l v. 21.1 a .munfiwvrlo .9 "u...ed....na..v¢._e.v....il . - ‘l‘il'blu’lr— (.oil‘.’ Janna!" ‘1‘: it... v 7'. "H"... ”t... n “U“ uvvnhpnu‘ EHJMDJRP?! vnnlt fir.“ .. 10?: pl»... 0!. 3 .u , ’c' E «5131' e '7 W92'F1‘» U 'a 1;" 23 'ufi'v :‘3 he): Iva-11.": Ant?! .3 I 2 1.! pr . ‘33. .‘3 in; H: . W" I. D. -‘Ilfittuilél‘ 3...? 1|!" 1| »DOD\$|. til-t F.\, 11. . "H o C '1 IA. DID? t0 : . -,.‘ 2 ’o”".'c"' ." e' 0”,4‘ n,¢§vf ‘2 .. 1. -. . 1 .b.:f..u.1 A 44 7 .- V a. -233.-. .. . ... 4.3.1.. nary . .2315...“ :1 . .Il , p . , . . f‘lllrhn I 3 .aa 12.1»? an‘i. .17. Y; , L . ‘9r41.",1&.v.r-|0.| . . . Us" I: 1.. u . . . . ‘.!.I.I.y V.Avl"..l!lyb 4 ... ‘ . :...‘...-u., {’11:le : 1.... . x... 113.... . .. 3‘9... It: xtln... 17.1. I. 10......th 4E. . . . .. . .1..i..€t-.1..0 :. : ‘.3 . .v.. I . . v t! ‘ Hogvl .71. L itllu‘-‘ I I. A f .1... £0) 5...: 9.1.5.... .I. ‘ I o».)ill“l..h'l. ‘ I. .f ‘l —”°v 1:3..' I cll . . x..f.§..l. ‘ . A ‘ T. .Plu . liluv. .‘ri v | , b "Y‘u Ir. 0 . ' ‘ 1| w -t u<..n . . 1.3.3.1. nu... .: .. ~15... ‘ . _ . 2. 1 f . ..H.L .v. . . _ .uans llv fil“... Lu I. .vl’c Jvoo ciihn. :11! . ‘ t ‘ . . 94:. 1-01.‘ V‘l.(1u....o...olu.u... Jan ”5&7"? .v. y 0 . V —J . I'll! b.Y..t....lo....ovr§1; .1. . . cl .1! 03.]. n.‘ . y . . .i 1 o. wanri|qt5txzz \ ulv. 0. “.4 . .hal-"léujll‘flfl55w .‘4 . , ‘u .. - ' ‘ I'll- l, I II‘ [vilg _ It! - I i llllllll“Nullllllllllll \llllllllllll 31 29300 f \ LIBRARY Michigan State University K I This is to certify that the dissertation entitled ORGANOCLAY AS TRIPHASE CATALYSTS presented by Chi-Li Lin has been accepted towards fulfillment of the requirements for Ph.D. degree in Inorganic Chemistry %a/M/ fiat”: 0" ‘ W j r professor Date Aug 15 1988 MS U is an Affirmative Action/Equal Opportunity Institution 042771 PLACE N RETURN BOXto movomochockunfrom your mood. TOAVOD FlNESMunonorbduoddodn. DATE DUE DATE DUE DATE DUE IIEQ __J [5% MSU In An Ailinndlvo ActioNEqunl Opponmiiy Institution Walla-9.1 ORGANOCLAY AS TRIPHASE CATALYSTS By Chi-Li Lin A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1988 ABSTRACT ORGANOCLAY AS TRIPHASE CATALYSTS By Chi-Li Lin A new type of triphase catalysts has been synthesized by intercalation of quaternary onium cations in smectite clays. These clay-supported phase transfer catalysts, as well as other triphase catalysts, are easily recovered by filtration and ideally suited for continuous flow methods. However, the onium-ion functionalized smectite clays have certain advantages over conventional tn'phase catalysts: i) they are stable under reaction conditions and provide efficient interfacial surfaces for the catalysis of organic reactions. ii) their catalytic prOperties are comparable to analogous biphase catalysts. iii) they can avoid using polar solvent in the catalytic process. The mechanistic investigation of this organoclay catalysis for alkylbromide displacement has indicated that the reaction mixtures form an oil-in-water emulsion in which the organoclay plays an essential role in stabilizing the emulsion. Three fundamental kinetic steps of the organoclay catalysis have also been proposed. In additional, the organoclays have been applied to a variety of useful synthetic transformations in which the organoclays were shown to be more efficient than the analogous polymer-supported phase transfer catalysts. The catalytic properties of several commercial organoclays and organo-layered compounds were also studied in comparison with those of our organoclay. Furthermore, the intercalation of chiral phase transfer cations in smectites has developed a new series of triphase chiral catalysts. The stereoselectivity and mechanism of these organoclays in the borohydride reduction of ketones and in the epoxidation of chalcone were studied. Although a great deal of research needs to be. done before the kinetic steps of clay-supported chiral catalysis can be proposed. It has to be emphasized that the clay-supported chiral catalysts are pracncally superior to conventional biphase catalysts. Moreover, for a given chiral cation we are able to alter its Optical selectivity by intercalating the cation on various clay hosts. TO MY FAMILY iv Acknowledgments I would like to thank Dr. 'I'J. Pinnavaia for his guidance and support though my graduate study and to Dr. CH. Brubaker for his guidance in editing this dissertation and Dr. N. Jackson and CK. Chung for helpful discussions. My special gratitude is extended to my family for their encouraging support. I also would like to thank all past and present group members for the friendship. TABLE OF CONTENTS Chapter Page LIST OF TABLES ....................................... VII LIST OF FIGURES ........................................ X CHAPTER I. INTRODUCTION A. Structure and Properties of Layered Silicates ....................... 1 1. Structure .......................................... 1 2. Ion Exchange ....................................... 3 3. Swelling and Catalytic Properties of Smectites ....................................... 3 B. The Principles of Phase Transfer Catalyst ......................... 9 1. The Nature of Phase Transfer Catalyst ........................ 9 2. Mechanism of Phase Transfer Catalysis ...................... 10 3. Applications of Phase Transfer Catalysis ...................... 12 4. Systems Related to Phase Transfer Catalysis ................... 12 a) Reactions at the interface ............................ 12 b) Micelle-Catalyzed Reactions .......................... 12 C. Some AspeCts of Chiral Phase Transfer Catalysis ................... 13 1. Nature of Chiral Phase Transfer Catalysis ..................... l3 2. Mechanism ....................................... 13 a) Epoxidation of Elecrron.Poor Olefins ..................... 19 b) Borohydride Reductions ............................. 19 3. Other Reactions .................................... 22 D. Development of Triphase Catalysis ........................... 22 1. Nature of Triphase Catalysis ............................. 22 2. Polymer. Supported Phase Transfer Catalysis ................... 32 a) Mechanism .................................... 35 b) Applications of Polymer-Supported Phase Transfer Catalysts ................................ 3S 3. Inorganic.Based Catalysis .............................. 35 a) Synthesis, Structure Characteristics and Behavior of Inorganic.Based Phase Transfer Catalysts ................................ 36 b) Mechanism and Applications .......................... 38 4. Polymer-Supported Chiral Phase Transfer Catalysts ................................... 38 a) Nature of Polymer-Supported Chiral Phase Transfer Catalysts ............................ 38 b) Mechanism and Applications .......................... 39 E. Research Objectives ..................................... 4 4 vi Chapter Page CHAPTER II EXPERIMENTAL A. Material ............................................ 46 1. Natural Sodium Hectorite ............................... 46 2. Laponite R ....................................... 46 3. Fluorohectorite ..................................... 47 4. Reagents ......................................... 4 7 B. Preparation of Organoclay Phase Transfer Catalysts ........................................... 48 C. Preparation of Clay-Supported Chiral catalyst ..................... 48 D. Organoclay Phase Transfer Catalysis ........................... 4 9 E. The Applications of Organoclay Phase Transfer Catalyst 1n Organic Synthesis ......................... 50 1. Synthesis of Ethers .................................. 50 2. Oxidation of Benzyl Alcohol ............................. 50 3. Synthesis of n. Pentyl Thiocyanate ......................... .51 4. Synthesis of Dipentyl Sulfide ............................ 51 5. C-Alkylation of Benzyl Cyanide .......................... .51 6. Dehalogenation of vis. Dibromides ......................... 51 7. Halogen Exchange ReaCtion ............................. 52 8. Blank Reactions and Control Reactions ...................... 52 F. Comparison of the Catalytic Pr0perties of Organoclays ......................................... 52 G. Comparison of the Catalytic Properties of . Some Organo-Layered Compounds ........................... 53 H. Asymmetric Reduction of Ketones ............................ 53 I. Asymmetric Epoxidation of Chalcones ......................... 54 J. The Determination of Enantiomeric Excess of Reaction by 1H NMR .................................. 55 K. Physical Measurements .................................. .55 1. Infrared Spectroscopy ................................. 55 2. X-ray Diffraction Studies ........................ . ....... 55 3. Gas Chromatography ................................. 56 4. Proton NMR Spectroscopy .............................. 56 5. Specific Rotation .................................... 56 6. Orbial Shaking Water Baths ............................. 56 vii Chapter Page CHAPTER III RESULTS AND DISCUSSION A. Preparation of Organoclay Phase Transfer Catalysts ........................................... 57 B. The Catalytic Properties of Organoclays 1n Alkylbromide Displacement by Cyanide Ion ...................... 63 The Dependence of Catalytic Reactivity on the Loading of Onium Cation ............................ 64 2. The Dependence of Catalytic ReaCtivity on the Molecular StruCture of the Onium Cation ................... 69 3. The Dependence of Catalytic Reactivity on the Molecular Structure of the Organic Substrates ................... 7 S 4. The Dependence of Catalytic Reactivity on the Concentration of Substrates and Catalysts ............. ' ........ 75 5. The Dependence of Catalytic Reactivity on the Layer Charge of Clay Host ............................ 76 6. The Dependence of Catalytic Reactivity on the Polarity of Organic Solvent ......... L ................. 84 7. The Mechanism of Organoclay Catalysis in Alkylbromide Displacement by Cyanide Ion .................. 88 C. Organic Synthesis Catalyzed by Organoclay ...................... 88 1. Synthesis of Aromatic Ether from Phenol ..................... 91 2. Oxidation of Alcohols ................................. 94 3. Synthesis of Alkyl Thiocyanate ........................... 94 4. Synthesis of Sulfide .................................. 94 5. C—alkylation of Nitriles ................................ 99 6. Dehalogenation of vic.Dibromides ......................... 100 7. Halogen Exchange .................................. 100 D. Comparison of the Catalytic Properties of Organoclay ......................................... 100 E. Comparison of the Catalytic Properties of Some Organo-Layered Compounds ........................... 108 F. Preparation of Clay-Supported Chiral Catalysts .................... 112 G. The Stereoselectivity and Mechanism of Organoclay in the Asymmetric Reduction of Ketones ....................... 112 l. The Dependence of Enantiomeric Excess on the Layer Charge of Clay Host ........................... 115 2. The Dependence of Enantiomeric Excess on the Molecular Structure of Chiral Cation and Organic Substrates ................................ 121 3. The Dependence of Enantiomeric Excess on the Concentration of Catalyst and Substrates ................... 123 viii Chapter Page 4. The Dependence of Enantiomeric Excess on the Temperature .................................. 130 5. The Dependence of Enantiomeric Excess on the Polarity of Organic Solvent ........................... 131 6. Conclusion ....................................... 131 H. The Stereoselecdvity and Mechanism of Organoclay in Asymmetric Epoxidation of Chalcones ....................... 132 1. The Dependence of Enantiomeric Excess on the Layer Charge of Clay Host ........................... 135 2 The Dependence of Enantiomeric Excess on the Temperature .................................... 136 3 The Dependence of Enantiomeric Excess on the Polarity of Organic Solvent ........................... 143 4. The Dependence of Enantiomeric Excess on the Concentration of Chalcone ............................. 144 5 The Dependence of Enantiomeric Excess on the Volume Ratio of Aqueous and Organic Phases ................. 144 6 Conclusion ....................................... 145 CHAPTER IV RECOMMENDATIONS .......................... 151 ix LIST OF TABLES TABLE Page 1. Products Formed by Ion Exchange Reaction of Hexadecyltrimethylammonium Chloride with Na+-Hectorite and Their Activity for Triphase Catalysts ..................................... 71 2. Produc1s Formed by Ion Exchange Reac1ion of Tetrabutylammonium Chloride with N a+—Hectorite and Their Acrivity for Triphase Catalysts ..................................... 72 3. Homoionic Organohectorites as Phase Transfer Catalysts for the Displacement Reaction of Alkylbromide with Cyanide ........................ 73 4. The Dependence of kobs Values on Alkyl Chain Length for Conversion of . Alkylbromides to Cyanide ................................ 77 5. Dependence of kobs Values on Potassium Cyanide Concentration for Conversion of Pentylbromide to Pentylcyanide ............................. 78 6. Dependence of kobs Values on Organoclay Concentration for Conversion of Pentylbromide to Pentylcyanide ............................. 81 7. Dependence of kobs Values on the Clay Host for Conversion of Pentylbromide to Pentylcyanide .................... 85 8. Dependence of kobs Values on the Polarity of Organic Solvent for Conversion of PentylbromidetoPentylcyanide. . . . . . . . . . . . . . . . . . . . .. ....... 86 9. Synthesis of Phenyl n-Pentyl Ether ........................... 95 10. Oxidation of Benzyl Alcoho ............................... 96 11. Synthesis of n-Pentyl Thiocyanate ........................... 97 12. Synthesis of n-Dipentyl Sulfide ............................. 98 13. C-Allrylation of N itriles ................................. 