ABSTRACT Title of dissertation: HYPERCOORDINATE SILICON COMPOUNDS IN ORGANIC SYNTHESIS: IMPROVED METHODS FOR THE SYNTHESIS OF ARYL(TRIALKOXY)SILANE DERIVATIVES; AND TRIMETHYLSILYL CYANIDE AS A CYANIDE SOURCE FOR NUCLEOPHILIC SUBSTITUTION Amy Slover Manoso, Doctor of Philosophy, 2004 Dissertation directed by: Professor Philip DeShong Department of Chemistry and Biochemistry Palladium-catalyzed cross-coupling reactions are versatile methods for the synthesis of carbon-carbon bonds. The Stille and Suzuki cross-coupling protocols have achieved prominence in the synthesis of pharmaceuticals and agrochemicals because of the high yields, tolerance for functional groups, and excellent stereoselectivities. However, there are features associated with each of these processes that limit the generality: the Stille tin reagents and byproducts are toxic; and the Suzuki boron reagents can be difficult to synthesize and purify. Environmentally benign arylsilane derivatives have emerged as powerful alternatives to conventional arylmetalloids for the Pd(0)- catalyzed aryl-aryl coupling reaction with organohalides and organo(pseudo)halides because they avoid the inherent limitations associated with traditional methodologies. It was previously reported that aryl(trialkoxy)silanes (also called siloxanes) are substrates for fluoride-promoted, Pd(0)-catalyzed coupling reactions with allylic ester and aryl derivatives providing cross-coupling products in high yields under mild conditions. However, few detailed studies of the synthesis of these useful cross-coupling reagents have been reported in the literature. In chapter one of this thesis, two methods for the synthesis of aryl siloxanes are studied and the optimal reaction conditions and the scope determined: (1) general reaction conditions for the synthesis of aryl(trialkoxy)silanes from aryl Grignard and lithium reagents and functional silanes have been developed; and (2) the scope of the palladium-catalyzed silylation of aryl halides with triethoxysilane to generate aryl(trialkoxy)silane derivatives has been expanded. In tandem, these two methods provide ready access to a wide range of aryl siloxane reagents for use in Pd(0)-mediated cross-coupling reactions, including highly functionalized siloxane intermediates in the synthesis of useful biologically active compounds. In the first part of chapter one, the synthesis of aryl siloxanes from the corresponding aryl organometallic reagent and tetraalkoxysilanes is described. Although examples in the literature have reported the use of a range of silicon electrophiles (including SiCl4 and Cl?Si(OR)3), tetraalkyl orthosilicates (Si(OR)4) allow for the most direct and convenient synthesis of arylsiloxanes, in that they are commercially available, inexpensive, and air and moisture stable. Using the reaction conditions developed herein, o-, m- and p-substituted bromoarenes underwent equally efficient metallation and silylation. Mixed results were obtained with heteroaromatic substrates: 3-bromothiophene, 3-bromo-4-methoxypyridine, 5-bromoindole, and N-methyl-5-bromoindole all underwent silylation in good yield, whereas a low yield of siloxane was obtained from 2-bromofuran, and 2-bromopyridine failed to be silylated. The synthesis of siloxanes via organo lithium and magnesium reagents is limited by the formation of di- and triarylated silanes (Ar2Si(OR)2, and Ar3SiOR, respectively), and dehalogenated (Ar-H) by-products. Lower temperatures allowed for the formation of predominantly monoaryl siloxanes, without requiring a large excess of the electrophile. Optimal reaction conditions for the synthesis of siloxanes from aryl Grignard reagents entailed addition of aryl magnesium reagents to 3 equiv of tetraethoxy- or tetramethoxysilane at ?30 ?C in THF. Aryl lithium species were silylated using 1.5 equiv of tetraethoxy- or tetramethoxysilane at ?78 ?C in ether. The proposed mechanism of silylation involves formation of the anionic pentacoordinate monoaryl(tetraalkoxy)silicate (ArSi(OR)4-), which unexpectedly is susceptible to nucleophilic attack by a second equivalent of the aryl metalloid to form diaryl(dialkoxy)siloxane by-products. The reductive dehalogenation of the aryl halide starting material presumably occurs during the metallation step, and is an inherent limitation of the use of organometallic reagents. The second part of chapter one discusses an alternative to the preparation of arylsilanes from organomagnesium or lithium intermediates: the silylation of aryl halide derivatives by triethoxysilane (H?Si(OEt)3) in the presence of a Pd catalyst. As initially reported in the literature, the silylation reaction was limited to p-substituted, electron-rich aryl iodide substrates. As described in this thesis, a more general Pd(0)-catalyst/ligand system has been developed which activates bromides and iodides: palladium (0) dibenzylideneacetone (Pd(dba)2) is activated with 2-(di-tert-butylphosphino)biphenyl (Buchwald's ligand) (1: 2 mole ratio of Pd : phosphine). Electron-rich, para- and meta- substituted aryl halides (including unprotected anilines and phenols) underwent silylation to form the corresponding aryl(triethoxy)silane in fair to excellent yield; however, ortho- substituted aryl halides failed to be silylated. Aryl chlorides were inert under the reaction conditions, and triflates were poor substrates for silylation, instead undergoing highly efficient reductive deoxygenation. The optimum silylation reagent is triethoxysilane; hexamethoxydisilane failed to be activated under a range of conditions. The major by- product of this reaction is reductive dehalogenation of the aryl halide starting material. Probable mechanisms for the silylation reaction and the reduction side-reaction are presented and discussed. The Pd-catalyzed silylation method is an excellent companion to the more traditional organometallic approach to the formation of the Ar-Si bond. Case in point, ortho-substituted aryl siloxanes are readily synthesized from the Grignard or lithium reagent. Unlike the metallation approach, the Pd-catalyzed silylation technique can be performed in the presence of a wide range of functional groups, including carbonyl- containing electrophiles, and protic moieties such as phenols or primary amines. In addition to the fluoride-promoted transfer of aryl moieties presented in chapter one, silanes have also been shown to transfer nucleophiles such as azide and cyanide anion. Chapter two presents the development of a high yielding silicon-based method for the preparation of alkyl nitriles, which serve as precursors to a variety of useful functional groups. Hypercoordinate cyanosilicate, prepared in situ by the reaction of cyanotrimethylsilane (Me3SiCN) with tetrabutylammonium fluoride (TBAF), is an effective source of nucleophilic cyanide. Primary and secondary alkyl halides and sulfonates underwent rapid and efficient cyanide displacement in the absence of phase transfer catalysts with the silicate derivative; inversion of configuration was observed for optically active alkyl halide substrates. Tetrabutylammonium fluoride was the optimum activating agent, and a full equivalent of fluoride ion was required for reaction completion. A nearly 1 : 1 stoichiometry of substrate to cyanosilicate affected formation of alkyl nitriles in acetonitrile or dioxane; in contrast, traditional methodologies typically employ a large excess of reagents, toxic phase transfer catalysts or solvents such as DMSO, or heavy- metal cyanide salts. Relative to other cyanide sources, the hypercoordinate cyanosilicate was much less basic, thereby mitigating the formation of elimination (alkene) by-products. The Me3SiCN/TBAF system is significantly less reactive and less basic than tetrabutylammonium cyanide (TBA-CN), therefore the mechanism of reaction most like involves the in situ generation of a hypercoordinate cyanosilicate, rather than disproportionation of Me3SiCN and TBAF to form TBA-CN in situ. HYPERCOORDINATE SILICON COMPOUNDS IN ORGANIC SYNTHESIS: IMPROVED METHODS FOR THE SYNTHESIS OF ARYL(TRIALKOXY)SILANE DERIVATIVES; AND TRIMETHYLSILYL CYANIDE AS A CYANIDE SOURCE FOR NUCLEOPHILIC SUBSTITUTION by Amy Slover Manoso Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2004 Advisory Committee: Professor Philip DeShong, Chair/Advisor Assistant Professor Lyle Isaacs Professor Paul Mazzocchi Professor Donald Nuss (Center for Biosystems Research) Professor Lawrence R. Sita ?Copyright by Amy Slover Manoso 2004 ii DEDICATION To my parents, Tina and Bill, husband, Mark sister and her partner, Betsy and Tanya and baby Spence ACKNOWLEDGMENTS I would like to express my sincerest thanks to my advisor, Professor Philip DeShong. Everything that I am today has in some way been influenced by Phil: I am a better bench chemist, scientific writer, teacher, thinker, chemical professional, public speaker and?oddly enough?daughter and spouse due to his example. Thanks for helping me keep my priorities in order, for insisting that I take myself seriously as a chemist, and for teaching me not to panic. iii My husband, Mark, has provided constant encouragement; I can definitively say that he has put in double the effort of any other spouse in the history of the DeShong group, because it has taken me twice as long. Without him, I could not have made it to this point. Thanks for making me finish what I started. I would also like to thank Dr. Jim Demas (University of Virginia) for encouraging me to do undergraduate research and to drop my Latin American Studies major. You were right: organic chemistry is a highly creative and artistic endeavor, and a liberal art in its own right; of course you are a physical chemist and I think at the time you meant it as a snub. I forgive you. Dr. James Whitney (Washington and Lee University) taught me that, when in doubt, the answer is "steric hindrance." Thank you for introducing me to organic chemistry and for being the first on a short list of phenomenal science teachers to show me that excellent science teaching is possible in a large lecture setting. You are deeply missed. I am grateful for the support of my colleagues in the DeShong group. You have been a valuable source of friendship, advice, and commiseration over the past years. I have seen some thirty-odd faces grace the DeShong laboratories, too many to name: ?I don?t know half of you half as well as I should like; and I like less than half of you half as well as you deserve.? ?J. R. R. Tolkien. From the Department, I would like to thank the organic faculty for outstanding teaching and guidance on matters both chemical and personal. Dr. Yui-Fai Lam and Caroline Ladd have provided invaluable assistance in obtaining NMR and mass spectral data. And finally, to my parents: this is the culmination of all of your love and effort, and without you I could not have made it this far. Thank you. iv TABLE OF CONTENTS List of Tables v List of Figures vii List of Schemes viii List of Abbreviations xiii Chapter 1: Improved Synthesis of Aryl(triethoxy)silanes for Use in Palladium-Mediated Cross-Coupling Reactions Introduction 1 Significance of Aryl(trialkoxy)silanes 1 Synthesis of Natural Products Using Aryl(trialkoxy)silane 7 Derivatives Research Goal: Synthesis of Aryl(trialkoxy)silane Derivatives 18 Overview of the Synthesis of Aryl Group Transfer Reagents 20 Preparation of Aryl Stille (Tin) Reagents 23 Preparation of Aryl Suzuki (Boron) Reagents 25 Preparation of Aryl Silicon Reagents 30 Results and Discussion 40 Improved Synthesis of Aryl(trialkoxy)silanes via Treatment of Aryl 40 Grignard or Lithium Reagents with Tetraalkoxysilane Improved Methods for the Synthesis of Aryl(triethoxy)silanes via 59 Palladium(0)-Catalyzed Silylation of Aryl Iodides and Bromides with Triethoxysilane Conclusions 77 Epilogue 79 Experimental 81 Chapter 2: Trimethylsilyl Cyanide As A Cyanide Source For Nucleophilic Substitution Introduction 118 Fluoride Ion Activation of Silicon Bonds 118 Atomic and Molecular Properties of Tetracoordinate and 122 Hypercoordinate Organosilicates Synthesis, Stability and Characterization of Hypercoordinate 127 Organosilicates Mechanisms of Nucleophilic Substitution 130 Reactions at Silicon Fluoride Ion Activation of Silicon Bonds in Organic Synthesis 135 Fluoride Activation of the Si-O Bond 137 Fluoride Activation of the Si-H Bond 140 Fluoride Activation of the Si-C Bond 141 Results and Discussion 145 Background 145 Cyanide Displacements Utilizing Hypercoordinate Cyanosilicate 153 Exploration of the Use of Catalytic Amounts of Fluoride 157 Investigation of Alternative Fluoride Sources 160 Proposed Mechanism 161 Conclusions 170 Epilogue 170 Experimental 175 References 191 v LIST OF TABLES Chapter 1 Table 1. Synthesis of Aryl(trialkoxy)silanes via Treatment of Aryl 45 Grignard or Lithium Reagents with Silicon Electrophiles: Review of the Literature to Date. Table 2. Optimization of the Synthesis of Arylsiloxanes Using 46 4-Bromomagnesium Anisole. Table 3. Synthesis of Aryl(trialkoxy)silanes Using Grignard Reagents. 48 Table 4. Optimization of the Synthesis of Arylsiloxanes Using 4-Lithiotoluene. 51 Table 5. Synthesis of Aryl(trialkoxy)silanes Using Lithium Reagents. 53 Table 6. Optimization of the Silylation of 4-Bromoanisole. 61 Table 7. Silylation of Aryl Bromides. 62 Table 8. Silylation of Heteroaryl Bromides. 63 Table 9. Silylation of Aryl Bromides: Effect of Substituent Position. 64 Table 10. Palladium-Catalyzed Silylation of Aryl Iodides. 66 Table 11. Silylation of Aryltriflate Derivatives. 69 Table 12. Silylation of Arylhalides Using Hexamethoxydisilane. 74 Chapter 2 Table 1. Comparison of Atomic Properties of Si and C. 123 Table 2. Atomic and Group Allred-Rochow Electronegativities. 124 Table 3. Approximate Average Bond Dissociation Energies and 123 Bond Lengths for Tetracoordinate C and Si. Table 4. Comparison of NMR Data for Tetracoordinate and Pentacoordinate 130 Silicates. Table 5. Reactivity of a Series of Fluorosilicate Complexes in the 147 Fluorination of 1-Bromododecane at 85 ?C. Table 6. Reaction of Hypercoordinate Silicate 5 with Alkyl Halides. 154 Table 7. Reaction of Benzyl Bromide (64) with Trimethylsilyl 158 Cyanide (3) and Varying Amounts of TBAF. vi Table 8. Reaction of Benzyl Bromide (64) and Trimethylsilyl 159 Cyanide (3) with Tetrabutylammonium Nucleophiles. Table 9. Reaction of Benzyl Bromide (64) and Trimethylsilyl 160 Cyanide (3) with Alternative Fluoride Sources. Table 10. Reaction of Secondary Alkyl Halides with Trimethylsilyl 161 Cyanide (3) and TBAF or TBAT (1). Table 11. Comparison of Nitrile Synthesis Using Trimethylsilyl 163 Cyanide (3)/TBAF or Tetrabutylammonium Cyanide (89). Table 12. 29Si-NMR Spectral Data for Authentic Samples Used In 166 Mechanistic Studies. vii LIST OF FIGURES Chapter 1 Figure 1. Silicon-Based Aryl Group Transfer Reagents. 2 Figure 2. Hypercoordinate Aryl Group Transfer Reagents. 3 Figure 3. Biaryl Natural Product Targets. 7 Figure 4. Model System For the Synthesis of Colchicine: Fitzgerald?s. 8 Compound (23) Figure 5. Natural Product Targets Featuring the 2,4-Arylpyridine Moiety. 11 Figure 6. Siloxane Derivatives for Application in Natural Product Syntheses. 17 Figure 7. Biaryl Compounds Synthesized Using the Phenyl(trimethoxy)silane 18 Cross-Coupling Methodology. Figure 8. Silicon-Based Aryl Group Transfer Reagents. 31 Figure 9. Siloxanes Prepared via the Grignard Reaction for Use in 50 Natural Product Syntheses. Figure 10. Complexes Formed in the Silylation of Pyridinyl Metal Reagents. 55 Figure 11. 5-Indole Siloxanes Prepared from Aryllithium Reagents for Use 56 in Natural Product Syntheses. Figure 12. Proposed Effect of Substituent Position. 65 Figure 13. Proposed Transition States for the Metathesis Reaction Mechanism. 72 Figure 14. Yields of Siloxanes Using Arylmetalloid Reagents or Palladium- 78 Mediated Silylation. Figure 15. Siloxane Intermediates for Use In Natural Product Syntheses.. 79 Chapter 2 Figure 1. Hybridization of Tetra-, Penta-, and Hexacoordinate Silicon. 125 Figure 2. Comparison of Rates of Expansion of Coordination. 126 Figure 3. Increase in Silicate Stability Due Chelating Ligands. 128 Figure 4. Reaction Coordinates for Single-Step and Two-Step 133 Substitution Processes. Figure 5. Angle of Attack by Nucleophiles on Silanes. 135 Figure 6. Triphenylsilyl Cyanide (93) and TBAF (1 equiv), 29Si Spectrum. 