”57 We observed that the addition of a small amount of concentrated sulfuric acid during ether crosslinking between polymer adlayers led to a marked increase in adlayer absorbance compared to the results for adlayers where no H2804 was added.56 We have undertaken a study to understand this effect more fully and report our findings here. In addition to the acid-mediated enhancement in the ellipsometric thickness of the polymer adlayers, we observe a significant increase in absorbance per layer when interlayer linking chemistry is performed in the presence of H2S04. The thickness measurements taken by themselves could be indicative of either an increase in amount of material adsorbed or simply an increase in the extent of structural disorder within the layers. When these findings are viewed in light of the absorbance data, it is clear that the change we observe is an increase in adlayer density, and not simply disorder. The enhancement of adlayer deposition with the addition of acid is not the same for all crosslinking reactions. To address the comparative importance of acid catalysis and dehydration in these reactions, we have studied adlayer deposition and crosslinking -84- ‘."I using H2804 and HCl. HCl enhances acid-catalyzed reactions, but cannot play a role in dehydration. The presence of H2S04 in these systems could drive these reactions toward product formation according to LeChatelier’s principle. In the ester and amide formation reactions, adipoyl chloride is used as the crosslinking agent. In these cases, it is likely that the reaction proceeds via the elimination of HCl, calling our assertion of adlayer thickness enhancement by dehydration into question. To better understand this reaction, we have prepared a separate substrate in a manner consistent with experimental conditions used in polymer adlayer deposition, except we omitted the polymer layer. We found that the surface acid chloride groups were converted to the corresponding carboxylic acid almost exclusively, indicating that the amide and ester formation reactions were between the amine or alcohol and a carboxylic acid, consistent with the explanations we offer for the enhanced deposition. The urea and urethane formation reactions proceed with the formation of a carbamic acid intermediate. The presence of water in the system would lead to decomposition to an amine and C02. In this case, the presence of H2804 leads to the removal of adventitious water in the reaction medium. In all cases, we observed measurably greater deposition of polymer with the addition of sulfuric acid than with the addition of hydrochloric acid, suggesting that, for these reactions, dehydration and protonation are both important, but dehydration plays a dominant role in mediating the reaction efficiency. We have also examined polymer adlayer deposition as a function of acid concentration. We assembled ester-linked multilayers of poly(NPM-VOB) using adipoyl chloride as the crosslinker and varied the concentration of acid added to the reaction. The adlayer deposition process depends linearly on acid concentration to a point, with the -35- ..i-FIKIESE A. win-‘5 1.17.." growth appearing to saturate at high acid concentrations. We also note that there is a marked difference between the deposition of the initial polymer layer for the reactions performed with HCl compared to those done with H2804. We consider the similarities and differences found for the addition of the two acids during the adlayer deposition reaction and interpret these effects in terms of the reactions responsible for adlayer crosslinking. 4.2 Experimental Synthesis: We use the alternating copolymers poly(N-phenylmaleimide—co-1- vinyloxy-4-butanol), poly(NPM-VOB), poly(3-chlorophenylmaleimide-co-3- aminopropyl-l-vinylether), poly(3CPM—APVE) and poly(4-bromophenylmaleimide-co- 3-aminopropyl-l-vinylether), poly(4BPM-APVE), (scheme 4.1 top panel) in this study. The synthesis and layered interfacial assembly of these polymers has been detailed elsewhere.”57 These polymers are capable of participating in ester, urethane, ether (poly(NPM-VOB)) and amide and urea interlayer linkages (poly(3CPM-APVE) and poly(4BPM-APVE)). Substrate preparation: Gold coated substrates and silicon and quartz wafers were cleaned using piranha solution (3:1 HZSO4zH202) for ~15 minutes. These substrates were then removed and rinsed with ethanol, then water, and dried under a stream of nitrogen. Silicon and quartz substrates were used in subsequent adlayer deposition reactions without further modification. For these substrates, the surface silanol groups are used directly in adlayer deposition reactions. Prior to adlayer deposition, gold substrates were exposed to 0.01 M 6-mercapto-1-hexanol in 60:40 ethanolezO solution at 40°C for ~30 -86- rrrinutes. Following activation of the surface with 6-mercapto-l-hexanol, the terminal hydroxyl groups were used in adlayer deposition reactions. The substrates used for the formation of ester and amide linkages were first reacted with adipoyl chloride in the presence of 4-methylmorpholine for 30 minutes at 35°C. The substrates used for the formation of urea and urethane linkages were reacted with 1,6-diisocyanatohexane in the presence of 4—methylmorpholine. Following surface modification, the substrates were exposed to approximately 5 mL of the appropriate polymer solution at a concentration of 10 mM in reagent grade dimethyl sulfoxide. For reactions where acids were added, the concentration range examined was between 0.073 M and 0.290 M for HCl and between 0.108 M and 0.432 M for H2S04 (Table 4.1). It should be noted that these are the final concentrations after addition of the acid to the polymer solutions. Optical Ellipsometry: Ellipsometric thickness data on the deposited polymer multilayer assemblies were acquired using a Rudolph Auto-EL II optical ellipsometer operating at 632.8 nm. The software used for data collection and processing was acquired from the manufacturer (Rudolph DAFIBM). The refractive index of the polymer adlayers was taken to be n = 1.54 + 0i in all cases. The thickness of the polymer multilayer assembly was measured after each deposition cycle. Regression of the ellipsometric data was performed using Microcal Origin® V.6.0 software. UV—Visible Spectroscopy: Absorption spectra of the polymer adlayers were acquired using a Cary 300 UV-visible spectrophotometer. For these measurements, the adlayers were deposited on quartz substrates and absorption spectra acquired after each deposition cycle. Data were collected between 190 nm and 500nm at a scan rate of 600 nm/min with 1 nm spectral resolution. -87- puma-E] .7 NHz NH2 0 O O O N O O N O O N O o l O O O . . . l? Cl l4 Br ‘ O O /U\ + R'NHZ ——’ /U\ ,R' + H20 3 R OH R 171 H 11 ° b + R'OH -——> /U\ .+ H0 R OH R Q’R 2 O )L C R—NZCZO + R'NHz ——> R\N N’R I l H H O ' d R—N=C=o + R’OH ———> R\N/U\O/R l H Scheme 4.1. Top panel, left to right: Structures of poly(3CPM-APVE), Poly(NPM- VOB) and poly(4BPM-APVE). Bottom panel: Reaction schemes for the several crosslinking strategies employed in this work. -88- Infrared Spectroscopy: FTIR spectra of the polymer adlayers on gold substrates were collected using a N icolet Magna 750 FTIR spectrometer. Spectral resolution was 4 cm'1 for all measurements. An external reflectance sample mount set at an incidence angle of 80° was used for data collection. Spectra of samples assembled on silicon were acquired using a Nicolet Magna-IR 550 set up to acquire data in transmission mode at Brewster’s angle. 4.3 Results and Discussion We report here on the role(s) of HCl and H2804 in mediating the crosslinking reactions used in the deposition of covalently linked maleimide-vinyl ether alternating “'57 The addition of a small amount of acid to the reaction vessel copolymer multilayers. mediates the deposition of polymer adlayers. We have investigated the role of acid in the mediation of amide, ester, urea and urethane bond formation, and we observe adlayer deposition enhancement in each case. Both spectroscopic and ellipsometric data point to the acid-enhancement of adlayer thickness, by a factor of between two and three, depending on the functionality being formed. The chemistry we use for adlayer growth is deliberately self-terminating at each layer, and the addition of subsequent layers requires an activation step. For the cases where the thickness and absorbance both increase as a result of adding acid to the interlayer linking reaction, the data are the result of an increase in the density of the polymer adlayers. We consider the results for specific systems next. The uv-visible and ellipsometric data for each system are summarized in Tables 2 and 3. -89- v, Y‘-‘-’T"‘"_‘.. . - . 2‘ Amide crosslinked polymer multilayers: Poly(3CPM-APVE) is bound to the substrates we use here by first reacting the surface functionality with adipoyl chloride, then reacting the polymer amine functionalities with terminal acid chlorides. Spectroscopic and ellipsometric data for this system are shown in Figure 1. Multilayers of these amine-terminated polymers can be assembled using adipoyl chloride as the crosslinking agent to form amide interlayer bonds (scheme 4.1a). This reaction is performed under anhydrous conditions, proceeds with or without the addition of acid and, for both experimental conditions, we obtain reproducible behavior. Ellipsometry shows regular layer growth with a slope of 19 Allayer, and uv-visible absorbance spectroscopy shows a linear increase in adlayer absorbance with a slope of 0.019 a.u.llayer. The addition of concentrated H2804 to the reaction vessel during layer deposition gives rise to a measurable change in the layer growth characteristics. In the presence of 0.216 M HZSO4, we recover ellipsometric data showing 30 Allayer and uv-visible absorbance data showing 0.042 a.u.llayer. As discussed above, the adlayer deposition chemistry we use ensures the deposition of individual polymer layers. Because there is an increase in both ellipsometric thickness and absorbance per layer with the addition of HZSO4 to the layer deposition reaction, the data indicate an increase in adlayer density. In an attempt to resolve whether or not protonation of a reaction intermediate accounts for our findings, we have repeated these experiments with the addition of HCl instead of H2504. In the presence of 0.145 M HCl during the adlayer crosslinking reaction also gives rise to enhanced adlayer growth, with uv-visible absorbance spectroscopy yielding a slope of 0.0294 a.u.llayer, intermediate between the anhydrous and HZSO4-mediated reactions. We were unable to acquire ellipsometry data for the HCl- -90- —-— vim-mtmal mediated reaction because HCl damaged the gold substrate surface, likely due to the destruction of the chromium adhesion layer. The FTIR spectrum of the crosslinked adlayers is shown in Figure 4.1. We observe the stretching resonances characteristic of amides and free NH functionalities at approximately 1655 cm'1 (amide I) and 3300 cm], respectively, and the bands are identical for adlayers grown without acid and with H2804 and HCl. In all cases, the reaction products are the same. It appears to be counterintuitive that the addition of HCl (or H2804 for that matter) would enhance the deposition due to the fact that we use adipoyl chloride as crosslinking agent, which generates HCl as a reaction product. In order to resolve this issue, we prepared a substrate under similar experimental conditions to those for polymer deposition and observed that the acid chloride groups were almost exclusively converted to the corresponding acid (Figure 4.2). Thus, the layer deposition reaction is between an amine and a carboxylic acid, and the explanations of dehydration and protonation for acid addition follow logically. We believe the mechanism of the reaction to be simple nucleophilic substitution by addition-elimination, where the reaction proceeds by attack of the acid from the polymer nucleophilic terminal amine, with the elimination of H20. -91- w 1-"! at!" up :1 .0 w E m 0.20 -§ 0'“ h z 0.1» ”“015 - l g _ 0900 250 300 350 400 <1“ wavelength (nm) (Cl/0.10 - < . 0.05 - 0.00' . . . ' 4 ' ' m .. 1 2 3 4 8 125 .. layer number — g 100. ,o 75 - _ g .. E 50 - * 8 I- .9 25 '- E 0 h l n l n I 1 l _ 1 2 3 4 .4" 3.5 :- layer number 2 3.0 _- g 2.5 _- $2: .0 ' . g 1.0 :- < 0.0 ‘ 4 1 . L .mwr‘“ ‘. 1 . 1 . 1 4000 3500 3000 2500 2000 1500 1000 frequency (cm'l) Figure 4.1. Data for amide-linked poly(3CPM-APVE). (a) Dependence of optical absorption at 245 nm on number of polymer adlayers. I are data points for acid-free growth, 0 are data for adlayer growth with 0.145 M HCl and A are data for adlayer growth with 0.216 M H2804. Inset: UV-Visible absorption spectrum for a four layer assembly of poly(3CPM-APVE). (b) Ellipsometric thickness of poly(3CPM-APVE) adlayers as a function number of layers deposited for growth without acid added (I) and with 0.216 M H2804 added (A). (c) FTIR spectrum of a two-layer assembly of poly(3CPM-APVE). -92- 1.4 - 1.2 - ,- 1.0 - MS? r 5 0.8 — 8 . g 0.6 - 35 r g 0.4 :- 0.2 0.0 L 3500 3000 2500 2000 1500 1000 freqency (cm'l) Figure 4.2. FT IR spectrum of substrate treated with adipoyl chloride showing essentially complete conversion of the terminal acid chlorides to carboxylic acids. Ester crosslinked polymer multilayers: The attachment of poly(NPM-VOB) to the substrate and subsequent adlayer deposition proceeds in the same manner as described above for poly(3CPM-AVPE), with poly(NPM-VOB) reacting with adipoyl chloride to produce an ester moiety (scheme 4.1b). Spectral and ellipsometric data are shown in Figure 4.3 a and b. For ester formation, we observe enhanced adlayer density when either H2804 or HCl are added to the reaction. For the deposition of the polymer layer in the absence of acid, we recover a growth rate of 22 A/layer from ellipsometric data, 34 Allayer for growth when the HCl concentration is 0.145 M, and 45 A/layer for growth when the final H2S04 concentration is 0.216 M For adlayers formed either -93- without addition of acid or with H2SO4 addition, UV-visible absorbance results are consonant with the ellipsometric data, yielding 0.0309 a.u.llayer at 220 nm for acid-free growth, and 0.0603 a.u.llayer for 0.216 M H2SO4-mediated growth. The absorbance data for adlayers grown in the presence of 0.145 M HCl yield a slope of 0.0303 a.u./layer, the same to within the experimental uncertainty as that measured for acid-free growth. This finding, in conjunction with the ellipsometry data, indicates that the adlayer density does not change for ester formation, but the polymer morphology does change upon adlayer exposure to HCl. The experimental data for ester formation suggest that it is not acid catalysis, but rather dehydration that dominates the ester formation reaction. The function of H2804 here is to sequester H2O as it is formed during the reaction. We note that this reaction is somewhat more difficult to control than the other reactions presented here. In some instances, it was difficult to deposit more than three molecular layers before large deviations from linearity were noticeable. The FT IR spectrum of the ester- crosslinked polymer layers is shown in Figure 4.30. Our findings for the growth of ester- crosslinked adlayers with the addition of acid are consistent with dehydration and protonation as discussed above for the amide-crosslinked adlayers. We also examined the growth of ester-crosslinked polymer adlayers without adding acid and with the addition of morpholine. We observed no significant enhancement for this reaction, consistent with our assertion that the acid chloride groups are predominantly converted to the corresponding carboxylic acid (Figure 4.2). Had these groups been acid chlorides, the presence of the base would have led to an enhancement according to Le Chatelier’s principle. -94- 13. :95 “In-1 I "’8 0.4 3 0.25 - ° 03’ ,., b .502 0.20 " 2% 01 _ < 0.0». - L . . 0.15 - 200 250 300 350 400 l. wavelength (11m) F4 (0 0.10 - V I < 0.05 - P 0.00 1 2 3 4 0:5: 200 ' layer number 35’ 150 - i E . .- 3 100 - E ' / E 50 - 8 . g 0 l J I n l n l in" l 2 3 MA 2'0 P layer number o F E 1.5 :- D g 1.0 :- § 0.5 :- 2? 0.0 1 l 1 l 1 I n l n J n l 4000 3500 3000 2500 2000 1500 1000 frequency (cm'l) Figure 4.3. Data for ester-linked poly(NPM-VOB). (a) Dependence of optical absorption at 231 nm on number of polymer adlayers. I are data points for acid-free growth, 0 are data for adlayer growth with 0.145 M HCl and A are data for adlayer growth with 0.216 M H2504. Inset: UV-Visible absorption spectrum for a four layer assembly of poly(NPM-VOB). (b) Ellipsometric thickness of poly(NPM-VOB) adlayers as a function number of layers deposited for growth without acid added (I), with 0.145 M HCl (0) and with 0.216 M H2SO4 added (A). (c) FTIR spectrum of a two-layer assembly of poly(NPM-VOB). -95- — ‘3’“ .