LIBRARY Michigan State University PLACE IN RETURN Box to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cJCIRC/DateDuo.pB§—p.15 THE CATALYTIC AND MECHANISTIC PROPERTIES OF ORGANOCLAYS FOR TRIPHASE CATALYSIS by Ton Lee A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1992 ABSTRACT The Catalytic and Mechanistic Properties of Organoclays for Triphase Catalysis by Ton Lee Organoclays are smectites that are ion-exchanged by cationic organic compound such as tetraalkylammonium or -phosphonium ions. For the use in triphase catalysis, the onium ions are usually the cationic surfactants which have one long alkyl chain. Differences in layer charge density result in different orientation of surfactant cations in the clay interlayers. In laponite, the cationic surfactants adopts monolayer structures in which the alkyl chains lie parallel to the silicate surface whereas lateral bilayer structures are formed in hectorite and Wyoming montmorillonite interlayers. As the layer charge density of the clays increased as in organo montmorillonite and F-hectorite, a pseudo- trimolecular structure of the cationic surfactant in Arizona montmorillonite interlayer and a paraffin-like structure in F-hectorite gallery were observed. Organo laponites with a monolayer structure do not stabilize emulsion mixtures of aqueous and organic solutions, causing them to be poor triphase catalysts. Organo hectorite with lateral bilayer structure do stabilize the water-in-oil emulsion and are good triphase catalysts. The mechanism for this organo hectorite triphase catalysis has been elucidated by studying the effect of solvent polarity on reactivity and by determining the dependence of reaction rates on the reactant concentrations. The relative rates of the alkylation for benzyl bromide and naphthoxide also provide mechanistic information. The most plausible mechanism is one in which the organic reactant is first adsorbed at the boundary of the clay surface and organic solution. Subsequent nucleophilic substitution reaction occurs at the aqueous-catalyst interface. Arizona montmorillonite organic derivatives with a pseudo -trimolecular structure and organo F-hectorite with a paraffin-like structure can also stabilize emulsion mixtures of aqueous and organic solution. However, for lipophilic surfactants, such as hexadecyltributyl phosphonium ion in the high layer charge density clay interlayer, the surfactants are sometimes ion-exchanged by metal cations and dissolved into organic solution, causing the organoclays to be poor catalysts for recycling. Surfactant desorption can be reduced by using non-polar organic solvents or a hydrophilic nucleophile in the triphase catalysis system. Hydrophilic surfactants such as hexadecyltrimethyl ammonium in these high layer charge density clay hosts do not desorb after triphase catalysis reaction. The desorption of hydrophilic surfactant occurs after triphase catalysis reaction when a co-surfactant, such as octanol, is present in the reaction system. Organoclays exhibit mechanistic phase transfer properties different from those of polymer supported catalysts. Chemoselectivity and regioselectivity different from that of the typical triphase catalysts can be obtained by using organoclays for the triphase catalysis reaction. To MY PARENTS ACKNOWLEDGMENTS I would like to express my appreciation to Professor T. J. Pinnavaia for his patience, encouragement, guidance and support during the course of the work and the preparation of this dissertation. Gratitude is also extended to Professor H. A. Eick for his carefully correcting the dissertation. I would also like to thank all my colleagues for fruitful discussions and their cooperation. Financial support from National Science Foundation and Department of Chemistry is also acknowledged. Finally, I would like to thank my family for their faith and encouragement which made all of this possible. TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES CHAPTER I INTRODUCTION A. Structure and Properties of Layer Silicates 1 B. Principle of Phase Transfer Catalysis 6 l. Biphase Catalysis 6 2. Triphase Catalysis 13 (a) Nature of triphase catalysis 13 (b) polymer supported catalysts 15 (c) Inorganic-based triphase catalysts -- 20 (d) Poly(ethylene glycol) derivatives as phase transfer catalysts 23 (e) Organoclay as triphase catalysts 25 C. Research Rationale and Objectives 25 1. Nature of Organoclays as Triphase Catalysts 25 (a) The structure of organoclays 25 (b) Colloidal properties of the organoclays - 29 (c) Mechanism of organoclay triphase catalysis 31 2. Longevity of the Clay Derivatives for Phase Transfer Catalysis 34 (a) Desorption of interlayer organic compounds 34 (b) Thermal and chemical instability of the organoclay ------- 37 (c) Co-surfactant influence on surfactant desorption ---------- 39 CHAPTER II EXPERIMENTAL A. Material - 43 1. Sodium Hectorite 43 2. Sodium Montrnorillonite 43 3. Laponite R 44 4. Fluorohectorite (F—Hectorite) 44 5. Rectorite 44 6. Organic Reagents 46 7. Inorganic Reagents 46 8. Synthesis of Organic Precursors 46 Preparation of Organic Clay Derivatives --- 46 1. Organoclays 46 2. Crown Ether Clay Complex 47 Organoclay Triphase Catalysis 47 1. Determination of Cyanation kobs(Equation 5) ==— 47 2. Determination of Cyanation k'obs (Equation 6) when the Amount of Pentyl Bromide Is Much larger Potassium Cyanide 48 3. Solvent Effect on Iodination 48 4. O/C Alkylation of Benzyl Bromide and Sodium Naphthoxide 50 5. Dependence of kobs on the Volume Ratio - 50 6. Catalytic Cyanation in the Absence of Water — 51 L011 ngevities of Organoclay Derivatives 51 l. Recyclability of Organoclays for Cyanation 51 2. Recyclability of Organoclay for Chlorination 51 3. Surfactant Desorption of Organoclays 52 4. Effect of Co-surfactant on Phase Transfer Catalysis Reactions 55 5. Conversion of n-Bromoalkan-l-ol to N-cyanoalkan—l-ol by Using Phase Transfer Catalysts 55 6. Conversion of n-Bromoalkane to N-cyanoalkane by Using Phase Transfer Catalysts 56 7. Recyclability of Crown Ether Clays for Cyanation 56 Application of Organoclay Triphase Catalysis for Organic Synthesis -57 1. Synthesis of Symmetrical Formaldehyde Acetals 9:59!" (Equation 10-12) —— 57 Synthesis of Alkyl Bromide by Dehydration of Alkyl Alcohol in Strongly Acidic Conditions (Equation 13) ---------- 57 Dehydrohalogenation of 2-Bromoethylbenzene (Equation 14) 57 Oxidation of trans-Stilbene (Equation 15) 58 Synthesis of 2,4-Dinitrophenyl Ether (Equation 16) ------------ 58 F. Physical Measurements 59 1. Infrared Spectroscopy — 59 2. Gas Liquid Chromatography 59 3. NMR Spectroscopy 59 4. X-Ray Diffraction 59 5 UV-Vis Spectroscopy 60 6. BET Surface Area Measurement 60 CHAPTER III RESULTS AND DISCUSSION A. The Preparation and Structure of Clay Derivatives - 61 1. The Preparation of Organoclays 61 2. The Structure of Organoclays - 64 3. The Preparation and Structure of Crown Ether Clay Complexes 72 B. The Catalytic Properties of Organoclays = 75 1. Wettability and Catalytic Capacities of Organoclays—---------- 75 2. Dependence of Catalytic Reactivity on the Bulk Reactant Concentration of Liquid Phases 83 3. Dependence of the Organoclay Catalytic reactivity on the Polarity of Organic Solvent 89 4. Dependence of O/C Alkylation on Biphase and Triphase Catalysts 92 5. The Dependence of Organoclay Catalytic Reactivity on the Volume of Two Liquid Phases 95 6. Organoclay Catalytic Reaction in the Absence of Water ------- 98 7. Mechanism of the Organoclay Triphase Catalysis 99 C. The Longevity of Organoclays for Triphase Catalysis 104 1. Recyclability of Organoclays as Triphase Catalysts-----------104 2. Surfactant Desorption 111 (21) Influence of aqueous electrolyte solution on t surfactant desorption of organoclays 111 (b) Dependence of the surfactant desorption on the surfactant structure in the presence of the sodium chloride aqueous solution and toluene 114 (c) Desorption of surfactant from clay hosts in the presence of aqueous sodium bromide solution and toluene -------- 117 (d) The dependence of surfactant desorption on electrolyte concentration 118 (e) Dependence of the surfactant desorption on the organic solvent polarity -- 122 iii 3. Co-Surfactant Effects for the Triphase Catalysis -124 (a) Influence of co-surfactant on [C15H33NMe3+] desorption 124 (b) Influence of co-surfactant on biphase and triphase catalysis 125 (0) Catalytic properties of organoclays and onium ion surfactant for the cyanation of alkyl bromide and . bromoalkyl alcohol 130 4. The Catalytic Properties of Crown Ether Clay Complexes-u 133 D. The Applications of Organoclays in other Triphase Catalysis Reactions —135 1. The Synthesis of Symmetrical Formaldehyde Acetals -------- 135 2. Synthesis of Alkyl Bromide 141 3. Dehydrohalogenation of 2-Bromoethylbenzene 144 4. Oxidation of trans-Stilbene 144 5. Synthesis of 2,4-Dinitrophenyl Ether 146 CHAPTER IV CONCLUSIONS 151 REFERENCES 159 iv LIST OF TABLES TABLE PAGE Surfactant-intercalated clay Catalyst. 62 The BET surface area of sodium clays and the organoclays in which the surfactant, [C16H33NMe3+], is intercalated in the interlayers. 65 The d-spacings of crown ether-clay complexes derived from X-Ray diffraction measurement. 73 The catalytic activities and the colloidal properties of the organoclays or the onium ion salts as phase transfer catalysts.----76 Solvent effect for the iodination reaction of pentyl bromide (Equation 5). 91 The product ratios of the O/C alkylation of sodium naphthoxide with benzyl bromide when organoclays and onium salts were used as the phase transfer catalysts. 93 The catalytic activity of the higher layer charge density C16H33PBu3+organoclay after multiple reaction cycle for cyanation and chlorination of pentyl bromide. — 108 The desorption of onium ion from the organoclays in the presence of 2.5N NaCl aqueous solution. 112 Desorption of surfactant from the organoclays in the presence of NaCl aqueous solution and toluene. 1 15 10. 11. 12. 13 14. 15. 16. 17. 18. 19. 20. 21. Desorption of surfactant from the organoclays in the presence of aqueous 2.5N Hafiz aqueous solution and toluene. 1 19 The desorption of the surfactant under the triphase condition with other organic solvent. 123 Desorption of [C15H33NMe3+] from the organoclays in the presence of octanol as co-surfactant. 126 Influence of octanol and tetradecanol co-surfactants on the cyanation of pentylbromide by byphase catalysts and triphase catalysts. 128 The cyanation of bromoalkanes and bromoalkanols under biphase and triphase catalysis condition. 132 Chemical yields for the cyanation of benzylbromide to benzyl cyanide. 134 The chemical yields of the acetal synthesis (Equation lO-l2).---139 The chemical shifts in 1H NMR of the starting materials and the reaction products of the acetal synthesis reaction (Equation 10-12). 140 The synthesis of octyl bromide from octanol by hydrobromic acid dehydration (Equation 13). 145 Synthesis of styrene from 2-bromoethylbenzene under alkaline solution (Equation 14). — 147 The formation of benzoic acid by the oxidation of trans-stilbene (Equation 15). 148 The synthesis of 2,4-dinitrophenyl ether from l-chloro-2,4- dinitrebenzene and phenolate (Equation 16). 149 LIST OF FIGURES Figure Page 1. The structure of a 2:1 dioctahedral alumina silicate mineral. ------- 3 2. The structure of a 2:1 trioctahedral magnesium silicate mineral.---4 3. The structure of rectorite. -7 4. Mechanism of biphase catalysis with a nucleophilic substitution reaction. 10 5. A general process of triphase catalysis for a nucleophilic substitution reaction. --14 6. Linear and cross-linked structures of polystyrene-supported triphase catalyst. 16 7. The O/C alkylation of benzyl bromide and naphthoxide in protic or aprotic solvents. 18 8. Mechanism of the polymer-supported triphase catalysis. ---------- 19 9. The synthesis of silica gel supported catalysts. 21 10. (a) The structure of polymer-supported poly(ethylene glycol) (b) The function of the PEG atom as oxygen chelating sites for the metal cation. 24 11. The surfactant orientation in the interlayers of various charge vii 12. 13. 14. 15. 16 17. 18. 19. 20 21. density clay. 26 A droplet dispersion of oil-in-water microemulsion. 41 The standard curve of the intensity area ratio of [C5H11CN]/[0.100g C10H22] in GLC analysis versus [C5H11CN] in 6 mL toluene as the organic solvent. 49 dependence of the UV-Vis absorption of C151133NMe3+-methyl orange chloroform solution at 418 nm on concentration of C15H33NMe3Br aqueous solution. ------------ 53 Dependence of UV-Vis absorption of C15H33PBu3+-methyl orange chloroform solution at 418 nm on concentration of C15H33PBu3Br aqueous solution. ------------- 54 The X-ray diffraction pattems of various organoclays: (a) L24AA; (b) H24AA; (c) W24AA; (d) A24AA; (e) F24AA.-- 66 The orientations of crown ethers in hectorite and Arizona montmorillonite interlayers. 74 Classification of the colloidal behavior of 0.100 g of organoclay in liquid mixtures containing 3 mL of 6.25M potassium cyanide aqueous solution, 6 mL of toluene and 2 mole of pentylbromide. 82 (a) Determination of a kobs from the slope of a plot of [C5H11Br]/[C5H11Br]o versus the reaction time in the presence of 0.100 g of C15H33NMe3hectorite. (b) The determination of a kobs' from the slope of a plot of [CN']/[CN']0 against the reaction time in the presence of 0.100 g of C15H33PBu3hectorite (26AA). 85 (a)Dependence of the observed rate constant (kobs) on nucleophile concentration for the cyanation reaction at 90°C in the presence of C15H33PBu3hectorite (H26AA) as the triphase catalyst. (b)Dependence of the observed rate constant (kobs') on the organic electrophile concentration for the reaction at 90°C in the presence of 0.100g C15H33PBu3hectorite as the triphase catalyst. 87 (a) The dependence of kobs on the volume of toluene for the 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. reaction of pentylbromide with KCN. (b) The observed rate plotted against the volume of aqueous solution while the volume of toluene solution was kept constant at 2 mL. 96 The emulsion of aqueous and organic solution stabilized by an amphiphilic organo hectorite. — 100 The mechanism of the triphase catalysis reaction with organo hectorites as the catalysts. 102 The recyclability of C15H33PBu3+hectorite (H26AA) over 10 reaction cycles for the cyanation of pentyl bromide at 90°C.---- 105 The recyclability of C15H33NMe3+hectorite (H24AA) over 10 reaction cycles for the cyanation of pentyl bromide at 90°C.---- 106 The recyclability of C16H33NMe3+F-hectorite (F24AA) over 10 reaction cycles for the cyanation of pentyl bromide at 90°C.---- 109 Dependence of [C15H33PBu3+] desorption from organo F-hectorite (F26AA) on the concentration of sodium chloride in an aqueous-toluene emulsion. 120 The X-ray diffraction pattem of C15H33PBu3+F-hectorite (F24AA) reaction products formed by reaction of aqueous-toluene emulsions at various NaCl concentration. ------ 121 The IR spectra of (a) sodium hectorite, dool=12.1A; (b) 18C6Hect, d001=15.6A; (c) recycled 18C6Hect after one cycle of reaction, dool=l3.2A. 136 The X-ray diffraction pattem of C15H33NMe3+hectorite (H24AA) after the catalysis reaction in which the reaction mixture contains 50% NaOH aqueous solution and 10 mL of methylene solution. 142 The X-ray diffraction pattern of sodium hectorite treated with 50% NaOH aqueous solution. 143 The 1H NMR spectra of (a) 1-chloro-2,4-dinitrobenzene (b) 2,4-dinitrophenyl ether. 150 CHAPTER I INTRODUCTION A. Structure and Properties of Layer Silicates The layer silicates, hectorite and montmorillonite, described in this dissertation are smectite minerals which are a class of naturally occurring minerals. The term 'clay mineral' refers to specific silicates with particle size less than 2 pm and with definite stoichiometry and crystalline structure. Smectites are composed of units made up of two silica tetrahedral sheets and a central octahedral sheet of magnesia or alumina”. The silicate tetrahedra are usually oriented so that the three basal oxygen atoms of each tetrahedron lie on the same plane, while the fourth oxygen atom defines a second common plane. The octahedral sheet contains a cation, usually Al or Mg, surrounded by six oxygens in an octahedral arrangement. The tetrahedral and octahedral sheets are combined so that the tips of the tetrahedrons of each silicate sheet and the oxygens of each octahedral sheet form a common layer. Smectite clays are 2:1 layer minerals which are divided into two structures: dioctahedral aluminum silicate minerals (Figure 1) and trioctahedral magnesium silicate minerals 2 charge arises predominantly from isomorphous substitution in the octahedral layer or from the substitution in the tetrahedral layer. Cations at particular locations of the silicate structure can be replaced by some other cation with similar ionic radius without changing the structure of the minerals. If the replacing cation has lower valence, a net negative charge will be deve10ped. The negative charge is then balanced by the presence of hydrated metal cations in the interlayer region of the structure. These hydrated metal cations are usually located adjacent to the point of anion charge on the basal plane. ' The trioctahedral silicate minerals used in the work are laponite, hectorite and F-hectorite. The idealized unit cell composition for laponite is Lio.36[Lio.36Mgs.64]“(Si8.oo)1V020(0H)4 in which the superscripts (IV) and (VI) refer to the respective cation in the tetrahedral and octahedral sites. The first "Lio,36" in the formula designates the exchangeable hydrated lithium in the interlayer region, and the second one represents the octahedral sheet lithium. Laponite is a synthetic low-charge smectite. The idealized unit cell composition of hectorite is Mo.67[Lio.67Mgs.331(Si8.oo)Ozo(OH. F)4 3 Figure 1 The structure of a 2:1 dioctahedral alumina silicate mineral. Interlayer Region 0 Al O O 031 @011 Figure 2 The structure of a 2:1 trioctahedral magnesium silicate mineral. 