167 viii LIST OF SCHEMES Chapter 1 Scheme 1 1 Scheme 2 2 Scheme 3 3 Scheme 4 4 Scheme 5 5 Scheme 6 5 Scheme 7 5 Scheme 8 6 Scheme 9 7 Scheme 10 9 Scheme 11 9 Scheme 12 10 Scheme 13 10 Scheme 14 12 Scheme 15 12 Scheme 16 13 Scheme 17 14 Scheme 18 15 Scheme 19 15 Scheme 20 16 Scheme 21 19 Scheme 22 20 Scheme 23 20 Scheme 24 21 Scheme 25 22 ix Scheme 26 23 Scheme 27 23 Scheme 28 24 Scheme 29 24 Scheme 30 25 Scheme 31 25 Scheme 32 26 Scheme 33 27 Scheme 34 27 Scheme 35 28 Scheme 36 28 Scheme 37 29 Scheme 38 30 Scheme 39 32 Scheme 40 32 Scheme 41 33 Scheme 42 33 Scheme 43 34 Scheme 44 34 Scheme 45 35 Scheme 46 36 Scheme 47 36 Scheme 48 37 Scheme 49 37 Scheme 50 38 Scheme 51 39 Scheme 52 40 Scheme 53 41 x Scheme 54 47 Scheme 55 49 Scheme 56 50 Scheme 57 54 Scheme 58 56 Scheme 59 57 Scheme 60 59 Scheme 61 67 Scheme 62 70 Scheme 63 70 Scheme 64 71 Scheme 65 72 Scheme 66 73 Scheme 67 73 Scheme 68 75 Scheme 69 75 Scheme 70 76 Scheme 71 77 Scheme 72 80 Chapter 2 Scheme 1 119 Scheme 2 119 Scheme 3 120 Scheme 4 120 Scheme 5 127 Scheme 6 131 Scheme 7 131 xi Scheme 8 132 Scheme 9 133 Scheme 10 134 Scheme 11 136 Scheme 12 137 Scheme 13 138 Scheme 14 138 Scheme 15 139 Scheme 16 139 Scheme 17 139 Scheme 18 140 Scheme 19 140 Scheme 20 141 Scheme 21 141 Scheme 22 142 Scheme 23 142 Scheme 24 143 Scheme 25 144 Scheme 26 146 Scheme 27 148 Scheme 28 148 Scheme 29 149 Scheme 30 150 Scheme 31 151 Scheme 32 152 Scheme 33 157 Scheme 34 158 Scheme 35 162 xii Scheme 36 164 Scheme 37 164 Scheme 38 165 Scheme 39 168 Scheme 40 168 Scheme 41 169 Scheme 42 171 Scheme 43 171 Scheme 44 172 Scheme 45 173 xiii LIST OF ABBREVIATIONS 18-crown-6 1,4,7,10,13,16-Hexaoxacyclooctadecane Ac acetyl ap apical aq. aqueous Ar aryl Bn benzyl BOC t-butyloxycarbonyl BTAF benzyl(trimethyl)amino fluoride Bu butyl Bz benzoyl cat catecholate cod cyclooctadienyl cy cyclohexyl dba dibenzylideneacetone DBU 1,8-diazabicyclo[5.4.0]undec-7-ene DDQ 2,3-dichloro-5,6-dicyano-1,4-benzoquinone DIEA N,N-diisopropylethylamine DMAP 4-dimethylaminopyridine DMF N,N-dimethylformamide DMSO dimethyl sulfoxide DOM directed ortho metallation dppf 1,1'-bis(diphenylphosphino)ferrocene dppp 1,3-bis(diphenylphosphino)propane EDG electron donating group ee enantiomeric excess eq equatorial equiv equivalent (s) xiv Et ethyl Et2O diethyl ether EWG electron withdrawing group ? gyromagnetic ratio G. C. gas chromatograph h hour(s) HMPA hexamethylphosphoramide HOMO highest-occupied molecular orbital Hz Hertz i-Pr isopropyl IR infrared J coupling constant Ln ligand LUMO lowest-unoccupied molecular orbital m meta M+ molecular ion m/z mass-to-charge ratio Me methyl MeCN acetonitrile MEM 2-methoxyethoxymethyl MHz megahertz min minute(s) MOM methoxymethyl mp melting point MS mass spectrometry Ms methanesulfonyl Naph naphthyl NMP 1-methyl-2-pyrrolidinone xv NMR nuclear magnetic resonance o ortho OAc acetate p para Ph phenyl Pr propyl pyr pyridinyl PZ phosphazenium Rf retardation factor rt room temperature SET Single-electron transfer SN1 substitution nucleophilic unimolecular SN2 substitution nucleophilic bimolecular t-Bu tertiary butyl T1 spin-lattice relaxation time TAS-F tris(dimethylamino)sulfur(trimethylsilyl)difluoride TBAF tetrabutylammonium fluoride TBAT tetrabutylammonium triphenyldifluorosilicate TBP trigonal bipyramidal Tf trifluoromethanesulfonyl THF tetrahydrofuran TLC thin layer chromatography TMS trimethylsilyl tol tolyl UV ultraviolet 1 Chapter 1. Improved Synthesis of Aryl(triethoxy)silanes for Use in Palladium- Mediated Cross-Coupling Reactions Introduction Significance of Aryl(trialkoxy)silanes The metal-catalyzed cross-coupling reaction for the formation of carbon-carbon bonds between unsaturated centers is an indispensable synthetic tool for the preparation of useful industrial1 and pharmaceutical2-9 materials. Of the many combinations of organometallic nucleophiles and organic electrophiles in the literature,10-12 the Stille (organostannane) and Suzuki (organoborane) coupling methodologies are the most widely employed for the synthesis of unsymmetrical biaryl derivatives and substituted alkenes due to the generally excellent yields, high stereoselectivities, and superior functional group tolerance (Scheme 1).12-20 Nonetheless, arylsilane derivatives have emerged as powerful alternatives to conventional arylmetalloids for the Pd(0)-catalyzed aryl-aryl coupling reaction with organohalides and organo(pseudo)halides because they avoid the inherent limitations associated with traditional methodologies: the Stille tin reagents and byproducts are toxic; and the Suzuki boron reagents can be difficult to synthesize and purify.8,21-59 Scheme 1 X CH3 SnBu3/B(OH)2 CH3 Si(OEt)3 SnBu3/B(OH)2 Si(OEt)3 H3C X H3C Pd(0) Pd(0), F-X = I, Br, Cl, OTf Pd(0) Pd(0), F- X = I, Br, Cl, OTf 2 Previous studies from our group21-25 as well as Hiyama,8,26-33 Ito,34,35 Denmark,36-44 and others45-56 have shown that a variety of silicon derivatives (i.e., compounds 1-4, Figure 1) undergo fluoride-mediated, Pd(0)-catalyzed aryl group transfer reactions.57-59 SiMeF2 OCH3 Si(OH)2Me Me Si Cl MeO Si(OEt)3 O2N 1 2 3 4 Figure 1. Silicon-Based Aryl Group Transfer Reagents. Each of the above silicon compounds participates in the Pd-catalyzed cross- coupling reaction only upon activation by a nucleophile, typically fluoride ion. The mechanism of action of these silicon reagents is similar. As illustrated in Scheme 2, the process is proposed to involve the in situ formation of the reactive hypercoordinate intermediate 6 by attack of fluoride on the tetracoordinate aryl silicon reagent 5.9,53,54 Conversion of the neutral tetracoordinate silane 5 to anionic hypercoordinate silicate 6 both lengthens and polarizes the silicon-ligand bonds, placing a large partial positive charge on silicon and a large partial negative charge on the arene and the other ligands.60-62 Scheme 2 Si R1 R2 R3 Si R1R2 R3 F F 65 The overall catalytic cycle for aryl-aryl coupling begins with formation of Pd(II) species 7 as a result of oxidative addition of Pd(0) into the aryl?X bond of the aryl halide substrate (Scheme 3). As described above, unlike the weakly polarized tetracoordinate precursor 5, the anionic hypercoordinate fluorosilicate 6 undergoes transmetallation with palladium complex 7 to form the bis-aryl Pd(II) species 9, with loss of R3SiFX? (8). Lastly, reductive elimination of Pd(0) regenerates the catalyst, and liberates the biaryl product. 3 Scheme 3 Si R1 R2 R3 Si R1R 2 R3 F F 79 X Pd(II) XPd(II) R RR 6 5 Si R1R2 R3 X F 8 Ln L n Pd(0)Ln R The most compelling evidence of the participation of a hypercoordinate silicate intermediate such as 6 is the ability of analogous isolable organosilicates to participate in the cross-coupling reaction in the absence of additional fluoride. Silicates 10 and 11 have been shown to be effective phenylating22 and vinylating63-65 reagents, respectively (Figure 2). In addition, catecholate derivatives such as 11 have been employed for the transfer of other groups, for example alkenyl and aryl moieties.65-68 Si OO O O Si PhPh F F Bu4N+ Ph Et3NH 11TBAT (10) Figure 2. Hypercoordinate Aryl Group Transfer Reagents. 4 The initial studies in the DeShong laboratories concerning the cross-coupling reactions of silicon reagents employed compound 10, tetrabutylammonium triphenyldifluorosilicate ([Ph3SiF2]?[Bu4N]+, TBAT).22,67 Under palladium catalysis, good yields of coupled products were obtained with aryl iodides, most aryl triflates, and electron-deficient aryl bromides (Scheme 4). In addition to forming the desired heterocoupled product 12, small amounts of homocoupled product 13 were also isolated.22 The major limitations of the TBAT cross-coupling methodology are (1) poor atom economy (under standard reaction conditions only one phenyl group out of three is transferred); and (2) substituted difluorotriaryl analogs of 10 are not easily synthesized.67 Scheme 4 Si PhPh F F TBAT (10) Ph Bu4N+ I Ph Pd(dba)2 O O O O + + 12 13 96% 4% In an effort to overcome the limitations associated with TBAT, phenyltrimethoxysilane (15, Scheme 5) was investigated as a less wasteful, and more easily derivatized arylating reagent.21,23-25 Siloxane derivatives such as 15 are non-toxic, hydrolytically stable compounds, whose synthesis and characterization have been reported most notably in the fields of polymer and materials chemistry.69,70 Recently, the DeShong research group described the synthesis of unsymmetrically substituted biaryls via phenylation of aryl iodides,21,24 bromides and chlorides23,24 with commercially available phenyltrimethoxysilane (15). For example, 2-bromopyridine (14) was cross-coupled with phenyltrimethoxysilane (15) to form biaryl 16 in 76% isolated yield (Scheme 5).23 5 Scheme 5 Si(OMe)3 Pd(OAc)2 / PPh3 TBAFN Br N+ 14 15 16 76% In congruence with the general mechanism depicted in Scheme 3, the assumed reactive silicon species is hypercoordinate fluorosilicate 17 (Scheme 6).21,67 Scheme 6 SiMeOMeO OMe F 17 H3C I Pd(dba)2 H3C F- Si(OMe)3 15 Recent efforts have focused on the cross-coupling reaction of arylsiloxane derivatives with aryltriflate substrates. Aryl triflates, readily derived from the phenol,71 are often more available than the corresponding aryl halides; however, siloxanes have failed to undergo efficient cross-coupling with aryl triflates (Scheme 7). Under standard cross- coupling conditions, the base TBAF induces triflate hydrolysis by liberating methoxide from the siloxane starting material. Preliminary studies indicate that hydrolysis can be mitigated by the addition of water, which presumably dilutes and solvates the offending alkoxide.72 Scheme 7 Si(OMe)3 Pd(dba)2 / PR3 TBAF+ 15 MeO OTf MeO + MeO OH 10% 90% 6 Very recently, siloxane derivatives bearing chelating alkoxy ligands were found to undergo Pd-catalyzed cross-coupling with aryl and heteroaryl triflates in the presence of a fluoride source in good to excellent yields.68,72 It is proposed that the alkoxide ligands are stabilized by the chelate effect, thereby attenuating triflate hydrolysis. Triethylammonium (bis)catechol phenyl silicate ([PhSi(cat)2]?[Et3NH]+, 18) and phenyl silatrane (19) derivatives are synthesized in high yield from the corresponding trialkoxysilane via transesterification to form crystalline, air and moisture stable complexes (Scheme 8). The stability of compounds 18 and 19 to hydrolysis and alcoholysis is well established.73-79 Scheme 8 Si(OMe)3 HO HO NEt3 MeOH N OH 3 toluene SiOO N O 19 Si OO O O 18 HNEt3 15 Silicates 18 and 19 were found to participate in the fluoride-activated, Pd-mediated cross-coupling reaction with aryl iodides, bromides and most notably triflates.68,72 The reaction tolerates a wide range of electron-withdrawing and electron-releasing substituents, as well as heteroaromatic systems. For example, aryl triflate 20 was cross- coupled with catecholate 18 to form biaryl 21 in 96% isolated yield (Scheme 9). 7 Scheme 9 Si OO O O 18 HNEt3 + OMe OTf 20 Pd(dba)2 / PR3 TBAF OMe 96% 21 Synthesis of Natural Products Using Aryl(trialkoxy)silane Derivatives To further demonstrate the generality of the hypercoordinate silicate cross-coupling reaction, preliminary studies are underway to apply this technique to the synthesis of complex biaryl natural products. Targets include the anti-inflammatory and anti-mitotic drugs colchicine (22),80,81 and Fitzgerald?s compound (23)82 and the antitumor 2-quinoline-2-pyridyl derivatives streptonigrin (24),83,84 and lavendamycin (25) (Figure 3).85,86 N O O H2N MeO N CH3H2N CO2H OMe OMe OH N O O H2N N CO2H Streptonigrin (24) Lavendamycin (25) CH3HN A B C D A B C D E 2 2 4 4 MeO M eO M eO NHAc OMeO Colchicine (22) O OMe OMe OMeMeO Fitzgerald's Compound (23) Figure 3. Biaryl Natural Product Targets. 8 Each of these natural product syntheses will have as the key biaryl carbon- carbon bond forming step a silicon-based cross-coupling reaction. This, in turn, will necessitate the synthesis of highly functionalized aryl(trialkoxy)silane and aryl halide coupling partners. The biaryl skeletons of colchicine,87-97 Fitzgerald?s compound,98 streptonigrin99-101 and lavendamycin102-109 have been synthesized previously by classical Stille and Suzuki cross-coupling methodologies; for each target, the previously reported synthetic approach will serve as a standard for our synthetic plan. Colchicine (22) is a commonly prescribed drug for the treatment of gout.80 This alkaloid has more recently been found to have taxol-like antitumor properties, in that it binds to ?-tubulin preventing cellular spindle formation and division.81 As a synthetic target, the colchicine skeleton was particularly intriguing to the DeShong group: the siloxane cross-coupling methodology had never before been used for the formation of carbon-carbon bonds between arenes and alternative aromatic systems, such as the tropolone moiety. Approaches to the synthesis of colchicine (22) in our laboratories have focused on 5-(trimethoxyphenyl)tropolone methyl ether (Fitzgerald?s compound, 23, Figure 4) as a model system for the formation of the aryl-tropolone bond. The synthesis of 23 is a worthwhile endeavor in its own right because Fitzgerald?s compound displays anti-mitotic properties comparable to those of colchicine.82 O OMe OMe OMeMeO Fitzgerald's Compound (23) Figure 4. Model System For the Synthesis of Colchicine: Fitzgerald?s Compound (23). 9 Compound 23 was prepared in excellent yield by Banwell through a Suzuki reaction employing 5-bromotropolone methyl ether (26) and the trisubstituted aryl boronic acid 27 (Scheme 10).98 It is noteworthy that attempts to synthesize compound 23 and simpler analogs via the Stille method consistently gave lower yields than the corresponding Suzuki approach. Aryl boronic acids remain the reagent of choice for the formation of the aryl-tropolone bond.98,110 In most cases, failure of the Stille reaction to provide the cross-coupled adduct was attributed to steric hindrance by the methoxy substituent ortho to the desired aryl-tropolone bond. Scheme 10 O OMe Br B(OH)2 OMe OMe MeO Pd(PPh3)4 Na2CO3 92% + 26 27 O OMe OMe OMeMeO 23 A similar tactic is proposed for the synthesis of 23 via the siloxane cross-coupling technology (Scheme 11). This approach requires the synthesis of complex aryl siloxane derivative 28; bromotropolone 26 has been prepared previously.98 Scheme 11 O OMe Br Si(OR)3 OMe OMe MeO+ 26 O OMe OMe OMeMeO 23 28 10 Handy111 and Seganish68 in the DeShong laboratories have probed the viability of using silicon-based reagents for the formation of the phenyl-tropolone bond via Pd(0)- catalyzed cross-coupling. To alleviate the steric demand on the cross-coupling of hindered siloxane 28 (Scheme 11), the simple phenylation of tropolone derivatives was investigated (Scheme 12). Handy demonstrated that MEM-protected bromotropolone 29 was unreactive to cross-coupling with phenyltrimethoxysilane (15) under standard and modified reaction conditions; in all cases, starting material was recovered.111 Scheme 12 O OMEM Br 29 Si(OMe)3 Pd(0) / PR3 + 15 O OMEM 30 TBAF X Seganish reported the high yielding phenylation of 2-tropolone trifluoromethanesulfonate (31) with triethylammonium (bis)catechol phenyl silicate (18).68 Seganish continues to explore the use of silicon catecholates for the arylation of substituted tropolone derivatives, with the ultimate goal of completing the total synthesis of colchicine as outlined in Scheme 11. Scheme 13 Si OO O O 18 HNEt3+ Pd(dba)2 / PPh3 TBAF 89% O OTf 31 O 32 11 Streptonigrin (24)83,84 and lavendamycin (25)85,86 are structurally similar antitumor antibiotics: both have at their core a highly substituted pyridine moiety (ring C), bound at C?2 and C?4 to a functionalized quinone and a substituted arene, respectively (Figure 5). As a synthetic target, this skeleton was of particular interest to the DeShong group: the siloxane cross-coupling methodology had not yet been investigated for the formation of carbon-carbon bonds between two heteroaromatic systems, such as pyridines and quinones. More importantly, our goal was to address a major limitation of the Suzuki reaction, which has been demonstrated to be ineffectual for the cross-coupling of heteroaromatic derivatives.112 Approaches to the synthesis of streptonigrin (24) and lavendamycin (25) in our laboratories have concentrated on nitramarine (33) as a model system for the construction of the 2-quinoline-2-pyridyl systems. In addition, nitramarine and lavendamycin have in common the ?-carboline (C?D?E) ring system. N O O H2N MeO N CH3H2N CO2H OMe OMe OH N O O H2N N CO2H Streptonigrin (24) Lavendamycin (25) CH3HN A B C D A B C D E 2 2 4 4 N N Nitramarine (33) HN A B C D E 2 2 4 4 2 2 4 4 Figure 5. Natural Product Targets Featuring the 2,4-Arylpyridine Moiety. Focusing on the formation of the quinolinyl-pyridyl (C?2/B?2) bond of nitramarine, compound 33 was prepared in good yield by Queguiner via Stille coupling of 2-chloropyridine derivative 34 with 2-(trimethylstannyl)quinoline 35 (Scheme 14); ?-carboline ring (ring D) closure of adduct 36 completed the synthesis.113 12 Scheme 14 NCl F NH(BOC) 34 N SnMe3 Pd(0) Ba(OH)2 78% N F NH(BOC) N 36 ?, pyr?HCl 83%+ N N 33 HN 35 A similar tactic is proposed for the synthesis of 33 via the siloxane cross-coupling technology (Scheme 15). This approach requires the synthesis of the quinoline siloxane 38; the ?