“'.‘.".'.'-"“1 . . -1 Urea crosslinked polymer multilayers: Poly(4BPM-APVE) adlayers are bound to substrates crosslinked by formation of a urea functionality (Scheme 4.1c). Spectral and ellipsometric data for this system are shown in Figure 4.4. The reaction to form urea linkages is between the polymer amine group and an isocyanate functionality, present in the adlayer structure using 1,6-diisocyantohexane. The function of 1,6- diisocyanatohexane in the formation of adlayer assemblies is analogous to adipoyl chloride in the reactions described above. When adlayers are grown under anhydrous conditions, we observe modest adlayer growth, with ellipsometry indicating a thickness of 10 Allayer and uv-visible absorbance at 238 nm yielding a linear dependence on adlayer deposition cycles with 0.0043 a.u.llayer. We note that the thicknesses and absorbance values we recover for this reaction scheme are lower that that recovered for acid-free ester and amide layer growth, indicating a relatively low polymer adlayer density under these reaction conditions. We observe adlayer density enhancement with the addition of HCl or H2SO4 to the crosslinking reaction, with ellipsometry indicating 39 Allayer for growth with 0.145 M HCl and 35 Allayer for growth with 0.216 M H2804. As with the ester formation, the polymer morphology does play some role in the adlayer growth, because the uv—visible absorption data yield slopes of 0.0052 a.u.llayer for 0.145 M HCl and 0.0102 a.u.llayer for 0.216 M H2SO4. Thus, while the ellipsometric data point to similar adlayer thicknesses, the absorbance data point to substantially different loading densities, even after the difference in acid concentration for the two reactions is taken into account. We have established that the reaction chemistry used for forming covalent multilayers using isocyanate/amine chemistry depends sensitively on the presence of -96- “—I‘l. xiii—m! water.‘6 Under anhydrous conditions, single layers can form in a regular manner and with the addition of water, adlayer growth becomes much less well controlled. The addition of concentrated H2SO4 or HCl to the reaction vessel during layer deposition appears to facilitate urea formation. We understand this effect in the context of H2804 removing adventitious water that would lead to the formation and subsequent degradation of the unstable carbamic acid functionalities, precluding their participation in urea formation. The isocyanate carbon is highly electrophilic, making it susceptible to attack by nucleophiles such as water and amines. Attack at this site by water leads to the formation of the carbamic acid, which is unstable. Reaction of the carbamic acid with an amine leads to the formation of a urea, and in the absence of acid this reaction appears to be inefficient. In these cases, the addition of acid to the reaction likely serves to protonate the isocyanate group, making it more electrophilic and increasing its reactivity toward nucleophiles such as alcohols and amines. The enhancement observed with H2SO4 is, again, probably due to its ability to sequester water that would lead to the degradation of the carbamic acid. As noted above, the enhancement in absorbance data is almost a factor of two greater for H2SO4 addition than for HCl addition, even though the concentration of H2804 is only 1.5 times greater than that of HCl. In an anhydrous environment, we expect H2SO4 to function predominantly as a monoprotic acid. -97- .1 ) 3’. «(L11 m 0'10 - E 0.10 0.08 -§ 0,, - s :30'06 '< 00900 250 300 350 400 a " wavelength (nm) _ V0.04 - <2 . ’./_//_' 0 02 P 4/4 * 000 l' I . I_ I L I I .2 1 2 3 4 E: 160 - layer number 8 - i g 120 r- .o 80 ' g . E 40 - o . E: O I 4 I n I n I E 1 l 2 b 3 4 m.“ 4.0 _ ayer num er 2 5 3.0 8 g 2.0 .0 § 1.0 .D < .9 o I a I l I I I. I n I 4000 3500 3000 2500 2000 1500 1000 frequency (cm") Figure 4.4. Data for urea-linked poly(4BPM-APVE). (a) Dependence of optical absorption at 255 nm on number of polymer adlayers. I are data points for acid-free growth, 0 are data for adlayer growth with 0.145 M HCl and A are data for adlayer growth with 0.216 M H2804. Inset: UV-Visible absorption spectrum for a four layer assembly of poly(4BPM-APVE). (b) Ellipsometric thickness of poly(4BPM-APVE) adlayers as a function number of layers deposited for growth without acid added (I), with 0.145 M HCl (0) and with 0.216 M H2504 added (A). (0) FTIR spectrum of a two- layer assembly of poly(4BPM-APVE). -98- Urethane crosslinked multilayers: The reaction of poly (NPM-VOB) with 1,6- diisocyanatohexane produces urethane interlayer linkages (Scheme 4.1d). The substrate, containing terminal -OH functionality, is exposed to 1,6-diisocyanatohexane in the presence of methylmorpholine in anhydrous acetonitrile to produce an isocyanate- terrninated surface. When depositing poly(NPM-VOB) on this activated substrate, we observe an enhancement in layer density when the adlayer deposition reaction is performed with the addition of HCl or H2804. The ellipsometric data yield a slope of 12 Allayer for acid-free deposition, 20 Allayer in the presence of 0.145 M HCl and 32 Allayer in the presence of 0.216 M H2804. UV-visible absorbance data of the polymer adlayers shows a similar trend, with a slope of 0.0093 a.u.llayer for acid-free growth, 0.0265 a.u.llayer for HCl-mediated adlayer deposition and 0.0348 a.u.llayer for H2804- mediated deposition. This reaction is qualitatively similar to that used for urea interlayer formation, save for the formation of a urethane moiety. In contrast to the urea interlayer linking chemistry, both the ellipsometry and absorbance data point to an adlayer density enhancement that is greater for H2SO4 than for HCl. Similar to acid-free deposition for the urea crosslinked adlayers, we recover modest adlayer growth, indicating that the surface reactivity of isocyanate functionalities is not as facile as that of acid chlorides. As discussed above, the reaction of isocyanates with nucleophiles such as amines and alcohols is susceptible to competitive reaction with water, resulting in the presence of carbamic acid terminal functionalities, which would likely decompose to amines. However, in the presence of concentrated acid we observe an increase in the ellipsometric thickness. In the case of concentrated H2804, the fact that this species is an efficient dehydrating agent minimizes the competitive effects of water, allowing for the efficient -99- .A" 1:“!- formation of the urethane moiety. In the case of hydrochloric acid, we would again expect to see an enhancement due to acid catalysis either in protonation of the imine or in stabilization of an alkoxide reaction intermediate. The effect of adding H2804 or HCl to the crosslinking reactions used in the formation of several polymer adlayer structures is, in general, to increase either the density and/or disorder of the material deposited at the interface. The crosslinking reactions we use are simple addition-elimination reactions that rely on nucleophilic attack. These reactions are typically efficient and in many cases can be catalyzed by the presence of acid, or driven to completion by the abstraction of water. Both H2804 and HCl can serve as a source of protons while H2SO4 can also function as a dehydrating agent. Thus, for reactions that are enhanced more by the presence of H2804 than HCl, such as ester and urea formation, it is reasonable to expect both acid catalysis and dehydration contribute to our experimental findings. Because both amide and ester reactions proceed with the elimination of water, the dehydrating capability of H2SO4 plays a major role in these reactions. While the formation of urea and urethane linkages does not proceed by water elimination, it should be noted that the intermediate (carbamic acid) formed in these reactions is susceptible to nucleophilic attack by water, leading to the decomposition of this species. In this case, the ability of H2804 to remove water would lead to enhancement in the deposition of these polymers. We can view these findings in the context of Le Chatelier’s principle, where the removal of either reaction products (H2O for these reactions) would tend to drive the crosslinking reactions to completion. These findings thus provide a simple means of controlling the density of covalently linked polymer adlayers at a variety of interfaces. -100- ‘1 are??? -_._...‘ - _.._..q - u“: 0.20» g; 0.15» 0-15 “g 0.10» E E 0.05» < . .4010 l- 0'0 00 250 300 350 400 a + wavelength (nm) <20.05 - 0.00 I I I I I I I 0;; 1 2 3 4 ‘0'; 150 .. layer number 8 g I g 100 - .0 I. .1. .5. E 50 - * 8 .9 ' a O I I I I I J I 1 2 3 4 MA 2'0 F layer number i 1.5 r- § 1.0 - B . ‘5 0.5 - 3 . < 0.0 I I I L I I I I I I I 4000 3500 3000 2500 2000 1500 1000 frequency (cm'l) Figure 4.5. Data for urethane crosslinked poly(NPM-VOB). (a) Dependence of optical absorption at 231 nm on number of polymer adlayers. I are data points for acid-free growth, 0 are data for adlayer growth with 0.145 M HCl and A are data for adlayer growth with 0.216 M H2SO4. Inset: UV-Visible absorption spectrum for a four layer assembly of poly(NPM-VOB). (b) Ellipsometric thickness of poly(NPM-VOB) adlayers as a function number of layers deposited for growth without acid added (I), with 0.145 M HCl (0) and with 0.216 M H2804 added (A). (0) FTIR spectrum of a two-layer assembly of poly(NPM-VOB). -101- We have also measured adlayer deposition density as a function of acid concentration. In this work, we have assembled ester—linked multilayer systems of poly(NPM-VOB) using adipoyl chloride as the crosslinking agent, where we varied the concentration of HCl or H2SO4 as a function of deposition cycle. We performed these reactions with HCl concentrations in the range of 0.073 to 0.290 M and with H2804 concentrations in the range of 0.108 to 0.432 M. We find that, for both HCl and H2804, increasing acid concentration initially yields an increase in the slopes of the absorbance data, and for the highest concentrations of acid used here, we observe a modest decrease in the slope of the absorption data (Table 4.1). These data suggest that there is an optimal pH range for the deposition reactions, indicative of a balance between multiple phenomena which contribute to the adlayer deposition process. For both acids, we observe an enhancement in the thickness of the initial layer that is related linearly to acid concentration, with H2804 producing thicker initial adlayers than HCl (Figure 4.6). In each case, the growth of the initial polymer adlayer is enhanced to a greater extent than the growth of subsequent adlayers. We understand this effect in the context of the greater availability of the reactive groups on the substrate surfaces and possibly the reactivity of the surface-bound functionalities being greater than the functionalities contained in the crosslinking agents. This latter issue remains open to investigation but, from a phenomenological point of view, the experimental acid concentration-dependence of adlayer growth demonstrates the ability to control polymer adlayer density by synthetic means. -102- 9 H N 1 “:11 p p—I O I 0.08 - m_\ “.- p E Absorbance maximum of first adlayer p o 8 8 I I I I I I I 0.1 0.2 0.3 0.4 0.5 acid concentration (M) p 8 .9 0 Figure 4.6. Dependence of ester-bound poly(NPM-VOB) monolayer absorbance at 231 nm on acid concentration for the initial polymer layer. (a) HCl added, (b) H2804 added. The different concentration dependencies for each acid indicates the importance of both protonation and dehydration in the ester formation reaction. —103- 4.4 Conclusions We have demonstrated the ability to control the density of polymer adlayer deposition at interfaces using a simple modification to the crosslinking reaction(s) used to connect adlayers. This new capability provides three distinct levels of control over interfacial adlayer preparation. We have control over the identity of the functional moieties incorporated into the polymer used in adlayer deposition by synthetic means. This level of control is important in the preparation of materials where specific chemical functionalities are required. We have also demonstrated previously that polymer adlayers can be assembled in discrete layered structures where each layer is of molecular dimension. We can connect individual polymer adlayers to one another by means of either ionic complexation or the formation of covalent crosslinks between polymer side groups. The third level of control, which is the focus of the work we have reported here, is the ability to control the density of the polymer adlayers. This new capability bears directly on our ability to control mass transport properties within these ultrathin interfacial materials. The control that we have established over adlayer density is determined by the mechanism(s) of the crosslinking reactions used in multilayer construction. The addition of H2804 or HCl to the crosslinking reactions gives rise to an enhancement in the extent of reaction because of dehydration and/or acid catalysis, depending on the reaction used. For all of the systems examined here we recover significant enhancement in the absorbance as well as ellipsometric data for the polymer deposition reactions to which sulfuric acid is added, relative to the results for the corresponding acid-free reactions. The addition of HCl gives rise to smaller enhancements in thickness and, sometimes absorbance as well, suggesting that the -104- W ‘u‘ “—q presence of HCl influences the extent of disorder in the polymer adlayer as well as mediating the crosslinking reaction to adjacent layers. Because these adlayer deposition reactions are self-terminating at each step, the observed enhancements in ellipsometric thickness and absorbance are necessarily the result of control over the density of the polymer adlayer. We anticipate that the addition of H280; or HCl to the polymer adlayer crosslinking reactions will play a role in determining the interface morphology, and this is an issue that we intend to address. -105— mm “2.1 -m '91?qu Table 4.1. Slopes of the UV-Visible absorbance data for the addition of HCl and H2804 as a function of concentration for the crosslinking of poly(NPM-VOB) with adipoyl chloride to form an ester linkage. Volume of acid [HCl] Adlayer [H2SO4] (M) Adlayer added (mL) (M) absorbance absorbance (a.u.llayer) (a.u.Ilayer) 0.03 0.073 0.0135 0.108 0.0154 0.06 0.145 0.0157 0.216 0.0296 0.12 0.290 0.0097 0.432 0.0273 Table 4.2. Dependence of ellipsometric thickness on acid addition for amide, ester, urea and urethane crosslinked adlayers. Crosslink Without acid With 0.073 M With 0.108 M functionality (Allayer) HC] (A/layer) H2804 (A/layer) Amide 19 30 Ester 22 34 45 Urea 10 35 39 Urethane 12 20 32 -106- “ s «‘57 ‘ 4.“ 7.19“? I Table 4.3. Dependence of absorbance data on acid addition for amide, ester, urea and urethane crosslinked polymer adlayers. Crosslink chromophore Am, Without With With acid 0.073M HCl 0.108 M (a.u.llayer) (a.u.llayer) H2S04 (a.u.llayer) o (:1 Amide 245nm 0.0188 0.0294 0.0424 0 o Ester m—Q 231nm 0.0309 0.0303 0.0603 0 o Urea IE“ O B" 255nm 0.0043 0.0052 0.0102 0 o Urethane m© 231nm 0.0093 0.0265 0.0348 0 -107- 1“: 'M‘nl 4.5 Literature Cited 1. Langmuir, I. J. Am. Chem. Soc. 1917,39, 1848. N . Blodgett, K. B. J. Am. Chem. Soc. 1935, 57, 1007. 3. Allara, D. L.; Nuzzo, R. G. Langmuir 1985, I, 45-52. 4. Allara, D. L.; Nuzzo, R. G. Langmuir 1985, I, 52-66. 5. Chidsey, C. E. D.; Porter, M. D.; Allara, D. L. J.Electrochem.Soc. 1986, I33, C130- C130. 6. Dubois, L. H.; Zegarski, B. R.; Nuzzo, R. G. J. Vac. Sci. & Technol. A-Vacuum Surfaces and Films 1987, 5, 634-635. 7. Dubois, L. H.; Zegarski, B. R.; Nuzzo, R. G. Proc.Nat. AcadSci. 1987, 84, 4739- 4742. 8. Nuzzo, R. G.; Fusco, F. A.; Allara, D. L. J. Am. Chem. Soc. 1987, 109, 2358-2368. 9. Nuzzo, R. G.; Zegarski, B. R.; Dubois, L. H. J. Am. Chem. Soc. 1987, 109, 733-740. 10. Porter, M. D.; Bright, T. B.; Allara, D. L.; Chidsey, C. E. D. J.Am. Chem. Soc. 1987, 109, 3559—3568. 11. Whitesides, G. M.; Troughton, E. B.; Bain, C.; Holmesfarley, S. R.; Wasserman, S. R.; Strong, L. H. J.Electrochem. Soc. 1987, 134, C1 lO-Cl 10. 12. Bain, C. D.; Whitesides, G. M. Science 1988, 240, 62-63. 13. Strong, L.; Whitesides, G. M. Langmuir 1988, 4, 546-558. 14. Sagiv, J. J. Am. Chem. Soc. 1980, 102, 92—98. 15. Kohli, P.; Blanchard, G. J. Langmuir 2000, 16, 8518-8524. 16. Kohli, P.; Blanchard, G. J. Langmuir 2000, 16, 4655-4661. 