1 :5, ; ; «it: i ' Interlayer Region WW ()Mg 00 031 (3301-1 5 in which the "M" designates a hydrated monovalent cation. Fluorohectorite is also a synthetic clay in which the hydroxyl groups are substituted by fluoride. The idealized unit cell composition of hectorite is Li1.6[Li1.6Mg4.4l(Sis.oo)020F4. The dioctahedral silicate minerals used in the dissertation are montrnorillonites from Arizona and Wyoming. The idealized unit cell compositions for both montrnorillonites are M1.161A12.34Mg1.16]Sis.00020(0H)4, and M0.67[A1333M80.67](Si8.00)020(0H)4, respectively Rectorite3 as shown in Figure 3 consists of a regular alternation of a mica-like layer and an expandable layer having a smectite composition. Rectorite also belongs to the family of regularly interstratified clay minerals as well as to the family of smectite clays. Its basal spacing is greater than 19A which is approximately double that of the smectite basal spacing. The formula of rectorite is [Nao.72Ko.02Cao.os)(Cao.24Nao.07)l(Al4.ooMgo.02)[Si6.ssAlr.62)022] The charge on the mineral arises from isomorphous substitution in the tetrahedral particle sheet of the 2:1 smectite clay layer. The cation exchange capacity (CEC) of rectorite is 60 milliequivalents per gram. 6 surface charge density of this material is approximately equal that of a smectite clay in which the CEC is 120 mqu 100g. Smectite clays can be swelled by adsorption of water or some organic solvents4‘7. With multiple layers of solvent, the galleries become liquid-like and accessible for chemical reactions. As the solvent content increases, the basal spacing of smectite clays also increase. The swelling capacity of the clay can be reduced by introducing electrolyte to the solution. The electrolyte reduces of the electric double layer repulsion of two neighbor silicate layers. Generally, higher layer charge density clays are less swellable4 than the lower layer charge clays. Smectite clays and their derivatives have shown catalytic properties for many reactionsg'“. The acidic nature provides the source of the catalytic capacity. Both Lewis and Bronsted acidity have been noted, the former derived from aluminum or iron species located in the crystal. The Bronsted acidity resulys from dissociation of the interlayer water molecule coordinated to polarizing interlayer exchangeable cations”. B. Principles of Phase Transfer Catalysis 1. Biphase Catalysis Chemists frequently encounter the problem of bringing the reactants in two mutually immiscible solutions into proximity to attain rapid reaction rates. The traditional procedure is to Figure 3. The structure of rectorite. Inter layer Region Expandable Interayer O M g Q 0 . K+ . Si(Al) © (XI 9 dissolve the reactants in a homogeneous medium. However, a suitable solvent is not always available and it is usually expensive and difficult to remove after reactions. Solvents such as THF or DMF sometimes present environmental problems in large scale operations. The technique of phase transfer catalysis can permit or accelerate reactions between ionic compounds and organic, water-insoluble substrates in solvents with low polarity. The concept of " phase transfer catalysis" is to transfer the ions, neutral molecules or free radicals from one phase to another where reaction can occur. It is clear that a phase transfer catalyst has considerable advantages over conventional procedures since it eliminates the requirement for expensive anhydrous or aprotic solvents, improves reaction rates, and lowers the reaction temperature13'16. The first type of phase transfer catalyst is the biphase catalyst13’14. Quaternary ammonium or phosphonium halide is typically used for typical biphase catalysis. The amphiphilic character of quatemary onium salts allows them to be wetted in both aqueous solution and organic solvents with low polarity. Stark and Owens13 first developed the use of ternary ammonium and phosphonium salts for phase transfer catalysis in organic synthesis. The mechanism of biphase catalysis was studied by Stark and Owens14 and Brandstonn”. A nucleophilic substitution reaction using onium salts as biphase catalysts13 is represented in Figure 4. An onium cation pairs with an anionic nucleophile Y' in the aqueous phase and the 10 Figure 4 Mechanism of biphase catalysis with a nucleOphilic substitution reaction. - Q+ - Rx(org) 4' Y (aqu) ’RY(org) + X (aqu) RY + Q“X‘ Organic X' + Q”Y‘ ' Y‘ + Q“X‘ Aqueous Q+ = NR4+; PR4+; NaT-Crown Ether 11 ion pair is extracted into the organic phase. Once the anion has been transfered into the organic phase, the anion and organic electrOphile RX undergo nucleophilic substitution and form RY and a new salt, Q+X’. The new salt then returns to the aqueous phase, where Q+ pairs with a new anion, Y’, for the next cycle. If there is one long alkyl chain in the onium ion, the biphase catalyst may also function as a surfactant. The surfactant molecules form small aggregates of 10 to 50 organic molecules dispersed in the aqueous phase. The small aggregations are called micelles, wherein the nonpolar organic parts of the molecules occupy the intemal volume and the highly polar group of the surfactant occupy the outer surface”. For micelle-catalyzed reactions, the positively charged outer surface attracts and concentrates anions from the bulk aqueous solution into the so-called Gouy layer near the surface of the micelle. Reaction then occurs at the Gouy layer19‘24. As comparable volumes of the organic solutions are added to the aqueous micellar solution, an emulsion mixture of aqueous and organic solution stabilized by surfactant is formed. If a co-surfactant, which is typically an alcohol with more than five carbon atoms, is added to the micellar solution, the emulsion is converted to a transparent microemulsion25'27. The microemulsion has a smaller droplet size and a lower surface tension than those of the emulsion27'30. Moreover, a microemulsion is thermodynamically stable, unlike an emulsion which is only kinetically stable25'29. As the microemulsion is utilized into the 12 biphase catalysis, a more efficient catalytic reaction can be obtained due to increased interfacial area between aqueous and organic phases. f Crown or macrocyclic ethers usually containing the basic unit (-0- CH2-CH2)n have also been met with interest for phase transfer catalysis31-35. These are exemplified by l8-crown-6 in which‘18 indicates the number of atoms in the ring and 6 represents the number of oxygens. Other commercially available crown ethers are dibenzo-18- crown-6, dicyclohexano-lS-crown-6 and 15-crown-5. A common feature of all crown and related compounds is a central hole or cavity which can chelate the substance. The cation complex of general interest for phase transfer catalysis are those formed with potassium, sodium cations, hydronium, ammonium, and diazonium. Complex formation between a crown ether and an anionic nucleOphile has two important characteristics. The first is the organic masking of the alkali metal providing an "onium ion" like entity that can be extracted into organic solvents with the accompanying anion. The second is that the anion part of the ion pair in the organic solvent is activated. In a dipolar aprotic solvent, this effect is most notable because the anion-solvent interactions are weak. Such a system has been described as involving the "reaction of naked anions". Since the discovery of phase transfer catalysis, numerous applications have been described, not only in organic chemistry, but also in inorganic chemistry“, analytical applications”, electrochemistry33, 13 photochemistry39 and polymer chemistry“0 have utilized phase transfer catalysis. In organic synthesis , phase transfer catalysis has emerged as a broadly useful tool for: a) nucleophilic substitution reactions b) alkylation and condensation c) reaction of dihalocarbenes and‘other carbenes d) ylide mediated reactions e) oxidations and reductions. In those reactions, the technique of phase transfer catalysis provides a method which avoids the use of a polar aprotic solvent and also improves the reaction rate. 2, Triphase Catalyssis a) Nature of Triphase Catalysis There is a new type of heterogeneous catalysis called " triphase catalysis" , in which the catalysts and each member of a pair of reactants are located in separate phases. Figure 5 illustrates the general features of a simple reaction of a triphase catalysis system“, Generally, the reactivity of triphase catalysis is lower than that of biphase catalysis based on the same equivalent of catalysts being used“. However, triphase catalysis exhibits the advantage of catalyst recovery and convenience of workup after catalytic reactions. Moreover, some triphase catalysts can provide regioselectivity due to the spatial constraint of the catalyst and the reactants. Four major types of triphase catalysts have been developed. Three of these utilize onium ions as the catalytic center. These onium ions are supported on polymers43-45, inorganic matrices47'50, or layer 14 Figure 5 A general process of triphase catalysis for a a nucleophilic substitution reaction. - Catalyst RX + Y (aqu) RY(org) "' X-(aqU) AQUEOUS ORGANIC RY 15 silicates51'54. Another type of the triphase catalysis uses polyethylene glycol as the catalyst32a55'57. b)Polymer supported catalysts Polymer supported catalysts are the first studied and best defined triphase catalysts. The backbone of the catalysts is polystyrene (Figure 6). Some of the phenyl rings are linked to onium salts. Also, the linear polymer can be expanded to a two-dimensional or three-dimensional network by cross linkage of two phenyl groups. In general, the higher the degree of cross linkage, the lower the diffusion of organic substance into the catalysts53»59. Tomoi and Ford53o50, Regen and Reese“, and Montanari et al.,51 have reported extensively on the mechanism and properties of triphase catalysis. The fundamental mechanism of the polystyrene-supported catalysts for phase transfer catalysis is verified by solvent effects45:51’52, O/C alkylation reactions of benzyl bromide and sodium naphthoxide51:53'55, and the dependence of rate on bulk cyanide concentrations55. Also, N MR studies of the swellability of the triphase catalysts in two immiscible solvents“, the rate dependence on cross- linking of the polystyrene backbone53o59, and the rate dependence on percentage of onium salts attached to the phenyl groups provide information on the mechanistic process57. The location of the catalytic reactions in polymer supported catalysts is indicated by the following phenomena. a)The polymer-supported catalysts show higher catalytic capacity in the presence of polar organic solvents. The dependence of reactivity on the polarity of the organic solvent increases in the order decane 16. Figure 6 Linear and cross-linked structures of polystyrene-supported triphase catalyst. Backbone / i // /" 7 ,u Ph 0 C O U U Q X X Ph Ph Ph Ph x = (CH2),PR3"X‘ or (CH2),,NR3*X' 17 < toluene < o-dichlorobenzeneGlfiZ. This suggests that solvation of the ionic pair formed by the onium cation and nucleophile anion in the organic solution dominates the catalytic reaction. b) For the O/C alkylation reaction of naphthoxide by benzyl bromide (Figure 7)“, the O-alkylation product is favored in aprotic solvent such as DMF, toluene etc., and the C-alkylation product is favored in protic solvents such as water, methanol etc. Polymer-supported catalysts afford mainly O- alkylation products ( Figure 7)51,53'65. Therefore, the catalytic reaction for this triphase catalysis occurs mainly in an aprotic, polymer environment. 0) The reactivity of a polymer supported catalyst is not sensitive to the bulk concentration of nucleophile“, indicating that the catalytic reaction does not occur at the aqueous interface of polymer supported catalysts. Therefore, the nucleophilic substitution reactions occur at the organic phase, but the anion exchange at the onium cation site for the next cycle may occur at the aqueous interface. The properties of polymer-based triphase catalysts suggest the fundamental kinetic processes6o shown in Figure 8. The first step is the mass transfer of electrophilic reactant, RX, from bulk solution to the catalytic surface. The second step is diffusion of the reactant through the polymer matrix to the active nucleophilic site. Nucleophilic substitution occurs at the active site in the third step, followed by the diffusion of products to the surface of the catalyst and mass transfer of the products to the bulk solution as the fourth step. The ion pair may go back to the aqueous solution to exchange a new nucleophile for the next reaction cycle. 18 Figure 7 The O/C alkylation of benzyl bromide and naphthoxide in protic or aprotic solvents. m. In P ti Sol t In Aprotic Solvent Evita) ven (DMF) Intermediate Intermediate (Energetically Stable) (Stabilized by H-Bond Formation with Solvent) CHzPh OCHZPh on O-Alkylation (100%) C-Alkylation (89%) 19 Figure 8 Mechanism of the polymer-supported triphase catalysis. (3) 4:. a “.2 41. 2° ‘82 O . l “’ 1(4) piii‘é‘“ §§_P+Buay’ g :Rx: EE—P+BU3X. g :RY: <—— + A ueous E—P’BuaY' + x _. BuaX' +Y 331111 My Phase 20 c) Inorganic-based triphase catalysts. The synthesis of inorganic-based triphase catalysts involves the anchoring of an alkyltriakoxysilane (either a haloalkyl or aminoalkyltrialkyoxy silane) to the surface of silica gel or alumina to afford an alkyl functionalized inorganic matrix (Figure 9)“. Several types of onium salts can be immobilized by means of various hydrocarbon chain length containing different functionalization57'48. Important considerations for effective catalyst design are (i) the porosity and type of the inorganic support; (ii) the length of spacing of the alkyl chain; (iii) the chemical structure of the onium salt. Silica gel with a titer of 0.5 mmole Cl‘lg, is designated as 6OSiC3PBu+Cl'(0.50), where 60Si indicates a silica gel support with a mean pore diameter of 60A, C3 indicates the length of the alkyl chain supporting the phosphonium salts, and the value in parenthesis represents the loading in millirnoles of the onium counterion per gram of catalytic support. The main difference between inorganic solid and polymer supported catalysts is the influence of solvent polarity on the triphase catalysis reaction49'69. In the case of the silica gel triphase catalysis, the activity increases with decreasing solvent polarity. On the contrary, polystyrene supported catalysts show higher activity in more polar solvents. The catalytic reactions are carried out in different microenvironments. The functionalized inorganic matrices are wettable both by the aqueous and the organic solution. This property allows these 21 Figure 9 The synthesis of silica gel supported catalysts. OH HCI 0” .. OH Reflux OH 311103 Gel ' Active Silica Gel Br(CH2)nSi(OEt)3 O PBU3 \ - CH PBu Br 0 0734 2).. a «— o>s;(CH2),,Br O o 22 systems to be active in liquid-liquid phase transfer catalysis even without stirring59, since the inorganic matrix acts as a supply and a pump for the aqueous phase which contains the nucleophile anions. As the catalytic reactions proceed very close to the catalyst surface, the exhausted onium salt can be regenerated easily by means of exchange with the aqueOus ’ solution. The large surface of the support promotes the anion exchange process70, The exact structure of these systems during catalytic reactions remains unknown and not all the questions concerning the deposition of the aqueous and organic phases into the functionalized pores have been answered. The adsorption of the organic substrate on the catalyst pores or surface may play an important role in the mechanism. However, the diffusion of the organic reactants onto the catalyst surface or pore does not play an essential role for inorganic material-supported phase transfer catalysis. Various applications concerning the use of silica gel or alumina based triphase catalysts in nucleophilic substitution reactions have been reported48’49: halogen exchange, synthesis of phenyl ethers and sulfides, reduction of carbonyl compounds with aqueous sodium borohydride, synthesis of nitrile, thiocyanate, nucleophilic, and N-alkylphthalimides. Depending on the reaction conditions, however, the functionalized inorganic matrices may lose some of their functionalization under the strong alkaline solution69, and then reactions can actually be catalyzed both by the phase transfer catalyst and the silica gel or alumina-based triphase catalysts as well as the hydrolyzed catalytic groups free in solution. 23 d). Poly(ethylene glycol)derivatives as phase transfer catalysts The polyethers, specifically poly(ethylene glycol)'s (PEG), are widely used in phase transfer catalysis. The function of this reagent in phase transfer catalysis reactions is similar to that of crown ether. PEG's openly chelate the metal cation (Figure 10)"'2 and carry the nucleophile to the organic phase to undergo the nucleophilic substitution reaction. Therefore, PEG may be used in most applications where crown ethers are currently used. While it is true that the reactivity of PEG's is often less than that of crown ethers, this may be compensated for by increasing the concentration of PEG. Many applications of organic reactions using PEG's as catalysts have been reviewed by Totten and Clinton”: the Biltz synthesis of phenyltoin; the synthesis of triaryl phosphate; the N-alkylation of nitrogen heterocycles; alkyloxymethyl and formation of alkoxypolyethoxymethyl derivatives of acylanilines; dehydrohalogenation of 2-bromoethylbenzene by hydroxide; carbonyl reduction; synthesis of phenyl ethers and sulfides; Williamson ether synthesis; polyether synthesis; aryldiazonium salt reactions; and oxidation chemistry. However, there is no report about the recyclability of these catalysts. 24 Figure 10 (a) The structure of polymer-supported poly(ethylene glycol) (b) The function of the PEG atoms as oxygen chelating sites for the metal cation. (a) §_._ CH20(CHZCH20),,H O-.~~ O R "K+ E o """ 9 ~~~~ OH---OH- \_/ L/ 25 e) Organoclay as triphase catalyst Organoclays, which are the smectite clays intercalated by ternary ammonium or other organic cations, have been used as triphase catalysts. Kadahodayan and Pinnavaia used clay derivatives with metal complexes such as M(phen)32+/XO42' ion pairs in the interlayer to facilitate the liquid-liquid phase transfer reactions51. Comelius, Laszlo and P. Pennetreau utilized montmorillonite derivatives as triphase catalysts for alkylation of dihalomethane54. More recently, Choudary, et al. used the acidified clays linked with onium salts to accelerate cyanation reactions52. Lin, Lee and Pinnavaia contributed extensively to understanding of the nature and the application of the phase transfer catalysis by using organoclays as the triphase catalysts53. This work is continued by focusing on the mechanism, colloidal properties and longevity of organoclays as triphase catalysts. C. Research Rationale and Objectives 1. Nature of Organoclays as Triphase Catalysts a) The structure of organoclays ' The orientation of the cations in smectite clay interlayers depends on the onium structure and the layer charge density of the layered silicate71' 73. Generally, a good organoclay for triphase catalysis should have an onium cation with more than one long alkyl chain. This onium ion can also be used as a surfactant. Lagaly summarized the surfactant orientations in the galleries of various layer charge density clays (Figure 11)“. For layered silicates with an extremely low layer charge density, 26 Figure 11 The surfactant orientation in the interlayers of various layer charge density clay. (a) Monolayer: [C15H33NMe3]+ Laponite (Cation exchange capacity, 55meq/ 100g) I WW1 T i 14.5A (a) Lateral Bilayer, [C16H33NMe3]+Hectorite (Cation exchange capacity, 73meq/100g) 27 Figure 11(Continued) (c) Pseudo Trimolecular Structure: [C15H33NMe3]+ Montrnorillonite(Arizona) (Cation exchange capacity, ll8meq/100g) (d) Paraffin Structure, [C16H33NMe3]+ F-hectorite (Cation exchange capacity, l40meq/ 100g) 28 Figure 1 1(Continued) (e) Lipid Structure, [(C12H25)2NM62]+ F'hCCtorite (Cation exchange capacity, 140meq/ 100g) (t) [(C12H25)2NMe2+]Laponite (Cation exchange capacity, 55meq/ 100g) \ \ NM MM ‘. WW3 ~~ M 17.5A 29 the surfactant adopts a monolayer structure with the chain parallel to the silicate surface (Figure 11a). As the layer charge density of the layer silicate increases, the ion will form a two layer structure with the chain parallel to the silicate layer, the so-called bilayer or lateral bilayer structure (Figure 11b). When the layer charge density of the layered silicate increases further, a pseudo trimolecular (Figure 11c) or paraffin structure (Figure 11d) of the surfactant will be adopted in the interlayer of the smectite clay. If a high layer charge density organoclay contains surfactants with two long alkyl chains, a lipid-like structure of the clay can be found (Figure lle). The orientation of the surfactants in the interlayer can influence the wettability of the organoclay on the aqueous and organic phase. Also, the efficiency and longevity of the organoclay as triphase catalyst are affected by the surfactant orientations, as will be discussed in the later chapter. b) Colloidal properties of the organoclays The amphiphilic character of the organoclays is a potentially important property for efficient triphase catalysis. Sodium clays are only wettable in aqueous solution or polar aprotic solvents such as methanol. The purpose of intercalating an onium ion surfactant into the clay interlayers and extemal surface is to increase the hydrophobicity of the material. The hydrophilicity or hydrophobicity of the organoclay are easily judged by observing the wettability of the material. Also, BET surface area measurements can provide information on the hydrophilic portion. The surfactant alkyl group will fill the interlayer gallery leaving no empty space accessible for the nitrogen adsorption. 