PLACE IN RETURN BOX to remove this Checkout from your "cord. TO AVOID F INES mum on or More data duo. DATE DUE DATE DUE DATE DUE Mun-9.1 ———— SYNTHESIS AND CHARACTERIZATION OF POLYMERIC PSEUDOCROWN ETHERS AND REVERSIBLE BLOCK/CRAFT COPOLYMERIC EMULSIFIERS BASED UPON POLYMER COMPLEXATION By Arvind Mohan Mathur A DISSERTATION Submitted to Michigan State University in partial fulfillment Of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemical Engineering 1996 ABSTRACT SYNTHESIS AND CHARACTERIZATION OF POLYMERIC PSEUDOCROWN ETHERS AND REVERSIBLE BLOCK/GRAFT COPOLYMERIC EMULSIFIERS BASED UPON POLYMER COMPLEXATION By Arvind Mohan Mathur The complexation properties of poly(ethylene glycol) (PEG), arising due to its Lewis base character, have been exploited for the development of two materials. First, a scheme for the synthesis of polymeric pseudocrown ethers has been devised based upon ion-induced templatization of oligomeric ethylene glycol diacrylates (PEGDA) in which the oligomer assumes a circular conformation with the end groups in close proximity. A combination of experimental and molecular modeling studies was used to characterize the templatization process and design a one-step free-radical copolymerization of templated PEGDA oligomers with a hydrophilic comonomer to produce crosslinked pseudocrown ether (PCE) resins. Experimental evidence of templatization included the PEGDA- induced solubilization of heavy metal salts into organics, proton NMR peak shifts, and excimer fluorescence of pyrene end-labeled PEG oligomers. Molecular dynamics simulations provided visualization of the templated conformations, and demonstrated a decrease in the mean end-to-end distance in the presence of a cation for both the PEGDA and the pyrene end-labeled oligomers. Ion-binding studies revealed the effects of the monomer to solvent ratio. and the presence of a template cation (N i”) during synthesis on the ion-binding capacity of the resins. The experimentally observed capacities of 0.3 mg of Ni2+ per gram of polymer are comparable to that of previously reported polymer-bound crown ethers. The polymer complexation properties of PEG were also exploited in the design of polymeric block/graft copolymeric emulsifiers comprising of a poly(methacrylic acid) (PMAA) backbone and PEG grafts. The swelling properties of crosslinked PMAA-g- PEG hydrogels revealed that changes in pH or solvent composition may be used to disrupt the complex. The reversible oil emulsification and interfacial tension properties of uncrosslinked PMAA-g—PEG copolymers were studied as a function of pH and molecular architecture. Under complex-promoting (acidic) conditions these copolymers exhibited surface active behavior (as evidenced by their ability to lower the oil-water interfacial tension and form oil-in-water emulsions) while under basic conditions the complex was broken and the surface activity was lost. The polarity sensitive fluorescence of solubilized pyrene was used to study aggregate formation as a function of pH and to determine a critical aggregate concentration as a function of molecular architecture. To my parents and my wife iv ACKNOWLEDGMENTS I would like to express my gratitude to several persons who have contributed to the successful completion of this dissertation and have made my stay at Michigan State University enjoyable. My major professor and mentor, Dr. Alec Scranton, has contributed in more than one way towards the accomplishment of the objectives of this research and my evolution as a scientist. I will be forever grateful for his guidance, encouragement and enthusiasm. I would like to thank the members of my committee DrS. Steven Boyd, Daina Briedis and Bob Ofoli for their help and use of equipment in their laboratories. Dr. John Klier at Dow Central Research, deserves special thanks for his interest, input and collaboration. Thanks are also due to the Department of Chemical Engineering, specially to Julie Caywood, Candy McMaster, and Faith Peterson, for being such good friends. I would like to acknowledge financial support by a grant (# CTS-9110163) from the National Science Foundation; the Research Excellence Fund of the State of Michigan; and the Environment Enhancement Grant from the Dow Chemical Company. Several members of my research group have contributed towards a gratifying and healthy research environment. For this, I wish to express my sincere thanks to Bharath, Dave, Eric, Jeff, Julie, Katy, Scott, Shail and Vijay. I would also like to thank several undergraduates who I had the pleasure of working with and in particular, Brenda Becker and Amanda Meeks, for their valuable help in the laboratory. I am grateful to Dilum, Himanshu, Irma, Murty, Ram, Sanjay and their significant others, for their friendship and all the good times we shared. I will always remember the trips to the Upper Peninsula, the Broadway shows and other such activities that made my stay at Michigan State much more complete. I cannot even begin to express my gratitude to my wife Shalini, whose love, support and understanding made this dissertation possible. I am grateful for her patience and words of encouragement when I needed them the most. Finally, this dissertation could not have been completed without the blessings and good wishes of our families, for which I am eternally grateful. vi TABLE OF CONTENTS LIST OF TABLES xiii LIST OF FIGURES xiv CHAPTER 1 BACKGROUND AND MOTIVATION l 1.1. INTRODUCTION .......................................................................................................... 1 1.2. POLYMERIC PSEUDOCROWN ETHERS - MOTIVATION ................................................. 2 1.2.1. Heavy Metal Ions in Aqueous Industrial Wastes ............................................. 2 1.2.2. Current Methods for Removing Heavy Metals from Water ............................ 4 1.2.3. Need for New Separation Methods for Heavy Metal Ions ............................... 7 1.3. POLYMERIC PSEUDOCROWN ETHERS - BACKGROUND ............................................... 8 1.4. MOLECULAR MODELING OF MACROMOLECULES AND LIGANDS .............................. l 1 1.4.1. Molecular Mechanics ..................................................................................... 12 1.4.2. Molecular Dynamics ...................................................................................... 13 1.5. REVERsIBLE COPOLYMERIC EMULSIFIERS - BACKGROUND AND MOTIVATION ........ 15 1.5.1. Need for the Development of Multiblock Reversible Emulsifiers ................ 19 1.5.2. Polymer Complexation Through Hydrogen Bonding .................................... 22 1.6. REFERENCES ............................................................................................................ 25 Vii CHAPTER 2 OBJECTIVES OF RESEARCH 29 CHAPTER 3 ION-SOLUBILIZATION AND ION INDUCED TEMPLATIZATION BY OLIGO(ETHYLENE GLYCOL) AND ITS DIACRYLATES 33 3.1. INTRODUCTION ........................................................................................................ 33 3.2. EXPERIMENTAL METHODS AND TECHNIQUES .......................................................... 38 3.2.1. Ion—Binding Studies of Oligomeric PEG and PEGDA .................................. 38 3.2.2. NMR Studies .................................................................................................. 39 3.2.3. Molecular Modeling ....................................................................................... 39 3.3. RESULTS AND DISCUSSION ...................................................................................... 40 3.3.1. Ion-Binding Studies with Oligomeric PEG ................................................... 40 3.3.2. NMR Studies .................................................................................................. 41 3.3.3. Molecular Dynamics Simulations .................................................................. 43 3.4. CONCLUSIONS .......................................................................................................... 51 3.5. REFERENCES ............................................................................................................ 53 CHAPTER 4 EFFECT OF ION-INDUCED TEMPLATIZATION ON THE END-TO-END DISTANCE OF PYRENE END-LABELED TETRA- AND PENTA-ETHYLENE GLYCOL 54 4.1. INTRODUCTION ........................................................................................................ 54 4.2. EXPERIMENTAL METHODS AND TECHNIQUES .......................................................... 57 4.2.1. Synthesis of Pyrene End-Labeled Oligomeric Ethylene Glycols .................. 57 4.2.2. Fluorescence Studies ...................................................................................... 59 4.2.3. Molecular Dynamics Simulations .................................................................. 6O viii 4.3. RESULTS AND DISCUSSION ...................................................................................... 61 4.3.1. Fluorescence Studies ...................................................................................... 61 4.3.2. Molecular Dynamic Simulations ................................................................... 65 4.4. CONCLUSIONS .......................................................................................................... 73 4.5. REFERENCES ...................................................................... ~ ...................................... 77 CHAPTER 5 SYNTHESIS AND ION-BINDING PROPERTIES OF POLYMERIC PSEUDOCROWN ETHERS 79 5.1. INTRODUCTION ........................................................................................................ 79 5.2. EXPERIMENTAL METHODS AND TECHNIQUES .......................................................... 83 5.2.1. Synthesis of Polymeric Pseudocrown Ethers ................................................. 83 5.2.2. Removal of Template Cation And Conditioning Of Resin ............................ 85 5.2.3. Ion-Binding Studies ....................................................................................... 85 5.2.4. Column Regeneration and Determination of Concentration of Eluted Solution ......................................................................................................... 87 5.3. RESULTS AND DISCUSSION ...................................................................................... 88 5.3.1. Polymeric Pseudocrown Ether Synthesis ..................................................... 88 5.3.2. Ion-binding Studies ...................................................................................... 89 5.3.3. Polymer Regeneration .................................................................................. 93 5.4. CONCLUSIONS .......................................................................................................... 94 5.5. REFERENCES ............................................................................................................ 96 CHAPTER 6 SWELLING PROPERTIES OF POLY(METHACRYLIC ACID-G-ETHYLENE GLYCOL) HYDROGELS - EFFECT OF SOLVENT 97 ix 6.1. INTRODUCTION ........................................................................................................ 97 6.2. EXPERIMENTAL METHODS AND TECHNIQUES ........................................................ 101 6.2.1. Hydrogel Synthesis ...................................................................................... 101 6.2.2. Equilibrium Swelling Studies ...................................................................... 103 6.2.3. Molecular Mechanics Studies ...................................................................... 104 6.3. RESULTS AND DISCUSSION .................................................................................... 105 6.3.1. Hydrogel Swelling Studies .......................................................................... 105 6.3.2. Molecular Modeling of PMAA-PEG Complexation ................................... 113 6.4. CONCLUSIONS ........................................................................................................ 116 6.5. REFERENCES .......................................................................................................... 119 CHAPTER 7 SYNTHESIS AND EMULSIFICATION PROPERTIES OF REVERSIBLE BLOCK/CRAFT POLYMERIC EMULSIFIERS BASED UPON POLYMER COMPLEXATION 120 7.1. INTRODUCTION ...................................................................................................... 120 7.2. EXPERIMENTAL METHODS AND TECHNIQUES ........................................................ 126 7.2.1. Copolymer Synthesis ................................................................................... 126 7.2.2. Emulsification Studies ................................................................................. 128 7.2.3. Interfacial Tension Measurements ............................................................... 128 7.3. RESULTS AND DISCUSSION .................................................................................... 130 7.3.1. Emulsification Studies ................................................................................. 130 7.3.2. Interfacial Tension Studies .......................................................................... 130 7.4. CONCLUSIONS ........................................................................................................ 137 7.5. REFERENCES .......................................................................................................... 139 CHAPTER 8 A FLUORESCENCE SPECTROSCOPY STUDY OF THE AGGREGATE FORMATION BEHAVIOR OF REVERSIBLE COPOLYMERIC BLOCK/GRAF T EMULSIFIERS 140 8.1. INTRODUCTION ...................................................................................................... 140 8.2. EXPERIMENTAL METHODS AND TECHNIQUES ........................................................ 144 8.3. RESULTS AND DISCUSSION .................................................................................... 146 8.3.1. Excitation Spectra ........................................................................................ 146 8.3.2. Estimation of the Critical Aggregate Concentration from the Excitation Spectra ......................................................................................................... 150 8.3.3. Emission Spectra .......................................................................................... 154 8.3.4. Estimation of the CAC from Emission Spectra ........................................... 157 8.4. CONCLUSIONS ........................................................................................................ 160 8.5. REFERENCES .......................................................................................................... 161 CHAPTER 9 CONCLUSIONS - 163 9.1. SUMMARY OF RESULTS .......................................................................................... 163 9.1.1. Synthesis and Characterization of Polymeric Pseudocrown Ethers ............. 163 9.1.2. Synthesis and Characterization of Reversible Block/Graft Copolymeric Emulsifiers Based Upon Polymer Complexation ....................................... 167 CHAPTER 10 RECOMMENDATIONS FOR FUTURE WORK - -170 10.1. SYNTHESIS AND CHARACTERIZATION OF POLYMERIC PSEUDOCROWN ETHERS 1 70 10.1.1. Ion-induced Templatization Studies ......................................................... 171 xi 10.1.2. Molecular Modeling ................................................................................. 171 10.1.3. Synthesis and Ion-Binding Studies ........................................................... 172 10.2. SYNTHESIS AND EMULSIFICATION PROPERTIES OF REVERSIBLE BLOCK/GRAFT COPOLYMERIC EMULSIFIERS BASED UPON POLYMER COMPLEXATION ............... 174 10.2.1. Study of Reversible Complexation ........................................................... 174 10.2.2. Copolymeric Emulsifier Synthesis ........................................................... 175 10.2.3. Inclusion of Comonomers ........................................................................ 176 10.2.4. Copolymer Characterization and Emulsification Studies ......................... 176 10.3. ALKALl-SWELLABLE THICKENERS COMPRISING POLY(METHACRYLIC ACID-G- ETHYLENE GLYCOL) COPOLYMERS BASED UPON POLYMER COMPLEXATION ...... 178 10.4. REFERENCES ........................................................................................................ 183 xii LIST OF TABLES Chapter 1 Table 1.1. Classification of surfactants based upon chemical structure. .......................... 17 Chapter 3 Table 3.1. Solubilities of salts upon the addition of PEG 300 .......................................... 41 xiii Figure 1.1. Figure 1.2. Figure 1.3. Figure 3.1. Figure 3.2. Figure 3.3. Figure 3.4. Figure 3.5. Figure 3.6. Figure 3.7. Figure 3.8. LIST OF FIGURES Chapter 1 a) Structure of 18-crown-6 and b) Structure of cryptand-222 .......................... 9 Crown ethers attached as pendent groups and as part of a polymer backbone. .......................................................................................................... 9 Potential application for reversible emulsifiers as an aqueous cleaner useful in the clean up of greased components in service stations. ............................ 22 Chapter 3 Scheme for the synthesis of polymeric pseudocrown ethers based upon a one- step free-radical polymerization of templated polyethylene glycol diacrylate with a comonomer such as hydroxy ethyl methacrylate ................................. 35 Schematic showing the cation complexation ability of oligomeric polyethers. ...................................................................................................... 35 Synthesis scheme of Warshawsky et al. for the synthesis of polymeric pseudocrown ethers. ....................................................................................... 36 ‘H NMR spectra of a) PEG 300 and b) PEG 300 in the presence of chromium chloride.(hydroxyl) protons. ........................................................................... 42 MD 50 ps simulation results for the mean end-to-end distance of PEGDA with different number of EG repeating units. ................................................. 44 Molecular representation of a PEGDA (n = 4) conformation templated around a Na“ ion showing charged surface of the molecule. Shown here is one conformation extracted from a 50 ps run. ................................................ 46 Effect of different simulation times on the mean end-to-end distance for PEGDA with Na+ with various numbers of EG repeating units. ................... 48 Normalized frequency distribution of end-to-end distance for PEGDA (n = 4) with and without Na+. .................................................................................... 49 xiv Figure 3.9. Figure 4.1. Figure 4.2. Figure 4.3. Figure 4.4. Figure 4.5. Figure 4.6. Figure 4.7. Figure 4.8. Figure 4.9. Cumulative frequency distribution of end-to-end distance for PEGDA (n = 4) with and without a Na+ ion. ........................................................................... 50 Chapter 4 Scheme for synthesis of pyrene end labeled oligoethylene glycols. .............. 58 Fluorescence emission spectra of (A) pyrene end-labeled pentaethylene glycol and (B) l-pyrenebutyric acid. The excitation was at 322 nm and both samples were at a concentration of 10‘5 M. .................................................... 62 Fluorescence profiles of pyrene end-labeled tetraethylene glycol in THF solvent (10'5 M) in the (A) absence and (B) the presence of nickel chloride. The excitation frequency is 322 nm. .............................................................. 64 Fluorescence emission spectra of pyrene end-labeled pentaethylene glycol (10'5 M) in chloroform in the (A) absence and (B) presence of tin chloride. Excitation is at 322 nm. .................................................................................. 66 Comparison of excimer-to-monomer ratio for pyrene end-labeled pentaethylene glycol (10'5 M) in chloroform and THF in the presence and absence of nickel chloride. ............................................................................. 67 Starting conformation of end-labeled pentaethylene glycol in the presence of a sodium cation. .............................................................................................. 69 Pyrene end-labeled pentaethylene glycol conformation with templating cation showing pyrene excimer and the end-to-end distance (ETED). ..................... 70 Comparison of the normalized frequency distribution of end-tO-end distance for TEGDA and Py-TEG-Py demonstrating the effect of pyrene end- labeling. .......................................................................................................... 71 Normalized frequency distribution of end-to-end distance for pyrene end- labeled pentaethylene glycol in the presence and absence of a templating sodium cation .................................................................................................. 72 Figure 4.10. Comparison of normalized frequency distribution of end-to-end distance Figure 5.1. obtained from molecular dynamics simulations for Py-TEG-Py and TEGDA, both with a templating Na+ cation ................................................ 74 Chapter 5 Schematic of free-radical polymerization scheme for the synthesis of polymeric pseudocrown ethers. ...................................................................... 81 XV Figure 5.2. Figure 5.3. Figure 6.1. Figure 6.2. Figure 6.3. Figure 6.4. Figure 6.5. Figure 6.6. Figure 6.7. Figure 7.1. Figure 7.2. Figure 7.3. Experimental setup for ion-binding and regeneration studies. ....................... 86 Experimental ion-binding capacities for polymeric pseudocrown ethers (PCES) and “control” p(HEMA) hydrogels as a function of the monomer to solvent ratio during synthesis. ........................................................................ 90 Chapter 6 Synthesis scheme for the preparation of poly(methacrylic acid-g-ethylene glycol) self-associating hydrogels. ............................................................... 102 Effect of solvent on the equilibrium degree of swelling for a PMAA-g-PEG gel synthesized in 50% solvent (50 % monomers). ...................................... 106 Equilibrium swelling profiles showing the effect of polymerization conditions for swelling in ethanol. ............................................................... 108 Equilibrium swelling profiles of crosslinked PEG-g-PMAA (50/50) and PMAA homopolymer (50/50) in methanol. ................................................. 1 10 Equilibrium Swelling profiles of crosslinked PMAA-g-PEG (50/50) and PMAA (50/50) in acetone. ........................................................................... 1 12 A molecular representation of a PMAA segment with four repeating units complexed with a PEG segment of the same length showing 1:1 hydrogen bonding between the acid protons and the ether oxygens. ........................... 115 A POLYGRAF molecular simulation depicting a 1:1 hydrogen bonding between a PMAA and PEG segment both with 14 repeating units. ............. 117 Chapter 7 (A) Reduction of the viscosities of heavy oils by the formation of oil-in-water emulsions, and (B) A schematic showing a treatment process that may be used for the pumping of heavy oils by forming oil-in-water emulsions. ..... 121 Schematic and mechanistic depiction of the block/graft copolymer, consisting of a poly(methacrylic acid) backbone and poly(ethylene glycol) grafts, under acidic or complex promoting conditions and under basic or complex breaking conditions. .................................................................................................... 123 Hydrogen bonded complexes formed between poly(methacrylic acid) and poly(ethylene glycol) units in (A) non-grafted and, (B) grafted configurations. .............................................................................................. l 25 xvi Figure 7.4. Figure 7.5. Figure 7.6. Figure 7.7. Figure 7.8. Figure 8.1. Figure 8.2. Figure 8.3. Figure 8.4. Figure 8.5. Figure 8.6. Figure 8.7. Schematic showing the free-radical polymerization scheme for the synthesis of reversible block/graft emulsifiers. ............................................................ 127 Schematic of the methodology adopted for the oil emulsification studies... 129 Emulsification behavior of two different block/ graft copolymers (10:1 and 20: 1) as a function of pH showing the percent oil emulsified (V/v) as a function of pH ............................................................................................... 131 Interfacial tension between the oil (methyl laurate) phase and the aqueous phase containing the emulsifiers, as a function of pH. Data is presented for the 10:1 and the 20:1 copolymeric emulsifiers ............................................. 133 Schematic representation of the possible emulsification mechanism adopted by the reversible emulsifiers (A) under acidic and basic conditions, and (B) the steric stabilization that may result. ......................................................... 135 Chapter 8 Pyrene excitation spectra for 20:] PMAA-g-PEGI 000 copolymer solutions of various concentrations with constant pyrene concentration (6 x 10'7 M). Emission was monitored at 397 nm .............................................................. 147 Plot of the pyrene excitation ratio (1342/1337) as a function of pH for 20:1 PMAA-g-PEGIOOO copolymer solutions with 6 x 10'7 M pyrene. The emission was monitored at 397 nm. ............................................................. 149 Plot of percent pyrene emulsified versus pH for a 20:1 PMAA-g-PEGIOOO copolymer solutions calculated from excitation ratio data. .......................... 151 A plot of the pyrene excitation ratio as a function of copolymer concentration for 10:1 and 20:1 PMAA-g-PEGlOOO copolymers at pH 2. ........................ 153 Pyrene emission spectra for 20:1 PMAA-g-PEG1000 copolymer solutions at pH 2. Excitation was at 322 nm and pyrene concentration in each sample was6x 10'7 M. ............................................................................................. 155 A plot of the pyrene emission ratio 1373/1339 (II/I3) as a function of pH for 20:1 copolymer solutions with a pyrene concentration of 6 x 10'7 M. Excitation was at 322 nm. .............................................................................................. 156 A plot of the pyrene emission ratio (excitation at 322 nm) versus concentration for 10:1 and 20:1 PMAA-g-PEGIOOO copolymers. .............. 158 xvii Chapter 10 Figure 10.1. Conformational changes resulting from going from an acidic (A) to a basic (B) environment causing the polymeric chain to expand. .......................... 180 Figure 10.2. A plot of Brookefield viscosity as a function of pH for 1:1 PMAA-g- PEG1000 (with a MAA to EG ratio of 1 :1) copolymer solutions at two concentrations. ............................................................................................ 1 82 xviii CHAPTER 1 BACKGROUND AND MOTIVATION 1.1. Introduction This chapter presents the background and motivation behind this research. The motivation behind development of novel polymeric pseudocrown ethers, which have potential applications in the separations industry, will be presented first. Since molecular simulations have been utilized to elucidate phenomenon at the molecular level, an overview of molecular modeling techniques and a brief review of their applications in macrocyclic ligand simulation will be discussed next. Following this a brief overview of potential applications which provided the motivation for the development of novel reversible block/graft emulsifiers will be presented. This will be followed by a short introduction into the phenomenon of hydrogen-bonded complexation between polymeric acids and bases. The unifying thread in all this work is the utilization of the complexation properties of poly(ethylene glycol). Other than its commonly knownl uses (such as, in cosmetics and toiletries, detergents and cleaners, inks, paints and coatings, etc.) poly(ethylene glycol) has the ability to form complexes with electron accepting species because of its Lewis base character. The cation complexation ability of poly(ethylene glycol) (PEG) has been used to design a novel technique for the synthesis of polymeric pseudocrown ethers. The ability of PEG to form reversible hydrogen bonded complexes with species possessing a Lewis acid character (such as poly(methacrylic acid)) has been utilized to design reversible block/graft emulsifiers. This chapter will lead into the specific objectives of this research which are addressed in the following chapter. 1.2. Polymeric Pseudocrown Ethers - Motivation 1.2.1. Heavy Metal Ions in Aqueous Industrial Wastes The control of heavy metal discharges to the environment promises to be a significant industrial challenge in the coming years. Although heavy metals such as lead, chromium, mercury, nickel, tin, copper, zinc, and cadmium are commonly present in ionic form in aqueous waste streams, their toxicity to plants, animals, and humansz'5 has led to increasingly stringent industrial emission standards.5 Heavy metals accumulate in living organisms and can cause chronic diseases and even death. Awareness of the potential dangers of heavy metals in industrial wastes was aroused over twenty-five years ago when aqueous industrial eflluents containing mercury and cadmium ions poisoned the food chain in Japan, causing the deaths of thousands of animals and over one hundred people.2 More recently, there has been mounting concern over the possibility of adverse health effects from exposure to heavy metal ions at relatively low concentrations previously thought safe. For example, lead is suspected of causing subtle, irreversible brain damage in growing children,5 while heavy metal ions in general are under investigation for their possible gene-mutation effect in living organisms.6 Heavy metal ions appear in industrial waste streams from a variety of sources. Relatively large concentrations of heavy metal ions occur in the waste streams from plants in which a heavy metal is an integral part of the manufacturing process. For example, waste streams from battery manufacturers and electroplating facilities typically contain heavy metals in concentrations of over one hundred parts per million.5 Other industries which produce waste water with relatively large concentrations of heavy metal ions include mining,4 steelworks,s and a variety of specialty chemical plants.6 Many processes which do not directly involve heavy metals may produce waste streams which contain significant concentrations of heavy metal ions. For example, zinc and chromium are often used as corrosion inhibitors in cooling towers, and are present in concentrations approaching ten parts per million in the blowdown streams? Furthermore, corrosion of catalysts or other metals may lead to Significant amounts of heavy metal ions in the aqueous waste streams from any chemical plant. The United States Environmental Protection Agency has established standards for industrial discharges to inland waters. For heavy metals, the discharge limits are typically in the range of fifty parts per billion.2 The heavy metal ions which enter natural bodies of water often concentrate in the sediments of lakes and oceans.5 Ground water often exhibits a very low concentration of heavy metal ions because they are immobilized in the top layers of the soil. This fact has led to increasing concern over the accumulation of heavy metals in the soils.7 In fact, at the Sapp Battery Superfund site in northern Florida, the upper soil contains lead in concentrations approaching 135,000 parts per million.8 Extensive ground water contamination has been observed in areas in which the soil is very sandy9 and unable to immobilize the metal ions. Finally, if heavy metal laden water is routed to a municipal water treatment plant, the metal ions concentrate in the sewage Sludge. The subsequent use of the sludge for agricultural fertilizer further contributes to the accumulation of heavy metals in the soil.7 Furthermore, heavy metals have been observed to inhibit biological sewage treatment processes if present in concentrations above a few parts per million.5 1.2.2. Current Methods for Removing Heavy Metals fiom Water The most common technique for removing heavy metal ions from aqueous streams is precipitation.'