-carboline halides and sulfonates (37) have been prepared previously.113 Alternatively, the complex pyridyl siloxane 39 could be prepared and coupled with the quinoline halide or sulfonate (40). Scheme 15 N N Nitramarine (33) HN N 37 HN N 38 + TfO/Br Si(OR)3 N 39 HN N 40 + (RO)3Si Br/OTf 13 Handy,24,111 and Seganish68 in the DeShong laboratories have probed the viability of using silicon-based reagents for the formation of the C?2/B?2 bond via Pd(0)- catalyzed cross-coupling. For these exploratory studies, the simple phenylation of pyridine and quinoline derivatives was investigated. Mowery had previously shown that 2-bromo and 3-bromopyridine (41 and 43) readily underwent phenylation using the fluoride-activated, Pd(0)-catalyzed siloxane methodology (Scheme 16, eqs. a and b).23 In parallel, 2- and 3-pyridyl triflates were cross-coupled in good yield by Seganish using the catecholate reagent 18 (Scheme 16, eqs. c and d).68 Scheme 16 [PhSi(cat)2]-[Et3NH]+, 18 TBAF N OTf N N OTf N 45 46 42 44 Pd(dba)2 / PR3 73% [PhSi(cat)2]-[Et3NH]+, 18 TBAF Pd(dba)2 / PR3 78% N Br N N Br N 41 43 42 44 (a) (b) (c) (d) PhSi(OMe)3, 15 TBAF Pd(OAc)2, PPh3 76% PhSi(OMe)3, 15 TBAF Pd(OAc)2, PPh3 62% 14 As illustrated in Scheme 17, Handy recently demonstrated the efficient cross- coupling of phenyltrimethoxysilane (15) and 2-bromoquinoline (47) to form 2-phenylquinoline (48, Scheme 17, eq. a).111 In turn, 2-quinoline triflate 49 was cross- coupled in excellent yield with (bis)catechol phenyl silicate 21 by Seganish (Scheme 17, eq. b).68 Scheme 17 N Br 47 PhSi(OMe)3, 15 TBAF Pd(OAc)2, PPh3 80% N 48 N OTf 49 [PhSi(cat)2]-[Et3NH]+, 18 TBAF Pd(OAc)2, PPh3 91% N 48 (a) (b) Encouraged by the initial results demonstrating the successful phenylation of simple pyridine and quinoline compounds, efforts are underway to develop the cross- coupling methodology for the formation of the C4?D4 and C2?B2 heteroaryl-heteroaryl bonds present in both streptonigrin (24) and lavendamycin (25). For streptonigrin, the proposed synthesis will have as its key steps (1) the silicon-based pyridyl-aryl cross- coupling of highly substituted 4-bromopyridine 52 with sterically encumbered arylsiloxane derivative 53; and (2) the subsequent silicon-based cross-coupling of the complete C?D ring system (51) with 2-(trialkoxysilylpyridyl)-5,8-quinone 50 (Scheme 18).112 15 Scheme 18 N OMe OMe BOCHN MeO N CH3BOCHN CO2R OMe OMe OMOM TfO/Br Si(OR)3 + N CH3BOCHN CH3 OMe OMe MeO Br Si(OR)3 + Streptonigrin (24) 50 51 52 53 OMOM N O O H 2N MeO N CH3H2N CO2H OMe OMe OH A B C D C D A B C D For lavendamycin (25), the proposed synthesis will have as its key steps (1) the silicon-based cross-coupling of highly substituted 4-bromopyridine 57 with ortho- substituted arylsiloxane derivative 58; and (2) the subsequent silicon-based cross- coupling of the complete C?D?E ring system (56) with 2-(trialkoxysilylpyridyl)-5,8-quinone 55 (Scheme 19).112 Scheme 19 N O O H2N N R 25, R = COOH, Lavendamycin 54, R = CH3 CH3HN A B C D E + + N OMe OMe R2N N CH3 CH3HN A B C D E Si(OR)3 55 TfO/Br 56 N CH3 Br CH3F C Cl 57 E 58 BOCHN Si(OR)3 16 Toward the goal of constructing the C?D ring system (51) of streptonigrin, McElroy has extended the scope of Pd-catalyzed siloxane coupling to include the efficient phenylation of functionalized 4-bromopyridines.112 Under palladium catalysis, hindered 4-bromopyridine derivative 59 was cross-coupled in good yield with TBAT (10) (Scheme 20, eq. a); however, when phenyltrimethoxysilane and TBAF were employed as the phenylating reagent with electron-deficient substrate 59, low yields of adduct 60 were obtained. Gratifyingly, electron-rich substrate 61 underwent high yielding phenylation with phenyltriethoxysilane (Scheme 20, eq. b). Scheme 20 NH3CO O2N CH3 Br CH3 TBAT (10) NH3CO O2N CH3 CH3 Pd(OAc)2, PPh3 67% 59 60 PhSi(OMe)3, 15 TBAF Pd(OAc)2, PPh3 89% NH3CO H3C CH3 Br CH3 61 NH3CO H3C CH3 CH3 62 (a) (b) In conclusion, for these proposed natural product syntheses to be successful, a general, efficient synthesis of highly functionalized aryl siloxanes is required. The siloxanes needed for the synthesis of colchicine (22), Fitzgerald?s compound (23), streptonigrin (24), lavendamycin (25), and nitramarine (34) are summarized, in Figure 6. 17 N 38 Si(OR)3 N 39 NH Si(OR)3 Si(OR)3 OMe OMe MeO 28 OMe OMe Si(OR)3 53 OMOM 58 BOCHN Si(OR)3 N OMe OMe R2N Si(OR)3 55 N OMe OMe BOCHN MeO Si(OR)3 50 Figure 6. Siloxane Derivatives for Application in Natural Product Syntheses. To date, the DeShong group has demonstrated the Pd-catalyzed fluoride- activated phenylation by phenyl(trialkoxy)silane derivatives of a range of electron-rich and electron-deficient aryl halides and triflates, including sterically hindered and heteroaromatic substrates.21,23-25 A sampling of the biaryl compounds that have been synthesized in our laboratories is shown in Figure 7. The syntheses of heteroatomic, aryloxy, and Lewis basic biaryl derivatives are particularly noteworthy because these types of compounds are difficult to access via the corresponding Suzuki reaction: the synthesis and reactions of arylborane reagents in the presence of Lewis basic moieties are often stymied by their interaction with the highly Lewis acidic boron center. 18 H3C CH3 O O OCH3 CH3 NHR NMe2 HN R R = Ac, COOMe NO2, Cl, CHO CN, t-Bu R = H, Ac, Ts N S o, m, p o, m, p 2, 32, 3 N O O C16H33 N OMe R CH3 CH3 R = NO2, H, Me Figure 7. Biaryl Compounds Synthesized Using the Phenyl(trimethoxy)silane Cross- Coupling Methodology. The research described above has already established the relative mildness, wide ranging functional group compatibility, and ease of work-up of the siloxane-based cross-coupling process for the transfer of phenyl.21,23-25 However, until the facility of synthesizing of aryl(trialkoxy)silane derivatives, and the ability of our siloxane methodology to transfer aryl groups other than phenyl has been established, the generality of the siloxane-based cross-coupling methodology remains unproven.41 Research Goal: Synthesis of Aryl(trialkoxy)silane Derivatives The goal of the research described herein has been the development and optimization of methods for the synthesis of aryl(trialkoxy)silane reagents. This study necessitated the synthesis of ortho, meta, and para-substituted, electron-rich and electron-deficient aryl siloxanes. Additionally, aryl(trialkoxy)silanes are needed as 19 precursors to the corresponding aryl(biscatecholate) or arylsilatrane derivatives required for the cross-coupling with aryl triflates. The penultimate goal is the development of methods for the synthesis of complex aryl siloxane intermediates needed for the construction of biaryl natural products. Scheme 21 X R X = I, Br,Cl, OTf Si(OEt)3 R SiOO N O Si OO O O HNEt R R R R' R' x Pd(0) X = I, Br, Cl, OTf In their dissertations on the Pd-mediated reactions of hypercoordinate silicates, Mowery67 and Handy111 reviewed the mechanism and scope of the cross-coupling reactions of aryl silicon, boron, and tin reagents. The aim of this chapter is to review existing methods for the synthesis of aryl trialkoxysilane, boron, and tin reagents for use in the Pd-mediated cross-coupling reaction. Finally, the results of studies in the DeShong lab on improved methods for the synthesis of aryl(trialkoxy)silanes are presented and discussed. 20 Overview of the Synthesis of Aryl Group Transfer Reagents Focusing on aryl silicon, boron, and tin reagents, two general approaches exist for the synthesis of these materials: (1) the most common and economical method is the treatment of an aryl Grignard or lithium reagent with a silicon, boron or tin electrophile (Scheme 22); Scheme 22 H3CO M H3CO B(OEt)2 M = MgBr, Li H3CO SiXnR3-n H3CO SnBu3 SiXnR4-n Cl SnBu3 X = F, Cl, Br R = alkoxy, alkyl B(OEt)3 and (2) when an incompatibility with organic functional groups arises, the transition metal- mediated silylation, borylation or stannylation of an aryl halide (vide infra) (Scheme 23).114 Scheme 23 H3CO X H3CO B(OMe)2 H3CO Si(OEt)3 H3CO SnBu3 Pd(0) or Rh(I) X = I, Br, Cl, OTf Pd(0) Pd(0) (MeO)2B-B(OMe)2 H-Si(OEt)3 Bu3Sn-SnBu3 Other miscellaneous methods have been reported, however, the above methods remain the most widely employed pathways to the synthesis of aryl cross-coupling reagents.10 The availability of multiple methods for the synthesis of any given reagent adds to the versatility and overall strength of the corresponding cross-coupling methodology. 21 Both approaches (Schemes 22 and 23) have their inherent strengths and limitations.10 Most notably among their strengths, these methods are regiospecific, allowing for the direct installment of the desired tin, boron, or silicon electrophile at the site of the halogen. Metallation?either via formation of the aryl Grignard or aryl lithium reagent?is a relatively inexpensive, scalable standard laboratory technique, and the desired arylmetalloid can be generated conveniently, in typically fair to excellent yields using previously reported procedures.115 Subsequent treatment of the aryl metalloid intermediate with a silicon, boron, or tin electrophile gives the desired product in fair to excellent yield, as described below. That said, the metalloid reaction is limited by the problems associated with the synthesis of aryl metalloids having electrophilic functional groups (i.e. esters, ketones, etc.).116-122 Also, protic functional groups must be protected prior to metallation. Highly reactive aryllithium species must be generated at low temperature (between ?78 and ?110 ?C), and often decompose before undergoing further transformation. The reaction of multifunctional silicon, boron, or tin electrophiles with highly reactive aryl metalloid derivatives is typically complicated by the formation of polyarylated products (Scheme 24, eq. a).123 In addition, homocoupling of the arylmetalloid can occur, resulting in the formation of biaryl contaminants (eq. b), and lithium-halogen exchange can be complicated by competing alkylation (eq. c).115 This overview will focus on known methods for the preparation of tin, boron, and silicon cross-coupling reagents; methods for the generation of aryl Grignard124,125 and lithium126-129 intermediates are reviewed elsewhere.115,130 Scheme 24 B(OMe)3 H3O + ArLi Ar B(OH)2 (Ar)2B OH+ +-78 ?CEt 2O 2 Ar-MgBr -78 ?CEt 2O Ar-Ar Ar-Br + RLi -78 ?CTHF Ar-R (a) (b) (c) 22 In turn, the Pd-mediated approach summarized in Scheme 23 (above) is relatively expensive and often limited by the lack of generality (vide infra); for example, different conditions (catalyst, solvent, base) are typically required for electron-deficient and electron-rich aryl halide substrates. In addition, chloroarenes are notoriously unreactive under typical cross-coupling conditions. Nevertheless, the Pd-mediated approach is an invaluable foil to the organometallic approach, given its relative mildness and functional group tolerance. Lastly, both methods require an aryl halide starting material: generation of the aryllithium by metal-halogen exchange, the Grignard under standard or Barbier reaction conditions, or the reactive Ar?Pd(II)?X intermediate necessitates the synthesis of the required aryl iodide or bromide precursor (Scheme 25). The regiospecific halogenation of a functionalized aromatic system can be problematic, and often requires multiple steps. Lastly, it is difficult to achieve the regioselective monolithiation of polyhalogenated aromatics; consequently, multiple products?both regio- and polyfunctionalized isomers?are obtained upon the addition of the electrophile.126-129 Scheme 25 Br LiR-Li MgBrMg0 Pd(0) Pd(II) X Ln E+ E+ EMeO MeO MeO MeO MeO Y E Aryllithium reagents can be generated in the absence of a halogen substituent on the arene substrate by directed ortho metallation (DOM).131-137 The DOM reaction involves chelation of a lithium transfer reagent to a Lewis basic group; this chelation results in deprotonation and lithiation at a position ortho to the directing group (Scheme 26). 23 Scheme 26 MeO R-Li E+MeO Li MeO E As with lithium-halogen exchange, DOM is a powerful method for the regioselective formation of aryllithium reagents; however, DOM is limited to substrates with a geometrically available heteroatom.132 Also, the presence of bulky substituents meta to the desired site of lithiation can compromise the regioselectivity of the reaction through steric interference;133-137 lastly, other substituents on the substrate which are capable of chelation can compete as metal-directing groups. Graduate student Michael Seganish is currently exploring the synthesis of ortho-substituted aryl(trialkoxy)silane derivatives utilizing directed ortho metallation. Preparation of Aryl Stille (Tin) Reagents Despite their toxicity, organotin reagents are prized for their ease of synthesis, handling and storage; they are stable to moisture and oxygen, and can be purified by distillation or C?18 flash column chromatography prior to use.138,139 In addition, aryl stannanes are compatible with a wide variety of organic functional groups, making the use of tedious protecting group strategies unnecessary. Aryltributyl? and ?trimethytin reagents are typically synthesized from the corresponding aryl halide by the in situ conversion to a reactive organolithium or organomagnesium species followed by treatment with trialkyltin chloride (Scheme 27).19,138,140 The synthesis of organotin compounds via lithium and Grignard reagents has been recently reviewed.138,141-143 Scheme 27 Me2N Br 1. BuLi THF, -78 ?C 2. Bu3SnCl 90% Me2N SnBu3 24 When electrophilic organic functional groups preclude the organometalloid method, the aryl halide substrate may be cross-coupled under Pd(0) catalysis with hexaalkyldistannanes (Scheme 28).19,138,144 The coupling approach is generally high- yielding, albeit limited to aryl iodides and bromides and intolerant of p-nitro and p-amino substituents on the aryl ring. The only detectable side reaction is homocoupling of the electrophile.19 Scheme 28 MeO I Me3Sn-SnMe3 Pd(PPh3)2Br2 toluene, 115 ?C 96% MeO SnMe3 In place of a hexaalkyldistannane, tributyltin hydride was recently shown to couple with most aryl iodides under Pd(0) catalysis in the presence of a weak base (Scheme 29).145 . Two byproducts were observed: the homocoupled product (Ar?Ar), and the reduced (dehalogenated) arene (Ar?H). Aryl bromides were unreactive. Scheme 29 H2N I Bu3Sn-H PdCl2(PMePh2)2 NMP, 25 ?C 73% H2N SnBu3 Less common methods for the synthesis of aryl tin derivatives include the photo- induced stannylation of mono and polychlorobenzenes using trimethylstannylsodium,146-149 and a Diels-Alder approach.143,150 A variety of ortho, meta and para-substituted aryl, pyridyl, and quinolinyl chlorides are substituted by Me3Sn? ions in liquid ammonia under irradiation to give the substitution product in high yields (Scheme 30).146 The reaction is limited to aryl chlorides; aryl bromide and iodide derivatives undergo dehalogenation (reduction) under the reaction conditions.151 Despite its limited scope, the formation of the 25 tin reagent from the corresponding chloride is noteworthy in that aryl chlorides?which are more readily obtained than the corresponding bromide or iodide?cannot be readily converted to the stannane using either the metallation or the coupling approach. Scheme 30 Z Cl + Na SnMe3 h?NH 3 Z SnMe3 Cl SnMe3 Z = CH, o-, m-, p- Z = N, 2,5-, 3,5-, 3,6- 58%, 90%, 88% 88%, 80%, 86% Lastly, the Diels-Alder reaction between methyl tributylstannylpropiolate (64) and alkyl substituted 1,3-butadienes (63) forms 1,4-cyclohexadienyl tin derivatives (65) in good yields; subsequent aromatization (elimination) gives the aryl tin reagent (66) (Scheme 31).143,150,152 The scope of this reaction is severely narrowed by a number of factors including (a) the often low regioselectivity of the cycloaddition reaction, and (b) only simple alkyl-substituted dienes and electron-deficient dienophiles undergo cyclization. Scheme 31 + CO2Me SnBu3 125 ?C, 50h 71% CO2Me SnBu3 DDQ CO2Me SnBu3 63 64 65 66 Preparation of Aryl Suzuki (Boron) Reagents Organoboron reagents are a non-toxic alternative to organotin reagents, and have essentially superceded the use of tin compounds in the Pd-mediated cross-coupling reaction for the formation of biaryls.153 As with tin reagents, Suzuki reagents are valued for their ease of handling and storage, due to their stability to moisture and oxygen.139 Unlike 26 tin reagents, organoboranes are more difficult to synthesize and purify (they are often used as the crude isolate), however much work has been dedicated toward the development of efficient routes to the formation of these highly useful synthetic intermediates.10,20,153-155 Methods for the synthesis of arylboronic acids and esters for use in Suzuki couplings are fundamentally the same as for Stille reagents. Aryl boron derivatives are classically synthesized from Grignard or lithium reagents and trialkyl borates; acid hydrolysis then gives the arylboronic acid (Scheme 32), or where feasible the ester may be coupled directly.156 The organometalloid approach to the synthesis of organoboron compounds has been thoroughly reviewed.20,153,155 Unlike the tin electrophile R3Sn?Cl, which bears only one leaving group, the typical methyl, ethyl or butyl borate electrophile B(OR)3 can undergo bis- or tris-alkylation leading to the formation of borinic acid (i.e., Ar2B(OMe)) or trialkylborane (Ar3B) derivatives, respectively. Triisopropyl borate has been shown to sterically temper multialkylation of the borate, and it has become the electrophile of choice