30 An amphiphilic organoclay can stabilize an emulsion of water and aprotic organic solution. Therefore, the nucleophile and the organic electrophile from two immiscible phases can be brought together on the phase boundary to undergo nucleophilic substitution reaction. Generally, an organoclay with a monolayer structure is a poor emulsifier for the water and oil mixture. However, the surfactants adopted in pseudo trimolecular and paraffin-like structure sometimes desorb from the reaction condition, making the high charge density organoclays poor catalysts. The amphiphilic nature of organoclay surfaces has been applied already in photochemistry74'77 and as absorbents for environmental pollutants73i79. Some organoclay surfaces may adsorb neutral chromophores, such as pyrene dissolved in aqueous solution. The adsorbed surfactants in clay hosts provide a hydrophobic site for the adsorption of luminescent organic molecules such as pyrene. In these colloids a large amount of pyrene excirner, in the form of dirnerized pyrene, is generated at a pyrene concentration of 5x10'5M, significantly below that needed to form an excirner in homogeneous solution. Hence, pyrene excirner formation is a result of localization of pyrene on the surface of a clay particle, which has the effect of dramatically increasing the local concentration. The pyrene excirner has an emission spectrum different from that of the pyrene momomer. Also, the probe molecule (pyrenylbutyl)trimethylammoniurn bromide (PN+) fluoresces well on the clay surface containing hexadecyltrimethyl ammonium bromide (CTAB). 31 However, PN+ fluorescence is quenched by dimethylaniline, nitrobenzene and nitromethane in the CTAB-Clay system. The hydrophobic clay surface adsorbs the organic quencher and reduces the life time and the intensity of PN+. This hydrophobic property of the organoclay has also been utilized for environmental applications. Boyd and co-workers have used amphiphilic organoclays to eliminate organic contaminants, such as phenyl derivatives from aqueous solution73i79. The sorption isotherms of benzene, toluene, ethylbenzene, propylbenzene, butylbenzene, naphthalene and biphenyl on the CTAB- clays indicate that sorption occurrs by partitioning interaction with the C15H33NMe3+-derived phase. In general, increasing the C16H33NMe3+ content and basal spacings with increasing clay charge density increased the non—ionic organic compound sorption on CTAB-Clays. Increased sorption of alkylbenzenes by high charge density organoclays can be attributed to the ability of large basal spacing to accommodate larger solute molecules. c) Mechanism of organoclay triphase catalysis The mechanism of organoclay triphase catalysis are investigated by (i) observing the influence of organic solvent polarity on catalytic reactivity (ii) observing the relative O/C alkylation reaction of naphthoxide and benzyl bromide by using organoclays as catalysts (iii) studying the dependence of catalytic reactivity on the concentration of both anionic nucleophiles and organic reactants, and (iv) studying dependence of catalytic reactivity on the volume ratio of the two immiscible phases. 32 For polymer supported catalysts the nucleophilic substitution reaction proceeds in the organic phase. Since polar solvents have good polymer swelling ability, the catalytic reactivity for the ionic polymer catalysts increases with solvent polarity. If the swellability of organoclays by organic solvents plays an important role for triphase catalysis, the features of organoclay triphase catalysis should be similar to the catalytic reaction using polymer supported catalysts“. Otherwise, the nucleophilic substitution reactions may occur in the aqueous phase or at the boundary of clay-liquid phases. O/C alkylation (Figure 5) can provide information about whether nucleophilic substitution proceeds in either aprotic or protic media55963. An O-alkylation product results from reactions in an aprotic organic solution, and a C-alkylation product is favored from protic media such as aqueous solution. The O/C alkylation has been well studied for triphase catalysis using polymer supported catalysts“. Basically, the reaction products are predominately O-alkylation when using polymer supported catalysts, indicating that the nucleophilic substitution reaction occurs in the aprotic solvent. The organoclay triphase catalysis will be studied by the same reaction to determine the reaction microenvironment. The reaction location can also be elucidated by the dependence of catalytic reactivity of cyanation (Equation 1) on the C5H11Br + CN' ------ > C5H11CN 4» Br' (1) bulk concentration of liquid phases. Since the nucleophiles and 33 substrates are not in the same phase, there should be one or two of the reactants adsorbed on the organoclay where the nucleophilic substitution reaction may occur. Three different plausible mechanisms can be described by the three following Rate = k[organoclay. RX] [CN'](aq) (2) Rate = k[organoclay. CN'][RX](org). or (3) Rate = k[organoclay. CN' RX] (4) equations. Equation 2 can be applied when the organic electrophile is adsorbed on the organoclay catalyst before reacting with the nucleophile located in the bulk aqueous solution. Equation 3 is appropriate only when the nucleophile is adsorbed on the organoclay catalyst and then contacted with the organic electrophile in bulk organic solution. If neither Equation 2 nor Equation 3 applies, the mechanism may be interpreted by Equation 4. If the mechanism for Equation 2 is favored, the pseudo first order rate constant, kobss can be determined from Equation 5. The Robs will be equal to k[CN']. That is, the observed -d[RBr]/dt = kobisBrl (5) rate constant should be proportional to the bulk nucleophile concentration, [CN'], when the [CN'] is much larger than the [RX]. However, if the mechanism discussed by Equation 3 is true, another 34 pseudo first order rate constant, kobs’, can be determined from Equation 6. This kobs' will be proportional to the bulk -d[CN']/dt = kobs'ICN'] (6) concentration of organic electrophile, [RX], because kobs' is assumed to be equal to k[RBr]. The dependence of the pseudo first order rate constant on the bulk concentration of potassium cyanide (Equation 5) has been studied by Lin”. There is a linear relationship between the kobs and [CN‘], indicating the nucleophilic substitution reaction occurs at the catalyst-aqueous boundary. We further investigate the relation between the reaction rate, kobs'. and [RBr]. A non-linaer relationship between kobs'. and [RBr] can prove the the reaction is not occuring in organic phase. The rate dependence on the volume ratio of organic and aqueous phase also provides information about the emulsion structure and the wettability of organoclays in the two liquid solutions. If the three phases, organoclay, aqueous and organic solution, adopt oil-in-water emulsion, a milky emulsion and excess of organic solution will be observed. Likewise, an water-in-oil emulsion will perform a milky emulsion and excess aqueous phase when the volume of the aqueous solution is more than the maximum content of aqueous droplets in the emulsion phase. This emulsion structure can be observed by changing the volume ratio of the two liquid phases. In a water-in—oil emulsion system, organoclays are mainly suspended in organic solution and aqueous droplets are immersed in the aqueous solution. As the volume of the aqueous solution increases, 35 a redundant aqueous solution appears and does not participate in the catalytic nucleophilic substitution reaction. That is, the reaction rate will not be influenced by the volume of bulk aqueous phase if the an excess quantity of aqueous solution appears. However, increasing the volume of organic solution will decrease the organoclay concentration in the organic suspension for the water-in-oil emulsion. This will result in low efficiency for the adsorption of organic electrolyte and catalytic reactivity in this water-in-oil emulsion system. 2. Longevity of the Clay Derivatives for Phase Transfer Catalysis. a) Desorption of interlayer organic compounds The most advantageous property of triphase catalysts is their potential recyclability. Polymer and inorganic matrices covalently link the onium salts so the catalyst structure will be maintained after the catalytic reactions. However, the polymer supported catalysts have yet to find industrial applications because of their diffusion limitations of mechanism and chemical instability30o31. Several inorganic supports suffer the same general disadvantage, namely, low reactivity on structural instability under reaction conditions. For example, silica gel as a support is destroyed by alkaline solution59. The decrease in organoclay recyclability for triphase catalysis is potentially limited by the ion exchange of the surfactants in the clay hosts with the counter cations of the nucleophile. The dissociated surfactants may dissolve into the organic phase, so the surfactant desorption may be dependent on the polarity of the organic solvent, the surfactant hydrophobicity and the affinity of the layered silicate for the surfactant. 36 Generally, polar organic solvents exhibit higher solubility for surfactants. Consequently, the surfactant may dissociate from the clay hosts to join the organic phase. To reduce the surfactant desorption, non-polar organic solvents are suggested for use in triphase catalysis. We expect organoclays suspended in non-polar organic solvents to exhibit good recyclability. Also, the organoclay triphase catalysis reactivity should be greater in non-polar solvents than polar solvents. The desorption or ion exchange properties of an organoclay under triphase reactions should depend on the nature of the onium ion itself. For example, the carbon-rich surfactant, hexadecyltributyl phosphonium, may desorb from F-hectorite but hexadecyltrimethyl ammonium may not desorb from the clay host during the triphase catalysis reactions. The three bulky butyl groups shield the positively charged phosphonium leading to a surfactant with more organic than ionic character. Therefore, the ion pair of the carbon-rich surfactant and nucleophile should have good solubility in organic solvents, and this should favor desorption of the surfactant into the organic phase. Since the charge radius ratio of the positively charged nitrogen in hexadecyltrimethyl ammonium is large and this cation is only shielded by three methyl and one long chain alkyl groups, leaving the surfactant more ionic than organic in nature. The solubility of the surfactant derivative should be poor in organic solvents. The attraction between the layered silicate and the surfactant should also depend on the layer charge density of the silicate host. The charge density of the host will determine the orientation of the organic cation in the gallery. Higher layer charge density clays should give paraffin structures and this should result in more surfactant desorption from the 37 layer sheet. However, surfactant desorption is not expected on organoclays with low layer charge densities causing these materials to be good triphase catalysts in terms of recyclability. Behaving differently from intercalated surfactants, cationic crown ether complexes can also pillar the clay interlayer“. However, croWn ether complex are poor candidates for triphase catalyst. The X-ray diffractogram diffraction data suggest thst a crown ether clay complex after a triphase catalysis reaction undergoes complex desorption by ion exchange. However, the recycled clay derivatives show IR absorptions of characteristic of the crown ether and reactivity for the catalytic reaction. Thus crown ethers may remain partially bonded to the extemal surface of the layered silicate instead of being intercalated in the gallery. b) Thermal and chemical instability of the organoclay Natural layered silicates are sometimes destroyed by strongly acidic or alkaline solutions, a potential limitation for their use in catalytic reactions which involve acidic or basic reagents. Surfactants based on the silicate surface can change the chemical reactivity of the smectite clay hosts. Structure changes due an acid or base reaction can easily be observed from the X-Ray diffraction measurement. It is very surprising that some of the organoclay derivatives with medium layer charge density such as organo hectorite are not affected by alkaline solution. This resistance to strong base is important and valuable since many nucleophilic substitution reactions employ strong bases such as cyanide or hydroxide as nucleophiles. In contrast, high layer charge density clay 38 derivatives do not seem to tolerate strong alkaline solutions since surfactant desorption produces an unprotected layered silicate surface. For triphase catalysis systems in strongly acidic environments all of the clays and their organoclays are expected to decompose at elevated temperature. However, typical phase transfer catalysis condition seldom utilizes strong acid reactants. The layer charge density of dioctahedral smectite clays (Figure 1) with small metal cations in the interlayer causes them to be thermally unstable at temperature above 250°C. An interlayer small cation such as lithium or sodium may migrate into the empty octahedral site at elevated temperature, resulting in lower layer charge density. Since there is no octahedral empty site in trioctahedral smectite clays (Figure 2), the layer charge density of the layered silicate is not influenced by the thermal treatment. Under triphase catalysis conditions, the aqueous phase typically contains metal cations which may affect the layer charge density of the dioctahedral smectite clays33’34 and cause surfactant desorption from the clay host. Thus, clay layer charge density reduction is a potential mechanism for surfactant desorption and the influence of thermal treatment on the the reduction of layer charge density is a potential disadvantage for dioctahedral organoclays as phase transfer catalysts. However, almost all phase transfer catalysis reactions proceed at temperatures below 250°C so this effect should not be significant in organoclay triphase catalysis. 39 Alkyl ammonium surfactants in clay interlayers may be thermally unstable under alkaline solution reaction conditions Quaternary hydroxides, for instance, are known to undergo Hofrnann degration35-37. Thermal decomposition of the ammonium surfactants occurs at elevated temperatures. [SH-elimination of the alkyl ammonium is the major cause for decomposition to olefin and amine compounds. The alkyl ammonium surfactants in the clay interlayer may be decomposed under reaction condition. The Hoffman degration reaction could be reduced if the alkyl ammoniums do not contain B-hydrogen. Fisk at Dow Chemical Company employs N-alkylarylpyridinium salts as the surfactant adsorbed on the metal surface for corrosion inhibition, since the surfactants can tolerate higher temperature treatment due to the absence of a B-hydrogen. c) Co-surfactant influence on surfactant desorption Hexadecyltrimethyl ammonium bromide (CTAB) is a typical cationic surfactant that forms an emulsion of two immiscible liquid phases. Adding an alcohol co-surfactant causes microemulsion formation. When this same surfactant is intercalated in layered silicate hosts for a triphase catalysis reaction, it is not expected to desorb even in the presence of concentrated electrolyte owing to the low solubility of the surfactant-anion pair in organic and aqueous solutions. Dobias found that adsorption of hexadecyl pyridinium chloride on minerals such as quartz decreases dramatically when the surfactant concentration is higher than the critical micelle concentration (CMC)33. It was suggested that surfactant in the form of a three dimensional micelle is more stable than a two-dimensional film adsorbed on the mineral surface. It is of interest to characterize the Istability of an organo smectite clay when a 4o thermodynamically stable microemulsion is formed in the presence of a co-surfactant such as an alcohol with more than five carbon atoms. Microemulsions are dispersions of oil and water made with surfactant and co-surfactant molecules2535. Either type of dispersion, oil-in-water or water-in-oil (Figure 12) is possible. The droplet size are very small, typically 100A, about 100 times smaller than typical emulsion droplet sizes27923. Also, the interfacial tension of the fluid is lower27- 2939, leading to larger contact surface between two immiscible liquids90. Moreover, microemulsions are thennodynamically stable, compared to emulsions25,25. Because the smaller surface tension results in smaller particle dispersion and an entropy increase. The negative entropy may compensate the surface energy to minimize the free energy. The co- surfactant plays an important role in partitioning between two neighboring polar groups of cationic surfactants to reduce the surface charge density on the surfactant-aqueous boundary9oi91. Microemulsions are very useful in industry. Microemulsions have been used to improve oil recovery when oil prices reached levels where tertiary recovery methods became profitable”. . Nowadays, microemulsion application is focused on solar energy conversion, liquid-liquid extraction, detergency and lubrication. Besides these applications, microemulsions can enhance the catalytic reactivity for micelle catalysis or biphase transfer catalysis. Introducing the co- surfactant to the oil\surfactant\water reaction mixture improves the catalytic reactivity of the biphase catalysis, because the smaller dispersion droplet leads to greater interfacial area for the catalytic reactions. Figure 12 A droplet dispersion of oil-in-water microemulsion. 42 On the other hand, co-surfactant may facilitate desorption of a cationic surfactant such as hexadecyltrimethyl ammonium from some layered silicate surfaces by forming a more thermodynamically stable microemulsion containing the co-surfactant, the two immiscible liquids and the ammonium surfactant. The recyclability of the organo clays would then be reduced if the reaction mixture contains the co-surfactant. Alternatively the co-surfactant could improve dispersion of the clay particles in the triphase reaction and thus enhance reactivity. CHAPTER II EXPERIMENTAL A. Material. 