101 14. Dehelogenation of vie-Dibromides .......................... 102 TABLE Page 15. Synthesis of Alkylchloride from Alkylbromide ................... 103 16. Infrared Data and 1H NMR Data for the Products by Organoclay Triphase Catalysis ........................... 105 17. Catalytic Reactivities of Commercial Organoclay in the Synthesis of n-Pentylcyanide from n-Pentylbromide under Triphase Reaction Conditions .......................... 109 18. Catalytic Reacrivities of Organolayered Compounds for the Synthesis of n-Pentylchloride from n-Pentylbromide under Triphase Catalysis Conditions ............................. 110 19. X-ray Basal Spacings Data of Clay-supported Chiral Catalysts ...................................... 113 20. Infrared Data and 1H NMR Data for Benzyl t-Butyl Ketone .................................. 116 21. The Blank Reactions of Borohydride Reduction of Butyl Phenyl Ketone .......................... 120 22. Asymmetric Borohydride Reduction of Phenyl Ketones in the Presence of Clay-Supported Chiral Catalysts ............................ 124 23. The Dependence of Excess Enantiomer on the Concentration of Clay-Supported Chiral Catalysts in Asymmetric Borohydride Reduction of Ketones .................................. 129 24. The Dependence of Enantiomeric Excess on Temperature in Asymmetric Borohydride Reduction of Ketones with Clay-Supported Chiral Catalysts ...................................... 133 25. The Dependence of Enantiomeric Excess on the Polarity of Organic Solvent in Asymmetric Borohydride Reduction of Ketones with Clay-Supported-Chiral Catalysts ......................... 134 26. Dependence of Enantiomeric Excess on the Molecular Structure of Chiral Catalysts for Asymmetric Epoxidation of Chalcone with Clay-Supported Chiral Catalysts ......................... 138 27. The Blank Reaction of Epoxidation of Chalcone .................. 139 xi TABLE Page 28. Dependence of Enantiomeric Excess on the Layer Charge of the Clay Host for Asymmetric Epoxidation of Chalcone with Clay-Supported Chiral Catalysts ............................ 142 29. Dependence of Enantiomeric Excess on the . emperature for Asymmetric Epoxidation of Chalcone with Clay-Supported Chiral Catalysts ................... 146 30. Dependence of Enantiomeric Excess on the Polarity of Organic Solvent for Asymmetric Epoxidation of Chalcone ........................ 149 31. Dependence of Enantiomeric Excess on the Concentration of Substrates and Volume Ratio of Aqueous and Organic Phase for Asymmetric Epoxidation of Chalcone with Clay-Supported Chiral Catalysts ............................ 150 xii | 1C LIST OF FIGURES FIGURE 10. The schematic structure of a 2 : l diocmhedral aluminum silicate minerals .............. The schematic structure of a 2 : 1 trioctahedral magnesium silicate minerals ............. Smectite clays classification ..................... The mechanism of phase transfer catalysis with 8N2 displacement reactions ................... Chiral phase transfer catalysts : salts of quinine (a-f), salts of ephedrine (g-j), 8 S. 9 R-quinine (k), 8 S, 9 R-cinchonidine (l), 8 R, 9 S-quinidine (m) and 8 R, 9 S-cichonine (n) ..... h ....... Preferred conformation of benzquuinium cation (1) Preferred conformation of benzquuinium cation about C4-C9 bond (A and B) Preferred conformation of benzquuinium cation about C3-C9 bond (C and D) (63, 64) ............ The transition states for asymmetric epoxidation of cyclohexenones catalyzed by (-) N-benzquuininium cation (63) ................ Asymmetric methylation of 6,7-dichloro- 5-methoxy-2-pheny1-l-indanone catalyzed by 8-R, 9-S, N—(p-trifluoromethyl benzyl) cinchonium bromide catalysts (the catalyst structure is in the Figure 5f) (73) ................... Asymmetric alkylation of 2-carbomethoxy- l-indanone catalyzed by chiral crown ethers (67) and quinine (The catalyst structure is in Figure 51:) (63, 74) ........................ Michael addition of thiols to a B unsaturated ketones catalyzed by cinchonidine (The catalyst Structure is in Figure 51) (63. 75) .......................... xiii Page ........... 4 ........... 6 ........... 8 ..... 11 ......... 14 ......... 17 ......... 20 ......... 23 ......... 24 ......... 26 FIGURE Page 11. 2+2 Cycloaddition reaCtions catalyzed by quinidine (The catalyst structure is in Figure 5k) (63, 76) ................................. 27 12. Asymmetric epoxidation of electron-poor olefins catalyzed by benzyl quininium chloride (The catalysts structure is in figure 5a) (63, 83-69. 77-78) .............................. 28 13. The process of triphase catalysis for 8N2 displacement reacrions ............................... 30 14. The mechanism of polymer..supported phase transfer catalysis : a) mass transfer or reactant form bulk liquid to the surface of the catalyst particle b) diffusion of reactant through the polymer matrix to the active sites. c) intrinsic reaction at the active sites, d) diffusion of product through the polymer matrix to the particle surface and e) mass transfer of product from the surface of the catalyst to the bulk solution. A Phosphonium substituted catalyst is used for illustratioon here, but this basic mechanism can apply to most polymer-supported-supported PT C’s ................. 33 15. Alkyl-functionalization of silica gel and alumina ........................................ 37 16. The strucmre of polymer..bound alkaloids (AC) and polymeric amines (DE) (48) ........................ 40 17. The mechanism of Michael addition catalyzed by polymer-bound cinchonal alkaloids (48) ...................... 42 18. Schematic representation of onium cation intercalation in smectite clay .............................. .58 19. Arrangements of alkylarnmonium ions in the layer silicates with different loading of the onium cation. 3) very low cation loading b) medium cation loading c) high cation loading d) homoionic compound ........................ 60 20. A combination system of sodium hectorite and homoionic organoclay ................................ 62 xiv FIGURE Page 21. The three..component system toluene/ water/ hectorite supported [C16H33PBu3]+ is presented by mass ratio at room temperature. 0 , emulsion;A, wet solid pastem, gel;o , suspension; I, emulsion with extra toluene ............................. 65 22. A proposed model for the uniform organoclay colloid emulsion formed under triphase reaction conditions ............................... 67 23. Dependence of kobs values on potassium cyanide concentration for the conversion of pentylbromide to pentylcyanide catalyzed by [C16H33PBu3]+-hectorite ........................ 79 24. Dependence of kobs values on organoclay concentration for conversion of pentylbromide to pentylcyanide catalyzed by [C 16H33PBu3]+-hectorite .............................. 82 25. The fundamental kinetic steps of the organoclay in the alkylbromide displacement by cyandie ion : i) the organic substrates are attracted to the alkyl chain of catalysts by hydrOphobic interaction. and the inorganic anions are attracred to the polar end of catalysts by electrostatic interactions ii) displacement reaction occur in the emulsion phase. iii) the products are transferred from the emulsion phase back to the organic phase or aqueous phase ........................... 89 26. The 1H NMR Spectrum of n..butyl phenyl alcohol (0.06 M in CDC13) in presence of various amount of chemical shift reagent Eu(hfc)3. The molar ratios of chemical shift reagent and the alcohol are 0. 0.32, 0.43, 0.52 and 0.87 in (a)-(e), respectively ...................... 117 27. Configuration of (-) Epoxychalcone .......................... 137 28. The possible conformation of transition state epoxidation of the chalcone catalyzed by 151AA ........................... 140 29. The hypOthesis of the dependence of the reaction rate on the temperature A) Epoxidation of chalcone by F151AA B) Epoxidation of chalcone by 15 1AA ....................... 147 XV Chapter I INTRODUCTION A. Structure and Properties of Layered Silicates 1. Structure The layered silicate described in this dissertation belongs to a Class of clay minerals known as smectite. The term" clay minerals" refers to specific structural types silicate with particle size less than 2 11m and with a definite stoichiometry and crystalline structure. Recent advances in X-ray crystallography and structural analysis have affirmed that completely ordered clay structures represent no more than ideal models (1,2). Smectites are composed of units made up of two silica tetrahedra sheets and a central octahedral sheet of magnesia or alumina. The silicate tetrahedral 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 silica sheet and the oxygens of the octahedral sheet form a common layer. Smectite clays are 2:1 layer minerals which are divided broadly into two structures : dioctahedral aluminum silicate minerals (Figure 1) and trioctahedral, magnesium 2 silicate rrrinerals (Figure 2). These can subdivided into groups in which the layer charge arises predominantly from isomorphous substitution in the octahedral layer and from substitution in tetrahedral layer. Further subdivisions can be made on the basis of layer charge density (Figure 3). Cations at particular locations of the silicate structure can be replaced by various Other cations with similar ionic radii without changing the structural characteristics of the mineral. If the replacing cation has a lower valence, a net negative charge will develop. Charge neutrality is achieved either by an opposing substitution with cations of higher valence or, as is usually the case, by the presence of additional cations, normally, arrays of hydrated alkaline earth or alkali metal cations in the interlayer region of the structure. The charge balancing cations are usually located adjacent to the points of anionic charge on the basal planes. However, small anhydrous cations, mainly H+ or Li+, can migrate through the oxygen sheet to the neighborhood of the substitution, where the anionic charge arises. Hectorite, Laponite and Fluorohectorite are the three smectites of principal interest in the present work. The idealized unit cell composition for hectorite is Mo.67[Mgs.33Li0.67lV1(Si8.o)1V020(0H.F)4 in which the superscripts (IV) and (VI) refer to the respective cation in tetrahedral and octahedral sites, and M represents a univalent or equivalent compensating cation (1). The commercial filler, Laponite is a synthetic low-charge hectorite (about 0.4 e- per Ozo[OH,F]4) which, unlike the natural mineral, can be obtained with a negligible iron content. Fluorohectorite is anOther synthetic high-charge smectite (about 1.8 e- per 020F4) in which the octahedral lattice hydroxyl groups have been replaced with fluoride ions. The idealized unit cell composition for Laponite and Fluorohectorite are Lio.36[LiO.36M85.64](Si8.00)020(0H)4 Lil.6[Li1.6Mg4.4](5i8.00)020F4 Interlamellar spacings in general depend upon the species of exchangeable cation present, the nature of solvent, and whether or not any electrolytes are present. 2. Ion Exchange The hydrated compensating cations of minerals are exchangeable. However, there is a size limitation for ion replacement. Hectorite, for example, exhibits an average distance about 8.7 A between exchange sites based on the calculation of a cation exchange capacity (73 milliequivalents per 100 g of air dried clay) and_the a and b unit cell parameter of the mineral (5.25 x 9.18 A) (3). Thus, cations with cross-section diameters greater than this value will fill the interlamellar spaces before an homoionic clay is achieved. And although the interlamellar surface is very large (750 mZ/g), the size of the exchanging ion can be a limiting factor in determining ion loading. The kinetics and equilibria of the exchange reaction depend on several variables (4,5). In general the exchange equilibrium favors a) cations with higher valence charge; b) the larger cation between species of a particular valence. Certain cations because of their size can take up a very favorable interlayer position. The rate of exchange reaction is determined mainly by the swelling properties of the mineral, the nature and concentration of cations. In smectite the position of exchangeable cations are about 80 percent on the basal plane surfaces, with the remainder on the edges. 