°"' This method is based upon the fact that the aqueous solubility of the heavy metal ions depends strongly upon pH. Under basic conditions, the metals typically precipitate as the corresponding hydroxide salts. The minimum attainable concentration of a heavy metal ion is determined by the magnitude of the solubility minimum, typically about 500 parts per billion.” If the waste stream contains two or more ions which exhibit solubility minima at different values of pH, the process must be run at a compromise pH value, thereby undermining the efficiency of the purification.'0 The sedimentation of the heavy metal precipitates usually takes several hours to complete, even with the addition of flocculating agents.5 Often precipitation is used only for an initial treatment of a waste stream, with fiIrther treatment by other methods being required to decrease the heavy metal ion concentration to desirable levels.5'IO A variety of other methods have been suggested for removing heavy metal ions from aqueous wastes, including activated carbon adsorption}10 ion exchange,”"'2 reverse osmosis?“ liquid-liquid extraction}l5 evaporation,5"6 freezing,” foam fractionation or 18.19 I flotation, and micellar enhanced ultrafiltration?“2 The latter two methods use surfactants in the separation process. In foam or bubble fractionation, positive heavy metal ions are attracted to anionic surfactant molecules at an air-liquid interface. Typically air is bubbled through the solution, allowing the ion laden surfactants to adsorb on the interface to be carried to the surface foam. In micellar enhanced ultrafiltration the surfactant forms micelles around the contaminant, rendering it too large to pass through an ultrafiltration membrane. Although this technique has been used primarily for removal of organic contaminants,”2| with an appropriate surfactant it could possibly be used for removing ions as well. All of the purification techniques developed to date have significant limitations when considering the removal of heavy metal ions from aqueous industrial wastes. Because heavy metal ions are typically present in relatively small concentrations, methods such as evaporation and freezing which inherently remove the water rather than the ions from the water, are relatively unattractive. Similarly, because the distribution coefficient of the metal ion is essentially insensitive to its concentration, liquid-liquid extraction is not well suited for removing heavy metals from waste streams.22 A major shortcoming of all of the techniques developed to date is a lack of selectivity, which is probably most pronounced for activated carbon. Although activated carbon will absorb most metals which are organically complexed,5 it will also absorb any other organic molecules present in the water. This necessitates frequent regeneration, and undermines the cost effectiveness of the method. For the methods which involve surfactants, the lack of selectivity results in a Significant decrease in the separation efficiency with increasing inert salt concentration.'8"9 Another shortcoming of the present purification techniques is the fact that they do not decrease the total quantity of heavy metal wastes, but merely concentrate the ions into a smaller volume of liquid or solid.2'5 Therefore, under the current technology, the problem of disposal of the concentrated wastes still remains. In many cases, the heavy metal ions themselves have some value, and ideally they would be recycled rather than discarded. However, since most industrial wastes contain several different types of ions, the concentrated wastes will contain a broth of many types of ions, and a separations problem still exists. For this reason, recycling of heavy metal ions has been employed only in a few special cases, such as the recycling of chromium ions in cooling tower blowdown.6 Ion exchange is the technique which is perhaps best suited to the removal of small quantities of an ionic component from water,”23 and has been used extensively for the removal of heavy metals from waste streams.5’lo A particularly attractive feature of ion exchange is that the cost of the separation decreases with decreasing concentration of the ion to be removed. This is important because the concentrations of heavy metals in waste streams are typically relatively small; on the order of hundreds of parts per million. Unfortunately, innocuous ions present in the water tend to reduce the capacity of the resins as well as the efficiency of the separation.24 This fact has led to a great deal of work in the area of selective chelating resins.”26 However, current chelating resins have not met the selectivity and performance requirements”26 for aqueous waste treatment. A second drawback of ion exchange is the problem of disposal of the regeneration wastes. In the ideal case, regeneration results in a doubling of the number of ions which must be discarded. However, in most cases an excess of regenerant is required,23 and the total quantity of waste ions is much higher than the ideal value. 1.2. 3. Need for New Separation Methods for Heavy Metal Ions The future promises to bring increasing concern over the accumulation of heavy metals in the environment due to industrial wastes. Undoubtedly, the threat of heavy metal pollution will lead to the adoption of more stringent effluent standards for industrial processes. This realization has caused many authorsz‘3'5"° to discuss the "zero discharge" goal for heavy metal ions. While absolutely zero discharge is intrinsically impossible, the separation methods in use today are not well suited to this goal. The current methods do not decrease the total quantity of heavy metal wastes, but merely concentrate the wastes. Although recovering and recycling of heavy metal ions has been practiced in a few specialized cases, the current technology is not amenable to this approach because the concentrated wastes that are generated typically contain a mixture of many ions. Further separations would be required for recycling to be feasible. There is a need for a heavy metal ion separation technique which would allow the ions to be removed from the waste stream and recycled without an extra separation step. Such a technique could enhance the economic attractiveness of recycling heavy metal wastes rather than Simply disposing of them, especially in the face of increasing regulatory pressures. A desirable technique should have the following properties: i) it should be capable of selectively removing one distinct type of metal ion from an aqueous mixture containing many ions, and ii) it Should allow the heavy metal to be reclaimed without creating additional ionic wastes. Furthermore, the method Should be cost effective and, ideally, not require existing waste treatment equipment and technology to be completely abandoned. 1.3. Polymeric Pseudocrown Ethers - Background Since the pioneering work of Pedersen”28 cyclic oligomers of ethylene glycol known as crown ethers have been extensively studied for their unique ion-binding properties. Pedersen demonstrated the ability of crown ethers to form stable complexes with metal cations and to solubilize inorganic salts in nonpolar solvents. This ion- binding ability arises from the electron lone pairs on the oxygen atoms of the cyclic 29.30 ether, with the most stable complexes formed with ions that fit snugly in the macrocyclic cavity thereby interacting optimally with all donor Oxygens (see Figure 1.1). Since Pedersen's work on crown ethers several related compounds have been 31-33 synthesized and studied, including cryptands, which are tricyclic ligands; 34.35 coronands, in which nitrogen, sulfur or phosphorous replace oxygen as the electron donor; podands,”36 which are open-chain analogs of the cyclic ligands; and podandocoronands34 which combine the structural properties of podands and coronands. Several reviews on the synthesis and ion-binding of multidentate macrocyclic compounds”39 and host-guest chemistry‘o‘“ provide an excellent overview of the area of supramolecular chemistry that has burgeoned since Pedersen's efforts in 1967. The unique ion-binding properties of macrocyclic ligands make them attractive for a variety of applications, most notably phase transfer catalysis. However these compounds are generally synthesized in low yields and through relatively long reaction (“Do m [0 O] @3637 ROJ k‘L/OJ Figure 1.1. a) Structure of 18-crown-6 b) Structure of cryptand-222 o o _ (OI—tr? 1 [O O] {MILD/CHOW— OH OH Figure 1.2. Crown ethers attached as pendent groups and as part of a polymer backbone. 10 sequences,”36 making them rather expensive. For this reason, investigators developed schemes for attaching these ligands to polymeric supports thereby allowing the ligands to be efficiently retrieved and recycled (see Figure 1.2). Methods for incorporating macrocyclic ligands either into the polymeric backbone or as pendant side groups36‘42’“ have been reported. Synthetic methods which have been used to produce macrocycle-containing polymers include condensation of di-functional macrocycles with appropriate comonomers, free-radical polymerizations of Vinyl crown ethers, and substitutions of macrocycles for leaving groups on existing polymer chains. Macrocycle-containing polymers have been used in a variety of applications in analytical and preparative chemistry, including anion and cation separations, column and thin layer chromatography, salt conversions, and phase transfer catalysis.3"3("“2“““"5“47 Macrocyclic ligands attached to polymeric chains may exhibit different ion-binding characteristics than the corresponding monomeric species. Interactions which affect the affinity and selectivity of polymer-bound macrocycles can arise from the polymer backbone, the point of attachment, neighboring ligands, or other pendant groups. Some of the limitations of crown ethers, such as toxicity (since they readily absorb through the skin) and ease of recovery, could be overcome by immobilization on polymeric chains. However, these materials were also synthesized in low yields with long reaction sequences resulting in a potentially expensive resin. One alternative to immobilizing crown ethers onto polymeric supports is to form crown ethers in-situ during a polymerization reaction. In such a scheme the crown ethers would be produced only during the polymerization process and would thus overcome limitations associated with ll toxicity and recovery. In addition, this scheme would be more attractive if it would involve a one-step and easy to perform reaction which would result in high yield and low cost. Materials, called pseudocrown ethers, offer all of the above advantages. A more detailed description of work related to the development of polymeric pseudocrown ethers and the scheme adopted in this research will be discussed in Chapter 3. 1.4. Molecular Modeling of Macromolecules and Ligands Molecular modeling has emerged as an invaluable tool in understanding the micro-environment of various systems. It has been extensively used in conformational analysis of proteins and for other bio-molecules. For example, conformational analysis of macromolecules have been performed and energies of various conformations calculated using semi-empirical potential energy functions.48 Earlier, involved computer programs were written in order to calculate and minimize the energy of a certain molecule and the visualization of its conformation was done by interpreting radial distribution fimctions in the case of molecular dynamics calculations, or based upon bond lengths and distances in the case of molecular mechanics calculations. With the advent of excellent computing techniques it is not only possible to decrease computation time and thus perform simulations previously considered tedious, but also to visualize the resulting conformations graphically on a computer terminal. Molecular modeling, however, can also have its own limitations49 and is limited by the size of the molecule under study. Molecular modeling may be useful as a tool for the scientific visualization of molecules under a certain set of conditions. It has been broadly divided into the 12 following categories: i) ab initio calculations on atoms and molecules using quantum mechanical fundamentals, ii) semi-empirical and empirical calculations which are extensions of ab initio calculations performed on larger molecules after making certain assumptions while performing the quantum mechanical calculations, iii) molecular mechanics calculations which are designed for larger molecules and are based on force fields, and, iv) molecular dynamics calculations, which are concerned with determining the trajectories of atoms within a molecule over a period of time by solving Newton's equations of motion subject to certain potential energy functions and constraints. While the theoretical chemist uses the first two methods of molecular modeling quite extensively to predict charge densities, extract molecular orbital information, and determine energies of various molecules, the latter two methods are most widely used for macromolecules like polymers and proteins for conformation analysis and energy calculations. 1 . 4. I . Molecular Mechanics Molecular mechanics (MM), which is often referred to as the force field method, is one of the most commonly used methods available to predict geometry. It requires limited computer resources and its results have the same quality with regard to geometry as those obtained by sophisticated quantum methods. The molecule is viewed as a collection of points (atoms) connected by springs (bonds) with different elasticities (force constants). The force holding the atoms together can be described by potential energy functions of structural features like bond lengths, bond angles, non bonded interactions and so on. The combination of these potential energy functions is called a force field. 13 There exists "natural" lengths and angles for bonds and a molecule will assume a geometry with those values in the simple cases. In general, steric, electrostatic and other strain forces must be included too. Because of the intimate connection between structure and energy, molecular mechanics involves both. To find the structure, one necessarily has to examine the energy to find where the energy minima occur. The energy, E, of a molecule in a force field arises from deviations from ideal structural features, and can be approximated by a sum of energy contributions due to bond stretching, bond bending, out of plane bending, torsional energy due to twisting about bonds, van der Waals non bonded interactions, electrostatic interactions and other user specified constraints. The value E is the difference in energy between the real molecule and a hypothetical molecule where all the structural features such as, bond angles and bond lengths, are exactly at their ideal or natural values. These equations are then solved using a minimization routine such as, the steepest descent method, the conjugate gradient method or the Broydon Fletcher steepest gradient method, in order to attain an energy minimum. Various force fields are currently available to describe different sets of molecules. AMBER50 and MM25| were two of the first models or "force fields" to be described and translated into an efficient computer program. Since then various other force fields have been developed, including CHARMM,52 DRIEDING,53 and MM3.54 I . 4. 2. Molecular Dynamics As the name suggests, molecular dynamics simulations are used to perform conformational analysis by solving Newton's equations of motion for the whole molecule. The force between two atoms is the gradient of the potential between them and, knowing 14 various potential energy expressions, one can integrate Newton's equation twice to calculate the new coordinates for an atom. The potential is assumed to be continuous and no hard Sphere forces are considered. Several algorithms, like predictor corrector methods, can be used to solve the resulting differential equations by finite difference techniques. Typical simulations are performed on the order of pico (10“) seconds due to the immense number of calculations involved. Periodic boundary conditions are typically used in molecular modeling with solvents in order to simulate a system having a constant volume and constant number of molecules. Periodic boundary conditions imply that if any molecule moves out of a face of the cube containing the system of molecules, another moves in from the opposite face to keep the number of molecules in the periodic boundary constant. Thus, molecular dynamics Simulations may provide information about molecular behavior and is widely used to estimate energies of molecules. Computer software such as BIOGRAF and POLYGRAF55 (Molecular Simulations Inc.; these programs now have been combined and the currently marketed software is known as CERIUS2) have been developed thereby making it easy to perform MM and MD calculations on a host of biological, organic and polymeric molecules. These packages also possess the ability to graphically display, in real time, changes in conformations of the molecules under study. This consolidation of a graphical interface with algorithms that perform MM and MD calculations has redefined molecular modeling, thereby making it a user-friendly tool to pursue investigations at the molecular level. Various studies have utilized the ability of this software to perform molecular investigations in an easy and visual manner. For example, the binding of phosphines to 15 Cr(CO)5 was studied using MM calculations performed56 on BIOGRAF. MD calculations using SYBYL (Version 5.3, Tripos Associates) have been used to predict equilibrium statistics and study the conformational dynamics of poly(dialkylsiloxanes).57 Molecular Simulation techniques are now being used to characterize the structure and properties 0f POIYmeric systems,”59 thereby making this technique viable for predicting mechanical and thermal properties which may lend a new dimension to the design and synthesis of new materials. POLYGRAF has been used to derive the phase diagrams of binary mixtures59 and to test the F lory-Huggins theory for a number of binary systems and to characterize their miscibility behavior. Molecular dynamics Simulations have also been used to obtain the diffusion coefficient of a penetrant gas in a polymeric material."0 Finally, molecular dynamics of aromatic polyesters performed using the AMBER code have yielded experimentally accessible properties such as persistence length.6| 1.5. Reversible Copolymeric Emulsifiers - Background and Motivation Surfactants, or surface active agents, are among the most versatile products of the chemical industry. They are used as detergents and cleaners, and in pharmaceuticals, drilling muds, motor oils, floatation agents, etc. The surface active nature of these materials arises from their ability to lower the interfacial tension between two interfaces thereby permitting miscibility by partitioning partially into both phases. The unique chemical structure that permits such behavior is the presence of two distinct groups within the same surfactant molecule, thereby imparting to them an amphipathic structure. 16 Usually, the two active segments in a surfactant which account for its surface activity are a hydrophilic (or lyophobic) group and a hydrophobic (lyophilic) group. AS the terms suggest, the hydrophilic nature is essential for allowing miscibility in the aqueous phase, whereas the hydrophobic character is required for partitioning into the oil phase. Typically above a certain concentration, known as the critical micelle concentration (CMC), surfactants form aggregates of molecules due to the amphiphilic nature of each molecule. These aggregates, or micelles as they are commonly referred to, may consist of an inner hydrophobic core (in aqueous solutions) while the hydrophilic moieties are extended into the continuous phase. Most surfactants can thus emulsify oil by entrapping small oil droplets into the hydrophobic domains. The emulsions so formed are stabilized by the presence of surfactant molecules on the interface which inhibit and often prevent coalescence of two droplets. Numerous texts are available which describe various kinds of surfactants commercially available. In general surfactants may be broadly classified as follows: anionic, cationic, and non-ionic based upon the surface active portion of the molecule (see Table 1.1). While most common types of surfactants fall into the first two categories most research is currently focused on the development of novel non-ionic surfactants. Non-ionic surfactants are usually either block or graft copolymers possessing amphiphilic nature. Block and graft copolymers are used as dispersing agents and stabilizers for particles and emulsions in water. In these systems, one block of the dispersing agent is designed to absorb onto the surface of the stabilized particle, and the other block is designed to extend into the continuous phase. Therefore, each particle possesses a layer of absorbed polymer chains which provides steric stabilization, ultimately arising from 17 Table 1.1. Classification of surfactants based upon chemical structure. Type of Surfactant Structure/Example a) Anionic O // R—C Sulfates, sulfonates, carboxylates and \ - + O ...Na phosphates Fatty Acid Soap 1i R—O—Ir—O' ...Na O....Na+ Alkyl Phosphate O R—O—isl—Q’...Na (I; Alkyl Sulfate b) Cationic Alkyl amines (Cg-Cm), Primary, secondary, tertiary and quaternary amines R I + R—N —R'...Cl RI Quaternary Amine c) Non-ionic Block(diblock, triblock), Graft Hydrophilic block -oxyethylenes Hydrophobic block - alkyl or aryl chains Rfo —CH2-CHzi‘OH n-Alkyl polyoxyethylene ether 18 the crowding of polymer chains which would occur if two particles approached one another. Polymeric surfactants have gained considerable importance in recent years resulting in the development of several di-and tri-block copolymers. Several polymeric surfactants are now commercially available. Among these, di- and tri-block copolymeric surfactants are the most widely used as emulsifiers and agrochemicals.”64 Diblock and triblock copolymers of ethylene oxide (EO) and propylene oxide (PO) have received the most attention due to their relatively simple architecture and the ability to “tailor-make” the copolymers by controlling the lengths of the hydrophilic (EO) and hydrophobic (PO) blocks. Block copolymers are most commonly synthesized using complex anionic polymerizations which are constrained by a variety of factors. For example, block copolymers may be formed by sequential monomer addition to a living anionic polymerization and while this technique is extensively utilized to produce a variety of di- and triblock copolymers, it has several limitations. The anionic polymerizations are sensitive to trace amounts of impurities such as oxygen, carbon dioxide, and water and therefore must be carried out with rigorously clean and dry reagents and reactors. For these reasons these reactions are typically carried out in small volumes and the resulting polymers are expensive. Anionic polymerizations may also be applied to only a limited group of monomers, and may not be used to polymerize vinyl monomers containing electron donating substituents in the DI position or polar monomers containing substituents that are reactive towards nucleophiles. Important classes of monomers such 19 as a-Olefins, vinyl esters, vinyl ethers, as well as some acrylates cannot be polymerized anionicaly. In contrast, free radical polymerizations are much more versatile and may be applied to nearly any vinyl monomer. Graft or “comb” copolymer surfactants consisting of a hydrophilic backbone and a hydrophobic graft or vice versa have also been synthesized and evaluated for their surfactant properties. This area of non-ionic surfactants consisting of block and comb or graft copolymers has been the focus of many reviews.”67 1.5.]. Need for the Development of Multiblock Reversible Emulsifiers Simple diblock copolymers have also been used as polymer compatibilizers and dispersion stabilizers for many years. Theoretical work by Noolandi68 has suggested that multi-block copolymers would be more efficient as compatibilizers than di- or triblock copolymers. His results suggest that less of the multi-block copolymer would be lost into the bulk phase as micelles or mesophases than the di- or triblock systems. Similarly, multi-block dispersing agents and stabilizers should be more efficient than the common di or triblock systems. This multiblock molecular design strategy is aimed at reducing the ability of the surfactant to form a separate phase, thereby resulting in an increase in the interfacial activity in a non-equilibrium system with a large interfacial area. Noolandi also suggests that multiblock systems will have less tendency to form micelles particularly if randomness is allowed in the structure of the block. Also, if well defined 20 blocks are not necessary then the costs involved in synthesizing these systems will also be considerably lower. The above discussion provides the motivation for an investigation into the feasibility of multiblock copolymers as emulsifiers. As mentioned before, most di-block and tri-block copolymers contain ethylene oxide (glycol) as the hydrophilic repeat unit and in many cases propylene oxide (glycol) as the hydrophobic repeat unit. The synthesis of a multiblock copolymer can be easily achieved since it does not require the tedious and multi-step synthesis needed to produce most di- and tri-block copolymeric emulsifiers. Most surfactants emulsify oil and exhibit consistent properties over a wide range of pH, temperature and other micro-environmental variables. Ionic surfactants are however susceptible to changes in pH and their efficiency is affected with changes in ionic strength. However, in some applications it may be desirable to expect the same surfactant molecule to behave as an emulsifier under a set of conditions and not behave as one under a different set of conditions. Some common applications that may benefit from this feature include most applications requiring the use of de-emulsifiers at some point in the processing scheme. A good example of such an application is the use of drilling “muds” in oil drilling operations and the concerns involved with forming and reforming emulsions during the entire oil recovery process. A stable emulsion (usually oil dispersed in water) is usually used to lubricate the cutting bit during drilling and to carry cuttings up to the surface. An emulsion may be desirable in one part of the oil production process and undesirable at the next stage. In oil fields, an in situ emulsion that is purposely created in a reservoir as a part of an oil recovery process may change to a different, undesirable type of emulsion (water-in-oil) when produced at the well-head. This 21 emulsion may have to be broken and reformulated as a new emulsion suitable for transportation by pipeline to a refinery. At the refinery the new emulsion will have to be broken and water from the emulsion removed, otherwise the water would cause problems during the processing of the crude oil. Another related application that requires emulsification followed by de-emulsi- fication after a series of operations have been performed is the use of surfactant solutions in the formation of an oil-in-water emulsion for the pumping of heavy crude oils and bitumens. Most heavy crudes and bitumens have viscosities of the order of thousands of centipoise rendering them impossible to pump without the use of heated pipelines. An emerging technology involves the use of emulsifiers to form low Viscosity Oil-in-water (comprising 70 wt.% oil-in-water) emulsions which facilitate easy pumping at mostly ambient temperatures. Once the emulsion has reached its destination the emulsion needs to be broken and the water removed for further processing of the crude. Other possible applications include the need for an aqueous cleaner which may be used in clean up operations in most service stations (Figure 1.3). Most routine clean up of greasy parts or service station floors is accomplished by either cleaning the greasy part in organic solvents, using soap solutions, or by impingement by high velocity water jets. In all of these methods the waste liquid that is generated contains the oil phase either dispersed in an aqueous phase or suspended in a two-phase mixture. Either way the problem of separation of the emulsified oil phase by a de-emulsification process still poses a challenge. An aqueous cleaner that would allow the greased part to be efficiently cleaned followed by easy separation of the resulting oil-in-water emulsion would be of commercial interest. 22 Aqueous Recycle Oil Disposal I —> ApH e . be _. o _. . . e e AT e ' C7 6 e Emulsification Phase Separation Figure 1.3. Potential application for reversible emulsifiers as an aqueous cleaner useful in the clean up of greased components in service stations Hence, there clearly exists great potential for the development of reversible emulsifiers. In the next section, background on polymer complexation, which forms the basis for the design of novel reversible block/graft copolymeric emulsifiers, will be presented. 1.5.2. Polymer Complexation Through Hydrogen Bonding Polymer complexation occurs due to the association of two or more complementary polymers, and may arise from electrostatic forces, hydrophobic interactions, hydrogen bonding, van der Waals forces or various combinations of these interactions.”72 Once one pair of complementary repeating units associate to form a segmental complex, many other units may readily associate without the loss of translational degrees of freedom owing to the long-chain structure of the polymers. Complexation between polymeric acids and polymeric bases formed as a result of electrostatic forces may occur by proton transfer resulting in poly(cation):poly(anion) 23 pairs. However, complexes formed by hydrogen bonding between two complementary polymeric Lewis acids and bases typically dissociate under a wide variety of conditions. Hydrogen bonded complexes are formed between a Lewis acid containing an electron deficient proton and a Lewis base with lone pairs of electrons. Hydrogen bonds are Specific and directional and are more localized than any other type of weak intermolecular interaction. Hence, the equilibrium between complexed and uncomplexed segments for two hydrogen bonded species in solution may be highly sensitive to conditions such as solution pH, temperature, solvent composition and polymer composition. Such complexes may be easily broken and reformed by controlling the surrounding conditions. Hydrogen bonded complexes between poly(methacrylic acid) (PMAA) and poly(ethylene glycol) (PEG) have been studied by several investigators (see Klier et al.,73 Scranton et al.,74 and references therein). Poly(ethylene glycol) is a Lewis base by virtue of the lone pairs of electrons on the ether oxygen, whereas poly(methacrylic acid) is a Lewis acid. While both PEG and PMAA are themselves soluble in water and are highly hydrophilic in nature, the hydrogen-bonded complex is hydrophobic. If a solution of PMAA is added to an aqueous solution of PEG an insoluble phase begins to form instantly. However, the complex is formed when the acid is protonated and can be disrupted by changes in surrounding conditions such as pH.”74 This interesting feature of these reversible complexes between PMAA and PEG may be used to design novel materials and polymeric systems. In this chapter the motivation behind the development of two novel materials: polymeric pseudocrown ethers and reversible block/ graft emulsifiers, has been presented. 24 Though the two proposed materials may have potential applications in two distinct areas, their development is an outcome of the unique complexation properties exhibited by poly(ethylene glycol). The next chapter will address the specific objectives of this research and will provide a brief outline of the layout of this thesis with an overview of each chapter. 10. 11. 12. 13. 14. 25 1.6. References Powell G.M. in, Handbook of Water-Soluble Gums and Resins, R. L. Davidson Ed., McGraw Hill, New York, Chapter 18, (1980). Mance G., Pollution Threats of Heavy Metals in Aquatic Environments, Elsevier Applied Science, London, (1987). Fergusson J.E., The Heavy Elements: Chemistry, Environmental Impact and Health Eflects, Pergamon Press, New York, (1990). Kelly M., Mining and the Freshwater Environment, Elsevier Applied Science, London, (1988). Harrison RM. and Laxen D.P.H., Lead Pollution: Causes and Control, Chapman and Hall, London (1981). Anderson R.E., in Traces of Heavy Metals in Water Removal Processes and Monitoring, J. E. Sabadell Ed., United States Environmental Protection Agency, Washington DC, pp 43 (1973). Alloway B.J., Heavy Metals in Soils, John Wiley & Sons Inc., New York (1990). Tmovsky M., Oxter J.P., Rudy R.J., Hanchak M.J., and Hartsfield B., in Hazardous Waste: Detection, Control, Treatment, R. Abbou Ed., Elsevier Applied Science, New York, pp 1581 (1988). Czupyma G., Levy R.D., MacLean AL, and Gold H., In-Situ Immobilization of Heavy Metal Contaminated Soils, Noyes Data Corp, Park Ridge, NJ (1989). Lanouette K.H., Chem. Eng, 73 (1977). Holl W. and Horst J., in Recent Developments in Ion Exchange, P.A. Williams and M.J. Hudson Eds., Elsevier Applied Science, London, pp 165 (1987). Lee LA. and Davis H.J., in Natural Waters, P.C. Singer Ed., Ann Arbor Science, pp 363 (1974). Houle P.C., in Traces of Heavy Metals in Water Removal Processes and Monitoring, J. E. Sabadell Ed., United States Environmental Protection Agency, Washington DC, pp 159 (1973). Furukawa D.H., in Traces of Heavy Metals in Water Removal Processes and Monitoring, J. E. Sabadell Ed., United States Environmental Protection Agency, Washington DC, pp 179 (1973). 