with highly reactive organometalloids.157 Scheme 32 1. n-BuLi, Et2O, -78 ?C 2. B(OBu)3 3. dil. HCl N Br N B(OH)2 79% Unlike the traditional organometallic approach, Pd(0)-catalyzed borylation strategies enable access to boronic acid and ester derivatives in the presence of electrophilic functionalities such as nitro, ester, ketone and cyano groups. Aryl halides and pseudohalides undergo borylation under palladium catalysis using two different types of boron nucleophiles: alkoxydiboron derivatives, (RO)2B?B(OR)2 or alkoxy boranes, H?B(OR)2 (Scheme 33). The Pd(0)-catalyzed cross-coupling approach to the synthesis of organoboron compounds has been reviewed.158-160 27 Scheme 33 Pd(0) B-nucleophile Ar X Ar B(OR)2 Miyaura has pioneered the use of bis(pinacolato)diborane for the borylation of aryl iodides,161,162 bromides,161,162 chlorides163,164 and triflates165 (Scheme 34). The reaction tolerates various functional groups, however ortho-substituents or electron-donating substituents slow down the reaction significantly, necessitating the use of specialized catalysts and a higher catalyst loading to achieve reasonable reaction times.163 The major reaction byproducts are the homocoupled starting material (believed to arise from the Suzuki reaction of the arylboronate product and unreacted aryl halide starting material), and reduced starting material (attributed to hydrolytic photodeboronation,166 a particular problem with boronic acids and esters adjacent to a heteroatom).163 Scheme 34 OMe X PdCl2(dppf), KOAc solvent, 80 ?C, 2-24 h 82%, X = I, DMSO 68%, X = Br, DMSO 93%, X = OTf, dioxane OMe B B O OO O B OO Pd(dba), PCy3, KOAc dioxane, 80 ?C, 6 h B B O OO O OMe Cl 92% Strongin has extended the scope of the methodology to aryldiazonium salts: the Pd-catalyzed cross-coupling of bis(pinacolato)diborane and most para-substituted aryldiazonium salts proceeds efficiently in moderate to excellent yields (42?96%)(Scheme 35).167 This reaction is notable in that the tetrafluoroborate salts are prepared from relatively inexpensive, readily available anilines. Also, this methodology features comparatively mild conditions, an environmentally-friendly alcohol solvent, and no added base; homocoupled contaminants were not observed (and no other byproducts were reported).167 28 Scheme 35 N2BF4 PdCl2(dppf) MeOH, 25-40 ?C, 8 h B B O OO O B OO 81%O OMe O OMe Masuda discovered that the coupling reaction of pinacolboron hydride with aryl iodides, bromides, or triflates in the presence of a catalytic amount of PdCl2(dppf) or PdCl2(PPh3), together with triethylamine afforded aryl boronates in good to excellent yields (Scheme 36).168,169 Since the initial report by Masuda, the reaction has not been greatly improved upon, although several variations of the reaction have appeared in the literature:170-173 Baudoin demonstrated that the reaction is catalyzed by Pd(OAc)2 and Buchwald?s phosphine (PCy2(o-biphenyl));170 and LeFloch employed a novel phosphinine-Pd complex as an alternative catalyst.171 Scheme 36 AcHN OTf 77% PdCl2(dppf), Et3N dioxane, 100 ?C, 4 hB H O O B O O AcHN+ The cross-coupling reaction employing pinacolboron hydride tolerates a variety of electron-rich and -deficient functional groups and is relatively insensitive to solvent (dioxane, toluene, dichloroethane, and acetonitrile were all equally accommodated).168-171 Ortho, meta, and para substituted substrates undergo equally efficient conversion to the boronate. As with the cross-coupling reactions using alkyl diboranes, aryl iodides are much more reactive than bromides or triflates, which require extended reaction times and elevated temperatures to reach completion. The major byproduct is the reduced substrate 29 Ar?H; when triethylamine is substituted by H?nig?s base, pyridine, KOAc, or DBU, the reduced species predominates. Less common methods for the synthesis of aryl boron derivatives include the rhodium or iridium-catalyzed borylation of arenes via C?H bond activation174 and the regioselective D?tz annulation of Fischer carbene complexes with alkynyl boronates.175 Both of these techniques are noteworthy in that each provides arylboronic acids without relying on the availability of the appropriate aryl halide. The former technique employs either pinacolboron hydride or bis(pinacolato)diborane, and was reviewed by Ishiyama in 2003.174 An example of this method is shown in Scheme 37.176 C?H bond activation suffers from the requirement for harsh reaction conditions, polyborylation, and often poor regioselectivity with mono or unsubstituted substrates; only disubstituted arenes and substituted heteroaromatics exhibit regiocontrol. Scheme 37 [IrCl(COD)]2+dibpy (3 mol%) B B O OO O N (i-Pr)3Si N (i-Pr)3Si B O O octane / 80 ?C / 16 h + H2+ N (i-Pr)3Si B O O B O O 79% 10% The benzannulation technique has some unexpected strengths: both the Fischer complexes and alkynyl boronates are readily accessible; and alkynyl boronates exhibit high regiospecificity in the D?tz annulation reaction, in stark contrast to the corresponding alkynylsilanes and -stannanes.175 The reaction tolerates alkyl and phenyl substituents on the alkynyl boronate substrate; however the presence of bulky alkynyl substituents leads to the exclusive formation of cyclobutenone products. Harrity developed and applied this technique to the synthesis of a novel class of quinone and hydroquinone boronic acids (Scheme 38).175 Aryl complex 67 and alkyne 68 underwent smooth 30 cycloaddition to provide arylboronate 69; regioisomer 71 was not detected. The remaining yield was of quinone 70, formed by the photodeborylation of the arylboronate product 69. Scheme 38 Cr(CO)5 OMe Ph B O O O O Ph B O O O O Ph B O O O O Ph 1. THF, 45 ?C 2. Ce(IV) 67 68 70 12% 69 57% 71 + + Preparation of Aryl Silicon Reagents Previous studies from our group21-25 as well as Hiyama,8,26-33 Ito,34,35 Denmark,36-44 and others45-56 have shown that a variety of silicon derivatives (i.e., compounds 72-78, Figure 7) undergo fluoride-mediated, Pd(0)-catalyzed aryl group transfer reactions.57-59 The Pd- catalyzed cross-coupling reaction of organosilicon compounds has largely been studied by Hiyama and Hatanaka (aryl(alkyl)dihalosilanes 72-73),8,26-33 and Shibata116 and DeShong21-25 (aryl(trialkoxy)silanes 74): all results to date indicate that these organosilicon derivatives participate in the cross-coupling reaction with similar efficiency and functional group tolerance as aryl stannanes and boranes.139 While initial reports demonstrate that Denmark?s36-44 arylsilacyclobutanes (75), and Mori?s28,31,52,56 arylsilanols (76-78) will participate in the aryl-aryl cross-coupling process, the scope and utility of these reactions have yet to be demonstrated. In addition, the potential application of the silanol 31 methodology is limited by the need for stoichiometric amounts of silver(I) oxide as an activator.139 Silacyclobutanes and silanols will not be discussed further. SiRX2 OCH3 Si(OH)2Et Me Si Cl MeO Si(OEt)3 O2N 72, X = F, R = Et, Me, i-Pr 73, X = Cl, R = Et 77 7574 Si(OH)(CH3)2 Me 76 Si(OH)3 78 Figure 8. Silicon-Based Aryl Group Transfer Reagents. Aryl(trialkoxy) and aryl(alkyl)dihalo silicon reagents (Ar?Si(OR)3 and Ar?Si(R)X2, respectively) possess many of the strengths, but few of the limitations associated with their tin and boron counterparts (vide infra).8,21-59 Cross-coupling prowess aside, it is the ease of working with these silicon reagents that has led to the flurry of interest in silicon- based aryl-group transfer reagents. Organosilicon derivatives?like boronic acids?are non-toxic relative to the tin compounds.26 Organosilanes are readily available, stable to air and moisture, and?unlike many boronic acids and esters?can undergo purification by distillation or column chromatography, and are stable to most reaction conditions employed in synthetic chemistry.26 Lastly, organosilicon compounds hold the promise of being more easily prepared than the notoriously problematic organoboron derivatives. Hiyama and Hatanaka have pioneered the use of aryl(alkyl)dihalosilanes in the Pd-catalyzed cross-coupling reaction with aryl halides (Scheme 39).26,177,178 A range of silanes (Ar?SiR3-nXn) has been shown to participate in the cross-coupling, among which the optimal reagents are the aryl(alkyl)difluoro-1,8,177,179,180 and aryl(alkyl)dichlorosilanes.29,30 The non-transferable ?dummy? alkyl ligand on silicon is typically methyl, ethyl or propyl, 32 although under certain conditions the methyl group is also capable of being transferred in the cross-coupling reaction, in competition with the aryl group.26,177,178 Scheme 39 Me DMF 100 ?C 67% + NC I NC Si(Et)F2 Me (?3-C3H5PdCl)2 KF The organofluorosilane derivatives are synthesized from the corresponding organochlorosilanes in moderate yield (50-60%), using antimony trifluoride (SbF3) (Swarts reaction)1,181-185 or CuF2 (Scheme 40).1,9 This conversion has also been reported to occur in moderate yield with HF, although few applications of this technique to the synthesis of aryl(alkyl)difluorosilanes have appeared in the literature.181,186-188 Miscellaneous other methods for the formation of organofluorosilane derivatives via the cleavage of Si?X bonds other than chloride have been reported (i.e., X = OH, OR, H), but are of limited scope.119,185,188-190 Hiyama and Hatanaka have since developed a cross-coupling protocol employing aryl(alkyl)dichlorosilanes directly, thus avoiding an additional synthetic step, and the handling of the toxic and moisture-sensitive SbF3 and CuF2.29,30 Scheme 40 H3C Si(Et)Cl2 1. SbF 3, -70 ?C ? 25 ?C, neat 2. 100 ?C, 2h 65% H3C Si(Et)F2 In parallel to the preparation of aryl tin and boron reagents, aryl(alkyl)dichlorosilane derivatives are traditionally synthesized from the corresponding aryl Grignard or lithium reagent and trichloro(alkyl)silane electrophiles (Scheme 41).1,29,33,181,183,187,190-195 No systematic study of the synthesis of aryl(alkyl)dichlorosilanes via the organometallic method has been reported, and little data exists in the literature: of the few reports stating 33 yields of purified product,33,193 a range of 33-88% (average yield of 50-60%) is given, and in no cases are the byproducts of the metallation reaction discussed or characterized. Hiyama and Hatanaka have demonstrated that crude aryl(alkyl)dichlorosilane derivatives prepared via metallation may be used without further purification for Pd-catalyzed cross- coupling, and report quantitative yields of crude silane via the metallation reaction.1,181 Scheme 41 1. Mgo, THF, 0 ?C 2. EtSiCl3 88% Br MeO Si(Et)Cl2 MeO Ortho substitution of the aryl Grignard or lithium species strongly impacts the yield of the metallation reaction (Scheme 42).33,193 Hiyama demonstrated that silylation of p- and m-bromotoluene with Et?SiCl3 occurred in good yield relative to o-bromotoluene.33 The more sterically compact alkyltrifluorosilane electrophiles (R?SiF3) have been shown to react with very hindered organometallic reagents where R?SiCl3 failed to react;191 however this approach has never been applied to the systematic synthesis of ortho-substituted aryl(alkyl)difluorosilanes. Scheme 42 1. Mgo, THF, 0 ?C 2. Et-SiCl3Me Br Me Si(Et)Cl2 o-Me, 35% m-Me, 72% p-Me, 62% A second, less common method for the synthesis of aryl(alkyl)dichlorosilanes is silylative decarbonylation of aryl acid chlorides, as reported by Rich.196 Again, this Pd- catalyzed cross-coupling technique is a good foil to the metallation reaction where electrophilic organic functional groups prohibit the organometalloid method. The palladium- catalyzed reaction of sym-tetrachlorodimethyldisilane(Cl2(Me)Si?Si(Me)Cl2) with benzoyl chloride derivatives was shown to form aromatic chlorosilanes in generally good yield (Scheme 43). The best yields were obtained with electron-deficient, p- and m-substituted 34 aromatic acid chlorides. The reaction failed for 2,6-pyridine dicarboxylic acid chloride (no reaction was observed), apparently due to complexation of the pyridine to the palladium catalyst. With few exceptions, the corresponding cross-coupling reaction of aryl halide substrates and Cl2(Me)Si?Si(Me)Cl2 fails.45,196-198 Scheme 43 O Cl + Cl2Si SiCl2 MeMe (PhCN)2PdCl2 PPh3 Si(Me)Cl2 + CO MeSiCl3+ 69%NO2 NO2 neat, 145 ?C, 13h Miscellaneous other preparations of aryl(alkyl)dihalo silane derivatives have been reported, but are of limited scope or are only practical in the industrial laboratory.184,199-203 An efficient synthesis of ortho-substituted aryl(alkyl)dihalo silane derivatives has not been described. In addition to functioning as aryl group transfer reagents, aryl(trialkoxy)silane derivatives (Ar-Si(OR)3) are important synthetic intermediates for the formation of hybrid organic-inorganic materials, such as synthetic glasses, ceramics, coatings or fibers prepared by sol-gel chemistry.118,204 These compounds are often called alkoxysilane ?monomers,? in reference to the polymerization of organosilanes into water-soluble oligomers.204 Despite the multidisciplinary utility of arylsiloxanes, few practical methods for the preparation of trialkoxysilyl or trihalogenosilyl derivatives have been reported. In analogy to the corresponding preparations of Stille and Suzuki reagents, the traditional method for the synthesis of aryl siloxanes is the treatment of an aryl Grignard or lithium reagent with a silicon electrophile (SiCl4, Cl-Si(OR)3, or Si(OR)4) (Scheme 44). Scheme 44 OCH3 M OCH3 Si(OR)3 M = MgBr, Li Si(OR)4 Cl-Si(OR)3 1. SiCl4 2. ROH 35 Aryl metals will react with silicon atoms bearing leaving groups other than chloride and alkoxy, including ?CN, ?H, ?OSiR3, ?NR2, and ?OAc, however these reactions are of limited practicality: not only are the electrophiles tedious to prepare and purify, but in the case of H?Si(OR)3, the pyrophoric byproduct silane (SiH4) is produced upon reaction with strong bases such as organometallic reagents.204,205 Other organometallic reagents generated by metal-halogen exchange have been employed in this transformation, as well, however they are of limited scope.121,204 The first organometalloid method for the synthesis of arylsiloxane derivatives involves a two-step process: (1) the silylation of an arylmetalloid with commercially- available, inexpensive silicon tetrachloride (SiCl4) to form the aryl(trichloro)silane intermediate; and then (2) alcoholysis (Scheme 45).116,206-214 Shibata synthesized a series of aryl(triethoxy)silanes in ?acceptable? yields using Grignard or lithium reagents (the choice of metal was dictated by the ease of forming the carbanion).116 Scheme 45 Pr SiCl3 MeOH pyridine 65% Pr Si(OMe)3 Pr Br 1. Mgo 2. SiCl4, 0-25 ?C, 17h Pr I 1. BuLi, hexane, -70 ?C 2. SiCl4, -70 to 25 ?C, 17h The use of SiCl4 as the silicon electrophile is of limited practicality for a number of reasons which include (a) moderate to poor product yield for the overall, two step conversion, (b) the difficult transfer and handling of moisture-sensitive reagents and intermediates, and (c) multiple substitution results in the formation of difficult to separate byproducts (Scheme 46).206,210 Large quantities of electrophile (20+-fold excess) have been employed to mitigate polyarylation, however this leads to difficulties in product isolation and HCl contamination.210 36 Scheme 46 SiCl4, 1.2 equiv heptane/THF reflux 2h MgCl SiCl3 + Cl2 Si 48% 17% Treatment of the arylmetalloid with Cl?Si(OR)3 should favor the formation of monoaryl siloxanes by preferential displacement of chloride from silicon.118,210,215,216 Not only are the yields somewhat better than the SiCl4/alcoholysis technique, but the reaction involves a single step from the formation of the organometalloid (i.e., Scheme 47).216 However, use of Cl?Si(OR)3 as the silicon electrophile still has several significant drawbacks: (1) unlike SiCl4, Cl?Si(OR)3 compounds are not commercially available, and are tedious to prepare;121,217,218 (2) while less reactive than the tetrachloride, chloro(trialkyl)silane reagents still require careful handling to avoid hydrolysis of the reactive Si?Cl bond; and (3) surprisingly, even when excess electrophile is used, or low temperatures are employed, over-arylation of silicon occurs, resulting in inseparable di- and tri-arylated byproducts.204,210 Scheme 47 Cl-Si(OMe)3, 1.0 equiv THF reflux, 1h MgBr Si(OMe)3 60% Br Br Treatment of the arylmetalloid with the less reactive electrophile Si(OEt)4 should favor the formation of monoaryl siloxanes because alkoxide is more difficult to displace from silicon than chloride; typically, the reaction of organometallic reagents with alkoxy silanes is more selective than the reaction with chlorosilanes.204 As with the arylation of Cl?Si(OR)3, the typical yields of the silylation of organometallic reagents with Si(OEt)4 are superior than the SiCl4/alcoholysis method, and this technique is more efficient because it involves only a single step from the formation of the organometalloid.116-120,209,214,219,220 37 For example in Scheme 48, conversion of the series of bromotoluene regioisomers to the Grignard reagent, followed by treatment