3. Swelling and Catalytic Properties of Smectites Smectite clays can be swelled by adsorption of water or organic solvent. With multiple layers of solvents the galleries become liquid-like, and accessible for chemical reactions. The osmotic swelling of smectite clays is limited by the electrostatic interactions between the anionic silicate layers and compensating interlayer cations. As the solvent content increases the basal spacings of smectite clays also increase. The Figure 1. The schematic structure of a 2 : 1 dioctahedral aluminum silicate mineral lnterloyer Region AAI OO ©OH 'Si Figure 2. The schematic structure of a 2 : l trioctahedral magnesium silicate mineral lnterloyer Region AMQ OSi 00 ©OH SMECTITE CLAYS DIOCTAHEDRAL TRIOCTAHEDRAL MONTMORI LLIONITE HECTORITE . BEIDELITE - SAPONITE NONTRONITE Figure 3 Smectite clays classification swelling and increases in basal spacing on treatment with solvent not only can be used for identification of these minerals but also plays a very important role in catalytic reaction as well. In the last few decades, smectite clays and their derivatives have shown catalytic activity in numerous reactions (68). Most of the reactions make use of the acidic nature of cation-exchanged or acid treated clay. Both Lewis and Bronsted activity have been noted, the former deriving from aluminum or iron species located at crystal edges (9). The Bronsted activity, however, results either from free acid or from the dissociation of interlayer water molecules coordinated to polarizing interlayer exchangeable cations. Catalytic organic reactions have often been shown to take place in the interlamellar space. We should expect, therefore, to find instances of unusual factors in these reactions since they occur in a region of high acidity and possibly, in a restricted 2D reaction space. B. The Principles of Phase Transfer Catalyst 1. The Nature of Phase Transfer Catalyst Chemists frequently encounter the problem of bringing together two mutually insoluble reagents in sufficient concentration to attain conveniently rapid reaction rates. The traditional procedure would involve dissolving the reactants in a homogeneous medium. But a suitable solvent is not always available and it is usually expensive and difficult to remove after reaction and may 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 of low polarity (10-12). The basic function of the phase transfer catalyst is to transfer ions, free radicals, neutral molecules, or even energy (in a chemical form) from one phase to another. It is clear that PTC has considerable advantages over conventional procedures since it can: a) remove the requirement for expensive anhydrous or aptotic solvents, b) 10 improve reaction rates, c) lower reaction temperatures, d) simplify work-up e) permit reactions that do not proceed using conventional methods, 1') modify selectivity, and g) increase yields. Reactions involving phase transfer phenomena were performed since 1913 (13). A considerable number of such reactions are buried in olderliterature (14-16) and especially in patents (17-29) in which quaternary ammonium or phosphonium salts were used as catalysts for two phase reactions. Since some of the authors entered the field more or less incidentally and did not reflect on the mechanisms involved in such catalytic reactions, few realized the potential and scope of the new technique. PT C techniques as we know them today originated in the work of Makosza and coworker in 1965 (30) and the term "Phase Transfer Catalysis" was coined by Starks and first used in patents in 1968 (11). Although PT C has emerged only in the last decade it already has become a widely useful synthetic technique. 2. Mechanism of Phase Transfer Catalysis The mechanism of phase transfer catalysis varies to some extent with the type of system involved (neutral, acidic or basic condition). All phase transfer catalyzed reactions involve at least two steps: a) transfer of one reagent from its "normal" into the second phase; and b) reaction of the transferred reagent with the nontransferred reagent. For the simplest 8N2 displacement reactions the mechanistic pathway is illustrated in Figure 4 (31, 32). In this Figure Q+ represents an onium salt cation which forms an ion pair with anion Y‘ in the aqueous phase and extracts the anion into the organic phase. Once the anion has been transferred into the organic phase it reacts with RX and forms a new salt [Q+X-]. The new salt [Q+X-] then returns to the aqueous phase, where Q+ picks a new Y- ion for the next cycle. It is necessary to consider the mechanism by which the anion is transferred when we study the system of phase transfer catalysis. There are several possible mechanisms for the phase transfer step (33) : a) simple ion exchange across the 11 23:23. 505323... «zm 5.2. 31.3.8 .832. 3a.... .0 ...».cagoo... 2; .e 932“. omega Amaoosc+21||I .X+2 ... ->+O III' ......................... k. . among .2590. -X+O ... >mtlll. xm ... .>+O 12 interface; b) transfer of Q+ back and forth across the interface with anion exchange in the aqueous phase; c) transfer of the inorganic salts into the organic phase for exchange d) transfer of anion at the organic-crystalline solid interface (at SL systems) e) formation of micelles in the; aqueous phase and transfer of anions across the micelle interface. PT C as it being defined is not based on any mechanism. Hence, extensive investigations are need to elucidate the mechanism of any PT C systems. 3. Applications of Phase Transfer Catalysis Since the discovery of phase transfer catalysis, numerous applications have been described (32—37), not only in organic chemistry, but also in inorganic chemistry (38), analytical applications (39), electrochemistry (40, 41), photochemistry (42, 43), and especially in polymer chemistry (44—50). In organic synthesis PT C has emerged as a broadly useful tool for : a) nucleophilic displacement reactions b) alkylation and condensation reactions c) reactions of dihalocarbenes and other carbenes d) ylide-mediated reactions e) oxidation and reduction reactions f) and various miscellaneous reactions. In those reactions the technique of phase transfer catalysis provides a method which can avoid the use of polar aprotic solvent and also improve the reaction rate. Although phase transfer catalysis has suffered difficulties of separation at some later stage it appears to have high potential along synthetic lines. 4. Systems Related to Phase Transfer Catalysis :1) Reactions at the interface (51) Reactions occurring at an interface tend to be rate limited by the amount of interfacial area available and are therefore highly sensitive to the amount of agitation and concentration of reactant species at the interface. b) Micelle-catalyzed reactions (52-54) Surfactant molecules possessing well defined diphilic properties usually form small aggregations of 10-50 organic molecules dispersed in the aqueous phase. These small aggregations are called micelles, wherein the nonpolar organic part of the molecules occupy the internal volume while the highly 13 polar groups of the surfactant occupy the outer surface. The positively charged outer surface attracts and concentrates anions from the bulk aqueous solution into a counter ion layer near the surface of'the micelle and reaction may occur at the micelle surface. C. Some Aspect of Chiral Phase Transfer Catalysis In the recent literature three different approaches have been used for asymmetric induction : chiral reagents (55, 56), chiral auxiliaries (57-62) and chiral catalysts (63-65). The disadvantage of the first two methods with respect to an economical process is that both are used in stoichiometric amounts and have to be recycled. Chiral phase transfer catalysts offer a potentially simple, one step solution to this problem. 1. Nature of Chiral Phase Transfer Catalysis Phase transfer catalysis involves ion pairs with a rather close association. Therefore with strongly orienting chiral catalyst cations it is possible to obtain significant asymmetric induction. To date virtually all the work has employed onium salts based on ephedrine and its relatives, the cinchona alkaloids and crown ethers (Figure 5). Some confusion still remains in the literature concerning the effectiveness of these catalysts (66). However, the recent results reported by Cram (67) show a great optical yield ( ~99%). Two generalizations seem to be emerging in the area of phase transfer chiral catalysis. An hydroxy substituent B to the quaternary ammonium center of the chiral catalyst is required for significant enantioselectivity. The enantiomeric excess increases with higher dilution of organic substrates, lower temperature, faster agitation and less polar solvents. The concentration of catalyst controls the rate of reaction but to have little effect on the enantiomeric excess. 2. Mechanism In order to establish a reasonable mechanism of asymmetric induction it is necessary to determine the preferred conformations of chiral phase transfer catalysts in solution . At present, the accurate determination of ground-state conformations of the 14 Figure 5. Chiral phase transfer catalysts : salts of quinine (a-f), salts of ephedrine (g-j), 8 S, 9 R-quinine (k), 8 S, 9 R- cinchonidine (l), 8 R, 9 S-quinidine (m) and 8 R, 9 S- cichonine (n). 15 nzo ...:aoazo azaoazo «I Pm Tu «C F: .x V‘\ I ":0. / K foam. ...... ma 1 :0 nuoqruo ”zoo . 8 foo o 2.120.410 «:00 a ”10...... foo o N02.2.0 nzoo .. nzao £00 a N... ... ..a ... .00 N... ... \ m .. ... H J. z 16 .”:oo ”:oo I _. ..... who .../2 a a . m I I_ ... 0 ... «U. ...: L 17 Figure 6. Preferred conformation of benzquuinium cation (I). Preferred conformation of benzquuinium cation about C4- C9 bond (A and B). Preferred conformation of benzquuinium cation about C3-C9 bond (C and D) (63, 64) OH ...»... H H o C. H H .6 N. 19 chiral phase transfer catalysts in solution is rarely feasible. Nevertheless, several preferred conformations of chiral catalysis in the transition state have been proposed by Wynberg (63) and Dolling (64) (Figure 6). In the following the mechanism of two particular reactions catalyzed by cinchona alkaloids will be discussed. a) Epoxidation of Electron-Poor Olefins So far the enantiomeric excesses (ee) obtained by using chiral phase transfer catalysts in epoxidation reactions have not been exceedingly high, except in a few cases (68,69). The reaction, however has a large scope and furnishes optically active epoxides inaccessible by other routes. Because of the lack of kinetic data and the complexity of phase transfer asymmetric catalysis it is very difficult to propose a mechanism. Wynberg has presented a hypothetical mechanism for the cyclohexenone epoxidation catalyzed by chalcone (63). By using a preferred conformation of the benzquuininium salt (Figure 6), assuming a tight ion pair between the catalytic cation and the peroxide anion with the hydrogen bonding between the hydroxyl of the alkaloid and the carbonyl oxygen of the ketone he predicted a transition structure fitting the description in Figure 7. When t-butyl hydroperoxide is the reagent, it forrrrs an ion pair with the quinuclidine nitrogen of the alkaloid in a fairly compact structure. This t- BuOO' quininium complex allows the approach of cyclohexenone or chalcone from one side and favors one of two possible enantiomers only. b) Borohydride Reductions There are a few references on borohydride reductions of ketones with chiral catalysts, mainly ephedrine and the cinchona alkaloids (68-71). These limited results imply that the catalysts decomposed under the reaction conditions (66, 72), however the decomposition is slow. No detailed mechanisms have been proposed. But most research groups recognize that the borohydride anion will form an ion pair with the catalyst cation, and the interaction of the B hydroxyl of the catalyst with the ketone will favor one of two 20 Figure 7. The transition states for asymmetric epoxidation of cyclohexenones catalyzed by (-) N-benzquuininium cation 63). 21 22 possible diastereomeric transition states. Further investigations are required before a clear picture emerges in this system. 3. Other Reactions The use of optically resolved catalysts for the direct synthesis of enantiomerically pure compounds remains one of the dreams of the chemist. It would seem that chiral phase transfer catalysts are particularly well adapted to this problem. Recently, Wynberg and Dehmlow have written detailed reviews of the applications of chiral phase transfer catalysis (34, 63). Several successful applications of chiral phase transfer catalysts in asymmetric reactions have been investigated. Dolling and his co-workers (73) demonstrated the methylation of a substituted indanone in 95% yield and 92% ee by using 8-R, 9-S, N-(p- trifiuoromethylbenzyl) cinchonium bromide as catalyst (Figure 8). For the Michael reaction, Wynberg (63, 74) found an 87% yield with an cc of 76% by adding methyl vinyl ketone (MVK) to 2-carbomethoxy-l-indanone catalyzed by quinine (Figure 9). Later studies by Cram showed that ee’s approaching 99% could be obtained using chiral crown ethers as catalyst (67). Similarly Michael additions of thiols to a,b unsaturated ketones catalyzed by cinchonidine obtained relatively high ee’s (Figure 10) (63, 75). The wide range of enones and thiols potentially amenable to this reaction and the versatility of the sulfur and carbonyl functionality (Figure 10) make this reaction useful in many ways. In Figure 11 it shows that a variety of aldehydes were able to react with ketene to form the corresponding b-lactones in excellent chemical and nearly quantitative enantiomeric yields (63, 76). Also chalcones (63, 77, 78), quinones (79-82) and cyclohexenones (83) are epoxidized successfully to optically active epoxyketones and epoxyquinones by 30% hydrogen peroxide,t-buty1hydroperoxide, and 28% sodium hypochlorite (Figure 12). 23 cr 7) CI H NaOH, Solvent .- CH X, PTC 01130 3 cr 0 Cl CH3 ee : 92 °/o c1130 Figure 8. Asymmetric methylation of 6,7-dichloro-S-methoxy-Z-phenyl-1- indanone catalyzed by 8-R, 9-S, N-(p-trifluoromethylbenzyl) cinchonium bromide catalysts (the catalyst structure is in the Figure 5f) (73). 24 Figure 9. Asymmetric alkylation of 2-carbomethoxy-1-indanone catalyzed by chiral crown ethers (67) and quinine (The catalyst structure is in Figure 5k) (63, 74). 25 $2. .xbm "zoo. :ouazo __ o 20 m3. magi; __ 559.0 ..mm P EEO nzouoo VS... azoaoo \ o ¥>2 26 O 0 R Clnehonidlne R SH . / ; 3' Solvent 0 R! R S RI Figure 10. Michael addition of thiols to a B unsaturated ketones catalyzed by cinchonidine (The catalyst structure is in Figure 51) (63, 75). 27 0 fl Quinidine II I Q—CHZ - c - c1 * CCI3CH - Toluene -4o°c o // QC" -<= l | H - c — o ccr3 enantiomeric excess > 95% Figure 11. 2+2 Cycloaddition reactions catalyzed by quinidine (The catalyst structure is in Figure 5k) (63, 76) 28 Figure 12. Asymmetric epoxidation of electron- poor olefins-catalyzed by benzyl quininium chloride (The catalysts structure is in figure 5a) (63, 83-69, 77-78) 29 OH I .... b , O I R none 0‘ b i ll 0 R4 Hz a. Q mo eonc “3 II o b BONC : benzquulnlnlum chloride : 30% H202 10% NaOH, or t-BuOOH, or 28% NaOCI 30 Figure 13. The process of triphase catalysis for 3N2 displacement reactions. 