1. Sodium Hectorite Naturally occurring California sodium hectorite (Bl-26) with a particle size of <2 um was obtained from Source Clay Mineral Depository, University of Missouri, in the pre-centrifuged and spray dried form. The mineral was purified by removing carbonates using pH5 acetate buffer solution and eliminating iron oxides by employing sodium hydrosulfate93i94. The idealized anhydrous unit-cell formula of hectorite is Nao,67[Lio,57Mg5,33(Si3,oo)020(OH,F)4], and the experimentally determined cation exchange capacity is about 73 meq/100g95 of the air dried clay. 43 44 2. Sodium montrnorillonites Two naturally occurring sodium montrnorillonites from Wyoming and Arizona with a particle size of <2 um were also obtained from Source Clay Mineral Depository, University of Missouri, in the pre-centrifuged and spray dried form. The purification method for the montmorillonite is the same as that stated above for hectorite. The idealized anhydrous unit-cell formula of montmorillonite (Wyoming) is Nao.7o[Mgo.7oAls.30(Sis.oo)02o(0H)4l. and the experimentally determined cation exchange capacity is about 75 mqu 100g of air dried clay95. The unit-cell of montmorillonite (Arizona) is Na1.16[Mg1.16Al2.s4(Si8.oo)02o(0H)4l, with a cation exchange capacity of 118 meq/l 00g of air dried clay. 3. Laponite R Synthetic laponite was obtained from Laporte Company, in England and was used without further purification. The unit-cell of this material is Lio,36[Lio,35Mg5,54(Si3,oo)Ozo(OH)4], with the cation exchange capacity of 55 meq./100g95 of air dried clay. 4. Fluorohectorite (F-hectorite) In this synthetic hectorite the octahedral lattice hydroxyl groups have been replaced by fluoride ions. The unit-cell formula of F-hectorite is L11,5o[Li1,soMg4,4o(Si3,oo)OzoF4]. The particle size of the material is larger than 2pm and the cation exchange capacity is approximately l40meq./100g95 of air dried clay. 45 5. Rectorite This mineral was from Ba-Tou, China3. The particles larger than 2pm were removed by suspending the mineral in an aqueous solution for 8 hours. Rectorite consists of a regular alternation of mica-like layers and expandable layers having the smectite composition. The chemical composition formula of the material is l(Nao.72Ko.02Cao.os)(Cao.24Nao.m)l(Al4.ooMgo.02)[Sio.58A11.121022. The negative charge on the layer arises from isomorphous substitution on the tetrahedral silica oxygen sheet. The cation exchange capacity is 60 meq./100g of air dried clay3. 6. Organic Reagents All reagents were obtained commercially and used without further purification. Hexadecyltributhyl phosphonium bromide; hexadecyltrimethyl ammonium bromide; tetrabutyl ammonium bromide; and didodecyldirnethyl ammonium bromide were obtained from Chemical Dynamics Company. Pentyl bromide, pentyl cyanide, decane, dodecane, benzyl bromide, 1,2-dibromo-l-phenylethane, trans-stilbene, n-octanol, n-decanol, n- tetradecanol, 1,8-octandiol, 1,10—decandiol, 6-bromohexanol, 8- bromooctanol, benzaldehyde, l-chloro-2,4-dinitrobenzene, l8-crown-6, dicyclohexa-chrown-6, and 2-bromoethylbenzene were obtained from Aldrich Chemical Corp. Toluene, methylene chloride, and dibromomethane were purchased from Mallinckrodt. Chloroform, methylene chloride, o-dichlorobenzene, butanol and naphthol were obtained from J. T. Baker Chemical company. Methyl orange, benzyl alcohol and isopentyl alcohol were purchased from Ficher-Scientific Company. 7. Inorganic Reagents Sodium hydroxide, sodium bromide, sodium bicarbonate, hydrobromic acid and sodium acetate were purchased from EM Science company. Potassium cyanide was obtained from Fisher Scientific company. Hydrochloric acid, magnesium sulfate and sodium chloride were available from Columbus Chemical Industries Inc. Acetic acid, sulfuric acid and sodium hydrosulfate were purchased from Mallinckrodt. Sodium citrate was obtained from I. T. Baker Chemical company. 8. Synthesis of organic precursors BrClonoOH97 was prepared by refluxing the mixture of 30 mmole of HOC10H200H and 20 mL of 48% HBr at 115°C in an oil bath for 5.5 hours followed by chromatographic separation using first hexane and then 1:1 hexanezether as eluent. The purity of the BrC10H200H was 45%. The remaining 55% of the products consisted of HOClonoOH and BrCronoBr- Sodium naphthoxide was prepared by mixing 0.20 mole of NaOH in 20 mL aqueous solution and 0.21 mole of 2-naphthol in 100 mL methanolic solution and allowing the solvent to evaporate under reduced pressure at room temperature. The yield of pure sodium naphthoxide was 75.2%. 47 B. Preparation of Organic Clay Derivatives l. Organoclays An aliquot of 1 % aqueous suspension of the sodium clay was added to an aliquot of aqueous solution of known amount of tetraalkyl ammonium or phosphonium at the room temperature. The amount of the onium salts was twice on the clay cation exchange capacity (CEC). For example, for sodium hectorite with a CEC of 73 meq/ 100g, an aliquot of aqueous solution containing 146 mole of onium salt was added to the suspension solution with 100 g clay to ensure that all the interalyer sodium cations were replaced by onium ions. After 24 hours of stirring, the products were purified by repetitively washing with ethanol to remove excess onium salt and then, resuspended in water until free of halide ion as tested by AgNO3. The pure products were collected by centrifugation and air dried at room temperature. 2. Crown Ether Clay Complex To sodium clay suspensions in methanol was added 2 CBC equivalent of l8-crown—6 or cis—dicyclohexal8-crown-6. These mixtures were stirred for 24 hours at room temperature. The excess crown ethers were removed by continuously washing with methanol. C. Organoclay Triphase Catalysis 1. Determination of Cyanation kobs (Equation 5) To a Coming culture tube was added 0.100 g of organoclay catalyst containing 0.06 mole of onium ion, an aliquot of 3 mL 6.67 M potassium cyanide aqueous solution, 6 mL toluene solution containing 2 mole of pentyl bromideand the internal standard, 0.15 g decane, the tube was sealed and stined vigorously in a 90°C oil bath. Aliquots of 5 to 10 48 microliter of the organic solutions were withdrawn at 20 to 30 minutes interval and diluted in 0.5 ml of ether for GLC analysis. The conversion of pentyl bromide to pentyl cyanide was calculated from the ratio changes of pentyl bromide to decane. Pseudo first order rate constants were calculated with a least square program. Generally, the total reaction time was 1.5 to 3.5 hours. 2. Determination of Cyanation k'obs (Equation 6) when the Amount of Pentyl Bromide Is Much Larger than Potassium Cyanide To a Corning culture tube, was added to a solution of 0.130g of potassium cyanide in 3 mL of water, 0.100 g of organoclay, and 5 mole of pentyl bromide in 6 mL toluene solution containing exactly 0.100 g decane. The same procedure for determining different k'obs was applied to pentyl bromide concentrations in toluene equal to 0.83M, 1.67 M, 2.50 M, and 3.33 M. Reactions carried out in a 90°C oil bath were monitored by withdrawing 5 to 10 microliter samples from the organic phase at 20 to 30 minute intervals. It was assumed that the loss of cyanide ion was equal to the formation of pentyl cyanide. A standard GLC curve was made by area integration of a standard solution containing 0.2 to 2 mole versus 0.100 g decane. The plot of mole of pentyl cyanide against the ratio of pentyl cyanide to decane is shown in Figure 13. The chemical yields of pentyl cyanide were determined by calculating the integration ratio of pentyl cyanide to decane from the GLC data and interpolating the ratio value in Fig 13 to find the pentyl cyanide concentration in the reaction mumm. Intensity Ratio of [CSHuCNMOJOOg C1011”) 49 Figure 13 The standard curve of the intensity area ratio of [CSHnCN]/[0.100 g C10H22] in GLC analysis versus [CSHuCN] in 6 mL toluene as the organic solvent. 1.6 I ' I ' I ' U I 0.0 0.4 0.0 1.2 1.6 2.0 [CsHuCN], mmole 50 3. Solvent Effect on Iodination To a Coming culture tube, 0.100 g of hexadecyltributyl phosphonium hectorite (H26AA) (equivalent of 0.06 mole of hexadecyltrimethyl phosphonium) was added, then 8 mole of potassium iodide in 3 mL water followed by 6 mL of desired organic solvent containing 2 mole of pentyl bromide. The deserved organic solvents could be decane, toluene, or o-dichlorobenzene. An internal standard, 0.15 g decane or dodecane, was added to the mixture, and the tube was sealed and stirred vigorously at 900C in an oil bath. Aliquots of 5 to 10 microliter of the organic solution were withdrawn at intervals of 20 to 30 minutes and diluted in 0.5 mL of ether for quantitative GLC analysis. The conversion of pentyl bromide to pentyl cyanide was calculated from the ratio of pentyl bromide to internal standard. Pseudo first order (rate constants were calculated with a least square program. 4. O/C Alkylation of Benzyl Bromide and Sodium Naphthoxide. A deserved amount of catalyst was added to a mixture of 0.5g (3 mole) of sodium naphthoxide in 5 mL water and 0.34 g (2 mole) of benzyl bromide in 5 mL of organic solution. The mixture was stirred at room temperature for 4 hours, then the organic phase was separated and the organic solvent was evaporated under reduced pressure. The ratio of O-alkylation and C-alkylation product (Figure 7) was determined by integration of the methylene proton absorption resonances in the 1H NMR spectra. The methylene proton resonances in the 1H NMR spectrum for C-alkylation and O-alkylation products occur at 4.48 ppm and 5.12 ppm, respectively. 5. Dependence of kobs on the Volume Ratio. 51 Aliquots of 6.25 M potassium cyanide aqueous solution, 0.33M pentyl bromide in toluene, and 0.100 g catalyst were stirred in a 90°C oil bath. The conversion of pentyl bromide to pentyl cyanide was calculated from the chromatographic ratio of pentyl bromide to the internal standard decane in GLC analysis. 6. Catalytic Cyanation in the Absence of Water A 6 mL quantity of toluene solution containing 2 mole of pentyl bromide, 20 mole potassium cyanide and 0.100 catalyst H26AA were stined at 900C in an oil bath. The conversion of pentyl bromide to pentyl cyanide was calculated from the chromatographic ratio of pentyl bromide to decane in the GLC analysis. The total reaction time was generally 1.5 to 3.5 hours. D. Longevities of Organoclay Derivatives l. Recyclabilities of Organoclays for Cyanation Aliquots containing 20 mole of potassium cyanide in 3 mL water and 2 mmole of pentyl bromide in 6 mL organic solution were added to 0.100 g of the organoclay catalyst. The organic solvent was either toluene or decane. The mixture was sealed in a Corning tube and stirred at 90°C in an oil bath. After the catalytic reaction was finished and the observed rate constant (kobs) was determined, the organoclay catalyst was filtered and then washed with 10 ml of water and ethanol. The catalytic reaction was then repeated using the recycled catalyst. 52 2. Recyclabilities of Organoclays for Chlorination Aliquots of 4 mL 2.5N NaCl aqueous solution and 4 mL of a toluene solution containing 2 mole of pentyl bromide and an intemal standard of 0.15 g decane were added to 0.100 g of the organoclay catalyst. The mixture was stirred at 90°C for 8 hours. The conversion was determined from the change in the integral GLC intensity for pentyl bromide and the intemal standard. The catalyst was filtered and washed with 10 mL fresh water and ethanol. The same catalytic reaction was repeated using the recycled catalyst. 3 Surfactant Desorption of Organoclays The desorption of surfactants from clay hosts was determined by measuring the concentration of cationic surfactants in bulk liquid phase after ion-exchange reactions. The desorbed cationic surfactant concentration in liquid solution was determined by the method developed by Wang and Langley93. Mixtures containing 0.03g organoclay, 4 mL aqueous solution with desired electrolyte concentrations and 4 mL of organic solvents were stirred at 90°C or room temperature for the desired time. After the ion-exchange reactions were finished, 1 mL of both organic and aqueous solutions were transfered to a 50 mL buffer solution containing citric acid and 1.0 mg methyl orange. If the desorption of onium ion was very high, the quantity of methyl orange was doubled or increased further to ensure an excess of indictor. The aqueous solutions were extracted using 25 mL of chloroform. Standard curves of UV-Vis absorptions in chloroform solution with a 10 mm length cell against known surfactant concentrations (Figures 14 and 15) were determined. The absorption for the surfactant-dye ion pairs in chloroform were measured at 418 run. If absorbancies were over the limit of the 53 Figure 14 Dependence of the UV-Vis absorption of C14133NMe3Brt methyl orange chlorofonn solution at 418 nm on concentration of C161133NMe3Br aqueous solution. 0.7 0.6 e 0.5 e 0.4 - 0.3 - Absorption 0.2 4 0.1 - 0.0 . . . . . 0.0 1.0 2.0 3.0 4.0 [CroHaaNM‘BsBrL 118/ml Various concentrations of C16H33NMe3Br in aqueous solution were added to excess amount of methyl orange. The ion pairs of the cationic surfactant and anionic dye were extracted into 25 mL of chloroform. The absorptions of the chloroform solutions in a 10 mm length cell were determined bya UV-Visible spectrophotometer at the wavelength 418 nm. 54 Dependence of the UV-Vis absorption of CIJI33PBU3BT+- methyl orange chloroform solution at 418 nm on concentration of C16H33PBu3Br aqueous solution. Figure 15 Absorption 0.0 I l I I ' I ' 0.0 1.0 2.0 3.0 4 0 [C16H33PB‘1331‘L 118/“11 Various concentrations of C161133PBu3Br in aqueous solution were added to excess amount of methyl orange. The ion pairs of the cationic surfactant and anionic dye were extracted into 25 mL of chloroform. The absorptions of the chloroform solutions in a 10 mm length cell were determined by a UV-Visible spectrophotometer at the wavelength 418 run. 55 instrumental measurement, the solution was diluted by adding more chloroform until the absorption was within the standard curve. According to Beer's Law, the concentration of the surfactant-dye adducts were proportional to absorption. Thus, desorption of surfactants from clay hosts to liquid phases were derived by measuring the UV-Vis absorption and finding the relative surfactant concentration from the standard curve. 4. Effect of Co-surfactant on Phase Transfer Catalysis Reactions. To a Coming culture tube was added 2 mole potassium cyanide in 3 mL of water, 0.100g of clay supported catalyst (equivalent to 0.060 mole of onium salt), and then 1.511 g (20 mole) of pentyl bromide and the desired amount of co-surfactants in 6 mL of toluene. Subsequently 0.100 g of decane, the internal standard, was added to the mixture, the tube sealed and vigorously stined at 90° for 4 hours. The conversions of pentyl bromide were determined from the GLC integrated intensity of pentyl cyanide relative to the decane integral intensity. A standard GLC curve for the integration ratio of pentyl cyanide to decane is determined (Figure 13) to quantify the amount of pentyl cyanide in the reaction mixtures. ' 5. Conversion of n-Bromoalkan-l-ol to n-Cyanoalkan-l-ol by Using Phase Transfer Catalysts To a Corning culture tube containing 0.100 g of the organoclay catalyst (equivalent to 0.06 mole of C15H33NMe3Br) was added on aqueous solution containing 10 mole potassium cyanide in 3 mL of water, and then 2 mole of n-bromoalkan-l-ol in 6 mL of toluene. After the internal standard, decane, was added to the mixture, the tube was sealed and stirred vigorously at 90°C in an oil bath for 4 hours. The 56 conversion of n-bromoalkan-l -ol was determined from the integrated intensities of n-bromoalkan- 1 -01 to decane peaks by GLC. 6. Conversion of n-Bromoalkane to n-Cyanoalkane Using Phase Transfer Catalysts To a Coming culture tube was added an aliquot solution containing 10 mole potassium cyanide in 3 mL of water, 0.100 g of the organoclay catalyst (equivalent to 0.06 mmole of C15H33NMe3Br) and 2 mole of n- bromoalkane in 6 mL of toluene. After 0.15 g of decane, the internal standard, was added to the mixture, the tube was sealed and stirred vigorously at 900C in an oil bath for 4 hours. The conversion of n- bromoalkan-l-ol was determined from the ratio change of the integrated intensities of n—bromoalkan-l-ol to decane peaks by GLC analysis. 7. Recyclability of Crown Ether Clays for Cyanation A toluene solution containing 2 mmole of benzyl bromide and 0.15 g ofdecane as the internal standard was added to 0.100 g ofthe crown ether clay and 5 mmole of potassium cyanide in 1 mL of water. THis mixture was placed in a Coming tube which was then sealed and stirred at 90°C in an oil bath for 12 hours. After the catalytic reaction was stopped, the conversion was determined by GLC. The solid material was filtered and washed using 10 mL of ethanol. The same catalytic reaction was repeated by using the recycbd material. 57 E. Application of Organoclay Triphase Catalysis to Organic Synthesis 1. Synthesis of Symmetrical Formaldehyde Acetals (Equation 10-12) Mixtures of 10 mL methylene chloride solution containing 10 mole of the desired alcohol, 10 g of dibromomethane and 10 mL of 50% NaOH aqueous solution were added to 0.2 g of the desired organoclay catalyst. The mixtures were refluxed with stirring at 90°C for 12 hours. After reaction, the liquid phases were easily filtered. The organic layer was dried over MgSO4 and the organic residue was obtained by evaporating the volatile organic solvent under reduced pressure. The structure and the yield of the product were determined by 1H NMR spectra. 2. Synthesis of Alkyl Bromide by Dehydration of Alkyl Alcohol in Strongly Acidic Conditions (Equation 13) To a Coming tube was added a mixture of 2.5 mole of octanol, 0.1 g of decane, 1 mL 47% HBr and 0.1 g of organoclay. The mixture was stirred at 90°C in an oil bath for 11 hours. The yield was determined by GLC. For the reaction with diluted reactants, 1 mL of water and 1 mL of toluene were added to the mixture indicated above. This reaction was also carried out at 90°C in an oil bath for 11 hours. 3. Dehydrohalogenation of 2-Bromoethylbenzene (Equation 14) A mixture of benzene containing 0.2 g of decane as the internal standard, 4 mole of 2-bromoethylbenzene, 2 mL 50% NaOH aqueous, and 0.1 g organoclay were introduced to a Corning tube which was sealed and heated at an elevated temperature. Yields were determined by GLC. 58 4. Oxidation of trans-Stilbene (Equation 15) A mixture of 3 mole of trans-stilbene in 10 mL benzene, 6 mole of KMn04 in 6 mL water and 0.200g organoclay catalyst, were refluxed with stirring at the room temperature for 8 hours. After filtering the reaction mixture, the acidic product in the organic layer was extraction with 20% NaOH aqueous solution. The aqueous layer was combined with the alkaline washing solution. The basic aqueous solution was acidified by hydrochloric acid, followed by extracted with 40 mL of methylene chloride. The pure benzoic acid was obtained by evaporating the organic solvent under reduced pressure and the yields were determined by weighing the product. 5. Synthesis of 2,4-Dinitrophenyl Ether (Equation 16) In a round bottomed flask a mixture of 5 mmole of 1-chloro-2,4- dinitrobenzene and 6 mole of phenol in 8 mL of benzene, 4 mL 1.5 N NaOH aqueous solution and 0.100 g of the organoclay catalyst were stirred at room temperature for 6 hours. After the organic layer was washed with deionized water, the product residue was obtained by evaporating the organic solvent under reduced presure. The purity of the organic residue was determined by 1H NMR spectra 1 F. Physical Measurements 1. Infrared Spectroscopy Infrared spectra were recorded using an IBM Single Beam FI‘ IR44 model spectrophotometer. Liquid spectra were obtained by using 0.1 mm NaCl cells and solid spectra were recorded by mixing the samples with KBr and pressing them into disks. 