15. l6. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 26 Konda K., Tsuneyuki T., Hiramatsu K., and Nakashio F., Sep. Sci. Technol, 1213 (1990) Probstein R.F., in Traces of Heavy Metals in Water Removal Processes and Monitoring, J. E. Sabadell Ed., United States Environmental Protection Agency, Washington DC, pp 71 (1973). Campbell R.J., in Traces of Heavy Metals in Water Removal Processes and Monitoring, J. E. Sabadell Ed., United States Environmental Protection Agency, Washington DC, pp 97 (1973). Lin J. and Huang S.D., Sep. Sci. Tech, 24, 1377 (1990). Huang S.D., Fann C.F., and HSieh H.S., J. Colloid Interface Sci, 89, 504(1982). Dunn RD. and Scamehom J .F., Sep. Sci. and Tech, 22, 763 (1987). Kandori K. and Schechter R.S., Sep. Sci. and Tech, 25, 83 (1990). Valdes-Krieg E., King C.J., and Sephton H.H., in Traces of Heavy Metals in Water Removal Processes and Monitoring, J. E. Sabadell Ed., United States Environmental Protection Agency, Washington DC, pp 189 (1973). Millar J.R., in Recent Developments in Ion Exchange, P.A. Williams and M.J. Hudson Eds., Elsevier Applied Science, London, pp 129 (1987). Calmon C., in Traces of Heavy Metals in Water Removal Processes and Monitoring, J. E. Sabadell Ed., United States Environmental Protection Agency, Washington DC, pp 7 (1973). Sahni SK. and Reedijk J., Coord. Chem. Revs., 59, l (1984). Akelah A. and Moet A., Functionalized Polymers and Their Applications, Chapman and Hall, London (1990). Pedersen C. J., J. Am. Chem. Soc., 89, 2495 (1967). Pedersen C. J ., J. Am. Chem. Soc. 89, 7017 (1967). Liotta C. L., in Synthetic Multidentate Macrocyclic Compounds, R. M. Izatt and J. J. Christensen Eds., Academic Press, New York, pp 111 (1978). Weber W. P. and Gokel G. W., Phase Transfer Catalysis in Organic Synthesis, Springer-Verlag, Berlin (1977). Dietrich B., Lehn J. M., and Savage J. P., Tetrahedron Lett,. 2885 (1969). 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 27 Lehn J. M., Pure Appl. Chem, 49, 857 (1977). Lehn J. M., Pure Appl. Chem, 50, 871 (1978). Weber E. and Vogtle F., in Host-Guest Complex Chemistry: Macrocycles, F. Vogtle and E. Weber, Eds., Springer-Verlag, New York, pp 1 (1985). Hancock R. D. and Martell A. E., Comments Inorg. Chem, 6, 237 (1988). Smid J., Ind. Eng. Chem. Prod. Res. Dev., 19, 364 (1980). Christensen J. J., Hill J. O., and Izatt R. M., Science, 174(4008), 459 (1971). Christensen J. J., Eatough D. J., and Izatt R. M., Chem. Rev., 74, 351 (1974). Bradshaw J. S. and Stott P. E., Tetrahedron, 36, 461 (1980). Cram D. J. and Cram J. M., Science, 183(4127), 803 (1974). Cram D. J., Angew. Chem. Int. Ed. Engl., 27, 1009 (1988). Blasius E. et al., J. Chromatogr., 147 (1980). Blasius E. and Janzen K. P., Pure Appl. Chem, 54, 2115 (1982). Gramain P., in Recent Developments in Ion-Exchange, P. A. Williams and M. J. Hudson Eds., Elsevier Applied Science, London, pp 300 (1987). Cinouuini M., Colonna S., Molinari H., and Montanari F., J. C. S. Chem. Comm, 394 (1976). Smid J., Varma A. J., and Shah S. C., J. Am. Chem. Soc., 191, 5764 (1979) Gramain P. and Frere Y., Ind. Eng. Chem. Prod. Res. Dev., 20, 524 (1981). Scott RA. and Schearga H.A., J. Chem. Phy., 45(6), 2091 (1966). Pensak D.A., Pure & Appl. Chem, 61(3), 601 (1989). Weiner P.K. and Kollman P.A., J. Comput. Chem, 2 (3), 287 (1981). Allinger N.L., J. Am. Chem. Soc., 99, 8127 (1977). Brooks B.R., Bruccoleri R.E., Olafson B.D., States D.J., Swarninathan S. and Karplus M., J. Comput. Chem, 4(2), 187 (1983). Mayo S.L, Olafson B.D., Goddard W.A., J. Phys. Chem, 94, 8897 (1990). 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 28 Allinger N.L., Yuh Y.H. and Li J.H., J. Am. Chem. Soc., 111, 8551 (1989). POLYGRAF, v 3.21, Molecular Simulations Inc., 16 New England Executive park, Burlington, MA 01803. Lee K.J. and. Brown T.L, Inorg. Chem, 31, 289 (1992). Neuburger N., Bahar I. and Mattice W.L., Macromolecules, 25, 2447 (1992). Fan CF. and Hsu S.L., Macromolecules, 25, 266 (1992). Fan C.F., Olafson B.D., Blanco M. and Hsu S.L., Macromolecules, 25, 3667 (1992) Muller-Plathe F ., Rogers SC. and van Gunsteren W.F., Macromolecules, 25, 6722 (1992). Depner M. and Schurmann B.L., Polymer, 33(2), 398, (1992). Schmolka I.R., J Am. Oil Chem. Soc., 54, 110-116 (1977). Schmolka I.R., J. Am. Oil Chem. Soc., 59(7), 322-327 (1982). Schmolka I.R., J. Am. Oil Chem. Soc., 68(3), 206-209 (1991). Piirma 1., “Polymeric Surfactants”, Marcel Dekker (1992). Schofield J.D., in Critical Reports in Applied Chemistry, M.R. Porter Ed., Elsevier Applied Science, New York, pp 35 (1990). Porter M.R., Handbook of Surfactants, Second Edition, Chapman & Hall, New York (1994). Noolandi J ., Makromol. Chem. Theory Simul., 1(5), 295-8 (1992). Tsuchida E. and Abe K., Adv. Polym. Sci., 45, 1 (1982). Bekturov E. and Bimendina L.A., Adv. Polym. Sci., 45, 100 (1980). Osada Y., Adv. Polym. Sci., 82, 1 (1987). Kabanov VA. and Zezin A.B., Vysokomol Soyed., A21, 243 (1979). Klier J ., Scranton AB. and Peppas N.A., Macromolecules 23, 4944 (1990). Scranton A.B., Klier J. and Aronson C.L., in Polyelectrolyte Gels, R. S. Harland and R. K. Prud’homme Eds., ACS Symposium Series 480, Washington, DC, pp 171(1992) CHAPTER 2 OBJECTIVES OF RESEARCH It is clear that there exists a need for an investigation into the development of polymeric pseudocrown ethers and reversible emulsifiers. The ability of poly(ethylene glycol) to form complexes with cations through ion-dipole interactions and with complementary Lewis acids through hydrogen bonding, will be utilized in the development of these materials. Essentially, the objective behind the development of both materials is the use of a Simple fundamental concept in the easy and inexpensive design of a polymeric system with potential commercial applications. However, a fundamental understanding of the complexation phenomenon in both cases is essential in order to better design the resulting materials. Broadly, this research may be divided into two sections: a) an investigation into the cation binding ability of oligo(ethylene glycol) and its use in the design of a polymeric pseudocrown ether through a one-step polymerization scheme, and b) utilization of the polymer complexation ability of poly(ethylene glycol) in the design of a novel reversible emulsifier. Molecular modeling techniques will be employed to guide the synthesis in both projects and to provide a fundamental understanding of the underlying complexation phenomenon and its effect on molecular conformations. These methods will not only be utilized to visualize various complexation behavior but will also be used to validate experimental phenomenon. 29 ii. iii. iv. Vi. 30 The specific objectives of this research are as follows: an experimental and theoretical study of the cation binding ability of oligo(ethylene glycol) diacrylates; characterization of the phenomenon of templatization using pyrene end-labeled tetra- and penta-ethylene glycol; synthesis of polymeric pseudocrown ethers based upon a template ion scheme and an investigation of ion-binding properties; synthesis of poly(methacrylic acid-g-ethylene glycol) hydrogels and an investigation into the complexation phenomenon through the study of equilibrium swelling in response to solvent type and composition and polymerization conditions; synthesis of novel reversible block/graft emulsifiers and a study of their oil emulsification and surface acitive behavior; and, use of the polarity sensitive fluorescence of pyrene to study the aggregate formation behavior of the reversible emulsifiers. Presented next is a brief description of the scope of each chapter, its contents and how the specific objectives of this research will be met. Chapter 3 is the first of three chapters which describe the synthesis and characterization of polymeric pseudocrown ethers. This chapter is devoted to the understanding of the ion-solubilization phenomenon observed experimentally through cation solubility studies in organic solvents. NMR spectroscopy studies along with ion- solubilization studies will be utilized to demonstrate the cation binding properties of 31 poly(ethylene glycol) and poly(ethylene glycol)-diacrylate. Further, molecular dynamics simulations will be employed to demonstrate the effect of the presence of a template ion on the end-to-end distance of poly(ethylene glycol)-diacrylates of varying chain lengths. These studies will provide a better understanding of the fundamental basis for the synthesis scheme which relies on template ion-induced cyclization of oligo(ethylene glycol) diacrylates. Chapter 4 will further build upon the experimental and modeling results obtained in Chapter 3. The cations for use in the design of the pseudocrown ethers will be identified and the size of the corresponding oligo(ethylene glycol)-diacrylates will be determined which will most effectively template the cation, thereby bringing the unsaturated end-groups into close proximity. For this purpose the synthesis of pyrene end-labeled tetra- and penta-ethylene glycol will be discussed and pyrene excimer fluorescence will be used to identify the oligomer which is most effectively able to template around a given cation. Once again, molecular dynamics Simulations will be employed to corroborate experimental results and demonstrate the effect of chain length on the decrease in the mean end-to-end distance observed in the presence of a template cation. Finally, in Chapter 5 the synthesis of polymeric pseudocrown ethers will be discussed. Various issues which were identified in previous chapters as being important in the design of polymeric pseudocrown ethers will be summarized and a synthesis scheme will be formulated. An experimental methodology will be developed for the study of the ion-binding capacities of the resins. The monomer to solvent ratio will be varied and its effect on the ion-binding capacity will be discussed. Finally, the ion- 32 binding properties of a polymeric pseudocrown ether for the binding of nickel cations from aqueous waste streams will be presented. Chapter 6 is the first of three chapters devoted to the development of reversible block/graft polymeric emulsifiers based upon polymer complexation. This chapter will describe the effect of polymer complexation on the swelling behavior of crosslinked poly(methacrylic acid-g-ethylene glycol) hydrogels in response to solvent type and composition. This work will build upon previous studies which have demonstrated the effect of pH on the complexation phenomenon exhibited by copolymers of poly(methacrylic acid) with oligo(ethylene glycol) grafts. In addition, molecular mechanics studies will be employed to investigate the stoichiometry of the complex. Chapter 7 will primarily describe the synthesis and emulsification properties of the reversible emulsifiers. Variables such as methacrylic acid to ethylene glycol repeat unit ratio, and their effect on the oil emulsification ability of the resulting copolymers, will be investigated. Interfacial tension measurements will be obtained as a function of pH for the block/graft copolymeric emulsifiers. In Chapter 8 the environment-sensitive fluorescence of pyrene will be employed to study the aggregate formation behavior of the emulsifiers as a function of concentration and pH. These studies will also be used to identify a “critical aggregate concentration” for the emulsifiers based upon peak intensity ratios in the excitation and emission spectrum of solubilized pyrene in aqueous copolymer solutions. Finally, a chapter each will be devoted to the presentation of the conclusions and recommendations for future work. CHAPTER 3 lON-SOLUBILIZATION AND ION INDUCED TEMPLATIZATION BY OLIGO(ETHYLENE GLYCOL) AND ITS DIACRYLATES' 3.1. Introduction In this chapter, the basis for a novel method for producing inexpensive polymeric pseudocrown ethers in situ during free-radical polymerizations will be investigated, both experimentally and by molecular dynamics Simulations. The synthetic scheme is based upon a template ion and exploits the tendency of oligomeric ethylene glycol diacrylates to form intramolecular cycles during the polymerization. Oligo(ethylene glycol) diacrylates undergo free radical polymerizations by virtue of their carbon double bonds. If a growing radical chain reacts with both double bonds within a few propagation steps, a cyclic "pseudocrown ether" will be formed. The probability of this occurrence depends upon the conformation of the diacrylate oligomer as well as the concentration of double bonds in the proximity of the free radical. This probability will be fairly small for bulk polymerizations and somewhat higher for dilute solution polymerizations. However, if the polymerization is carried out in a nonpolar solvent with a small concentration of a solubilized ion, the ion will bind with the oligo(ethylene glycol) diacrylate, inducing it to assume a cyclic conformation which brings the unsaturated end groups in close proximity ’ Adapted from: Mathur A. M. and Scranton A. 3., “Synthesis and Ion-Binding Properties of Polymeric Pseudocrown Ethers: A Molecular Dynamics Study”, Sep. Sci. & Tech, 30 (7-9), 1071 (1994). 33 34 (Figure 3.1). This template effect brought on by the complexed ion Should dramatically increase the probability of primary cyclization and should, therefore, increase the pseudocrown ether yield. Compared to traditional schemes in which a polymeric chain is functionalized with a preexisting crown ether, our scheme offers safety advantages (since crown ethers are not produced until they are polymer-bound) and is less expensive. This novel synthetic scheme for polymeric pseudocrown ethers builds upon previous research on the ion-binding properties of noncyclic oligo(ethylene glycol)s and polyglycol dimethyl ethers (glymes) (see Figure 3.2). The complex stability for alkali cations with glymes exhibits an increase with increasing chain length for relatively small chains' but levels off at a threshold value which depends upon the Size of the cation. Ethylene oxide oligomers have been reported to form 1:1 and 4:1 complexes with mercuric chloride,2 and the structures of the resulting complexes have been elucidated. Various studies have Shown that poly(ethylene glycol) (PEG) and glymes of various chain lengths may be used as a cosolvent phase in heterogeneous reaction systems, due to their ability to bring into contact reagents from aqueous and organic phases, and as agents for phase-transfer catalysts in organic syntheses.” This research also builds upon previous work on polymeric pseudocrown ethers. Warshawsky et al.7 first synthesized polymeric pseudocrown ethers by reacting chloromethylated polystyrene resins with oligomeric ethylene glycols (Figure 3.3). In their method, the alcohol moieties on the ethylene glycol reacted with the chloromethyl groups by a nucleophilic substitution reaction to produce pseudocrown ethers with varying ring sizes. The resulting polymeric pseudocrown ethers were found to bind various alkali and transition metal ions. 35 o I CHK \ __ + CHz—C—C 0 CH2 C—C—C “2 + Metal n Catio PEGDA CH3 _| CHZ=CH CH,=CH Cit—if) lb=o (i=0 — O (:D + < ‘6’? CH2 0 Ozttiaoln l PCE HEMA Templated PEGDA Figure 3.1. Scheme for the synthesis of polymeric pseudocrown ethers based upon a one- step free-radical polymerization of templated polyethylene glycol diacrylate with a comonomer such as hydroxy ethyl methacrylate. OH CHz/ " CH2 Metal Cation =OH \O. + / . CH2 CH2 " 6———-CH2 0 \CH2/ \ / CH2 “—0 H2 Figure 3.2. Schematic showing the cation complexation ability of oligomeric polyethers. 36 CHZCI + OH—CHz—CHfEO—CHz—CHZEl—O—CHz—CHz—OH l O_C H2 _C H2\ HZCI I M 0—0 0—0— CH2 Figure 3.3. Synthesis scheme of Warshawsky et al. for the synthesis of polymeric pseudocrown ethers. 37 The conformations of macrocyclic ligands and their complexes with ions have been studied using molecular modeling techniques. For example, a molecular mechanics study of 18-crown-6 and its alkali complexes8 demonstrated that the lowest energy conformers possessed the Ci symmetry, as verified by X-ray crystallography. Wipf et al.8 have used molecular mechanics simulations to explain the temperature-dependent dipole moment of 18-crown-6 and to investigate the stability of its complexes with Na+ and K+. A distance geometry approach was adopted by Weiner et al.9 to study the geometry of ring systems including 18-crown-6 using both the MM2 and AMBER force fields. Molecular mechanics has also been used to estimate complexation energies and to provide insight into host-guest interactions of alkali cations with anisole spherands.l0 Computer simulations illustrated that for some ligands, steric strain involved in wrapping the ligand around the ion is comparable to the free energy of complexation.‘| Kollman and collaborators'2 have used a host of simulations methods, including distance geometry, molecular mechanics, and molecular dynamics, to investigate complexes of spherands, crown ethers and porphyrins. These studies revealed that the simulations yielded satisfactory binding energies in good agreement with experimental values. Finally, molecular mechanics and molecular dynamics studies of cation complexes of a cyclic urea-anisole spherandl3 have yielded valuable information about host-guest preorganization. In this chapter we will describe a series of molecular dynamic simulations of various oligomeric ethylene glycol diacrylates in the presence and in the absence of cationic species. The molecular modeling studies were performed to corroborate experimental investigations which demonstrated ion solubilization in organic solvents 38 using PEG. In addition, 1H NMR studies were performed to confirm the complexation of the cations with the PEG. The goal of this modeling effort is to provide an underlying molecular understanding which will allow the development of an efficient reaction scheme which maximizes the yield of the polymeric pseudocrown ether. Clearly the number of ethylene glycol repeating units in the diacrylate is important because the optimal ligand must effectively envelop the cation in order to bring the unsaturated end- groups in close proximity. Therefore, simulations were performed for ligands containing between two and ten ethylene glycol repeating units, with and without the presence of a cation. In addition, a series of simulations were performed to investigate the effects of simulation time and boundary conditions. 3.2. Experimental Methods and Techniques 3.2.1. Ion-Binding Studies of Oligomeric PEG and PEGDA A series of ion-binding and solubility studies were performed in various solvents. The objective of the studies was to identify ion/ligand/solvent systems in which an initially insoluble ion was solubilized by the PEG ligand. This would ensure that the ligand and the cation complexed with one another. Metal salts that were investigated include cuprous, nickel, lead, tin and chromium chlorides, while the solvents in which the studies were carried out include chloroform, methylene chloride, carbon tetrachloride, and dioxane. All salts were dried in an oven at 60°C for 4 days and stored in a desiccator before use. PEG with an average molecular weight of 300 daltons (Polysciences Inc., Warrington, PA) was chosen as the representative oligomer for these solubility studies. 39 In these experiments, 1 wt % of salt was added to solvent in a 20-ml glass vial. The contents of the Vial were then mixed thoroughly in a vortex mixer for about 10 min. If the salt was not soluble in the solvent, then about 10 wt % PEG 300 was added to the vial and the Vial agitated for about 15 min. The solubilization of the salts (if any) upon the addition of the PEG 300 was noted. 3. 2.2. NMR Studies 1H NMR spectroscopy was used to characterize binding by the PEG 300 by comparing PEG 300 spectra collected in the absence and in the presence of the cations. These studies were performed in deuterated chloroform and deuterated methylene chloride, and the salts under study were tin chloride and chromium chloride. The NMR spectra were taken on a Varian 300 instrument at the Max T. Rogers NMR facility at Michigan State University. All samples had a concentration of 0.1 wt % in the deuterated solvent. The salt-containing samples were prepared by dissolving 0.1 wt % saturated PEG 300 in the NMR solvent. All spectra were taken at 25°C and were obtained by signal averaging 64 scans. The CHC13 residual peak at 7.24 ppm in CDCI3 was used as a chemical shift reference. 3. 2. 3. Molecular Modeling Molecular dynamics (MD) simulations were performed on a Silicon Graphics IRIS 4D/220GTX workstation using the molecular modeling software POLYGRAF.l4 These MD simulations were performed on PEGDA containing between two and ten ethylene glycol repeating units, with and without the presence of a cation (Na+ or Al3+). 40 The molecular structures were built using the polymer builder and were energy minimized before further MD calculations. Molecular mechanics (energy minimization) calculations were based on the DRIEDING force field parameters, and the steepest descent algorithm was used to perform the minimization. The Gasteiger algorithm was used to charge equilibrate the structures. Canonical TVN MD Simulations were performed on the oligomers for a period ranging from 50 to 200 ps and the molecular trajectories were stored over the entire run. The trajectory file was written at every 0.1 ps, and the temperature for all the runs was kept at 300 K. The 200-ps simulations were performed on a single PEGDA molecule within a periodic cell with an external pressure of 1 atm at 300 K with and without the cation (sodium). Plots of total energy versus time and of end-to-end distance (ETED) versus time were generated from the Simulations. In addition, the time-averaged values of the above two parameters were calculated. 3.3. Results And Discussion 3. 3. 1. Ion-Binding Studies with Oligomeric PEG Results from the ion solubility studies are Shown in Table 3.1. In each case shown in the table, the ion was insoluble before the addition of the PEG. A table entry of "Y" indicates that the addition of the PEG solubilized the metal salts. This solubilization indicates that the ligand and cation are complexed with one another. Similar ion-binding properties were observed for the diacrylates and dimethacrylates of PEG. Since the pseudocrown ether synthesis should be carried out in a solvent in which the PEGDA is 41 soluble but the salt is insoluble unless it is complexed with the PEGDA, chloroform and methylene chloride are good choices for the synthesis solvent. Table 3.1. Solubilities of Salts Upon the Addition of PEG 300 Solvents/ CHCl3 CHzClz CCI4 Dioxane Salts CuCl Y Y Y --- NiClz Y Y m --- PbC12 --- m m --- SnClz Y Y Y Y CTCI3 Y Y partial --- 3.3.2. NMR Studies 1H NMR spectra of PEG 300 in deuterated chloroform are shown in Figure 3.4. The spectrum in Figure 3.4a was collected in the absence of any metal salts, while that in Figure 3.4b was obtained in the presence of chromium chloride. In these spectra, the singlet peak located at 3.6 ppm corresponds to the PEG backbone methylene protons (CH2) while the end group hydroxyl protons appear at 2.4 ppm in Figure 3.4a and at 2.7 ppm in Figure 3.4b. Therefore, the presence of chromium chloride bound by PEG 300 induces a shift in the location of the peak corresponding to the end-group hydroxyl protons. A similar Shift was observed in deuterated methylene chloride and when tin chloride was substituted for chromium chloride. These shifts may be attributed to complexation of the metal ion with the PEG ligand. The ion may withdraw electron density from the donor oxygen atoms, resulting in inductive deshielding of local 42 (a) ‘1" V Y r V Y Y Y I (b) (I .3 A P) a O. O O 3 Figure 3.4. 1H NMR Spectra of a) PEG 300 and b) PEG 300 in the presence of chromium chloride.(hydroxy1) protons. 43 Similar shifts in NMR peak position induced by metal salts have been reported in the literature. For example, "B and 13C NMR Spectroscopy studies on crown ethers revealed that the chemical shift of a ligand was shifted to a higher field in the presence of an ion as compared to that of the uncomplexed host.l5 Proton NMR chemical shifts induced by ionic association on a poly(ethylene oxide) chain have been observed, and the effect of counter ions on the magnitude of the downfield shift has been studied.”17 3. 3. 3. Molecular Dynamics Simulations A series of simulations were performed to elucidate the effect of the presence of a cation on the conformation of a PEGDA ligand. To determine the effect of the ligand size, simulations were performed on PEGDA containing between two and ten ethylene glycol repeating units with and without the presence of a cation (Na+ or Al3+). The first set of experiments was carried out for a duration of 50 pS. For these simulations, the total system energy as a function of time fluctuated about the mean value typically with a standard deviation less than 5%. Therefore, all of the individual conformations had nearly the same energy and are equally probable. The PEGDA end-to-end distance (distance between unsaturated end groups), which is a parameter useful in determining whether the ion causes any templatization, was also plotted as a function of time for all the simulations. This value also fluctuated with a standard deviation as large as 20% of the mean. Figure 3.5 is a plot of simulation results for the mean end-to-end distance (ETED) as a function of the number of EG repeating units in the PEGDA chain. The figure illustrates that the addition of a sodium ion results in a decrease in the mean ETED 44 18 16 4- Mean End-to-End Distance (A) 4 ~1- +PEGDA 2 " +PEGDA with Na‘ 0 I a I I I I I 2 3 4 5 6 7 8 9 10 Number of Ethylene Glycol Repeat Units "n" Figure 3.5. MD 50 ps simulation results for the mean end-to-end distance of PEGDA with different number of EG repeating units. 45 for all of the PEGDA chain lengths. Similar trends were observed when the sodium ion was replaced with an aluminum ion. This suggests that there is indeed a propensity for cyclization owing to the presence of a cation since the oligomer tends to template around the cation thereby decreasing the mean ETED. The effect of the ion is illustrated in Figure 3.6, which contains a pictorial representation of one of the conformations adopted in the presence of a sodium ion. This figure indicates that the PEGDA assumes a crown ether-like conformation around the ion with the unsaturated end groups in close proximity. The results Shown in Figure 3.5 also suggest that there exists an optimum PEGDA chain length since the PEGDA ligand containing four EG repeating units exhibits the lowest ETED and therefore the maximum templatization. This result is in agreement with studies on Size-match selectivity reported in the literature. For crown ethers, it has been reported that maximum complexation is achieved when the size of the cavity is comparable to the ionic diameter. For Na“, the crown ether tert-butyl- cyclohexyl-lS-crown-S with 5 donor oxygens has been found to provide the required Size match selectivity.‘8 For PEGDA, our simulations indicate that the end group acrylate carbonyl oxygens participate in the ion binding, making the PEGDA with n = 4 the optimal ligand to bind with sodium. To investigate the effect of Simulation time, MD runs were performed for 50, 100, and 200 pS. In general, MD simulations of greater duration allow a larger number of conformations to be sampled and therefore provide more reliable averaging. However, the inevitable cost of longer simulations is greater computational requirements (in our studies a 50-ps simulation takes a few hours, whereas the 200-ps runs require days to complete). Therefore, it is important to determine the minimum simulation time which 46 Figure 3.6. Molecular representation of a PEGDA (n = 4) conformation templated around a Na“ ion showing charged surface of the molecule. Shown here is one conformation extracted from a 50 ps run. 47 provides reliable results. The effect of simulation time on the mean ETED of PEGDA in the presence of a sodium ion is illustrated in Figure 3.7. The figure illustrates that for PEGDA ligands containing seven or fewer EG units, the simulation results are very repeatable; however, the Simulations exhibit greater variations for larger ligands. This interesting behavior suggests that the simulation time has an effect only for relatively long ligands in which the length of the oligomer becomes considerably greater than that required to encompass the ion. For shorter ligands, the ion imposes an overwhelming constraint on the conformations adopted by the ligand due to binding interactions. For the longer ligands, the ion interacts only with a portion of the chain, placing milder constraints on the end-to-end distance. Differential and cumulative distributions for the PEGDA end-to-end distance from the 200-ps simulations are plotted in Figures 3.8 and 3.9, respectively. The points in Figure 3.8 correspond to the simulation data, while the curves represent Gaussian fits to the data. Both figures clearly illustrate that the presence of the sodium ion shifts the end-to-end distance to lower values while increasing the dispersion slightly. The cumulative distributions shown in Figure 3.9 may be used to estimate the fraction of the (equally probable) conformations with an end-to-end distance below a give value. By the ergodic theorem, this will correspond to the fraction of PEGDA ligands in the system with an end-to-end distance below this threshold value. Therefore, the Simulations can provide information about the fraction of "templated" ligands which could lead to the formation of pseudocrown ethers upon reaction. 48 18 16, Mean end-to-end distance (A) L If, , I, ,, LI L, m .1— L 2 3 4 5 6 7 8 9 Number of ethylene glycol repeat units "n" Figure 3.7. Effect of different simulation times on the mean end-to-end distance for PEGDA with Na+ with various numbers of EG repeating units. 49 0.045 - 0,040: .. A PEGDA (n=4) with Na“ PEGDA(n=4) 0.035 - 0.030 - 0.025 ; 0.020 — 0.015 - Normalized Frequency 0.010 — 0.005 - 0.000 - End-tO-End Distance (A) Figure 3.8. Normalized frequency distribution of end-to-end distance for PEGDA (n = 4) with and without Na+. 50 0.9 ,_ 0.8 «F 0.7 4- .0 w + PEGDA (n=4) -o— PEGDA (n=4) with Na“ Normalized Cumulative Frequency O 01 A I L l T I I I 25 30 35 40 End-to-End Distance (A) Figure 3.9. Cumulative frequency distribution of end-to-end distance for PEGDA (n = 4) with and without a Na+ ion. 51 3.4. Conclusions A novel scheme to synthesize polymeric pseudocrown ethers in situ during a free- radical polymerization has been presented. This scheme is based upon a template ion and exploits the tendency of oligomeric ethylene glycol diacrylates to form intramolecular cycles during polymerization. Compared to traditional schemes in which a polymeric chain is functionalized with preexisting crown ethers, our scheme offers safety advantages and is less expensive. It will thus be possible to economically synthesize polymers with unique ion-binding capabilities appropriate for applications such as phase- transfer catalysis, ion chromatography, and metal ion removal from waste streams. Preliminary experimental studies demonstrated that certain salts that were not soluble in nonpolar solvents, such as chloroform and methylene chloride, were made soluble upon the addition of oligomeric PEG 300. Further evidence of cation binding by oligomeric PEG was obtained by 1H NMR studies of PEG and its complexes with metal salts. In an effort to optimize the template ion synthesis approach, molecular modeling studies were performed on oligomeric PEGDA containing between two and ten ethylene glycol repeating units, with and without the presence of cations. Simulation results indicated that the presence of the templating cation significantly decreased the mean end- to-end distance, thereby bringing the unsaturated end-groups into close proximity. Although the presence of the cation decreased the ETED for all PEGDA chain lengths, the PEGDA ligand that resulted in the most effective templatization for Na+ contained four ethylene glycol repeating units. The simulation time had little effect on the results for relatively short PEGDA ligands containing seven or fewer ethylene glycol repeating 52 units. These ligands are highly constrained due to binding interactions with the ion. A higher degree of variation with Simulation time was observed for longer ligands in which only a portion of the chain interacts with the ion. These Simulation results have affirmed the role of a templating ion in the synthetic scheme. Moreover, the results provide insight into the selection of the templating ion and the PEGDA oligomer that will maximize templatization by bringing the unsaturated end groups into proximity. 