17. Zhang, Z. J.; Hu, R. S.; Liu, 2. F. Langmuir 2000, 16, 1158-1162. -108- w‘““r—r~q 18 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. Home, J. C.; Huang, Y.; Liu, G. Y.; Blanchard, G. J. J. Am. Chem. Soc. 1999, 121, 4419-4426. Home, J. C.; Blanchard, G. J. J. Am. Chem. Soc. 1999, 121 , 4427-4432. Kohli, P.; Blanchard, G. J. Langmuir 1999, 15, 1418-1422. Lee, H.; Kepley, L. J .; Hong, H. G.; Akhter, S.; Mallouk, T. E. J. Phys. Chem. 1988, 92, 2597-2601. Putvinski, T. M.; Schilling, M. L.; Katz, H. B.; Chidsey, C. E. D.; Mujsce, A. M.; Emerson, A. B. Langmuir 1990, 6, 1567-1571. Hong, H. G.; Sackett, D. D.; Mallouk, T. E. Chem. Mat. 1991, 3, 521-527. 'fu-muuq Katz, H. B.; Schilling, M. L.; Chidsey, C. E. D.; Putvinski, T. M.; Hutton, R. S. Chem. Mat. 1991, 3, 699-703. Vermeulen, L. A.; Thompson, M. E. Nature 1992, 358, 656—658. Katz, H. E.: Schilling, M. L. Chem. Mat. 1993, 5, 1162-1166. Schilling, M. L.; Katz, H. B.; Stein, S. M.; Shane, S. P.; Wilson, W. L.; Buratto, S.; Ungashe, S. B.; Taylor, G. N.; Putvinski, T. M.; Chidsey, C. E. D. Langmuir 1993, 9, 2156-2160. Vermeulen, L. A.; Snover, J. L.; Sapochak, L. S.; Thompson, M. E. J. Am. Chem. Soc. 1993, 115, 11767-11774. Katz, H. E. Chem. Mat. 1994, 6, 2227-2232. Katz, H. B.; Bent, S. P.; Wilson, W. L.; Schilling, M. L.; Ungashe, S. B. J. Am. Chem. Soc. 1994, 116, 6631-6635. Thompson, M. E. Chem. Mat. 1994, 6, 1168-1175. Vermeulen, L. A.; Thompson, M. E. Chem. Mat. 1994, 6, 77-81. -109- 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43 44 45 46 47 48 49 50 Hanken, H. G.; Corn, R. M. Anal. Chem. 1995, 67, 3767-3774. Hoekstra, K. J .; Bein, T. Chem. Mat. 1996, 8, 1865-1870. Dubas, S. T.; Farhat, T. R.; Schlenoff, J. B. J. Am. Chem. Soc. 2001, 123, 5368-5369. Dubas, S. T.; Schlenoff, J. B. Macromolecules 2001, 34, 3736-3740. Farhat, T. R.; Schlenoff, J. B. Langmuir 2001, I 7, 1184-1192. Schlenoff, J. B.; Dubas, S. T. Macromolecules 2001, 34, 592-598. Dubas, S. T.; Schlenoff, J. B. Macromolecules 1999, 32, 8153-8160. Graul, T. W.; Schlenoff, J. B. Anal. Chem. 1999, 71, 4007-4013. Schlenoff, J. B.; Ly, H.; Li, M. J. Am. Chem. Soc. 1998, 120, 7626-7634. Schlenoff, J. B.; Laurent, D.; Ly, H.; Stepp, J. Chem. Eng. Technol. 1998, 21, 757- 759. . 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. . Schlenoff, J. B.; Li, M. Ber. Bunsen-Ges. Phys. Chem. Chem. Phys. 1996, 100, 943- 947. . Kim, H. N .; Keller, S. W.; Mallouk, T. B.; Schmitt, J .; Decher, G. Chem. Mat. 1997, 9, 1414-1421. . Everhart, D. S. Chemtech 1999, 29, 30-37. . Dermody, D. L.; Peez, R. F.; Bergbreiter, D. E.; Crooks, R. M. Langmuir 1999, 15, 885-890. . Dermody, D. L.; Crooks, R. M.; Kim, T. J. Am. Chem. Soc. 1996, 118, 11912-11917. . Yang, H. C.; Dermody, D. L.; Xu, C. J.; Ricco, A. J.; Crooks, R. M. Langmuir 1996, 12, 726-735. -110- 51. Wells, M.; Dermody, D. L.; Yang, H. C.; Kim, T.; Crooks, R. M.; Ricco, A. J. Langmuir 1996, 12, 1989-1996. 52. Katz, H. E.; Scheller, G.; Putvinski, T. M.; Schilling, M. L.; Wilson, W. L.; Chidsey, C. E. D. Science 1991, 254, 1485-1487. 53. Katz, H. E.; Wilson, W. L.; Scheller, G. J. Am. Chem. Soc. 1994, 116, 6636-6640. 54. Batchelder, D. N .; Evans, 8. B.; Freeman, T. L.; Haussling, L.; Ringsdorf, H.; Wolf, H. J. Am. Chem. Soc. 1994, 116, 1050-1053. 55. Kohli, P.; Scranton, A. B.; Blanchard, G. J. Macromolecules 1998, 31, 5681-5689. 56. Major, J. S.; Blanchard, G. J. Chem. Mat. 2002, in press. 57. Major, J. S.; Blanchard, G. J. Chem. Mat. 2002, in press. —111- l- 41. 'u-no ;-.l-I..t1 m Chapter 5 Adsorption Behavior of Polymer Modified Interfaces Abstract We report on the vapor phase adsorption behavior of selected organic molecules onto/into polymer-modified interfaces. We use maleirrride-vinyl ether alternating copolymers in conjunction with layer growth chemistry to modify gold-coated substrates with mono- and multilayer polymer structures. The identities of the polymer side groups determine the properties of the polymer interface. We measure the mass uptake of the interface using a quartz crystal microbalance mounted in a chamber where the partial pressure of the organic solvent vapor can be controlled. An optical null ellipsometer is used to m0nitor interface thickness during the deposition process. We interpret the adsorption data in the context of the BET adsorption isotherm. Our results point to the adsorption process depending on both polymer identity and thickness, implying chemical control over the adsorption and desorption kinetics and thus the thermodynamics of the process. -112- i 77' 1___"‘ 5.1 Introduction The development of thin film chemistry over the past 15 years has enabled advances in several technologically important areas. Work in thin film materials design . . 4- 3 chermcal sensrng 7, and synthesis has benefited second harmonic generation," separations!“9 and molecular scale electronic device applications. The methods of choice for the controlled assembly of layered thin films include Langmuir-Blodgett deposition, 10-12 13-15 monolayer self-assembly, ionic and covalent layer growth.”18 The characterization of these materials has focused on layer thickness, morphology, wetability and constituent orientation. Among our reasons for interest in thin film materials is their potential for use in controlling chemical separations. We focus here on the thermodynamics and kinetics of interaction between vapor phase adsorbates and polymer interfacial layers. In addition to structural characterization by ellipsometry, and FTIR and UV-visible spectroscopies, we use an apparatus based on in situ quartz crystal microbalance nano-gravimetry and optical null ellipsometry that allows us to quantify the adsorption of organic vapor phase species onto/into polymer modified interfaces. The combination of in situ mass and thickness data on these layers allows measurement of the adsorption isotherm for these systems and modeling of the isotherm data provides direct information on the energetics of adsorption. In some cases, the kinetics of adsorption in these interfacial materials can be evaluated as well. For example, the net rate of adsorbate uptake by the polymer monolayers can range from ~1 ng/cmz-s to ~50 ng/cmz- 5, depending on the identity of the polymer. Whether this system-dependent variability represents a uniform change in both the adsorption and desorption rate constants or a —113- AM'-A-_m1'] I W l differential change in one relative to the other can be evaluated by measuring the adsorption isotherm behavior of these materials. We use maleirrride-vinyl ether alternating copolymers to construct layers, with covalent (amide or ester) or ionic (zirconium phosphonate) interlayer linkages to bond individual layers to one another. To elucidate the behavior of these polymer films in the presence of organic gas phase adsorbates, we monitor the thickness of the films in situ as a function of mass uptake using an optical null ellipsometer. The ellipsometric data point to the absence of measurable swelling of the polymer films, despite the functional form of the mass uptake data, which reveal substantial interaction between the polymer layer and the adsorbate. The experimental adsorption isotherm data can be understood in the context of the Brunauer-Emmett-Teller (BET) model.19 The enthalpy of desorption we recover from fits of the data to the BET isotherm model depends on both the chemical identity of the polymer layer(s) and on the adsorbate. The experimental values of AHdes all lie between 29 and 39 kJ/mol, indicating modest interactions between the adsorbate and the polymer-modified interface. While such energies fall at the border between what is considered to be cherrrisorption and physisorption, the reversibility of the adsorption process demonstrates physisorption to be the dominant mode of interaction. Our data indicate that, in some cases, capillary condensation occurs for high adsorbate vapor pressures, and this finding argues for the polymer layers being quite permeable. Of significance is that, in the case where we examine adsorption for monolayer and bilayer structures of the same polymer, we find the recovered AHdcs to change with number of polymer layers. This finding provides direct evidence for differential control over adsorption and desorption kinetics by means of interface structure. -114— v v. - .HIT‘T 01-001.? 1 5.2 Experimental The synthesis and layer-wise assembly of the several polymers we use in this 16"8 These materials were attached to the work has been described previously. (evaporated gold) surface of QCMs either covalently or ionically after the device surfaces had been modified with 6-mercapto-1-hexanol. Covalent interlayer linkages between polymer layers were formed using adipoyl chloride as a cross-linking agent to form ester or amide bonds. Ionic interlayer linkages were formed by ionic association between Zr4+ and phosphonate functionalities on polymer side groups. In addition to growth on QCM substrates, quartz slides were also used for layer assembly to allow characterization using UV-visible transmission spectroscopy. Substrates: The QCMs used here are AT-cut crystalline quartz with vapor- deposited gold electrodes, operating at a resonance frequency of ~6 MHz. The change in QCM oscillation frequency is related to adsorbate mass uptake through the Sauerbrey equation,20 = -2Am ( fo )2 n l A(flp)/2 where Af = the observed frequency change, Am = the mass change, f0 = the fundamental Af frequency of the crystal (~6 MHz), n = the harmonic of the fundamental frequency (n = 1 ' here), A = exposed area of the QCM (0.33 cmz), u = the shear modulus of quartz (2.947 x 101 I g cm'1 s") and p = the density of quartz (2.648 g cm'3). Substrate preparation: QCMs and quartz slides were cleaned using piranha solution (3:1 H2SO42H2O2) for ~15 rrrinutes. The QCMs were then rinsed with absolute ethanol, then water, and dried under a stream of nitrogen gas. The clean QCMs were exposed to a 10 mM solution of 6-mercapto-l-hexanol for approximately 30 minutes at -115- 5.1 mA-ww—gl 35°C. This reaction introduces terminal hydroxyl functionality to the QCM surface for subsequent polymer attachment. Following reaction with 6-mercapto-l-hexanol, the QCMs were rinsed with ethanol, then water, and dried under a stream of nitrogen gas. The substrates were then used in the polymer deposition reactions. Covalent QCM modification: QCMs with gold electrodes coated with o- mercapto-l-hexanol were exposed to 3% (v/v) 4-methylmorpholine and 3% (v/v) adipoyl chloride in anhydrous acetonitrile under an inert argon atmosphere for ~30 minutes at room temperature. The reacted substrates were then rinsed briefly with ethyl acetate, dried under a stream of nitrogen gas and exposed to a solution containing a maleimide- vinyl ether polymer synthesized with amine- or hydroxyl-terminated side groups. The reaction of the substrate acid chloride functionality with the polymer amine or hydroxyl side groups produces a layer of MVE polymer covalently bound to the QCM electrode, and we have described this chemistry in detail elsewhere.‘8 We determine the thickness of the resulting polymer layer using optical null ellipsometry. Phosphorylation and zirconation of QCMs: Phosphorylation of the QCMs was achieved by exposing the hydroxyl terrrrinated QCMs to 5% (v/v) collidine and 3.5% (v/v) phosphorus oxychloride in anhydrous acetonitrile under an argon atmosphere for 30 minutes at room temperature. These substrates were then rinsed with acetonitrile, acetone and water, and dried under a stream of nitrogen gas. Zirconation was accomplished by exposing the phosphorylated substrates to a 60:40 ethanolzwater solution of 10 r_n_M zirconyl chloride octahydrate at room temperature for ~30 minutes. The zirconated substrates were rinsed with ethanol and water, then dried using nitrogen. These substrates were then exposed to a solution of an MVE polymer containing -116- phosphonate-terminated side groups that had been partially deprotected for layer deposition, as described previously.‘6 Quartz and oxidized silicon substrates: Quartz and silicon substrates were prepared by first cleaning in piranha solution ~30 minutes followed by ethanol and water rinses and drying under a stream of nitrogen gas. The resulting surface was characterized by the presence of silanol groups that are amenable to ionic or covalent layer attachment chemistry. Preparation of these interfaces was similar to the methods described above except silanol groups were reacted directly with adipoyl chloride (covalent interlayer linkages) or POC13 (ionic interlayer linkages). The substrates were then reacted with the appropriate polymer solution under experimental conditions identical to those described above for QCM preparation. The quartz substrates were used to acquire UV-Visible spectra of the polymer films. The silicon and gold-coated (QCM) substrates were used for FTIR characterization of the polymer adlayers (not presented here). Adsorption isotherm measurements. The experimental apparatus used to acquire isotherm data was constructed in-house and has been described in detail elsewhere.2| This system consists of a QCM housed in an environmentally controlled stainless steel chamber. Control of the chamber environment is accomplished using a gas blending system that allows the partial pressure of adsorbate vapor to be varied in a predictable and reproducible manner. Quartz windows were mounted on the sides of the chamber to allow optical access to the QCM surface for ellipsometric measurements. ‘ The ellipsometer was assembled from discrete optical components and operated at 632.8 nm in the optical nulling mode. The temperature of the chamber, QCM and adsorbate reservoir was controlled with a circulating water bath (Neslab RTE-111). The frequency -117- of the QCM is proportional to the adsorbate mass uptake and the thickness of the polymer adlayer is monitored in situ and in real time using the ellipsometer. We detemrine the adsorbate volume from the experimental mass change data, assuming a bulk liquid density for the adsorbed species. Optical ellipsometry: The thickness of the polymer-modified interfaces prior to exposure to the vapor phase adsorbate was measured using a commercial optical ellipsometer (Rudolph Research AutoEL 11) operating at 632.8 nm. In all cases where comparisons could be made, the commercial ellipsometer and the home-built instrument yielded identical results to within the uncertainty of the measurement. Thickness change calculations for all ellipsometric measurements were performed using Rudolph software (DAFIBM). UV- Visible Spectroscopy: UV-visible spectroscopic data were acquired using a Cary 300 UV-visible spectrophotometer in the 190 — 500 nm range at a scan rate of 600 nm/min with 1 nm resolution. The software used to operate this instrument was provided with the instrument. 5.3 Results and Discussion We have assembled several mono- and multilayer structures using maleimide- vinyl ether alternating copolymers (Figure 5.1) and have characterized their adsorption behavior on exposure to methanol and hexane. The central issue in this work is achieving an understanding of how polymer adlayer identity and thickness influences the thermodynamics and kinetics of interaction with selected adsorbates. -118- "‘“ml ‘. t 2.. . I t 1,. s Figure 5.1 Structures of molecules bound to QCM surfaces. (a) 6-mercapto-l-hexanol, (b) poly(4HPM-VEP), (c) poly(4MPA-VEP) -119- we 1 n n O N O Q N O 0 : o : ‘c1 Br 0 o 9 o S s I D. . , . --------- ,I......-.,. ._..-..-. Figure 5.1 cont’d. Structure of molecules bound to QCM surfaces. (d) poly(4BPM- APVE), (e) poly(3CPM-APVE). -120- . “AAA—.1 To achieve this understanding we use adsorption isotherm measurements, with ellipsometry and steady state spectroscopy data being used to provide additional information on the materials. We use QCM gravimetry to relate the volume of the adsorbate taken up by thin polymer films and ellipsometry to measure any change in the thickness of the polymer layer as a result of adsorbate exposure. These data are of a functional form consistent with the BET adsorption isotherm, and fits of the data to this model allow us to extract information about the enthalpy of desorption (AHdes) of the adsorbate, the volume of adsorbate required to cover the interface one molecular layer thick (le) and, assuming a bulk liquid density of the adsorbate, the effective surface area '7... .-. ..