59 2. Gas Liquid Chromatography All product mixtures together with solvent and internal standard were analde by gas liquid chromatography either on a model 5880 Hewlett- Packard instrument filled with a flame ionization detector and a 25 m x 0.25 mm cross-linked dirnethylsilicone capillary column or on a model 5890 Hewlett-Packard chromatograph with a flame ionization detector and a capillary 60 m x 0.25 mm column. Products were identified by comparison of GLC retention times with those of authentic samples. The percentage yield of products was determined by integration of the “absorption peaks of the starting reagents and the internal standard. 3. NMR Spectroscopy Proton and carbon-13 nuclear magnetic resonance spectra were recorded on a Bruker WM-250 or Gemini-300 MHz spectrometer. Chemical shifts were usually measured relative to tetramethylsilane as intemal standard and reported in units of ppm. 4. X-Ray Diffraction X-ray d001 basal spacing were determined for oriented film samples with either a Philips X-ray or a Rigaku X-ray diffractometer using Cu-Ka radiation with wavelength equals to 1.5405 A. The fihn specimens were prepared by allowing an aqueous suspension of the samples to evaporate on a microscope glass and monitoring the diffraction over a 2-theta range from 1° to 40°. Peak positions in the angle 2 theta were converted to d- spacing with a standard chart. 5. UV-Vis spectroscopy UV-Vis spectra were recorded by a 9430 UV-Visible IBM Spectrophotometer. Sample concentrations were determined by . absorption measurements. Sample and reference solutions were contained in 10 mm quartz cells. 6. BET Surface Area Measurement The BET surface area were recorded by a QUANTASORB Jr. Sorption Analyzer. The specific surface area of the samples was determined by measuring sorption of nitrogen on the sample at liquid nitrogen temperature. CHAPTER III RESULTS AND DISCUSSION A. The Preparation and Structures of Clay Derivatives 1. The Preparation of Organoclays The reaction of sodium smectite in liquid suspension with stoichiometric amounts of onium salts dissolved in water, ethanol or acetone results in products in which the sodium ion is displaced by onium cations. A series of organoclays shown in Table l were prepared using this method. The first letter designates the clay source or type, e.g. "A" is Arizona montmorillonite and "F" represents the clay host, fluorohectorite. The number designates the onium ion in the clay interlayer and the reaction solvents are represented by the last two letters. For example, "AA" of A26AA means that the organoclay is synthesized in aqueous solution. Likewise, H24AB is synthesized in ethanol and H24AC is generated in acetone, and the number "26" designates the onium ion, C16H33PBu3+. 61 62 cos—€80 a: 92 in kmmuvznzao as? 8 2.533 <52 5 e: «.8 Emmovzazau as? a: 2582.”. «44$ 8 «.2 EN Emmovzazeu as? m: 2824 <39. 5 S 3: kamuvzszao as; 3 282.3 <39» 8 4.4 S: Esauvzszeo 253% R 3:98: USA: a. S w: 3:622:20 355m 2. areas: mesa: S S a: hmmovzazeo can? ma sees: 53: 3. as n: kmzovszeu as? an 3883 <33 2v €85.29 3; RSQEE A Baa—e85 Sou 880.95 Boar—em 08 >20 835.5 49:88 .33 can—«225-...Soatam _ 03am. 63 .Smcoficgeoa 282.3 mm .2884 :5 882E832: «5:5»? mm 4.58.? ... a on 0: 24349 .2; o: 2:28.: <32: an as a: 243:6: 883 2. 28.8: <32: 4mm 98 «.8 EEBZQREQ .243 a: 2:28;."— flan... 8 8.8 gm EEBZQQEB as? a: .282: 35?. m: :2 :8 +%:uvz§n:2uv 8a? 2 382.3 539» we we :8 EEBZNEEQ 883 ms 2:98: 5.8: a. S a: EEBZNEBQ .283 am 283 $33 «8 :2 Zn 3:23:20 833 8 2:28: <32 88 3: 3a +memaa:e_u as? c: 2:28.: <82 3 3: 3a 23:31.6 883 a: .2824 <32 2: am 2: 3:335 88>» 2. 582.3 2283 an S: B: $34316 883 2. 2:28: < C5H11CN + Br‘ ' (1) nucleophile, CN-, was held in ten fold equivalent excess over the organic substrate, CsHllBr, so that the observed rate constant (kobs) could be measured and calculated by Equation 5. -d[CsH1 lBr] /dt = koblesH 1 131'] (5) Organo clays that facilitate formation of water and organic emulsions were usually good phase transfer catalysts due to the efficient mixing of the two immiscible liquid phases. The organo laponites (L24AA and L26AA) containing C115H33NMe3+ and C15H33PBu3+, respectively, always show poor 76 Table 4 The catalytic activitiesa and the colloidal properties of the organoclays and the onium salts as phase transfer catalysts. Catalyst Wetting of the clay 105xkobs, see1 under reaction conditions L24AA water 0.31 H24AA emulsion 2.93 H24AB emulsion 3.07 H24AC water 0.25 W24AA emulsion 3.64 A24AA emulsion 4.07 F24AA emulsion 4.09 R24AA Paste in 1.40 toluene L26AA water 1.10 H26AA emulsion 12.9 W26AA emulsion 27.2 A26AA ion desorption 54.6 F26AA ion desorption 56.5 R26AA Paste in 8.20 toluene Continued 77 Table 4 (Continued) Catalyst Wetting of the clay 105xkobs, see-1 under reaction conditions L30AA partial emulsion 1.96 H30AA emulsion 2.92 W30AA emulsion 4.45 A30AA emulsion 8.39 H18AA water 2.01 C1¢5H33NMe3Brb emulsion 2.43 C15H33PBu3Brb toluene 62.2 NMeaBrb toluene 15.6 3. Reaction and wetting conditions: 2.0 mmole pentyl bromide in 6 mL toluene; 20.0 mmole potassium cyanide in 3 mL water; 0.100g organoclay; 90°C. b Biphase catalyst: The amount of C15H33NMe3Br, C15H33PBu3Br and NMe4Br are 0.60, 0.55 and 0.61 mole respectively. 78 catalytic activities with kobs' of 0.31 x10“5 and 1.10x10'5 sec-1, respectively, in triphase catalysis (Table 4), owing to inability to form the emulsion. The hydrophobicity of the material can not be effectively enhanced because of the large edge surface area versus the lamellar surface area of the clay. In Table 2, laponite derivatives are seen to exhibit larger surface areas than the other organoclays, resulting from the small particle size of the larnella laponite particles. The surface area receives significant contribution from the particle edge, which can not be covered by the alkyl chains of the surfactants. The incomplete coverage of the basal laponite surface by surfactant (Figure 7a) also results in poor amphiphilic properties. When the interlayer surfactant in hexadecyltrimethyl laponite was replaced by didodecyldimethyl ammonium( L30AA), kobs increased from 0.3lx10-5 to 1.96x10-5 sec-1, but the improvement was not satisfactory. For the lattice onium ions, the surfactant chains adopt a lateral bilayer structure in the clay interlayer and a monolayer structure on the external clay surface (Figure 7e), so that the extemal surface is totally covered by surfactant alkyl chains. As the clay layer charge density increases and the surface area of the organo clay decreases, the triphase catalytic activity of an organoclay increases (Table 4). Hectorite derivatives, H24AA and H26AA, with kobs values of 2.93x10'5 and 12.9x10-5 seC'1 respectively, facilitate the catalytic reactions more readily than laponite derivatives, L24AA and L26AA, with kobs values of 0.3lx10-5 and 1.10x10-5 sec-l respectively. For H24AA in which hexadecyltrimethyl ammonium is in the interlayer cation the triphase catalysis rate constant (2.93x10‘5sec-1) for cyanation is slightly larger than that kobs for the micellar biphase 79 catalysis when using the same equivalent surfactant, hexadecyltrimethyl ammonium bromide with kobs of 2.43x10‘5 3e0'1. The organo hectorite H26AA with hexadecyltributyl phosphonium as the interlayer onium ion (kobs = 12.9x10'5 seC'l) does not perform any better as a triphase catalyst than the corresponding biphase catalyst (kobs = 62.3x10‘5 sec-1). However, the recyclability and the greater convenience of workup after catalytic reaction more than compensates for the disadvantage of a lower reaction rate. The hectorite derivative with tetrabutyl ammonium in the interlayer (H18AA) is not a good triphase catalyst (kobs = 2.01x10-5 seC'l) owing to the instability of this clay to a toluene/water emulsion although tetrabutyl ammonium bromide is a powerful biphase catalyst (kobs = 1.56x10'4se0'1). This lack of amphiphilic character results from the short alkyl chain in the onium ion and the incomplete coverage of the external hectorite basal surface leading to poor enhancement on the material hydrophobicity. Therefore, the contact area of the two immiscible liquids partitioned by H18AA is much smaller than the emulsion stabilized by the high d-spacing organo hectorite, H24AA and H26AA. The solvent used for intercalation of the cationic surfactant into the hectorite interlayers also plays an important role in determining the extent of onium ion exchange and, hence, the colloidal and catalytic properties for the organo hectorite. Water and ethanol have high dielectric constants and produce good swellability for the clay silicate layers. This facilitates the interlayer exchange of the metal cations by the cationic surfactants. The d-spacings of organo hectorites, H24AA and H24AB, prepared in water and ethanol, respectively, exhibit no difference in d-spacing (17.8A) (Table 1). Also, the catalytic capacities 80 for the two materials are comparable. However, the C15H33NMe3+ surfactant can not be completely ion-exchanged into the hectorite interlayer using acetone (14.0A) as the solvent for the intercalation reaction and the d-spacing for the material synthesized in acetone is lower than that of H24AA and H24AB. Therefore, the catalytic capacity of H24AC with kobs 0.25le5 see1 is very poor. The silicate layer does not swell very well in low dielectric organic solvents. The triphase catalytic reactivities of the organo montrnorillonites and F-hectorite, A24AA and F24AA interlayered by hexadecyltrimethyl ammonium are also high (kobs = 4.07x10'5 and 565le5 sec-1 respectively). These clays have high layer charge densities. However, F26AA and A26AA do not stabilize emulsion formation even though they show unusually high kobs values. In this case, however, the surfactant is desorbed from the clay hosts. In fact, the observed reaction occurs by biphase catalysis. That is, A26AA and F26AA catalysis occurs by the desorbed and soluble surfactant instead of by solid organoclays. . Although rectorite derivatives, R24AA and R26AA, are also amphiphilic materials, they only form a paste-like gel with the water phase of the triphase catalyst system instead of forming a milk-like emulsion. A paste-like gel retards dynamical stirring during reaction. Thus the catalytic activities for the two rectorite derivatives with kobs 1.40x10-5 (R24AA) and 8.20le5 sec-1 (R26AA) respectively, are not as good as those of the other high layer charge density layered silicate derivatives. The thickness of the rectorite layer (19011) is twice that of smectite layer (9.6A) because rectorite layer stacking consists of a mica and an expandable smectite sheet3. Rectorite particles with a thickness 81 double that of the corresponding smectite particle may have a lower suspension extent in liquid solution than the smectite particle. Therefore, the rectorite particle aggregated and a paste formed. The colloidal properties of organoclays for triphase catalysis reactions as listed in Table 1 can be classified into three categories, as shown in Figure 18. The first class of organoclays, designated Type A clays, is amphiphilic and stabilizes emulsions of 3 mL 2.5 N NaCl aqueous and 6 mL toluene solutions (Figure 18a). The density of unsolvated organoclays should be greater than the density of water and toluene. However, when toluene is adsorbed by the amphiphilic organoclay, the density of the organoclays become intermediate between the density of water and toluene. After the amphiphilic clay/water/toluene mixtures are centrifuged, the solid organoclays are located between the aqueous and organic phases (Figure 18a). The second type of the organoclays, designated by Type B in Figure 18b, does not support emulsions of aqueous and organic solutions. Organo F-hectorite undergoes surfactant desorption and organo laponites become suspended in the water liquid phase. Toluene is not adsorbed on the hydrophilic organo laponite or ion exchanged sodium F-hectorite formed by the surfactant desorption. Upon centrifugation, the Type B clay materials collect at the bottom of the two liquid phases shown in Figure 18b. Type C organoclay is unique in rectorite derivatives. A paste is formed in the organic phase which retards dynamic stirring (Figure 18c). Since Type C materials also absorb toluene, these organoclays will be located at the boundary of the aqueous and organic solutions after the mixtures of organo rectorite, toluene and aqueous phases are centrifuged. 82 Figure 18 Classification of the colloidal behavior of 0.100 g of organoclay in liquid mixtures containing 3 mL of 6.25M potassium cyanide aqueous solution, 6 mL of toluene and 2 mole of pentylbromide. 1101A 12129.9 H24AA:I-126AA;H30AA; gal-2mm , R24AA; R26AA W24AA; W26AA; W30AA; A24AA; A30AA; F24AA a—Organic Layer j—(hganic Layer ~:L-(Jrganic Layer VJ—Aqueous Layer * __Aqueous Layer Nil—Aqueous Layer Organoclay EId—‘Organoclay V—- Organoclay Before Stirring Before Stining Before Stirring U U U —Milky Emulsion —Immiscible Liquids ‘_ P33“ v _Aqueous Layer Dims“ sumn‘ ' a Drum stirring mm 3 stirring I Organic Layer Organic Layer —Milky Emulsion __PaS‘c U SClay-Water Aqueous Layer uspensron After standing for 1 hour After standing for 1 hour After standing for 1 hour U U U _1— Organic Layer 4— Organic Layer — Organic Layer 07833001” Aqueous La er Organoclay Aqueous Layer :J_.. Organoclayy Aqueous Layer After Centrifugation After Centrifugation After Centrifugation 83 2 . Dependence of Catalytic Reactivity on the Bulk Reactant Concentration of Liquid Phases Under triphase reaction condition, the nucleophile and organic substrate are segregated into separate phases. Thus, one or both of the reactants must be adsorbed on the organoclay to interact with each other either at the solid liquid interface or within the solid phase. As described in the Introductory Chapter three different triphase process are possible and are characterized by the three following equations. Rate = k[organo clay- RX] [CN'](aq) (2) , Rate = k[organo clay- CN'][RX](org). or (3) Rate = k[organoclay- CN' RX] (4) Equation 2 describes the catalytic reaction occurring at the clay- aqueous interface. If the equation holds for the pseudo first order condition, i.e., [CN'] >> [RX], a pseudo first order rate constant, is designated kobs. is equal to k[CN']. The observed constant, kobs. was calculated from Equation 5 in which the [RX] value was traced by GLC at different reaction time during the biphase reaction. leXl/dt = kobisXl (5) then. lanXI/[RXIO = 'kobst Figure 19a is an example of determining the -kobs value by finding the slope of ln[RX]/[RX]0 against the reaction time, t. The observed rate constant, Robs. is proportional to the bulk cyanide concentration if the triphase reaction occurs at catalyst aqueous interface. 84 For this assumption, four aliquots of cyanide solution with 5, 10, 15 and 20 mole in 3 mL aqueous solution versus 2 mole of pentyl bromide in 6 mL of toluene were prepared. According to Lin's investigation”, kobs is linear to [CN‘] (Figure 20a), indicating that equation 2 is an appropriate description of the reaction rate under pseudo first order rate condition and the catalytic reaction are suggested to occur at the catalyst-aqueous boundary. For polymer supported phase transfer catalysts, the catalytic reactivity is insensitive to the bulk concentration of cyanide“. Organoclay triphase catalysis reactions are carried out in different reaction environment from the polymer supported triphase catalysis. If the condition for organoclay phase transfer catalysis are first order in electrophile (Equation 5) by varying pentyl bromide from 5 to 20 mole in 6 mL organic phase which keeping cyanide constant at 2 mole in 3 milliliter of aqueous solution, the kobs' was not linear with the concentration of [RX] (Figure 20b). This pesudo orer rate constant, k'obSr is derived from Equation 6. d[CN']/dt = kobs'lCN'l (6) Then. Ln([CN'l/[CN']o) = kobs't 0r. Ln((2-[C5H11CN])/2) = kobs't where the [C5H11CN] designates the miniequivalent of C5H11CN. 85 Figure 19 (a) Determination of a k0.” from the slope of a plot of [CanBr]/[C5HnBr]° versus reaction time ill the presence of 0.100 g C16H33NMe3hectorite. -0.20 - Ln([C5HuBr]/[C5HuBr]o) -0.40 ....,. 0 5000 Time, sec The initial concentration of CN' was 20 mole in 3 mL aqueous solution and [CsHuBr] was 2 mole in 6 mL toluene solution. The equivalent of the CN' was ten times that of CsHuBr', that was recognized that the nucleophile, CN', was far more than electroophile, CsHuBr. Thus, the observed rate constant was first order in [CsHuBr] and the kinetic equation was represented by d[C5HuBr]/dt = -kob,[C5HuBr]. ThereforeJtob. could be derived from ln([C5HuBr]/[C5H11Br]o) = -kobst. Ln{(0.67-[05H11Br])/067} 86 Figure 19 (Continued) (b) Determination of a kob,’ from the slope of a plot of [CN']/[CN]0 against reaction time in the presence Of 0.100g C16H33PBU3hCCtOI'itC (H26AA) . 0.00 -Slope = kob; = 6.271110'5sec'l -0.20 - -0.40 - 70-60 ' r ' I ‘ I ' l 0 2000 4000 6000 8000 Time, sec The initial concentration of CN' was 2 mole in 3 mL aqueous solution and [CsHuBr] was 10 mole in 6 mL toluene solution. The equivalent of the CanBr was five times that of CN'. Thus, the reaction was first order in the concentration of the nucleophile [CN'] and the kinetic equation was represented by d[CN']/dt = kom't. Then the observed rate constant (kobs') could be derived from In( [CN']/ [CN']0) = «out. The depletion of CN' was equal to the production of CSHHCN, which could be analyzed by GLC. The equivalent of CN' was equal to (2 - CsHuCN). 87 Figure 20 (a) Dependence of the observed rate constant (Robs) on nucleophile concentration for the cyanation reaction (Equation 1) at 90°C in the presence of 0.100 g . C15H33PBu3hectorite (H26AA) as the triphase catalyst. losntomcec“) r I 1 r 0 2 4 6 Concentration(Mole/Liter) The pseudo order rate constant kg.” was determined under conditions when the reactant is first order in electrophile; i.e., the nucleophile concentraction was much larger than the organic electrophile concentration, which was kept at 0.33M in toluene. 88 Figure 20 (continued) (b)Dependence of the observed rate constant (kobS') on the organic electrophile concentration for the reaction at 90°C in the presence of 0.100 g C16H33PBu3hectorite as the triphase catalyst. 12 Pentyl Bromide Concentration (Mole/Liter) The pseudo first order rate constant kob,‘ was determined under conditions when the reaction is first order in nucleophile. That is, the electrophile concentration was far larger than the nucleOphile concentration, which was kept at only 0.67 M in aqueous solution. 