10. ll. 12. 13. 14. 15. 16. 17. 18. 53 3.5. References Chan L. L., Wong K. H., and Smid J., J. Am. Chem. Soc., 92, 1955 (1970). Iwamoto R., Bull. Chem. Soc. Jpn., 46, 1114 (1973). Lee D. J. and Chang V. S., J. Org. Chem, 43, 1532 (1978). Regen S. L., Besse J. J., and McLick J., J. Am. Chem. Soc., 101, 116 (1979). Zupancic B. and Kakalj M., Synthesis, 913 (1981). Sukata K., Bull. Chem. Soc. Jpn., 56, 280 (1983). Warshawsky A., Kalir R., Deshe A., Berkovitz H., and Patchomik A., J. Am. Chem. Soc., 101, 4249 (1979). Wipf G., Weiner P. and Kollman P., J. Am. Chem. Soc., 92, 666 (1982). Weiner P. K. et al., Tetrahedron, 39, 1113 (1983). Kollman P. A., Wipr., and Singh U. C., J. Am. Chem. Soc., 107, 2212 (1985). Hancock R. D. and Martell A. E., Comments Inorg. Chem, 6, 237 (1988). Kollman P. A., Grootenhuis P. D. J., and Lopez M. A., Pure Appl. Chem, 61, 593 (1989) Maye P. V. and Venanzi C. A., J. Comput Chem, 12, 994 (1991). POLYGRAF, v 3.21, Molecular Simulations Inc., 16 New England Executive park, Burlington, MA 01803. Live D. and Chan S. I., J. Am. Chem. Soc., 98, 3769 (1976). Ono K. and Honda H., Macromolecules, 25, 6368 (1992). Yanagida S., Takahashi K., and Okahara M., Bull. Chem. Soc. Jpn., 51, 1294 (1978). Christensen J. J., Hill J. O., and Izatt R. M., Science, 174 (4008), 459 (1971). CHAPTER 4 EFFECT OF ION-INDUCED TEMPLATIZATION ON THE END-TO- END DISTANCE OF PYRENE END-LABELED TETRA- AND PENTA-ETHYLENE GLYCOL‘ 4.1. Introduction In the previous chapter a novel schemel was presented for the synthesis of polymeric pseudocrown ethers which have potential applications in cation binding, separations, and waste management. This scheme is based upon the tendency of oligomeric ethylene glycol diacrylates to assume circular conformations in the presence of templating cations. This templatization phenomenon may induce the diacrylate unsaturated end-groups to be in proximity. If a free-radical polymerization is initiated with a comonomer such as hydroxy ethyl methacrylate, both double bonds of such a templating oligomer may be incorporated into the same kinetic chain resulting in the formation of pendent loops. The probability of cyclization is increased if the polymerization is carried out in non-polar solvents in which the cation itself is insoluble thus ensuring maximum templatization and hence primary loop formation. Furthermore, if the free-radical polymerization is performed in dilute solution, the probability of ' Adapted from: Mathur A. M. and Scranton A. 8., “Synthesis and Ion-Binding Properties of Polymeric Pseudocrown Ethers II: Ion-Induced Templatization of Oligomeric Ethylene Glycol Diacrylates”, Sep. Sci. & Tech, (in Press). 54 55 cyclization is enhanced and would result in reduced crosslinking and more cyclization as shown in Chapter 3, Figure 3.1. Our previous solubility and 'H NMR studies demonstrated that various cations were solubilized by oligomeric ethylene glycols' by virtue of ion-dipole interactions between the cation and the electron lone pairs of the ether linkages. Further, molecular dynamics simulations were performed on oligomeric ethylene glycol diacrylates both in the presence and absence of templating cations. A reduction in the mean end-to-end distance in the presence of a cation was observedl due to ion-induced constraints on the possible number and type of conformations adopted by the oligomer. These Simulations affirmed the role of the templating ion in the synthetic scheme. Experimental corroboration of the theoretical results suggesting that templatization brings the chain ends into proximity is presented in this chapter. Pyrene has been extensively used as a molecular probe due to its excellent fluorescence properties and its ability to form bi-molecular complexes (excimers) which fluoresce in a different spectral region than the individual “monomer” pyrene molecules. Cuniberti and Perico2 were among the first to study polymer cyclization using steady- state pyrene excimer fluorescence. These investigators synthesized pyrene end-labeled poly(ethylene oxide) and studied the influence of molecular weight on the extent of intramolecular excimer formation. They observed an increase in the intramolecular excimer fluorescence intensity with decreasing molecular weight. Winnik et al.3'4 studied the cyclization dynamics of pyrene end-labeled polystyrene and demonstrated that intramolecular excimer formation could be observed only for chains containing less than 1000 to 2000 bonds. For higher molecular weight chains the rate constant for cyclization 56 decreases rapidly. Winnik’s group has also studied the cyclization dynamics of pyrene end-capped polystyrenes using transient fluorescence experiments. For example, researchers from Winnik’s group reported the determination of cyclization rate constants from fluorescence decay data,5‘6 asymmetric end-labeling of polystyrene and determination of cyclization rate constants from intramolecular exciplex emission decay data6 and fluorescence quenching in pyrene/benzil end-labeled poly(tetrarnethylene oxide).7 A review of their earlier work8 on the cyclization of polymer chains in solution and luminescence techniques to study the morphology of prototype industrial materials, is a good source of literature in this area. Recently, Martinho et al.9 studied the effect of hydrostatic pressure on the cyclization of pyrene end-labeled polystyrene in various solvents and observed that the cyclization rate decreases monotonically with pressure. Frank et al. '0'” studied the complexation of poly(acrylic acid) with pyrene-labeled poly(ethylene glycol) by monitoring the excimer-to-monomer fluorescence intensity ratio. These authors used this technique to study the hydrophobic attraction between pyrene-end-labeled polyethylene glycols in water and water-methanol mixtures, and the effect of the hydrophobic interaction in the poly(methacrylic acid)/pyrene end-labeled poly(ethylene glycol) complexes. They also used the monomer/excimer ratio to demonstrate the effect of pH on complexes of poly(acrylic) or poly(methacrylic acid) with pyrene end-labeled poly(ethylene glycol).'3'” Frank’s group has also developed a new anionic synthesis schemel5 for the synthesis of pyrene end-labeled polystyrene having no ester linkages. The resulting polymers demonstrated enhanced thermal and hydrolytic stability as compared to those synthesized by previous methods. 57 The preceding two paragraphs illustrate the utility of pyrene excimer fluorescence experiments for conformation studies in cyclic or complexing polymer systems. In this study, this technique will be used to investigate ion-induced cyclization of relatively short-chain oligo(ethylene glycol) molecules. The monomer to excimer fluorescence ratio will be utilized to characterize the extent of oligomer cyclization. Comparison of the fluorescence ratio obtained in the presence of the templating ion to that observed in the absence of the ion, will provide insight into the conformation of the templated ligand. Therefore, these studies serve as an experimental test of the previously reported molecular modeling results.l In addition, we present new molecular dynamics Simulations of pyrene-labeled ligands to investigate the effect of the fluorescent chromophore on the simulation results. These simulations will provide molecular insight into the fluorescence results by elucidating the effects of the templating ion and/or the pyrene end groups on the conformations adopted by the oligo(ethylene glycol) chains. 4.2. Experimental Methods and Techniques 4. 2. 1. Synthesis of Pyrene End-Labeled Oligomeric Ethylene Glycols Pyrene end-labeled tetraethylene glycol (TEG) and pentaethylene glycol (PEG) were prepared based on the procedure reported by Cuniberti and Perico2 (Figure 4.1). One-pyrenebutyric acid (3.11 g), was mixed with tetraethylene glycol (0.899 g) or pentaethylene glycol (1.013 g) and para-toluenesulphonic acid monohydrate (0.268 g) in a 500 ml round bottom flask equipped with a Dean Stark trap as well as a condenser. These chemicals were used as received from Aldrich. To this mixture was added 250 ml 58 CHZCHZCHZCOOOH O Toluene PTSA + OH I -H20 l-pyrenebutyric acid Oligo-ethylene glycol (n = 4,5) I (CH2)3—C—O<-CH2-CH2-Oirfl) —-(CH2)3 I 06 End- to-end- distance (ETED) a0 Figure 4.1. Scheme for synthesis of pyrene end labeled oligoethylene glycols. 59 of dry toluene. The mixture was stirred and allowed to reflux overnight. The next day, the reaction was stopped and TLC (silica gel) revealed one strong spot at Rf = 0.13 (hexane: ethyl acetate 1:1). The starting material, l-pyrenebutyric acid, appears at Rf = 0.43. The product was isolated as follows. The contents of the round-bottom flask were emptied into a separatory funnel and extracted with 3 x 100 ml saturated sodium bicarbonate solution to remove the catalyst para—toluenesulphonic acid and reactants. The toluene layer was extracted once with distilled water (100 ml) and once with brine (100 ml), then stirred over anhydrous sodium sulfate for 20 minutes and filtered. Finally the solvent was evaporated on a rotary evaporator at 60°C. The dried product was placed on a high vacuum line for one hour and pumped at room temperature. The product yield was 96% and the proton and carbon-13 NMR spectra were consistent with the desired products: tetraethylene glycol and pentaethylene glycol which are pyrene-labeled on both ends (henceforth abbreviated Py-TEG-Py and Py-PEG-Py, respectively). 4. 2. 2. Fluorescence Studies The fluorescence studies of pyrene end-labeled oligomers were performed using an Aminco-Bowman Series AB2 Luminescence Spectrometer. An excitation wavelength of 322 nm was chosen based upon absorption spectrum of pyrene, and the steady-state fluorescence emission was collected for wavelengths between 330 and 630 nm. The solvent to be used for the fluorescence studies had to meet several criteria. First, it should be a solvent in which the templating ions themselves are insoluble, but are solubilized upon addition of the oligo(ethylene glycol) chains.l This ensures that all ions in solution 60 are bound by the oligomers, and therefore ensures maximum templatization. In addition, the solvent must meet the optical requirements of low absorption and emission in the Spectral window between 300 and 630 nm. With these criteria in mind, chloroform and THF were selected for the studies. Dilute solutions of the pyrene end-labeled oligomers (Py-PEG-Py and Py-TEG-Py) were prepared in HPLC grade chloroform and THF at concentrations varying from 10“ to 10'7 M. For studies performed in the presence of a templating ion, the cation (tin, chromium or nickel) was introduced into the polymer solution as its hydrated chloride salt.'6 These salts were added in excess and the resulting solutions were agitated for 2-3 days to ensure maximum templatization and equilibration. The solutions were then filtered prior to the fluorescence studies. 4. 2. 3. Molecular Dynamics Simulations Molecular dynamics (MD) simulations were performed on pyrene end-labeled tetraethylene glycol and pentaethylene glycol in the presence and in the absence of the templating ion. The pyrene end-labeled tetra- and pentaethylene glycols were modeled using molecular modeling software POLYGRAF.l7 The molecular structures were energy minimized using the DRIEDING force field and were charge equilibrated using the Gastieger algorithm prior to the MD simulations. For Simulations in the presence of the templating ion, sodium (N ai) was introduced in proximity to the ether oxygens before energy minimization and charge equilibration. TVN MD simulations were performed in the absence of any solvent18 at 300K for either 100 or 200 ps and their trajectories were recorded at every 0.1 ps. Our previous simulations' indicated that for oligomers of this relatively short chain length, there is little or no effect of simulation time (above 50 pS) 61 on the mean end-to-end distance. However, in the case of tetra- and pentaethylene glycol simulations were performed for 100 and 200 ps respectively, to allow for the bulky pyrene end-groups to sample more conformations and hence provide a better average. Plots of end-to-end distance versus time were generated and plots of energy versus time yielded almost a horizontal line thereby ensuring that all the assumed conformations were equally probable and possessed almost the same energy. 4.3. Results And Discussion 4. 3. 1. Fluorescence Studies Figure 4.2 contains fluorescence profiles of two distinct systems: a dilute solution (10’5 M) of 1-pyrenebutyric acid (curve A) and a similarly dilute solution (10'5 M) of the pyrene-end labeled oligomer Py-PEG-Py (curve B), both in chloroform. The figure illustrates that these two systems have markedly different characteristic fluorescence profiles. While both spectra exhibit Sharp peaks at approximately 382 and 402 nm (characteristic of the uncomplexed, monomeric pyrene), the Py-PEG-Py exhibits a second broad peak centered at around 480 nm (characteristic of the pyrene excimer). Therefore, Figure 4.2 illustrates the effect of placing two pyrene molecules in proximity on the two ends of an oligomeric ethylene glycol chain. At this low concentration, the l- pyrenebutyric acid exhibits essentially no excimer formation, indicating that the pyrene moieties all exist as uncomplexed “monomers.” However, at the same low concentration, the Py-PEG-Py system exhibits significant excimer formation, indicating that pyrene moieties have formed sandwich complexes with one another. A series of studies were 62 .._A Py-PEG-Py __B l-PBA Emission Intensity (arb. units) ...2 ° 330 380 430 480 530 580 630 Emission Wavelength (nm) Figure 4.2. Fluorescence emission spectra of (A) pyrene end-labeled pentaethylene glycol and (B) l-pyrenebutyric acid. The excitation was at 322 nm and both samples were at a concentration of 10'5 M. 63 performed as a function of concentration which suggested that, at these low concentrations, the ratio of the fluorescence intensity of the excimer peak to that of the monomer peak (the excimer-to-monomer ratio) is essentially constant, while at higher concentrations (10“ M and above) this ratio exhibits an upward Slope as the concentration is increased. This suggests that at low concentrations (10‘5 M and below) the excimer formation is primarily intramolecular, with the pyrene moieties on either end of the oligomeric chain coming together to form an excimer, while at higher concentrations the intermolecular excimer formation between pyrene moieties on two different oligomers becomes Significant (hence the dependence on concentration). Therefore, these studies suggest that some oligomer cyclization occurs in the pyrene end-labeled PEG even in the absence of a templating ion. This is presumably due to the fact that the oligomeric chains are rather small, leading to a locally high concentration of the two pyrene moieties attached to a Single chain. Figure 4.3 illustrates the effect of a templating cation on the fluorescence profile of Py-TEG-Py. The figure illustrates that in the presence of the templating ion (Nizi in this case) the fluorescence spectrum of Py-TEG-Py exhibits a significantly enhanced pyrene excimer-to-monomer ratio relative to spectrum obtained in the absence of any ion. Similar results were obtained using Sn2+ or Cr” as the templating ion and with chloroform as the solvent. We attribute the observed enhancement of the excimer peak upon addition of a templating cation to ion-induced cyclization of the oligo(ethylene glycol) chain. In agreement with our previous molecular modeling studiesl these fluorescence studies indicate that the templating ion binds with the oligo(ethylene glycol) ligand, and facilitates the formation of cyclic conformations which bring the two chain 64 10 9 _A Py-TEG- Py 8 _B Py-TEG-Py with Ni Emission Intensity (arb. units) LII 4 3 2 ‘ J 0 . 330 380 430 480 530 580 630 Emission Wavelength (nm) Figure 4.3. Fluorescence profiles of pyrene end-labeled tetraethylene glycol in THF solvent (10.5 M) in the (A) absence and (B) the presence of nickel chloride. The excitation frequency is 322 nm. 65 ends in proximity. Similar trends were observed for the fluorescence profiles of the longer oligomeric ligand Py-PEG-Py, as shown in Figure 4.4. Experimental results for the pyrene excimer-tO-monomer fluorescence ratio for Py-PEG-Py in two different solvents (chloroform and THF) are illustrated in Figure 4.5. The figure contains data Obtained both in the presence and in the absence of Ni” as a templating ion for both solvents, and illustrates some interesting results. For both solvents, in the absence of the templating ion, the excimer-to-monomer ratio (Iex/Im) was found to be 0.8, suggesting the probability of cyclization was essentially the same. However, in the presence of the nickel ions the excimer-to-monomer ratio was considerably higher in chloroform than in THF. This suggests that more oligomer cyclization occurs in chloroform than THF. This result, combined with the fact that the oligo(ethylene glycol) ligands more effectively solubilize the nickel ions in chloroform,‘ provides further indirect evidence for ion-induced cyclization. In addition, these studies indicate that the pyrene excimer-to-monomer fluorescence ratio provides a sensitive measure of the ability of the oligo(ethylene glycol) to solubilize the ions into an organic solvent. 4. 3. 2. Molecular Dynamic Simulations In the previous chapter,l we reported molecular dynamic Simulations of oligo(ethylene glycol) diacrylates both in the presence and in the absence of a templating cation. These studies revealed that the addition of the templating cation leads to a significant decrease in the average end-to-end distance (in the case of tetra(ethylene glycol) diacrylate (TEGDA), the shift is from ~ 19 A in the absence of the ion to ~ 7 A in 66 8 B —A Py-PEG-Py _B Py-PEG-Py with Sn Emission Intensity (arh. units) I J 0 I I 330 380 430 480 530 580 630 Emission Wavelength (nm) Figure 4.4. Fluorescence emission spectra of pyrene end-labeled pentaethylene glycol (10'5 M) in chloroform in the (A) absence and (B) presence of tin chloride. Excitation is at 322 nm. 67 '0 Cl Py-PEG-Py . Py-PEG-Py with Ni Excimer to monomer peak ratio .9 .O O :- .- 45 O5 00 -- N A .O N Chloroform Solvent Figure 4.5. Comparison of excimer-to—monomer ratio for pyrene end-labeled penta- ethylene glycol (10'5 M) in chloroform and THF in the presence and absence of nickel chloride. 68 the presence Of the ion). In the present study, molecular dynamic simulations were performed on pyrene end-labeled ligands to elucidate the effect of the large fluorescent chromophores on the chain conformations (Figure 4.6 and 4.7). These new simulations are compared to our previous simulation results in the following discussion. Figure 4.8 contains simulation results for the normalized frequency distribution of the end-to-end distance for two distinct systems: pyrene end-labeled tetra(ethylene glycol), and the analogous unlabeled diacrylate (TEGDA). These results indicate that replacing the acrylate group with the pyrene chromophore results in a marked decrease in the mean end-to-end distance (ETED) while leaving the dispersion relatively unchanged (the ETED is ~19 A for the TEGDA, and ~8.5 A for Py-TEG-Py). These trends may be attributed to the associative interactions between two pyrene moieties (these interactions are responsible for the formation of the sandwich complex). These interactions tend to bring the chain ends in proximity, resulting in the decreased mean end-to-end distance (Figure 4.7). These simulation results suggest that even in the absence of a templating cation, the pyrene end-labeled oligomers exhibit an increased tendency to form cycles compared to the TEGDA. This result is consistent with the significant excimer formation even in the absence of the templating ion exhibited in Figures 4.2-4.4. Figure 4.9 demonstrates the further reduction in the mean end-tO-end distance upon addition of the templating cation to Py-TEG-Py. The figure contains simulation results for Py-TEG-Py in the absence and in the presence of a templating ion. These molecular dynamics simulations demonstrate that the mean end-to-end distance decreases to ~56 A from ~8.5 A in the presence of the cation. Again the simulation results are consistent with the ion-induced enhancement in the excimer fluorescence shown in Figures 4.3 and 4.4. 69 .‘K‘cf‘ua -Z . ' Figure 4.6. Starting conformation of end-labeled pentaethylene glycol in the presence of a sodium cation. 70 Figure 4.7. Pyrene end-labeled pentaethylene glycol conformation with templating cation showing pyrene excimer and the end-tO-end distance (ETED). 71 0.06 - 0.05 - 0.04 - 0.03 -l 0.02 - Normalized Frequency 0.01 — 0.00 ~ 25 04 or A o _l or N o End-to—End Distance (A) Figure 4.8. Comparison of the normalized frequency distribution of end-to-end distance for TEGDA and Py-TEG-Py demonstrating the effect of pyrene end- labeling. 72 0.08 - o D Py'TEGPy o o 0 Py-TEG-Py Wllh Na+ g“ 0.06 - a E '8 0.04 - £1 '3' E 0.02 - Z 0.03 - I 1 I I I fit I 0 5 10 15 20 FJld-tO-FJld Distance (A) Figure 4.9. Normalized frequency distribution of end-to-end distance for pyrene end- labeled pentaethylene glycol in the presence and absence of a templating sodium cation. 73 Figure 4.10 demonstrates the effect of the pyrene end-labeling on the possible conformations of the templated tetra(ethylene glycol). The figure contains simulation results for both Py-TEG-Py and TEGDA in the presence of the templating ion. Figure 4.10 once again illustrates that the templating ion greatly restricts the conformations adopted by the chain as illustrated by the reduced mean end-to-end distance compared to the ion-free simulation results shown in Figure 4.8. It is interesting that the Simulation results suggest that, while the presence of the pyrene end groups has only a slight effect on the mean value of the end-to-end distance, it has a marked impact on the dispersion. This suggests that the combined presence of the templating ion and the pyrene end groups places considerable restrictions on the conformations adopted by the oligomer. 4.4. Conclusions In this chapter we have demonstrated the use of pyrene excimer fluorescence experiments to investigate the ion-induced cyclization of relatively short-chain, pyrene end-labeled oligo(ethylene glycol) molecules. The excimer fluorescence results demonstrated that even in the absence of any templating cation, a significant excimer emission was observed, indicating that the locally high concentration of the two pyrene moieties attached to the same oligomer leads to intramolecular cyclization. The addition of templating cations such as Ni2+ and Sn2+ leads to further enhancements in the excimer fluorescence. These results suggest that the templating ion binds with the oligo(ethylene glycol) chain and facilitates the formation of cyclic conformations which bring the two chain ends in proximity. For pyrene end-labeled penta(ethylene glycol), the excimer 74 0.08 D O .. o Py-TEGPy with Na+ a TEGDA with Na+ 0.06 - >: 2 3 8' a 0.04 - E ,g E g 0.02 - z 0.00 - l ' l ' l ' l ' l ' l 0 5 10 15 20 25 End-to-End Distance (A) Figure 4.10. Comparison of normalized frequency distribution of end-to-end distance Obtained from molecular dynamics simulations for Py-TEG-Py and TEGDA, both with a templating Na+ cation. 75 fluorescence enhancement induced by nickel ions was more pronounced in chloroform than in tetrahydrofuran. Therefore the pyrene excimer-to-monomer fluorescence ratio was found to provide a sensitive measure of the ability of the oligo(ethylene glycol) mixture to solubilize the ions into an organic solvent. Molecular dynamics Simulations of pyrene-labeled oligo(ethylene glycol) ligands were performed to provide molecular insight into the fluorescence results by elucidating the effects of the templating ion and/or the pyrene end groups on the conformations adopted by the oligo(ethylene glycol) chains. Simulation results indicated that even in the absence of a templating cation, the pyrene end-labeled oligomers exhibit an enhanced tendency to form cycles compared to the analogous unlabeled ligands. This result is consistent with the fluorescence results in which significant excimer formation was observed even in the absence of any cation. Simulation results further indicated that the addition of a templating cation leads to a further reduction in the mean end-to-end distance for pyrene end labeled tetra(ethylene glycol) in a manner consistent with the observed enhanced excimer fluorescence. Finally, for two systems containing the templating ion, the presence of the pyrene end groups on the ethylene glycol oligomer has only a slight effect on the mean end-to-end distance, but leads to a marked reduction in the dispersion of the distribution. Hence, the combined presence of the templating ion and the pyrene end groups appears to place considerable restrictions on the conformations adopted by the oligomer. The studies reported in this contribution provide considerable insight into the ion- induced cyclization of oligomeric ethylene glycols, and therefore enhance our .understanding of the synthesis of polymeric pseudocrown ethers by the template ion 76 method. These studies have contributed to the selection of a synthesis solvent and the template ion-PEGDA oligomer length that may be used for the synthesis of polymeric pseudocrown ethers. In the next chapter these results will be used to develop a synthetic scheme for the synthesis of polymeric pseudocrown ethers. 10. ll. 12. 13. 14. 15. 16. 77 4.5. References Mathur A. M. and Scranton A. B., Sep. Sci. & T ech., 30 (7-9), 1071 (1994). Cuniberti C. and Perico A., Eur. Polym. J., 13, 369 (1977). Winnik M. A., Redpath A. E. C. and Richards D. H., Macromolecules, 13, 535 (1980) Redpath A. E. C. and Winnik M. A., Ann. N. Y. Acad. Sci., 75, 366 (1981). Winnik M. A., Redpath A. E. C., Paton K. and Danhelka J., Polymer, 25, 91 (1984) Winnik M. A. and Sinclair A. M., Can. J. Chem, 63, 1300 (1985). Slomkowski S. and Winnik M. A., Macromolecules, 19(2), 500 (1986). Winnik M. A., in Photophysical and Photochemical Tools in Polymer Science: Conformation, Dynamics and Morphology, M. A. Winnik Ed., NATO ASI Series C: Mathematical and Physical Sciences 182, pp 15, 293 and 611 (1986). Martinho J. M. G., Catanheira E. M. S., Reis e Sousa A. T., Saghbini 8., Andre J. C. and Winnik M. A., Macromolecules, 28, 1167 (1995). Oyama H. T., Tang W. T. and Frank C. W., Macromolecules, 20, 474 (1987). Char K., Frank C. W., Gast A. P. and Tang W. T., Macromolecules, 20, 1833 (1987) Oyama H. T., Tang W. T. and Frank C. W., Macromolecules, 20, 1839 (1987). Oyama H. T., Hemker D. J. and Frank C. W., Macromolecules, 22, 1255 (1989). Hemker D. J ., Garza V. and Frank C. W., Macromolecules, 23, 4411 (1990). Tang W. T., Hadziioannou G., Smith B. A. and Frank C. W., Polymer, 29, 1718 (1988) Several authors have studied the extraction of alkali metal and alkaline earth metal cations from aqueous media by macrocyclic and polymer bound macrocyclic compounds. For overviews see (i) Sahni S. K. and Reedijk J ., Coordination Chemistry Reviews, 59, 1 (1984) (ii) Christensen J. J ., Eatough D. J. and Izatt R. M., Chem. Rev., 74 (3), 351(1974). Specifically, the binding of hydrated salts of nickel and chromium by macrocyclic ligands has been reported by (iii) Prabhakar 17. 18. 78 L. D. and Umarani C., J.M.S.-Pure Appl. Chem, A 32(1), 129 (1995), and (iv) Blasius E. and Janzen K. P., Pure & Appl. Chem, 54(11), 2115 (1982). POLYGRAF, v 3.21, Molecular Simulations Inc., 16 New England Executive Park, Burlington, MA 01803. The primary reason for not including any solvent in our simulations was the fact that the solvent-cation interactions would be minimal since the cation is itself insoluble in the solvents used (chloroform and THF). Also, since in our laboratory we have solubilized salts in bulk oligoethylene glycols, the modeling may be interpreted to reflect those observations. CHAPTER 5 SYNTHESIS AND ION-BINDING PROPERTIES OF POLYMERIC PSEUDOCROWN ETHERS 5.1. Introduction In the previous chapters, the cation binding properties of oligomeric ethylene glycol diacrylates has been studied, both experimentally and by molecular modeling. It has been demonstrated1 that oligomeric ethylene glycol diacrylates solubilize metal salts, such as nickel chloride and chromium chloride, in organic solvents such as chloroform. Ion-solubilization studies along with NMR Spectroscopy provided evidence for the cation binding properties of these oligomeric polyethers. Further, molecular modeling studies have corroborated these experimental findings by demonstrating that there is a decrease in oligomeric ethylene glycol diacrylate mean end-to-end distance in the presence of a templating cation. In order to investigate the template effect further, pyrene end-labeled tetra— and penta-ethylene glycol were synthesized and cation induced templatization was observed2 by studying excimer formation in the presence and absence of cations. Molecular modeling studies demonstrated that there is a decrease in the mean end-to-end distance in the presence of cations such as Na” and Al”. Studies with tetra- and penta- ethylene glycol diacrylate suggested that the templating efficiency is maximum with cations such as nickel and chromium as evidenced by a large increase in pyrene excimer fluorescence. These studies form the basis of a template ion scheme for the synthesis of 79 80 polymeric pseudocrown ethers. In this method, the templatization ability of short chain oligomeric ethylene glycol diacrylates in the presence of a metal cation, is exploited to synthesize polymeric pseudocrown ethers via a single step polymerization. As described earlier,l if a free-radical polymerization is performed with a hydrophilic comonomer such as hydroxy ethyl methacrylate, the probability of cyclization is greatly enhanced due to ion-induced templatization. Moreover, if the polymerization is carried out in an organic solvent in which the template effect is maximized (a solvent in which the metal salt is itself insoluble) and under dilute conditions, the probability of cyclization is maximized. Based upon these studies a model cation-oligo(ethylene glycol)-diacrylate system can now be studied for investigating the cation binding properties of polymeric pseudocrown ethers synthesized by this scheme. In this chapter, the optimization of this synthesis scheme will be discussed. Tetra- ethylene glycol diacrylate with nickel as the templating cation will be the system under investigation, and polymeric pseudocrown ethers (PCES) synthesized from this model system will be investigated for their cation binding properties. Moreover, issues such as PCE regeneration and effect of monomer to solvent ratio on the binding capacity, will be addressed. The ion-binding properties of “control” hydrogels synthesized in the absence of a templating cation have been compared to the binding capacity of the PCES. A schematic for the PCE synthesis by the free-radical polymerization of hydroxy ethyl methacrylate with templated oligo(ethylene glycol)-diacrylate is shown in Figure 5.1. Synthesis of the PCE hydrogels was achieved by performing a photopolymerization with chloroform as the solvent. The choice of solvent was based upon previous studies which demonstrated that salts such as nickel chloride were solubilized by oligo(ethylene 81 a; 3‘,“ é‘w'im‘i ,-,., ,‘F‘ 4". 3 Solvent . “N , “- ”-t V:- r..' ' ., 1.8,,- ' ‘ i~ "57.". ' that . ”a I 't' t 5‘29‘”:QED55‘£E&3$' “v war-“a ”' 'a 0' 4? 0:2; .1. . -., . " ___> 35.5», :1; $.‘1‘2fir‘fifi a ‘ .a-p . 53% .v- 9', - ogfi’g’fii nght CH3 \ CH2=C es 1;“ 0 CH2 8:4“, I C H—CH—o ’ €le 0” fC 2 2 2ft \ PEGDA / k HEMA / Figure 5.1. Schematic of free-radical polymerization scheme for the synthesis of polymeric pseudocrown ethers. 