with 2 equiv of tetraethoxysilane afforded acceptable yields of the corresponding ortho- and para-(triethoxysilyl)toluene isomers; the low yield of m-(triethoxysilyl)toluene was attributed to the lowered nucleophilicity of the carbanion (electronic effect) .120 Scheme 48 Si(OEt)4, 2.0 equiv THF reflux, 16 h o-, 68% m-, 29% p-, 48% MgBr CH3 Si(OEt)3 CH3 The use of Si(OR)4 for the one-step preparation of arylsiloxanes via the organometallic approach has several advantages including the commercial availability, low cost, and ease of handling and storage of tetraalkoxysilanes.204 Polyarylation still occurs, however, lowering the yield of monoarylated silane, and necessitating careful fractional distillation or chromatography to isolate the desired product. For example, Frohn was able to make perfluorinated siloxane 80 in 55% yield from the Grignard reagent and Si(OEt)4, where the modest yield was due to the formation of di- and triarylated products 81 and 82, respectively (Scheme 49).220 Scheme 49 1. Mgo 2. Si(OEt)4, 2.0 equiv Br F F F F F Si(OEt)3 F F F F F + (C6F5)2Si(OEt)2 (C6F5)3Si(OEt)+ 80 55% 81 35% 82 trace 79 Masuda has reported the Pd(0)-catalyzed silylation of aryl iodides using triethoxysilane (H?Si(OEt)3) (Scheme 50).58,221 The reaction was achieved with 38 triethoxysilane and Pd2(dba)3?CHCl3 with the addition of 2 equiv of tris(o-tolyl)phosphine and a base such as i-Pr2NEt. Excellent yields of aryl siloxane were obtained with electron-rich, para-substituted aryl iodides. In contrast to aryl iodides, aryl bromides were reported to provide low yields of aryl siloxane under identical conditions. In addition, the silylation of ortho- or meta-substituted aryl iodides/bromides was not reported. Scheme 50 H3CO Si(OEt)3 H3CO I H3CO H + 83 84 85 H-Si(OEt)3 i-Pr2NEt / NMP Pd(dba)2?CHCl3 P(o-tol)3 The major byproduct was the reduced substrate Ar?H (85); in contrast to the polyarylated silane byproducts formed using the organometallic technique, the reduced byproduct is readily separated by distillation or column chromatography. When H?nig?s base was substituted by triethylamine, pyridine, KOAc, or left out of the reaction completely, the reduced species predominated.19 Electron-withdrawing groups not only slowed the reaction significantly, necessitating higher reaction temperatures and longer reaction times, but facilitated the transfer of hydride from the silane to give the reduced arene as the major product.19,159 With few exceptions, the corresponding cross-coupling reaction of aryl halide substrates and Cl3Si?SiCl3 fails,45,196,197 and silylation employing hexaalkoxydisilanes ((RO)3Si?Si(OR)3) has not been reported.198 The Masuda method is an interesting alternative to the organometalloid approach in that the yields of the cross-coupling reaction are far superior than the metallation reaction with para-substituted electron-rich aryl halide substrates.221 In addition, where the cross- coupling technique fails with ortho-substituted arenes, the metallation method works optimally. The Masuda technique is limited to aryl iodides, unlike the metallation method, which typically employs the more available aryl bromides (and sometimes chlorides). Unfortunately, neither the organometallic approach nor the silylation technique is effective for the silylation of arenes bearing electron-withdrawing, electrophilic moieties. 39 In conclusion, existing methods for the synthesis of aryl siloxanes fall into two categories (Scheme 51): (1) treatment of an aryl Grignard or lithium reagent with a silicon electrophile ; and (2) silylation of an aryl iodide by triethoxysilane (H?Si(OEt)3) in the presence of a palladium catalyst as reported by Masuda.221 Scheme 51 H-Si(OEt)3 Pd(0), PR3 Masuda OCH3 Si(OEt)3 OCH3 I OCH3 M M = MgBr, Li OCH3 X X = I, Br Si(OEt)4 Metallation Both methods, as presently constituted, have their limitations. The metalloid reaction is limited by the generally inferior yields of monoarylsiloxane, as well as the problems associated with the synthesis of aryl metalloids having electrophilic functional groups (i.e. esters, ketones, etc.).116-122 A general method for the synthesis of arylsiloxanes from aryl lithium or magnesium reagents has not been reported, nor has a systematic study of this reaction been performed. Our first goal was to evaluate the metallation method, and to develop a practical, general synthetic technique utilizing standard organometallic reagents for the synthesis of aryl(trialkoxy)silanes. Ideally our method would employ the inexpensive, commercially available tetraethoxysilane, or its methoxy analog. The Pd-mediated silylation protocol of Masuda is limited to electron-rich iodo arenes.221 Masuda reported that electron-rich, para-substituted aryl iodides gave excellent yields of aryl siloxane. In contrast to aryl iodides, aryl bromides were reported to provide low yields of aryl siloxane under identical conditions. In addition, the silylation of ortho- or meta-substituted aryl iodides/bromides was not reported. Accordingly, the goals of this study were to extend the Masuda silylation reaction to aryl bromides and chlorides, and to evaluate the Masuda methodology for the synthesis of ortho- and meta-substituted aryl siloxanes. 40 Results and Discussion Improved Synthesis of Aryl(trialkoxy)silanes via Treatment of Aryl Grignard or Lithium Reagents with Tetraalkoxysilane Numerous methods for the formation of the aryl C?Si bond appear in the literature, the vast majority of which describe the trimethylsilylation of organic molecules.204,222,223 In contrast, few systematic studies of the synthesis of aryl(trialkoxy)silanes have been described, partly because the synthetic usefulness and interesting material properties of aryl(trialkoxy)silanes have only recently been discovered (vide supra).204 Our laboratory has focused specifically on aryl(trialkoxy)silanes as surrogates for Stille or Suzuki reagents (Scheme 52). Although initially described in a single paper by Shibata in 1997,116 the Pd-catalyzed fluoride-activated phenylation by phenyl(trialkoxy)silane derivatives has been independently cultivated into a practical synthetic technique by the DeShong group.21,23-25 The scope of the cross-coupling reaction now encompasses phenylation of a range of electron-rich and electron-deficient aryl halides and triflates, including sterically hindered arenes and heteroaromatic substrates.21,23-25 Having demonstrated the efficient phenylation, we wished to study the transfer of substituted aryl groups to ascertain the scope and limitations of the methodology. This study necessitated the synthesis of ortho-, meta-, and para-substituted, electron-rich and electron-deficient aryl siloxanes. Scheme 52 X H3C SnR3/B(OH)2 H3C Si(OR)3 Pd(0) Pd(0), F-X = I, Br, Cl 41 The most common approach to the formation of siloxanes is silylation of an aryl lithium or magnesium (Grignard) reagent (Scheme 53); Table 1 summarizes the literature to date. Although definitive conclusions cannot be drawn from this data, some generalizations can be made including (a) the yields are typically poor to modest, most likely due to competing formation of di- and tri- arylated products (over-arylation) (entries 1 and 26); (b) the choice of metal (Li or Mg) is dictated by the ease of forming the carbanion, although aryllithium reagents are much more reactive than the corresponding arylmagnesium reagents, requiring lower temperatures and shorter reaction times; (c) although SiCl4, Cl?Si(OR)3 and Si(OR)4 all perform similarly in the silylation reaction, the modest yield of the alcoholysis reaction of aryl(trichloro)silane derivatives limits the utility of SiCl4 (entry 30); (d) heteroaromatic systems are particularly problematic (entries10-12 and 31) and (e) the reaction tolerates a variety of aryl halides (o-, m-, and p-substituted, electron-rich as well as deficient arenes undergo silylation). Scheme 53 OCH3 M OCH3 Si(OR)3 M = MgBr, Li Si(OR)4 Cl-Si(OR)3 1. SiCl4 2. ROH Toward our goal of synthesizing a variety of highly functionalized siloxane derivatives for use in aryl coupling reactions, it was necessary to determine the scope and limitations of the synthesis of aryl siloxanes from standard organometallic reagents. For our studies, 4-bromomagnesiumanisole 87, Table 2) was selected as the initial nucleophile to be investigated in the metallation methodology. The inexpensive, commercially available electrophiles tetramethoxysilane (Si(OMe)4), tetraethoxysilane (Si(OEt)4) and tetrachlorosilane (SiCl4) were each evaluated for their efficacy in the formation of the Si?Ar bond. The results are summarized in Table 2. 42 Ar X metallationconditions Ar M Ar Si(OR)3silylationconditions SiL4 Conditions Product Substrate Metallation Silylation Yielda Entry Ar-Si(OR)3 Ar-X Ar-M Ar-M SiL4 (equiv) % 1224 Si(OMe) 3 Br Ar-MgBrb Mg?/I2 Et2O/THF reflux Si(OMe)4 (1.5) Et2O/THF 25 ?C 35c 2225 Si(OEt) 3MeO Br Ar-MgBrb Mg?/I2 THF 65 ?C/2 h Si(OEt)4 (4.0) THF 65 ?C/12 h 74 3225 Si(OEt) 3F I Ar-MgIb Mg?/I2 THF 65 ?C/2 h Si(OEt)4 (4.0) THF 65 ?C/12 h 68 4118 Si(OEt)3 (EtO)3Si 1-Br 4-Br (1,4)-bis MgBrb Mg?/I2 THF 65 ?C/2 h Si(OEt)4 (5.0) THF 65 ?C/1 h 55 5118 Si(OEt)3 (EtO)3Si 1-Br 4-Br (1,4)-bis Ar-Li t-BuLi Et2O -78 ?C/4 h Cl-Si(OEt)3 (2.5) Et2O -78 to 25 ?C 32 6118 Si(OEt)3 (EtO)3Si 2 4-Br 4?-Br (4,4?)-bis MgBrb Mg? THF 65 ?C/2 h Si(OEt)4 (5.0) THF 65 ?C/5 d 34 7118 Si(OEt)3 (EtO)3Si 3 4-Br 4?-Br (4,4?)-bis Ar-Li t-BuLi Et2O -78 ?C/4 h Cl-Si(OEt)3 (2.5) Et2O -78 to 25 ?C 24 8118 Si(OEt)3 Si(OEt)3 4-Br 4?-Br (4,4?)-bis Ar-Li t-BuLi Et2O -78 ?C/4 h Cl-Si(OEt)3 (2.5) Et2O -78 to 25 ?C 24 9226 Si(OEt) 3 N N Ph I Ar-Li BuLi THF -105?C Cl-Si(OEt)3 (1.2) THF 0?C/12 h 75 10227 N Si(OEt) 3 Br ArLi BuLi Et2O -78?C/45min Si(OEt)4 (2.0) Et2O -78 to 0?C 10 43 Ar X metallationconditions Ar M Ar Si(OR)3silylationconditions SiL4 Conditions Product Substrate Metallation Silylation Yielda Entry Ar-Si(OR)3 Ar-X Ar-M Ar-M SiL4 (equiv) % 11227 N Si(OEt) 3 Br ArLi BuLi Et2O -78?C/1h Si(OEt)4 (2.0) Et2O -78 to 0?C 16 12227 N Si(OEt)3 Br ArLi BuLi Et2O -78?C/15min Si(OEt)4 (2.0) Et2O -78 to 0?C 21 13119 NMe2 Si(OEt)3 Hd ArLi BuLi Et2O 25?C/60 h Si(OEt)4 (1.3) Et2O 25?C/12h 56 14219 Si(OMe) 3 Br ArMgBr Mg? Et2O 25?C/4h Si(OMe)4 (1.0) Et2O 25?C/ 3 h 43 15216 Si(OMe) 3Br Br ArMgBr Mg? THF Cl-Si(OMe)3 (1.0) THF 25?C/12h 65?C/1h 50 16120 Si(OEt)3 Br ArMgBr Mg? Et2O/THF reflux Si(OEt)4 (2.0) Et2O/THF reflux/16 h 69 17120 Si(OEt)3 Br ArMgBr Mg? Et2O/THF reflux Si(OEt)4 (2.0) Et2O/THF reflux/36 h 29 18120 Si(OEt) 3 Br ArMgBr Mg? Et2O/THF reflux Si(OEt)4 (2.0) Et2O/THF reflux/36 h 48 19208 SiCl 3Me2N Br ArMgBr Mg? THF 40?C SiCl4 (2.0) THF/ heptane 40?C 40 20208 SiCl 3MeO Br ArMgBr Mg? THF 40?C SiCl4 (2.0) THF/ heptane 40?C 32 44 Ar X metallationconditions Ar M Ar Si(OR)3silylationconditions SiL4 Conditions Product Substrate Metallation Silylation Yielda Entry Ar-Si(OR)3 Ar-X Ar-M Ar-M SiL4 (equiv) % 21208 SiCl 3Cl Br ArMgBr Mg? THF 40?C SiCl4 (2.0) THF/ heptane 40?C 39 22214 SiCl3 Me N NMe Hd ArLi BuLi Et2O -30 to 0?C SiCl4 (1.3) Et2O 25?C/12h 91 23214 (MeO)3Si Me N NMe Hd ArLi BuLi Et2O 25?C/60 h Si(OMe)4 (1.3) Et2O 25?C/12h 17 24214 (MeO)3Si Me N NMeCl Hd ArLi BuLi Et2O 25?C/6 d Si(OMe)4 (1.3) Et2O 25?C/12h 17 25207 SiCl 3 Br MgBr Mg? Et2O SiCl4 (1.5) Et2O/ benzene 25?C/12h 59 26206 SiCl 3 Cl MgCl Mg? THF SiCl4/(1.1) THF/ heptane reflux/2h 47e 27116 Si(OMe) 3R R = 4-(Pr)-cyh- Br Ar-MgBr Mg? THF 50?C/2h 1. SiCl4 (1.5) THF 25?C/17h 2. MeOH pyr/0?C 64f 28116 Si(OMe) 3R R = 4-(Pr)-cyh- Br Ar-MgBr Mg? THF 50?C/2h Si(OMe)4 (1.5) THF 25?C/17h 35 45 Ar X metallationconditions Ar M Ar Si(OR)3silylationconditions SiL4 Conditions Product Substrate Metallation Silylation Yielda Entry Ar-Si(OR)3 Ar-X Ar-M Ar-M SiL4 (equiv) % 29116 Si(OMe) 3R R = 4-(Pr)-cyh- I Ar-Li BuLi THF -70?C/2h 1. SiCl4 (2.0) THF 25?C/17h 2. MeOH pyr/0?C 66f 30213 NMeNMe2 Si(OEt)3 Hd Ar-Li BuLi/TMEDA Hexane 25?C/2d 1. SiCl4 (1.1) hexane 25?C/12h 2. NaOEt THF/2d 6f (22)h 31210 N Si(OEt)3 Br Ar-Li BuLi Et2O -78?C/30min 1. SiCl4 (2.0) Et2O 25?C/3h 2. EtOH/TEA 15f a Yields are after distillation. Byproducts were not reported. Unless otherwise indicated, the organometalloid (Ar-MgBr or Ar-Li) was first generated, and then combined with the indicated silicon electrophile SiL4 (Si(OR)4, Cl-Si(OEt)3, or SiCl4). b Grignard generated under Barbier conditions: Mg?, I2 (if used), the silicon electrophile SiL4 (Si(OR)4, Cl-Si(OEt)3, or SiCl4) and solvent were combined and brought to the indicated temperature with stirring; the arylbromide was added at such a rate as to maintain reflux, and then the reaction was stirred at the given temperature until the reaction was complete. c A 41% yield of the diaryl(dialkoxy)silane and trace amounts of the triaryl(alkoxy) silane were isolated. d Aryl lithium generated by directed ortho metallation. e A 17% yield of the diaryl(dialkoxy)silane and a 10% yield of biphenyl were isolated. f The overall yield of Ar-Si(OR)3 after alcoholysis of Ar-SiCl3, based on the amount of substrate (Ar-X or Ar-H). g The major product was the diaryl(dialkoxy)silane. h The yield in parentheses indicates the yield of Ar-SiCl3. Table 1. Synthesis of Aryl(trialkoxy)silanes via Treatment of Aryl Grignard or Lithium Reagents with Silicon Electrophiles SiL4: Review of the Literature to Date. 46 MgBr Si(OR)3 Si(OR)4, THF Br OCH3 OCH3 OCH3 86 88 + 89 Mg?, I2 (cat) 87 1. SiCl4, THF 2. EtOH, pyr, 0 ?C ORTHFreflux Ar2Si(OR)2 Conditionsa Yield (%)b,c Entry Silane / Equiv T (?C) 88 89 1 Si(OMe)4 / 3.0 25 47 47 2 Si(OMe)4 / 3.0 -30 76 3 3 Si(OMe)4 / 1.5 -30 65 4 4 Si(OEt)4 / 3.0 0 72 23 5 Si(OEt)4 / 3.0 -10 70 23 6 Si(OEt)4 / 3.0 -30 83 3 7 Si(OEt)4 / 1.5 -30 76 2 8 Si(Cl)4 / 3.0d -30 70d 3 a Reactions of p-methoxy magnesium bromide 87 (1.0 equiv) with Si(OR) 4 orSiCl 4 (1.5 to 3.0 equiv) were allowed to stir at the given temperature in THF.The reaction mixture was stirred at the indicated temperature for 1 h, and then at room temperature for 12 h. b GC yields are based on amount of 4-bromoanisole (86). c The remainder of the product was of anisole. d Yield of 4-(triethoxysilyl)anisole (88). The crude, concentrated 4-(trichlorosilyl)anisole was converted to the siloxane 88 by dropwise addition of the chlorosilane to EtOH/pyridine at 0 ?C. Table 2. Optimization of the Synthesis of Arylsiloxanes Using 4-Bromomagnesium Anisole. The temperature at which the Grignard reagent was added to the electrophile had a marked effect on the reaction. As seen in entries 1, 4 and 5, at temperatures above ?30 ?C a significant amount of the diaryl(dialkoxy)silane impurity 89 was formed, although no triaryl(alkoxy)silane was observed over the tested temperature range. The remainder of the reaction mixture was anisole, which can result from the aqueous quench of unreacted arylmagnesium bromide. However, longer reaction time or elevated temperature did not improve the yield. Also, the amount of anisole remained constant irrespective of the electrophile. It is likely that reduction occurs at least in part during the in situ formation 47 of the Grignard reagent, and is an inherent limitation of the metallation approach to the synthesis of siloxanes. The best result was obtained when the Grignard reagent was treated at ?30 ?C with 3.0 equiv of the electrophile Si(OEt)4 (83% yield, entry 6); under identical reaction conditions, Si(OMe)4 (76% yield, entry 2) and SiCl4 (followed by alcoholysis) (70%, entry 8) also performed well (the yields were determined by G.C. analysis). In contrast, Shibata observed a 60% yield of siloxane using the two-step SiCl4/alcoholysis method for the silylation of p-(4-propylcyclohexyl)phenylmagnesium bromide (Scheme 54, also above in Table 2, entries 27 and 28), but only a 35% yield of siloxane via direct silylation with Si(OMe)4. The reaction conditions employed by Shibata were addition of the Grignard reagent to 1.5 equiv of silane at 0 ?C.116 In our study, a two-fold decrease in the concentration of the electrophile slightly decreased the yield of the desired monoarylated silane, but surprisingly did not stimulate the formation of diarylated byproduct (Table 3, entries 3 and 7). Scheme 54 Pr Si(OMe)3 Pr MgBr 1. SiCl4 (1.5 equiv) THF, 0-25 ?C, 17h 64% THF 0-25 ?C, 17h 35% Si(OMe)4 (1.5 equiv) 2. MeOH, pyridine Preliminary studies performed by Arash Soheili (an undergraduate) using commercial phenylmagnesium bromide indicated that increasing the concentration of the electrophile from 3 to 5 to 30 equiv did not improve the reaction outcome, and only served to complicate product isolation. Soheili also observed little reaction improvement with alternative ether solvents, and poorer results with toluene or hexane. Ultimately, THF 48 rather than lower-boiling Et2O was employed as the solvent in order to facilitate in situ formation of the Grignard reagent (which required heating to progress efficiently).228 In order to demonstrate the scope of the Grignard reaction, a variety of substituted aryl bromides was examined (Table 3). The isolated yields obtained in our laboratories were generally better than those reported in the literature. Only trace amounts of diaryl(dialkoxy)silane contaminants were observed for all entries. Again, the remaining yield was of reduced (dehalogenated) aryl halide. As expected, tetramethoxysilane and tetraethoxysilane worked equally well (entries 1 and 2, and 6 and 7). MgBr Si(OR')3 R R THF Si(OR')4 Br R THF Mg? 90 9291 Entry Aryl Siloxane Yield (%)a,b Entry Aryl Siloxane Yield (%)a,b 1 Si(OEt)3 OCH3 88 6 Si(OEt)3 CH3 81 2 Si(OMe)3 OCH3 83 7 Si(OMe)3 CH3 71 3 Si(OEt)3 H3CO 84 8 Si(OEt)3 H3C 75 4 H3CO Si(OEt)3 82 9 H3C Si(OEt)3 86 5 O O Si(OEt)3 74 10 Si(OEt)3 55 a Reactions of arylmagnesium bromide 91 (1.0 equiv) with Si(OEt) 4 or Si(OMe)4 (3.0equiv) were allowed to stir at ?30 ?C in THF. The reaction mixture was stirred at ?30 ?C for 1 h, and then at room temperature for 12 h. b Yields of 92 are after distillation (purity >95%). The remainder of the product was of the reduced (dehalogenated) arene. Table 3. Synthesis of Aryl(trialkoxy)silanes Using Grignard Reagents. 