- 2...? of the (polymer) adsorbent. The BET equation relates the adsorbate volume to its partial pressure according to V ,cz m V,,,, = g 5.2 ‘ (1-2){1-(1-C)z} where V422, is the volume of the adsorbate, Vm; is the monolayer volume, c is a term related to the energetics of adsorption and z is the normalized vapor phase concentration of the adsorbate. c = exp((AHd,, — AHW )/RT) 5 3 z = p/p' The quantity AHdes is the enthalpy of desorption of the adsorbate molecule from the surface, AHvap is the enthalpy of vaporization at the temperature of the experiment, p is the vapor pressure of the adsorbate and p* is its saturation vapor pressure. The term c represents the balance between the enthalpies of desorption and vaporization, and is thus related directly to the energy of interaction. We control the value of z experimentally by varying the volume ratio of dry nitrogen and “wet” nitrogen using the gas blending -121- system. The “wet” nitrogen is saturated with adsorbate by passing dry nitrogen through a gas-washing bottle filled with the adsorbate liquid. The ratio of these two gas streams is controlled and introduced to the chamber containing the QCM. We use methanol and hexane as adsorbates in this report. We consider the behavior of each interface individually. Before discussing the experimental data, a word is in order about our choice of the BET adsorption isotherm model. While there are a variety of adsorption isotherm models, we have found that the BET isotherm is sufficiently general to allow the range of experimental data we present here to be considered within the framework of a single model. Brunauer22 considered that there are five main categories of adsorption isotherms, ranging from the Langmuir isotherm (Type I) for monolayer adsorption to systems that exhibit capillary condensation at high adsorbate concentrations (Types IV and V). Isotherms classified as Types II and III are characterized by multilayer growth that, at high adsorbate concentrations, are consistent with bulk liquid formation at the surface. In effect, the difference between bulk liquid condensation and capillary condensation lies in the porosity of the interface under examination. For the interfaces we examine here, those characterized by relatively little porosity (e. g. clean gold) exhibit Type II or Type III behavior while the layered polymer interfaces are characterized by Type IV or Type V saturation effects. Not surprisingly, the thiol-modified gold interface manifests intermediate behavior. Uncoated gold QCM: The dependence of the adsorbate volume on partial pressure for an uncoated gold QCM substrate is shown in Figures 5.2. By fitting the data to Eq. 2 and using published values of AHvap for the adsorbate,23 we recover the enthalpy -122- of desorption, AHdcs = 37.6i0.l kJ/mol for methanol, with the volume of a monolayer le = 38i2 pL. For hexane, we measure AHdcs = 29.2:03 kJ/mol and le = 179i35 pL. From these monolayer volume data and with information on the QCM surface area (0.33 cmz),2| and the surface roughness factor (3.7)24 for evaporated gold, we can estimate the thickness of the adlayers. For methanol, we calculate a monolayer thickness of 3.1 A and for hexane we calculate a thickness of 14.6 1- 3 A. Molecular mechanics calculations predict ~3 A for methanol and 8.1 A for all-trans hexane. The recovered estimates for monolayer thickness are clearly not precise, but they are in reasonable agreement with expectations, given the nature of the measurements. The functionality of the adsorption isotherm is most consistent with a BET Type H or Type 111 model for both adsorbates (Figures 5.2a and 5.2b). One interpretation of such an isotherm is that the surface is porous, but that is clearly not applicable for this system. We note that the isotherm is fitted by the BET model for z < 0.6, so determination of Type II (or Type IH) behavior, which is most strongly characterized for high z, is somewhat speculative. For adsorption onto gold, no distinct monolayer region is observed, likely because of the fluxional and associative nature of the adsorbates. The onset of bulk liquid condensation at high 2 is indicated by the continued increase in Vads. Our data indicate that gold interacts more strongly with methanol than with hexane, consistent with the well-established hydrophilic nature of clean gold. -l23- 300 250' 200* 150 100- r SOLM 1 0'- 0.0 ‘ 0.2 ‘ 0.4 I 0.6 ' 0.8 A 1.0 b Z=P/P* 300. 250_ 200 150 100 I 1 t——o——t V 1' I Vads (pL) rm F—O—d I 1 I V8,, (pL) I L I A l l J A I A I 0.0 0.2 0.4 0.6 0.8 1.0 2 = p/p* Figure 5.2. Adsorption isotherm data for uncoated gold QCM. (a) Data and best-fit line for adsorption of methanol and (b) data and best-fit line for adsorption of hexane. The data were fitted to the BET model as given in Eqs. 2 and 3. Results of the fits are presented in Table 1. -124- 6-Mercapto-1-hexanol modified QCM: The adsorption isotherms of methanol and hexane on a QCM modified with a monolayer of 6-mercapto-1-hexanol (Figure 5.13) reveal substantially different behavior than that seen for uncoated gold. The data shown in Figures 5.3 (a and b) reveal a significant increase in the volume of a monolayer for methanol, le = 650 i 55 pL as well as for hexane, th = 896 i 46 pL. In effect, the addition of the 6-mercapto-l-hexanol layer has increased the “surface area” of the interface available for interaction with methanol by a factor of 17 and for hexane by a factor of 5. The implication of this experimental observation is that there is substantial penetration of the adsorbate into the thiol monolayer and, surprisingly, this effect is more pronounced for methanol than it is for hexane. We also observe changes in the enthalpies of desorption for the modified interface. For methanol vapor, we recover AHdeS = 35.4 i 0.3 kJ/mol and for hexane we recover AHdcs = 32.9 i 0.2 kJ/mol. The strength of interaction between methanol and 6-mercapto-l-hexanol is less than between methanol and gold, and the opposite is true for interactions with hexane. This is not an unexpected result based on simple polarity arguments. It is likely that the short reaction time used for deposition of 6-mercapto-l-hexanol yields a monolayer with significant structural disorder. A well-ordered surface would present a relatively dense array of hydroxyl groups at the surface, which would give rise to more favorable interactions with methanol than with hexane. The increased energy of interaction observed for the hexane indicates adsorbate interactions with the monolayer methylene groups, a condition that can exist only in the limit of a disordered monolayer. -125- p—a O O O I 0.8 10 l - ' 100m ' 0.6 T 0.8 ' 1.0 2 = p/p* Figure 5.3. Adsorption isotherm data for a QCM modified with 6-mercapto-1-hexanol. (a) Data and best-fit line for adsorption of methanol and (b) data and best-fit line for adsorption of hexane. The data were fitted to the BET model as given in Eqs.2 and 3. Results of the fits are presented in Table 5.1. -126- ”mtg-"a Poly(4HPM-VEP) modified QCM: Poly(4-hydroxyphenylmaleirrride-vinyl ether diisopropyl phosphonate) was attached covalently to QCM surfaces modified with 6- mercapto-l-hexanol using surface-bound adipoyl chloride to react with the terminal -OH group of 4-hydroxyphenylmaleimide (Figure 5.1b). The behavior of this polymer toward methanol and hexane adsorbates is strikingly different (Figure 5.4). For methanol adsorption we recover le = 894 i 40 pL and AHdes = 36.0 i 0.2 kJ/mol, corresponding to a factor of 23 increase in effective surface area relative to clean gold. These results, over the region where the BET isotherm is applicable, point to relatively high permeability of the polymer and modest strength of interaction with methanol. Consistent with these findings is the observation that the equilibration required for a monolayer of poly(4I-IPM-VEP) is fast relative to many of the other polymers we report here. For high adsorbate partial pressures, we observe a saturation phenomenon, which is typically taken to indicate the onset of capillary condensation within the polymer layer. The single poly(4HPM-VEP) layer, where no change of ellipsometric thickness is seen upon methanol adsorption, appears to be a highly porous entity with a relatively large internal surface area. The onset of capillary condensation at high adsorbate partial pressures for such a structure is not surprising owing to the propensity of the adsorbate to hydrogen-bond. The behavior of this polymer layer upon exposure to hexane is fundamentally different. There is a small adlayer volume for low adsorbate partial pressures, with a saturation seen for 2: ~ 0.4. Above z ~ 0.5, the QCM fails to resonate, an indication of bulk liquid condensation on the surface. This is not surprising since the phosphonate groups will likely be near the surface, making interactions with nonpolar adsorbates relatively weak. The picture that emerges for interactions with hexane is that -127- Ian-nit 1m?“— *3 I l this adsorbate cannot penetrate the polymer surface substantially, but condenses on the surface. The average dimension of the pores formed by the polymer layer may be such that adsorption of methanol into the matrix is favorable while hexane is too large to penetrate the matrix substantially. For methanol, the value of le for this interface is large compared to that for poly(3CPM-APVE) and poly(4BPM-APVE), (vide infra), likely due to the presence of the phosphonate groups with which the methanol can interact. Poly(4MPA-VEP) modified QCM: Poly(3,3’-methylphenylazomaleimide-vinyl ether phosphonate) was bound to the QCM substrate by ionic association chemistry. The QCM was modified with 6-mercapto-l-hexanol and the terminal hydroxyl groups were reacted with POC13, then water to produce a phosphate surface. Exposure to Zr4+ then the polymer produced the modified surface (Figure 5.10). Our intent in using this interface was to design a material that is highly porous, presenting a large surface area. Given the rigidity of the azobenzene moiety, we expect the polymer layer to possess significant void volume and that the strongest interactions should be with a relatively non-polar adsorbate. The absorption spectrum of this polymer shows that both cis and trans azobenzene conformers are present (Figure 5.5a), with the ratio being consistent with that reported previously for multilayers of the azobenzene-substituted MVE alternating copolymer.25 The UV-visible spectrum thus points to a relatively open polymer structure. Exposure of the poly(4MPA-VEP) interface to methanol vapor produces an isotherm that, when fit to the BET model, gives a best-fit le = 744 i 12 pL and AHdes = 35.6 i 0.1 kJ/mol (Figure 5.5b). These results are consistent with several of the other polymer interfacial structures we present here, indicating significant permeability of the -128- polymer matrix to methanol despite the presence of the nonpolar dimethylazobenzene side groups. The phosphonate functionality within the polymer layer is a likely site for methanol interaction, as suggested above for poly(4HPM-VEP). Exposure of the poly(4- MPA-VEP) interface to vapor phase hexane yields an adsorption isotherm which gives best-fit values of le = 4752 i 68 pL and AHdcs = 29.1 i 0.1 kJ/mol (Figure 5b). We recover a Type V isotherm for exposure of poly(4MPA-VEP) to methanol, with the data being less amenable to modeling for hexane. For hexane exposure, we were unable to collect data beyond z ~ 0.4, at which point the QCM ceases to oscillate. This is a reversible condition; when we reduce 2, we recover the original resonant frequency of the poly(4MPA-VEP)-modified QCM. Such behavior is taken to be consistent with surface wetting, i.e., capillary condensation and/or bulk condensation and is consistent with the high value of le. The confirmation of capillary condensation could be made by measuring hysteresis in the isotherm. We were unable to acquire hysteresis data due to the very sharp dependence of Vad, on z. Despite the limits on our ability to apply the BET model to these data, it is clear that adsorption of hexane is efficient for this polymer adlayer, in agreement with our assertion that providing a relatively nonpolar and open structure will favor adsorption of nonpolar adsorbates. -129- 1500 1000 V,,, (pL) 500 0.8 1 1.0 .5 .8 V... (PL) 8 O I o e o I I I I I I I I I I I 0.0 0.2 0.4 0.6 0.8 1.0 z=p/p* Figure 5.4. Adsorption isotherm data for a QCM modified with poly(4HPM-VEP). (a) Data and best-fit line for adsorption of methanol and (b) data and best-fit line for adsorption of hexane. The data were fitted to the BET model as given in Eqs. 2 and 3. Results of the fits are presented in Table 5.1 -130- I 1'}.- "IA-un' 1. a 0.12 I 0.10 0.08 l 0.06 - 0.04 - 0.02 absorbance (a.u.) I 'I' 350 ‘ 400 450 A 0.00 . r . . 200 250 300 h _ wavelength (nm) 800} 600} .O . ° 3 ; - E: 400 - 2.8. > 200} 0;] I .I I I I I I I I I 0.0 0.2 0.4 0.6 0.8 1.0 z=p/p* Figure 5.5. Absorption spectrum of poly(4MPA-VEP) on a quartz substrate. The band centered near 325 nm is the S2t—So transition for the trans side groups and the band centered near 240nm is the S2t—So transition for the cis side groups. (b) Data and best-fit line for adsorption of methanol (0) and hexane (O). The data were fitted to the BET model as given in Eqs. 2 and 3. Results of the fits are presented in Table 5.1. -131- IA.“ 6 I .‘l‘: ‘ Poly(4BPM-APVE) modified QCM: Poly(4-bromophenylmaleimide-arrrinopropyl vinyl ether) was bound covalently to the QCM by formation of an amide bond to an underlayer of adipoyl chloride that had been bound to the 6-mercapto-1-hexanol (Figure 1d). Exposure of this interface to methanol yields an adsorption isotherm characterized by th = 425 i 72 pL and AHdcs = 37.1 i 1.1 kJ/mol (Figure 5.6a), similar to that seen for the uncoated gold substrate and somewhat more than seen for a single layer of the less polar poly(3CPM-APVE) (vide infra). We observe a decrease in le relative to the 6- mercapto-l-hexanol modified QCM, presumably due to the blocking of some buried hydroxyl functionality by reaction with adipoyl chloride. We anticipate some interaction between methanol and the polymer because of the permanent dipole moment of the 4- bromophenyl maleimide monomer unit. Semi-empirical calculations using HyperChemqD show 11. = 3.76 D for 4-bromophenyl maleimide. Poly(3CPM-APVE) modified QCM: Poly(3-chlorophenylmaleimide-arninopropyl vinyl ether) was bound covalently to the QCM by formation of an amide bond to an underlayer of adipoyl chloride that had been bound to the 6-mercapto-1-hexanol (Figure 5.1e). A second layer of polymer is linked to the first using adipoyl chloride as the covalent cross-linking agent.18 The adsorption isotherm data for poly(3CPM-AVPE) interfaces provides insight into the structure-dependent adsorption kinetics that characterizes such systems. Upon addition of a single layer, the adsorption isotherm for methanol (Figure 5.6b) indicates a monolayer volume of 209 i l pL, which is substantially less than the 650 pL seen for the 6-mercapto-1-hexanol surface. The enthalpy of desorption is AHd,s = 35.9 i 0.2 kJ/mol, similar to that measured for the 6- mercapto-l-methanol surface and less than that seen for the more polar poly(4BPM- -l32- APVE) layer. These data suggest that the dominant interaction of the methanol adsorbate is with buried hydroxyl and/or carboxylic acid functionality beneath the polymer layer, with the reduced monolayer volume being an indication of the fraction of unreacted hydroxyl and carboxylic acid groups present. The addition of a second polymer layer yields the same adsorbate monolayer volume, th = 216 i 14 pL, but a larger enthalpy of desorption, AH“, = 38.5 i 0.7 kJ/mol (Figure 5.6b). We note no ellipsometric thickness change of the polymer layers on exposure to methanol, consistent with the notion that the adsorbate interacts most strongly with unreacted portions of the underlayer. The finding of the same monolayer volume argues for the buried hydroxyl and carboxylic acid groups acting as the dominant interaction sites for the adsorbate, with the difference in AHdes being reflective of the structural properties of the polymer. In this picture, the second polymer layer reduces the permeability of the interface. Because there is a substantial vapor phase adsorbate concentration, the rate of adsorption is likely quasi-zeroth order while the rate of desorption is mediated by the ability of the methanol to escape the polymer layer(s) after adsorption. This model does not argue for a change in the strength of interaction of the adsorbate with the underlayer, but rather that the apparent change in AH“,s is a consequence of the change in the rates (not necessarily the rate constants) of adsorption and desorption into/out of the polymer film. Because we are measuring mass uptake within the polymer interface, we do not have the experimental resolution to determine where within the interface the adsorbate is interacting most strongly. Once in the polymer layer, the methanol has more difficulty escaping from the bilayer than from the monolayer. In this manner we can use layer structure to mediate the kinetics of adsorbate uptake