82 glycol)-diacrylate. Further tetra- and penta-ethylene glycol were most efficient in templating nickel and chromium as determined from the decrease in end-to-end distance observed in both experimental and modeling results. For the purposes of this model system under study, poly(ethylene glycol)-diacrylate (PEGDA 200) was chosen as the templating oligomer and nickel chloride or nitrate as the salt. (The PEGDA 200 contains on an average 4.5 ethylene glycol (EG) repeat units since the molecular weight of one EG repeat unit is 44 and the number 200 represents the average molecular weight of the EG segment in the PEGDA oligomer). The comonomer chosen was hydroxyethyl methacrylate (HEMA) due to its hydrophilic nature. Both the HEMA and the PEGDA undergo free-radical polymerizations. The free-radical polymerization was initiated by light instead of thermal means because of the low boiling point of chloroform and the unavailability of efficient thermal initiators with a low half life at room temperature. For purposes of comparison and to demonstrate the enhancement in cation-binding, crosslinked homopolymers of HEMA containing the same amount of PEGDA (in the absence of any salt) were synthesized. Their ion-binding properties have been compared to those exhibited by the PCES. Further, various regeneration schemes have been studied for the removal of the template cation after synthesis and for recovery of bound cations during ion-binding studies. 83 5.2. Experimental Methods and Techniques 5. 2. I . Synthesis of Polymeric Pseudocrown Ethers Polymeric pseudocrown ethers were synthesized based upon a template ion scheme as described earlier in Chapter 3 (Figure 3.1). All chemicals and materials were used as received. All organic solvents were HPLC grade and water used for making salt solutions was also HPLC grade. First, salt containing the template cation was added to PEGDA 200 (Polysciences, Warrington, PA) with constant stirring until saturation was reached and no more salt could be added to the PEGDA without precipitation. A saturated PEGDA-nickel solution was formed upon adding upto 0.7 g of Ni(NO3)2.6HZO to 5 g of PEGDA 200. This mixture was calculated to contain 0.4814 moles of salt per gram of PEGDA 200 or 0.09715 moles of Ni2+ per gram of oligomer. Chloroform was the solvent in which the synthesis was performed and the ratio of monomers to solvent was varied from 20/80 to 50/50 (by weight). In the case of the PCE hydrogels the PEGDA ZOO-nickel solution was used while for the poly(HEMA) hydrogels the pure PEGDA 200 was used in the synthesis. The PEGDA 200 was selected to be 5 wt.% of the monomers. This was based upon several studies in which the amount of PEGDA 200 was varied from 1% of monomers (by weight) to 10% of monomers which showed that 5% of monomers was the optimum amount. Values lower than 5% resulted in minimal ion-binding while an amount greater than 5% may result in higher crosslinking density and hence a decrease in the ion-binding capacity. The following procedure was optimized for the synthesis of the PCE hydrogels and the “control” 84 crosslinked poly(HEMA) hydrogels. In a 4 02. square glass bottle, 0.2148 g of PEGDA 200 was added to 3.8472 g of HEMA. Next, 0.0419 g of Irgacure 651 (Ciba-Giegy) was added to the monomer mixture as the free-radical photoinitiator. The photoinitiator was present at 1 wt.% of the monomers. For a dilute synthesis (20/80 monomer/solvent ratio), 16.0388 g of chloroform was added to the reaction mixture. The solution was mixed well for about 5 minutes in the sealed glass bottle using a vortex mixer. A Black-Ray UV lamp (200 W) was used to initiate the photopolymerization by exposing one face of the horizontally placed glass bottle to the lamp. The reaction mixture was exposed to light for 30 minutes in which time the clear reaction mixture turns into an opaque whitish solid mass. In order to remove unreacted monomer, initiator molecules, and residual solvent trapped within the gel, 40 ml of methanol was added to the polymer in the glass bottle. This would facilitate the swelling of the gel and the removal of unreacted monomer, initiator, and chloroform from the gel into the bulk solvent phase. The methanol solvent was replaced with fresh methanol every day for 3 days. To facilitate the removal of methanol from within the gel, the gel was swollen in deionized water next. The water was replaced everyday for 4 days in order to extract all the methanol and any remaining monomer. The gel was then removed from the bottle and placed in an oven at 80°C for drying. The drying process ensured complete removal of any trapped solvent in the gel. After drying the gel for about 4-5 days the polymer was then ground in a mortar and pestle, and for further particle size reduction, in a coffee grinder. The polymer was now weighed and the yield recorded. This polymer was used for further regeneration and ion- binding studies. 85 5.2.2. Removal of Template Cation And Conditioning Of Resin Before using the polymer for ion-binding studies, complete removal of the template cation was essential. The ground polymer particles were soaked in 0.1M HCI solution for 2 days after which the solution was replaced with fresh acid solution. The polymer was then filtered over a vacuum filter and stored in deionized water for further 1186. 5. 2. 3. Ion-Binding Studies An experimental setup that was used for ion-binding studies is Shown in Figure 5.2. A 25 cm tall jacketed glass column with an internal diameter of about 10 mm was used for the ion-binding studies. The ion-binding experiments were carried out at 25°C by circulating water from a controlled temperature water bath. The polymer particles swollen in water were filtered and loaded into the column. The column was gently compacted so as to decrease the void volume and channeling. To condition the column before ion-binding studies were performed, 50 ml of 1M HCI was pumped through the column at a rate of 2 ml/min followed by 50 ml of deionized water. The metering pump was capable of delivering flow rates from 1 ml/min to 9 ml/min. For ion-binding studies, atomic absorption spectroscopy was employed for the determination of ion concentrations. A Perkin Elmer AA-20 atomic absorption spectrometer was used for this purpose. For the determination of concentrations of various aqueous nickel solutions, the 352.7 nm wavelength from a nickel lamp (Varian Scientific) with a slit width of 0.5 was used. A 100 ppm (based upon Ni”) stock solution was prepared in HPLC grade water using nickelous nitrate hexahydrate as the stock solution. Various dilutions of this Jacketed Column Temperature controlled water bath 86 Atomic Absorption Spectrophotometer Metering pump Figure 5.2. Experimental setup for ion-binding and regeneration studies. 87 solution were used as standards for generating calibration curves for the AA spectrometer. Ion-binding studies were performed by pumping 100 ml of the 100 ppm Ni” stock solution through the column at a rate of 2 ml/min. The concentration of the solution after passing through the column was determined for each 5 ml of solution collected and a breakthrough curve was generated for each run. A similar experiment was performed in which case the concentration of the total volume of solution collected at the column outlet was determined. The difference between the inlet concentration and outlet concentration was used to calculate the amount of Ni” bound by the column. This value was represented as mg of Ni” bound per gram of polymer. However, the ion- binding capacity was calculated from the concentration of the eluted solution as described below. 5. 2. 4. Column Regeneration and Determination of Concentration of Eluted Solution Column regeneration was achieved by eluting HCl solutions with concentrations varying from 0.1M to IM at temperatures ranging from 40°C to 60°C. It was observed that complete elution of the bound cations is achieved by a minimal volume of an HCI solution at a concentration of 1M and at 60°C. The water bath temperature was set to 60°C and once the temperature equilibrated in the bath, a graduated cylinder with 20 ml of 1M HCl was placed in the bath to raise the temperature of the eluting acid solution to 60°C. The elution was then performed by pumping the acid solution through the column at 2 ml/min. The eluent was collected at the column exit and its concentration was 88 determined by AA spectroscopy. From this calculation an accurate determination of the ion-binding capacity of the polymeric resin could be made. 5.3. Results and Discussion 5.3.1. Polymeric Pseudocrown Ether Synthesis The synthesis scheme described earlier and pictorially represented in Figure 5.1 was optimized through a series of syntheses. First, based upon previous studies,2 it was demonstrated that Ni” and Cr3+ can be optimally bound by tetra- and penta—ethylene glycol thereby bringing the end-groups in close proximity. Since the synthesis scheme relies on efficient templatization, it was decided to select nickel and chromium as the cations for ion-induced templatization, and oligomeric poly(ethylene glycol)-diacrylate with 4-5 ethylene glycol repeat units as the templating oligomers. PEGDA 200, with an average ethylene glycol molecular weight of 200 daltons representing an average 4.5 number of ethylene glycol repeat units, was readily available from Polysciences, Warrington, PA. Further, in order to develop a methodology for the study of the synthesis and ion-binding properties of polymeric pseudocrown ethers by the ion-induced templatization scheme, nickel was chosen as the candidate cation. Nickel is one of the many heavy metal ions found in many waste streams in inorganic form. Several methods for the binding of nickel from aqueous waste streams have been employed. Precipitation through chemicals such as soda ash,3 use of chelating agents such as EDTA,4 and separation by ion-exchange resins5 are examples of some techniques used conventionally. More recently, liquid surfactant and emulsion liquid 89 membranes employing poly[poly(ethylene glycol) phosphate] as a macromolecular, multisite carrier of nickel cations has been investigated.6 This polymer, dissolved in 1,2- dichloroethane, has been shown to be an efficient carrier of Ni” ions in bulk liquid membranes with a pH gradient as the driving force. From the above discussion it is apparent that nickel is an important heavy metal cation for which an efficient separation technology would be desirable. The synthesis of the pseudocrown ethers was achieved after several trial runs to determine conditions that would maximize PCE yield and ion-binding capacity. One of the variables investigated was the amount of PEGDA 200 to be incorporated into the PCE. As discussed before, an increase in the templated PEGDA concentration may increase the final ion-binding capacity but will also result in an increase in the fraction of PEGDA lost towards crosslinking. Our previous modeling studies2 have indicated that the probability of cyclization even in the absence of a templating cation is increased as the length of the oligomeric ethylene glycol chain decreases. This may be attributed to the short chain length of the oligomer and the limited number of equilibrium conformations that may be adopted resulting in a low entropy. Based upon the optimization studies a PEGDA concentration of 5 wt.% of the monomers was chosen for this study. 5. 3. 2. Ion-binding Studies The results of the ion-binding properties exhibited by the PCE synthesized with nickel as a templating cation are shown in Figure 5.3. The data presented is an average of two experimental runs for each polymer. This figure illustrates the ion-binding capacity 90 0.35 a PCE CI P(HEMA) 0.3 _- l Ii .lil‘ iim‘ hi i I . C 1542?; T T ‘ ”T a, E _—L_ g 0.25 ._.. ;:—=:_—::: E% L L :5” : _‘~—:—_——._—— 5 0 2 :=—_—: 2 - * -‘ * ELL ' E E or E: :— V ————=:-—'L_—__LL"" :————*‘—“—_.__ b 0.15 :%—f '—'——— —_—‘ — — ID :~__—_—.=—:T: EE-z _____ g, ; E U) 0'1 Li 7? ‘_"_" T ........... ’ .5 ~—- _ ' F __ _ __ 'o E: —:___L_ 5 TE ‘ m 0'05 T :53? — -- _:_~ — — L i .m—F ...._..__.__._ ...___.__— _____.—. l l ll lllllllllll .,. l 1‘ l r l 20/80 35/65 50/50 Monomer/solvent ratio during synthesis Figure 5.3. Experimental ion-binding capacities for polymeric pseudocrown ethers (PCES) and “control” p(HEMA) hydrogels as a function of the monomer to solvent ratio during synthesis. 91 of the PCB as a function of the monomer to solvent ratio during synthesis. Also shown in the figure is a comparison between the ion-binding ability of the PCB and the p(HEMA) hydrogel synthesized with the same amount of PEGDA 200 but in the absence of any template cation. There are two primary observations to be made from this data. First, as the ratio of monomer to solvent increases the average ion-binding capacity decreases. A PCE gel synthesized in a 20/80 monomer to solvent ratio exhibited an ion-binding capacity of about 0.30 mg/g polymer while that synthesized in monomer to solvent ratio of 50/50 had an average ion-binding capacity of 0.17 mg/g polymer. This is consistent with our hypothesis that the probability of cyclization would be enhanced under dilute solution polymerization conditions since interactions between two templated oligomers would be minimal thereby preserving the cyclic conformation. The data also suggests that a plateau may be reached on the maximum attainable ion-binding capacity as the monomer to solvent ratio is decreased. Also, experiments demonstrated that if the monomer to solvent concentration is decreased to 10/90, the mechanical integrity of the polymeric material may also be compromised since a highly swellable hydrogel is produced. Since strength is of importance when considering loading and repeated ion-binding operations both in the lab and potentially in industry, a lower ratio than 20/80 was not investigated for ion-binding properties. Since the synthesis of the first crown ether almost three decades ago, most cation binding studies reported in the literature have dealt with the alkali or alkaline earth metals. Fewer studies have focused on transition metal binding from aqueous streams. However it is mostly transition metals that present a challenge in the separations industry 92 due to their toxicity. While conventional ion-exchange resins have ion-binding capacities of upto a few hundred milligrams per gram of resin, most polymeric crown ethers and other polymer bound cyclic ligands have binding capacities at least an order of magnitude lower. For example, Kahana et 01.7 have reported the ion-binding capacities of various polymeric pseudocrown ethers ranging from as low as 0.035 mg/g to 0.245 mg/ g polymer for Li” and from 1.32 mg/g to 13.2 mg/g for Csi. By comparison, in a study of the ion- binding capacities of several chelating ion-exchange polymers for Ni”, capacities ranging from 0.23 mg/g for picolinic type resins8 to 31.7 mg/g for ethyleneimine type resins,9 have been reported. The PCEs synthesized and investigated in this study exhibit ion- binding capacities comparable to the polymeric pseudocrown ethers mentioned above. The second important result of these studies was the observed enhancement in the ion-binding properties demonstrated by the PCB as compared to the p(HEMA) hydrogel. The PCE gel synthesized at ratio of 20/80 exhibited an increase in ion-binding capacity by a factor of 7.5 while that synthesized under a monomer to solvent ratio of 50/50 possessed about double the ion-binding capacity of its p(HEMA) counterpart. This enhancement may be attributed to the ion-induced templatization of PEGDA in the case of the PCB, which results in an increase in cyclization as compared to crosslinking. Another trend that is observed 'is the increase in ion-binding capacity of the p(HEMA) polymers as the monomer to solvent ratio is increased. This may be attributed to the increased cyclization of the PEGDA as the monomer to solvent ratio is increased possibly resulting from chain interactions and dynamics. Templatization increases the fraction of PEGDA 200 oligomer chains that have a low mean end-to-end distance thereby bringing the unsaturated acrylate end-groups in 93 close proximity. This may facilitate the incorporation of the PEGDA into a growing polymer chain, resulting in the formation of pendent loops in a crown ether type conformation. This method not only results in the incorporation of pendent cyclic oligo- ethers onto a hydrophilic polymeric support by a one-step inexpensive free-radical polymerization, but also ensures the formation of a crown ether like structure only during synthesis when such a structure is immobilized onto the polymeric backbone. This overcomes the health hazards associated with crown ethers and crown ether derivatives (since they ma readily absorb through the skin). This, along with a high ion-binding capacity, makes PCEs suitable for use as an easy to synthesize and inexpensive ion- binding resin. 5. 3. 3. Polymer Regeneration 7,10 It was observed by Kahana and Warshawsky that the ion-complexation properties of polymeric crown ethers are temperature dependent. They observed that, for ions bound by crown groups by a salt-coordination mechanism, a temperature increase from 20°C to 60°C caused enhanced spontaneous elution of the bound salt. These investigators have also studied the binding of protic and Lewis acids to polymeric pseudocrown ethers.ll They observed that complexation of the pseudocrown ether with acids such as HCl in water starts at HCl concentrations exceeding 1M. Further, other researchers have also demonstrated that a combination of acidic conditions and high temperature'z"3 is most efficient in eluting cations bound by polymeric crown ethers. Our regeneration scheme was developed based upon all these considerations. The first 94 occasion when regeneration was employed was after synthesis, for the removal of the template cation. Measurement of the concentration of the eluted solution revealed that upto 95% of regeneration of the PCB could be achieved. Further, during ion-binding studies, regeneration was performed in order to reuse the column and to determine the ion uptake. Values obtained for the ion-binding capacity based upon the difference between the initial and final concentrations of the solution passed through the bed were in good agreement with those calculated from the eluent concentration after regeneration. In most cases, upto 100% of the bound nickel was eluted from the column upon regeneration. This thermo-regulated decomplexation ability of the PCB makes it attractive for repeated loading and re-use. 5.4. Conclusions In this chapter the synthesis and ion-binding properties of polymeric pseudocrown ethers has been presented. Specifically, the synthesis of a pseudocrown ether hydrogel with hydroxy ethyl methacrylate as the comonomer and nickel as the template cation has been discussed. Based upon earlier studies, poly(ethylene glycol)-diacrylate 200, with an average of 4.5 ethylene glycol repeat units, was chosen as the templating oligomer. Hydrogels with hydroxy ethyl methacrylate as the comonomer and PEGDA 200 as the crosslinking agent were also synthesized, but in the absence of any templating cation. The conditioning, ion-binding properties and regeneration of the PCEs was discussed. The monomer to solvent ratio during synthesis were varied from 20/80 to 50/50 and its effect on the ion-binding capacity was investigated. 95 The PCE synthesized with a monomer to solvent ratio of 20/80 was observed to possess the highest ion-binding capacity (0.3 mg Nip/g polymer) which is comparable to capacities of polymer bound crown ethers reported in the literature. The binding capacity of the PCEs was observed to decrease with an increase in the monomer to solvent ratio during synthesis. This was attributed to a possible decrease in templatization efficiency and inter-oligomer interactions resulting in greater crosslinking at the expense of pendent crown ethers. Also, the effect of the template ion on the ion-binding capacity of the PCE was demonstrated by comparing the binding capacity of the PCE with that exhibited by a p(HEMA) hydrogel synthesized similarly but without any templating cation. The PCE exhibited an enhanced ion-binding capacity for nickel. Finally, an efficient regeneration scheme was optimized which resulted in complete regeneration of the PCB. These experiments have demonstrated the promise of the template ion induced cyclization scheme for the synthesis of polymeric pseudocrown ethers. With an easy one- step synthesis scheme and comparable binding properties for Ni”, this research could result in the development of an inexpensive resin for the binding of nickel from aqueous waste streams. Further, this scheme could be extended to synthesize PCEs for the binding of other toxic heavy metal ions of interest such as chromium, copper, lead, zinc among others. 10. 11. 12. 13. 96 5.5. References Mathur A. M. and Scranton A. B., Sep. Sci. & T ech., 30 (7-9), 1071 (1994). Mathur A. M. and Scranton A. B., “Synthesis and Ion-Binding Properties of Polymeric Pseudocrown Ethers II: Ion-Induced Templatization of Oligomeric Ethylene Glycol Diacrylates,” Sep. Sci. & Tech., (in press). Lanouette K.H., Chem. Eng., Oct. 17 (1977). Lee LA. and Davis H.J., in Trace Metals and Metal-Organic Interactions in Natural Waters, P. C. Singer Ed., Ann Arbor Science, pp 363 (1974). Hudson M., in Ion Exchange: Science and Technology, A.E. Rodrigues Ed., NATO ASI Series E, 107, pp 463 (1986). Wodzki R., Wyszynska A. and Narebska A., Sep. Sci. & Tech., 25(11 &12), 1175- 1187(1990) Kahana N. and Warshawsky A., J. Polym. Sci. .' Polym. Chem. Ed., 23, 231-253 (1985) Warshawsky A., in [an Exchange: Science and Technology, A.E. Rodrigues Ed., NATO ASI Series E, 107, pp 67 (1986). Warshawsky A., Die Angew. Makromol. Chem, 109/110, 171-196 (1981). Warshawsky A. and Kahana N., J. Am. Chem. Soc., 104, 2663-2664 (1982). Warshawsky A. and Borer B., Hydrometallurgy, 4, 83-92 (1979). Koshima H. and Onishi H., Analyst, 114, 615-617 (1989). Warshawsky A., Deshe A., Shechtman L., and Kedem O., React. Polym., 12, 261- 268 (1990). CHAPTER 6 SWELLING PROPERTIES OF POLY(METHACRYLIC ACID-g- ETHYLENE GLYCOL) HYDROGELS - EFFECT OF SOLVENT 6.1. Introduction In this chapter, a brief discussion of hydrogels and their potential applications will be followed by a study of the effect of solvent in the disruption of the complex formed between poly(methacrylic acid) and poly(ethylene glycol). For this purpose, crosslinked hydrogels with a poly(methacrylic acid) backbone with poly(ethylene glycol) grafts were synthesized and their swelling behavior in response to several conditions was investigated. These studies further our knowledge of reversible complexation and the ability of a solvent to disrupt the complex. Hydrogels are macromolecular networks which swell, but do not dissolve, in water. The ability of hydrogels to absorb water arises from hydrophilic functional groups attached to the polymeric backbone, while their resistance to dissolution arises from crosslinks between network chains. Many materials, both naturally occurring and synthetic, fit the definition of hydrogels. Crosslinked dextrans and collagens are examples of natural polymers which are modified to produce hydrogels. Classes of synthetic hydrogels include poly(hydroxyalkyl methacrylates), poly(acrylamide), poly(N- vinyl pyrrolidone), poly(acrylic acid), and poly(vinyl alcohol). 97 98 Synthetic hydrogel networks are useful for applications which require a material that has good compatibility with aqueous solvents, yet will not dissolve. Such applications include biomaterials, controlled release devices, chromatographic packings, and electrophoresis gels. Many properties of synthetic hydrogels make them suitable for biomedical applications which require contact with living tissue. The ability to absorb and retain aqueous media not only gives hydrogels a strong superficial resemblance to living tissue, but also makes them permeable to small molecules such as oxygen, nutrients, and metabolites. The soft, rubbery consistency of swollen hydrogels minimizes fiictional irritation felt by surrounding cells and tissue, while the low interfacial tension with aqueous solvents reduces protein adsorption and denaturation. Furthermore, synthetic hydrogel networks can be swollen and washed of undesirable byproducts, residual initiators and monomers, and may be fabricated in a variety of shapes and geometries. Wichterle and Liml first suggested that a hydrogel based on poly(2-hydroxyethyl methacrylate) (PHEMA) could be a biocompatible synthetic material. Since this initial work, hydrogels have been investigated for a wide variety of biomedical applications. Sofi contact lenses are an example of a biomedical application in which PHEMA hydrogels have found extensive use. The requirements of a polymer to be used in this application are quite severe. A suitable material must show optical clarity, strength, and dimensional stability under a variety of conditions. Furthermore, the material must be compatible with the cornea and must be permeable to oxygen and carbon dioxide. PHEMA hydrogels meet these requirements. The same properties have rendered these hydrogels useful for many other biomedical applications. Hydrogels have been suggested 99 for use as coatings for catheters and sutures, membranes for hemodialysis, burn dressings, artificial cartilage, heart valves, and vitreous humor, as well as many other applications. The high solute permeability of hydrogels has led to their use in devices for controlled release of drugs or other active agents. One mechanism of controlled release involves first dispersing a water-soluble drug into the hydrogel network, then drying the device. In the dry polymer matrix, the drug is essentially immobile. When the device is placed in an aqueous environment, the hydrogel swells, and the drug diffuses out of the network. The rate of drug release depends on two simultaneous processes, water migration into the device and drug diffusion out of the swollen network. The net effect is a gradual release of the drug over the course of a well defined time period, thereby maintaining close to optimum therapeutic medication levels while avoiding intermittent and massive dose effects. Furthermore, controlled release from hydrogel microspheres allows the possibility of targeting the drug to specific sites in the body where the drug is needed. This targeting may make optimum use of expensive drugs while minimizing system-wide toxic effects. Hydrogels exhibiting extremely high swelling capacities may be synthesized by free radical crosslinking polymerizations of ionogenic monomers such as acrylic and methacrylic acid (or their sodium salts). For example, poly(acrylic acid) hydrogels may exhibit a maximum swelling of more than 99% water, and are used in applications which require “superabsorbing” polymers. These superabsorbing polymers have received considerable attention in recent years, and have been extensively reviewedz‘4 Polymeric hydrogels that exhibit swelling transitions in response to external stimuli such as changes in pH, temperature, or composition have been investigated as 100 candidate materials for controlled release, separations, control of enzyme-substrate reactions, and sensor applications. Poly(methacrylic acid) and poly(ethylene glycol) form hydrophobic hydrogen bonded complexes when the acid is protonated. This hydrogen- bonded complex can be reversibly formed by species that may compete favorably for the formation of hydrogen bonds with each of the constituents. Self-associating networks of poly(methacrylic acid-g-ethylene glycol) have been investigated by Klier et al.5 and Scranton et al.6 These researchers studied the synthesis and environmentally sensitive swelling characteristics of poly(methacrylic acid-g-ethylene glycol) networks. The complex-forming constituents were covalently linked to one another and swelling was largely regulated by control of complexation. The oligomeric ethylene glycol grafts were capable of forming hydrogen bonded complexes with the PMAA backbone. The complex was formed when the acid was protonated (low pH) and could be broken by neutralizing the acid (high pH). The dependence of equilibrium network swelling on external pH, temperature, composition and network structure was examined.5 The P(MAA-g-EG) hydrogels exhibited a sharp swelling transition in aqueous solutions at an external pH of 4. These swelling characteristics were attributed to the complexation and decomplexation induced by changes in the surrounding conditions. Nuclear Overhauser Effect (N OE) studies were used to demonstrate complexation between dilute solutions of poly(methacrylic acid) and poly(ethylene glycol) (PEG) at lower PEG molecular weights than previously reported.6 This work builds upon previous work on the study of self-associating networks of poly(methacrylic acid-g-ethylene glycol) which exhibit reversible swelling transitions in response to pH and temperature. The objective of this study is to extend the 101 understanding of the reversible complexation phenomenon in response to solvent type and composition. Synthesis conditions were also varied in order to study their effect on the maximum equilibrium degree of swelling. In addition, crosslinked homopolymers of poly(methacrylic acid) have been synthesized and their swelling behavior in various solvents has been compared to the graft copolymers. Molecular modeling (molecular mechanics) techniques were also used to investigate the stoichiometry of the complex. 6.2. Experimental Methods and Techniques 6. 2. 1. Hydrogel Synthesis Poly(methacrylic acid-g-ethylene glycol) hydrogels, henceforth referred to as P(MAA-g—EG), were synthesized by copolymerization of methacrylic acid (MAA, Aldrich) with methoxy poly(ethylene glycol) methacrylate 400 (MPEGMA 400, Polysciences, Warrington, PA), in the presence of ethylene glycol dimethacrylate (EGDMA, Polysciences, Warrington, PA). Homopolymers of poly(methacrylic acid), henceforth known as PMAA, were synthesized by polymerizing MAA in the presence of EGDMA. All monomers and solvents were high purity and were used as received. The graft copolymers were synthesized by the free-radical solution polymerization of MAA and MPEGMA 400 (with about 9 EG repeat units) in the presence ethylene glycol dimethacrylate (EGDMA) as crosslinking agent (Figure 6.1). The molar ratio of ethylene glycol repeating unit to methacrylic acid was chosen to be 1:1 for all copolymers synthesized. The crosslinking agent (EGDMA) was 0.2 wt % based on monomer weight for all syntheses. The reaction was initiated by a mixture (1 :1 by 102 (EH3 (EH3 (EH3 CH2:IC CH2=C Csz'C :0 I _ C =0 3 ‘F‘° cs . + 0 + A H Ah CH2 9 2 6H Methacrylic Acid (MAA) CH2 . 2 l O 9, Y" n " c = 0 CH3 l Methoxypoly(ethylene GlycoI)-methacrylate (MPEGMMA) CH2: (,3 11:9 CH3 Ethylene glycol di-methacrylate (EGDMA) n=1 lnitiators Free Radical Solvent Polymerization 40 C V Backbone PMAA 0000 {‘9 EGDMA Crosslink £76) Complexed MPEGMMA Grafts Figure 6.1. Synthesis scheme for the preparation of poly(methacrylic acid-g-ethylene glycol) self-associating hydrogels. 103 weight) of sodium bisulfite and ammonium persulfate (0.01 wt % of monomer). The polymerization was carried out in a solution of a 50/50 (w/w) mixture of distilled water and ethanol (200 proof) to ensure proper miscibility of the initiators with the water soluble monomers. The ratio of monomers to solvent was varied between 20/80 to 80/20 (w/w) to study the effect of solvent content during synthesis upon the swelling properties of the resulting hydrogels. The reaction mixture was placed in 6 ml polyethylene vials and were thoroughly mixed using a vortex mixer for 10 minutes. The vials were sealed tight and placed vertically in a circulating water bath at 45°C for 24 hr. The resulting gels were removed from the vials and cut into small cylindrical pieces about 1 cm. long. Each of the gels was then placed in a container of distilled water for 24 hr. In order to remove any residual initiator, the gels were swollen in acidic water (pH=4) which was periodically changed for 3 days. 6. 2. 2. Equilibrium Swelling Studies Polymer gels were first weighed to record their equilibrium mass after being swollen in distilled water. Next, the hydrogels were placed in containers with various solvent-water mixtures ranging from pure water to pure solvent. The solvents studied included methanol, ethanol, acetone, n-propanol and iso-propanol. The swollen mass of the hydrogels was determined by periodically removing them from the container and weighing the drip-free sample until equilibrium had been achieved. Usually it took about a week to achieve an equilibrium degree of swelling. The hydrogel samples were then dried in an oven maintained at 65°C for 7-8 days until there was no change in mass 104 recorded. The equilibrium solvent weight fraction was determined by dividing the difference between the swollen mass and the dried mass by the swollen mass. 6. 