89 Figure 19b is an example of detennining kobs' from the slope of a plot of ln((0.67-[C5H11CN])/0.67) against reaction time, t. This result demonstrates that the mechanism represented by Equation 2 does not hold. This behavior is explained by a model in which the organic electrophile is initially adsorbed on the boundary of the clay and the organic solution and then the clay is transferred to the aqueous phase to react with the nucleophile in the aqueous solution. The catalytic reaction rate depends on the nucleophile bulk concentration and the surface concentration of the organic electrophile on the organoclay (Equation '2). The surface concentration of adsorbed pentyl bromide on the clay may be increased by increasing the bulk electrophile concentration, but the adsorption relation is not linear. Consequently, the kobs' curve in Figure 19b is not a straight line. 3 . Dependence of Organoclay Catalytic Reactivity on the . Polarity of the Organic Solvent The solvent polarity dependence of organoclay triphase catalysis can also afford information on the catalytic reaction mechanism. Lin has investigated the influence of polarity on the cyanation reaction by using C15H33PBu3+hectorite as the catalyst”. Organic solvents of lower polarity result in higher catalytic reactivity for the organoclay triphase catalysis. In contrast, Montanari, et al.51452 using polystyrene- supported phosphonium bromide as a triphase catalyst, found that higher polarity organic solvents led to higher iodination reaction rates. The organo hectorite, H26AA, when used for iodination (Equation 7), exhibits a reactivity dependence on the solvent polarity which is opposite that derived for polymer supported catalysts. The data in Table 5 indicate that with H26AA as the triphase catalyst solvents of low 90 polarity produce high reactivity for the iodination of pentyl bromide. If the onium ion is supported in C5H11Br + I' ----- > C5H11I + 31" (7) inorganic matrices such as 8102 or A1203, the reactivity of cyanation increases with decreasing polarity of the organic solvents47r49. The dependence of solvent polarity on the inorganic based triphase catalysts parallels that of organoclays. Because of the good swellability of the polymer-based catalysts in polar organic solvents, anionic nucleophiles are carried more easily to the ionic center of the polymer. Consequently, nucleophilic substitution reactions proceed easily in the polar organic medium. While organoclays swellability increases with organic solvent polarities, catalytic reactivity under the triphase catalysis does not improve. This is because the reaction site for an organoclay triphase catalysis is not located in the organic medium. Organoclays as catalysts for nucleophilic substitution reactions require pre-adsorption of the organic electrophiles on the organoclay surfaces. Highly polar organic solvents exhibit strong affinity for organic reactants. Also, adsorption of organic electrophiles on the organoclay is retarded by competitive adsorption of the polar organic solvent. Thus, the surface concentration of organic reactants will be lowered by increasing the polarity of the organic solvent and the catalytic reactivity of organo clays for triphase catalysis will decrease. 91 Table 5 Solvent effect on kobs for the iodonation reaction of pentyl bromide (Equation 8) under condition with pseudo first order in nucleophile Catalyst Solvent 104kobs. sec-1 H26AA o-Dichlorobenzene 0.81 H26AA Toluene 1.29 H26AA Decarle 3.42 Reaction condition: 8 mole of potassium iodide; 2 mole of pentyl bromide; 0.15 g decane as internal standard; 3 ml water and 6 mL organic solvent; 0.100g of H26AA; 90°C 92 4. Dependence of O/C Alkylation on Biphase and Triphase Catalysts There are two possible products for the alkylation of naphthoxide by benzyl bromide55o58 as shown in Figure 7. In protic solvents such as water, trifluoroacetic acid, methanol or ethanol, the C-alkylation product is predominant. C-alkylation is favored in protic solvents because the enolate of the C-alkylation precursor has an carbonyl group as shown in Figure 7 which can be stabilized by hydrogen bonding between the solvent and the intermediate anionic complex, whereas the negative charge is not favored to locate on the low electroaffmitive carbon. On the other hand, the O-alkylation product is favored in aprotic solvents, such as toluene, benzene, DMF, DMSO and dichloromethane. The enolate resonance structure of the O-alkylation precursor is energetically stable because the negative charge is located on the high electroaff'mitive oxygen. Since the aprotic solvent can not form a hydrogen bond with the hydroxy group of the potential intermediate, the intermediate favors the energetically stable form which will result in the O-alkylation product. The ratio of the two alkylation products in a triphase catalysis system can provide information about the location of the reaction. As shown in Table 6, the major products for the alkylation of naphthoxide by benzyl bromide with organo hectorites (H26AA and H24AA) as catalysts are C-alkylated derivatives. This suggests that the catalytic reaction occurs in the aqueous environment. In contrast to organoclays, nucleophilic substitution reactions for polymer supported 93 .040 .3 000003 00000000 320:: 05 0:0 0220510000 .00 00:00 00300305 05 mam—0:002 .3 00220300 203 02205 30:09 .0 020.0000 0000— 0000300 2:. .0 .022 :2 B 855.28 so; 8:0. 8:230 05 2F .0 .020 000002 .50: v .003 3030000 .00 3022 end 00.03 .00 _E n 5 02:05:00: 23000 .00 00028 M 200200 02020 _E n 5 0220.5 30:00 .00 302.: N ”00020000 000000“ .0 2-800- 2 0 02 0.00 4.00 0.00:0 6440: 2-80.0- 2 0 00.0 0.00 0..:. 0.00:0 58:22:06 2-800- 2 0 00.: 0.00 0.00 88:8 4.4.40: 2-80.0- S 0 00.0 0.04 0.00 8843 58:22:20 0.: 0.00 0.00:0 5022220 20 0.04 0.00:0 500: :0 0.04 88:3 8.25220 0.00 0.: 88:8 500: 822202800 08:20.: 280202: 802088022. -0 0o .0 -0 0o .0 .8200 20.200 00030000 000000.: 00000 05 00 000: 203 0:00 2:20 :00 003000020 00:3 02205 30:00 .23 02:05:00: 22000 .00 0000002 00:03:00 Us 0:. .00 0000. 000005 05. 0 030.: 94 catalysts proceed in organic instead of aqueous medium51.53v64. For alkylations with C16H33NMe3Br and C15H33PBu3Br as the biphase catalysts, O-alkylation process dominates the reactions (Table 6), since the biphase catalyst carries the nucleophile from the aqueous phase to the organic phase where the nucleophilic substitution reaction occurs. The O-alkylation percentage of the biphase catalysis is sensitive to organic solvent polarity. The ion pair of the O-alkylation intermediate and onium ion in polar organic solvents such as methylene chloride shows greater solubility than in less polar solvents such as benzene. The results is a higher O-alkylation percentage in methylene chloride (72%) than in benzene (60%) when C15H33NMe3Br is the biphase catalyst. However, for organo clay triphase catalysis, the O/C alkylation ratios are not significantly influenced by the polarity of organic solvent as determined by C15H33NMe3+hectorite (H24AA) as triphase catalyst. The O-alkylation percentages using H24AA as catalyst are 21% and 22% in benzene and methylene, respectively. For C15H33PBu3+hectorite (H26AA) as the triphase catalyst, the 0- alkylation yield is high when reacted in methylene chloride (49%) but low when in benzene (18%). Organoclays afford predominate by C-alkylation. This result is very unusual. Organoclays provide a unique selectivity for reaction products which can only be rationalized if the reaction occurs at the clay-aqueous interface. 9S 5 . The Dependence of Organoclay Catalytic Reactivity on the Volume of the two Liquid Phases A stirred mixture of the organoclay, aqueous and organic solutions forms a water in oil emulsion. Organo hectorite is homogeneously distributed in the emulsion and the water droplets are incorporated into the organic phase. When the volume of water is more than needed for an emulsion, a clear aqueous phase appears in the triphase system. A typical emulsion of 6 mL of organic liquid, 3 mL aqueous solution and 0.1g organo hectorite formed by stirring is milky white in appearance and no excess organic phase or aqueous phase is evidenced. A small excess organic liquid segregates from the emulsion when the mixture is allowed to remain for more than five minutes. As the volume of aqueous phase is increased higher than 3 mL and the volume of organic solution phase is decreased lower than 6 mL, a clear aqueous solution phase can be observed during stirring. The excess aqueous solution can not be incorporated into the emulsion of supported organo hectorite in the organic phase. Since the triphase system is a water-in-oil emulsion, a part of the aqueous phase does not participate the emulsion system. The activity of an organoclay triphase catalysis is only influenced by the volume of organic solution and not affected by the volume of aqueous solution (Figure 21). The catalytic activity decreases when the volume of the organic phase is increased while the concentration of reactants, as the total volume of aqueous and organic solution and the amount of organoclay catalyst, H26AA were kept constant (Figure 21a). Based on the same amount of organoclay catalyst, increasing the volume of the 96 Figure 21 (a) The dependence of km on the volume of toluene for the reaction of pentylbromide with KCN. 104%, Sec. 1 i 1 r u f u 2 3 >4 . s I 6 Volume of Organic Phase, m1. Reaction was carried out under a condition that was pseudo first order in electrophile. The total volume of organic and aqueous phase was 9 mL. The concentration of cyanide in aqueous solution was 6.25 M and the pentyl bromide concenu'ation was 0.33 M in toluene. 97 Figure 21 (Continued) (b) The observed rate plotted against the volume of aqueous solution while the volume of toluene solution was kept constant at 2 mL. 4.5 4.0 « q—t . .0 3.5 ‘ * l l F 0 U) ‘T 3.0 - O —1 q E 2.5 . 8 1 o 2.04 U o ‘5 1.5 « ad 1 E 1.0 '4 -° 0.5 4 O 0.0 f I ' r V I ' I a 4 5 e 7 Volume of Auqeous Phase, mL Reaction was carried out under a condition that was pseudo first order in electrophile. The concentration of cyanide in aqueous solution was 6.25 M and the pentyl bromide concentration was 0.33 M in toluene. 98 organic phase will decrease the suspended catalyst concentration in the organic solution. Thus, a low efficiency of adsorbing organic reactants on organoclay results in low catalytic reactivity when the volume of the organic phase is increased. However, when the volume of organic phase is kept constant, the catalytic reactivity is not apparently changed by changing the volume of aqueous solution (Figure 21b). Since only a certain amount of aqueous solution can be incorporated into the emulsion, the excess aqueous phase which can not be incorporated into the emulsion does not participate in the organoclay triphase catalysis. Therefore, the reactivity is not influenced by the volume of aqueous solution if the concentration of the nucleophile, CN‘, is kept constant. 6. Organoclay Catalytic Reaction in the Absence of Water. Nucleophilic substitution reactions are not catalyzed by the organoclay H26AA in the absence of water. For the cyanation of pentyl bromide, potassium cyanide can not be transferred by the organoclay from the solid phase into the toluene solution to react with the organic reactant. Potassium cyanide is hydrophilic, but it can be suspended in the organic solution. However, a phase transfer catalysis reaction is known to occur in the absence of water when a polymer supported catalyst or an onium salt is employed as the phase transfer catalyst62,99. A polymer supported catalyst can transfer the nucleophile by ion- exchanging the anion on the polymer-attached onium group. This onium-nucleophile ionic pair is accessible to the electrophile in organic solution for nucleophilic substitution reaction. An organoclay does not have an ionic center for ion-exchange reaction with a nucleophile. Instead, the organoclay facilitates the nucleophilic substitution by forming an emulsion mixture of aqueous and organic solution. The 99 nucleophile can not be transferred from the solid state into the organoclay. Therefore, if the nucleophile can not be dissolved in the liquid solution of the phase transfer catalysis reaction, catalytic reaction can not proceed. Consequently, water is required for catalytic nucleophilic substitution reactions in the presence of organoclay. 7. Mechanism of the Organoclay Triphase Catalysis The assembly of an organoclay emulsion is the decisive factor for the efficient triphase catalysis. Levine and Spence and co- workersloo,101 have investigated emulsions of organic and aqueous solution stabilized by fine clay particles containing adsorbed surfactant. In our organoclay triphase catalysis reaction system, the organoclays act as an emulsifier to minimize the particle size of the dispersion droplet (Figure 22). At the aqueous interface, the onium surfactant orient vertically on the surface to explore the hydrophilic silicate surface and increase hydrophilic interactionloz. At the organic liquid catalyst boundary, the surfactant may orient horizontally on the silicate surface to shield the polar clay surface and maintain the hydrophobicity of the material surface. For the purpose of forming an emulsion, the organo clay should have a high layer charge density and be interlayered by an onium ion with at least one long alkyl chain. Therefore, the inherent hydrophilic layered silicate will be modified to an amphiphilic material. Hectorite derivatives (H24AA and H26AA) with long chain alkyl ammonium can form emulsions and are good triphase catalysts. 100 Figure 22 The emulsion of aqueous and organic solutions stabilized by an amphiphilic organo hectorite. Membrane-like Organrc Pl Organo clay 86 Assembly E OrganicPhase g E WWWJ I fi JV 'WLWWW, WWW : WWWMMM :Sondqvvemgm'" ‘ E l 32 ~35 M' «we 101 However, not all organoclays are efficient triphase catalysts. Hectorite derivative, H18AA, with tetrabutyl ammonium in the interlayer, fails to form an emulsion and is a poor triphase catalyst. The onium ion in this case fails to shield the clay surface for wetting by the organic phase. Laponite derivatives do not form emulsions due to the low layer charge density of the layer silicate, which limits adsorption of onium ion on the silicate surface and the degree of hydrophobicity. Also, the inherently small particle size of laponite affords a large edge surface area which i can not be covered by the hydrophobic surfactant tail, and this also contributes to the hydrophilicity of the surface. The mechanism of efficient triphase catalysis by organo hectorites (H24AA and H26AA) as triphase catalysts is elucidated by the solvent effect and the dependence on reactant concentrations. Also, the experiment on O/C alkylation and the study of the volume ratio of the two liquid phases provides mechanistic information for the organoclay triphase catalysis reaction. All these results suggest that the reaction occurs at the clay-aqueous liquid boundary. Figure 23 shows the proposed mechanism for nucleophilic substitution reaction by organo hectorites as triphase catalysts. An organo hectorite particle which stabilizes the organic-aqueous emulsion is a part of the assembled organo clays. They behave as emulsifiers, or better, as membranes between the aqueous and organic phases. The catalyst swellability in organic solvents is not a significant factor for catalytic reactivity. 102 Figure 23 The mechanism of the triphase catalysis reaction with organo hectorites as the catalysts. Binding of RX occurs at the clay-organic interface. Organic Solution F®Q~3W WCH“ Mar IJW Clay Membrane | - Who Br Aqueous Solution «3%: Q CH3 Nucleophilic attack occurs at the clay-aqueous interface. Hl 103 However, the adsorption of organic electrophile on the organo clay surfaces and the concentration of anionic nucleophile in aqueous solution are the two most important factors for the catalytic reactivity of nucleophilic substitution reaction. Asorption of the organic molecule is occurring at the organic- aqueous interface. Solvents of low polarity produce high adsorption of the organic electrophile. Thus, the surface concentration of organic reactant is high and the catalytic reaction is more efficient when non- polar organic solvents are used in the triphase catalysis system. The nucleophilic substitution reaction occurs at the boundary of the catalyst and the aqueous solution. The nucleophile in the aqueous solution is directly reacting with the organic electrophile which has been adsorbed on the organoclay surface, since the reaction rate is proportional to the bulk concentration of the nucleophile in the aqueous solution. 104 C. The Longevity of Organoclays for Triphase Catalysis 1. Recyclability of Organoclays as Triphase Catalysts The structural stability and catalytic recyclability of organoclays for TPC are dependent on the layer charge densities of the silicate layer hosts, and on the solubilities of the ion-pair formed between nucleophile anions and surfactant cations in the organic solvents. In order to study stability and catalytic recyclability, organoclays containing either C15H33NMe3+ or C15H33PBu3+ as the intercalated cations were employed as the catalysts. The hectorite derivative, H26AA, containing the interlayer C15H33PBu3+ cation shows excellent recyclability for triphase catalysis (Figure 24). Another hectorite derivative, H24AA, with C15H33NMe3+ gallery cation shows fair catalytic recyclability (Figure 25) for the probe cyanation reaction (Equation 1). After 10 cycles of cyanation reaction, in which the total reaction time was 30 hours and the reaction temperature 90°C, both of the hectorite derivatives retained their d- spacings, as judged by the XRD. The loss of catalytic activity after use for vigorous cyanation reaction was approximately 10% and 55% for H26AA and H24AA respectively. The catalysts were recycled by simple filtration and resuspension in fresh reaction mixture. This recycle procedure could result in incomplete recovery of the solid catalyst and an apparent loss of reactivity. The loss of the catalytic reactivity of H26AA is most likely due to incomplete catalyst recovery by the filtration procedure. However, the 55% loss in activity for H24AA Figure 24 loskobs, Sec-l O-‘NQ‘hUIO’VGC 105 The recyclability of C16H33PBu3‘hectorite (H26AA) over 10 reaction cycles for the cyanation of pentyl bromide at 90°C. Reaction conditions for each cycle were the same as those described in Figure 19a. _s (n 1 _0 _L Q «ht l n I l _0 .0 d N 111 A _5 o l lllllll+llllllll A ll 'l O '0 fi $1 m .0 0 Number of catalytic cycles 106 Figure 25 The recyclability of C16H33NMe3+hectorite (H24AA) Rate, 105 x 10,500“) over 10 reaction cycles for the cyanation of pentyl bromide at 90°C. Reaction conditions for each cycle were the same as those described in Figure 19a. 4.0 ‘ 1.0-l 0.0 V V I W V I V V I V V ' V V I ff I V Y I V U I i r' I V V V T 012 3 4 5 6 7 a 91011 Number of catalytic cycle 107 after 10 reaction cycles can not be due to loss of catalyst by the filtration. Quaternary ammonium ions may undergo Hoffmann elimination, the result of which is production of trialkyl amines and alkenes in strongly alkaline solution“. For example, the interlayer surfactant, C16H33NMe3+, which contains b-hydrogens may decompose into an amine and alkene in the basic solution at elevated temperature35‘37. The structural stability of the extremely high layer charge density onium ion F-hectorite derivative F26AA, in which the surfactant adopts 'a paraffin-like structure, is almost negligible. More than 98% of the catalytic capacity is lost after one reaction cycle (Table 7). The X-ray pattern shows the catalyst is transformed to the potassium form of F- hectorite, an indication that the surfactant does not remain bound to the clay host. For C15H33NMe3+-F-hectorite F24AA which adopts a paraffm-like structure in the interlayer, the catalytic recyclability is satisfactory (Figure 26) but the solid catalyst is converted to a mixture of potassium F-hectorite and an interstratified form of F-hectorite in which the gallery are alternately occupied by potassium and surfactant cations. This regularly interstratified form of F-hectorite will be discussed in the later section. The difference in recyclability between F26AA and F24AA is due to the nature of the surfactants themselves. The carbon-rich surfactant, C16H33PBu3+ dissolves into the liquid phase but the other surfactant, C1¢5H33NMe3+ remains intercalated in the silicate host. This desorption mechanism will be discussed in more detail in the section D.3. 