31 E . ..x xx + ..., >m 030m -x xx ..> U_Z20 02 m ..e.ee.8_.-+.mm=e.ozmeze 38505 0 ... ... use; .8 ...m 5 025.8 833.80.. 0.088 0.0m ”0:030. .8 0 ... 00.80.00.800 0.088 0.~ .00.... 0.03 08030. 083000.. 2F .0000 .0 0.8... 0800008800 0... 0. 008000.003 ..0 80.80800 00.. 88800 0.8 .080 .8... 0000.... w. 3.0 0.0. 00.0. 0.00. - 00m .25. 02.0 0.... 00.m 0.0m - 00m 320 mm. .0 0.0. 00.N 0.0m - 00m E 0m. .0 0.... 00.m - 0.0m 00m <<0 mm . .0 N0. 00.~ - 0.0N 00m <6. 00. . - 0.0. 00m <. m 0x 00 00:00:09.9 h 030,—. 86 Table 8 Dependence of kobs Values on the Polarity of Organic Solvent for Conversion of Pentylbromide to Pentylcyanide?! Organic Solvent kobs. hr-1 Chlorobenzene 0.254 Toluene 0.575 Decane 0.950 a Reaction conditions: 2.0 mmole pentyl bromide in 6 ml toluene; 20.0 mmole potassium cyanide in 3 ml water; 0.10 g [Bu3PC16H33]+- hectorite; 900C. 87 Three groups of solvents have been classified in this respect : 1) Polar protic solvents readily solvate both anions and cations and give a high degree of dissociation of ion pairs. 2) Polar aprotic solvents readily solvate cations but not anion. Ion pairs are highly dissociated in such solvents. 3) Nonpolar aprotic solvents exhibit good solubilities toward onium salts but negligible solubility of inorganic salts . In these solvents the dominant species are ion pairs and the interaction of ion pairs and solvent is weak. In PT C reactions it has been found that nonpolar and dipolar aprotic solvents function well. In general, reactions are faster in the solvents of greater dielectric constant. In polymer-based triphase catalyst systems the nature of the organic solvent will affect the catalytic activity for different reasons. Regen has proposed that the organic solvent can determine the absorption of reactant on the polymer and control the extent of swelling, which in turn will detemrine the number of active sites available for catalysis, and establish the nature of the microenvironment at the active site, thus affecting the free energy of activation. In the reaction of an alkylbromide with aqueous sodium cyanide catalyzed by polystyrene-bound phase transfer catalysts, Regen’s work has shown a modest dependence of rate on the type of organic solvent used (89), and Ford has demonstrated a reaction rate decreasing with solvent in the order Chlorobenzene > toluene > decane over wide ranges of particle sizes and polymer cross-linking (95). An analogous reaction catalyzed by the hectorite-supported [C16H33PBu3]+ showed the opposite dependence of catalytic reactivity on the polarity of solvent (Table 8). The pseudo first rate constant increased as the polarity of the solvent decreased (Chlorobenzene < toluene < decane). By using the reaction model shown in Figure 22, we can predict that the kinetics of bromide displacement by cyanide occurring in organoclay systems are determined partially by the hydrophobic interactions between the emulsion phase and the organic substrates (alkylbromide), and partially by the 88 electrostatic interactions between the the emulsion phase and inorganic anion (cyanide anion). Increasing the polarity of organic solvent utilized will increase the interaction between the organic solvent and organic substrates and consequently weaken the interactions between the emulsion phase and the organic substrates. This will decrease the partition of the organic substrates in the reactive center and decrease the catalytic reacrivity. 7. The Mechanism of Organoclay in Alkylbromide Displacement by Cyanide Ion On the basis of the investigations discussed above, the organoclay catalysis mechanism for alkylbromide displacement can be summarized as follows. The reaction mixtures formed an oil-in water type emulsion in which the organoclay plays an essential role in stabilizing the emulsion. The ability of organoclay to serve as an emulsion former of reaction mixtures is controlled by three factors : the loading of onium ion, the orientation of the onium cation inside the gallery (both are affected by the preparation method and layer charge of the clay host) and the molecular structure of the intercalated onium cation. The fundamental kinetic steps of the organoclay systems for the cyanide displacement reaction are : i) the alkybromide are attracted to the alkyl chain of catalysts by hydrophobic interaction, and the CN' are attracted to the polar end of catalysts by electrostatic interactions; ii) displacement reaction occur in the emulsion phase; iii) the products, alkylcyanide and Br- are transferred from the emulsion phase back to the organic phase and aqueous phase respectively (Figure 25). C. Organic Synthesis Catalyzed by Organoclay In the previous studies we have demonstrated the feasibility of triphase catalysis for the attack of cyanide ion on alkylbromides catalyzed by organoclay. In addition to the cyanation reactions, we now demonstrate that organoclays can be applied to a variety of useful synthetic transformations, including : 1) ether synthesis, 2) oxidation of alcohols, 3) synthesis of thiocyanates and sulfides, 4) C-alkylation of nitriles, 5) dehalogenation of 89 Figure 25. The fundamental kinetic steps of the organoclay in the alkylbromide displacement by cyanide ion : i) the organic substrates are attracted to the alkyl chain of catalysts by hydrophobic interaction and the inorganic anions are attracted to the polar end of catalysts by electrostatic interactions. ii) displacement reaction occur in the emulsion phase. iii) the products are transferred from the emulsion phase back to the organic phase or aqueous phase. zzzzzz 0000000000000 91 vic-dibromides, and 6) halogen exchange. Analogous applications of other triphase catalysis systems have been surveyed by Dehmlow (34). In the following examples [MeN(C8H17)]3+-hectorite (25AA) and [Bu3PC16H33]+-hectorite (26AA) were utilized as catalysts. The X-ray d001 values of the recovered catalysts used in the above reactions showed that the catalysts retained their initial d-spacing after the reactions. This indicates that the organoclays are very stable and do not lose their functionization under the reaction conditions. The slight difference in X-ray d001 values between the‘original catalysts and recovered catalysts are probably due to the adsorption of organic species from the reaction mixture. The products were identified by comparing their 1H NMR spectrum and IR spectra with those for authentic samples (Table 16) or their GLC spectra with those of Standard samples. 1. Synthesis of aromatic ethers from phenol Conventional methods for the preparation of alkyl and aryl ethers are many in number. However, considerable effort is still being expended in developing new and more convenient procedures. Organoclays were found to be useful triphase catalysts in the synthesis of ethers. A phenoxide ion generated by treatment of phenol with aqueous sodium hydroxide (2.5 M), was reacted with n-pentylbromide at 900C. The reactions were catalyzed by organoclays 25AA and 26AA (Table 9). Phenyl l-pentylether was produced in 83% yield within 1.5 and 2 h respectively (Table 9). The molar ratio of nucleophile, pentylbromide, and catalyst used was 2/1/0.01. In this reaction, unlike the previous triphase reaction, catalyst 25 AA exhibited better catalytic reactivity than 26AA. Reeves has reported a similar reaction catalyzed by a polymer-supported phase transfer catalyst that gave 81% conversion in 1 hr at 1100C and the molar ratio of nucleophile, pentylbromide and catalyst used was 3/1/0.01 (120). Since the reaction conditions were not identical, it is hard to compare the efficiency of those two catalyst systems. 92 1. Ether Synthesis 2.5 M NaOH 05H11Br + Q—OH ’ Q—O— 05”" 90° 2. Oxidation of Alcohol H 107 NaOCL Q—w ° 50° 3. Synthesis of Thiocyanates and Sulfides 90° 90° 93 4. C-Alkylation at Activated CH-Bonds 50%NaOH f" 50 - 5. Dehalogenation of vic-Dihalide Brl-l H l | Mel + Nazszo3 l ©-°-°- - Q-c-‘°-© I I . l H Br 9° H 6. Halogen Exchange RBr + NaCl ——"’ RC1 + NaBr 90° 94 However, in other studies of similar reaction conditions, the organoclay exhibit a higher catalytic activity than polymer—supported phase transfer catalysts. 2. Oxidation of Alcohols Reaction of benzyl alcohol dissolved in toluene with 10% aqueous sodium hypochlorite in the presence of catalysts 25AA or 26AA at 500C gave benzaldehyde in 83.2% and in 98.4% yield within 10.0 h (Table 10). Further oxidation to benzyl acid occurred for a longer reaction timer. Regen has attempted a similar reaction catalyzed by a polymer-supported phase transfer catalysts but it afforded only 51% conversion after 50 h at 500C (86). These results demonstrate that for the oxidation of alcohols, a higher catalytic reacrivity was found tor the organoclay than for the polymer-supported phase transfer catalyst. 3. Synthesis of alkyl thiocyanate The reaction between aqueous sodium thiocyanate solution and n-pentylbromide was conducted at 900C with organoclay as the triphase catalyst. With a 2/l/0.01 molar . ratio of thiocyanate, n-pentylbromide and catalyst used, n-pentylthiocyanate was obtained in 99% and 95% chemical yield after reaction time of 1.5 h and 0.5 h in the presence of organoclay 25AA and 26AA, respectively (Table 11). By using a polymer- supported catalyst, Reeves carried out the same reaction under an analogous reaction condition except at a higher reaction temperature (1100C). He observed a 94% chemical yield of n-pentylthiocyanate after 2 h (120). This example shows that an organoclay exhibits better catalytic properties than a polymer-supported phase transfer catalysts. 4. Synthesis of sulfide The conversions of n-pentyl bromide to n-dipentyl sulfide were catalyzed by organoclay 25AA or 26AA at 900C. The chemical yields obtained for the two catalysts were 91% and 97%, respectively after a reaction time of 0.5 h (Table 12). Reeve used polymer as support in preparation of his triphase catalysts (polystyrene supported- (CH3)3-PBu-3+). A similar alkylsulfide synthesis with a 1.2/1/0.01 molar ratio of sulfide, 9S Table 9 Synthesis of Phenyl n-pentyl ether9 Catalyst Time, hr Yield,% doo1,Ah BlankQ 1.0 38.5 - [McN(C8H17)3]+- 1.5 83.0 21.0 hectorite [Bu3PCl6H33]+ 2.0 83.0 22.9 hectorite a Reaction conditions: 5.0 mmole n-pentylbromide; 10.0 mmole phenol and 6.0 ml of 2.5 M sodium hydroxide solution; no organic solvent is used; 0.05 mole homoionic organoclay; 900C. X-ray d001 spacing for recovered catalyst. No homoionic organoclay was used in this blank run. 1010" 96 Table 10 Oxidation of Benzyl Alcoholfi Catalyst Time, hr Yield,% d001,/XL Blanks 10 10.7 - [MeN(C3H17)3]+- 10 83.2 22.5 hectorite [Bu3PC16H33]+ 10 98.4 22.5 hectorite Reaction conditions: 2 mmole benzyl alcohol in 3 ml toluene; 5 ml 10% sodium hypochlorite; 0.05 mmole homoionic organoclay; 500C. X-ray d001 spacing for recovery catalyst. No homoionic organoclay was used in this blank run. IOO’ I” 97 Table 11 Synthesis of n-Pentyl Thiocyanateg Catalyst Time, hr Yield,% d001,Ab BlankQ 1.5 28.7 - [MeN(C3H17)3]+- 1.5 99.0 19.2 hectorite [Bu3PC16H33]+ 0.5 95.0 19.0 hectorite a Reaction conditions: 10.0 mmole sodium thiocyanate in 4 ml water; 5.0 mmole n-pentyl bromide; no organic solvent is used ; 0.05 mole homoionic organoclay; 900C. X-ray d001 spacing for recovered catalyst. No homoionic organoclay was used in this blank run. lOlO‘ 98 Table 12 Synthesis of n-Dipentyl Sulfideg Catalyst . Time, hr Yield,% d001,A.b BlankQ 1.0 20.0 - [MeN(C3H17)3]+- 0.5 91.0 21.0 hectorite [Bu3PC16H33]+ 0.5 97.0 23.2 hectorite a Reaction conditions: 6.0 mmole sodium sulfide in 3.0 ml water; 5.0 mmole n-pentylbromide; no organic solvent is used; 0.05 mole homoionic organoclay; 900C. X-ray d001 spacing for recovered catalyst. No homoionic organoclay was used in this blank run. 10 IO‘ 99 n-pentylbromide and catalyst was undertaken by Reeve at 110°C. He observed a 98% chemical yield after a reaction time of 1.5 h (120). The reaction rates of organoclays are three time as great as the rate observed for a polymer-supported phase transfer catalyst when the reaction is run of the same temperature. Organoclays proved to be highly efficient catalysts in the application of sulfide synthesis. 5. C-alkylation of nitriles The alkylation of phenylacetonitrile with n-pentylbromide was carried out under triphase system with 50% sodium hydroxide aqueous solution at 500C. The molar ratio of phenylacetonitrile, n-pentylbromide and catalyst was 1/ 1/0.01. After 6 h reaction time a 85% monalkylation was obtained when it catalyzed by 25AA while a 82% chemical yield was received after 2 h when it catalyzed by 26AA (Table 13). Only monalkylation was obtained in both reactions. Organoclay 25AA inherits a lower catalytic activity than 26AA as usual. Dou attempted analogous reactions under both biphase and triphase condition by using tributyhexadecyl phosphonium bromide and the corresponding polymer-supported cation as catalysts (142). Under biphase condition, 88% monoalkylation and 9% dialkylation (in percent for 100% of substrate) was achieved after 10 h at 700C with a 1/ 1/0.01 molar ratio of phenylacetonitrile, n-pentylbromide and catalyst. Increasing the amount of catalyst by altering the molar ratio to l/ 1/0.1, gave 70% monalkylation and 13% dialkylation after reaction time of 5 h under the analogous conditions. Triphase reaction conditions gave a best overall yield, up to 72%, and 4.5% dialkylation at 700C for 10 h reaction time. Comparisons of the catalytic properties of [Bu3PC16H33]+- hectorite (26AA), polymer supported [Bu3PC16H33]+ and Bu3PC16H33Br indicated that the organoclays are not only better catalysts than the corresponding polymer supported catalysts but sometimes they exhibit higher catalytic reactivities and specificities than the analogous biphase catalysts. 