2. 3. Molecular Mechanics Studies To study the stoichiometry of the PMAA-PEG complex, molecular mechanics simulations were performed using the molecular modeling software POLYGRAF.7 The polymer builder module was used to build syndiotactic PMAA and PEG segments (each of 4, 8 and 14 repeating units). The structures were energy minimized and charge equilibrated to facilitate the assumption of realistic minimum energy conformations with realistic molecular geometries. The two polymer segments were then placed in close proximity and energy minimization calculations (molecular mechanics, MM) were performed until an energy minimum was reached and the calculations converged. In order to ensure that the minimum attained was not a local minimum, molecular dynamics (MD) calculations were performed on the system for a period of 1 ps, followed by further energy minimization calculations. (For a brief description of both MM and MD techniques, please refer to Chapter 1). Hydrogen bonds between the ether oxygens and the acid protons were said to be formed by setting a cutoff distance of 3.5 A and highlighting the oxygen-proton pairs that satisfied the criteria. 105 6.3. Results and Discussion 6. 3. 1. Hydrogel Swelling Studies The equilibrium swelling profiles for P(MAA-g-EG) gels synthesized in a 50/50 monomer/solvent ratio are shown in Figure 6.2. The three curves represent data for swelling in aqueous mixtures of solvent, from pure water to pure solvent. A homologous series of alcohols was chosen for this study and the data shown in Figure 6.2 illustrates the equilibrium swelling behavior of the P(MAA-g-EG) hydrogels in methanol, ethanol and n-Propanol. The data are represented by the symbols, and lines drawn through the data points represent trends. As shown in Figure 6.2, an equilibrium solvent weight fraction just over 0.3 is observed in the limit of infinite dilution (pure water or 0 weight fraction). In addition, the equilibrium degree of swelling increases as the weight fraction of solvent in the water-solvent mixture increases. This trend is consistent with the idea that solvents with Lewis base character may compete with PEG for hydrogen bonding sites on the backbone PMAA, resulting in an increased degree of swelling due to decomplexation of the covalently bonded grafts with the acid backbone. The effectiveness of a solvent in breaking the complexes is probably related to its ability to disrupt the hydrophobic interactions as well as compete for hydrogen bonding sites. Also, other than in methanol, in which an equilibrium solvent weight fraction of about 0.8 is observed in pure methanol, both ethanol and n-propanol curves go through a maximum and decrease to a lower degree of swelling as the amount of solvent in the aqueous mixtures increases. 106 1.0 0.9.3 0.8-I A -. .. -_..-._._.........., 0.71 0.6: 0.5; 0.3 - Eq. Solvent Wt. Fraction 0.2 a ~——l-——- Methanol -—O—-— Ethanol 0.1 - —V— n-Propanol 0.0 l l T l I T l T 0.0 0.2 0.4 0.6 0.8 1.0 Wt. Fraction Solvent in Water Figure 6.2. Effect of solvent on the equilibrium degree of swelling for a PMAA-g-PEG gel synthesized in 50% solvent (50 % monomers). 107 The swelling results at low concentrations of solvent demonstrate that n-propanol is most effective in breaking the complex (as indicated by the slope of the trendline at lower concentrations). However, gels swollen in ethanol or methanol achieve a higher equilibrium degree of swelling (between 0.3 and 0.7 weight fraction solvent in water) as compared to n-propanol. This trend may be explained by the ability of n-propanol to disrupt the complex at lower concentrations by better stabilizing the hydrophobic interactions due to the propyl group. Swelling transitions of P(MAA-g-EG) gels in alcohol solvent mixtures may be attributed to two competing phenomena: i) the ability of the solvent to disrupt hydrogen bonding and stabilize hydrophobic interactions, and ii) the dissociation/ionization of the polymeric acid in the solvent. As shown in Figure 6.2, swelling in the homologous series of alcohols with increasing hydrophobic character results in a decrease in the equilibrium degree of swelling at high solvent concentrations. This downward trend may be explained by the re-association of the PMAA in environments that are more non-polar and inhibit ionization. The pK, of PMAA may be affected by the ease with which the acid proton can exchange with the solvent molecules. For example, the pK, of PMAA in water varies between 4 and 4.8 and allows for rapid exchange between the acid proton and the dissociated water protons. At higher solvent concentrations PMAA ionization may be the controlling factor which affects complexation and thus swelling behavior. In environments such as high concentrations of n-propanol or ethanol, the complexes may be re-formed due to the re-association of the acid, resulting in a lower equilibrium degree of swelling. Figure 6.3 illustrates the effect of polymerization conditions on the maximum equilibrium degree of swelling achieved by a P(MAA-g-EG) hydrogel. In this figure, 108 1.0 0.81 0.7- 0.6: 0.53 0.4 - 0.3- Eq. Solvent Wt. Fraction b 0.2_ —‘-— 20% Monomer . —O— 50% Monomer 0.1 - + 80% Monomer 0.0 I I I I I T I I 0.0 0.2 0.4 0.6 0.8 1.0 Wt. Fraction Ethanol in Water Figure 6.3. Equilibrium swelling profiles showing the effect of polymerization conditions for swelling in ethanol. 109 data are shown for three different (20/80, 50/50 and 80/20) monomer to solvent ratios during synthesis. The swelling profiles are for swelling in various concentrations of aqueous ethanol mixtures. For all three cases, the equilibrium degree of swelling increases as the concentration of ethanol in the ethanol-water mixture increases and reaches a maximum at around 0.6 wt. fraction ethanol in water. As shown in Figure 6.3, gels synthesized at a lower monomer to solvent ratio (20% monomer) are capable of achieving an equilibrium solvent fraction of 0.9 at intermediate solvent concentrations, while those synthesized at a high monomer to solvent ratio (80% monomer) reach an equilibrium solvent weight fraction of only about 0.7 at similar solvent concentrations. These trends demonstrate that the gels synthesized in more solvent exhibit a correspondingly higher equilibrium degree of swelling as compared to gels synthesized in a higher monomer to solvent ratio. This may be explained by considering the conformations of the oligomers and growing polymer chains during polymerization. In a dilute system, the monomers and crosslinking agents may be relatively far apart resulting in a decrease in the crosslinking density (and entanglement) and an increase in the sol fraction. However, at higher monomer to solvent ratios, close proximity of growing polymer chains to crosslinking agents may result in a higher degree of crosslinking (and entanglement) and incorporation of more monomer units into the hydrogel network, thereby decreasing the sol fraction. To demonstrate the effect of complexation on the equilibrium degree of swelling of a P(MAA-g-EG) hydrogel, comparisons were made with the swelling characteristics of a crosslinked PMAA gel synthesized under the same conditions. Figure 6.4 shows the swelling behavior of a PMAA hydrogel compared to a P(MAA-g-EG) hydrogel 110 c .9. *6 9 LL E E (D 2 O (I) IB- 4 0.2: —l-— PMAA Q1 _ “~0— 50/50 PMAA-g-PEG 0.0 I I I I I I I I I 0.0 0.2 0.4 0.6 0.8 1.0 Wt. Fraction Methanol in Water Figure 6.4. Equilibrium swelling profiles of crosslinked PEG-g-PMAA (SO/50) and PMAA homopolymer (SO/50) in methanol. lll (synthesized in a 50/50 monomer solvent ratio). The swelling profiles shown are for swelling in increasing concentrations of aqueous methanol solutions. The swelling characteristics of the P(MAA-g-EG) gel follow a sigmoidal trend with a low equilibrium solvent uptake at infinite dilution (0 weight fraction methanol) and ~0.8 equilibrium solvent weight fraction at higher concentrations of methanol. On the other hand the swelling characteristics of PMAA show little variation with increasing methanol concentration. The slight increase (from a value of 0.7 solvent weight fraction) followed by a nominal decrease to an equilibrium solvent weight fraction of 0.75 may be explained by the varying degree of dissociation of the polyacid in different concentrations of aqueous methanol solutions. The effect of complexation can be easily demonstrated by the difference in the equilibrium degree of swelling between the PMAA and the P(MAA- g-EG) gels at low methanol concentrations. This difference can be attributed to the compact configuration (low degree of swelling) adopted by the complexed hydrogel at low methanol concentrations. However, as the concentration of methanol increases it is able to compete effectively for hydrogen bonding sites on the backbone PMAA and disrupt the hydrogen bonding between the grafted PEG segments and the PMAA. In the absence of any grafts the equilibrium degree of swelling of the PMAA hydrogel is a function of the degree of dissociation in a particular aqueous methanol solution, and does not exhibit the lower degree of swelling demonstrated by the P(MAA-g-EG) hydrogel. Figure 6.5 illustrates the difference in swelling behavior in acetone between a PMAA gel and a P(MAA-g-EG) gel synthesized in a 50/50 monomer solvent ratio. The two curves once again demonstrate the effect of complexation on the equilibrium swelling behavior of these hydrogels. However, since acetone is a comparatively more 112 1.0 I: .9 ‘6 ‘3 LL 3 E (D .2 O (0 . U" 0.24 LU 4 —l_ PMAA 0.1 - -O- PMAA-g-PEG 0.0 I I I I I I I 1 I 0.0 0.2 0.4 0.6 0.8 1.0 Wt. Fraction Acetone in Water Figure 6.5. Equilibrium Swelling profiles of crosslinked PMAA-g—PEG (50/50) and PMAA (50/50) in acetone. 113 hydrophobic solvent (with two methyl groups) compared to methanol, at high acetone concentrations the pKa of the PMAA may decrease and the acid may undergo re— association resulting in a more compact configuration evidenced by a lower equilibrium solvent weight fraction. Figure 6.5 brings out the two effects of complexation and dissociation quite clearly as seen by, i) the difference in equilibrium swelling at low concentrations between the two hydrogels and ii) a lowering in the equilibrium degree of swelling at high acetone concentrations shown by both the PMAA and the P(MAA-g—EG) hydrogels. 6. 3. 2. Molecular Modeling of PMAA-PEG Complexation It has been demonstrated by numerous investigators (see reference 5 and references therein) that PMAA and PEG form a 1:1 complex based on a molar ratio of methacrylic acid and ethylene glycol repeat units. Klier et al.5 demonstrated that P(MAA-g-EG) hydrogels synthesized with a 1:1 molar ratio exhibited the least equilibrium degree of swelling at a pH of 4 (complex promoting conditions). It has also been proposed that deviations from this stoichiometry may lead to free polymer chains or increased hydration of the complex (due to the presence of uncomplexed hydrophilic units within a complexed chain). Experimental work by various researchers provides evidence of the formation of the complex in a 1:1 stoichiometry. However, these observations leave significant questions about the fundamental nature of the complexes, and do not conclusively establish that the complexes form in a strict 1:1 stoichiometry with each acidic moiety bound with each successive ether linkage. 114 In this study, we have attempted to demonstrate the stoichiometry of the PMAA- PEG complex (under complex promoting conditions when the acid is protonated) by performing molecular simulations on segments of PMAA and PEG of varying lengths. Syndiotactic PMAA was built using the POLYGRAF polymer builder, and segments consisting of 4, 8 and 14 repeat units were studied for complexation with segments of PEG with exactly the same number of repeat units. Molecular mechanics calculations were first performed on a tetrad of PMAA in close proximity to a tetrad of PEG. The objective of these calculations was to demonstrate that sterically it is possible to form stable 1:1 complexes between four repeating units of PMAA and PEG. Molecular mechanics calculations demonstrated the formation of such as complex by satisfying the hydrogen bonding criteria. The ratio of energy contribution due to hydrogen bonding to the total energy of the minimized structure was higher, compared to the starting conformation. Figure 6.6 shows the PMAA tetrad complexed with a PEG segment with 4 EG repeat units thereby validating the 1:1 complex stoichiometry. This conformation was the lowest energy conformation hence most probable and favorable. In addition, the same complexed state was attained after perturbing the system by molecular dynamics calculations at 25°C followed by energy minimization. A molecular dynamics run ensured that the conformation observed earlier was not at a local minimum and was indeed the favored conformation. It can also be observed that, though the PEG repeat unit contains three atoms (-C-C-O) in the backbone and the PMAA repeat unit contains two atoms (-C-C-) in the backbone, it is possible to form a 1:1 complex. To further extend this study to longer segments, 115 ( -bonds it \ I. , II PEI} Figure 6.6. A molecular representation of a PMAA segment with four repeating units complexed with a PEG segment of the same length showing 1:1 hydrogen bonding between the acid protons and the ether oxygens. 116 calculations with 8 and 14 repeating units yielded similar results. Figure 6.7 represents an equilibrium conformation of a segment consisting of 14 repeat units of PMAA complexed with a PEG segment with the same number of repeating units. The curved shape of the complex points to the possibility of a helical structure when the number of repeat units are higher. This would be consistent with the fact that the PEG chain may have to wrap around the PMAA chain due to an extra atom present in its repeat unit, in order to form a 1:1 complex. 6.4. Conclusions The swelling characteristics of graft copolymers of poly(methacrylic acid) with poly(ethylene glycol) grafts were investigated. The swelling studies were performed in increasing concentrations of water-solvent mixtures and the equilibrium solvent weight fraction was recorded. Swelling behavior in a homologous series of alcohols (methanol, ethanol and n-propanol) demonstrated the effectiveness of the solvent in disrupting the complex and stabilizing the hydrophobic interactions. The ability of a solvent to break the complex (at low solvent concentrations) may be related to the ease with which it could compete in the formation of hydrogen bonds with the PMAA backbone. However at high solvent concentrations the degree of dissociation of the PMAA could possibly govern the equilibrium degree of swelling. For alcohols with longer alkyl groups (n- propanol) the degree of swelling decreased gradually at higher solvent concentrations. The ratio of monomer to solvent during synthesis affected the maximum equilibrium degree of swelling attained by a hydrogel. Hydrogels synthesized in a 20/80 monomer 117 Figure 6.7. A POLYGRAF molecular simulation depicting a 1:1 hydrogen bonding between a PMAA and PEG segment both with 14 repeating units. 118 solvent ratio achieved a higher degree of swelling which may be attributed to a lower degree of crosslinking and entanglement. Comparisons between the swelling behavior of a crosslinked homopolymer of PMAA with a P(MAA-g-EG) gel in methanol and acetone demonstrated the effect of complexation as evidenced by the difference in the equilibrium degree of swelling at low solvent concentrations. A lowering in the degree of swelling at high concentrations of acetone or n-propanol was attributed to a lower degree of dissociation of the polyacid (PMAA). Molecular simulations strengthened the evidence of the existence of a 1:1 complex stoichiometry, by demonstrating 1:1 complexation at the molecular level. Molecular mechanics simulations on PMAA and PEG segments with 4, 8 and 14 repeat units exhibited the formation of stable conformations with a 1:1 complex stoichiometry. 119 6.5. References Wichterle o. and Lim 1)., Nature, 185, 117 (1960). Gross J. R., in Absorbent Polymer Technology, L. Brannon-Peppas and R. S. Harland Eds., Elsevier Science Publishing Company Inc., New York, pp 3-22 (1990) Buchholz F. L., in Absorbent Polymer Technology, L. Brannon-Peppas and R. S. Harland Eds., Elsevier Science Publishing Company Inc., New York, NY, pp 23- 44 (1990). Buchholz F. L., Trends in Polymer Science, 2(8), 277-281 (1994). Klier J., Scranton A. B., and Peppas N. A., Macromolecules, 23, 4944-4949 (1990) Scranton A. B., Klier J., and Aronson C. L., in Polyelectrolyte Gels, R. S. Harland, R. K. Prud’homme Eds., ACS Symposium Series 480, Washington DC, pp 170-189 (1992). POLYGRAF, v 3.21, Molecular Simulations Inc., 16 New England Executive park, Burlington, MA 01803. CHAPTER 7 SYNTHESIS AND EMULSIFICATION PROPERTIES OF REVERSIBLE BLOCK/CRAFT POLYMERIC EMULSIFIERS BASED UPON POLYMER COMPLEXATION 7.1. Introduction In this chapter, the synthesis and emulsification properties of block/graft polymeric emulsifiers based upon polymer complexation, is presented. The motivation behind the development of reversible block/graft emulsifiers has been presented in Chapter 1. These emulsifiers could potentially impact enhanced oil recovery1 and pipeline emulsion transportation of heavy oils.2 The reversible emulsification ability of these polymers may be useful in both applications. For example, oil-in-water emulsions provide a cost-effective alternative to heated pipelines or the addition of diluents for the pumping of heavy oils. Typical transport emulsions are composed of 70 wt.% oil, 30 wt.% aqueous phase and 500-2000 ppm of a stabilizing surfactant formulation.2 The resulting emulsion has a viscosity in the range of 50-200 cP at the pipeline operating conditions as compared to heavy oil viscosities that may range from 10,000 to 100,000 cP (see Figure 7.1A). Nonionic surfactants may be used for this purpose due to their relative insensitivity to the salt content of the aqueous phase. However, once pumping has been completed the oil-in-water emulsion needs to be broken for the recovery of oil (Figure 7.1 B). This may be achieved by either the addition of de-emulsifiers3 or by boiling off 120 121 100,000 ‘ ~ - ~ ~ Heavy Crude Oils and Bitumens 10,000 , 1,000 ‘ * ~ ‘ ‘ Operating Limit (Crude) Viscosity (cP) 100 Oil-in-water emulsions 10 l 1 g I l l l 40 60 80100120140160 T(°F) (A) Heavy Crude Mixer Emulsion Storage Pipeline System Water 0» Surfactant Emulsion Treatment Dry Crude ‘ ............ ......... —> Water (B) Figure 7.1. (A) Reduction of the viscosities of heavy oils by the formation of oil-in-water emulsions, and (B) A schematic showing a treatment process that may be used for the pumping of heavy oils by forming oil-in-water emulsions. 122 the water and in some cases, a combination of both techniques. Either way, this step represents enormous energy costs and the addition of a de-emulsification agent. If the emulsification and de—emulsification can be performed by the same emulsifier without the encumbrance of high energy costs, it would result in substantial savings and relatively simple operating conditions. The block/graft emulsifiers presented in this chapter could have potential applications in such situations requiring emulsification and de- emulsification under different sets of conditions. Simple diblock copolymers have also been used as polymer compatibilizers and dispersion stabilizers for many years. Theoretical work by Noolandi4 has suggested that multi-block copolymers would be more efficient as compatibilizers than di- or triblock copolymers. His results suggest that less of the multi-block copolymer would be lost into the bulk phase as micelles or mesophases than the di- or triblock systems. Similarly, multi-block dispersing agents and stabilizers should be more efficient than the common di or triblock systems. In this research, a potentially new kind of block copolymer, the block/graft copolymer, has been investigated. Block/graft copolymers consist of a polymer backbone that has side chain grafts extending linearly off the polymer backbone at regular intervals. The graft side chains are capable of complexing with the backbone under certain conditions to form blocks of hydrophilic and relatively hydrophobic polymer as shown in Figure 7.2. The complexation occurs by hydrogen bonding under acidic conditions. Because of the regular side chains, block/graft copolymers are often referred to as “comb” copolymers in their uncomplexed state. 123 Basic Conditions / Poly(ethylene glycol) grafts Hydrophilic methacrylic acid backbone Acidic Conditions Hydrophobic chai sections due to omplexation of PEG gr fts with backbone Hydrophilic sections of backbone not complexed with PEG grafts Figure 7.2. Schematic and mechanistic depiction of the block/ graft copolymer, consisting of a poly(methacrylic acid) backbone and poly(ethylene glycol) grafts, under acidic or complex promoting conditions and under basic or complex breaking conditions. 124 In this chapter, a simple one-step method for producing reversible block/graft copolymers for applications as dispersing agents and stabilizers, is presented. These novel polymers have been formed by free radical polymerization of a vinyl monomer with small amounts of a macromonomer. The polymers consist of a methacrylic acid backbone and oligo(ethylene glycol) grafts. The oligo(ethylene glycol) grafis are incorporated into the chain during the polymerization of the methacrylic acid, eliminating the need for a multi- step synthesis. The backbone Lewis acid repeat units are capable of forming hydrogen bonded complexes with the Lewis base ether oxygens of the oligomeric grafis (Figure 7.3B). These complexes are formed when the acid is protonated or under acidic conditions and result in the complexed regions of the backbone becoming hydrophobic. By choosing the ratio of methacrylic acid to ethylene glycol repeat units it is possible to form pseudo-multiblock copolymers consisting of a hydrophilic acid backbone interspersed with hydrophobic segments due to complexation. Such a polymer behaves like a multiblock polymeric emulsifier under acidic conditions and as a hydrophilic graft copolymer under basic conditions. Thus this reversible emulsification behavior can be turned “on” or “off’ by controlling the pH or other variables that affect complexation. The emulsification behavior of these novel reversible emulsifiers has been investigated as a function of pH. The effect of hydrophilic segment length has also been studied by varying the ratio of the methacrylic acid to ethylene oxide repeat units. Variables such as emulsifier concentration and chemical structure and their effect on the emulsification capacity, have been investigated. Interfacial tension measurements were performed to elucidate the surface active behavior of these polymers as function of solution pH. 125 —> rrrrvyw I I l o b (5 o b b o WWW/W _/\_/\_/\_/\_/\_/\_/\_ (A) CH3 CH3 CH3 CH3 CH3 CH3 CH3 I 10“ CH3\. / \ / \ / \“m o o o 0 .. 2 : : : 9 i i i i *i l o o o o o O I I I I I I C=O C=O C=O C=O C=O C=O CH3 CH3 CH3 CH3 CH3 CH3 (B) Figure 7.3. Hydrogen bonded complexes formed between poly(methacrylic acid) and ' poly(ethylene glycol) units in (A) non-grafted and, (B) grafted configurations. 126 7.2. Experimental Methods and Techniques 7. 2. 1. Copolymer Synthesis The graft copolymers were synthesized via a free-radical solution polymerization of methacrylic acid (MAA) with methoxy poly(ethylene glycol) methacrylate (MPEGMA) using hydrogen peroxide (30% solution) as initiator and a 50/50 mixture of ethanol/water as the solvent (see Figure 7.4). The macromonomer employed was MPEGMA 1000 consisting of an average 22.5 number of ethylene glycol (EG) repeating units. Polymers with different molar ratios of MAA to EG repeat units (such as 5:1, 10:1 and 20:1) were synthesized. All chemicals were high purity and were used as received. For a repeat unit ratio of 20:1 MAAzEO the synthesis is described below. MAA (Polysciences, Warrington, PA) (1.945 g) was added to 0.055 g of MPEGMA 1000 (Polysciences) in a clean glass bottle. To this mixture, 8 grams of solvent was added comprising a 50/50 mixture of ethanol (200 proof) and water (by weight). The reaction was initiated by adding 40 uL of a 30% solution of hydrogen peroxide as a thermal initiator. The sample was placed in a controlled temperature water bath at 80°C and polymerized for 36 hours, after which it was dried in an oven at 80°C. The resulting solid polymer was ground using a mortar and pestle. Polymer solutions were made by dissolving required amounts of dry polymer in HPLC grade water with constant stirring. The solutions were stirred overnight or until a clear solution was obtained. The solutions were filtered using a coarse #4 qualitative (Whatrnan) filter paper under vacuum in order to remove any suspended impurities. 127 . . tawny” . C O . . ' g ‘ . . 50/50 EtOHNVater 37,; O .‘r.¢"s,% ’ we fit , . 0 ' H202 i; ’ . ' é ' Heat . O I‘D Q W \ . CH3 HzCi 0 OH \ MAA / kMPEGMA1000J Figure 7.4. Schematic showing the free-radical polymerization scheme for the synthesis of reversible block/graft emulsifiers. 128 7. 2. 2. Emulsification Studies Emulsification studies were performed to determine qualitatively and quantitatively the emulsification ability of the copolymers. For these studies, a 0.1 wt.% solution of the copolymer was prepared in HPLC grade water under constant stin'ing. Methyl laurate (Aldrich) was chosen as the “oil” phase. Samples, each consisting of 10 ml of the polymer solution, were adjusted to different values of pH by the addition of small volumes of either lM HCl or IM NaOH solution. In 10 ml stoppered graduated cylinders, 2 m1 of methyl laurate was added to 8 ml of the pH-adjusted polymer solutions (Figure 7.5). The mixture was agitated vigorously for about 2 minutes and then allowed to stand for 4 days or until no change in the emulsion quality and volume was noted. The percent oil emulsified was calculated by measuring the volume of un-emulsified clear oil in the oil phase and the volume of the cream (stable emulsion). 7. 2. 3. Interfacial Tension Measurements All interfacial tension measurements were performed on a Kriiss Processor Tensiometer K12 (Krfiss, USA) equipped with a controlled temperature water bath. The Wilhelmy plate method was used to determine the interfacial tension between the oil phase (methyl laurate) and the aqueous phase. Measurements were performed in a 49 ml (small) cup at 25°C as per the instrument procedure for interfacial tension measurements. Three readings per sample were recorded and the average of the three was reported as the interfacial tension. Polymer solutions at various pH for the 10:1 and 20:1 (MAA to EG repeat unit ratio) polymers were prepared from a 0.1 wt.% stock solution. As before the 129 Oil Emulsion “Cream” Agitate Emulsifier we“ —> solution N \ Microemulsion or water Emulsion Methyl laurate - oil phase Figure 7.5. Schematic of the methodology adopted for the oil emulsification studies. 130 pH of the samples was adjusted by addition of small quantities of 1M HCl or IM NaOH solutions. Interfacial tension measurements were performed for solution pH ranging between 2 and 12 for both the 10:1 and 20:1 copolymer solutions. 7.3. Results And Discussion 7. 3. I . Emulsification Studies The emulsification properties of two different block/ graft copolymeric emulsifiers is illustrated in Figure 7.6. The figure represents results from tests performed with 20 % (v/v) methyl laurate dispersed in a 0.1 wt.% polymer solution. The figure demonstrates that even at fairly low concentrations under acidic conditions, the copolymers are able to emulsify 20 vol.% oil by forming a stable emulsion. In addition, the emulsions may be readily broken by raising the pH, with the percentage emulsified dropping to zero under basic conditions. It is interesting to note that these properties are reversible, and that a system cycled from acidic to basic conditions will reform a stable emulsion if the pH is returned to acidic conditions. The data in Figure 7.6 also illustrates that molecular architecture plays an important role in the performance of the emulsifiers. The figure shows that the emulsification ability, as characterized by the percent oil emulsified, depends upon the methacrylic acid to ethylene glycol repeat unit ratio. The 20:1 MAA to EG polymer exhibits a decrease in the emulsification capacity at a considerably lower pH than the 10:1 system. The lines in Figure 7.6 are drawn to illustrate the emulsification trends for the two repeat unit ratios under investigation. 131 1001 80* 60- 40‘ Percent Oil Emulsified Figure 7.6. Emulsification behavior of two different block/graft copolymers (10:1 and 20:1) as a function of pH showing the percent oil emulsified (v/v) as a function of pH. 132 7.3. 2. Interfacial Tension Studies The data presented in Figure 7.7 illustrates the trends observed for the interfacial tension between the oil (methyl laurate) phase and the aqueous phase containing the polymeric emulsifiers. Once again, data for both the 10:1 and the 20:1 copolymers is represented by the symbols whereas the lines show trends. Under acidic conditions, it can be observed that the interfacial tension is lowered to a value of about 10 mN/m as compared to a value of about 21 mN/m under neutral conditions. The interfacial tension between methyl laurate and pure water, in the absence of any polymer, was determined to be 21.5 mN/m at 25°C. Both copolymeric emulsifiers under investigation exhibit similar trends, however, the 20:1 copolymer exhibits a marginally lower interfacial tension under acidic conditions. We attribute the observed emulsification and lowering of interfacial tension properties exhibited by the reversible block/graft copolymers to a self associating phenomenon in which the grafts form complexes with the backbone. In this system, the self-complexation of the block copolymer occurs because of hydrogen bonding between the ether linkages of the poly(ethylene glycol) (PEG) and the acid moieties of the poly(methacrylic acid) (PMAA). In the complex, the hydrophilic moieties are bound with one another as illustrated in Figure 7.2. Therefore the complex is significantly more hydrophobic than its constituent polymers, and the complexed structure mimics a pseudo- multiblock copolymer architechture. The fact that the complex is significantly more hydrophobic than the uncomplexed polymers has been established by the study of the swelling properties of crosslinked poly(methacrylic acid-g-ethylene glycol) networks 133 22‘ - 10:1 19‘ 2 E, C .9 (I) C m I... 33 8 ‘l: 9. 5 10‘ ' Wilhelmy Plate Method ‘ (Methyl Laurate - Water) 8 I I I I l I 1 l I l r O 2 4 6 8 10 12 pH Figure 7.7. Interfacial tension between the oil (methyl laurate) phase and the aqueous phase containing the emulsifiers, as a function of pH. Data is presented for the 10:1 and the 20:1 copolymeric emulsifiers. 134 reported by Klier and collaboratorss'6 For example, these authors concluded that the aqueous swelling minimum observed under acidic conditions for gels with 1:1 MAAzEG repeat unit ratios can be attributed to the formation of hydrogen bonded hydrophobically stabilized complexes. They also observed a reduction in the equilibrium degree of swelling with temperature for non—stoichiometn'c gels due to the increase in strength of the hydrophobic interactions between the complexing polymers.‘5 In addition, the hydrogen- bonded hydrophobic complex formed between hydrophilic Lewis acid carboxylic groups and hydrophilic Lewis base groups can be disrupted by a change in pH, solvent type or temperature. Due to this phenomenon the block/graft copolymers described in this chapter can be used as reversible emulsifiers which can be turned “on” or “off” by triggering a “switch” such as pH, solvent or temperature. The emulsification trends can be explained by the formation and disruption of the complex between the hydrophilic backbone and the ethylene glycol grafts. For the MAAzEG molar ratios being discussed here, the long hydrophilic sections of the complexed polymer would sterically stabilize oil-in-water emulsions as shown in Figure 7.8. The hydrophobic sections of the pseudo-block copolymer would preferentially partition onto the surface of an oil droplet while the hydrophilic segments would impart solubility in the bulk aqueous phase. This would prevent droplet coalescence and thus result in a stable emulsion (Figure 7.8). Since the complexes are formed under acidic conditions and can be disrupted under basic conditions, the copolymers are able to emulsify oil only under complex promoting conditions. For the 20:1 copolymer, a decrease in the percent oil emulsified with an increase in pH may be attributed to an increase in the hydrophilic carboxylic acid segments, resulting in inefficient 135 Hydrophilic Sequences ‘— Water Insoluble Complex Sequence Basic conditions Acidic conditions (A) KEV Figure 7.8. Schematic representation of the possible emulsification mechanism adopted by the reversible emulsifiers (A) under acidic and basic conditions, and (B) the steric stabilization that may result. 