49 Surprisingly, the results were comparable with o-, m-, and p-substituted anisyl and tolyl Grignard reagents (entries 2-4, and 7-9, respectively); all yields were approximately 80%. This result contradicts the findings of Selin and co-workers (Scheme 55, also above in Table 2, entries 15-17), who observed acceptable yields of o- and p-, but not m-(triethoxysilyl)toluene; the reaction conditions employed by Selin entailed addition of the Grignard reagent to 2.0 equiv of tetraethoxysilane at reflux.120 Scheme 55 Si(OEt)4, 2.0 equiv THF reflux, 16 h o-, 68% m-, 29% p-, 48% MgBr CH3 Si(OEt)3 CH3 Several researchers have reported acceptable yields of arylsiloxane using Barbier reaction conditions?the in situ formation of the reactive Grignard intermediate in the presence of the silicon electrophile (Scheme 56, and Table 2, entries 1-4, and 6).118,130,224,225 Ideally, Barbier conditions are a method of controlling multiple substitution at silicon, because the Grignard reagent is immediately consumed, and never present in excess. For example, Shea and co-workers combined magnesium metal, a crystal of iodine, 4 equiv of tetraethoxysilane and THF, and brought the mixture to reflux (Scheme 56). A solution of p-bromoanisole (86) in THF was then added slowly, and the mixture was allowed to reflux for 12 h; a 74% isolated yield of 4-(triethoxysilyl)anisole (88) was obtained without mention of byproducts.225 This result was not reproducible in our laboratories: under identical Barbier reaction conditions, only 50% yield at 80% conversion was achieved; although no diarylated silane was observed, the remaining yield was of reduced arene. 50 Scheme 56 Si(OEt)3 Br H3CO H3CO 88 1. Mg?, I2, THF, reflux 2. THF, reflux, 12h 86 74% (50%)Si OEt OEt EtO OEt As shown in Table 3 (entry 5), 3,4-(methylenedioxy)bromobenzene underwent smooth conversion to the siloxane (93, 74% yield). Siloxane 93 was chosen as a model for siloxane 28, which is a synthetic intermediate in the synthesis of Fitzgerald?s compound 23 (Figure 4). Since the completion of this study, Correia in the DeShong laboratories has applied the Grignard methodology described herein toward the preparation of siloxane 94 in 50% yield from the corresponding bromide.229 Si(OEt)3 OMe OMe MeO 28 Si(OEt)3 O O 93 Si(OEt)3 OMe O O 94 Figure 9. Siloxanes Prepared via the Grignard Reaction for Use in Natural Product Syntheses. In continuing pursuit of our goal of synthesizing a variety of highly functionalized siloxane derivatives for use in aryl coupling reactions, we turned our efforts to determining the scope and limitations of the synthesis of aryl siloxanes from lithium reagents. For our studies, 4-lithiotoluene (96, Table 4) was selected as the initial nucleophile for investigation of the formation of the Si?Ar bond. The electrophile tetraethoxysilane (Si(OEt)4) was chosen for initial study, in lieu of tetramethoxysilane (Si(OMe)4), which freezes at the low temperatures required for the formation of aryllithium reagents, or tetrachlorosilane (SiCl4), because the subsequent alcoholysis step had proven to be problematic (low yielding and complicated by polymerization). Both ether and THF were 51 evaluated for their suitability in the silylation of aryl lithium intermediates, although ether was the solvent of choice for practical reasons. The higher-boiling ethereal solvent was preferred for the Grignard protocol because it allowed for heating the reaction to initiate formation of the Grignard, or to drive silylation. The more reactive lithium reagents decompose above low temperatures and do not require higher-boiling solvents. The results of the silylation reactions are shown in Table 4. Li Si(OEt)3 Si(OEt)4 Br CH3 CH3 CH3 95 97 + Ar2Si(OEt)2 + Ar3Si(OEt) 98 99 n-BuLi 96 Conditionsa Yield (%)b,c Entry Equiv Si(OEt)4 Solvent T (?C) 97 98 99 1 1.5 THF 0 7 25 54 2 1.5 THF -30 74 12 1 3 1.5 Et2O -30 77 6 0 4 1.5 THF -78 81 9 0 5 1.5 Et2O -78 82 5 1 6 3.0 Et2O -78 86 3 1 a Reactions of p-tolyl lithium 96 (1.0 equiv) with Si(OEt) 4 (1.5 to 3.0 equiv) wereallowed to stir at the given temperature; the reaction was complete in 1 h. b GC yields are based on amount of 4-bromotoluene (95). c The remainder of the product was toluene. Table 4. Optimization of the Synthesis of Arylsiloxanes Using 4-Lithiotoluene. Like the Grignard approach, the temperature at which the lithium reagent was added to the electrophile dramatically effected the reaction outcome. As seen in entries 1- 3, at temperatures above ?78 ?C a significant amount of the diaryl(dialkoxy)silane impurity 98 was formed. In addition, trialkylated product 99 was observed. Note that the analogous transformation employing Grignard reagents did not lead to the formation of 52 trialkylated products, because of the significantly lower reactivity of Grignard reagents. At 0 ?C (entry 1), the major product was tris-p-tolyl(ethoxy)silane (99). In contrast, at ?30 ?C (entries 2 and 3) and ?78 ?C (entries 4-6) formation of the trialkylated silane 99 was effectively suppressed, although diarylated silane 98 was formed. As in the Grignard reaction, reduction (dehalogenation) of the starting material also was observed. At low temperature, the reaction was essentially insensitive to the solvent (Et2O and THF were tolerated, entries 4 and 5) and only a small excess of the electrophile was needed (1.5-3.0 equiv were equally effective, entries 5 and 6). The best result was obtained when the lithium reagent in ether was treated at ?78 ?C with 3.0 equiv of the electrophile Si(OEt)4 (86% yield, entry 6). Experience garnered applying these reaction conditions to other substrates would show that more than 1.5 equiv were unnecessary to achieve the optimum yield of siloxane. The electrophile and the lithium reagent were allowed to stir at ?78 ?C until the reaction ceased to progress. Visiting Scientist Dr. Chuljin Ahn made the discovery that quenching the reaction at ?78 ?C before allowing the mixture to slowly warm to room temperature mitigated formation of di- and tri- arylated byproducts. In order to demonstrate the scope of the optimized silylation reaction, a variety of substituted aryl bromides was examined (Table 5). The isolated yields obtained were generally better than those reported in the literature. Only trace amounts of diaryl(dialkoxy)silane contaminants were observed for most entries; again, the remaining yield was of reduced (dehalogenated) aryl halide. The best results were obtained with less reactive (less electron-rich) aryl lithium derivatives: the less electron-rich tolyl series (entries 1-3) and electron-neutral phenyl (entry 4) outperformed the more electron-rich anisyl/thioanisyl series (entries 5-8). Bearing two electron-donating substituents, the 3,4-(methylenedioxy)phenyllithium intermediate (entry 9) proved to be highly reactive: the major product was the diarylated silane. Reducing the temperature to ?110 ?C, or increasing the concentration of Si(OEt)4 failed to improve the yield. Note that the analogous Grignard reagent underwent clean silylation (Table 3, entry 5), forming 74% of 53 triethoxy(3,4-methylenedioxyphenyl)silane. Again as with magnesium reagents, the reaction using aryllithium intermediates was insensitive to the position of substituents: o-, m-, and p-substituted arenes underwent silylation with equal efficiency (Table 5, entries 1-3, and 5-8). Li Si(OEt)3 R R Si(OEt)4 Br R Et2O -78 ?C n-Bu Li 90 92100 Et2O -78 ?C Entry Aryl Siloxane (92) % Yielda,b Entry Aryl Siloxane (92) % Yielda,b 1 Si(OEt)3 CH3 79 7 H3CO Si(OEt)3 67 2 Si(OEt)3 H3C 71 8 Si(OEt)3MeS 50 3 H3C Si(OEt)3 85 9 O O Si(OEt)3 30c 4 Si(OEt)3 74 10 S Si(OEt)3 50 5 Si(OEt)3 OCH3 60 11 O Si(OEt)3 22d 6 Si(OEt)3 H3CO 66 12 N Si(OEt)3 0 a Reactions of aryl lithium 100 (1.0 equiv) with Si(OEt) 4 (1.5 equiv) were allowed to stir at?78 ?C for 1 h in Et 2O. b Yields of 92 are after distillation or chromatography (purity >95%). c The major product was Ar2Si(OEt)2. d G.C. analysis indicated a 2:1:1 ratio of mono:di:tri arylalkoxysilanes. Table 5. Synthesis of Aryl(trialkoxy)silanes Using Lithium Reagents. 54 Mixed results were obtained for silylation of heteroaromatic systems (entries 10- 12). For example, an acceptable yield of 3-(triethoxysilyl)thiophene was obtained from the corresponding bromide (entry 10), whereas silylation of 2-bromofuran was inefficient and 2-bromopyridine failed to give any of the desired siloxane (entries 11 and 12). The yield of 2-(triethoxysilyl)furan was compromised by the formation of di- and triarylated products, indicating that 2-lithiofuran was successfully formed in situ, but was highly reactive. In contrast, attempted silylation of 2-bromopyridine led to exclusive formation of pyridine, indicating successful formation of the organolithium intermediate, but unsuccessful silylation. Dr. Chuljin Ahn confirmed the nearly quantitative formation of 2-lithiopyridine under our reaction conditions by capture with the more reactive electrophiles TMS-Cl and benzaldehyde.228 In an effort to increase the nucleophilicity of the pyridine metalloid, silylation of 3-bromo-4-methoxypyridine (101) was attempted with dramatic results. Handy isolated 25% of the desired monopyridinyl siloxane 102, in addition to significant quantities of dipyridinyl- and tripyridinyl siloxanes 103 and 104.111 Scheme 57 NMeO Br NMeO Si(OEt)3 1. BuLi, Et2O, -78 ?C 2. Si(OEt)4, -78 ?C 3. H2O, -78 ?C to 25 ?C + 102, 25%101 NMeO Si(OEt)4-n n 103, n = 2, 25% 104, n = 3, 20% Scheme 57 The results with pyridine derivatives were not surprising given the literature precedence. Using analogous reaction conditions, Schmidbaur and co-workers synthesized 2- and 3-(triethoxysilyl)pyridine in 10% and 16% yield, respectively (Table 2, entries 10 and 11).227 Zeldin carefully examined the synthesis of 3-pyridinyl substituted ethoxysilane monomers, focusing on the preparation of diethoxy(methyl)(3-pyridinyl)silane (PyrSiMe2(OEt)).210 A variety of silicon electrophiles were evaluated, most notably MeSiCl3 (followed by alcoholysis), MeSi(OEt)2Cl, and 55 MeSi(OEt)3, under different metallation conditions. Three significant conclusions were drawn: (1) MeSi(OEt)3 (3 equiv) was ideal for the methyl(diethoxy)silylation of 3-pyridinyl magnesium bromide, yielding 45% of the desired mono substitution product, and no diarylated adduct; (2) in contrast, a 20-fold excess of MeSiCl3 was required to mitigate multiple substitution by 3-lithiopyridine, and as with MeSi(OEt)2Cl, still yielded a significant amount of the diarylated silane and led to the formation of mixed silanes and silicates, for example 105 and 106 (Figure 10); and (3) the use of pyridinyllithium reagents led to the formation of a pentacoordinate anionic species between the silane product and LiOEt formed in the displacement reaction (107). The decomposition of complexes 105- 107 during distillation was blamed for the very low product yields.210 N Si N SiMeX2 X Me Cl- Si N X Me X SiMeX2 Si OEt Me EtO Li+ N 105, X = OEt, Cl 106, X = OEt, Cl OEt N 107 Figure 10. Complexes Formed in the Silylation of Pyridinyl Metal Reagents. Zeldin obtained the best results when the pyridinyl Grignard reagent was generated under Barbier conditions (Scheme 58, eq. a); in fact, when the preformed 3-pyridinyl magnesium bromide was treated with MeSi(OEt)3, no reaction was observed.210 Given that pyridinyl Grignard reagents are reputed to be unreactive toward most electrophiles,230,231 Zeldin?s result is striking. Unfortunately in our laboratories, using the Barbier reaction conditions reported by Zeldin, but substituting Si(OEt)4 as the electrophile, preparation of 3-(triethoxy)pyridine failed. The only products obtained were the homocoupled product 3,3?-dipyridyl (109) and pyridine (Scheme 58, eq. b). Homocoupling (Wurtz-type coupling) is a common side-reaction of organomagnesium reagents.115 56 Scheme 58 N SiMe(OEt)2 N Br 108 45% 1. Mg?, I2, THF, reflux 2. THF, reflux, 8 h 43 Si OEt OEt Me OEt (a) N N Br 109 30% 1. Mg?, I2, THF, reflux 2. THF, reflux, 8 h 43 Si OEt OEt EtO OEt (b)N Since the completion of this study, Handy and Correia applied the general method for the synthesis of arylsiloxanes from aryllithium reagents outlined in Table 5 to prepare indole derivatives 110 and 111 in 70% and 60% yield, respectively.24,229 These 5-indole siloxanes are synthetic intermediates in the preparation of biologically-interesting tryptamines and tetrahydropyridylindoles.111,232 For the synthesis of siloxane 110, Handy first converted 5-bromoindole to the potassium salt using KH; subsequent lithiation and treatment with Si(OEt)4 afforded the siloxane.24,111 NH Si(OEt)3 110 NMe Si(OEt)3 111 Figure 11. 5-Indole Siloxanes Prepared from Aryllithium Reagents for Use in Natural Product Syntheses. With both Grignard and lithium reagents, the formation of di- and triarylated silanes was best controlled by employing reduced reaction temperatures. In contrast, change of solvent or the use of large excesses of the electrophile had relatively little impact on the reaction outcome. It is unlikely that the temperature effect observed in the arylation is due solely to the attenuation of organometalloid reactivity at reduced temperature. Contrary to 57 expectation, lower temperatures have been shown to increase the reactivity of the organometallic species toward electrophiles by causing deaggregation of the organometalloid.115 This phenomenon would clearly favor multiple arylation of the silane. Instead, a reduction in diarylation was observed when the temperature was reduced; therefore, the reactivity of the nucleophile must not be greatly altered within the temperature range studied. In fact, it has been observed that phenyllithium does not change its aggregation state until temperatures as low as ?100 ?C are reached.115 The relative rates of reaction of alkoxysilanes with organometallic reagents follows the order of silane electrophilicity: Si(OEt)4 > RSi(OEt)3 > R2Si(OEt)2 > R3Si(OEt); the same order is observed with chlorosilanes.204 Given the relative reactivities, polyarylation should be stoichiometrically controllable, or mitigated by slow addition of the organometallic reagent to the electrophile. On the contrary, even at low temperature, with less than 1 equiv of organometallic reagent, diarylated and even triarylated byproducts are observed, suggesting that other factors are at play than organometalloid reactivity or electrophile electronics.122 The excepted mechanism of alkylation comprises formation of pentacoordinate adduct 113 as shown in Scheme 59.122,204,233 Diarylation is the result of either direct attack of a second aryl nucleophile on the pentacoordinate silicate 113 to form hexacoordinate intermediate 115, or attack on the tetracoordinate monoaryl silane 114. Scheme 59 OEt Si EtO OEtOEt Si OEt OEt OEt OEt Si OEtOEt OEt OEt Ph Ph Ph Ph-Li fast 113112 115 Li+ 2Li+ 2- slow Ph Si Ph OEtOEt 116 OEt Si Ph OEtOEt 114 + + 2 LiOEt LiOEt Ph-Li fastPh-Li slow 58 Again, it was observed that at lower temperatures the formation of diaryl products was minimal. We propose that at reduced temperature, the pentacoordinate silicate 113 formed by initial attack of the metalloid reagent does not decompose (with loss of alkoxide) to give the neutral monoaddition product 114. The stable, anionic monoaryl silicate 113 is less reactive than its neutral monoaryl silane counterpart 114; in effect, the monoaryl silane 114 is ?protected? as the pentacoordinate silane against multiple arylation. This result is quite different than what has been observed for chlorosilanes, where the analogous pentacoordinate tetrachloroaryl silicate (Ar-SiCl4-) has been found to be more reactive than the neutral tetracoordinate monoaryl silane (Ar-SiCl3).122,204,233 In conclusion, general reaction conditions for the synthesis of aryl(trialkoxy)silanes from aryl Grignard and lithium reagents and functional silanes have been developed. Although examples in the literature have described the use of a range of silicon electrophiles (including SiCl4 and Cl?Si(OR)3), tetraalkyl orthosilicates (Si(OR)4) allow for the most direct and convenient