108 Table 7 The catalytic activity of the high layer charge density C15H33PBu3+-Clay after multiple reaction cycle for cyanation and chlorination of pentyl bromidea. Organoclay Organic Nucleophile Pseudo First Order Rate Solvent Constant (105 x kobs. See'l) Reaction Run 1 2 3 W26AA Toluene CN‘ 27.2 10.7 4.33 A26AA Toluene CN' 54.6 9.44 2.1 1 F26AA Toluene CN' 56.5 1.91 0.98 A26AA Decane CN' 71.8 49.6 F26AA Decane CN‘ 46.9 4.96 A26AA Toluene CI' 24.3 %b 22.7%b a. Reaction Conditions: 2 mole pentyl bromide; 20 mole potassium cyanide or 7.5 mmole sodium chloride; 3 mL aqueous solution; 6 mL organic solution; reaction temperature 90°C. b The reactivity is determined by the chemical yield of the chlorination product after 8 hour reaction. 109 Figure 26 The recyclability of C16H33NMe3+-F-hectorite (F24AA) Rate, 105xkob,(sec'l) over 10 reaction cycles for the cyanation of pentyl bromide at 90°C. Reaction conditions for each cycle were the same as those described in Figure 19a. 5 4-' 3. l .1 I- .I I 211 ' ' . + 1- Ottrvrrrrr' l *r'r 0 2 4 6 8 10 12 Number of catalytic cycles 110 The two montmorillonite derivatives, W26AA and A26AA, which contain C15H33NMe3+ and C15H33PBu3+ onium ions respectively, are between F-hectorite and hectorite supported catalysts in structural stability and catalytic recyclability (Table 7). Although the Wyoming montmorillonite derivative, W26AA, shows the same lateral bilayer surfactant structure as the hectorite derivative, H26AA, W26AA does not exhibit catalytic recyclability as good as that of H26AA. For the Arizona montmorillonite derivative (A26AA) which adopts a pseudo trimolecular structure, the catalytic recyclability for pentyl bromide cyanation is also poor (Table 7). The surfactant (C15H33PBu+) appears to desorb from the montnorillonite host after the phase transfer catalysis reaction. The extent of desorption of C1¢5H331PBu3+ from the clay hosts is correlated with the clay layer charge density. ' When the reaction conditions are changed, the recyclability properties of organoclays for PTC are also changed. The recycled C15H33PBu3+-Montmorillonite(Arizona) (A26AA) after pentyl bromide cyanation is a poor catalyst for the next cycle of the same reaction if toluene is employed as the organic solvent. However, if a non-polar solvent such as decane is used on a reaction solvent, the catalytic stability for cyanation can be greatly improved. The poor solubility of the surfactant-cyanide ion pair in the non-polar organic solvent undoubtedly contributes to the stability of the intercalate. This low solubility of the ion pair in the non-polar organic solvent can limit the desorption of the surfactant from the clay hosts. Also, the recyclability 111 of A26AA for catalytic chlorination is much better than for the cyanation reaction with the same solvent. The surfactant-chloride ion pair is not as soluble as surfactant-cyanide pair in the toluene solution. Therefore, the instability of the organoclay structure and the loss of catalytic recyclability is due to the desorption of the surfactants by ion pairing with the nucleophiles in the organic solvent. 2. Surfactant Desorption The desorption of onium ion surfactants from clay interlayers is the main cause of the loss of catalytic capacity for organoclay catalysts. This desorption process has been investigated by treating the organoclays under various condition and analyzing the concentration of the desorbed surfactants dissolved in liquid solution. (a) Influence of aqueous electrolyte solution on surfactant desorption of organoclays. Onium ion surfactants are expected to remain adsorbed on the clay hosts in the presence of concentrated electrolyte aqueous solution without an organic solvent. The affinity between the surfactant-anion pair and water is less than the electrostatic force between the surfactant and silicate layer. The results of the surfactant desorption for different organoclays in 2.5 N sodium chloride solution is shown in Table 8. The amount of surfactant desorption was determined by spectrometric analysis of the surfactant in liquid solution”. The cationic surfactants in the aqueous phase were extracted into the chloroform layer in the 112 Table 8. The desorption of onium ion from the organoclays in the presence of 2.5N NaCl aqueous solutiona. Organoclay onium ion emp. Original d-Spacing Change of Desorption 0C d-Spacing after Surfactant treatment Orientation with 2.5N after NaCl treatment with N aCl C161133NM¢3+Laponit¢ < 0.3% 250 14.51: 14.5: None 3 2 2)2 2 2/ 2 250% NaOH C8H17OH + HBr(47%) O’gmday *CanB' Organoclay Q c Bf ’ 01-3611, .011, Hz 50% NaOH (‘iibt—EL=Ern- .. .. 0 Or anocla 15% NaOH 0 N0: N0, (11) CH2[O(CH2)2CH(CH3)2]2(12) (13) (14) (15) (16) 139 Table 16. The chemical yields of the acetal synthesis (Equation 10-12)3. Entry R Group Catalyst Chemical Yield (%) 1 Benzyl H24AA >95 2 Benzyl H26AA >95 3 Benzyl A24AA >95 4 Benzyl A26AA >95 5 Benzyl C15H33NMe3Br >95 6 Benzyl C15H33PBu3Br >95 ,7 Benzyl No 18 8 Benzyl Recycled H24AA >95 9 Benzyl Recycled H24AA >95 10 Butyl H24AA 37 1 l Butyl H26AA 50 12 Butyl A24AA 63 13 Butyl A26AA 58 14 Butyl C15H33NMe3Br 55 15 Butyl C15H33PBu3Br 56 16 Butyl No 0 17 Butyl Recycled H24AA 35 18 Butyl recycled H26AA 49 19 iso-Pentyl H24AA 32 20 iso-Pentyl H26AA 28 21 ' iso-Pentyl A24AA 58 22 iso-Pentyl A26AA 38 23 iso-Pentyl C15H33NMe3Br 36 24 iso-Pentyl C16H33PBu3Br 33 25 iso-Pentyl No 0 26 iso-Pentyl Recycled H24AA 30 27 Iso-Pentyl recycled H26AA 24 a. Reaction condition: 10 mole of alcohol; 103 CHzBrz; 10g CH2C12; 10 mL 50% NaOH aqueous solution; reaction temperature, 90°C; reaction time 12 hours. Table 17 The chemical shifts in the 1H NMR of the starting materials and the reaction products of the acetal synthesis reaction (Equation 10-12). Chemical Chemical shift (ppm) C6H5CH20CH20CH2C6H5 W 4.87 ppm (s, 2H) 7.3-7.4 ppm (m, 10H) C4H90CH20C4H9 0.86-0.94 ppm (t, 6H) [(CH3)2CH(CH2)20]2CH2 C5H5CH20H C4H90H (CH3)2CH(CH2)20H 1.3-1.4 ppm (m, 4H) 1.45-1.55 ppm (m, 4H) 4.62 ppm (s, 2H) 0.90-0.92 ppm ((1, 12H) 1.40-1.51 ppm (q, 4H) 1.60-1.80ppm (m, 4H) - a 4.62 ppm (s, 2H) a 7.3-7.4 ppm (m, 5H) 0.86-0.94 ppm (t, 3H) 1.3-1.4 ppm (m, 2H) 1.45-1.55 ppm (m. 2H) W 0.90-0.92 ppm ((1, 6) 1.40-1.51 ppm (q, 2H) 1.60-1.80ppm (m, 2H) W a. The underline marks are the absorption peaks used for measuring the integration ratio of starting materials and reaction products. 141 and converted to smaller particles suspended in the aqueous phase. The trace amounts of recycled Arizona montmorillonite derivatives, A24AA and A26AA, do not retain their original structure following reaction, as evidenced by X-Ray diffraction. As is mentioned in this chapter, the surfactants in the high layer charge density smectite clay group, such as Arizona montmorillonite, will dissociate from the interlayer if the clay derivatives are suspended in electrolyte solutions. Since the surfactants do not coat the surface of the silicate layer, hydroxide ion can destroy the unprotected surface and decompoe the mineral. The surfactant in a relatively low layer charge density hectorite will not dissociate so the silicate layer surface is always protected by the surfactant. 2. Synthesis of Alkyl Bromide The long chain alkyl bromide can be synthesized by refluxing a mixture of alkyl alcohol in concentrated hydrobromic acid. Montanari et al.113 have utilized hexadecyltributyl phosphonium bromide as a phase transfer catalyst to facilitate this reaction. The organo clay (H26AA) is employed as the phase transfer catalyst for the alkyl bromide synthesis (Equation 13). The catalytic reaction conversions are better than the conversion for the blank reaction (Table 18). However, the organo clay is totally decomposed by the strong acid and the recycled solid material does not show any catalytic activity. The XRD pattern of the recycled material is amorphous, with the original d001 absorption peak of H26AA absent. The hectorite supported catalysts can tolerate alkaline solution but are destroyed by strongly acidic solution. Fortunately, most phase transfer catalysis reactions are carried out under neutral or basic condition. 142 Figure 30 The X-ray diffraction pattern of C16H33NMefhectorite (H24AA) after the catalysis reaction in which the reaction mixture contains 50% NaOH aqueous solution and 10 mL of methylene chloride solution. The peaks at 19A arises from H24AA. The 10A, 4 7.A and 2.4A peaks belong to the reactants. 1000 19A 4711 10A 400 (1 2AA Figure 31 143 The X-ray diffraction pattern of sodium hectorite treated with 50% NaOH aqueous solution. No reflection peak at 12.5A ccould be found. The reflection peaks at 4.7A and 2.4A are due sodium hydroxide. 4.7A 2.4.4 144 3. Dehydrohalogenation of 2-Bromoethylbenzene Styrene, which is a very important industrial chemical for polymerization, can be generated by treating 2-bromoethylbenzene in strong alkaline aqueous solution using triphase catalysts such as polyethylene glycol or organoclay (Equation 14). The chemical yields of the reaction are determined by the GLC. Commercial styrene from Aldrich Chemical Company shows the same retention time in the GLC as the product obtained in the catalytic reaction. The chemical yield of the styrene in the reaction catalyzed by H26AA and H24AA are shown 1 in Table 19. No styrene is formed in the absence of organo clay as triphase catalyst. The organo hectorite preserves its original layer structure after this catalytic reaction. 4 . Oxidation of trans-Stilbene Organo clays can facilitate the oxidation of trans-stilbene under the triphase catalysis condition (Equation 15). The product, benzoic acid, is identified by. its melting point and NMR spectrum. Table 20 shows that the chemical yields of the oxidation using organoclays as catalysts are much better than the blank reaction without a phase transfer catalyst. However, the organo clay catalysts are difficult to recycle due to the large amount of manganese dioxide precipitate mixed-in with the catalyst. 145 Table 18 The synthesis of octyl bromide from octanol by hydrobromic acid dehydrationa (Equation 13). Catalyst Concentration Reaction Chemical yieldb of HBr (%) Time C16H33PBu3+ 47% 11 Hours 79% Hectorite (H24AA) None 47% 11 Hours 37% C15H33PBu3+ 23% ' 11 Hours 0% Hectorite (H24AA) None 23% 11 Hours 0% a. Reaction Condition: 1 mL of 47% HBr or 2 mL or 23% of HBr; 2.5 mmole octanol; 0.060 mole catalyst; Reaction temperature 90°C. b. The chemical yields were detennined by GLC. 146 5 . Synthesis of 2,4-Dinitrophenyl Ether Organo clay triphase catalyst can also be employed for the synthesis of 2,4-dinitrophenyl ether by treating l-chloro-2,4-dinitrobenzene and phenol in 1.5N NaOH solution (Equation 16). The chemical yield of the nucleophilic substitution reaction is almost 100% (Table 21). The reaction product and chemical yield were determined by 1H NMR and 13C NMR (Figure 32). Tundo and Venturello47 have used silica gel supported onium ions as the triphase catalysts to synthesize 2,4- dinitrophenyl ether. However, the reaction occurs in strongly basic solution. Since organo hectorite is stable in alkaline solution, this hectorite derivative is a more suitable triphase catalyst for this synthesis reaction. 147 Table 19 Synthesis of styrene from 2-bromoethy1benzene under alkaline solutiona (Equation 14). Catalyst Reaction Time temperature Yieldb doo1(A)¢ C16H33PBu3+ 13 hours 65°C 73% 21 -Hectorite (H26AA) C15H33PBu3+ 13 hours 65°C 70% 18 -Hectorite (H24AA) C15H33PBu3+ 10 hours 75°C 76% 21 -Hectorite (H26AA) C151133PBu3+ 14 hours 75°C 83% 21 -Hectorite (H26AA) None 13 hours 65°C 9% a. Reaction Condition: 4 mole 2-bromoethylbenzene; 0.300g Decane (Internal Standard); 4 mL Benzene; 0.100g Organoclay; 3 mL 50% NaOH aqueous solution. b. The chemical yields were determined by GLC. c. The d-spacing of the recovered catalyst. 148 Table 20 The formation of benzoic acid by the oxidation of trans-stilbenea (Equation 15). Catalyst Reaction Time Yield (%)b C15H33PBu3+Hectorite 10 Hours 78 (H26AA) None 10 Hours 63 C16H33PBu3+Hectorite 18 Hours 94 (H26AA) None 18 Hours 81 a. Reaction Condition: 0.200 g H26AA; 3 mole trans-Stilbene; 6 mole KMnO4; 10 mL Benzene; 6 mL Water. b. The yields were obtained by weighing the isolated benzoic acid after the reactions. 149 Table 21 The synthesis of 2,4-dinitrophenyl ether from l-chloro-2,4-dinitrobenzene and phenolatea (Equation 16). Catalyst Reaction Time Chemical Yieldb (%) C1611133PBu3+ 4 Hours >95 Hectorite (H26AA) C15H33PBu3+Br 4 Hours >95 None 4 Hours 0 a. Reaction Condition: 6 mole phenol and 1-chloro-2,4- dinitrobenzene in 5 mL benzene; 6 mole sodium hydroxide in 5 mL of aqueous solution; 0.060 mole catalyst; reaction temperature 25°C. b Yields were determined by the 1H NMR integration area of starting material and reaction product (Figure 34). The chemical product, 2,4-dinitrophenyl ether, was also characterized by its melting point (68-7l°C). 150 Figure 32 The 1H NMR spectra of (a) 1-chloro-2,4-dinitrobenzene and (b) 2,4-dinitrophenyl ether. (a) \07” 1 gag? .ft/ *1 —7.2657 IYIIIIY'IIIIIIIIYIVVIIIvv'II'I'rIII'IIIIIIIII[ IIIIIIII 'Y'VT'I'I" IIIIIIII 'IIIIIVIvllvvvaIIIv'vIYI'IIIIIIIIIIIII; 90 88 86 8.4 8.2 80 7.8 76 7.4 7.2 70 68M ——1 Syd—rd W 007 4.10 l.“ 1.40 4.70 {k f ' 3 ' I I I ll . r 1 ‘ I l i ’— i 1 , ' M J ‘1’ ..M-_~ urlIIYr.tYII[ VVVVVVVVV rVYIYll'VVITI'IlYYYUI IIIIIIII IIIUVIIIVITIY'Y[IIIIIVY'Y'IVYYIVVYV.YYYVIV'V'] VVVVVV “I! R” 80 84 8d 3) 7.8 76 7.4 7.;.' 709914 0H CHAPTER IV CONCLUSIONS Organoclays provide different properties in triphase catalysis from those of typical polymer supported triphase catalysts. The mechanistic property of the clays in which the onium salts are immobilized on polystyrene is similar to that of onium ion biphase catalysis. In this mechanisml3'14’1‘5'17'61'62'l 14 the anionic nucleophile is paired with an onium cation in the aqueous phase followed by extraction into the organic phase to undergo the nucle0philic substitution reaction. With the polymer supported catalyst the reactant must undergo mass transfer to the catalyst surface, diffuse to the active site which is close to the region of the onium ion, proceed with chemical reaction, and the reaction product must leave the site. Any or all of these fundamental steps may limit the catalytic reactivity. Organo hectorite does not remove the nucleophile from the aqueous phase to the organic phase. The clay organic derivative provides an amphiphilic property which can stabilize a water in oil emulsion. Because the nucleophilic substitution reaction can only occur at the liquid-liquid interface in the absence of a phase transfer catalyst, formation of an 151 152 emulsion can dramatically increase the interfacial area of the oil and water phase resulting in a higher probability for chemical reaction. In organo hectorite triphase catalysis, only the organic reactant must undergo the process of mass transfer and diffuse to the active site on the catalyst surface. The nucleophile can freely transfer to the active site without the two limiting steps. This is evidenced by the linear relationship between the reaction rates and nucleophile concentration in bulk aqueous phase (Figure 19a). However, the stronger solvation of the nucleophile in aqueous solution results in a low nucle0philic ability for the anion. The activation energy of reactions catalyzed by organo hectorites is theoretically equal to that of the reaction proceeding in aqueous solution. The function of the organo hectorite is to increase reactant concentration in the active site. The nucleophilic substitution reaction catalyzed by polymer supported catalyst is carried out in the organic phase61’63'64’115. The function of the polystyrene supported catalyst is (a) to lower the activation energy of the reaction pathway and (b) to provide a phase with a high effective concentration of the potential reactantllé:1 17. The first function, which can be achieved by changing the chemical reaction environment from the aqueous to the organic phase, results in enhanced nucleophilic ability of the anionic nucleophile. The second can be fulfilled by increasing the lipophilic property of the polymer supported catalyst. The catalytic ability of polystyrene supported and free onium salts sometimes depend on the extraction coefficient of the nucleophile between the aqueous phase and the organic phase15'17. Hydrophobic anions such as SCN- and I- have high extraction coefficients. However, high hydrophilic nucleophiles such as hydroxide or chloride which have low extraction coefficients in these two liquid 153 solutions mediated by phase transfer catalysts, lead to a low effective concentration of the nucleophile at the reaction site. The typical phase transfer catalysts used for reactions that employ the hydroxide ion as the nucleophile are expensive crown ether, cryptates or derivatives of poly(ethylene glycol)31'42'55'57'118. These analogues can hydrogen bond with the hydroxide, and then be extracted into the organic phase. Organo hectorite is also a good triphase catalyst for reactions in this strongly basic system. The hydrophilic anion does not have to be transferred from the aqueous to the organic phase so the reaction is not limited by the extraction coefficient. The organic solvents in triphase catalysis systems also influence catalytic reactivity. In a triphase catalysis reaction with a polystyrene supported catalyst, use of non-polar organic 46:61. The extraction solvents results in low catalytic reactivity coefficients of anionic nucleophiles are low in non-polar organic solvents. In addition, the low swellability of the ion pair, onium cation and anionic nucleophile, in the non-polar organic solvent leads to difficulty with mass transfer and diffusion of reactants to the active site. Use of inexpensive and non-toxic non-polar organic solvents such as decane in organoclay triphase catalysis reactions can improve the chemical reactivity relative to that of more expensive polar organic solvents, such as toluene and o- dichlorobenzene. The low affinity of the non-polar organic solvent and the organic reactant results in ease of mass transfer of the organic electrophile from the bulk organic phase to the organoclay surface. Then the catalytic reactivity is high because of the high concentration of the organic reactant on the organoclay surface. Although an organoclay triphase catalyst does not significantly reduce the activation energy of the chemical reaction, the independence of the extraction coefficient of the 154 anionic nucle0phi1e provides the advantage that a larger variety of nucleophiles and non-polar organic solvents can be appropriately used in the triphase catalysis system. Another kind of triphase catalyst is that in which the onium ions are supported in inorganic matrices such as silica gel or a1umina47'48. Chemical reactions using these inorganic-based materials as triphase catalysts are