and/or desorption. -l33- t—1 r— O N O O O O r r o 0.8 ‘ 1.0 b z=pm* 1200 1000 800 I O I O I (10 (12 (14 (I6 (18 1L0 z = p/p* Figure 5.6. Methanol adsorption isotherm data and best-fit lines for a QCM modified with (a) poly(4BPM-APVE) and (b) poly(3CPM-APVE) with one polymer layer (0) and two polymer layers (O). The data were fitted to the BET model as given in Eqs. 2 and 3. Results of the fits are presented in Table 5.1. -134- 5.4 Conclusions The range of substrates and polymer structures we have examined using adsorption isotherm measurements with methanol and hexane as adsorbates demonstrates that structural control over selectivity and adsorption/desorption kinetics is feasible. We interpret our data for single polymer layers in the context of polymer side group and adsorbate polarity. In many cases we observe adsorption isotherm behavior consistent with capillary condensation, pointing to the structurally heterogeneous and/or porous nature of the polymer layers. Modification of our gold QCM electrode interfaces with 6- mercapto-l-hexanol leads to a significant increase in monolayer coverage relative to the uncoated gold surface, suggesting disorder in the monolayer. The construction of an interface using poly(4MPA-VEP) with ionic layer linking chemistry yielded the expected result for a system possessing substantial free volume. In this case we recover a BET Type V isotherm, consistent with capillary condensation within the internal matrix of the polymer. For the case where adsorption of methanol is compared for one and two layers of poly(3CPM-APVE), we find that the addition of the second layer influences the effective adsorption and desorption rates, leading to an apparent change in desorption enthalpy. These experiments demonstrate the feasibility of kinetic control in diffusion/adsorption over molecular length scales. We anticipate that the use of polymer multilayers comprised of several different substituted MVE polymers will afford even greater kinetic control and adsorbate selectivity. -l35- Table 5.1. Layer thickness, monolayer adsorbate volume (le) and enthalpy of desorption (AI-Ides) extracted from experimental data and fits to BET isotherm. Methanol Hexane Interface Thickness AHads AHads (A) Vm’ (9L) (kJ/mol) Vm‘ 9’” (kJ/mol) Gold -- 38 1 2 37.6 1 0.1 179 1 35 29.2 1 0.3 6-mercapto-1—hexanol 911 650 1 55 35.4 1 0.3 896 1 46 32.9 1 0.2 Poly(4HPM-VEP) 3512 894 1 40 36.0 1 0.2 -- -- Poly(4MPA-VEP) 4913 4752 1 744112 35.6101 29.1101 68 Poly(4BPM-APVE) 3011 425 1 72 37.1 1 1.1 -- -- Poly(3CPM-APVE) 1 34:2 209 1 1 35.9 1 0.2 -- -- layer Poly(3CPM-APVE) 2 5613 216114 38.5107 -- -- layers -136- 5.5 10. ll. 12. 13. Literature Cited Bakiamoh, S. B.; Blanchard, G. J. Langmuir, 2001, 17, 3438. Katz, H. E.; Scheller, G.; Putvinski, T. M.; Schilling, M. L.; Wilson, W. L.; Chidsey, C. E. D. Science, 1991, 254, 1485. Li, D.; Ratner, M. A.; Marks, T. J.; Zhang, C. H.; Yang, J.; Wong, G. K. J. Am. Chem. Soc.,: 1990, 112, 7389. Demas, J. N.; Degraff, B. A.; Coleman, P. B. Anal. Chem. New & Features, 1999, 793A. Yoon, H. C.; Kim, H-S. Anal. Chem. 2000, 72, 922. Delamarche, E.; Sundarababu, G.; Biebuyck, H.; Michel, B.; Gerber, C.; Sirgrist, H.; Wolf, H.; Ringsdtirf, H.; Xanthopoulos, N.; Mathieu, H. J. Langmuir, 1996, 12, 1997. Brousseau, L. C. III; Aurentz, D. J.; Benesi, A. J .; Mallouk, T. E. Anal. Chem. 1997, 69, 688. Vrancken, K. C.; VanderVoort, P.; Gillisdhamers, 1.; VanSant, E. F.; Grobet, P.; J. Am. Chem. Soc. Faraday Trans. 1992, 88, 3197. Pfleiderer, B.; Albert, K.; Bayer, E. J. Chromatogr. 1990, 506, 343. Ulman, A. Chem. Rev. 1996, 96, 1533. Dubois, L. H.; Nuzzo, R. G. Annu. Rev. Phys. Chem. 1992, 43, 437. Skulason, H.; Frisbie, C. D. Langmuir, 1998, 14, 5834. Lee, H.; Kepley, L. J.; Hong, H. G.; Akhter, S.; Mallouk, T. E. J. Phys. Chem. 1988, 92, 2597. —137- 14. 15. l6. l7. 18. 19. 20. 21. 22. 23. 24. 25. Lee, H.; Kepley, L. J.; Hong, H. G.; Mallouk, T. E. J. Am. Chem. Soc. 1988, 110, 618. Putvinski, T. M.; Schilling, M. L.; Katz, H. E.; Chidsey, C. E. D.; Mujsce, A. M.; Emerson, A. B. Langmuir, 1990, 6, 1567. Major, J. S.; Blanchard, G. J. Langmuir, 2001, 17, 1163. Major, J. S.; Blanchard, G. J. Chem. Mater., in review. Major, J. S.; Blanchard, G. J.Chem. Mater., in review. Brunauer, S.; Emett, P. H.; Teller, E.; J. Am. Chem. Soc., 1938, 60, 309. Sauerbrey, G. Z. Z. Phys. 1959, 155, 206. Karpovich, D. S.; Blanchard, G. J. Langmuir 1997, 13, 4031. (a) Brunauer, S.; The Adsorption of Gases and Vapors, Vol. 1., Princeton University Press, Princeton, NJ, 1945. (b) Adamson, A. W.; Physical Chemistry of Surfaces, Fifth Edition, John Wiley & Sons, Inc., New York, 1990, p. 6097f. Lide, D. R., Ed.; CRC Handbook of Chemistry and Physics, 71” Edition, CRC Press, Inc., Boca Raton, FL, 1990. Abbott, N. L.; Kumar, A.; Whitesides, G. M. Chem. Mater. 1994, 6, 596-602. Kohli, P.; Rini, M. C.; Major, J. S.; Blanchard, G. J.; J. Mater. Chem, 2001, 11, 2996-3001. -138- Chapter 6 Molecular Control of Adsorption Properties of Polymer Modified Interfaces Abstract In this work we report the assembly of several polymer-modified interfaces where we attempt to generate a chemical potential gradient in the interface along the surface normal. To achieve this goal, we have prepared a suite of copolymers capable of multilayer assembly, where these polymers have been designed to be self-terminating at the deposition of one molecular layer. The polymers are prepared by radical copolymerization of various vinyl ether monomers with maleimide monomers that have been prepared in-house and present a variety of functional side-groups. The functionalities present on the vinyl ether side groups are used in layer assembly, while the maleimide side groups are used to generate the chemical potential gradient. In order realize the chemical potential gradient, we have generated a gradient in the dipole moment of the maleimide side groups. Measurement of the adsorption isotherms of several polymer bilayers upon exposure to methanol and hexane reveals behavior that depends on both the identity and layer order of the interfaces, and we interpret these data in the context of polymer layer polarity. 139 6.1 Introduction Separation science is one area of chemistry where the chemical nature of the interface is intimately related to its bulk behavior. In most separation schemes, there are two different phases (typically one stationary and the other mobile) where the change in the chemical potential that a solute experiences as it is transferred between the two phases serves as the driving force for separation. In most cases when the resolution of solute molecules is poor, attempts at achieving improved separation/selectivity have focused on applying some external gradient such as temperature”, pressureg’12 or voltage13 programming. Attempts to vary the chemical potential of system have focused on solvent 2'4‘9’14'16 where creating a gradient in solvent “strength” leads to resolution of gradients, poorly resolved solutes. In all of these cases, the application of a gradient presents several challenges to the experimenter, such as the need for additional equipment, equipment stability (in the case for solvent gradients) costly solvents, and the time required to optimize the separation. Additionally, the use of high pressures and voltages could also prove unsafe. According to Giddings, the generation of a gradient in some property of a system would lead to a chemical potential gradient within the material.17 In this work, we have designed and assembled a variety of interfaces where we attempt to create a chemical potential gradient within the interfaces along the surface normal. To realize this structural goal, we have employed layer-by-layer assembly of maleirrride-vinyl ether alternating copolymers onto our interfaces (quartz crystal microbalance, hereafter referred to as QCM) where we have varied the identity (dipole moment) of the functional constituent incorporated into each layer. In this way, we generate a gradient in the dipole 140 moment of the interface. Interfaces possessing an internal gradient in some property should lead, in principle, to enhanced separations and would eliminate many of the problems associated with the generation of gradients according to the schemes mentioned above. In order to characterize these interfaces, we perform adsorption isotherm measurements and fit our experimental data to the BET model. We have employed an in- house assembled apparatus composed of QCM electronics and discrete optical components that allow simultaneous mass uptake and in situ ellipsometric thickness 18"9 of this system has been described measurements. The design]8 and operation elsewhere. Using this apparatus we were able to determine mass uptake as a function of adsorbate vapor pressure. We have reported previously (chapter 5) on the adsorption behavior of various polymer-modified interfaces"). In that work, we established that the predominant interaction between the adsorbate and the interface occurred at the underlayer as a result of the adsorbates ability to partition into the interface. We also demonstrated that the polymer overlayer influences this process, apparently by influencing the desorption kinetics. In the work we report here, we have set out to demonstrate our ability to selectively control the partitioning process through design. As mentioned above, these interfaces have been designed where we vary the identity of the functional groups incorporated into each layer thereby manipulating the overall property of the interface. While this body of work represents a substantial departure from the mainstream of separation science, the results of our studies are clearly indicative of the ability to design interfacial materials where we are able to control selectivity of the interface by synthetic 141 IN.“ means. We demonstrate that, in addition to affecting the properties of the interface based on the functional groups incorporated during polymer synthesis, varying the order of deposition of the polymers leads to dramatic changes in the form of the adsorption isotherm. At this point, the relationship between polymer side group dipole moment and adsorption behavior remains to be explored fully, but these preliminary data suggest that we have control over interface adsorption behavior. Further studies are underway to determine more qualitatively the role of dipole moment on the adsorption properties of interfaces. 6.3 Experimental In this work we employ several alternating maleimide vinylether copolymers 20’2“. We assemble whose synthesis and multilayer assembly was reported previously multilayers of these polymers onto quartz crystal rrricrobalance (QCM) electrodes (Au) and have used these substrates to measure the vapor phase adsorption isotherm for these systems. Substrate preparation. QCM acquired from McCoy Electronics (P/N 78-18-4). Their surface is characterized as microcrystalline gold domains that are predominantly (111). These substrates were cleaned in piranha solution (3:1 H2SO4:H2O2) followed by rinsing in ethanol, then water. These substrates were then dried under a stream of nitrogen gas. In order to introduce reactive functionalities onto the surface of the QCM, we use well-established alkanethiol/gold chemistry. In this scheme, we expose the clean QCM to a 10mM solution of 6-mercaptohexanol in prepared in a 60:40 ethanol:water solution. The QCM and solution were warmed to ~40°C for 30-60 minutes following by 142 ethanol and water rinses, and drying by nitrogen gas. Once the QCM electrode surfaces have been hydroxylated they are then exposed to a solution of 3% (v/v) adipoyl chloride and 3% (v/v) 4-methylmorpholine in anhydrous acetonitrile under an inert argon atmosphere for approximately 30 minutes at 40°C. After reaction with adipoyl chloride, the QCMs are removed from the solution and rinsed with ethyl acetate followed by exposure to the polymer solution. Multilayer assembly. Subsequent to surface preparation, the QCM is then exposed to the appropriate polymer solution and warmed to 40°C for 30 minutes. Polymer solutions are prepared at a concentration of 10 mM in either dimethyl sulfoxide (DMSO) or a 75:25 DMSO:acetonitrile poly(NPM-VOB). These solutions were reused throughout the entire experiment with no observed loss of reactivity. Multilayer assembly is accomplished by reacting the initially deposited polymer layer with adipoyl chloride as described above followed by exposure to the second polymer solution. This scheme can be employed to deposit the desired number of polymer layers. In the following discussion we use the following nomenclature; S-xxx-yyy, where S indicates the substrate, and the terms xxx and yyy are abbreviations for the polymers used. BPM = poly(4BPM-APVE), CPM = poly(3CPM-APVE), NPM = poly(NPM-VOB) and HPM = poly(HPM-VEP). In all cases, adipoyl chloride is used as the interlayer linking moiety. S-BPM-NPM: This interface (scheme 6.1a) was prepared by first linking one layer of poly(4BPM-APVE) to the (u-hydroxythiol-functionalized substrate using adipoyl chloride under the conditions reported above. The initial layer of poly(4BPM-APVE) was then reacted with adipoyl chloride followed by exposure to a solution of poly(NPM- VOB). 143 S-NPM-BPM: This interface (scheme 6.1b - the structural complement to the above interface) was prepared in an identical manner to S-BPM-NPM except poly(NPM- VOB) is the initial polymer layer, followed by the addition of a layer of poly(4BPM- APVE). S-BPM-CPM: This interface (scheme 6.1 c) was prepared by first depositing one layer of poly(4BPM-APVE) then capping with poly(3CPM-APVE), using adipoyl chloride as the crosslinking agent. S-CPM-BPM: This interface (scheme 6.1d) is the structural complement to the above system prepared under identical experimental conditions. Here, the layer deposited initially is poly(3CPM-APVE), which is reacted with adipoyl chloride, then with poly(4BPM-APVE) to generate the bilayer. S-BPM-HPM: This substrate (scheme 6.1e) is prepared by initially depositing a layer of poly(4BPM-APVE) followed by reaction with adipoyl chloride and deposition of poly(4I-IPM-VEP). S-HPM-BPM: This interface (scheme 6.1f — the structural complement to the above interface) was prepared by first depositing one layer of poly(4HPM-VEP) followed by one layer of poly(4BPM-APVE). It should be noted that in these two interfaces, the deposition of poly(4HPM-VEP) was accomplished by reaction of the hydroxyl group of the phenyl maleimide substituent with adipoyl chloride, leading to the formation of an aromatic ester. 144 0N0 Eoi O 0 Br 0 0 ii: ii ‘i l 1 Scheme 6.1. Interface structures. (a) S-NPM-BPM, (b) S-BPM-NPM. 145 ”—1 -l Edit—jam Mal? "' "l ‘ .4 — Scheme 6.1 cont’d. Interface structures. (c) S-BPM-CPM, (d) S-CPM-BPM. 146 Scheme 6.1 cont’d. Interface structure. (e) S-BPM-HPM, (f) S-HPM-BPM. 147 We L‘I'I.v: Inn-91m” . - I Optical ellipsometry. Measurements of polymer adlayer thickness on the QCM prior to exposure to the vapor phase adsorbates, were accomplished using a Rudolph Research AutoEL II commercial optical nulling ellipsometer. This ellipsometer was equipped with a HeNe laser operating at 632.8 nm. The software (DAFIBM) used to operate the ellipsometer was provided by Rudolph. Adsorption isothenn measurements. The apparatus we use for the measurement of adsorption isotherms has been described in Chapter 5. We note that the results we obtain with this system are highly reproducible. In almost all cases, reproducibility between samples is better than 5% and only at very high values of z do deviations sometimes become significant. In most such cases, these deviations are within 10% from run to run, even if separated in time by months. This reproducibility is seen for measurements taken on the same substrates but on different days. For substrates prepared at different times (often weeks apart) we obtain isotherm data that are reproducible to within ~5% or less. 148 'l 1. P 6.4 Results and Discussion The focus of the work we report here is the construction and evaluation of interfaces where we demonstrate the ability to manipulate the adsorption/desorption properties of the interface in some controlled manner. We have chosen to prepare these interfaces using various maleimide-vinyl ether copolymers (Chapter 2) and have assembled these polymers into bilayers using covalent crosslinking strategies that we have reported on in Chapters 2 and 3. We have chosen to work with these polymers due to the fact that we are able to introduce various functional groups into the matrix of the polymer through simple substitution of the maleirrride side-groups, where the preparation of the maleirrrides have been reported on elsewherezz’zz. Additionally, the ability to vary the crosslinking chemistry used to assemble the multilayers adds another means of tuning the properties of the resulting adlayer. A layer-by—layer assembly strategy allows us to vary the functional constituents incorporated into each individual layer — a feature that we use to control interfacial behavior. As mentioned earlier, the ultimate aim of this body of work is the design of interfaces where we attempt to generate a chemical potential gradient, with the goal of controlling the selectivity of the interactions between specific analytes and these interfaces. To construct the chemical potential gradient, we have varied the identity of the polymer side groups within each layer, where the gradient to be generated is controlled by the dipole moments of the unreacted side groups. We have used semi- empirical calculations (PM3 parameterization) to calculate the dipole moments of the maleirrride side groups and we present these results in Table 6.1. We present the adsorption isotherm data for the bilayer structures reported here in Table 6.2. 