100 6. Dehalogenation of vic-dibromides Dehalogenation of meso-1,2-dibromo-1,2diphenylethane to stilbene in toluene was canied out through the use of sodium thiosulfate and sodium iodide. When the reaction was catalyzed by organoclays 25AA or 26AA, completely stereospecific conversion to trans-stilbene with 95% and 98% yield, respectively, was obtained at 90°C after 7 h reaction time (Table 14). Under similar reaction conditions Regen observed the same product in 100% chemical yield after 12 h by utilizing polymer-supported phase transfer catalysts (86). As expected, organoclays are better catalysts than other triphase catalysts. 7. Halogen Exchange The reactions of aqueous sodium chloride and alkybromide in toluene were catalyzed by [MeN(C8H17)]3+-hectorite (25AA) and [Bu3PC16H33]+-hectorite (26AA) at 900C. Table 15 illustrates the conversions for a series of alkybromides to the corresponding alkylchlorides. The reaction rates were not altered drastically by increasing the hydrocarbon chain length of the alkylbrornides. These reactions were limited by the thermodynamic equilibrium values at 3 : 10 for a Br- : Cl- ratio. By using excess sodium chloride, Regen was able to obtain a modest reaction rate when the chloride-bromide exchange reaction was catalyzed by polymer-CH2-Bu3P+ at elevated temperature, 110°C (86, 95). We can expect our reaction rates to increase more than 11 times if we use an analogous concentration of sodium chloride. D. Comparison of the Catalytic Properties of Organoclay The catalytic properties of a few commercial organoclays were studied in comparison with that of our organoclay (Table 17). In the cyanation reaction organoclay, Benton 34, exhibited no catalytic activity and gave 0% chemical yield after 24 h reaction time. Under analogous reaction conditions organoclay ECCT 40 , ECCT AF and ECCT PS showed modest catalytic activity and received a similar result of 73% chemical conversion within 8 h. These results, in comparison with our previous work listed in 101 Table 13 C-Alkylation of N itrilesa Catalyst Time, hr Yield,% acclaim Blankfl 6.0 3.5 [MeN(C3H17)3]+- 6.0 85.8 21.5 ‘ hectorite [Bu3PC16H33]+ 2.0 82.0 21.9 hectorite a Reaction conditions: 5.0 mmole n-pentylbromide; 5.0 mmole benzylcyanide; 2.0 ml of 50% sodium hydroxide solution; no organic solvent is used; 0.05 mmole homoionic organoclay; 500C. X-ray d001 spacing for recovered catalyst. No homoionic organoclay was used in this blank run. 1010‘ 102 Table 14 Dehologenation of vic-Dibromidesfi Catalyst Time, hr Yield,% doo1,A!2 Blankfi 21.0 0.0 - [MeN(C3H17)3]+- 7.0 95.0 22.5 hectorite [Bu3PC16H33]+ 7 .0 98.0 20.0 hectorite a Reaction conditions: 1.0 mmole meso- 1,2-dibromo—1,2 diphenylethane in 2.0 ml toluene, 0.246 mmole sodium iodide; 32.0 mmole sodium thiosulfate in 2.0 ml water; 0.05 mmole homoionic organoclay; 90°C. 1; X-ray d001 spacing for recovered catalyst. 52 No homoionic organoclay was used in this blank run 103 Table 15 Synthesis of Alkylchloride from Alkylbromideil Catalyst alkylbromide Time, Yield domh (hr) (%) (A) Blanle C5H11Br 48.0 0.0 - [MeN(C3H17)3]+- C5H11Br 78.0 73.4 21.8 hectorite [Bu3PC16H33]+ C5H1 1Br 48.0 64.5 22.4 hectorite Blank2§ C9H19Br 48.0 0.0 - [MeN(C3H17)3]+- C9H198r 54.0 73.5 21.8 hectorite [Bu3PC15H33]+ C9H19Br 48.0 62.8 22.4 hectorite Blank3§ C12H258r 48.0 0.0 - [MeN(C3H17)3]+- Cleszr 54.0 63.3 21.8 hectorite [Bu3PC15H33]+- C12H258r 48.0 56.4 22.4 hectorite 104 Table 15 continued Reaction conditions: 2.0 mmole alkylbromide in 6.0 ml toluene; 6.67 mole sodium chloride in 3.0 ml water", 0.05 mmole homoionic organoclay; 900C. 1: X-ray d001 spacing for recovered catalyst. Q No homoionic organoclay was used in this blank run. 105 .8003N .8003.— .0 800083 J 833.0 m 583.8 .5888 .8888 .3 0N2 .3 v0: .3 0mN~ .3 05m" .3 $2 .8 mm: .m 003 .m 3% .3 Ram .8 anon 055.8035. 3:08-: a 8... .. 8e .. 8e .. m: .. 8w .... 82 .3 82 .3 82 .e we: a 8.: .... 8N. .. 22 .8 82 a 8: .. 82 a 82 .8 8: a 8: .e 2 8 .... 88 .... 88 .3 88 .... 88 8282 :28 m 000 .m 03. .3 m; .3 n5 .3 cow .3 :00— .3 m2: .3 be: .m 002 .3 33 .3 32 .8 on: .m 002 ...m 002 .m #02 .8 whwm .sm RQN 9 8% .3 Son .3 38 Sam 5:08-: .08.. QMEZT: #780 > 2ng 039350 05:00... 3300590 3 8:8... c5 .8 an: :22 3 es. 3.5 8285 2 use... 106 58:2 Eaamms .Eamcmfi .m 8&2. Sam was .. Sauna .888: .683: .u 588.: .. 588.8 .Eaaov.m .Eqanné .: 5&8: .. 6888 Own .5 ovn .m ooh m o2. .m con .5 OS .m 8o .3 CR: .5 an»; .8 com" .3 82 .3 8m: .3 anon .3 89m .3 95m .m> wave .5. cm. .8 own .3 59 .3 Sb— .3 num— m 33 .E 82 .E NOS .5 cmNN .m nhwm .m> Son .5 3% .5 08m .8 Neon E a; .5 8n .3 2a .3 3: .E 82 am $318 2.2 .5 SN— .3 mmg .E n52 .m mm: .m Vnwm .m> Row .5 «man 23.8-85 0:25:30: _.9:om-~ 08:8 3:25 3.27:: fl—qu> 0385 3.880 2 22¢ 107 .8020: m: 8:80.520 03:88:00 38: 3 8:058 3825 :: m: 0028388882 0. 28:3: 02:38 203 :85: .8825 £80 : 0.5 82: wimmpa 0:: Ex .83 388:: 05 838 .3 38:50 203 800% Bow 8:8 5:2 88 80 3%: .3 38:30 203 :80on :0::_0m 8082.88 .3 098:8 0:: 2:23. 8308 3 not: 8: N 88:5 5 03880: 80508 on. 3:808: 3:33 803 30:00.8 25. w 0| .Dl 388800 2 03:9 108 Table 3, indicated that similar catalytic properties are obtained for the commercial organoclay of the ECCT series, and our 19AA-25AM organoclays. However, [Bu3PC15H33]+-hectorite exhibited a reaction rate which is about four times as fast as that of the organoclay of ECCT series. E. Comparison of the Catalytic Properties of Some Organo-Layered Compounds A series of layered compounds and their intercalation with alkylamonium ions were synthesized . The six major layered compound are alpha-zirconium phosphate (Zr(HP04)2), sodium molybdenum bronze (NaxMoO3), sodium titanate (NaTi307), sodium tintanium sulfides (NaTiSz), Zn-Cr layered double hydroxide (anCr(OH)6) and layered chloro tin phosphate (SnCl(OH)(HPO4). The intercalated onium ions were (C12H25)NH(CH3)2+. (C16H33)N(CH3)3+ and (C12H25)2N(CH3)2+. except for the layered double hydroxide in which case the intercalated ion was dodecylsulfate. The catalytic properties of organo-layered compounds were investigated for bromide displacement both by cyanide and chloride. All of the organo—layered compounds in these studies were either deintercalated or decomposed in the cyanation reactions, as indicated by the X-ray data of the recovered catalysts. In the chloride displacement of pentylbromide reaction, only (C12H25)NH(CH3)2+-zirconium phosphate gave a positive results (Table 18). Emulsion formation was observed in the (C12H25)NH(CH3)2+- zirconium phosphate system and an analogous chemical conversion, 62.8% yield after 48 h, was observed for organoclay 26AA. The catalytic conversions, observed for (C12H25)2N(CH3)2+-[Zr2Cr(0H)6]-[C12H25504]. (C12H25)NH(CH3)2+-Ch10r0 tin phosphate and (C12H25)2N(CH3)2+-sodium tintanium sulfides were due to the de- intercalation of the onium cations and their subsequent involvement as biphase catalysts, and can be proved by the continuation of the reaction after removing the solid catalysts. Although the catalytic reactivities of the organo-layered compounds were disappointing, these triphase technique might be able to improve by varying the crystallinity of the 109 Table 1? Catalytic Reactivities of Commercial Organoclay in the Synthesis of n-Pentylcyanide from n-Pentylbromide under Triphase ReaCtion Conditionsfl Catalyst Time, hr Yield,% Benton 34h 24.0 0.0 ECCI‘ 40!: 8.0 73.0 ECCT AFh 8.0 7 3.0 ECCT PSD 8.0 72.0 [MeN(C3H17)3]+- 8.0 77.0 hectorite [Bu3PC16H33]+- 3.5 95.0 hectorite a Reaction conditions: 2.0 mmole n-pentylbromide in 6.0 ml toluene; 20.0 mmole sodium cyanide in 3.0 ml water; 0.13 g of organoclay; 900C. Benton 24 was donated by Baroid Co. and ECCT organoclay series were obtained from English China Clay Co. 110 AVOEEOXUS 3: 03:8: 2 : o3 IBEUVZQENGV 302032 0: 03:3. od oém +NAm=UVZAnNm§UV 8°25 0: 03:; ed oém +mAm5vammm£UV 3.955 8: 03:: 3... 3:. [423823328 538:0": 8:36: 0:53: $.20; 208$ 933:0 magma—0:00 «6.33:0 8:..th :02: 02:.0..£b:0m-: 80:: our—02035:: :0 305:»: 05 :0: «2:09:00 :80»£0::w.5 :0 8230:3— 833“: 3 03a. .008 "0:509:00 00:93—05:30 w o: .o ".033 :8 Qm E 02:93 833.80: 0:088 36 ”0:020. :8 c 5 02805359: 0:088 QN 600.: 0:03 8:030: $32.0: 0.5. w 111 53.2 9. 03% o: 0.3 IRRBVZQNINGV $52 2. 03925 3:. 0% 83523920: 8 032,25 0.: 3m .vomnmmfiofifioromé .vommmmfiufifiorumé 2. 03.28: on: gum Ilkamovzfimmfiuv con—3:0": 83:5 0:330 $.20; 2.06: 9336 3.5.8 E 2%... 112 layered host compounds and by altering the molecular structure of the intercalated onium ion. F. Preparation of Clay-Supported Chiral Catalysts Asymmetric synthesis by using chiral phase transfer catalysts or polymer- supported chiral phase transfer catalysts has been a subject of interest to several authors in the last few years. Although the optical purity of the resulting products is low, and few reaction mechanism have been studied, they are important as an imaginative new series of catalysts. In our previous studies of clay-supported phase transfer catalysts we successfully have translated the solution reactions into triphase systems. With thisbackground, we attempt to develop a clay-supported chiral phase transfer catalysis. The intercalation of (-) N-dodecyl-N-methyl ephedrinium and (-) N-benzquuininium on hectorite or flurohectorite was achieved following an analogous preparation procedure as for the clay-supported phase transfer catalysts reported earlier. Table 19 illustrates that the d- spacing of the clay supported (-) N -dodecyl N -methyl ephedrinium increases from 18.57 A to 25.80 A when the layer charge density of clay host increase from 0.6 e' per O4[OH, F14 (hectorite) to 1.8 e- per O4F4 (F-hectorite). A similar dependence of d-spacing on the layer charge density of clay support was observed for the clay-supported (-) N- benzquuininium. This dependence, as previous discussed, is due to the different orientations of the chiral cation on the clay host of different layer charge density. Cation orientation appears to be a determinative factor on the efficiency and Stereoselectivity of the asymmetric synthesis. G. The Stereoselectivity and Mechanism of the Asymmetric Reduction of Ketones Catalyzed by Organoclay There are few studies on asymmetric borohydride reduction of carbonyl compounds catalyzed by chiral phase transfer catalysts. Colonna and his co-workers reported that (-) N-dodecyl N-methylephedrinium bromide and (-) N-benzquuininium 113 Table 19 X-ray Basal Spacings Data of Clay-supported Chiral Catalysts?- Samples Number d001, A (-) N-Dodecyl N-Methyl 131AA 18.57 ephedrinium hectorite (-) N -Dodecyl N -Methyl F131AA 25.80 ephedrinium F-hectorite (-) N-Benzquuininium 151AA 19.60 hectorite (-) N-Benzquuininium F151AA 22.84 F-hecton'te a Clay-supported chiral catalysts were prepared under following condition: 1% aqueous clay solution was stirred vigorously with 0.073 M aqueous chiral onium salt solution for 24 hours. The products were washed free of excess onium salt. and collected by centrifugation, and air dried at room temperature; for l meq of clay, 2 meq of chiral onium ion was used. 114 chloride gave 10.6% and 22% ee, respectively, in the reduction of t-butyl phenyl ketone under biphase condition (69, 143, 141) (the ee numbers given in the literatures were 13.9% and 28.5% according the maximum value of [aJDzo = 30.0 in acetone, but we recalculated the ee number assuming the value of [a]D20 = 39.6 in acetone (143)). However, neither chiral catalyst was able to promote asymmetric induction in the reaction of ketones with less hindered reactive center such as phenyl n-propanone. So far, no detailed mechanism of asymmetric borohydride reduction catalyzed by (-) N-dodecyl N-methylephedrinium bromide and (-) N-benzquuininium chloride has been proposed. However, most research groups have recognized that in order to obtain a high optical activity in the borohydride reduction, a tight ion pair between the borohydride anion and the chiral cation has to form, and the B hydroxy group of the chiral cation should interact with the ketone by oxygen H-bonding. Several authors have recognized that the Stereoselectivity of chiral catalysis is affected by the polarity of the solvent employed, the molecular structure of the chiral phase transfer catalyst, the molecular structure of the substrate and the conditions of the reaction (temperature, time and catalyst-substrate molar ratio). We have investigated the behavior of clay-supported chiral phase transfer catalysts in the asymmetric reduction of ketones by consideration of similar experimental parameters. In the following experiments, t-butyl phenyl ketone was synthesized according to the reported literature. The identification of the t-butyl phenyl ketone and the reaction products was achieved by comparing their 1H NMR spectra and IR spectra with those of authentic samples (Table 20). The enantiomeric excess of products were determined by 1H NMR in the presence of Eu(hfc)3 as a chiral shift reagent (Figure 26). The non- catalyzed reactions run under analogous conditions gave 11% chemical yield after 48 h and 75% conversion after 24 h for n-butyl phenyl ketone and t-butyl phenyl ketone respectively (Table 21 blank 1-2). Other blank reactions run without ketones exhibited a zero specific rotation after 72 h under similar reaction 115 conditions (Table 21 blank 3-6). These results indicated that the organoclay of this study are stable under the reaction conditions and the onium cation is not desorbed into solution. 