136 emulsification. A decrease in the MAA to EG ratio may result in shorter hydrophilic segments per chain, thereby facilitating better partitioning into the oil phase under acidic conditions. However, both copolymers display a trend in which they emulsify oil under acidic conditions and act as poor emulsifiers under basic conditions. When base was added to the acidic samples, the emulsion was broken instantly upon agitation resulting in a two-phase system. Upon lowering the pH and vigorous mixing thereafter, the emulsion was re-formed thereby elucidating their reversible nature. Hence pseudo-multi block/graft copolymers may be “tailor made” to work as reversible emulsifiers in the pH range of choice by controlling the MAA to EG repeat unit ratio. The emulsification behavior may also be controlled by addition of hydrophilic or hydrophobic oligomers as grafts to impart additional hydrophilicity or hydrophobicity. The lowering of interfacial tension may also be explained by the formation of a pseudo-multi block structure under complex promoting conditions. With an alternating hydrophobic-hydrophilic architecture, the polymeric emulsifiers may form aggregates much like micelles. However, these aggregates may be formed either by a single chain if it can assume a suitable conformation or by the cooperative effort of several chains. The aggregates may consist of a hydrophobic interior and a hydrophilic domain extending into the bulk aqueous phase. This picture is consistent with the emulsification studies, since oil droplets may be preferentially incorporated into the interior of such aggregates. Similarly, the lowering of interfacial tension may be attributed to the preferential partitioning of the hydrophobic segments into the oil phase all along the oil-water interface. In the following chapter, a fluorescence spectroscopy study will be performed 137 with pyrene as a fluorescent probe to study the aggregates formed by these copolymers as a function of pH and concentration. The block/graft copolymers may be useful in a variety of applications which may require emulsification under certain conditions, while phase separation may be desired under other conditions. One such example is an aqueous cleaning system which offers all of the separation advantages of an emulsifying system but also allows (by breaking the emulsion) the oil and water to be readily separated from one another. After the emulsion is broken, the oil phase may be readily skimmed from the water, and the aqueous phase (including the surfactant) may be recycled. These types of emulsifiers would ultimately affect a variety of industries ranging from petroleum processing (for wastewater treatment and separations, and enhanced oil recovery) to automotive service stations. 7.4. Conclusions The synthesis and emulsification behavior of novel block/graft copolymeric emulsifiers based upon a reversible hydrophobic architecture has been presented. In these systems, the graft copolymer assumes a multiblock copolymer configuration under certain conditions. The copolymers exhibit reversible emulsification behavior and may be useful in a variety of applications. Emulsification studies demonstrated that the copolymers can emulsify oil under acidic conditions in a reversible fashion. As the pH of the emulsifier solution is raised towards neutral, the percent oil emulsified decreases, and under basic conditions no emulsification is observed. Interfacial tension studies illustrated the surface active nature of the copolymers under acidic conditions. A 138 lowering of the interfacial tension was observed under acidic conditions for the copolymeric emulsifiers under investigation. These properties exhibited by the block/graft copolymers can be attributed to the formation of hydrogen bonded hydrophobic complexes between the methacrylic acid backbone and the ethylene glycol repeat units on the graft. This may result in an alternating hydrophilic-hydrophobic pseudo-multi block structure which may be responsible for the surface active properties of these copolymers. 139 7.5. References Taylor KC. and Hawkins BE, in Emulsions: Fundamental Applications in the Petroleum Industry, L.L. Schramm Ed., ACS Advances in Chemistry Series 231, pp 263 (1992). Rimmer D.P, Gregoli A.A., Hamshar J.A. and Yildirim E., in Emulsions: Fundamental Applications in the Petroleum Industry, L.L. Schramm Ed., ACS Advances in Chemistry Series 231, pp 295 (1992). Grace L., in Emulsions: Fundamental Applications in the Petroleum Industry, L.L. Schramm Ed., ACS Advances in Chemistry Series 231, pp 313 (1992). Noolandi, J., Makromol. Chem. Theory Simul. 1(5), 295-8 (1992). Scranton A.B., Klier J ., Aronson C.L., in Polyelectrolyte Gels, R. S. Harland and R. K. Prud’homme Eds., ACS Symposium Series 480, Washington, DC, pp 171 (1992) Klier J., Scranton AB, and Peppas N.A., Macromolecules 23, 4944 (1990). CHAPTER 8 A FLUORESCENCE SPECTROSCOPY STUDY OF THE AGGREGATE FORMATION BEHAVIOR OF REVERSIBLE COPOLYMERIC BLOCKIGRAFT EMULSIFIERS 8.1. Introduction In the previous chapter, the synthesis and emulsification properties of block/graft copolymeric emulsifiers were presented. The chemical nature of the reversible hydrogen bonded complexes between backbone methacrylic acid repeat units and the oligomeric ethylene glycol grafts, was also discussed. The oil emulsification properties of the copolymers were presented as a function of pH. The copolymers emulsified oil under acidic conditions since the grafts form hydrophobic complexes with the backbone resulting in an alternating hydrophobic-hydrophilic structure. It was also demonstrated that these complexes were reversible and that, upon raising the pH, the emulsion could be broken, and reformed thereafter upon agitation if the pH was lowered. Interfacial tension measurements between a model oil (methyl laurate) phase and the aqueous polymer solution phase exhibited a lowering of the interfacial tension for the samples at acidic pH as compared to those under neutral or basic pH. The studies mentioned above, have established the surface active nature of the block/graft copolymers under acidic pH. In this chapter, the aggregate formation behavior of the copolymeric emulsifiers under acidic conditions, has been investigated. It 140 141 is hypothesized that the aggregates may consist of a hydrophobic interior which may be responsible for the emulsification properties observed in the previous chapter. The environment-sensitive fluorescence of pyrene has been utilized to study the formation of the aggregates with hydrophobic interiors. Pyrene not only would serve as a fluorescent probe but also as the model “oil” phase for these studies. Pyrene has been used as a polarity sensitive probe due to the effects of iosmeric solvents on the vibronic band intensities in the fluorescence spectrum of pyrene' (also known as the Ham2 effect). This environmental effect on the vibronic band intensity in pyrene monomer fluorescence was utilized by Kalyanasundaram and Thomas3 in studies of micellar systems. The strong perturbation of the vibronic band intensities was used as a probe to accurately determine critical micelle concentrations and also to investigate the extent of water penetration in micellar systems. An excellent review article by Turro et al.4 discusses the use of luminescence probes as sensors for probing the microenvironment of micellar systems. More recently, both solubilized pyrene as well as pyrene end-labeling has been used to study micellar solutions formed by block copolymer surfactants. For example, the use of pyrene end labeled poly(ethylene oxide) to study its interactions with sodium dodecyl sulfate (SDS) in aqueous solution, has been reported.’ These investigators observed an increase in the excimer/monomer emission intensity with an increase in SDS concentration which was explained by the incorporation of the pyrene end—labeled poly(ethylene oxide) into aggregates comprising two or more end-labeled chains.5 Similarly, Prochazka et al.° and Kiserow et al.7 have utilized the time-resolved fluorescence of naphthalene to study the chain dynamics of micelles formed by napthalene-labeled diblock and triblock copolymers in aqueous media. It was observed 142 that intramolecular excimer formation (which is controlled by the mobility of the pendant fluorescent groups and the polymer chain dynamics) is sterically hindered in the micellar cores as they become more compact.° Bakeev et al. have used l-pyrenecarboxylaldehyde as a fluorescence probe to study the complexation of amphiphilic polyelectrolytes with surfactants of the same charge in aqueous solutions.8 In addition, the pH and electrolyte related chain dynamics of a water soluble terpolymer based on acrylamide and the surface active monomer sodium 11 acrylamidoundecanoate has been investigated by incorporation of 2-(l-pyrenylsulfonamido) ethyl acrylamide into the monomer.9 The steady state fluorescence studies conducted by these investigators revealed significant constrictions of the polymer chains as pH decreases or electrolyte concentration increases. Maltesh et al. studied the effect of binding of cations to pyrene end-labeled polyethylene glycol (PEG) on its interactions with sodium dodecyl sulfate (SDS).'0 The conformational changes of pyrene end-labeled PEG during its association with SDS has also been investigated by monitoring the excimer formation between the pyrene end- Il,l2 groups. The applications of fluorescence spectroscopy to the study of polymer- surfactant interactions and polymer association in water have been well presented in recent reviews.'3'” Other aromatic luminescent probes solubilized in aqueous media have also been used as probes for the study of micellar solutions.” Intrinsic excimer fluorescence of p0lystyrene-block-poly(ethylene propylene) has also been used to study the block copolymer micelles and to determine the critical micelle concentration (CMC).'° The determination of the critical micelle concentration using pyrene as a solubilized 143 fluorescent probe for diblock copolymers,”l8 block electrolyte solutions,”20 and triblock copolymers,” has been recently investigated. Investigators have also used other aromatic 22.23 fluorescent probes such as 1,6-Diphenyl-1,3,5-hexatriene (DPH) and dypyme24 for the investigation of micelle formation by the Pluronic (BASF) class of triblock copolymeric surfactants. These investigators have utilized the environment-sensitive fluorescence of the solubilized fluorescent probe to obtain estimates for the CMC and to study micelle formation in the copolymer systems. The CMC of many of the block copolymer systems is too small to be determined by light-scattering techniques.” In such cases a method based on the use of pyrene as a fluorescent probe has been employed to determine the CMC.”“‘9 In the presence of micelles, the pyrene probe partitions between the aqueous and the micellar phases. This partitioning leads to a number of interesting changes in the (0,0) band in both the excitation and the emission spectra of pyrene. There is a red shift in the excitation spectrum, change in the vibrational fine structure of pyrene fluorescence (the ratio Il/I3 decreases) and an increase in the fluorescence decay time (manifested by an increase in intensity), accompanying the transfer of pyrene from the aqueous to the hydrophobic micellar phase. The excitation peak maximum for the broad peak centered at around 332 nm was observed to shift to about 338 nm in hydrophobic environments.”“9 This shift was also reported to be accompanied by an increase in the emission intensity attributed to an increase in the fluorescence lifetime of the excited pyrene. In the case of the emission spectrum of pyrene, the ratio of the peak at ~378 nm (1,) to the peak at ~389 nm (13) is very sensitive to the polarity of the environment.'8"9'2| The Il peak, which arises from the (0,0) transition from the lowest excited electronic state, is a “symmetry forbidden” 144 transition that can be enhanced by the distortion of the n-electron cloud. Whereas, the electronic transition that results in the I3 peak is not forbidden and thus is relatively solvent-insensitive.‘9 Thus, the ratio (I,/I3) can serve as a measure of the polarity of the environment experienced by pyrene. The CMC values estimated using this ratio have been reported3"8"9'21 to be in agreement with those obtained from conventional techniques such as surface tension measurements. In this chapter, we will utilize the environment-sensitive fluorescence of solubilized pyrene to study the aggregate formation behavior of block/graft copolymers of poly(methacrylic acid-g-ethylene glycol) as a function of pH and concentration. The ability of the copolymers to emulsify oil under acidic conditions points to the formation of hydrophobic domains. Steady state fluorescence measurements of the excitation and emission spectra of solubilized pyrene, as it partitions between the hydrophobic domains and the bulk aqueous phase, will be analyzed. In addition, possible methods to determine the critical aggregate concentration will be presented. 8.2. Experimental Methods and Techniques The synthesis of reversible block/graft copolymeric emulsifiers of poly(methacrylic acid) (PMAA) with poly(ethylene glycol) (PEG) grafts (referred to as PMAA-g-PEGIOOO) was discussed in detail in the previous chapter. Two PMAA-g- PEG1000 copolymers investigated in this study contained MAA to EG ratios of 10:1 and 20:1. Fluorescence spectroscopy was performed on 00polymer samples at pH values 145 ranging from <2 to 7 and concentrations ranging from 0.0 wt.% (no polymer) to 0.5 wt.%. The studies were performed for two sets of copolymeric emulsifiers, those containing 10:1 and 20:1 ratio of methacrylic acid to ethylene glycol repeat units. Polymer solutions were prepared by dissolving known amount of the dry polymer into HPLC grade water. The solutions were filtered by a coarse #4 filter paper (Whatman) in order to remove any suspended impurities. Typically a 100 ml sample with a polymer concentration of 0.5 wt.% was prepared and successive dilutions were made from this stock solution. Pyrene was used as the polarity sensitive fluorescent probe for all the experiments. The solubility limit of pyrene in water is reported at a concentration of 6x107 M. To prepare a saturated pyrene solution in water, 0.010 g of pyrene was added to 100 ml of HPLC grade water and stirred for 4 days. The solution was then filtered using a medium #2 filter paper (Whatrnan) to remove any undissolved pyrene. The concentration of this pyrene solution was assumed to be 6x10'7 M, which is the saturation concentration of pyrene in water. For all the studies reported in this chapter, the pyrene concentration was kept constant at a value of 6x10'8 M for each sample. This was to ensure that no excimer formation would occur since the concentration of pyrene was 10% of its saturation concentration. Sample concentrations varied from 0.0 wt.% (no polymer) to 0.5 wt.% and each 10 ml sample contained 6x10'8 M pyrene. The pH of the samples was varied from <2 to about 7 by adding a few microliters of either a 1M HCl or IM NaOH solution. The fluorescence spectroscopy was performed on an Aminco-Bowman Series 2 Luminescence Spectrometer. Both the excitation and emission spectra were recorded for 146 every sample. For the excitation spectra, emission was at 397 nm while the excitation was varied between 290-360 nm. For emission spectra, excitation was at 322 nm and the emission was collected between 370 and 450 nm. The slit width was kept at 0.2 nm and the spectra were collected at intervals of 0.2 nm at a rate of 1 nm/s. The detector voltage was kept constant for the excitation and emission scans over all concentrations. 8.3. Results and Discussion 8. 3. 1. Excitation Spectra Representative excitation spectra of solubilized pyrene for a series of copolymer samples are shown in Figure 8.1. This figure is a plot of the emission intensity (arbitrary units) at 397 nm as the excitation intensity is varied from 290 nm to 360 nm. The plot consists of four spectra representing aqueous solutions of a 20:1 (methacrylic acid (MAA) to ethylene glycol (EG) molar repeat unit ratio) copolymer with concentrations ranging from 0.005 wt.% to 0.1 wt.%, each at pH 2. This figure illustrates the red shift associated with the partitioning of pyrene into hydrophobic domains. In addition, as the copolymer concentration increases from 0.005 wt.% to 0.1 wt.% there is an increase in emission intensity seen in the excitation spectra. The excitation peak maximum shifts from 337 nm (corresponding to pyrene in bulk aqueous phase) to 342 nm (representing pyrene in a hydrophobic phase). As discussed in the previous chapter, PMAA-g—PEGIOOO block/graft copolymers form a pseudo-block copolymer structure under acidic conditions. This occurs possibly due to the formation of hydrophobic segments arising from hydrogen bonded MAA-EG 147 1.8 . 337 342 1.6 -_ Polymer concentration -0.005 wt. % (A) 1.4-— —0.01 wt. % (B) —0.05 wt. % (C) 1.2-- —0.1 wt. % (D) D a 1 ‘" 2 0 3.- o . E C _ 0.6-- o 4 "3 0.2-- A H, - A“... ., —” 0 : 290 300 310 320 330 340 350 360 Wavelength (nm) Figure 8.1. Pyrene excitation spectra for 20:1 PMAA-g-PEGIOOO copolymer solutions of various concentrations with constant pyrene concentration (6 x 10‘7 M). Emission was monitored at 397 nm. 148 units along the copolymer backbone (see Figures 7.2 and 7.3). Since the ratio of MAA to EG repeat units is 20:1, under complex promoting conditions this may result in the formation of alternating hydrophobic and hydrophilic segments within the same chain. One or more chains may then associate to form aggregates with hydrophobic cores in the interior with the hydrophilic MAA groups providing solubility in the bulk aqueous phase (Figure 7.8). Most researchers have used the 11/13 ratio in pyrene emission spectra to study micellization and estimate the CMC of block copolymer solutions.3"°'2‘ However, the excitation spectrum of pyrene may also be used to study aggregate formation and to estimate the critical aggregate concentration (CAC). Figure 8.2 illustrates the pH dependence of the ratio of the peak at 337 nm to that at 342 nm (1337/1342). The data presented in this figure is for a 20:1 MAzEG copolymer for concentrations ranging from 0.005 wt.% to 0.1 wt.%. The data are represented by the symbols while the lines drawn through them illustrate trends. There are two important observations that can be made from this figure. First, as the pH is raised from 2 to 7 the excitation ratio 1337/1342, decreases and approaches a value of ~0.4 at pH 7. Second, this decrease is more pronounced as the concentration of the copolymer is increased. In addition, under acidic conditions the excitation ratio 1337/1342 increases from about 0.6 for a concentration. of 0.005 wt.% to about 1.8 when the copolymer concentration is raised to 0.1 wt.%. It was also observed that for a pH greater than 7 the ratio remained constant at about 0.4. This was also the value observed from the excitation spectrum of pyrene in water without any copolymer present. 149 2.0 1 8 2 . Polymer concentration ' ~ . 0.005 wt. % ’7‘ 1-6 ‘ A l 0.01 wt. °/o 5 . ‘ A 0.05 wt. % : 1-4 ‘ o 0.1 wt. % N . :3 1.2 - .9 l E 1.0 - c 0.8 - I A .9 . E 0.6 - o--———-ILN\\ '6 , \\\ g 0.4 4 :MM 0.2 - 0.0 . , , I . I . pH Figure 8.2. Plot of the pyrene excitation ratio (1342/1337) as a function of pH for 20:] PMAA-g-PEGIOOO copolymer solutions with 6 x 10'7 M pyrene. The emission was monitored at 397 nm. 150 This data may be used to plot the percent pyrene “emulsified” as a function of pH by considering a simple two-phase model for the partitioning of pyrene in the hydrophobic and hydrophilic domains. In our studies, pyrene may be considered as the “oil” phase due to its hydrophobic nature. If an excitation ratio of 0.4 is considered to represent pyrene in the aqueous phase and a value to 2.0 represents pyrene in the hydrophobic phase (based upon experiments with a polymer concentration of 0.5 wt.%), then a simple linear relationship may be used to calculate the percent pyrene “emulsified” by the copolymer. Figure 8.3 is plot of the percent pyrene emulsified calculated in the above manner versus pH for the data presented in Figure 8.2. This figure demonstrates aggregate formation exhibited by the 20:1 copolymer surfactant under acidic conditions. An increase in the polymer concentration results in the formation of an increased number of polymeric aggregates under acidic conditions. This may be concluded from the increase in the excitation ratio resulting from pyrene molecules experiencing a hydrophobic environment. This figure also illustrates the surface inactive behavior of the block/graft copolymer under neutral and basic pH due to the decomplexation of the PEG graft from the PMAA backbone. 8. 3. 2. Estimation of the Critical Aggregate Concentration fiom the Excitation Spectra The PMAA-g-PEGIOOO copolymers employed in this study have shown evidence of aggregate formation which was first demonstrated by oil emulsification studies discussed in the previous chapter. While most ionic and di-and tri-block copolymeric emulsifiers form micelles above the critical micelle concentration, the PMAA-g- PEG1000 copolymers may not form micelles in the strict sense. Due to the proposed 151 100 90 o Polymer concentration ‘ 0 0.005 wt. % 5c: 807, ‘ I 0.01 wt. % 93 A A 0.05 wt. % 5': 7O - g . e 0.1 wt. % E 60 - I5” .. g 50 1 . S 40 ‘ I 0. I *5 30 - 8 20 - d.’ ‘ I “We “\ 10 - \\-\\ 0 I I . l ' r T l . TS“ I . I ' 0 1 2 3 4 5 6 7 8 pH Figure 8.3. Plot of percent pyrene emulsified versus pH for a 20:1 PMAA-g-PEGIOOO copolymer solutions calculated from excitation ratio data. 152 pseudo-multiblock architecture under acidic conditions, aggregates may be formed by one or more chains depending on the length of the polymeric chain and the size of the aggregate. In this study, a critical aggregate concentration will be identified for the PMAA—g—PEGIOOO copolymers. The CAC may be defined as the concentration at which the first copolymer aggregate is formed analogous to the definition of the CMC. While most researchers have used the pyrene emission spectrum to estimate the CMC of copolymeric surfactants, a few investigators have used the pyrene excitation spectrum to estimate the CMC of block copolymer surfactants.‘8"9 In general, a ratio of peaks in the excitation spectrum has been employed. For example, the peak ratio 1333/1333 has been employed to estimate the CMCs of polystyrene-b-poly(ethylene oxide) diblock copolymers'g"9 In this study, the peak ratio 1337/1342 has been employed to estimate the CMC of the reversible block/graft emulsifiers. Figure 8.4 demonstrates the effect of copolymer concentration on the excitation ratio (1337/1342) at a pH of 2 for the PMAA-g- PEGlOOO copolymeric emulsifiers. This figure illustrates the trends exhibited by two different copolymers with MAA to EG ratios of 10:1 and 20:1. The data are represented by the symbols whereas the lines drawn through them illustrate trends. The sigmoidal nature of the trends in this figure is similar to plots of properties such as surface tension plotted as a function of surfactant concentration.'7'22'25‘26 For example, most surfactants exhibit a decrease in surface tension with an increase in the aqueous concentration. However, a leveling-off is observed after a certain concentration and the CMC may be estimated from the intersection of tangents drawn along the two slopes.‘7'22'25'26 A similar procedure can be adopted to estimate the critical aggregate concentration (CAC) from this data for the two copolymeric emulsifiers. From Figure 8.4 the estimated values for the 153 I 1021 2'0‘ . 20:1 7,, 1.5- .36 .9 g 1.0- C .2 4 iii 2 0.5— LlJ I: I I l I l i 0.0 IIIHII I I IIIIIII r Trnlnl I r IIIIII' I rrrnrfl r rrrrnr] n rrrrrul 1E—6 1E-5 1E-4 1E-3 0.01 0.1 1 Polymer Concentration (wt. %) Figure 8.4. A plot of the pyrene excitation ratio as a function of copolymer concentration for 10:1 and 20:1 PMAA-g-PEGIOOO copolymers at pH 2. 154 CAC for the 10:1 and the 20:1 copolymeric emulsifiers were 8 x 10‘4 wt.% and 2 x 10'3 wt.% respectively. For di-block and tri-block copolymeric surfactants, the CMCs reported in literature range from 4 x 10“1 wt.% to 6 x 10'4 wt.% for poly(ethylene oxide)- polystyrene tri- and diblock copolymers respectively.” The values obtained from the excitation spectrum for the PMAA-g—PEGIOOO block/graft copolymers under acidic conditions are of the same order of magnitude as those reported in literature for diblock and triblock copolymeric surfactants. 8. 3. 3. Emission Spectra The steady state pyrene emission spectra for a 20:1 PMAA-g-PEGIOOO copolymeric emulsifier at a pH of 2 is shown in Figure 8.5. The figure contains spectra for copolymer solutions from 0.005 wt.% to 0.5 wt.% with the same pyrene concentration (6 x 10‘7 M). The two trends that can be readily observed are an increase in overall emission intensity as the copolymer concentration increases, and the change in the peak intensity ratio (1373/1339) commonly referred to as the II/I3 ratio. As in the case of the excitation spectrum, these trends may be explained by the partitioning of the pyrene between the aqueous and hydrophobic domains created by the copolymer aggregates under acidic conditions. The intensity increase may be attributed to an increase in the lifetime of the excited state pyrene, while the increase in the ratio Il/I3 arises from forbidden vibronic band transitions which exhibit marked intensity enhancement under the influence of solvent polarity. Figure 8.6 illustrates the dependence of the Il/I3 ratio (ratio of peak at 378 nm and 389 nm respectively) for a 20:1 copolymer as a function of 155 1 378 389 0,9 -_ Polymer concentration -- 0.005 wt. “/0 (A) 0.8 -— —0.01 wt. % (B) 0.7 .L . — 0.05 wt. % (C) D — 0.1 wt. % (o) 0.6 -«-— 3. -_ . .5.) 0.5 I C c ,9 0.4 ar- , E . 0.3 +- ' A 0.2 -— - 0 1 -.~ “N .K 0 I i i ‘ i ‘ i i l I I 370 380 390 400 410 420 430 440 450 Wavelength (nm) Figure 8.5. Pyrene emission spectra for 20:1 PMAA-g-PEGIOOO copolymer solutions at pH 2. Excitation was at 322 nm and pyrene concentration in each sample was 6 x 10'7 M. 156 1.8- A 1.6-1 g :0, 1.4- I?) . E 1.2- E c 1.0— . .9 Polymer concentration <0 o -‘-’-’ 0.8- . 0.005wt.% UEJ . - 0.01wt.% 06— A 0.05 wt.% ' 0 0.1wt.% 0.4 I I I I r I I 0 2 4 6 8 Figure 8.6. A plot of the pyrene emission ratio 1373/1339 (I,/I3) as a function of pH for 20:1 copolymer solutions with a pyrene concentration of 6 x 10'7 M. Excitation was at 322 nm. 157 pH. The figure contains data for copolymer solutions with concentrations ranging from 0.005 wt.% to 0.1 wt.%. The data are represented by the symbols and the lines drawn through them illustrate trends. From this figure it can be noted that, under acidic conditions, the emission ratio (II/I3) decreases as the copolymer concentration is increased. From experiments performed on solutions without any copolymer, the II/I3 ratio for pyrene was found to be ~1.8 and independent of pH. In addition, the pyrene emission spectrum for a solution containing 0.5 wt.% of 20:1 copolymeric emulsifier yielded a Il/I3 ratio of ~0.8. Values for the emission ratio I,/I3 for pyrene in water reported in literature range from 1.58 to 1.9.3'17’” The I,/I3 values for pyrene in a hydrophobic environment range from 0.55-0.6 in various hydrocarbon solvents,3 0.95 for polystyrene films,l7 1.15-1.20 in polystyrene-b—poly(ethylene oxide) solutions,‘8 and between 1.05-1.12 in polystyrene-b-poly(sodium acrylate) solutions.19 By considering a Il/I3 value of 1.8 to represent pyrene in the aqueous phase and a value of 0.8 to represent pyrene in a predominantly hydrophobic phase, a plot of percent pyrene emulsified as a function of pH may be constructed in a manner similar to that described in the pyrene excitation spectrum studies. 8. 3. 4. Estimation of the CA C from Emission Spectra A plot of the emission ratio (II/I3) as a function of concentration at a pH of 2 is presented in Figure 8.7. The figure contains data for two PMAA-g-PEGIOOO copolymers with MAA to E6 repeat units ratios of 10:1 and 20:1. The data is represented by the symbols and lines drawn through them represent trends. Both the copolymers exhibit similar sigmoidal trends and the technique used to estimate the CAC from excitation data 158 1.6- . _, I \\ _ “a _<:> 1.4- }; 1.2~ 76 (I ‘5 .2! E LlJ 0.8- I ‘ I l l I I lWIFl 1 r1 rlllll fiT 1 Illlll l l l llllll I l T llllll l l l lllll 1E-6 1E-5 1E-4 1E-3 0.01 0.1 1 Polymer Concentration (wt. %) Figure 8.7. A plot of the pyrene emission ratio (excitation at 322 nm) versus concentration for 10:1 and 20:1 PMAA-g-PEGIOOO copolymers. 159 can be similarly applied to determine the CAC from the data in Figure 8.7. This may be achieved by drawing tangents to the two faces of the sigmoidal curve as shown in the figure and determining the CAC as the concentration corresponding to the point where the tangents intersect. Using this method the CAC for the 10:1 and 20:1 copolymeric emulsifiers was found to be 8 x 10“ wt.% and 2 x 10'3 wt.% respectively. These values are in perfect agreement with those obtained from the excitation spectra for the copolymers. The differences in aggregate formation and the estimated CAC between the two copolymeric emulsifiers are clearly evident in Figures 8.4 and 8.7. The 10:1 PMAA-g- PEGIOOO copolymer contains 10 times the number of hydrophobic sites under complexed conditions as compared to the 20:1 copolymer. It is therefore correct to presume that on an average the 10:1 copolymer will contain twice the number of hydrophobic segments per chain as compared to the 20:1 copolymer. The lower critical aggregate concentration exhibited by the 10:1 copolymer may be attributed to the higher number of hydrophobic units per chain formed under complex promoting conditions. This may facilitate the formation of hydrophobic aggregates at a lower concentration than the 20:1 copolymer. However, a ratio lower than 10:1 and closer to 1:] may not exhibit a lower CAC since the excess of hydrophilic methacrylic acid units is essential for the solubilization of the aggregates into the bulk aqueous phase. 160 8.4. Conclusions In this chapter, the aggregate formation behavior of reversible block/graft copolymers has been presented. Fluorescence spectroscopy has been used to study the formation of hydrophobic aggregates. The polarity sensitive fluorescence of solubilized pyrene was used to study the aggregate formation as a function of pH and concentration. Both the excitation and emission spectra of pyrene have been used to estimate the critical aggregate concentration for 10:1 and 20:1 copolymers. Specifically, the 1337/1342 ratio in the excitation spectrum and the 1373/1339 ratio in the emission spectrum were plotted as a function of pH for various copolymer concentrations. As the copolymer concentration was increased, the excitation and emission ratios indicated increased partitioning of the pyrene into hydrophobic domains. This may be interpreted as direct evidence of the increase in the number of aggregates formed as a result of an increase in copolymer concentration. Further, the critical aggregate concentration for the 10:1 and 20:1 copolymers was estimated from both the excitation and emission spectra to be 8 x 10“ wt.% and 2 x 10'3 wt.% respectively. 10. 11. 12. 13. 14. 15. 161 8.5. References Nakajima A., J. Mol. Spec, 61, 467-469 (1976). Ham J. S., J. Chem. Phys., 21, 756-758 (1953). Kalyanasundaram K. and Thomas J. K., J. Am. Chem. Soc., 99, 2039 (1977). Turro N. J., Gratzel M., and Braun A. M., Angew. Chem. Int. Ed. Engl., 19, 675- 696 (1980). Quina F., Abuin E. and Lissi E., Macromolecules, 23, 5173-5175 (1990). Prochazka K., Kiserow D., Ramireddy C., Tuzar Z., Munk P., and Weber S. E., Macromolecules, 25, 454-460 (1992). Kiserow D., Chan J ., Ramireddy C., Munk P., and Weber S. E., Macromolecules, 25, 5338-5344 (1992). Bakeev K. N., Ponomarenko E. A., Shishkanova T. V., Tirrell D. A., Zezin A. B., and Kabanov V. A., Macromolecules, 28, 2886-2892 (1995). Kramer M. C., Steger J. R., and McCormick C. L., in Multidimensional Spectroscopy of Polymers: Vibrational, NMR and Fluorescence Techniques, M. W. Urban and T. Provder Eds., ACS Symposium Series, 598, Washington DC, pp 379 (1995). Maltesh C. and Somasundaran P., Langmuir, 8, 1926-1930 (1992). Maltesh C., Somasundaran P., and Ramachandran R., J. Appl. Polym. Sci. .' Appl. Polym. Sym., 45, 329-338 ( 1990). Hu Y.