149 Table 6.1. Calculated dipole moments of selected polymer side groups. Calculations performed using Hyperchem v. 6.0 with PM3 parameterization. Maleimide side acronym Calculated SO drpole moment gm) (D) 0 O 0 Cl 0 O | -—.~Bt BPM 2.6 O The adsorption of vapor phase adsorbates onto the interfaces has been characterized in the context of the BET model. Using this model, we were able to extract useful information about our interfaces such as energies of interaction (enthalpies of desorption), and monolayer volume over limited ranges of adsorbate partial pressure, and have recovered substantially different values that depend on (1) the functional groups incorporated into the layers, and (2) the order of layer deposition. Based on these data, we have demonstrated our ability to manipulate adsorption in a qualitatively predictable manner. We have chosen methanol (polar protic) and n-hexane (non-polar) as adsorbates. We also attempted these studies using acetone (polar aprotic), but were unable to acquire useful data for most of our systems, so these studies were discontinued. The adsorption isotherm data and fits to Eq. 6.2 are presented in Figures 6.1 — 6.6. S-BPM-NPM/S-NPM-BPM: The adsorption isotherms for these systems reveal some interesting features. At low methanol vapor pressure, we observe that the isotherms for both interfaces are similar. As p approaches p*, the isotherm for S-BPM-NPM can be 150 approximated as a BET Type H isotherm, due to the apparent onset of saturation, while S-NPM-BPM yields behavior that is more Type H1 in character (Figure 6.1). A Type II isotherm indicates a strong(er) enthalpy of adsorption, i.e., stronger interaction of the adsorbate with the adsorbent, while a Type HI isotherm is typically taken to reflect a weaker interaction between the adsorbate and adsorbent. Fits of these data to the BET equation produce values of le = 91 1 5 pL and 371 1 24 pL for S-BPM-NPM and S- NPM-BPM, respectively. Values of c = 6.0 i 1.1 for S-BPM-NPM and 0.45 i 0.05 for S-NPM-BPM for exposure of these interfaces to methanol, reveal substantially different enthalpies of desorption; A112,es = 41.9 1 0.5 kJ/mol for S-BPM-NPM and AHdeS = 35.5 1 0.3 kJ/mol for S-NPM-BPM. When these interfaces are exposed to n-hexane, we measure adsorption isotherms that are qualitatively different than those seen for methanol (Figure 6.2). The isotherm for S-BPM-NPM exhibits a distinct region of saturation followed by one of efficient adsorption. Beyond 2 ~ 0.45, data collection is not possible due to the amount of adsorbate present at the interface. For S-NPM-BPM, we observe the onset of saturation near 2 ~ 0.5, consistent with capillary condensation. Fits of these data for low values of z to the BET equation yield values of Vm. = 113 1 15 pL and c = 9.68 1 3.57 for S-BPM- NPM and Vm. = 955 1 90 pL and c = 0.78 1 0.10 for S-NPM-BPM. The corresponding values of AHdes are 37.2 1 0.8 kJ/mol for S-BPM-NPM and AHdes = 30.9 1 0.3 kJ/mol for S-NPM-BPM. This behavior is the converse of that seen for methanol adsorption when the data are compared to AHvap for each solvent. For methanol, AHyap = 37.43 kJ/mol at 298 K and for n-hexane, AHvap = 31.56 kJ/mol at 298. K23 This finding indicates that the top-most polymer layer mediates, but does not 151 control completely, the adsorption behavior of these adlayers. This control over the effective thermodynamics of adsorption implies differential control over adsorption and desorption kinetics for these systems that is based on adlayer chemical structure. 16(1): 0 SUB-NPMBPM . I SUB-BPM—NPM 1400 - 1 . L nethano adsorption 1200 r O A 1000- 7i 1 Va, 8(111- '3 . o > 600.” f . 400 - 9 ~ 8 20) - 0 e 9 G 0 . . . 0 09' O 900 I I I I I I I I 1 0.0 0.2 0.4 0.6 0.8 1.0 2 (1113*) Figure 6.1. BET isotherm for the adsorption of methanol on S-BPM-NPM and S-NPM- BPM. 152 - " I" r 4]. I) W. —Wr _Ia K00- . ll SUBIWNHWEA 14m_ . O SUB-WW ' hameaqufiar EDO- HDO- jo} wL o O 0 g I 00 > 6D- 09 ' o e 403- 00 0 ' OO .0 1D- 000 .0 - 00090000. 0 00 1 1 1 1 1 00 02 04 06 08 10 Figure 6.2. BET isotherm for the adsorption of hexane on S-BPM-NPM and S-NPM- BPM. 153 S-BPM-CPM/S-CPM-BPM: The adsorption isotherm data we recover for these polymer adlayers reveal complex, adsorbate-dependent behavior (Figures 6.3 and 6.4). When S-BPM-CPM is exposed to methanol we recover an isotherm that is characterized by a plateau region near 2 ~ 0.5, followed be rapid adsorption and saturation. This behavior appears to be consistent with the two polymer layers functioning as independent entities, with saturation and possibly capillary condensation occurring for each layer separately (Figure 6.3). Exposure of S-CPM-BPM to methanol yields adsorption behavior similar to a BET Type III isotherm, with only a suggestion of saturation at higher z-values. While saturation at high 2 values occurs at almost the same point for both systems, the adsorption behavior up to that point is substantially different for the two systems. Fits of the experimental data to the BET equation yield c values of 0110.08 (AHdes = 31.7 1 4.0 kJ/mol) and 1.081036 (AHdes = 37.6 1 0.7 kJ/mol) for S- BPM-CPM and S-CPM-BPM, respectively. The monolayer volumes of le = 2800 1 1965 pL for S-BPM-CPM and le = 128 1 28 pL for S-CPM-BPM suggest a porous polymer structure. S-BPM-CPM appears to manifest BET Type HI behavior in regions of low z, where we fit these data to Eq. 6.2. We note that fitting these data proved difficult, as reflected in the significant uncertainty in the resulting values of th and c (Table 6.2). When these interfaces are exposed to hexane, we recover behavior that is characterized by saturation at low 2 values and BET Type H characteristics up to saturation. Fits of data for both polymer assemblies reveal large le values (Table 6.2) and small c values; AHdcs = 31.8 1 0.4 kJ/mol for S-BPM-CPM and AHdes = 25.3 1 0.3 kJ/mol for S-CPM-BPM. These interactions are weaker in absolute strength, but similar 154 in magnitude relative to AHvap as those seen for methanol. These findings reveal that 3CPM is experimentally less polar than 4BPM, in agreement with the semi-empirical calculations of side group dipole moment (Table 6.1). A point worth noting in this particular case is the fact that the energies of interaction relative to AHvap for each solvent are reverse for methanol and hexane, as expected. 14(1)— _ O SUB-BPM—CPM 1201- O SUB-CPM—BPM o rretlnnoladsorptim o 1000- ~ 0 o A 80)— o ....l 3 * o >§ 600- ' ' 0 4m- 0000. o ' 0o 0 21D- 00 ... r- 900.. .0 00.0 OHj-f-i.l I I I l 00 02 0.4 06 08 10 Z (1113*) Figure 6.3. Isotherms for the adsorption of methanol on S-BPM-CPM and S-CPM-BPM. 155 W4 lays“ 'mfl - O SUB-Bl’M-(PM O SUB-(PM-BPM Vad,(PL) s 8 es . s Figure 6.4. Isotherms for the adsorption of hexane on S-BPM-CPM and S-CPM-BPM. 156 S—BPM-HPM/S-HPM-BPM: Exposure of these interfaces to methanol yields BET Type II behavior for S-HPM-BPM with saturation occurring at z > 0.6 (Figure 6.5). The isotherm of S-BPM-HPM however, is slightly more complex. There is a significantly stronger interaction of S-HPM-BPM with methanol, as seen from the low-z behavior for both adlayers. Fitting the data yields c values of 0.66 1 0.1 (AHdcs = 36.4 1 0.4 kJ/mol) and 73.5 1 9.0 (AHdcs = 48.1 1 0.3 kJ/mol) for S-BPM-HPM and S-HPM-BPM, respectively. The fact that the apparently stronger interaction occurs for the case where the polar adlayer is confined between the substrate and a relatively less polar layer again indicates differential control over adsorption and desorption kinetics at these interfaces. 1600'- o SUB-BPM-HPM o e SUB-HPM—BPM 1400 - . _ methanol adsorptron 1200 - 0 1000 - 0 (a . v“ 8(1) r O 8 - o o > 600 - o ' . .. O 400 1- 00 ° . O 0 - o 200 - coo , o" -- 88833” ' O W I I I I 4 0.0 0 2 0 4 0.6 0 8 1 O z (p/p*) Figure 6.5. Isotherms for the adsorption of methanol on S-BPM-HPM and S-HPM-BPM. 157 Attempts to acquire data for the exposure of both S-HPM-BPM and S-BPM-HPM to hexane proved unsuccessful. In both cases, QCM oscillation was not consistent with adsorption. In some instances, we observed that the frequency of the QCM increased, behavior consistent with mass decrease, suggesting the possible destruction of the polymer layers. At high values of z, the QCM oscillation frequency was unstable, and upon removal of the adsorbate stream, we recovered the original (z = 0) oscillation frequency. In contrast to the other adsorbate/polymer adlayer systems, S-HPM-BPM and S-BPM-HPM do not appear to achieve a reproducible equilibrium condition with hexane adsorbate. Clearly, further work will be required to understand the reason for this anomalous behavior. One issue that requires consideration is the assumption that we are working in the thermodynamic limit with these adsorption measurements. The experimental data point to differential control over the adsorption and desorption kinetics of these polymer adlayers based on chemical structure and ordering. The issue of differential control over kinetic processes calls into question our assumption of operating in the thermodynamic lirrrit. At the very least, it is clear that we are operating under a steady-state condition, but our experimental data indicate that we are, in fact, in the thermodynarrric limit. We show in Figure 6.6 the forward and reverse adsorption isotherms for methanol adsorption on S-BPM-CPM. These data do not demonstrate any significant hysteresis. While it is true that hysteresis in adsorption isotherms can be seen, even in the thermodynamic limit, depending on the nature of the adsorbate-adlayer interaction, the absence of hysteresis demonstrates that we are operating in the limit where the adsorption and desorption 158 processes are occurring under equilibrium conditions. 1400- 1200- 8 1000- 8004 Vads (DU 600- O 400- 0°93 1 200- O. , . . 0.0 0.2 0.4 0.6 0.8 1.0 Figure 6.6. Forward (P ) and backward (D) isotherms for the exposure of S- BPM-CPM to methanol. 159 Table 6.2. Results of fitting experimental adsorption isotherm data to Equation 6.2. methanol hexane V1111 C Vm] C S-BPM-NPM 9115 pL 6.01l.1 113115 pL 96813.57 S-NPM-BPM 371124 pL 0.451005 955190 pL 07810.10 S-BPM-CPM 280011965pL 0101008 9681113 pL 1.111018 S-CPM-BPM 128128 pL 1.081036 169218 0.081001 S-BPM-l-IPM 911193 pL 0.661010 S-I-IPM-BPM 11712 pL 73519.0 6.5 Conclusions In this work, we have demonstrated our ability to control the adsorption and desorption kinetics of polymer interfacial materials differentially by means of the chemical identity of the polymer adlayers and the order in which they are deposited. We have observed significant differences in the effective interaction energies of these systems and the shapes of the isotherms recovered in most cases these shapes are consistent with the expected trends. These data are a first demonstration that we can control the adsorption properties of these interfaces and, ultimately, their selectivity through careful design and ordering of polymeric adlayers. 160 6.6 Literature Cited 10. ll. 12. l3. 14. 15. 16. 17. 18. 19. 20. Kanazawa, H.; Kashiwase, Y.; Yamamoto, K.; Matsushima, Y.; Kikuchi, A.; Sakurai, Y.; Okano, T. Anal. Chem. 1997, 69, 823. . Lukulay, P. H.; McGuffin, V. L. Anal. Chem, 1997, 69, 2963. Li, J.; Carr, P. W. Anal. Chem.1997, 69, 837. Go, H.; Sudo, Y.; Hosoya, K.; Ikegami, T.; Tanaka, N. Anal. Chem. 1998, 70, 4086. Li, J .; Carr, P. W. Anal. Chem. 1997, 69, 3884. Li, J .; Carr, P. W. Anal. Chem. 1997, 69, 2202. Thompson, J. D.; Carr, P. W. Anal. Chem. 2002, 74, 1017. Lan, K.; Jorgenson, J. W. Anal. Chem. 1998, 70, 2773. MacNair, J. E.; Patel, K. D.; Jorgenson, J. W. Anal. Chem. 1999, 71, 700. Tolley, L.; Jorgenson, J. W.; Moseley, M. A. Anal. Chem. 2001, 73, 2985. MacNair, J. E.; Lewis, K. C.; Jorgenson, J. W. Anal. Chem. 1997,69, 983. McGuffin, V. L.; Chen, S-H. Anal. Chem. 1997, 69, 930. Hutterer, K. M.; Jorgenson, J. W. Anal. Chem. 1999, 71, 1293. Taylor, M. R.; Teale, P.; Westwood, S. A.;.Perrett, D. Anal. Chem. 1997, 69, 2554. Yan, C.; Dadoo, R.; Zare, R. N.; Rakestraw, D. J.; Anex, D. 8. Anal. Chem. 1996, 68, 2726. Huber, C. G.; Choudhary, G.; Horvath, C. Anal. Chem. 1997, 69, 4429. Gidding, C. G. Unified Separation Science; Wiley Interscience, 1991. Karpovich, D. S.; Blanchard, G. J. Langmuir, 1997, I3, 4031. Major, J . S.; Blanchard, G. J. Langmuir, in review. Major, J. S.; Blanchard, G. J. Chem. Mater. in press. 161 21. Major, J. S.; Blanchard, G. J. Chem. Mater. in press. 22. Kohli, P.; Blanchard, G. J. Langmuir 1999, 15, 1418. 23. Lide, D. R. CRC Handbook of Chemistry and Physics, 82nd Ed. ,' CRC Press, 2001- 2002. 162 Chapter 7 Covalently-Bound Polymer Multilayers for Efficient Metal Ion Sorption Abstract We report on the design, synthesis and characterization of a polymeric multilayer assembly that can act as an efficient sorbent for selected metal ions. We use a covalent layer-by-layer deposition scheme for the construction of multilayer assemblies using the alternating co-polymer of 4-hydroxyphenylmaleimide and ethylvinylether-Z- diisopropylphosphonate. The maleimide-vinyl ether (MVE) polymer layers are reacted at the maleimide hydroxyl functionality with adipoyl chloride to form an ester linkage, and the terminal acid chloride functionality is reactive toward subsequent deposition of another MVE polymer layer. The vinyl ether isopropylphosphonates remain protected during layer growth. Once layer growth is complete, the isopropylphosphonates are deprotected using bromotrimethylsilane (BTMS) to activate the multilayer for metal ion sorption. The uptake of Zr4+ by the multilayer is confirmed by ellipsometry, FI‘IR, X-ray photoelectron spectroscopy (XPS) and quartz crystal microbalance (QCM) gravimetry. The XPS data show ~ 13 atom % of the multilayer is Zr4+ after exposure to a 5 mM ethanolic solution of zirconyl chloride, and QCM data show that metal ion uptake is fast. The films also show changes in FTIR spectra and ellipsometric thickness, indicating substantial structural changes associated with metal ion uptake. -l63- 7.1 Introduction Gaining the ability to design and construct interfaces with significant control over their macroscopic properties has attracted substantial research interest. This effort has been aimed largely at optimizing molecular-level organization, with Langmuir-Blodgett films, alkanethiol-gold self-assembled monolayers (SAMs)“4 and metal phosphonate multilayers}7 being the most widely studied examples. The array of chemical functionalities that can be introduced into the layers, as well as the range of assembly and interlayer chemistry that can be accessed affords great versatility in the properties of the resulting materials. The promise of interfacial chemistry lies in its potential application in areas such 8,9 11 as bio-sensors/bio-recognition, optical second harmonic generation)“ chemical sensors”’13 and separations.'4"5 The ability to assemble thin films with molecular-scale control over layer thickness and uniformity necessitates the use of well-defined and well- controlled reaction schemes. We have reported previously on the growth of layered polymeric interfaces using maleimide-vinyl ether (MVE) chemistry, where Zr- bisphosphonate (ZP) interlayer linking chemistry is used to achieve layer-by-layer growth.