believed to occur at the liquid-catalyst interface49'50. The catalytic reactivity of the inorganic-based material also increases with decreasing the polarity of organic solvent47'49’69. The advantage of these materials are their physical strength and the lack of swelling119. Smectite clays also have strong rigiditylzo. Organoclays swell in organic solvents. A high polar organic solvent, such as o- dichlorobenzene, dramatically swells organo hectorite. However, a high polar organic solvent strongly solvates the organic electrophile and limits mass transfer of the organic molecule with the result of poor catalytic activity. The swellability of polystyrene supported onium salts in organic solvents plays an important role in the catalytic reactivity of the triphase catalysis reaction. All other factors being equal, reaction rates decrease with increased cross-linking whenever rate is limited by intraparticle diffusion 58:59:64’67’119. The more highly cross-linked the catalyst, the lower is the diffusion of reactants to the active sites. However, the most common 1% and 2% cross-linked resin in solvent-swollen form is too gelatinous to be recovered by filtration1 19. Although catalytic activity is most studied in phase transfer catalysis, recyclability is an aspect in triphase catalysis that is sometimes the most important. Good polymer supported catalysts with great swellability in organic solvent may be difficult to recycle. High cross-linked polystyrene supported catalysts 155 more easily recovered but are lower in catalytic activity than those of low cross-linked polymers. Organoclay is easily recovered by filtration or centrifugation after the chemical reaction. The main disadvantage of organoclay for recycling is that some of the onium ions in the clay interlayer may desorb from clay hosts with high layer charge densities. Because the bond between the clay layer and the onium ion guests is electrostatic, metal cations from the aqueous phase may substitute for the onium ions. However, hectorite as the host overcomes the problem of onium ion desorption. The interlayer onium ions in this medium layer charge density clay host are not ion-exchanged by metal cations from the aqueous solution. This makes the hectorite derivative a good triphase catalyst in terms of longevity. The chemical instability of the triphase catalyst under reaction conditions also results in poor recyclability. Ammonium and phosphonium salts may be destroyed during the reaction. This is especially true for quaternary ammonium salts, which undergo Hoffmann elimination in strongly alkaline solution and produce trialkylamine and alkene products”. Most triphase catalysts have encountered this problemfior121 and so does organoclay. The catalytic activity of hexadecyltrimethyl ammonium hectorite gradually decays in each cyanation reaction. Before 50% of the catalytic activity of the organo hectorite is lost, more than 100 turnovers of the chemical reaction have been achieved. This organo hectorite is still valuable in this triphase catalysis cyanation reaction. Quaternary phosphonium is generally more stable than quaternary ammonium under reaction conditions”. Thus hexadecyltributyl phosphonium hectorite can preserve its catalytic capacity better than the quaternary ammonium hectorite. More than 300 156 turnovers of the reaction have been achieved while the hectorite supported phosphonium still maintains 85 % of the catalytic activity. The 15 % loss of reactivity is believed to caused by loss of the solid catalyst during reaction work-up. The hectorite matrix can sometimes reduce the chemical decomposition of the interlayer guests. For example, triphenylbenzyl phosphonium and benzaldehyde in hydroxide solution will undergo the Witting reaction to form trans-stilbenelzz. When the triphenylbenzyl phosphonium ion was intercalated into the hectorite interlayer, the reaction could not be observed under the same reaction Conditions. Therefore, hectorite can more or less prevent the interlayer guest from participating in the chemical reaction. Silica gel and alumina are unstable in hydroxide solution and this may limit the use of the inorganic-based catalysts for some chemical reaction. Silica gel is even unstable in cyanide solutionl 19. Sodium smectite clays including hectorite are also unstable in strongly basic solutions. However, as the interlayer sodium cations of hectorite are ion-exchanged by cationic surfactants, the material is durable in this vigorous reaction environment. This organo hectorite can be recovered after use in a triphase catalysis reaction with hydroxide as a reagent and the recycled catalyst is still active. Generally, the reactivity of triphase catalysis is less than that of biphase catalysis or homogeneous reaction with different polar aprotic compounds as solvents. The catalytic functional groups anchored on the macromolecule supports can not move as freely as the analog dissolved in liquid solution. Hexadecyltributyl phosphonium ion performs catalytic activity five times more than that of the similar phosphonium group supported on polystyrene61 . This phosphonium ion also shows four times 157 more catalytic activity than the organo hectorite which contains the same equivalent of phosphonium in the hectorite interlayer. An exceptional situation is that in which the hexadecyltrimethyl ammonium hectorite is more catalytically active than the free surfactant. This surfactant proceeds by micellar catalysislsrz‘uzg”124 which is slower than the typical biphase catalysis that uses the onium ions with four bulky alkyl groups attached on the nitrogen or phosphine. The organo hectorite undergoes a different mechanistic process and results in a faster catalytic reaction. Polar aprotic solvents such as DMF or THF are used typically for homogeneous reactions. These solvents have strong solvation for metal cations but weak affinity for anionic nucleophiles. They can reduce the shielding of electrons in anions and strengthen the nucleophilicity of the anionic reactants. However, these kinds of chemicals are usually expensive and difficult to remove after reaction and may present environmental problems in large scale operation. Triphase catalysts not only have the advantage of catalyst recycling, but sometimes provide various selectivities for chemical reaction, re giochemistry, or even stereochernistry. Polystyrene supported onium salts proceed by a mechanistic process similar to that of biphase catalysis. This may result in analogous products to those obtained by using these two materials as. the phase transfer catalysts. For example, the O- alkylation product (Figure 7) predominates in the alkylation reaction of benzyl bromide and naphthoxide when ether polymer supported or free onium salts are used as catalysts since the alkylation reaction“64 occurs at the organic environment. With organo hectorite the reactions proceed in the aqueous instead of the organic phase, so the C-alkylation product predominates in this catalytic alkylation reaction. Also, when the cationic 158 surfactant hexadecyltrimethyl ammonium bromide is used as the micellar catalyst for the cyanization of n-bromoalkanol or l-bromoalkane, the former reactant is much more reactive than the latter. However, as the cationic surfactant is intercalated into the hectorite interlayer and used as the triphase catalyst for the same cyanization reactions, the reactivities of both reactants are comparable. The reactant, n-bromoalkanol can change the micelle structure in micellar catalysis but l-bromoalkane can not. However, neither of the two reactants can alter the mechanistic structure of organo hectorite triphase catalysis. Clay supported or free chiral ammonium such as the (-)-benzquuininium ion can be also used as chiral catalysts for asymmetric synthesis reactions. Very interestingly, these two materials always tend to produce the predominate enantiomeric products with opposite configuration. In borohydride reduction reactions of alkylphenylketone125'126, epoxidation of chalcone127 and Michael thio- addition of 2-cyclohexenone128'129, the major enantiomeric products that results from the tWo chiral catalysts have opposite configurations. The orientations of the onium ion are different on hectorite and in liquid solution, leading to a different spatial constraint and resulting in opposite configurational enantiomers. Moreover, clay-supported and free chiral onium ions proceed, via asymmetric synthesis reactions in two different liquid phases. Organic hectorite proceeds by phase transfer catalysis reactions in aqueous environment but free onium ions carry out the analogous reactions in the organic phase. The difference of reaction environment may also provide a different selectivity in the asymmetric synthesis. 10. 11. l2. 13. REFERENCES G. W. Brindley and G. Brown, Eds.,"Crystal Structures of Clay Minerals and Their X-Ray Identification", Mineralogical Society, London, 1980. D. M. Moore and R. C. Reynolds, Jr " X-Ray Diffraction and the Identificationand and Analysis of Clay Minerals" Oxford University Press, New York, 1989. J. Guan, E. Min and Z. Yu, Proc. 9th Int. Congr. Cata., Calgary Canada, Vol. 1, Edited by M. Phillips and M. Teman, Chem. Inst. Canada, Ottawa, 104 (1988). R. K. Marcelja, R. M. Pashley and J. P. Quirk, J. Phys. Chem. 22, 6468 ( 1988). J. J. Spitzer, Langmuir 3, 199 (1989). P. Richmond, B. W. Ninham, J. Colloid. and Interface Sci. 40, 406 (1972). H. Van Olphen,"An Introduction to Clay Colloid Chemistry" 2nd ed., Wiley, New York, 1977. F. Figueras, Catal. Rev. Sci. Eng. 30, 457 (1988). J. M. Adams, Appl. Clay Sci. 2, 309 (1987). M. Frenkel, Clays Clay Miner. 22, 435 (1974). G. Porcelet and A. Schultz, NATO ASI Ser., Ser. C: Chemical Reaction in Organic and Inorganic Constrained Systems, 165 (1986). B. M. G. Theng "The chemistry of Clay-Organic Reactions" Wiley, New York, 1974. C. M. Stark and R. M. Owens, J. Am. Chem. Soc. 23, 195 (1971). 159 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 26'. 27. 28. 29. 30. 31. 32. 33. 160 C. M. Stark and R. M. Owens, J. Am. Chem. Soc. 25, 3613 (1973). E. V. Dehmlow and S. S. Dehmlow, " Phase Transfer Catalysis", 2nd ed., Verlog Chemie, Weinheirn, 1983. C. M. Stark and C. Liotta, "Phase Transfer Catalysis, Principles and Techniques", Academic press, New York, 1978. A. Brandstonn, Adv. Phys. Org. Chem. 13, 267 (1977). J. M. Brown, Colloid Sci. 3, 253 (1979). J. M. Brown, P. a. Chaloner and A. Colens, J. Chem. Soc. Perkin Trans. 2, 71 (1979). J. M. Brown and J. A. Jenkins, J. Chem. Soc. Chem. Commun. 458 (1976). J. M. Brown, S. K. Baker, A. Colens, J. R. Darwent and D. Parrins, "Enzyrnic and Non-Enzymic Catalysis" Edited by P. Dunnill and A. Wiseman and N. Blackbrough, E. Horwood, 111, New York,. 1980. J. H. Brown and J. R. Jarwent, J. Chem. Soc. Chem. Commun. 169 (1979). E. H. Cordes and R. Bruce Dunlap, Accounts Chem. Res. 2, 329 (1969). E. J. Fendler and J. H. Fendler, Adv. Phys. Org. Chem. 18, 271 ( 1970). D. Langevin, Acc. Chem. Res. 21, 255 (1988). W. Stoeckenius, J. H. Schulman, L. M. Prince, Kalloid Z. 169, 170 (1960). J. E. Bowcott and J. H. Schulman, Z. Electrochem. 283 (1955). J. S. Cuo, E. D. Sudol, J. W. Vanderhoff, H. J. Yue and M. S. E1- Aasser, J. Colloid. and Interface Sci. 142, 184 (1992). E. Ruckestein and J. C. Chi, J. Chem. Soc. Faraday Trans. 2, 1690 (1975) M. J. Rosen, ACS Syrnp. Ser. 311, Phenomena in Mixed Surfactant Systems, 144 (1986). V. V. Grushin, I. S. Akhren and M. E. Vol'pin, J. Organomet. Chem. 321, 403 (1989). S. H. Korzeniowski and G. W. Gokel, Tetrahedron Lett. 12, 1637 (1977). S. H. Korzeniowski and G. W. Gokel, Tetrahedron Lett. 41), 3519 (1977). 34. 35. 36. 37. 38. 39. 41. 42. 43. 45. 47. 4s. 49. 50. 51. 52. 53. 54. 55. 56. 161 G. W. Gokel and D. J. Cram, J. Chem. Soc. Chem. Commun, 418 (1973). G. W. Gokel, M. F. Ahern, J. R. Beadle, L. Blum, S. H. Korzeniowski, A. Leopold and D. E. Rosenberg, Isr. J. Chem 26, 270 (1985). R. M. Izatt and F. F. Cristenseu, "Synthesis Multidentate Macrocyclic Compoun " Academic Press, New York, 197 8. N. A Gibson and J. W. Hosking, Aust. J. Chem. 18, 123 (1965). E. Laurent, R. Rauniyar and M. Tomalla, J. Appl. Electrochem. 14, 741(1984);15_, 121 (1985) Z. Goren, L Willner and H. Taniguchi, J. Org. Chem. 42, 4755 (1984). D. C. Sherrington, Macromol. Chem. (London) 3, 303 (1984). S. L. Regen, Angew. Chem. Int. Ed. Engl. 18, 421 (1979). G. E. Totten and N. A. Clinton, J. Macromol. Sci., Rev. Macromol. Chem. Phys. (223, 293 (1988). S. L. Regen, J. Am. Chem. Soc. 28, 6270 (1976). S. L. Regen, J. Org. Chem. .42 875 (1977). S. L. Regen and J. J. Besse. J. Am. Chem. Soc. 193, 4059 (1079). S. L. Regen, J. C. K. Heh and J. McLick, J. Org. Chem. 451,, 1961 (1979). P. Tundo and P. Venturello, J. Am. Chem. Soc. 191, 6606 (1979). P. T‘undo, P. Venturello and E. Angeletti, J. Am. Chem. Soc. 104, 6551 (1982). P. Tundo and M. Badiali, React. Polym. 19,55 (1989). P. T‘undo and P. Venturello, J. Am. Chem. Soc. 193, 856 (1981). A. Kadahodayan and T. J. Pinnavaia, J. Mol. Catal. 21, 109 (1983). B. M. Choudary, Y. V. Subbo Rao and B. P. Prasad, Clays Clay mimer. 32, 329 (1991). C. Lin, T. Lee and T. J. Pinnavaia, ACS Symposium Series 522, Supramolecular Architecture, 145 (1992). A. Comelis, P. Laszlo and P. Pennetreau, Clays Clay Miner. 18, 437 (1983). Y. Kimura and S. L. Regen, J. Org. Chem. 58, 195 (1983). D. E. Bergbreiter and J. B. Blanton, J. Org. Chem. 59, 5828 (1985). 57. 58. 59. 61. 62. 63. 65. 67 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 162 N. Yamazaki, A. Hirao and S. Nakahama, J. Macromol. Sci. Chem. A13, 321 (1979). M. Tomoi and W. T. Ford, J. Am. Chem. Soc. 193, 3828 (1981). W. T. Ford, J. Lee and M. Tomoi, Macromolecules 13, 1246 (1982). M. Tomoi and W. T. Ford, J. Am Chem. Soc. 193, 3821 (1981). F. Montanari, S. Quici and P. Tundo, J. Org. Chem. 48, 199 (1983). D. Landini, A. Mari and F. Montanari, J. Am. Chem. Soc. 199, 2796 (1978). S. L. Regen, Nouv. J. Chim. 6, 629 (1982). N. Ohtani, C. A. Wilkie, A. Nigam and S. L. Regen, Macromolecules 15, 516 (1981). Q. A. Reutov and A. LiKurts, Russ. Chem. Rev. (Engl. Trans), 46, 1964 (1977). F. Montanari and P. Tundo, J. Org. Chem. 46,, 2125 (1981). P. L. Anelli, F. Montanari and S. Quici, J. Chem. Soc. Perkin Trans. 2, 1827 (1983). ‘ N. Komblum, R. Seltzer and P. Haberfield, J. Am. Chem. Soc. 83, 1148 (1963). P. Tundo, P. Venturello and E. Angeletti, Isr. J. Chem. 26, 283 (1985). Y. Okahara, H-J. Lirn, G. Nakarnura and S. Hachiya, J. Am. Chem. Soc. 193, 4855 (1983). G. Lagaly, Solid State Ionics 22, 43 (1986). A. Weiss, Angew. Chem. Int. Ed. Engl. 2, 134 (1963). A. Justo, C. Maqueda, J. L. Perez-Rodriquez and G. Lagaly, Clay Miner. 22, 319 (1987). J. K. Thomas, Acc. Chem. Res. 21, 275 (1988) R. A. DellaGuardia and J. K. Thomas, J. Phys. Chem. 81, 3550 (1983) T. Nakamura and J. K. Thomas, J. Phys. Chem. 29, 641 (1986) J. K. Th0mas, J. Phys. Chem. 21, 267 (1987) S. A. Boyd, W. A. Jaynes and B. S. Ross " Organic Substances and Sedrnents in Water" edited by R. Baker, CRS Press, Baco Raton, Florida, 181 (1991). 79. 81. 82. 83. 85. 86. 87. 88. 89. 91'. 92. 93. 94. 95. 96. 97. 163 W. F. Jaynes and S. Boyd, Soil Sci. Soc. Am. J. 33, 43 (1991) W. T. Ford, Adv. Polym. Sci. 24, 201 (1984). E. V. Dehmlowand S. S. Dehmlow "Phase Transfer Catalysis" Verlog Chernie Winheim, P65 (1983). E. Ruiz-Hitzky and H. Casal, NATO ASI Ser. Ser. C 163: Chemical Reaction in Organic and Inorganic Constrained Systems, 213 (1986). D. M. Clementz and M. M. Mortland, Clays Clay miner. 22, 223 (1974). R. Greene-Kelly, J. Soil Sci. 4, 233 (1953). 1 D. A. Archer, H. Booth and R. D. Stangroom, J. Chem. Soc. 322 (1969). a T. Kametani, K. Kigasawa, M. Hiiragi, N. Wagatsuma and K. Wakisaka, Tetrahedron Lett. 8, 635 (1969). J. T. Bunus and K. T. Leefek, Can. J. Chem. 41, 3725 (1969). , B. Dobias, ACS Symp. Ser. 311, Phenomena in Mixed Surfactant Systems, 216 (1986). O. Abillon, D. Chatenay, D. Langevin and J. Meunier, ' Surfactants Solution 2, Edited by L. L. Mittle and B. Lindman, 1159 (1984). B. Zhu, G. Zhao and J. Cui, ACS Symp. Ser. 311, Phenomena in Mixed Surfactant Systems, 173 (1986). J. C. Brackrnan, J. B. F. N. Engbert, J. Colloid. and Interface Sci. 132, 250 ( 1989). D. O. Shah" Surface Phenomena in Enhanced oil Recovery" Plenum, New York (1981). S. Landau Ph. D. Thesis " Physical and Catalytic Properties of I-Irdroxy-Metal Intercalated Smectite Minerals" Chemistry Department, Michigan State University, 1985. J. M. Millsand M. A Zwarich, Clays Clay Miner. 29, 169 (1972). Chi-Li Lin, Ph. D. Thesis " Organoclays as Triphase Catalysts" Chemistry Department, Michigan State University, (1988) J. Butrulle and T. J. Pinnavaia " Characterization of Catalytic Material" 1. E. Wachs Ed. Butterworth-Heinemann, Boston pp. 149-163 (1992) P. Chuit, Helv. Chim. Acta. 2, 264 (1926) 98. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 164 L. K. Wang and D. Langley, Arch. Environ. Contamin Toxicol. 3, 447 (1977). A. Nag, S. K. Sarkar, S. S. Palit and S. Pandey, Indian J. Chem., Set. B 283, 64 (1989). J. R. Spence and J. H. Masliyah, Can. J. Chem. Eng. 61, 924 (1989). S. Levine and E. Sanford, Can. J. Chem. Eng. 62, 258 (1985). G. Lagaly, R. Witter and H. Sander, Adsorption from Solution, Symp. Academic Press, New York, 65 ( 1983). P. Mukerjee and K. J. Mysels, J. Am. Chem. Soc. 71, 2937 (1955). I. M. Klotz, R. K. Burkhard and J. M. Wrouhart, J. Am. Chem. Soc. 36, 77 (1952). Y. Hayarni and K. Motomura, ACS Symp. Ser.311, Phenomena in Mixed Surfactant Systems, 312 (1986). M. Gobbo, R. Fomaseir and U. Tonellato, Surfactants Solution, 2, Edited by L. L. Mittle and B. Lindman, Plenum, New York, 1169 (1984). J. M. Brown, J. L. Lynn, Jr, Ber. Bunsen-Ges. Phys. Chem. 85, 95 (1980). J. M. Brown, C. A. Bunton, S. Diaz and Y. lhala, J. Org. Chem. 45, 4169 (9180). R. W. Juriand R. A. Bartsch, J. Org. Chem. 55, 143 (1979). G. Manecke, A. Kramer, H. Winter and R. Reuter, Nouv. J. Chim. 6, 623 (1982). A. Comelis, P. Laszlo and P. Pemretreau, Clays Clay Miner. 18, 437 (1983). A. Cornelius and P. Laszlo, Synthesis, 162 (1982). D. Landini, F. Montanari and F. Rolla, Synthesis, 771 (1978). R. C. F. Jones, General Synthetic Method 1, 402 ( 1978). H. Molinari and F. Montanari, J. Chem. Soc. Chem. Commun, 639 (1977). S. L .Regen, J. J. Beese and J. Mclick, J. Am. Chem. Soc. 191, 116 (1979). S. L. Regen and L. Dulak, J. Am. Chem. Soc. 22, 623 (1977). Y. Kimura and S. L. Regen, J. Org Chem. 48, 195 (1983). W. T. Ford, CHEMTECH, 436 (1984). 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. 165 T. J. Pinnavaia and H. Kim, Zeolite Microporous: Synthesis, Structure and Reactivity, Edited by E. G. Derouane, Kluiver Acedamic Publishers, New Zetherland, 79, (1990), H. J -M. Dou, R. Gallo and P. Hassanaly and J. Metzger, J. Org. Chem. 4.2, 4275 (1977). W. Tagaki, T. Inoue, Y. Yano and T. Okonogi, Tetrahedron Lett. 39, 2587 (1974) L. J. Winters and E. Grunwald, J. Am. Chem. Soc. 81, 4608 (1965). K. G. van Senden and C. Koingsberger, Tetrahedron 22, 1301 (1965). S. Julia, A. Ginebreda, J. Guixer, J. Masana, A. Tomas and S. Colonna, J. Chem. Soc. Perkin Trans. 1, 547 (1981) S. Colonna, R. Fomasier and C. N. R. Centro, J. Chem. Soc. Perkin Trans. 1, 371 (1978) R. Helder, J. C. Hummelen, R. W. P. M. Laane, J. S. Wiering and H. Wynberg, Tetrahedron Lett. 21, 1831 (1976) S. Colonna, A. Re and H. Wynberg, J. Chem. Soc. Perkin Trans. 1, 547 (1981) H. Hiemstra and H. Wynberg, J. Am. Chem. Soc. 193, 417 (1981) 16 u11111131111111111111