1. The Dependence of Enantiomeric Excess on the Layer Charge of Clay Host In the reduction of n-butyl phenyl ketone, hectorite-supported (-) N-dodecyl N- methylephedrinium (131AA) afforded a low reaction rate but high optical activity compared with the corresponding chiral phase transfer catalysts 100% chemical yield with 799.7% enantiomeric excess was obtained after a reaction time of 24 h (Table 22). Replacing the hectorite support (sample 131AA) by F-hectorite (sample F131AA) resulted in enhancing the catalytic reactivity but decreasing the Stereoselectivity of organoclay, giving a 100% chemical conversion but essentially no optical yield after 16 h (Table 22). Similar influences of support on catalytic activity were observed for hectorite and F-hectorite-supported (-) N-benzquuininium (151AA and F151AA). However, both 151AA and F151AA show no significant Stereoselectivity (0~1.6% and 1.6~5.0% ee‘ respectively) (Table 22). When the layer charge densities of the clay host are increased, the active sites and organophilic characters of the organoclay are enhanced. This enhancement leads to immovement of the catalytic activity of the organoclay. The predominant enantiomer for the products had R configurations when the reactions were catalyzed by a hectorite-support chiral catalysts. In contrast, F-hectorite supported chiral catalysts favored the S isomer. This effect was seen clearly for low temperature reactions (Table 24). It is evident that the clay host plays an important role in asymmetric borohydride reduction. The important energy difference between hectorite and F-hectorite may express themselves in the "best fits" of the transition states. Since the hydroxy-group and reactive cation center of the catalysts are involved in the transition state of reaction, their positions and orientations relative to clay support will 116 was... .3 23:8. 350.... S. .... 05.3»...08050. 0. 050.0. 35808 803 3...... 1.0.52.0 ......00 .002 ES . .c was... .3 3.5.30 803 0.500% aces—em .8038 0.5%.... 0... ..o 5.8.59.5 .3 3.00:8 0.83 0.8300...— 0.5. 08...... 80:08 .3 BE. 9... 0:05.00 5.3 03.... 300...... 0... 89¢ 3.00.9.0 0.83 “.803...— 05. .8038 v... 8.80.820 3.0.5300 0| 'Ul .91 8.3.3.... .3 3.00:8 ...... 0.83... 8:68 .23 .003. :05 9... £3.38 0.2.3.005 8068 00.883. .0083 5.3 00:33 803 800.50 0;. es. .050 .33 3.08.5 33 0.5.0.. 3.5.. 3.3-. 2:. “I 8.38.5 .. Same... .... panama .833. .83.? .. 8&3... 89.2.5 .83 ow... m Samoa; .m 8.336 Ear—ems .m 8.33.. .m mos. .m 3K. .3 was. .3 gm .8 82 .E 3.2 .8 men. .3 3m. .8 mm: .8 em: .5 comm .:0 comm .m X3 .5 can .3 Omen .3 Noon .5 exam .3 ween .... 93 .5 >2. .m 8a .m on: .0 Na: .8 mum. .8 8m— .E man. .5. 03: ...... co: .0. 2.3 .E 82 .0 :2 ...m thm .... Cam .8 Gem 99.87.. 3.5.. 325. -.. 99.00? 380m 125.. 0.0.5.3— 3.5.. 19.5.. 322:. 3-80 a 29:8 3500?. 3.5.— 3.5... c5 .282 .35.. 3.5-. 2.93. 38.. 3....-. 8. sea ..22 m. as 38 85.... cm 2...... Figure 26. 117 The 1H NMR Spectrum of n-butyl phenyl alcohol (0.06 M in CDC13) in presence of various amount of chemical shift reagent Eu(hfc)3. The molar ratios of chemical shift reagent and the alcohol are 0, 0.32, 0.43, 0.52 and 0.87 in (a)-(e), respectively. - 118 3 3V - E r .3 .3 119 E i..- E .3 E 120 .2 83 2.: a 8:68 55 ~208th S: 332 do §-§om a no @8388 9.03 80:88 238% 22. 'Ol 6:653 3525 3 0:83 55 040 3 388% 93 0:98. math: no woman 33 3:83 mo 20% Roman—5 OI Oaths—85MB 2.8580; 295: Nd Us“? .8 06 3 35.383 83388 0.088 0 655.3 E o.m ”22.8.80 scanned m .Uovn v8: 33 zfioocawuo 25582. 2. ”.883 E. ofi 5 35283 833.52— oBEE 90:02.3 1: Qm E 0:98— 3088 Wm ”22:38 couoaom w. 06 ed 9Q. - (£3 E mo x55 ed od ad. - < 22. 22> 2 .22: 29:8 28.8 9.20-2 “Benz 3225 32506 8882 «:8. 282.8 225m 3.22.8 «a 225 127 2 3m 32 2 ca 3 225-. <32m 2 3.; 282 2 ed 3 23..-. <32m a 5m 32 2 o3 ad .25-. <32". 2 2m ix: 2 3 3 2.3-. <32”. m an Ex: 2 3 on 2.3-.. 3322 m 3 2.2 2 od 3 2:5-.. <32m m 3 0.2: 2 3 on 2.3-.. 3322 m Q». 32 2 3 3 EB-.. <32; u 2. .22» 2. .22» 2 .22... 28:: 28.8 380.2 38:2 22:5 2282.0 8882 .25. 28228 295m 255:8 33 2%» 128 ..cowao. 5.7. .823 2. 2853. go Seaman on. 5 AN: 2 and 522 E. .3 3:22.22. 83 «.398 utoEoucacm o .2265... BEBE 2. 2.83. .23 020 .3 3820.. 203 82.85 6:28. wag... no @023 2.3 .28... .8 22» 32825 n .Uovxsoocaws 2:20:52 088:. Nd .2225 .2. ed 6:359 ...: QM 3:38:00 .5288 3.5an: a 882.8 33 2%.» 129 Table 23 The Dependence of Excess Eenantiomer on the Concentration of Clay Supported Chiral Catalysts in Asymmetric Borohydride Reduction of Ketones; Sample Amount of Time, Chemical Optical Number Catalyst hr Yield, % Yield, % mmole 12 Q F131AA 0.2 48 100.0 12.1 S F131AA 10.0 36 100.0 11.4 S a Remaining reaction conditions: 2.5 mmole n-butyl phenyl ketone in 3.0 ml benzene; 6 mmole potassium borohydride in 5.0 ml water; -100C. 12 Chemical yield of alcohol was based on starting ketone. Product was detected by GLC with decanc as internal standard. g Enantiomeric excess was determined by 1H NMR (250 MHz) in the presence of Eu(hfc)3 as chiral shift reagent. 130 of the transition state. Thus, when the non-catalyzed reaction pathways are insignificant, the concentrations of substrates do not affect the enantiomeric purity of the products. 4. The Dependence of Enantiomeric Excess on the Temperature For both n-butyl phenyl ketone and t-butyl phenyl ketone reduction the enantiomeric purity of the corresponding alcohols increased with decreasing the temperature (Table 24). For F-hecton‘te-(-) N-dodecyl N-methylephedrinium (F131AA) as catalyst, the reduction of n-butyl phenyl ketone afforded 12.1%, 0.8%, 0%, 0% cc at the reaction temperatures of ~100C, 40C, 250C and 500C respectively. The reduction of t-butyl phenyl ketone gave 11.4% and 8.7% cc at ~100C and 40C respectively. For the reduction of n-butyl phenyl ketone hectorite-(-) N-dodecyl N-methylephedrinium (131AA) showed a similar dependence and afforded 25.1% and 9.7% cc at -100C and 40C respectively. Temperature can affect the degree of asymmetric induction in two ways. First, as the temperature increases, it promotes the reaction through the non- catalyzed pathway, which results in racemic mixtures. At lower reaction temperatures the reaction rate of the non-catalyzed pathway, which has a higher activation energy, is not comparable to that of the catalyzed pathway. Therefore, the catalyzed reaction becomes the major reaction pathway (according to the Arrehenius equation : rate coefficient k = A exp (-Ea/RT)) . Similarly the small energy difference between diastereomeric reaction paths will caused greater stereoselectivity at low temperature. Also, molecular motion increases with temperature and the conformation of the transition state through the catalyzed pathway will lose its rigidity and decrease the stereoselectivity. This second effect may not be as drastic for the more hindered t-butyl group as for the n-butyl group. Thus, in the reduction of t-butyl phenyl ketone, the temperature effects are higher than those in n-butyl phenyl ketone. 131 5. The Dependence of Enantiomeric Excess on the Polarity of Organic Solvent Another important generalization that seems to be emerging in enantioselectivity concerns the nature of the organic solvent employed. In a biphase catalyst system, more often than not, the highest inductions are observed with non-polar solvents. Ion pairs with strong association play a crucial role in the asymmetric induction. Media capable of hydrogen bonding or solvating chiral ion pairs generally inhibit induction because of a weakening of the ion pair interaction. This could also explain why in solvents with high dielectric constant, where the force between two charged species is weak, the enantiomeric purities are substantially lower than in solvents with low dielectric constant. Similarly, in clay-supported chiral catalysis one observed the dependence of stereoselectivity on the polarity of organic solvents employed. In the presence of benzene, hectorite-H N-dodecyl N-methylephedrinium (131AA) afforded a predominant enantiomer of R configuration with 9.7% optical yield in the n-butyl phenyl ketone reduction, whereas in THF, 131AA afforded a predominant enantiomer of opposite configuration with 8.6% cc (Table 25). The complexity of clay-supported chiral catalysts and the lack of kinetic data make it difficult to rationalize the results. However, from our knowledge of previous experiments, the organic solvents can affect the emulsion formation of reaction mixtures which in turn affect the reaction rate of catalyzed pathway. When the catalyzed reaction becomes the major reaction pathway we can expect to obtain a higher degree of enantioselectivity. The organic solvents can also alter the interaction of the chiral ion pair and influence the asymmetric induction. There is no generalization of the solvent effects in this system. It is necessary to study each clay- supported chiral catalyst individually for this subject. 6. Conclusion Although a great deal of research needs to be done before the mechanism of clay- supported chiral catalysis can be proposed, it appears worthwhile to summarize the 132 results of the study of ketone reduction : 1) For a given intercalated chiral cation, the organoclays of hectorite and F-hectorite supports tend to afford predominant enantiomers with opposite configurations; 2) The concentrations of substrates do not affect the degree of enantioselectivity; 3) The concentration of hectorite-supported (-) N-dodecyl N- methylephedrinium (131AA) catalyst affects the reaction rate but not the stereoselectiviy for the reduction n-butyl phenly ketone. However, this concentration effect of catalyst should be studied individual for different clay-supported chiral catalysts and organic substrates. 4) In general, low temperature favors the stereochemical course of asymmetric ketone reduction. However, the extent of the temperature dependence is different for various ketones. 5) Changing the solvent can afford products in favor of opposite isomer. However, the effects of organic solvents are not clear at this point. It has to be said that the clay-supported chiral catalysts are practically superior to conventional biphase catalysts. Moreover, for a given chiral catalyst we are able to alter its optical selectivity by intercalatin g the chiral catalysts on various clay hosts. H. The Stereoselectivity and Mechanism of Asymmetric Epoxidation of Chalcones Catalyzed by Organoclays The epoxide function plays an important role in metabolic processes. It is surprising therefore that the catalytic synthesis of optically active epoxides leaves much to be desired. Wynberg has attempted to epoxidize chalcone under the influence of quinine and ephedrine but the results were disappointing. However, chalcones and related compounds could be transformed in excellent chemical yields into optically active epoxides by using quaternary salts derived from quinine as chiral phase transfer catalysts in biphase systems (63). A limited study and the lack of kinetic data make it difficult to drive conclusions concerning the role of the catalyst structure and to propose reaction mechanisms. Nevertheless, Wynberg and Dollin g have stated that the cinchona alkaloids are by far the better all around catalysts; and that the absolute configuration at C-8-C-9 of 133 Table 24 The Dependence of Enantiomeric Excess on Temperature in Asymmetric Borohydride Reduction of Ketones with Clay-Supported Chiral Catalystsa Sample ketone Temp. Time, Chemical Optical Number R-Group 0C hr Yield, Yield, %._b %. 2 F131AA n-butyl -10 48.0 100.0 12.1 S F131AA n-butyl 4 16.0 100.0 0.8 S F131AA n-butyl 25 8.5 100.0 0.0 F131AA n—butyl 50 5.0 100.0 0.0 F131AA t-butyl - 10 22.5 100.0 11.4 F131AA t-butyl 4 10.0 100.0 8.7 131AA n-butyl -10 72.0 100.0 25.1 131AA n-butyl 4 24.0 100.0 9.7 Reaction conditions: 2.5 mmole ketone in 3.0 ml benzene; 6 mole potassium borohydride in 5.0 ml water“, 0.2 mmole homoionic organoclay. with decanc as internal standard. Eu(hfc)3 as chiral shift reagent. Chemical yield of alcohol was based on starting ketone and detected by GLC Enantiomeric excess was determined by 1H NMR (250 MHz) in the presence of 134 Table 25 The Dependence of Enantiomeric Excess on the Polarity of Organic Solvent in Asymmetric Borohydride Reduction of Ketones with Clay-Supported Chiral Catalystsfl Sample Organic Time, Chemical Optical Number Solvent hr Yield, % Yield,% 12 2 131AA CC14 24.0 100.0 5.4 R 131AA C6H6 24.0 100.0 9.0 R 131AA THE 24.0 100.0 8.6 S a Reaction conditions: 2.5 mmole benzyl n-butyl ketone in 3.0 ml organic solvent; 6 mmole potassium borohydride in 5.0 ml water; 0.2 mmole homoionic organoclay;40C. . Chemical yield of alcohol was based on starting ketone and detected by GLC. Enantiomeric excess was determined by 1H N MR (250 MHz) in the presence of Eu(hfc)3 as chiral shift reagent. 