‘6 In that work, the identity of the maleimide substituent determines the properties of the resulting multilayer interface. We report here on a different use of MVE alternating copolymers in the construction of multilayer assemblies. Our present focus is on the design, synthesis and characterization of a covalently-bonded multilayer structure where the polymer vinyl ether side groups are free to interact strongly with any metal ions they come into contact with. We construct the MVE copolymer using 4-hydroxyphenyl maleimide and -l64- ethylvinylether-2-diisopropylphophonate. The unique design aspect of this work is that we connect the polymer layers through the 4-hydroxyphenylmaleirnide side groups instead of using the well-established metal ion coordination to the phosphonate side groups. We use adipoyl chloride to form covalent diester interlayer linkages. In contrast to the conventional ZP interlayer linking chemistry, where the phosphonates are sequestered by the formation of metal-bisphosphonate complexes, we use the sidegroups in these polymer multilayers in their deprotected form as active sites for metal ion uptake, and report here on the interaction between deprotected polymer multilayers and Zr“. To circumvent the formation of phosphoesters during layer growth, the phosphonate groups of the copolymer are protected by isopropyl functionalities. Even if these linkages form to some extent, the deprotecting agent bromotrimethylsilane (BTMS), which we use to activate the multilayers, has been shown to de-esterify phosphoesters.l7 After the desired number of deposition cycles, four in this case, the phosphonate groups are deprotected using BTMS. We have chosen to demonstrate this chemistry using four layers as a matter of convenience. We have grown thicker layers and see no evidence to indicate a loss of reactivity beyond four layers. Deprotection results in the facile hydrolysis of the phosphonate groups, rendering a multilayer film capable of efficient metal ion uptake. Exposure of the deprotected multilayer to ZrOCl2 solution results in rapid metal ion uptake. Ellipsometry, FTIR and X-ray photoelectron spectroscopy (XPS) data each reveal significant changes in the multilayer structure after sorption of Zr“. XPS measurements show that ~13 atom % of the surface, is comprised of Zr“, consistent with the concentration of Zr“ we would expect for saturation of the multilayer assembly. -165- These data point to essentially irreversible uptake of the metal ion. XPS data on systems exposed to the ZrOC12 solution for extended periods of time yielded the same concentration as those exposed for short times. Quartz crystal rrricrobalance (QCM) data demonstrate the uptake of the Zr“ is fast, with a steady state loading density being achieved within seconds of exposure to metal ion solution. We report here on the design, synthesis and characterization of these multilayer assemblies that act as metal ion “sponges”. 7.2 Experimental Syntheses: The synthesis of maleimide-vinyl ether alternating copolymers has “5'18 In this work, we use 4-hydroxyphenylmaleimide as one been described before. monomer. It is synthesized from maleic anhydride and 4-hydroxyaniline and the reaction chemistry is the same as that described for similar, substituted maleimides.16 The vinyl ether phosphonate was prepared by reacting tri(isopropyl)phosphite with 2- chloroethylvinylether according to a method described by Rabinowitz.l9 This reaction was carried out under an inert argon atmosphere and refluxed at 170°C for five days. The maleimide-vinyl ether alternating copolymer was prepared by radical copolymerization using azobisisobutyronitrile (AIBN) as the initiator (scheme 7.1a).'8‘20 We observe no homopolymerization of the monomers used in this work. -l66- 0 o “ AIBN/CHC13 24 Hrs, 60°C iPrO/P OH OiOPr o } BTMS/CHgCN 2 Hrs, 45°C /P\‘\‘0 P\=o IPrO OiPr HO 0H CH1 (”1 Scheme 7.1. (a) Monomers used in the synthesis of the MVE alternating copolymer reported here. (b) Deprotection (activation) chemistry for the MVE copolymer. In this scheme, the isopropyl groups are hydrolyzed subsequent to layer assembly. -l67- Polymer Layer Deposition: Quartz and gold substrates, including QCMs, were cleaned using piranha solution (3:1 H2SO4:H2O2) and rinsed with ethanol, then with distilled water prior to layer deposition. The gold substrates were first exposed to a 10 mM solution of 6-mercapto-1-hexanol in a 60:40 ethanol:water solution for 1 hour at 45°C. Quartz substrates were exposed to a 5 M HCl solution for 30 minutes. Both substrates were rinsed with ethanol and water, then dried under a stream of nitrogen and exposed to adipoyl chloride in anhydrous acetonitrile under an argon atmosphere. The substrates were then removed and rinsed with ethyl acetate, dried under a stream of nitrogen and exposed to a 10 mM solution of the MVE polymer in DMSO for 1 hour at 45°C. The resulting polymer layer was reacted with adipoyl chloride, then immersed in the polymer solution and the process repeated as required to deposit the desired number of layers (scheme 7.2). Deprotection/hydrolysis of the diisopropylphosphonate groups: Once the multilayer synthesis is completed, the polymer layers, which contain diisopropylphosphonate side groups, were exposed to bromotrimethylsilane (BTMS) in anhydrous acetonitrile under an argon atmosphere for two hours. The substrates were then removed from the BTMS solution and immersed in acetonitrile for 30 minutes, rinsed with ethyl acetate and dried under a stream of nitrogen. After rinsing the substrates, ellipsometric measurements were taken, followed by the exposure of the substrates to a 5 mM ethanolic ZrOC12 solution for at least one hour. -168- o o .. 6% fl Q:~r\/\O o/\/P.. 0.. o I; O O O N 19.8319 0 o .H mm o R IV..\/\ o/\/p...6 0.. I, o 0 0 0 \/\/>\3 H0 Idealized schematic of the MVE polym system showing two layers -169- connected by a diester linkage. Scheme 7.2. Optical Null Ellipsometry: Ellipsometric thickness measurements of the layers deposited on gold were made with a Rudolph Auto-EL H optical null ellipsometer operating at 632.8 nm. Rudolph DAFIBM software was used for data collection and processing. For all films, the refractive index was taken to be n = 1.54 + 0i. Infrared Spectroscopy: For the multilayer assemblies on gold, FTIR spectra were acquired following each individual deposition cycle. IR spectra were also collected prior to multilayer reaction with BTMS and ZrOCl2. For all measurements, a Nicolet Magna 750 FT IR spectrometer was used, and for all data presented here, the instrumental resolution was set to 4 cm". The data were acquired using an external reflectance sample mount, set to an incidence angle of 80°. UV—Visible Spectrophotometry: UV-visible absorption data were acquired for samples grown on quartz substrates. Spectra were acquired after each deposition cycle using a Cary 300 UV-visible spectrophotometer. The wavelength range scanned was 190 nm - 500 nm, at a scan rate of 600 nm/min. The collected data were plotted using Microcal Origin 6.0® software. X-Ray Photoelectron Spectroscopy (XPS): XPS data were acquired using a Perkin-Elmer Physical Electronics PHI 5400 X-ray photoelectron spectrometer. This system was equipped with a Mg X-ray source operated at 300 W (15kV, 20 mA). The carbon (Cls) line at 284.6 eV was used as a reference in determining the binding energies of the various species. Quartz Crystal Microbalance (QCM): Time-resolved solution phase QCM data were acquired using a Hewlett-Packard 53131A 225 MHz universal frequency counter with data acquisition programmed using National Instruments LabVIEW® programming -170- language. The in-house built QCM holder was connected to a resonant oscillator circuit (Maxtek, Inc.) powered by a 5V DC supply. A Neslab RTE-111 refrigerated circulating water bath was used for temperature control at 303 K 1 0.01 K. The mounted QCM is immersed in a jacketed beaker containing 100 mL of mixed hexanes at 30°C. The solution is stirred vigorously. Once a baseline oscillation frequency is established, 1.0 mL of a 5 mM solution of zirconyl chloride in 60:40 ethanol:water is introduced. Data collection is terminated after a new steady-state frequency had been established, typically within a minute. 7.3 Results and Discussion The focus of this work is on a novel covalent growth strategy for MVE polymers that allows their use as efficient metal ion sorbent materials. In this work, the ZP chemistry usually employed in the connection of layers is used instead as a chemically reactive site for metal ion complexation. The novel aspect of the work we report here lies in our ability to assemble these systems in a discrete manner such that the phosphonate groups are available at a later time for metal ion sequestration. The maleirrride monomer we use in polymer synthesis possesses a terminal hydroxyl functionality which we react with adipoyl chloride to form interlayer linkages. Because we use a protected phosphonate-containing vinyl ether, formation of a phosphoester functionality does not compete efficiently with the formation of the diester. With this chemistry, the formation of porous multilayers is facile and subsequent reaction of the protected phosphonates with BTMS results in essentially complete deprotection (Scheme 7.1b). Several groups have explored the use of covalent ester and amide interlayer linking chemistry“24 for the -l71- growth of thin films. The method we report here uses diester formation for layer adhesion, resulting in rapid growth of robust layers. Using this chemistry, the assembly of a four-layered system takes about eight hours. Optical null ellipsometry and FTIR and UV-visible spectroscopies all confirm the deposition of the polymer layers. The ellipsometry data (Figure 7.1) demonstrate a linear dependence of interface thickness on number of layers, with a slope of 16 1 1 A per adipoyl chloride-MVE polymer layer growth cycle. Each ellipsometric data point is the average of 40 individual determinations at different locations on the sample surface. Semi-empirical calculations for the layer unit suggest a thickness of 24 A for the fully extended structure. The fact that we recover 16 Allayer experimentally, which is the same as seen for layers of phenyl-substituted maleimide-containing polymer,16 is likely the result of the substantial lack of order in these layers. The ellipsometric measurements represent the average of the thickness distribution for the sampled area, and there is likely heterogeneity over length scales smaller than those sampled by the ellipsometric experiment. We have also used FT IR to monitor the layer-by-layer deposition of the polymer, and present these data in Figure 7.2a. We monitor the carbonyl stretching region between 1770 cm”I and 1650 cm'1 because both the maleirrride monomer and diester interlayer linkages contain these functionalities. Unfortunately, these bands are not resolved. The C=O absorbance depends linearly on the number of layers deposited (Figure 7.2b). -l72- 8 slope = 16 1 l A/layer \) O ' 1 interoept=41lA ellipsometric thickness (A) 8 8 8 ES 8 10- 0 1 1 1 1 0 l 2 3 4 number of layers Figure 7.1. Ellipsometric data for a four-layered MVE polymer system. Each layer consists of the MVE polymer and the adipoyl chloride underlayer. The slope of the best fit line through the data is 1611 Allayer. -173- 0.010 I Layer 3 Layw m 0.000 - 4 0.005 Absorbance (a.u.) 3500 3250 3000 270,00 1750 1500 1250 1000 frequency (cm’l) 0.004 1- b 0.003 - 3? a Q) g 0.002 ~ :6 e O a 0.001 - 0.000 1 . 1 1 0 1 2 3 4 number of layers Figure 7.2 (a) FTIR absorbance data layers 1-4. The spectra are offset for clarity of presentation. (b) Carbonyl stretching band absorbance as a function of number of layers. -174- The UV-visible spectroscopic data were collected for each layer deposited and they also yielded a linear response per layer (Figure 7.3). The dominant absorption band is associated with the hydroxyphenyl succininride moiety in the polymer backbone and its maximum is centered at 231 nm. We are not aware of any reports of the extinction coefficient for this chromophore, precluding our ability to estimate the loading density of the polymer layers on the surface. We use both ellipsometry and FT IR to monitor the deprotection/hydrolysis of the isopropylphosphonate groups. After the deprotection reaction, the ellipsometric thickness of the four-layer interface decreased from 83 1 6 A to 44 1 1A (the uncertainties are 110). We believe that this decrease in thickness is the result of replacing the relatively bulky isopropyl groups with the smaller hydroxyl groups. This reduction in thickness can be accounted for only in part by the change in physical size of the phosphonate terminal functionalities. For the deprotected phosphonic acid functionalities, the hydroxyl groups are capable of participating in hydrogen bonding. Hydrogen bonding between the hydroxyl groups would naturally result in a more compact structure, leading to a decrease in the thickness, as we observe. A key issue relative to the structural freedom within the polymer layers is the ability of the deprotected phosphonate groups to come into close proximity with phosphonate groups from adjacent layers. This is because of the ability of deprotected phosphonates to H- bond. Interlayer H-bonding can affect the thickness of the films substantially, while not affecting the availability of the phosphonate groups for metal ion complexation. The dramatic change in thickness upon deprotection argues for significant structural freedom within the polymer matrix. -175- fl“_‘7‘“—j 0.020 » 0.035 . E 0015- C: 0.030 ,7, ’ 8 0.010- < . 0.025 :5 3 0.020 0 . 5 0.015 0 . 3 0.010 <1 0.005 0.000 -0'm5 I I I I I I L I I A I I I I L 4 J 200 225 250 275 300 325 350 375 400 Wavelength (nm) Figure 7.3. UV—Visible absorbance spectra of MVE polymer layers as a function of number of layers. Inset: Dependence of the 231 nm absorption intensity as a function of number of layers. The FTIR data also confirm the deprotection/hydrolysis reaction based on the significant changes seen in the organophosphorus spectral region (Figure 4a). Unambiguous assignment of the features in this spectral region is not possible due to extensive overlap in the ~1415 cm’l - 1085 cm'1 region, but it is known that the position of these resonances is sensitive to the nature of the substituents attached to these groups. -l76- A significant difference is observed in the spectrum after deprotection and exposure to the ionic solution (Figure 7.4). The -OH stretching region at ~3600 cm'l exhibits a 4 pronounced change upon exposure to Zr +. 0.015 - 1 0.010 - 32‘ :3 Q) g 0 ~ .005 - 8 E < ”A 0.000 fi/A‘“ I I I III I I I 3500 3000 2000 1500 1000 fiequency(cn15 Figure 7.4. FTIR absorbance data for a four-layered system prior to deptrotection of the phosphonate groups with BTMS (spectrum I), after deprotection with BTMS followed by hydrolysis (spectrum H), and after exposure to ZrOCl2 (spectrum H1). The spectra are offset for clarity of presentation. The interfacial assembly we report here changes substantially upon exposure to Zr“. After immersion in a 5 mM ethanolic ZrOCl2 solution, the polymer multilayer increases to 137 1 23 A, an increase of more than three times the original thickness of the deprotected polymer film suggesting substantial structural disruption upon metal ion sequestration. This increase in thickness resulting from Zr“ uptake is likely related to the incorporation of non-stoichiometric waters of hydration in the multilayer matrix. The FTIR spectrum of the deprotected and zirconated polymer multilayer is shown in Figure -l77- 7.5. After rinsing the zirconated multilayer with 100% ethanol, then with distilled water, we see a decrease in the thickness of the interface to 110 A, which we attribute to the removal of non-stoichiometric waters of hydration incorporated into the multilayer structure during Zr“+ uptake. An FTIR spectrum taken after the ethanol/water rinse confirms the loss of water, as seen by the decrease in the ~3400 cm'1 and ~1500 — 1600 cm'l regions (Figure 7.5). Once this decrease has occurred, however, no further thickness change was observed, even after heating the interfacial assembly at 50°C in 100% ethanol for one hour. FTIR spectra taken before and after heating in 100% ethanol were idenficaL One issue of potential concern is whether or not the polymer swells upon exposure to ethanolic solution. We measured the ellipsometric thickness before and after immersion of the sample in a 60:40 ethanol:water solution and observed no thickness change to within the experimental uncertainty, indicating that these polymers do not swell upon exposure to the solvent system when metal ions are not present. When these same substrates are exposed to a solution of ZrOCl2 in 60:40 ethanol:water, the 100 A thickness increase for the four layer interface is reproduced readily. Examination of the interface FTIR spectrum after Zr“ uptake revealed significant changes in the hydroxyl stretch region of the spectrum. One significant change is the shift in the peak maximum from ~3200 cm'1 to ~3400 cm’1 (Figure 7.5) with processing. -l78- I— 0.008 0.006 0.004 absorbance (a.u.) 0.002 0.000 " 1 I 1 JII . 