synthesis of arylsiloxanes, in that they are commercially available, inexpensive, and air and moisture stable. Using the reaction conditions developed herein, o-, m- and p-substituted bromoarenes underwent equally efficient metallation and silylation. Mixed results were obtained with heteroaromatic substrates: 3-bromothiophene, 3-bromo-4-methoxypyridine, 5-bromoindole, N-methyl-5-bromoindole all underwent silylation in good yield, whereas low yields of siloxanes were obtained from 2-bromofuran, and 2-bromopyridine failed to be silylated. Lower temperatures allowed for the formation of predominantly mono arylated siloxanes, without requiring more than 1.5- 3.0 equiv of the electrophile. Several important issues remain unresolved. Irrespective of the electrophile, the synthesis of aryl(trialkoxy)silane derivatives using Grignard or lithium reagents is plagued by modest yields caused by the formation of reduced (dehalogenated) substrate, and multiply arylated byproducts. These limitations are most likely inherent to the methodology: reduction of the organometalloid presumably occurs during the metallation step (and thus occurs in all reactions employing Grignard or lithium species); and multiple 59 alkylation stems from the high susceptibility toward nucleophilic attack of the hypercoordinate monoaryl(tetraethoxy)silicate intermediate. Improved Methods for the Synthesis of Aryl(triethoxy)silanes via Palladium(0)- Catalyzed Silylation of Aryl Iodides and Bromides with Triethoxysilane An alternative to the preparation of arylsilanes from organomagnesium or lithium intermediates was reported by Masuda and co-workers in 1997: the silylation of aryl iodide derivatives by triethoxysilane (H?Si(OEt)3) in the presence of a Pd catalyst (Scheme 60).221 Masuda reported that electron-rich, para-substituted aryl iodides gave excellent yields of aryl siloxane. The major byproduct of the reaction was the reduced arene (119). In contrast to aryl iodides, aryl bromides were reported to provide low yields of aryl siloxane under identical conditions. In addition, the silylation of ortho- or meta- substituted aryl iodides/bromides was not reported. Accordingly, the goals of this study were to extend the Masuda silylation reaction to aryl bromides and chlorides, and to evaluate the Masuda methodology for the synthesis of ortho- and meta-substituted aryl siloxanes. Scheme 60 H3CO Si(OEt)3 H3CO I H3CO H + 117 118 119 H-Si(OEt)3 i-Pr2NEt / NMP Pd(dba)2?CHCl3 P(o-tol)3 Toward our goal of synthesizing a variety of highly functionalized siloxane derivatives for use in aryl coupling reactions, it was necessary to determine whether Masuda conditions could be modified, and the scope extended to encompass silylation of aryl bromide derivatives. Standard optimization approaches involving change of solvent, reaction temperature, catalyst, base, and catalyst:ligand ratio failed to provide improved yields of siloxane. As reported by Masuda, the silylation reaction of phenyl bromide 60 under these conditions afforded the reduced arene rather than the desired siloxane derivative (see Scheme 60).221 Previous studies in our laboratory23 had shown that incorporation of Buchwald's phosphine 120 into Pd-catalyzed reactions resulted in dramatically improved yields of adducts.234-238 The use of Buchwald's and related phosphine derivatives in the silylation reaction was investigated and the results are summarized in Table 6. Entries 1 and 2 are the results obtained employing the standard conditions reported by Masuda and show that the silylation provided a low yield of the desired aryl siloxane. The major product was the reduced arene. Substitution of Buchwald's ligand (120) as the phosphine resulted in a marked improvement in the yield of siloxane obtained (entry 10). Similarly, inclusion of tributylphosphine (entries 6 and 7) , as reported by Fu,239 had a less significant effect, and showed an appreciable difference in yield depending on the solvent. Subsequent studies were performed with the Buchwald ligand because of its inherent ease of handling, thus avoiding the air-sensitive nature of tributylphosphine. A variety of other ligands was investigated including mono- and bidentate phosphines (dppp, dppf, etc.), Buchwald?s 2-(dicyclohexylphosphino)biphenyl, and triphenylarsine; however, these ligands failed to provide siloxane in acceptable yields. Similarly, additives such as tetraalkylammonium iodides, copper iodide, silver(I) salts, thallium acetate, or the inclusion of molecular sieves failed to improve the reaction outcome. Solvents that are either less polar than NMP, or that can coordinate with the catalyst resulted in dramatically reduced yields of the siloxane (Table 6, entries 12-14). An amine base was required for silylation. In the absence of base, slow reduction of the starting material was observed (entry 15). Triethylamine (entry 16) gave more reduced material when compared to diisopropylethylamine, possibly due to its ability to coordinate as a ligand. Pyridine (entry 17) and 2,6-lutidine (entry 18) slowed the reaction dramatically and resulted in reduction preferentially. Replacement of the amine base by weak alkali bases such as KOAc (entry 20) and CsCO3 (entry 21) was ineffective. (t-Bu)2P Ph 120 61 Br OCH3 H Si(OEt)3 Si(OEt)3 OCH3 H OCH3 + + Pd(dba)2 ligand base solvent 86 118 119 Conditionsa Yield (%)b,c Ligand Base Solvent Entry (10 mol %) (3 mmol) (4 mL) 118 119 1 P(o-tol)3 i-Pr2NEt NMP 21 79 2 P(o-tol)3 i-Pr2NEt DMF 15 70 3 none i-Pr2NEt DMF 14 86 4 PPh3 i-Pr2NEt DMF 0 5 5 dppf i-Pr2NEt NMP 0 8 6 P(t-Bu)3 i-Pr2NEt NMP 59 41 7 P(t-Bu)3 i-Pr2NEt DMF 45 55 8 P[(2,4,6-OMe)Ph]3 i-Pr2NEt NMP 8 92 9 P(cy)2(o-biphenyl) i-Pr2NEt NMP 17 83 10 P(t-Bu)2(o-biphenyl) i-Pr2NEt NMP 75 25 11 P(t-Bu)2(o-biphenyl) i-Pr2NEt DMF 36 64 12 P(t-Bu)2(o-biphenyl) i-Pr2NEt THF 6 54 13 P(t-Bu)2(o-biphenyl) i-Pr2NEt CH3CN 6 93 14 P(t-Bu)2(o-biphenyl) i-Pr2NEt toluene 3 19 15 P(t-Bu)2(o-biphenyl) none NMP 0 4 16 P(t-Bu)2(o-biphenyl) Et3N NMP 55 45 17 P(t-Bu)2(o-biphenyl) pyridine NMP 0 5 18 P(t-Bu)2(o-biphenyl) 2,6-lutidine NMP 32 26 19 P(t-Bu)2(o-biphenyl) DBU NMP 0 57 20 P(t-Bu)2(o-biphenyl) KOAc NMP 22 78 21 P(t-Bu)2(o-biphenyl) Cs2CO3 NMP 3 97 a Reactions of 4-bromoanisole (86) (1.0 equiv) with H-Si(OEt) 3 (1.5 equiv) wereallowed to stir at room temperature for 12 h in 4 mL of solvent by using Pd(dba) 2 (5 mol%), phosphine (10 mol %), and base (3.0 equiv). b GC yields are based on amount of 4-bromoanisole. c Remaining percentage was unreacted starting material. Table 6. Optimization of the Silylation of 4-Bromoanisole. In order to demonstrate the scope of the modified silylation reaction, a variety of substituted aryl bromides was examined (Table 7). The best yields of siloxane were obtained for electron-rich aryl bromides (entries 1-7) although the yields obtained are generally modest. This trend was not unexpected based on the report from Masuda with 62 iodides.221 Silylation of bromobenzene or electron-poor aryl bromides was unsuccessful (entries 8-10), exclusively resulting in dehalogenation of the starting material. Br H Si(OEt)3 Si(OEt)3 H R R R + +Pd(dba)2 P(t-Bu)2(o-biphenyl)(120) i-Pr2NEt NMP 90 92 121 Entry Aryl Bromide (90) % Yielda-c (92) Entry Aryl Bromide (90) % Yielda-c (92) 1 Br OCH3 68 6 Br OAc 30 2 Br CH3 43 7 Br OH 28d,e 3 O O Br 44 8 Br 5 4 Br NHAc 33 9 Br O 0 5 Br NH2 38d,e 10 Br Cl 0 a Reactions of arylbromide (90) (1.0 equiv) with H-Si(OEt) 3 (1.5 equiv) were allowed to stirat 60 ?C for 12 h in NMP by using 3 mol % Pd(dba) 2, (t-Bu)2P(o-biphenyl)(120) (6 mol %),and i-Pr 2NEt (3 equiv). b Isolated yield of purified (>95%) product. c Unless otherwise indicated, remaining percentage was reduced starting material (121). d Reaction allowed to stir at room temperature. e Reaction stopped at 2 h. Table 7. Silylation of Aryl Bromides. Mixed results were obtained for heteroaromatic systems (Table 8). For example, a satisfactory yield of 2-(triethoxysilyl)thiophene was obtained from the corresponding bromide (entry 1), whereas, silylation of 2- and 3-bromopyridine and 2- and 3-bromoquinoline failed (entries 2-5), probably due in part to nitrogen functioning as a ligand for the metal (vide infra). Support for this hypothesis is that use of Pd(0)/pyridine as a catalyst system for the silylation of 4-bromoanisole (Table 1, entry 17) resulted in a 63 dramatic reduction in the rate of silylation. Not surprisingly, the more electron-rich substituted pyridine derivative 3-bromo-4-methoxypyridine was converted to the siloxane in fair yield (Table 8, entry 6). H Si(OEt)3Ar-Br Ar-Si(OEt)3 Ar-H+ Pd(dba)2 P(t-Bu)2(o-biphenyl)(120) i-Pr2NEt NMP122 123 124 + Entry Heteroaryl Bromide (122) % Yielda-c (123) Entry Heteroaryl Bromide (122) % Yielda-c (123) 1 S Br 43 4 N Br 0 2 N Br 0d 5 N Br 0 3 N Br 0 6 N Br MeO 57 a Reactions of heteroarylbromide (122) (1.0 equiv) with H-Si(OEt) 3 (1.5 equiv) were allowed to stir at 60 ?C for 12 h in NMP by using 3 mol % Pd(dba)2, (t-Bu)2P(o-biphenyl)(120) (6 mol %), and i-Pr2NEt (3 equiv). b Isolated yield of purified (>95%) product. c Unless otherwise noted, the remaining percentage was reduced starting material (124). d Trace amounts of 2,2?-dipyridyl were observed by G.C. analysis. Table 8. Silylation of Heteroaryl Bromides. Given that the yield of silylated arene was dependent upon the electron-rich character of the arene system, it was expected that 4-bromo-1,2-(methylenedioxy) benzene (Table 7, entry 3) would give a comparable (or greater) yield of siloxane than 4-bromoanisole (entry 1). Surprisingly, the opposite was observed. It was proposed that the presence of a substituent at the meta position resulted in the lower yield. Further exploration of the effect of substituent placement on the silylation of bromoanisoles was undertaken and the results are summarized in Table 9. The yield of siloxane rapidly declined as the methoxy substituent shifted from the para (68%) to meta (25%) to ortho 64 (0%) position. Attempts to improve the yield of meta- or ortho-aryl siloxanes by altering the phosphine ligand were unsuccessful. Br H Si(OEt)3 Si(OEt)3 H R R R + +Pd(dba)2 P(t-Bu)2(o-biphenyl)(120) i-Pr2NEt NMP 90 92 121 Entry Aryl Bromide (90) % Yield (92)a-c Entry Aryl Bromide (90) % Yield (92)a-c 1 Br OCH3 68 4 Br CH3 43 2 Br OCH3 25 5 Br CH3 10 3 Br H3CO 0 6 Br H3C 0 a Reactions of arylbromide (90) (1.0 equiv) with H-Si(OEt) 3 (1.5 equiv) were allowed to stir at 60 ?C for 12 h in NMP using 3 mol % Pd(dba)2,(t-Bu)2P(o-biphenyl) (120) (6 mol %), and i-Pr2NEt (3 equiv). b Yields are isolated yields of purified (>95%) product. c Remaining yield was of reduced starting material (121). Table 9. Silylation of Aryl Bromides: Effect of Substituent Position. There are several possible explanations for the effect of ortho and meta substitution on the silylation reaction. It was initially thought that coordination of the Lewis basic oxygen lone pair of the methoxy substituent to palladium or silicon was responsible for this outcome (Figure 12). However, a non-coordinating ortho-methyl substituent also prevented silylation (Table 9, entry 6) indicating that steric interference also had a significant role. Additionally, coordination of the ortho-methoxy oxygen to silicon was not observed by either 29Si or 19F-NMR,111 although this result does not rule out the possibility that the dative bond in the hypercoordinate system 127 forms under the reaction conditions. Sterically hindered ortho- and meta-substituted aryl silanes have been shown to be more susceptible to acid240-244 and base229,245-248-catalyzed protodesilylation than the 65 less hindered para-substituted analogs, regardless of the nature of the ortho substituent. Truthfully, it is proposed that both steric and electronic factors are responsible for the decreased yields in this reaction. The potential effects of substituent position on the mechanism of the silylation reaction are discussed in detail below. OMe Si OMeOMe MeO OMe Si OMe OMe OMe Si OMe OMe 125 126 127 MeO MeO Figure 12. Proposed Effect of Substituent Position. The modified reaction conditions were also applied to a variety of substituted aryl iodides in order to demonstrate the generality of the new ligand system and to determine the impact of arene substituent location on the silylation of iodides (Table 10). In agreement with the report of Masuda, electron-rich para-substituted aryl iodides underwent silylation more efficiently than the corresponding aryl bromides.221 Similar yields of siloxanes were achieved with aryl iodides when either the Pd(dba)2/P(t-Bu)2(o-biphenyl)(120) system or Masuda's Pd2(dba)3?CHCl3/P(o-tol)3 system was employed. Aryl iodides were less sensitive to the electronic nature of the activating group than the corresponding bromides. For example, both 4-bromo (Table 10, entry 8) and 4-chloroiodobenzene (entry 9) gave comparable yields of siloxane to electron-rich aryl iodides (i.e., O-, N-, and alkyl-substituted). In contrast, only electron-rich aryl bromide analogs underwent silylation under these conditions. In the case of 4-bromo-iodobenzene (entry 8), the reaction required close monitoring for consumption of starting material because dehalogenation of the desired product, 4-(triethoxysilyl)-bromobenzene, occurred under these conditions. Protection of other functionalities was unnecessary as shown by the excellent yields obtained with 4-iodoaniline (entry 11) and 4-iodophenol (entry 13). Prolonged reaction of iodoaniline not only lowered the yield of siloxane, but also affected silylation of the unprotected amine nitrogen. 66 H Si(OEt)3Ar-I Ar-Si(OEt)3 Ar-H+ Pd(dba)2 P(t-Bu)2(o-biphenyl)(120) i-Pr2NEt NMP128 123 124 + Entry Aryl Iodide (128) % Yield (123)a-c Entry Aryl Iodide (128) % Yield (123)a-c 1 I OCH3 86 10 I NHAc 68e 2 I OCH3 52 11 I NH2 77e 3 I H3CO 10 12 I OAc 75e 4 I CH3 80 13 I OH 70d 6 I CH3 65 14 I O 24 7 I H3C 0d 15 N I 0d,g 8 I Br 68e,f 16 N I 10 9 I Cl 75e 17 S I 92 a Reactions of aryliodide (128) (1.0 equiv) with H-Si(OEt) 3 (1.5 equiv) were allowed to stirat 60 ?C for 12 h in NMP using Pd(dba) 2 (3 mol %), (t-Bu)2P(o-biphenyl) (120) (6 mol %),and i-Pr 2NEt (3 equiv). b Yields shown in parentheses are GC yields; all other yields are isolated yields of purified (>95%) product. c Unless noted otherwise, the remaining percentage was of reduced starting material (124). d Yield confirmed by GCMS. e Reaction allowed to stir at room temperature. f Reaction stopped at 2 h. g Dipyridyl was isolated in 15% yield. Table 10. Palladium-Catalyzed Silylation of Aryl Iodides. 67 In contrast to the electron-rich aryl iodide systems, 4-iodoacetophenone was silylated in only 24% yield (entry 14) (see other examples in Masuda221). Recall that the bromo analog failed to be silylated under these conditions (Table 7, entry 9). It is noteworthy that even the low yield was a triumph since this acetophenone analog cannot be synthesized by the organometallic approach (vide supra) due to the presence of the reactive carbonyl group. Once again, the position of the substituent greatly impacted the reaction outcome: ortho-iodoanisole (Table 9, entry 3) or toluene (entry 7) gave poor yields of product (10%, and 0%, respectively); whereas their meta- and para-iodo arene counterparts gave acceptable yields. The results from silylation of 2- and 3-iodopyridine are notable. While in all previous examples the Pd-mediated silylation reaction yielded either aryl siloxane or reduced starting material (Scheme 60), homocoupling of 2-iodopyridine was observed. The homocoupled product [2,2']Bipyridinyl (130, Scheme 61) was isolated in 15% yield from the attempted silylation of 2-iodopyridine (the reaction went to completion, with pyridine as the only other product). The homocoupling of 2-halopyridines in the presence of a palladium catalyst is a known reaction.249-251 As a control, 2-iodopyridine was exposed to the standard silylation conditions in the absence of triethoxysilane, and [2, 2']dipyridyl (130) was isolated in quantitative yield (Scheme 61). In contrast, when 2-bromopyridine was used in the control experiment, only a trace amount of the homocoupled product 130 was formed. The remainder of the reaction mixture was starting material. The results of the control experiments explain why no homocoupled product was observed in the attempted silylation of 2-bromopyridine, whereas homocoupling was a significant side- reaction in the silylation of 2-iodopyridine. Scheme 61 N X N Ni-Pr2NEt / NMP 129, X = I 42, X = Br 130 X = I, 99% X = Br, 6% Pd(dba)2 P(t-Bu)2(o-biphenyl)(120) 68 Attempts to silylate 3-iodopyridine were problematic also, although no homocoupled adduct was observed. We obtained only 10% of the desired product using either Masuda's conditions or our modified reaction conditions (Table 10, entry 16). These results were surprising because Masuda reported a 56% yield of 3-(triethoxysilyl)pyridine using Pd2(dba)3?CHCl3/P(o-tol)3 in NMP (1h, 80 ?C).221 Again it appears that the metal coordinating ability of the pyridine nucleus is the culprit since 2-iodothiophene was readily silylated (92% yield, Table 9, entry 17). Aryl triflates are appealing substrates for silylation since they are readily derived from the phenol,71 and are often more available than the corresponding aryl halides. However, aryl triflates failed to undergo efficient silylation under the standard conditions used for aryl iodides and bromides (Table 11). As seen in entries 1 and 9, electron-rich and deficient aryl triflates underwent reduction upon treatment with excess Hunig?s base and triethoxysilane with catalytic Pd(dba)2/P(t-Bu)2(o-biphenyl)(120). A base was required for complete reaction turnover (entries 2,4,10, and 12), and the reaction tolerated both polar and non-polar solvents. The electron-rich substrate 4-methoxyphenyl triflate underwent reduction most efficiently using the alternative Buchwald ligand, P(cy)2(o-biphenyl) (entry 8), whereas the electron-deficient substrate 4-nitrophenyl triflate tolerated a range of phosphine ligands: P(t-Bu)2(o-biphenyl) (120), P(cy)2(o-biphenyl), and PPh3 all affected efficient reduction. 