1010' 135 the alkaloids govern the absolute configurations of the products. A mechanism, as discussed in chapter 1, concerning a compact ion pair was presented by Wynberg (63). Epoxidation of chalcone catalyzed by benzyl quininium chloride affords a 34% optical yield in the presence of 30% hydrogen peroxide and 10% sodium hydroxide. The predominant configuration of epoxychalcone is S with an R configuration at the a-carbon and an 8 configuration at the B-carbon (Figure 27). Kobayashi has utilized a polystyrene supported quininium salt as a catalyst in chalcone epoxidations, and the results do parallel those of the analogous non—supported reactions (43). The phase transfer-mediated asymmetric epoxidation by using clay-supported ephedrinium salt (131AA) was attempted with little success (T able 26). However, clay- supported cinchona alkaloids (151AA, F151AA) exhibit a higher stereoselectivity than analogous non-supported or polymer supported cinchona alkaloids and afforded products with maximum optical yield, up to 46.7% (Table 29). The blank reactions were carried out under the conditions either without the chalcone or clay-supported chiral catalysts (Table 27). No specific rotations was observed by polarimeter when no chalcone was used. These results indicated that both 151AA and F151AA were stable under the reaction conditions. If organoclays lost their functionizations or the intercalated cation decomposed under the reaction conditions the specific rotations of the reaction residence should be greater than 0. The non-catalyzed reaction rate was very low and non significant product was observed after 3 days (Table 27). l. The Dependence of Enantiomeric Excess on the Layer Charge of Clay Host According to the previous discussion, the layer charge density of the clay host affects both the loading and the orientation of onium cation. This, in turn, influences the catalytic and stereoselective properties of clay supported chiral catalysts. For the epoxidation of chalcone catalyzed by the F-hectorite supported n-benzyl quininium (F151AA) the reaction rate was about 5 times as fast as that of the analogous reaction 136 catalyzed by hectorite supported n-benzyl quininium (151AA). Nevertheless, these two clay-supported chiral catalysts possess a similar enantioselectivity and afforded predominant enantiomers of opposite configuration. For instance, with F151AA as catalyst, the reactions afford a 100% chemical yield and 27.9% enantiomeric excess of epoxychalcone with a predominant isomer of S configuration (Table 28). In contrast, a predominant R enantiomer was found in the analogous reaction catalyzed by 151AA (21.0% ee). These results were consistent with the observations in the asymmetric borohydride reduction. For a given intercalated chiral cation, the F-hectorite support and hectorite support afforded analogous products with predominant enantiomers of opposite configuration. For a preferred conformation of benzyl quininium cation (Figure 6), tight ion pairing between the chiral cation and the peroxide anion, and hydrogen bonding between the hydroxyl of the alkaloid and the carbonyl oxygen of the chalcone, suggest the conformations of transition state shown in Figure 28. The various possible conformations of transition state introduced by diverse clay hosts have different activation energies. The various organophilic characteristics of diverse clay hosts also affect the catalytic activity of the organoclay. These may explain the influence of clay hosts on the catalytic and enantiomeric properties of clay-supported chiral catalysts. 2. The Dependence of Enantiomeric Excess on the Temperature Colonna has reported that in the presence of polymer-supported chiral catalysts, the degree of asymmetric induction in the epoxidation of chalcones decreases as the temperature increases (146). A‘ similar dependence of stereoselectivity on the temperature were reported for other asymmetric induction under biphase systems (63, 147). However, in the organoclay catalysis, we discovered, that fOr the chalcone epoxidation, the enantiomeric purity of the epoxychalcone increased with temperature. Table 29 illustrates that the reaction catalyzed by F511AA in presence of carbon tetrachloride afforded 46.7% ee at 250C and 17.0% ee at 40C. Analogous reactions in 137 Figure 27. Configuration of (-) Epoxychalcone. 138 Table 26 Dependence of Enantiomeric Excess on the Molecular Structure of Chiral Catalysts for Asymmetric Epoxidation of Chalcone with Clay Supported Chiral Catalystsé Sample Reaction Chemical Optical Number Time, hr Conversion, % Yield,% 12 9 131AA 102 100.0 0.7 a—S, B—R 151AA 102 100.0 21.0 a—S, B—R a Reaction conditions : 0.48 mmole chalcone in 3 ml benzene; 30% H202 1.5m] and 12% N aOH 1.5 ml; 250C; 0.2 mmole homoionic organoclay. b The reaction was monitored by TLC until no trace of the starting reagent (chalcone) was present. Q Enantiomeric excess was determined by 1H NMR (250 MHz) in the presence of Eu(hfc)3 as chiral shift reagent. 139 .. < 8% < 2.5 8.. 8.8.2: 5.3 368223 .3 .252 do hoe—macro; a :0 3.3368 0.83 20226.. 2.28% 05. .cm 035. E 3.86:. 0:5 .6282 on. .22. 258 8: 2.3 8.60:. .8228. 05 we 8...: OH“. 25. .003 $285.96 2.5682. 208:. Nd ..E n.— mOaZ gen. .2... .2. n4 NONE oeom "econ—.3 .8 Gm 6:026:60 23.83. .003 ”.8 2. 39.2 $3 9... .8 m.— NONE amen "econ—.3 .5 ca 5 0:835 2688 cad ”2.2.6.50 .8283. m OI'UI .Dl od - QS. <52". mm 25.5 od - ONE ()2 2 Am 6:85 ad ad o3» - m. 255 u u .2. .5233— :omm..0>=oU .2 .083. 23:52 056on 32:86 .5263”— .deEU .5203“ a 2.8.2.0 .o 8828...". ..o 888.. 28... 2F R 2...» 140 Figure 28. The possible conformation of transition state of the chalcone epoxidation catalyzed by 151AA 142 .8030. ......m .820 v... £0.53 ..0 00:30... 0... ... G: .2 SN. ”:27. ... .3 00885.00 2.3 3.00.8 0.580.235. M 0:80.... 83 6.50.5.0. 809.0. wig... 0... .0 008. 0.. ....5 Us... .... 029.58 33 80.5.3. 0.... m. $2008.30 080.080.. 0.088 Nd "U03 ”.8 n. :92 Q0... ...... .8m.. NON: $0.“ “8030.. 0.835 .8 m ... 0.50.20 0.088 5.... . 80.50.50 80.53.. 3.8.080... w. .... .mé ...... 08. 8. <<.n. m... ...-.. ...: 08. 8 <32”. m m o... .20; .0 80.80800 .... .08.... 508:2 .880 .8825 888.. oasam 8.9.3.0 ......U 0080.53-86 ....3 0.50.5.0 .0 8083300.". 08088....< 5.. .8: 8.0 0... .0 03.2.0 .0»... 0... .5 mmooxm 0.580382... .0 0080.52.00 mu 0...“... 143 the presence of benzene bear 27.9% ee at 25°C and 10.6% ee at 4°C. In contrast utilizing 151AA as the catalyst, a predominant isomer of opposite configuration was obtained with 25.3% cc at 250C and 14.2% ee at 40c. Although increasing the temperature will enhance the reaction through the non- catalyzed pathway, this is not significant for chalcone epoxidation at the experimental temperatures used here (Table 27). Figure 29 presents a hypothesis concerning the dependence of reaction rate on the temperature. For the expoidation catalyzed by F151AA the formation rate of R isomer is higher than that of S isomer at high temperature . However, the formation rate of the S isomer does not respond to temperature as drastically as that of the R isomer. Therefore, as .the temperature decreases the formation rates of R and S isomers become equal and no enantioselectivity will be observed. Furthermore at extremely low temperatures one can expect that the S isomer to become the predominant product. An analogous dependence of optical yield on the temperature was obeyed for the epoxidation catalyzed by 151AA. The reaction rate of S isomer instead of the R isomer responds more drastically to temperature change. Thus, the prevalent enantiomers have an R configuration at high temperature while a predominant S isomer is expected at low temperature. 3. The Dependence of Enantiomeric Excess on the Polarity of Organic Solvent Both Wynberg and Colonna have claimed that the quininium salt catalyzed phase transfer reaction are subject to strong solvent effects (63, 148). Non-polar aprotic solvents give higher asymmetric induction than solvents with high dielectric constants which inhibit the asymmetric induction by weakening the ion pair interaction (63, 148). However, in the polymer-supported chiral catalyst systems there is no direct correlation between the dielectric constant of the solvent and the enantiomeric excess of the epoxychalcone obtained (146). The epoxidation of chalcone catalyzed by F151AA 144 afford 46.7%, 27.9% and 8.2% cc in the presence of carbon tetrachloride, benzene and THF respectively (Table 30). While utilizing 151AA as catalySts the degree of stereoselectiviy in the epoxidation of chalcones was nearly the same in the presence of carbon tetrachloride and benzene (25.3% and 21.0%). Since CC14 and C6H5 have similar dielectric constant (2.23 and 2.28 respectively); An analogous solvent effect on the optical activity should be observed in CCl4 and C6H6, if there is a direct correspondence between the polarity of solvent and the optical activity of the reaction. It is too early to rationalize the solvent effects on the stereoselectivity of the organoclay systems. However, according to previous studies we can expect different organoclays to have varying dependencies of organic solvent on the enantioselectivity of the epoxidation reaction. 4. The Dependence of Enantiomeric Excess on the Concentration of Chalcone Since the concentration of chalcone does not affect the conformation of transition state of reaction, there is no substantial change of the stereoselectivity of the reaction when the different concentrations of chalcone are employed (Table 31). This observation is identical with that of the asymmetric borohydride reduction in the organoclay systems. 5. The Dependence of Enantiomeric Excess on the Volume Ratio of Aqueous and Organic Phases Table 31 illustrates the results concerning the dependence of enantiomeric excess of products on the volume ratio of the aqueous and organic phases. For a constant concentration of catalyst and substrate, the catalytic reactivity and stereoselctivity of the reaction decreases as the aqueous volume employed increases. The epoxidation of chalcone was complete after 44 h and the optical yield was 29.1% when the volume ratio of aqueous and organic phases equal to 1. It took 64 h and 102 h to complete the reactions when the volume ratio was increased to 5/3 and 7/3, and the optical yield were 20.8% and 15.4% optical purities, respectively. The volume ratio of aqueous and organic 145 phases is especially important in determining the characteristics of the emulsion for reaction. This, in turn affects the reaction through the catalyzed pathway and influences the degree of enantioselectivity. 6. Conclusions A few results of epoxidation of chalcone catalyzed by organoclay are summarized as follows : 1) For a given intercalated chiral cation the F-hectorite and hectorite supports yielded epoxychalcones with predominant enantiomers of opposite configuration. 2) The concentration of chalcone does not affect the optical purity of product. 3) Over the experimental temperature range, 250C to 40C, the stereoselectivity of the reaction increases with temperature. 4) Carbon tetrachloride is thus for the solvent of choice in the asymmetric epoxidation of chalcone catalyzed by organoclay 151AA or F151AA. 5) The volume ratio of aqueous and organic phases is especially important in determining a characteristics of the emulsion of reaction mixtures, which in turn affects the catalyzed reaction rate and influences the degree of enantioselectivity. .80w00. ...... .0....0 00 20.5.5 .0 00:05.0 0... ... .0: S. anv 0.22 ... .... 000.8550 003 0000.5 0.580.800... 0 80.0.0 00>. 0.50.0.5. 80w00. 8.0.0.. 0... .0 000.. 0.. .....0 0..... .... 005.808 003 00.500. 0..... m . 00.008030 005.080.. 0.088 Nd ”.8 n. .0007. c0N. 0:0 .80. NONE 008 ”80200. 0.8030 .8 m ... 0.50.0..0 0.088 mvd . 800.0000 80.500. 8.0.080... .0 a... .010 $0 0.8. S. :00 m... <<§ .... .010 N... 0.8. 00. 0.00 c <<.0. 01.. ...-.. a... 0.8. cm 0:00 mm <52. m... ...... 0.0. 0.8. N. 0:00 c <52. 6 M -- ........................................................................................................................... m... .....e 8.. 0.8. o... 300 n <52... 010.010 ...: 0.8. N. :00 ... <50... .0. 0 00.0.0; 00 80.05800 ....08.... 80>.0m 00 508:7. .00..0O 0.0.82.0 80.5000 0.0030 .080 ... 0.080m 8....an 80.0 00.500.54.50 ....3 0.50.0..0 .0 00000.50m 0.808803. . 5. 0.805080... 0... .5 mmooxm 0.580.808”. .0 00.50.5000 0N 0.00... 147 Figure 29. The hypothesis of the dependence of the reaction rate on the temperature A) Epoxidation of chalcone by F151AA B) Epoxidation of chalcone by 151AA Ink 148 HT (3) 1/T (A) Ink .4.“ 149 .8095. .....0 .085 00 20.5.5 .0 00:80.0 0.0 0. .0... .2 and M22 .... .3 000.8550 003 0008.0 0.580.885 8000.0 003 5000.000. 80w00. 00.0.0.0 0... .0 000.0 00 ..80 0..... .... 005.808 003 00.500. 0..... «DID 0.0500030 0.00.0800 0.088 N... .003 ..8 n. $002 002 000 .8n.. NONE 000m .8038 0.0030 .8 m 0. 0000.000 0.08... «0... . 800.008 00.500. 88.080... 0 0... .0... 0... 0.8. 8. 0000 <<.0. 0... .0... 0.0. 0.8. 8. 300 <<.0. 0... .00 0.0 0.8. 0. 0.... <<.0.. 0... .0... a... 0.8. on 0000 <<.0.0 0... .0-.. ...... 0.8. o. 300 <50... 0 ... 2. .0.0.> .00 80.0.0800 ... .08.... 80>.0m .00802 0.0000 .00.8000 00.5000 0.0030 0.080m 0050.000 .0 00.000505 0.008800... .0. 80>.0m 0.0030 .0 5.8.00 0... 00 0000.5 0.580.885 .0 0000000000 on 0.00... 80000. .....0 .0800 00 20.5.5 .0 00500.0 0... 0. A a: .2 03 v 022 .... .... 008.8050 003 0008.0 0.580.805 .0 8000.0 003 80000. 000.00 0... .0 000.0 00 ..80 0..... .3 0058008 003 00.500. 0..... 0 00300500030 0.00.0800 0.088 Nd 500000.. .8 0.... ”0000.050 00.500. 8.8080”. .0. 0... .010 ...0. 0.8. 8. 00.0 0.0 0.0 <<.0.. 0.0 .0... 0.0. 0.8. 5 00.0 0.0 0.0 <<.0.. 0... .0... .0. 0.8. 3 00.0 0.. 0.. <50... 0... .0-.. 0.0. 0.8. 8. 00.. 0.. 0.. <<.0.. O b 0... .0-... .0. 0.8. 3. 00.0 0.. 0.. <<.0.. min 5:0 0.8 0.8. ON 3.... 0.. n. <<.n.0 w .0.. a .0.. .8 .8 .05; 00.0.0800 ... .08.... 0.08... 00m. $0.” 50802 .5000 00.8000 00.5000 0000.000 1002 NON: 0.080m 000.000 ......0 80800.05 5.3 88.2.0 .0 000000.005 0.008800. .0. 00000 0.0030 000 00000.). .0 000m 080.0> 000 00.0.0085 .0 00000000000 0... 00 0000.5 0.5888005 .0 0000000000 .m 0.00... 151 CHAPTER IV RECOMMENDATIONS Although in this study we have successful translated solution reactions into a triphase organoclay system, there are a few studies worthy of exploring in the future. Several parameters which control the efficiency of the organoclays in nucleophilic substitution remain to be studied : 1) the dependence of the catalytic activity on the swelling properties of the organoclay in various organic solvents. 