1 I 1 3500 3000 2600 1500 1000 frecuency (cm") Figure 7.5. Comparison of the four-layered system, after exposure to ZrOCl2 (spectrum III), and after rinsing with absolute ethanol, then water (spectrum Iv). These spectra are offset for clarity of presentation. The reason for the prominence of this band in spectrum III of Figure 7.5 is the presence of liquid water in the polymer multilayer subsequent to exposure to metal ions. When rinsed with ethanol, we observe a decrease in this peak (spectrum IV, Figure 7.5), coincident with a decrease in the ellipsometric thickness, consistent with the removal of non-stoichiometric water from the polymer film. After heating to 50°C in 100% ethanol for 1 hour, no further change was observed in the FTIR spectrum, again consistent with the ellipsometric data. -179- We have studied the uptake of Zr“ in the polymer multilayers using both XPS and quartz crystal microbalance gravimetry. We consider the XPS data first and present a survey scan of a zirconated four layer assembly in Figure 7.6. For the area sampled, approximately 12.9 atom % of the interface is Zr“.25 This finding suggests that most of the Zr“ is complexed by one phosphonate group. XPS data also confirm the presence of C1 in the multilayer assembly in some of the samples, and the meaning of its presence is not completely clear. We offer two possible explanations for this observation. First, the presence of Cl may be consistent with the maintenance of macroscopic charge neutrality within the polymer matrix. We note that the observed Cl concentration is not present in a stoichiometric amount. Second, the Cl could be due to the presence of unreacted adipoyl chloride used as an interlayer linker. We note that hydrolysis of the acid chloride functionality is known to be very efficient, making this explanation unlikely. In some cases, Cl was observed prior to exposure to ZrOCl2 and in other cases no Cl was detected even after exposure to the metal ion solution, making a conclusive explanation impossible. We note the apparent absence of a P resonance in these data. XPS is relatively insensitive to P, however, 3 1P NMR data of polymer multilayers bound to silica and suspended in solution26 (not shown) reveal the presence of P in these multilayers. We note that the XPS indication of 1:1 complexation is consistent with the increase in layer thickness upon metal ion sorption. If 2:1 complexation dominated, it would cross- link the polymer matrix, precluding an increase in thickness. -180- 70000 C Auger O Auger 60000r 50000- 015 ' N 1 § 40000- S 8 \ Zr 3p 0 30000- 1 C Is Zr3d I c1 y 20000- \ P 2p 1 Au 4f 10000- 0 I I I I I I I I I I 1000 800 600 400 200 0 binding energy (eV) Figure 7.6. XPS survey scan of a surface with four layers of poly(4HPM- VEP) on a gold substrate. We consider the QCM data next. This measurement records the change in quartz crystal microbalance resonant frequency as a function of time. The injection of an aliquot of Zr“ causes a change in the resonant frequency of a polymer-coated QCM suspended in solution and we present these data in Figure 7.7. The response of a QCM to -181- changes in mass is typically quantitative in the gas phase, but not in solution due to the dielectric properties of the surrounding medium. Our interest in these data is not in terms of the quantitative mass uptake (this information is available from the XPS data), but rather in the time-course of the QCM response. The kinetics of mass uptake for these systems can be understood in terms of the BET adsorption isotherm and thus we can obtain valuable information on the interaction of the metal ions with the polymer matrix. We provide the QCM data in this paper only as a demonstration of rapid metal ion uptake and will treat these data in greater detail in a subsequent paper. We have found that the uptake of Zr“ and Ca2+ (not shown) is very fast for this polymer multilayer structure. We note that, among the complexities associated with extracting quantitative information from these data is the change in the thickness of the polymer layer as a result of metal ion complexation. Detailed modeling of this effect will need to be considered before useful kinetic information can be extracted from the data. -182- -1m 1- Af (Hz) -150 _ time (sec) Figure 7.7. QCM data showing rapid uptake of Zr“ in a four-layer film. 7.4 Conclusions We have reported on two key points in this paper. First, we have demonstrated facile covalent layer-by-layer growth for MVE alternating copolymers. Second, we have -183- demonstrated that these materials can be used as efficient metal ion sorbents based on the XPS loading data and the QCM kinetic data. We provide ellipsometric, spectroscopic (FTIR, UV-visible and XPS) and gravimetric (QCM) data to support facile covalent layer construction and metal ion uptake. The ellipsometric, FTIR and UV-visible spectroscopic data demonstrate linear layer growth and the XPS and QCM data provide the mass and kinetic uptake information for Zr“ sorption from solution. The XPS are consistent with predominantly 1:1 P03: : Zr“ complexation, with the presence of Cl supporting this assertion. We observe a large ellipsometric thickness increase upon uptake of Zr“ and speculate that the basis for this thickness change is (1) structural rearrangement within the polymer matrix to accommodate the presence of the metal ions and (2) the presence of nonstoichiometric water that is somehow associated with the incorporation of the metal ion. These data are not consistent with polymer swelling upon exposure to the solvent system we use. A substantial fraction of the thickness change can be eliminated by washing of the metal-containing polymer multilayer with ethanol, supporting the assertion that nonstoichiometric water is associated with the initial thickness increase on exposure to metal ions. Several issues remain to be explored for these polymer multilayers. One important consideration is that, for four polymer layers, the polymer matrix appears to be sufficiently open to allow ready access by metal ions. How the polymer matrix permeability will change with the growth of additional layers will ultimately determine the use of these materials for large-scale rapid metal ion sorption applications. A second vital issue that needs to be addressed is a detailed understanding of the metal ion uptake -184- kinetics. Gaining this understanding will be complicated by the changes in the polymer matrix associated with the incorporation of the metal ions. Understanding the forward and reverse kinetic processes will allow us to determine the thermodynamic properties of these layers and thereby understand the extent to which metal ion uptake is reversible. These issues are under investigation. 7.5 Literature Cited 1. Ulman, A. Chem. Rev. 1996, 96, 1533. 2. Dubois, L. H.; Nuzzo, R. G. Annu. Rev. Phys. Chem. 1992, 43, 437. 3. Zhang, Z. J.; Hu, R. S.; Liu, Z. P.; Langmuir 2000, I6, 1158. 4. Skulason, H.; Frisbie, C. D. Langmuir 1998, 14, 5834. 5. Lee, H., Kepley, L. J., Hong, H. G., Akhter, S., Mallouk, T. E. J. Phys. Chem. 1988, 92, 2597. 6. Lee, H., Kepley, L. J., Hong, H. G., Mallouk, T. E. J. Am. Chem. Soc., 1988, 110, 618. -185- 7. Putvinski, T. M.; Schilling, M. L.; Katz, H. E.; Chidsey, C. E. D.; Mujsce, A. M.; Emerson, A. B.; Langmuir 1990, 6, 1567. 8. Yoon, H. C.; Kim, H-S. Anal. Chem. 2000, 72, 922. 9. Delamarche, E.; Sundarababu, G.; Biebuyck, H.; Michel, B.; Gerber, C.; Sigrist, H.; Wolf, H,; Ringsdorf, H.; Xanthopoulos, N.; Mathieu, H. J .; Langmuir, 1996, 12, 1997. 10. Katz, H. E.; Scheller, G.; Putvinski, T. M.; Schilling, M. L.; Wilson, W. L.; Chidsey, C. E. D.; Science, 1991, 254, 1485. 11. Li, D.; Ratner, M. A.; Marks, T. J.; Zhang, C. H.; Yang, J.; Wong, G. K.; J. Am. 1 Chem. Soc. 1990, 112 7389. 12. Brousseau, 111, L. C.; Aurentz, D. J.; Benesi, A. J.; Mallouk, T. E. Anal. Chem. 1997, 69, 688. 13. Brousseau, 111, L. C.; Mallouk, T. E. Anal. Chem. 1997, 69, 679. 14. Vrancken, K. C.; VanderVoort, P.; Gillisdhamers, I.; VanSant, E. F.; Grobet, P.; J. Chem. Soc., Faraday Trans. 1992, 88, 3197. 15. Pfleiderer, B.; Albert, K.; Bayer, E. J. Chromatogr. 1990, 506, 343. 16. Kohli, P.; Blanchard, G. J. Langmuir, 1999, 15, 1418. 17. Chouinard, P. M.; Bartlett, P. A. J. Org. Chem. 1986,51, 75. 18. Kohli, P.; Scranton, A. B.; Blanchard, G. J.; Macromolecules 1998, 31, 5681. 19. Rabinowitz, R. J. Org. Chem. 1961, 26,5152. 20. Olson, K. G.; Butler, G. B.; Macromolecules 1984, I 7, 2480. 21. Kohli, P.; Blanchard, G. J .; Langmuir 2000, 16,4655. 22. Beyer, C.; Bohanon, T. M.; Knoll, W.; Ringsdorf, H.; Langmuir 1996, 12, 2514. -l86- 23. Duevel, R. V.; Corn, R. M. Anal. Chem. 1992, 64, 337. 24. Van Ryswyk, H. B.; Turtle, E. D.; Watson-Clark, R.; Tanzer, T. A.; Herman, T. K.; Chong, P. Y.; Waller, P. J.; Taurog, A. L.; Wagner, C. E.; Langmuir 1996, 12, 6143. 25. The concentration of Zr“ was calculated using the raw experimental data and the sensitivity factors for the elements found. The corrected data indicate that 12.9% of the atoms detected by XPS are Zr“ in the four-layer films. 26. Kohli, P.; Blanchard, G. J .; Langmuir, 2000, 16, 695. -l87- Chapter 8 Conclusions and Future Prognosis The work presented in this dissertation has focused mainly on designing materials and new assembly strategies, where we are able to manipulate material property in a controlled manner. We have demonstrated three distinct levels of control over these materials. First, we have demonstrated our ability to control the identity of the functional groups incorporated into the matrix of the material by simply varying the substituted aniline used to prepare the maleimides. Second, by choosing the appropriate vinyl or vinyl ether co-monomer, we have been able to demonstrate several simple covalent routes to multilayer assembly. The third level of control demonstrated in this work is the ability to control the loading density of these materials at the interface. This level of control is realized through simple acid catalysis and/or dehydration of the various covalent interlayer linking reactions. The ability to control the loading density of the functional moieties at the interface can prove useful in the design of materials that can be used as sensors and in conductive interfaces. With these levels of control demonstrated, we next focused on the design of specific interfaces where we attempt to control the adsorption/desorption behavior of the interface. In this body of work, we attempt to design interfaces where there is a “built-in” chemical potential gradient. Here, we create a gradient in the dipole moment of the side-groups incorporated into the layered polymer matrix, where the layer-by-layer chemistry employed to assemble these materials allows the 188 introduction of different moieties in each layer. Subsequent to polymer multilayer assembly, we measure the vapor phase adsorption isotherm for the interfaces in the presence of methanol, acetone and hexane. To determine the adsorption isotherms, we employ an apparatus designed and assembled in the Blanchard lab that allows the simultaneous acquisition of adlayer mass and thickness change as a function of adsorbate vapor pressure. These data in concert allows us to determine the energetics of adsorbate-adsorbent interaction, monolayer coverage and specific surface area. It should be noted that our inability to determine the specific surface area for our systems speaks directly to the porous/disordered nature of our interfaces. We have chosen to fit our adsorption data to the BET equation and recovered values consistent with physisorption of the adsorbate molecules on the interfaces. In all cases, the adsorption was reversible by simply shutting off the adsorbate vapor supply to the interface. We have shown from fits of our data to the BET isotherm model that we do have the ability to control the adsorption/desorption properties of the interfaces that we prepare (Chapter 6). This control was realized when we (1) varied the side group functionalities in the layers and (2) when we varied the order of layer deposition for our mixed polymer interfaces. Additionally, we also demonstrated our ability to design interfaces for specific purpose. One example is the assembly of 4-hydroxyphenylmaleimide-co-ethyl diisopropylphosphonate vinyl ether into multilayers, where the layer assembly is accomplished though covalent linkages. Once we have the layers assembled, we remove the isopropyl groups and use the interface to sequester metal ions from solution. 189 8.2 Future Prognosis The work we present in this dissertation serves as an opening to a variety of new areas that will be under further exploration within the Blanchard group. With respect to the preparation of interfaces characterized by a chemical potential gradient, the next logical step is the assembly of more complex structures and the measurement of their adsorption isotherms. The aim of this next step is the assembly of three and four layered stacks and varying the position of the different polymer layers. One key issue with respect to design, is how best to assemble the layers to create an interface that behaves in a predetermined manner. We want to determine, in a more qualitative way the effects of the dipole moment gradient on the adsorption properties of the interfaces. It should be noted that, while we represent the gradients as being one continuous profile, this is not the case due to the fact that the dipole moments of the side-groups can be significantly different from layer to layer and intralayer dipole- dipole interactions remain to be accounted for. The difference in the dipole moments of the side-groups will likely lead to a stepwise variation, and the real issue now becomes how large or small do these steps need to be before noticeable change or control is realized. While it may seem rather intuitive that the larger the difference between the dipole moment, the greater the effect should be, this may not necessarily follow. If the steps are significantly large, then the isotherms may actually become very complex as we may begin to observe mixed mechanisms (e.g. absorption, partitioning and/or capillary condensation) due to the fact that the layers begin to behave as individual entities within the matrix. To this end, a more smooth and regular decrease may lead to a more uniform gradient profile. Before moving on to 190 ram: further studies, the first step should be to estimate the dipole moments of the side groups using semi-empirical calculations. Based on the results of these calculations, interfaces can then be designed to vary the slopes of the gradient profiles and to determine the best profile to optimize interfacial behavior in a predetermined manner. A second area that needs to be explored is the use of a more extensive library of adsorbates to determine the isotherms. With a body of data that allows us to determine how best to assemble the interface to achieve a desired behavior, and having information that tells us how these interfaces respond to specific adsorbate molecules, we can begin to design specific separation systems. If we begin to think about designing chromatographic stationary phases, for instance, we can optimize the system in a quick and easy manner by performing these types of measurements rather than having to pack columns and optimize the separations. While these isotherm determinations are gas-solid measurements, it is also possible to perform liquid-solid measurements as well. In this scheme, a technique such as surface plasmon resonance (SPR) can be used to acquire the isotherm of liquid phase adsorption at the interface. Here, it is possible to determine mass transport of bulk liquid adsorbates in a manner similar to QCM measurement. Another issue worth exploring is surface organization as a function of adlayer density. In chapter 4 we noted the acid enhanced deposition of polymer materials through acid catalysis and/or dehydration. In order to get a better understanding of these interfaces it would be useful to determine how these materials assemble/organize on the surface. To perform these measurements, a technique such as AFM or SEM can be used. From these measurements, we would be able to 191 determine the formation (or not) of islands on the surface. Additionally, electrochemical measurements, as well as isotherm measurements could prove useful, where we can determine the role of the adlayer in mass transport of an electroactive of vapor phase species to the interface. These types of measurements would allow for the design and optimization of the interfaces with specific functions. 192 l11111111111I