-Z., Zhao C.-L., and Winnik M. A., Langmuir, 6, 880-883 (1990). Winnik F. M., in Interactions of Surfactants with Polymers and Proteins, E. D. Goddard and K. P. Anantpadmanabhan Eds., CRC Press, Ann Arbor, pp 367 (1992) Winnik M. A. and Winnik F. M., in Structure Property Relations in Polymers: Spectroscopy and Performance, M. W. Urban and C. D. Craver Eds., ACS Advances in Chemistry Series, Washington DC, 236, pp 485 (1993). Almgren M., Grieser F., and Thomas J. K., J. Am. Chem. Soc., 101(2), 279-291 (1979) 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 162 Yeung A. S. and Frank C. W., Polymer, 31, 2101-2111 (1990). Zhao C.-L. and Winnik M. A., Langmuir, 6, 514-516 (1990). Wilhelm M., Zhao C.-L., Wang Y., Xu R., and Winnik M. A., Macromolecules, 24, 1033-1040 (1991). Astafieva I., Zhong X. F., and Eisenberg A., Macromolecules, 26, 7339-7352 (1993). Astafieva 1., Khougaz K., and Eisenberg A., Macromolecules, 28, 7127-7134 (1995) Nivaggioli T., Alexandridis P., and Hatton T. A., Langmuir, 11, 730-737 (1995). Alexandridis P., Athanassiou V., Fukuda S., and Hatton T. A., Langmuir, 10, 2604-2612 (1994). Alexandridis P., Holzwarth J. F ., and Hatton T. A., Macromolecules, 27, 2414- 2424 (1994). Nivaggioli T., Tsao B., Alexandridis P., and Hatton T. A., Langmuir, 11, 119-126 (1995) Rosen M. J ., Surfactants and Interfacial Phenomenon, Second Ed., John Wiley & Sons, New York, 1989. Clint J. H., Surfactant Aggregation, Chapman Hall, New York, 1992. CHAPTER 9 CONCLUSIONS In this research, two polymeric materials have been designed based upon the complexation properties of poly(ethylene glycol), and their properties have been investigated. These materials have several potential applications and may be employed commercially to solve current engineering problems. During the course of this research, several fundamental issues have also been addressed in order to understand phenomena at the molecular level. Molecular modeling techniques have been employed for purposes of visualization and the study of polymeric conformations. These studies have provided validation of proposed hypotheses and have been used to guide the synthesis efforts. Finally, this research has accomplished its objectives of developing strategies for the synthesis and characterization of polymeric pseudocrown ethers and reversible copolymeric emulsifiers based upon polymer complexation. 9.1. Summary of Results 9.1.]. Synthesis and Characterization of Polymeric Pseudocrown Ethers Ion-induced Templatization of Oligomeric Ethylene Glycol Diacrylates. The scheme for the synthesis of polymeric pseudocrown ethers is based upon the ability of oligomeric ethylene glycol diacrylates to assume circular conformations when associated 163 164 with metal cations of appropriate size. Experimental studies demonstrated that certain salts that were not soluble in nonpolar solvents, such as chloroform and methylene chloride, became soluble upon the addition of oligomeric PEG 300. Further evidence of cation binding by oligomeric PEG was obtained by 'H NMR studies of PEG and its complexes with metal salts. Pyrene excimer fluorescence was also employed to investigate the ion-induced cyclization of relatively short-chain, pyrene end-labeled oligo(ethylene glycol) molecules. The excimer fluorescence results demonstrated that, even in the absence of any templating cation, a significant excimer emission was observed, indicating that the locally high concentration of the two pyrene moieties attached to the same oligomer leads to intramolecular cyclization. The addition of templating cations such as Ni” and Sn2+ lead to further enhancements in the excimer fluorescence. These results suggest that the templating ion binds with the oligo(ethylene glycol) chain and facilitates the formation of cyclic conformations which bring the two chain ends into proximity. For pyrene end-labeled penta-ethylene glycol, the excimer fluorescence enhancement induced by nickel ions was more pronounced in chloroform than in tetrahydrofuran. Therefore the pyrene excimer-to-monomer fluorescence ratio was found to provide a sensitive measure of the ability of the oligo(ethylene glycol) mixture to solubilize the ions into an organic solvent. Molecular Modeling Studies. In an effort to optimize the template ion synthesis approach, molecular modeling studies were performed on oligomeric PEGDA containing between two and ten ethylene glycol repeating units, with and without cations. Simulation results indicated that the templating cation significantly decreased the mean end-to-end distance, thereby bringing the unsaturated end-groups into close proximity. 165 Although the presence of the cation decreased the ETED for all PEGDA chain lengths, the PEGDA ligand that resulted in the most effective templatization for Na+ contained four ethylene glycol repeating units. The simulation time had little effect on the results for relatively short PEGDA ligands containing seven or fewer ethylene glycol repeating units. These ligands are highly constrained due to binding interactions with the ion. A higher degree of variation with simulation time was observed for longer ligands in which only a portion of the chain interacts with the ion. These simulation results have affirmed the role of a templating ion in the synthetic scheme. Moreover, the results provide insight into the selection of the templating ion and the PEGDA oligomer that will maximize templatization by bringing the unsaturated end groups into proximity. In addition, molecular dynamics simulations of pyrene-labeled oligo(ethylene glycol) ligands were performed to provide molecular insight into the fluorescence results by elucidating the effects of the templating ion and/or the pyrene end groups on the conformations adopted by the oligo(ethylene glycol) chains. Simulation results indicated that, even in the absence of a templating cation, the pyrene end-labeled oligomers exhibit an enhanced tendency to form cycles compared to the analogous unlabeled ligands. This result is consistent with the fluorescence results in which significant excimer formation was observed even in the absence of any cation. Simulation results further indicated that the addition of a templating cation leads to a further reduction in the mean end-to-end distance for pyrene end labeled tetra-ethylene glycol in a manner consistent with the observed enhanced excimer fluorescence. 166 Ion-binding Studies. The synthesis of a pseudocrown ether hydrogel with hydroxy ethyl methacrylate as the comonomer and nickel as the template cation was discussed. Based upon earlier studies, poly(ethylene glycol)-diacrylate 200 with an average of 4.5 ethylene glycol repeat units, was chosen as the templating oligomer. Hydrogels with hydroxy ethyl methacrylate as the comonomer and PEGDA 200 in the absence of any templating cation were also synthesized as a “control” for comparison purposes. The conditioning, ion-binding properties and regeneration of the PCEs was discussed. The monomer to solvent ratio during synthesis were varied from 20/80 to 50/ 50 and its effect on the ion-binding capacity was investigated. The PCE synthesized with a monomer to solvent ratio of 20/80 was observed to possess the highest ion-binding capacity (0.3 mg Nip/g polymer) which is comparable to capacities of other polymer bound crown ethers reported in the literature. The binding capacity of the PCEs was observed to decrease with an increase in the monomer to solvent ratio during synthesis. This was attributed to a possible decrease in templatization efficiency and inter-oligomer interactions resulting in greater crosslinking at the expense of pendant crown ethers. Also, the effect of the template ion on the ion- binding capacity of the PCB was demonstrated by comparing the binding capacity of the PCB with that exhibited by a p(HEMA) “control” hydrogel synthesized without any templating cation. The PCE exhibited an enhanced ion-binding capacity for nickel. Finally, an efficient regeneration scheme was optimized which resulted in complete regeneration of the PCB. These studies have demonstrated the promise of the template ion induced cyclization scheme for the synthesis of polymeric pseudocrown ethers. With an easy one 167 step-synthesis scheme and comparable binding properties for Ni“, this research could result in the development of an inexpensive resin for the binding of nickel from aqueous waste streams. Further, this scheme could be extended to synthesize PCEs for the binding of other toxic heavy metal ions of interest such as chromium, copper, lead and zinc, among others. 9.1.2. Synthesis and Characterization of Reversible Block/Graft Copolymeric Emulsifiers Based Upon Polymer C omplexation Swelling Properties of Poly(methacrylic acid-g—ethylene glycol) Hydrogels. To further the understanding of reversible complexation between poly(methacrylic acid) and poly(ethylene glycol), the equilibrium swelling of crosslinked PMAA-g-PEG hydrogels was investigated. The swelling studies were performed in increasing aqueous concentrations of the solvents and the equilibrium solvent weight fraction was calculated. Swelling behavior in a homologous series of alcohols (methanol, ethanol and n-propanol) demonstrated the effectiveness of a solvent in disrupting complex and stabilizing the hydrophobic interactions. The ability of a solvent to break the complex (at low solvent concentrations) was related to the ease with which it could compete in the formation of hydrogen bonds with the PMAA backbone. However, at high solvent concentrations, the degree of dissociation of the PMAA may govern the equilibrium degree of swelling. For alcohols with longer alkyl groups (n-propanol) the degree of swelling decreased gradually at higher solvent concentrations. The ratio of monomer to solvent during synthesis affected the maximum equilibrium degree of swelling attained by a hydrogel. Hydrogels synthesized in a 20/80 monomer solvent ratio achieved a higher degree of swelling which 168 may be attributed to a lower degree of crosslinking and entanglement. Comparisons between the swelling behavior of a crosslinked homopolymer of PMAA with a P(MAA- g—EG) gel in methanol and acetone demonstrated the effect of complexation as evidenced by the difference in the equilibrium degree of swelling at low solvent concentrations. These studies also illustrated the effect of dissociation of the PMAA on its swelling behavior at high solvent concentrations. A lowering in the degree of swelling at high concentrations of acetone or n-propanol may be attributed to a lower degree of dissociation of the polyacid (PMAA). Molecular simulations demonstrated the existence of a 1:1 complex stoichiometry between complexing PMAA and PEG segments. Molecular mechanics simulations on PMAA and PEG segments with 4, 8 and 14 repeat units exhibited the formation of stable conformations with a 1:1 complex stoichiometry. Synthesis and Emulsification Properties of Reversible Block/Grafi Copolymers Based Upon Polymer Complexation. The synthesis and emulsification behavior of novel block/graft copolymeric emulsifiers based upon a reversible hydrophobic architecture has been presented. In these systems the graft copolymer assumes a multiblock copolymer configuration under certain conditions of pH. These copolymers exhibit reversible emulsification behavior and may be useful in a variety of applications. Emulsification studies demonstrated that these copolymers can emulsify oil under acidic conditions in a reversible fashion. As the pH of the emulsifier solution is raised towards neutral the percent oil emulsified decreases, and under basic conditions no emulsification is observed. Interfacial tension studies illustrated the surface active nature of the copolymers under acidic conditions. A lowering of the interfacial tension was observed 169 under acidic conditions for the copolymeric emulsifiers under investigation. These properties exhibited by the block/graft copolymers can be attributed to the formation of hydrogen bonded hydrophobic complexes between the methacrylic acid backbone and the ethylene glycol repeat units on the graft. This may result in an alternating hydrophilic- hydrophobic pseudo-multi block structure which may be responsible for the surface active properties of the copolymers. Aggregate Formation by Reversible Copolymeric Block/Graft Emulsifiers. Fluorescence spectroscopy has been used to study the formation of hydrophobic aggregates. The polarity sensitive fluorescence of solubilized pyrene was used to study the aggregate formation as a function of pH and concentration. Both the excitation and emission spectra of pyrene has been used to estimate the critical aggregate concentration for 10:1 and 20:1 copolymers. Specifically, the 1337/1342, ratio in the excitation spectrum and the 1373/1389 ratio in the emission spectrum was plotted as a function of pH for various copolymer concentrations. As the copolymer concentration was increased, the excitation and emission ratios indicated increased partitioning of the pyrene into a hydrophobic domains. This may be interpreted as direct evidence of the increase in the number of aggregates formed as a result of an increase in copolymer concentration. Further, the critical aggregate concentration for the 10:1 and 20:1 copolymers was estimated from both the excitation and emission spectra to be 8 x 10“ wt.% and 2 x 10'3 wt.% respectively. CHAPTER 1 O RECOMMENDATIONS FOR FUTURE WORK Two novel materials have been developed based upon polymer complexation. While a significant contribution has been made in the understanding of fundamental concepts which led to the design of these materials, there is considerable potential for further investigations. Future work that is appropriate may be divided into two broad categories; further investigations into the synthesis and characterization of the materials and theoretical modeling studies for visualization of molecular behavior and conformations. In addition, other potential applications of these materials should be addressed based upon their unique properties. In this chapter, some recommendations for future work will be made for both the polymeric pseudocrown ethers and the reversible block/graft copolymer projects. 10.1. Synthesis and Characterization of Polymeric Pseudocrown Ethers The synthesis and ion-binding properties of polymeric pseudocrown ethers (PCES) have been investigated. The ion-induced templatization of oligomeric ethylene glycol diacrylates has been studied by ion-solubilization, NMR spectroscopy, molecular modeling and pyrene end-labeling studies. In addition, a synthetic scheme has been developed based on a free-radical photo-polymerization for the synthesis of the 170 171 pseudocrown ethers. However, further investigations need to be carried out in each of the above mentioned aspects of this research. 10.1.1. Ion-induced Templatization Studies Studies performed resulted in the selection of nickel and chromium as the two heavy metal candidates for the template ion-induced synthesis of PCES. This selection was based on ion-solubilization studies of salts of nickel and chromium including chlorides and nitrates of nickel(II), chromium (III and VI). Studies performed focused on PEGDA 200 as the oligomer of choice due to theoretical considerations based upon the optimum number of ethylene glycol units that may be required for templatization which resulted in the end-groups being in close proximity. Other salts and cations that should be investigated include lead, copper, tin and other platinum group metals. In addition, metal cations such as technetium and ruthenium may be good candidates in an attempt to apply this methodology to study the binding of radioactive metals such as plutonium and uranium. Experiments to be performed may include the determination of the optimum cation-oligomer combination that results in maximum templatization. The experiments should include ion-solubilization and NMR spectroscopy to determine the extent and nature of cation binding. Also, the effect of the counterion on templatization has not been addressed in this study and needs to be investigated. 10.1.2. Molecular Modeling Molecular simulations may be performed on each of the systems under investigation to visualize the configurations adopted by the oligomer in the presence of a 172 cation. End-to-end distance studies performed using molecular dynamics (MD) simulations will yield valuable information on the optimum oligomer chain length required for templatization. However, when performing MD simulations, the inclusion of solvent would account for oligomer solvent interactions. As a result the simulation time would have to be increased to a number which allows for the equilibration of the system. Simulations time of 200ps+ would be suitable for these studies. However, a final assessment can be made only after examining a plot of energy versus time. The estimation of binding energies is possible from the molecular dynamics simulations. The binding energy per mole may be estimated as the difference between the average total energy for the oligomer and the energy of the oligomer with the cation. These values may be compared to those reported in the literature or with the binding energy values obtained from experimental binding constant data. MD simulations performed have primarily employed sodium, aluminum, ruthenium, technetium and copper. However, other metal cations such as chromium and nickel that were used in the synthesis of PCEs were not available on POLYGRAF for modeling purposes. The new version of POLYGRAF currently available (CERIUS2, Molecular Simulations Inc.) should be used in further simulations since it may include several other metal cations and contains an efficient and faster MD module. 10.1.3. Synthesis and Ion-Binding Studies Based upon the ion-solubilization studies, nickel was chosen as the template cation and PEGDA 200 (with an average 4.5 ethylene glycol repeat units) was chosen as the oligomer that would optimally template the cation. Synthesis and ion-binding studies 173 were performed on PCEs based upon this model system. In addition, the ratio of PEGDA 200 to the comonomer hydroxy ethyl methacrylate (HEMA) was chosen to be 5 wt.%. This was based upon a 20/80 monomer to solvent ratio during synthesis. It may be possible to increase the PEGDA 200 concentration to 10 wt.% or more if a 10/90 monomer to solvent ratio (or lower) is chosen. That may result in PCEs with higher binding capacities. Other possible investigations include the use of other hydrophilic comonomers such as hydroxy ethyl acrylate, carboxylic monomers such as acrylic and methacrylic acid, and other hydrophilic monomers that may be readily polymerized with the templated PEG diacrylates. Requirements for the comonomers include hydrophilicity, strength in the polymerized form, and non-interference with the templatization process. It would be interesting to observe the characteristics of a PCE with an ionogenic backbone comprising acrylic or methacrylic acid. That may lead To thyevelopment of conductive polymers and enhanced cation uptake. Ion-binding studies were performed using atomic absorption (AA) spectroscopy and a difference technique was used in order to determine the ion uptake. While AA spectroscopy is a very sensitive technique, it requires calibration before each run and is sensitive to impurities. Further, since the ion-uptake per gram of polymer being investigated is very small, other techniques such as ICP mass spectroscopy may prove to be more accurate for the determination of cation concentrations in aqueous solutions. The present study investigated the binding properties of PCB synthesized with nickel as the templating cation. The ion-binding studies performed considered only the binding of nickel to the PCE. Further studies should include PCEs synthesized with other 174 cations as template ions based upon the strategy presented in this research. In addition, binding studies should include several cations in order to determine the selectivity of these materials. The hypothesis that the PCE will be selective for the template cation should be tested conclusively. 10.2. Synthesis and Emulsification Properties of Reversible Block/Graft Copolymeric Emulsifiers Based Upon Polymer Complexation Reversible block/graft copolymers were synthesized based upon the ability of poly(ethylene glycol) (PEG) to form hydrophobic complexes with hydrophilic polymers such as poly(methacrylic acid) (PMAA). Graft copolymers with a PMAA backbone and PEG 1000 (containing on an average 23 EG repeat units) grafts were synthesized by a one-step free-radical polymerization. The copolymers behaved as emulsifiers under complex promoting (acidic) conditions while under conditions when the complex was broken, the surface active behavior was also lost. Oil emulsification and interfacial tension measurements demonstrated their surface active behavior. The polarity sensitive fluorescence of pyrene was used to study aggregate formation and to estimate the critical aggregate concentration (CAC). Studies that may be useful in further understanding the emulsification behavior and aggregate formation by these copolymers are addressed below. 10.2.1. Study of Reversible Complexation The fundamental basis behind the reversible nature of these materials is the formation of a hydrophobic complex by two constituent hydrophilic polymers. 175 Complexation of PMAA with PEG has been investigated by several researchers and the variables that affect complexation have also been discussed. However, experimental evidence that points to the formation of 1:1 complexes between MAA and EG repeat units has not been supported conclusively via a molecular picture. Molecular dynamics simulations may be performed on PMAA-g—PEG grafi copolymers in the presence of solvent to study the complex stoichiometry. Molecular Simulations Inc. software (POLARIS or CERIUSZ) may be used to model polymers in solvent. These studies will provide a better understanding of the complex and the effect of solvent and pK, of the acid on complexation. Further, two-dimensional lH or l3C NMR studies will be useful in studying complexation of a PMAA-g-PEG copolymer. By performing 2-D 'H NMR on a sample in water under complex promoting conditions using a suitable pulse sequence and water-suppression, it may be possible to investigate the complex by evaluating the backbone acid proton peaks. 1 0. 2. 2. Copolymeric Emulsifier Synthesis The PMAA-g-PEGIOOO graft copolymer was synthesized by a free-radical solution polymerization. Variables that may be further investigated should include the composition of the solvent, choice and concentration of initiator, MAA to EG repeat unit ratio, and the length of the PEG graft. All of these variables will have an effect on the oil emulsification and aggregate formation behavior of the copolymer. For example, the effect of MAA to EG ratio has been investigated in this study and has been found to impact the CAC of the copolymer solutions under acidic pH. Finally, the synthesis 176 should be attempted under complex promoting conditions to determine the effect on the structure of the graft copolymer that is obtained. 10. 2. 3. Inclusion of Comonomers Studies dealing with the inclusion of comonomers to impart additional hydrophilic or hydrophobic properties to the PMAA-g—PEG copolymer, should be conducted. For example, a copolymer synthesized with a MAAzEG ratio of 1:1 and containing a few mole percent of a hydrophobic oligomer such as lauryl methacrylate would result in a graft copolymer that would be primarily hydrophobic under complex promoting conditions and would result in a graft copolymer with a hydrophilic backbone and both hydrophilic and hydrophobic grafts under basic conditions. Such a copolymer would behave as a graft copolymeric emulsifier under basic conditions and its surface active nature would be lost under acidic or complex promoting conditions. Preliminary studies in our laboratory have demonstrated the viability of such a scheme in which the emulsification behavior of the copolymer would be exactly opposite to that of the PMAA-g-PEG copolymers synthesized with MAA to EG ratios greater than unity. 10. 2. 4. C opolymer Characterization and Emulsification Studies The PMAA-g-PEG copolymers investigated in this research need to be characterized. Determination of the molecular weight distribution of these materials could lead to an understanding of the emulsification behavior of these materials. The universal calibration technique may be used to determine the molecular weight distribution of the graft copolymers by gel permeation chromatography (GPC). In 177 addition, molecular weight determination is also possible by first hydrolyzing the grafts followed by determination of the backbone molecular weight since the MW of the grafts would already be known. Proton NMR spectroscopy along with 13C NMR would be useful in characterizing the extent of grafting and determining the number of grafts per chain in conjunction with the molecular weight distribution. The oil emulsification studies presented in this thesis are adequate in determining qualitatively the emulsification ability of a particular copolymer solution under acidic conditions and for quantitatively estimating the amount of oil emulsified as the pH is raised. However, studies to be performed in the future should include emulsification studies for the determination of: i) the maximum amount of oil emulsified under acidic conditions as a function of copolymer concentration, ii) stability of the emulsion, iii) compositions of the aqueous and emulsion phases, and iv) nature of the emulsion phase under all conditions (whether oil or water continuous). Light scattering studies may be useful in determining the size of the polymeric aggregates under various conditions of pH. Centrifugation studies may be used to study the stability of the emulsions. Another important aspect that needs to be addressed is the surface active behavior of PMAA alone as a “control.” While PEG does not exhibit any appreciable surface activity, PMAA has been known to form hydrophobic aggregates under low pH conditions. This is due to the tight coiled conformations which arise from protonation of the acid. However, when the acid is neutralized (the pH is raised) the PMAA coils expand due to charge-charge repulsions. The contribution of the coiling and uncoiling of the PMAA as a function of pH and its oil emulsification properties are currently being investigated. The formation of hydrophobic aggregates by PMAA may be responsible for 178 its ability to behave as a surface active material. Hence, a thorough investigation needs to be conducted to study the properties exhibited by PMAA alone. 10.3. Alkali-Swellable Thickeners Comprising Poly(methacrylic acid-g-ethylene glycol) Copolymers Based Upon Polymer Complexation Potential new applications for block/graft copolymers comprising PMAA backbones and PEG grafts include alkali-swellable thickeners. Alkali-swellable or alkali soluble thickeners (ASTs) have found wide application for thickening paint, coating, textile, consumer product, and adhesive formulations. In general, these thickeners exhibit low viscosity under acidic conditions, but high viscosity under basic conditions making them easy to manufacture and blend into formulations while providing excellent thickening properties. As described in several reviews and monographs“2 these thickeners are typically produced by emulsion polymerizations of acrylic or methacrylic acid with a hydrophobic monomer such as ethyl acrylate. The acid groups may be positioned on the surface of the beads by semi-batch or multi-stage addition of the hydrophilic monomer in the latter stages of an emulsion polymerization.3'4 In general, these polymers contain more than 40 wt.% of the hydrophobic monomer. For example, Murakami et al.2 used compositions containing 48-51 wt.% of the hydrophobic monomer; Rodriguez and Wolfe5 report systems containing 66 wt.% of the hydrophobic ethyl acrylate monomer or 33 wt.% of the hydrophilic methylmethacrylate monomer, while Duprey6 suggests that at least 30% alkyl methacrylates with one to four carbon atoms must be used. Shay7 recommends compositions with 15 to 50% carboxy monomer, 10% surfactant monomer, and the balance a hydrophobic (ethyl acrylate) monomer. With high 179 amounts of the hydrophilic acidic monomer, it is often necessary to include crosslinking agents to maintain insolubility and high thickening efficiency.2 Water-insolubility in the application described above is provided via hydrophobic comonomers or functional groups (such as aliphatic esters of acrylic or methacrylic acid). Thickeners contain hydrophobic comonomers to provide water insolubility at acidic pH. This class of materials have never been synthesized without hydrophobic comonomers or hydrophobic functional groups. Thickeners which exhibit very low solution viscosities under acidic conditions while exhibiting an order of magnitude increase in viscosity under basic conditions are been currently investigated. These thickeners are graft copolymers of acrylic or methacrylic acid with oligomeric ethylene glycol grafts. Carboxylic acids such as acrylic or methacrylic acid (Lewis acids) form strong hydrophobic hydrogen-bonded complexes with Lewis base functionalities such as the ether oxygens of polyethers on polyethylene glycol (or oxide). The mechanism of these interactions and the nature of the complex are well understood (Scranton et al.“‘9 and references therein). Under complex promoting conditions, the acid is protonated and the grafts align along the backbone due to hydrogen bonding resulting in a coiled conformation as shown in Figure 10.1. However, when the pH is raised the acid groups are neutralized and the complex is broken resulting in a net volume expansion of the coiled polymer due to charge-charge repulsion and mobility of the graft. The class of polymers discussed above can be readily synthesized by a single step free radical polymerization reaction. However, alkali swellable microparticles consisting of a similar system can also be produced by a novel emulsion polymerization reaction of the two hydrophilic monomers. 180 (A) __ Hydrophilic Backbone (B) .................. Hydrophilic Graft Figure 10.1. Conformational changes resulting from going from an acidic (A) to a basic (B) environment causing the polymeric chain to expand. 181 Viscosity studies performed on a Brookfield viscometer for PMAA-g-PEGIOOO solutions for two solution concentrations as a function of pH are shown in Figure 10.2. The data shown in this figure is for a 1:1 MAA to EG copolymer of PMAA-g-PEGIOOO. The symbols represent data while the line drawn through them is to illustrate the trend. These studies demonstrate a large increase in viscosity as a function of pH for both the 0.05 wt.% and the 0.1 wt.% copolymers. This increase is however much more significant for the 0.1 wt.% copolymer in which the viscosity increases from about 1-2 cP under acidic conditions by over two orders of magnitude to ~ 700 cP at around neutral pH. Clearly, great potential exists in the development and further investigations into the synthesis and viscosity studies of PMAA-g-PEGIOOO copolymers with a 1:1 MAA to EG ratio. These copolymers may also be synthesized with other comonomers to impart hydrophobic or hydrophilic character to the thickeners. A host of applications exist for alkali-swellable thickeners and further studies into the synthesis and characterization of these materials should be undertaken. 182 800- —-— 0.05 wt.% 1:1.2 PMAA-g-PEG1000 . —o—O.1wt.% . E 9, 600- E U) 0 8 '5 400‘ 2 (D h: .92 o 200‘ 8 m ______._.———I 0___.=__,__.—qz—L , . , -.—=+—4 2 4 6 8 10 12 pH Figure 10.2. A plot of Brookefield viscosity as a function of pH for 1:1 PMAA-g- PEG1000 (with a MAA to EG ratio of 1:1) copolymer solutions at two concentrations. 183 10.4. References Shay G.D., Advances in Chemistry, 223, 457 (1989). Murakami T., Fernando R.H., Glass J.E., Surface Coating International, 76, 8 (1993). Eisenhart E.K. et al., US. patent 5266646 (1995) Eisenhart E.K. et al., US. patent 5451641 (1995). Rodriguez BE, and Wolfe M.S., Macromolecules, 27, 6642 (1994). Duprey J., US. patent 4351754 (1982). Shay G.D. et al., US. patent 4801671 (1989). Scranton A.B., Klier J., Aronson C.L., in Polyelectrolyte Gels, R. S. Harland and R. K. Prud’homme, Eds., ACS Symposium Series 480, American Chemical Society, Washington, DC, pp 171 (1992). Klier J ., Scranton A.B., and Peppas N.A., Macromolecules 23, 4944 (1990).