2) the dependence of the catalytic properties on the volume ratio of aqueous and organic phases. Preparation of a new series of organoclays by intercalation of crown ethers or cryptands in the clay may be worthwhile to attempt. In addition to the reactions that were investigated in this study, the organoclays can emerge as a broadly useful tools in other organic synthesis : 1) alkylation and condensation reactions 2) Ylide-mediated reactions 3) miscellansous reactions. The investigations of organo—layered compounds as triphase catalysts have just been initiated. The catalytic properties and mechanisms of organo-layered compounds remain obscure. However, this technique can be understood better by synthesis of various organo—layered compounds with different crystallinity and molecular structure of the intercalated onium ion. The mechanisms of catalysis of organo-layered compounds can be investigated by examining the parameters that control the efficiency of the 152 organoclay catalysts (e.g., the structure of intercalated onium cations, the concentration of substrates and catalySts. the degree of crystallinity of the layered compounds and the polarity of organic solvent). Furthermore, the intercalation of chiral catalysts in the layered compounds perhaps can be developed into other types of triphase chiral catalysts. It is evident that the clay-supported chiral catalysts are superior to corresponding chiral catalysts and polymer-supported chiral catalysts. Attempts to prepare a new series of clay-supported chiral catalysts by intercalating the (+) n-benzyl quinidium cation or chiral crown ether complexes (125) on the clay is worthwhile. For the borohydride reduction the mechanism can be understood better by further examinations of the dependence of catalytic activity and the stereoselectivity of the reaction on 1) the volume ratio of aqueous and organic phases; 2) the method of mixing the reagents; 3) the alkyl chain length of ketone; 4) the swelling of clay-supported chiral catalysts in various organic solvents; and 5) temperature. The epoxidation of chalcones, catalyzed by organoclays can be expanded into the epoxidation of electron-poor olefins ( e.g., chalcones, quinones and cyclohexenones) (68, 69, 77, 78, 81-83). The catalytic activity and the stereoselectivity of the reaction might be improved by using t-butyl hydroperoxide or 28% sodium hypochlorite instead of hydrogen peroxide.- Other parameters that might affect the catalytic activity and the stereoselectivity of the system are : 1) the method of mixing the reagents; 2) the swelling of clay-supported chiral catalysts in various organic solvents; 3) the molecular structure of the organic substrate; and 4) temperature. In addition to the borohydride reduction and epoxidation of electron-poor olelfins, several successful application of chiral phase transfer catalysts in asymmetric induction have been investigated : Michael reactions (34, 63, 73-76), 1,4-thiol and thiolacetate additions (34, 63, 75), selenophenol addition reaction (34, 63, 151), 2,2-cycloaddition reaction (34, 64, 76, 150) and 1,2-additions (34, 63, 151). From our previous studies, we 153 can also expect that those reactions may achieved high stereoselectivities by utilizing the clay-supported chiral catalysts. >09°>’.°‘.‘"PP° 11. 12. 13. 14. 15. 16. 17. 18. 154 LIST OF REFERENCES G. Brown, Ed., "X-Ray Identification and Crystal Structure of Clay Minerals", Mineralogical Society, London, 1961. G.W. Brindley and G. Brown, Eds, "Crystal Structures of Clay Minerals and Their X-Ray Identification", Mineralogical Society, London, 1980. TJ. Pinnavaia, Science, 1983 229, 365. S.L. Swartzen-Allen and E. Matijec, Chem. Rev., 1974 14, 385. E.T. Uskova, N .G. Vasilev, and LA. Uskov, Colloid J. USSR, 1968 3_Q, 118. JD. Bemal, "The Origin of Life", Widenfield and Nicholson, London 1967. D. Fishman LT. Klug and A. Shani, Synthesis, 1981, 137. LS. Newton, Speciality Chemicals, 1984, 17. B.I(.G. Theng, "The Chemistry of Clay—Organic Reactions", Adam I-Iilger, London 1974. CM. Starks, J. Am. Chem. Soc., 1971, 93, 195. (a) CM. Starks and DR. Napier, US. patent, 3992432, 1976; (b) C.M. Starks and DR. Napier, British patent, 1227144, 1976; (c) C.M. Starks and DR. Napier, French patent, 1573164 1969; (CI) CM. Starks and DR. Napier, Australian patent, 439286, 1969; (e) C.M. Starks and DR. Napier, Netherlands patent, 6804687, 1968. CM. Starks and RM. Owens, J. Am. Chem. Soc., 1973 25, 3613. E. Muller, O.Bayer, and H. Morschel, Germ. Offen1., 268621, 1913. J. Jarrousse, CR. Acad. Sci. Paris, 1951 232, 1421. (a) A. T. Babayan, N. Gambaryan, and NP. Gambaryan, Zh, Obshch. Khim., 1954 M. 1887; (b) A. T. Babayan, N. Gambaryan, and NP. Gambaryan, Chem. Abstr., 4910879, 1955; (c) A. T. Babayan, N. Gambaryan, and NP. Gambaryan, Chem. Zentralbl. Sonderb., 1950 SA, 4532. G. Maerker, J.F. Carmichael, and W. Port, J. Org. Chem, 1961 26, 2681. DuPont, British Patent 632346, 1949. P. Edwards, US. Patent 2537981, 1951. 19. 20. 21 . 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 41. 155 Pest Control, Ltd., British Patent 692774, 1953. R. Kohler and H. Pietsch, German Patent 944995, 1956. BB. Copelin and GB. Crane, US. patent 2779781, 1957. Farbenfabriken Bayer, German patent 959497, 1957. B. Graham, US. Patent 2866802, 1958. Gavaert Photo-Produce N.V., Beligian Patent 602793, 1961. BE. Jennings, British Patent 907647, 1962. W.S. Port, British Patent 912104, 1962. RW. Kay, British Patent 916772, 1963. F. Nerdel, British Patent 1052047, 1966. BC. Oxenrider and RM. Hetterky, US. Patent 3297634, 1967. M. Makosza and B. Serafinowa, Rocz. Chem, 1965 32, 1223 and subsequent papers. . E.V. Dehmlow, Angew. Chem. Inst. Edu., 1977 _1_6, 493. CM. Starks and CL. Liotta, "Phase Transfer Catalysis Principles and Techniques", Academic Press, New York, 1978. J.E. Gordon R.E. Kutina, J. Am. Chem. Soci., 1977 22, 3903. E.V. Dehmlow and 8.8. Dehmlow, "Phase Transfer Catalysis", Verlag Chemie, Basel, 1983. W.P Weber and G.W. Gokel, "Phase Transfer Catalysis in Organic Synthesis", Springer Verlag, New York 1977. R.A.B. Bannard, "Phase Transfer Catalysis and Some of its Applications to Organic Chemistry", US. Dept. of Commerce NIRS AD-AO30 503 July, 1976. H.H. Freedman, Pure and Appl. Chem, 1986 5_8, 857. RM. Izatt and FF. Christensen, "Synthetic Multidentate Macrocyclic Compounds", Academic press, New York, 1978. N.A. Gibson and J .W. Hosking, Aust. J. Chem, 1965 1_8, 123. (a) E. Laurent, R. Rauniyar and M. Tomalla, J. Appl. Electrochem 1984 1_4, 741; (b) E. Laurent, R. Rauniyar and M. Tomalla, ibid, 1985 _l_§, 121. SP. Ellis, D. Fletcher, W.M. Brooks and K.P. Healy, J. Appl. Electrochem, 1983 13, 735. 42. 43. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 61. 62. 63. 156 Z.Goren and I. Willner and H. Taniguchi, J. Org. Chem, 1984 42, 4755. T.Kitamura, S. Kobayashi and H. Taniguchi, J.Org. Chem, 1984 42, 4755. W.T. Ford, Adv. Polym. Sci., 1984 33, 49. W.T. Ford, Polym. Sci. and Tech., 1984 a, 201. DC. Sherrington, Macromol. Chem, 1984 3, 303. J.W. Verbicky Jr., and EA. O’neil, J. Org. Chem, 1985 30, 1786. E. Chiellini, R. Solaro and S. D’Antone, Polymer Sci. and Tech., 1984 24, 227. Y. Iwai and M. Ueda, Polymer Sci. and Tech., 1984 24, 121. (a) A.J. Dias and T.J. McCarthy, Polym. Mater. Sci. Eng., 1983 _4_9, 574; (b) A.J. Dias and T.J. McCarthy, Macromolecules, 1984 l_7, 2529. EM. Menger, Chem. Soc. Rev., 1972 _l_, 229. K. Shinoda, R. Nakagawa, B. Tamamushi and T. Isemura, "Some Physicochemical Properties of Colloidal Surfactants", Mir. Moscow, 1966. RA. Rehinder and Zhur. Jses, K him. Obschch. Mendeleeva, 1966 11, 362. P. Mukerjee, Adv. Colloid. Interface Sci., 1967 l, 241. P.G. Duggan and W.S. Murphy, J. Chem. Soc., Perkin Trans. 1, 1976, 634. P. Duhamel, J.Y. Valnot, E. Jamal, J. Tetrahedron Lett., 1983, 2863. M. Pfau, G. Revial, A. Guingant and J. D’Angelo, J. Am. Chem. Soc., 1985 l_QZ, 273. A. Enders, C hemtech, 1981, 504. A.l. Meyers, D.R. Williams, G. W. Erickson, S. White and M. Druelinger, J. Am. Chem Soc., 1981 _1_Q3, 3081. K, Tomioka, K. Ando, Y. Takemasa and K. Koga, J. Am. Chem. Soc., 1984 _1_0_6, 2718. S. Hashimoto and K. Koga, Tetrahedron Lett., 1979, 3495. D. Enders and H. Eichenauer, Chem. Ber., 1979 _l_l_2, 2933. H. Wynberg, Top. S tereochem., 1986 1_6, 87. U. H. Dolling, D.L. Hughes, A. Bhattacharya, KM. Ryan, S. Karady L.M. Weinstock and EU. Grabowski, ACS Symp. Ser., 1987 326, 68. 65. 67. 68. 69. 70. 71. 72. 73. 74 75. 76. 77. 78. 79. 80. 81. 82 83. 84. 85. 86. 87. 88. 89. 157 B. Bosnich, NATO ASI Series E "Asymmetric Catalysis", Martinus Nijhoff Publishers, Dordrecht 1986. E.V. Dehmlow, P.Singh and J. Heider, J. Chem. Research, 1981, 292. DJ. Cram, and G.D.Y. Sogah, J. Chem. Soc. Chem. Comm, 1981, 625. J. Balcells, S. Colonna and R. Fomasier, Synthesis, 1976, 266. S. Colonna and R. Fomasier, Synthesis, 1975, 531. R. Kinishi, Y. Nakajima, J. Oda and Y. Inouye, Agric. Biol. Chem, 1978 Q, 869. R. Kinishi, N. Uchida, Y. Nakajima, J. Oda and Y. Inouye, Agric. Biol. Chem, 1980 14;, 643. J. Heider, Doctoral Dissertation, T. U. Berlin, 1978. U.H. Dolling, P. Davis and E.J. Grabowski, J. Am. Chem. Soc., 1984 106, 446. . H. Wynberg and R. Helder, Tetrahedron Lett., 1975, 4057. H. Hiemstra, PH.D. thesis, University of Groningen, Groningen, The Netherlands, 1980. H. Wynberg and B.G.J. Staring, J. Chem. Soc. Chem. Comm, 1984, 1181. H. Wynberg and B.G. Marsman, J. Org. Chem, 1979 _4_4_, 2312. J .C. Hummelen and H. Wynberg, Tetrahedron Lett., 1978, 1089. H. Wynberg and B.G. Marsman, J. Org. Chem, 1980 43, 158. Y. Harigaya, H. Yamaguchi and M. Onda, Heterocycles, 1981 13, 183. R. Helder, R.W. Laane, J.C.Wiering, J.C. Hummelen and H. Wynberg, Tetrahedron Lett., 1976, 1831. . G. Snatzke, F.L. Feringa, B. Greydanus, H. Pluim, H. Wynberg and B.G. Marsman, J. Org. Chem, 1980 43, 4094. B.G. Marsman, Ph.D. Thesis, University of Groningen, The Netherlands, 1981. S.L. Regen, J. Am. Chem. Soc., 1975 21, 5956. S.L. Regen, J. Am. Chem. Soc., 1976 28_, 6270. S.L. Regen, J. Org. Chem. 1977, 42, 875. S.L. Regen and A. Nigan, J. Am. Chem. Soc., 1978 190, 7773. S.L. Regen and J.J. Besse, J. Am. Chem. Soc., 1979 _1_Q_l_, 4059. S.L. Regen, J.C.K. Heh and J. Mclick, J. Org. Chem, 1979 44;, 1961. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 113. 158 S.L. Regen, Nov. J. Chem, 1982 629. S.L. Regen, K. Fukunishi and B. Czech, J. Org. Chem, 1981 g, 1218. M. Cinquini, S. Solonna, H. Molinari,F. Montanari and P. Tundo, J. Chem. Soc. Chem. Commun., 1976, 394. H. Molinari, F. Montanari and P. Tundo, J. Chem. Soc. Chem. Commun., 1977, 639 . H. Molinari, F. Montanari, S. Quici and P. Tundo, J. Am. Chem. Soc., 1979 191. 3920. (a) M. Tomoi and W.T. Ford, J. Am. Chem. Soc., 1981 123 3821; (b) M. Tomoi and W.T. Ford, ibid, 1981 103 3828 . F. Montanari and P. Tundo, J. Org. Chem, 1981 4_6_, 2125. Y. Kimura and S.L. Regen, J. Org. Chem, 1983 _4_8 195 . Y. Kimura, P. Kirszensztejn and S.L. Regen, J. Org. Chem, 1983 4_8, 385. M. Tomoi, J. Poly. Sci. Polymer. Chem. Ed., 1985 21,49. P. Tundo, Synthesis, 1978, 315. P. Tundo and P. Venturello, J. Am. Chem. Soc., 1979 1m, 6606. P. Tundo and P. Venturello, J. Am. Chem. Soc., 1981 123, 856. (a) P. Venturello,P. Tundo and E. Angeletti, J. Am. Chem. Soc., 1982 M, 6547; (a) P. Venturello,P. Tundo and E. Angeletti, ibid, 1982 M, 6551 . A. Cornelius and P. Laszlo, Synthesis, 1982, 162. A. kadkhodayan and T. J. Pinnavaia, J. Molec. Cataly., 1983 21, 109. P. Tundo, J. Chem. Soc. Chem. Commun., 1977, 641. P. Tundo and P. Venturello, Tetrahedron Lett., 1980, 2581 . P. Tundo, P. Venturello and E. Angeletti, ls. J. chem, 1985 26, 283. W.T. Ford, J. Lee and P. Tundo, Macromolecules, 1982 Q, 1246. F. Montanari, S. Quici and P. Tundo, J. Org. Chem, 1983 4_8, 199. SL. Regen, Angew. Chem, Int. Ed. Engl, 1979 L8, 421. N. Oktani, C.A. Wikie, A. Nigan and S.L. Regen, Macromolecules, 1981 14, 516. S. Colonna, R. Fomasier and U. Pfeiffer, J. Chem. Soc. Perkin Trans. 1, 1978 2, 8. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 138. 139. 140. 141. 142. 143. 144. 145. 146. 159 N. Kobayashi and K. Iwai, Makromol. Chem. Rapid Commun., 1981 2, 105. S. Banfi, M. Cinquini and S. Colonna, Bull. Chem. Jpn., 1981 _5_4,, 1841. G. Manecke and H. Winter, Makromol. Chem. Rapid C ommun., 1981 2, 569. E. Chiellini, S. D’Antone and R. Solaro, Polym. Prepr. Am. Chem. Soc., Div. Polym. Chem, 1982 23, 179. DC. Sherrington and J. Kelly, Polym. prepr., Am. Chem. Soc., Div. Polym. Chem, 1982 23, 177. DC. Sherrington "Phase Transfer Catalysis in Polymer Chemistry", ed. C.E. Carraher and L. Mathias, Plenum Press, New York, 249 1983 . M. S. Chiles, D.D. Jackson and RC. Reeves, J. Org. Chem, 1980 45, 2915. F. Rolla, W. Roth and L. Homer, Naturwissenschaften, 1977 Q, 337. DC. Sherrington and J. Kelly, Polymer Sci.and Tech., 1984 24, 249. Y. Kawakami and Y. Yamashita, Polymer Sci. and Tech., 1984 24, 263. N. Kobayashi and K. Iwai, J. Org. Chem, 1981 16, 1823. H. Hiemstra and H. Wynberg, J. Am. Chem. Soc., 1981 123, 417. S. Landau Ph.D. thesis " Physical and Catalytic Properties of Hydroxy-Metal Interlayered Smectite Minerals", 1985. DE. Pearson, J. Am. Chem. Soc., 1950 22, 4169. W.A. Jacolis and M. Heidelberge, J. Am Chem. Soc., 1919 4_1_, 2090. EH. Cordes and RB. Dunlap, Acct. Chem. Res., 1969, 329. I.V. Berezin, K. Martinek and A.K. Yatsimirskii, Russian Chemical Reviews, 1973, 787. H. Komeili-Zadeh, H.J.-M. Dou, and J. Metzger, J. Org. Chem, 1978 43,156. 8. Colonna, S. Julia, A. Ginebreda, J. Guixer and A. Tomas, J. C. S. Perkin Trans. 1, 1981, 574. S. Colonna, J. C. S. Perkin Trans. 1, 1978, 371. (a) B. Boyer, G. Lamaty and J.-P. Roque and J. Solofo, Nouv. J. Chim, 1986 l_Q, 553; (b) B. Boyer, G. Lamaty and J .-P. Roque and J. Solofo, ibid , 1986 _1_0_, 559; (c) B. Boyer, G. Lamaty and J .-P. Roque and J. Solofo, ibid, 1986 _1_Q, 563. S. Colonna, S. Julia, J. Guixer J. Masana, J. Rocas, R. Annuziate and H. Molinari, J. Chem. Soc. Perkin Trans. 1, 1982, 1317. 147. 148. 149. 150. 151. 160 H. Wynberg, S. Colonna and ABerto Re, J. Chem. Soc. Perkin Trans. 1, 1981, 547. H. Wynberg and B. Greijdnus, J. C. S. Chem. Comm, 1978, 427. H. Wynberg and H. Pluim, Tetrahedron Lett., 1979, 1251. (a) D. Borrmann and R. Wegler, Chem. Ber., 1966 22, 1245; (b) D. Borrmann and R. Wegler, Chem. Ber., 1967 $9.. 1575. H. Wynberg and A.A. Smaardijk, Unpublished results.