69 OTf R H Si(OEt)3 Si(OEt)3 R H R + + Pd(dba)2 ligand base solvent 131 92 121 Conditionsa Yield (%)b,c Entry R Ligand Base Solvent 92 121 1 OMe P(t-Bu)2(o-biphenyl) i-Pr2NEt NMP 3 90 2 OMe P(t-Bu)2(o-biphenyl) none NMP 0 5 3 OMe P(t-Bu)2(o-biphenyl) i-Pr2NEt dioxane 2 96 4 OMe P(t-Bu)2(o-biphenyl) none dioxane 0 7 5 OMe PPh3 i-Pr2NEt NMP 0 5 6 OMe PPh3 i-Pr2NEt dioxane 0 24 7 OMe P(cy)2(o-biphenyl) i-Pr2NEt NMP 1 32 8 OMe P(cy)2(o-biphenyl) i-Pr2NEt dioxane 1 99 9 NO2 P(t-Bu)2(o-biphenyl) i-Pr2NEt NMP 0 100 10 NO2 P(t-Bu)2(o-biphenyl) none NMP 0 22 11 NO2 P(t-Bu)2(o-biphenyl) i-Pr2NEt dioxane 0 100 12 NO2 P(t-Bu)2(o-biphenyl) none dioxane 0 15 13 NO2 PPh3 i-Pr2NEt NMP 0 100 14 NO2 PPh3 i-Pr2NEt dioxane 0 100 15 NO2 P(cy)2(o-biphenyl) i-Pr2NEt NMP 0 100 16 NO2 P(cy)2(o-biphenyl) i-Pr2NEt dioxane 0 100 a Reactions of aryltriflate (131) (1.0 mmol) with H-Si(OEt) 3 (1.5 mmol) were performed atroom temperature for 12h in 4 mL of solvent by using Pd(dba) 2 (5 mol %), phosphine (10mol %), and base (3 mmol). b GC yields are based on amount of aryltriflate (131). c Remaining percentage was unreacted starting material. Table 11. Silylation of Aryltriflate Derivatives. The Pd-catalyzed reduction of aryl triflates using triethylsilane (Et3Si?H) was reported by Kotsuki in 1995 (Scheme 62).252 Under mild conditions, a variety of aryl triflates underwent deoxygenation in excellent yield. It was observed that triethylamine increased the reaction rate. With further optimization, triflate deoxygenation using H-Si(OEt)3 in dioxane (or other volatile, low-polarity solvents such as THF or toluene) under our conditions could prove to be synthetically useful, and more attractive than existing methods, which require DMF.252 70 Scheme 62 H-SiEt3 Pd(OAc)2 dppp DMF 100% OTf H Two mechanisms have been proposed for the palladium-catalyzed silylation reaction, both of which account for the formation of reduced arene byproducts.221 Both the C-X bond of aryl halides and the Si-H bond of silanes are known to undergo oxidative addition to Pd(0) complexes.45,253-258 Thus, two logical mechanisms have been proposed, one with Ar-Pd(II)-X, and the other with R3Si-Pd(II)-H as the key intermediate, respectively. Scheme 63 illustrates a reasonable mechanism for the oxidative addition of the aryl halide to Pd(0), to form Ar-Pd(II)-X complex 132. Transmetallation with the silane forms four-centered palladium intermediate 133, which in turn undergoes reductive elimination to liberate aryl silane and regenerate the Pd(0) catalyst. Support for this mechanism was provided by Chatgilialoglu, who demonstrated that organic halides underwent oxidative addition to Pd(0) catalysts more readily than hydrosilanes.255 Additionally, it was shown that the mechanism did not involve the intermediacy of carbenium ions or free-radical species. Scheme 63 H(EtO)3Si Ar I (EtO)3Si Ar Ln Pd(II) I Ar HI Pd(II) I Ln H (EtO)3Si Pd I H Ar Pd(0)Ln 132134 133 71 A mechanism involving oxidative addition of the hydrosilane followed by ?-bond metathesis has been proposed by Kunai and Ishikawa for the exchange of the Si-H and C-I bonds of alkyl and aryl iodides and hydrosilanes in the presence of a Pd(0) catalyst (Scheme 64).258 The mechanism for silylation begins with formation of a Si-Pd bond by oxidative addition of Pd(0) to the silane. Silylation of the aryl halide results from ?-bond metathesis, for overall Si-H/C-I bond exchange and regeneration of the Pd(0) catalyst. Scheme 64 H(EtO)3Si Ar I(EtO)3Si Ar Pd(0)Ln (EtO)3Si Pd(II) H Ln (EtO)3Si Pd H IAr Ln HI Pd(II) H Ln I 135134 136 Both mechanisms account for the formation of dehalogenated aryl halide substrate when ortho substituents are present on the arene. Reductive elimination of silylated arene (Ar-Si(OEt)3) from 4-coordinate Pd intermediate 133 requires that the ?Si(OEt)3 and the ?Ar ligands adopt a cis geometry, as shown above in Scheme 64. Steric bulk on the arene will disfavor the cis conformation, and isomerization of the complex reorients the two largest substituents to the trans geometry on the palladium complex (137, Scheme 65). As a consequence, the Ar- and H- ligands are placed preferentially cis, and Ar-H is formed upon reductive elimination from palladium. 72 Scheme 65 137133 (EtO)3Si Pd I H Ar (EtO)3Si Pd Ar H I Ar H (EtO)3Si I+ Similarly, for the proposed mechanism entailing ?-bond metathesis, steric bulk on the arene will disfavor the approach of Ar-X on the R3Si-Pd(II)-H intermediate which places the large arene adjacent to the bulky ?SiR3 ligand, as shown in transition state 136 in Figure 13. Instead, the less sterically demanding approach 139 should be favored, and formation of Ar-H in lieu of the desired Ar-SiR3 results. (EtO)3Si Pd H IAr Ln 136 (EtO)3Si Pd H ArI Ln 139 Figure 13. Proposed Transition States for the Metathesis Reaction Mechanism. In order to determine which of the above two mechanisms most likely is at play in this process, Masuda demonstrated that neither silylation nor reduction occurred when a preformed Ar-Pd(II)-I complex was treated with (EtO)3SiH and triethylamine, indicating that the mechanism of reaction likely does not involve initial oxidative addition to the aryl halide.221 Instead, the latter mechanism entailing initial oxidative addition into the silane, followed by ?-bond metathesis seems the most likely pathway for the formation of aryl siloxane product. In an effort to circumvent formation of reduced arene byproducts, the silylation reaction was attempted using hexamethoxydisilane ((MeO)3Si-Si(OMe)3) (Scheme 66). Alkylchloro- and alkylalkoxydisilanes had been shown previously to participate in the Pd- mediated silylation of aryl halides (vide supra).45,197,198,257,259-267 It is notable that the 73 analogous Pd-mediated process using hexachlorodisilane did not yield trichlorosilyl arenes (which could be converted to the siloxane upon alcoholysis).265 Scheme 66 (MeO)3Si Si(OMe)3 + X R Pd(0) Si(OMe)3 R X-Si(OMe)3 Unfortunately, aryl halides failed to undergo silylation using hexamethoxydisilane under the conditions used for the silylation of aryl iodides and bromides with hydrosilane (Table 12). Under standard silylation conditions, no reaction was observed with electron- rich p-haloanisole derivatives (entries 1-4). In contrast, the electron-deficient system of 4-bromoacetophenone underwent complete reduction using standard as well as modified silylation conditions (entries 5 to 9). Hatanaka reported the palladium-mediated, fluoride-induced trimethylsilylation of electron-rich and electron-neutral aryl iodides with hexamethyldisilane (Scheme 67). Under the same conditions reported by Hatanaka using hexamethoxysilane, no reaction was observed (Table 12, entries 10-12). Finally, the best results were obtained under forced reaction conditions (entry 13): treatment of bromobenzene with hexamethoxydisilane for 72 hours in a sealed tube at 140 ?C in the presence of catalytic Pd(0) yielded 27% of the desired silane, as well as reduced arene and biphenyl contaminants. Scheme 67 Me3Si-SiMe3 Pd(PPh3)4 (20 mol%) TASF HMPA, 25 ?C 100% IMe SiMe3Me 74 X (MeO)3Si Si(OMe)3 Si(OMe)3 H R R R + + Pd(0) ligand base solvent 139 92 121 Substrate Conditionsa Yield (%)b,c Entry X R Ligand Pd catalyst / base solvent T (?C) 92 121 1 I OMe P(t-Bu)2(o-biphenyl) Pd(dba)2 / i-Pr2NEt NMP 25 0 0 2 I OMe P(o-tol)3 Pd(dba)2?CHCl3/i-Pr2NEt NMP 25 0 2 3 Br OMe P(t-Bu)2(o-biphenyl) Pd(dba)2 / i-Pr2NEt NMP 25 0 0 4 Cl OMe P(t-Bu)2(o-biphenyl) Pd(dba)2 / i-Pr2NEt NMP 25 0 0 5 Br Ac P(t-Bu)2(o-biphenyl) Pd(dba)2 / i-Pr2NEt NMP 100 0 100 6 Br Ac P(t-Bu)2(o-biphenyl) Pd(dba)2 / i-Pr2NEt DMPU 100 0 100 7 I Ac P(t-Bu)2(o-biphenyl) Pd(dba)2 / KF (aq) DMPU 75 0 100 8 Br Ac P(t-Bu)2(o-biphenyl) Pd(dba)2 / KF (aq) DMPU 75 0 100 9 Br OMe P(t-Bu)2(o-biphenyl) Pd(dba)2 / K2CO3 (aq) DMPU 75 0 0 10 I OMe Pd(PPh3)4 / TBAF NMP 25 0 0 11 I Me Pd(PPh3)4 / TBAF DMF 25 0 3 12 I Me Pd(PPh3)4 / TBAF HMPA 25 0 3 13d,e Br H Pd(PPh3)4 toluene 140 27 10 a Reactions of aryl halide 139 (1.0 equiv) with (MeO) 3Si-Si(OMe)3 (2.0 equiv) wereperformed using the indicated temperature and solvent for 12h-48h by using the given palladium catalyst (5 mol %), phosphine (10 mol %), and base (3 equiv). b GC yields are based on amount of aryl halide 139. c Unless otherwise noted, remaining percentage was unreacted starting material (139). d Reaction was performed in a sealed tube for 72h. e Biphenyl was observed in 7% yield. Table 12. Silylation of Arylhalides Using Hexamethoxydisilane. The results of the silylation reaction employing hexamethoxydisilane are surprising. In the absence of a hydride source, it was anticipated that reduction of the aryl halide would be suppressed. One possible hydrogen atom source is methoxide, which can be generated upon decomposition of the disilane under the harsh reaction conditions.198 Previous researchers have demonstrated the catalytic dehalogenation of a large variety of aryl chlorides and bromides with sodium or potassium methoxide mediated by Pd(0) (Scheme 68).268 75 Scheme 68 Pd(dba)2, KOMe, 140 dioxane, 100 ?C, 1h 100% Cl O O N N 140 The proposed mechanism of dehalogenation by methoxide under palladium catalysis is depicted in Scheme 69.268 The mechanism begins with oxidative addition of the aryl halide to Pd(0) to form the Ar-Pd(II)-X intermediate 132, followed by methoxide attack with displacement of the halide to give key intermediate 141. Then, ?-hydride elimination and liberation of formaldehyde gives the palladium-hydride complex 134, which can reductively eliminate reduced arene and regenerate the catalyst. Scheme 69 Ar X Pd(0)Ln Ln Pd(II) X Ar 132 MeO XLn Pd(II) OMe Ar 141 H H O Ln Pd(II) H Ar 134 Ar H 76 Rich pioneered the use of aryl acyl chloride derivatives as substrates for silylation (Scheme 70).196,269,270 For example, treatment of m-nitrobenzoyl chloride (142) with 1,1,2,2-dichlorodimethyldisilane under palladium catalysis resulted in decarbonylative silylation to form aryl silane 143.196 Scheme 70 MeCl2Si-SiCl2Me (PhCN)2PdCl2 PPh3 145 ?C, 20h 69% SiCl2MeO2NO 2N Cl O + CO + MeCl2Si-Cl 142 143 Rich demonstrated the efficient silylation of para and meta-substituted electron rich and deficient aromatic acid chlorides utilizing methylchlorodisilanes; ortho-substituted acid chlorides failed to undergo carbonylative silylation.196 It is notable that the analogous method using hexachlorodisilane did not successfully form trichlorosilyl arenes (which could be converted to the siloxane upon alcoholysis). Rich reported that hexachlorodisilane underwent decomposition under the conditions of the reaction. Lastly, a number of byproducts were observed, including benzil and diarylketone derivatives, especially when electron-deficient substrates were employed. The decarbonylative silylation reaction as reported by Rich was attempted in our laboratories, using hexamethoxydisilane (Scheme 71, eqn. a). Benzoyl chloride was heated without solvent in a sealed tube with hexamethoxydisilane and (PhCN)2PdCl2 in the presence of triphenylphosphine. However, none of the desired product was formed; instead, quantitative formation of methylbenzoate was observed. Clearly, the harsh reaction conditions elicited formation of methoxide via decomposition of the disilane, resulting in the methanolysis of the acyl chloride moiety. The thermal decomposition of chloro196 and alkoxy198 disilanes was reported, previously. Formation of arylsilane was observed when a different catalyst was employed (Scheme 71, eqn. b), however formation of the methyl ester was still a significant side-reaction. Less harsh reaction 77 conditions resulted in no reaction. Efforts to employ the disilane for the synthesis of aryl siloxanes from aryl halides and acyl chlorides were abandoned in light of the poor results of these preliminary studies, the expense of the disilane reagent, and the harsh reaction conditions. Scheme 71 (MeO)3Si-Si(OMe)3 (PhCN)2PdCl2 PPh3 145 ?C, 20h Si(OMe)3 Cl O + OMe O 0 % 100 % (MeO)3Si-Si(OMe)3 Pd(PPh3)4 145 ?C, 20h Si(OMe)3 Cl O + OMe O 45 % 55 % (a) (b) In summary, the Pd-mediated silylation of electron-rich aryl halides has been extended to include both iodides and bromides. For bromo arenes, the inclusion of Buchwald's phosphine ligand (120) is crucial. Although aryl iodides remain the substrates of choice for silylation, an acceptable yield of arylsiloxane may be obtained from the corresponding bromide. This is an advantage since aryl bromides are generally less expensive and more widely available than the analogous aryl iodides. Conclusions The scope of the silylation protocol has been fully defined: the synthesis of aryl and heteroaryl siloxanes using H-Si(OEt)3 and a Pd catalyst is limited generally to electron-rich, para- and meta-substituted aryl bromides and iodides. Aryl chlorides were inert under the reaction conditions, and triflates were poor substrates for silylation, instead undergoing highly efficient reductive deoxygenation. The optimum silylation reagent is triethoxysilane. This silylation method is an excellent companion to the more traditional organometallic approach to the formation of the Ar-Si bond. Case in point, ortho- substituted aryl siloxanes are readily synthesized from the Grignard or lithium reagent. 78 Unlike the metallation approach, the Pd-catalyzed silylation technique can be performed in the presence of a wide range of functional groups, including carbonyl-containing electrophiles, and protic moieties such as phenols or primary amines. To date, a large variety of simple aryl and heteroaryl halides has been shown to undergo silylation by one or both of the above protocols (Figure 14).24,25,68,111,229 70% 64% 65% 74% 68% 71% Br OMe Br OMe Br OMe Br Me Br Me Br Me 43% 50% 22% 70% 60% S Br O Br N 57% Br OMe Br Br HN MeN 33% 38% 30% 28% 55% Br NHAc Br Br Br OH Br NH2 OAc S Br 86% 52% 10% 80% 65% I OMe I OMe I OMe I Me I Me 92% 68% 77% 75% 70% I NHAc I I I OHNH2 OAcS I N 10% I 80% I Ac Figure 14. Yields of Siloxanes Using Arylmetalloid Reagents or Palladium-Mediated Silylation. The siloxane derivatives synthesized by the methods described herein have been utilized in palladium-catalyzed cross-coupling. The synthesis of aryl siloxanes has been shown to be comparable in yield to the synthesis of aryl boron reagents. In 79 contrast to the Suzuki reagents, siloxanes are easily purified by chromatography, distillation, or even recrystalization (in the case of aryl catecholates and silatranes). Although aryl tin reagents remain the easiest to synthesize, the silicon-based cross- coupling technique is an attractive alternative to Stille coupling, due to the low toxicity of the siloxane cross-coupling reagents. Application of these methodologies to the synthesis of complex, biologically active materials has been investigated in our laboratory. For example, the synthesis of a variety of highly functionalized siloxane derivatives has been accomplished. Coupling of these siloxanes to provide intermediates in the total synthesis of pancratistatin, streptonigrin, lavendamycin, and colchicine are being investigated by other members of the group and these results will be reported in due course. A sample of highly functionalized siloxanes prepared by these methods includes 93, 94, 142 and 143 (Figure 15). O O Si(OEt)3 93, 74% O O Si(OEt)3 94, 50% OMe MeO MeO Si(OEt)3 142, 50% OMOM N OMeMe Me Si(OEt)3 143, 30% Figure 15. Siloxane Intermediates for Use In Natural Product Syntheses. Epilogue Since the publication of the above method for aryl siloxane synthesis via Pd(0)- catalyzed silylation of aryl bromides and iodides in 2001,271 Masuda reported the Rh(I)- catalyzed silylation of aryl iodides and bromides (Scheme 72).272 The silylation is analogous to the Pd(0) silylation, in that the reaction requires a tertiary amine and polar amide solvent. Again, dehalogenation of the aryl halide starting material was the major side-reaction. Significantly, the substitution of Rh(I) for Pd(0) as the catalyst allowed for the silylation of electron-deficient aryl halides, as well as hindered, ortho?substituted silanes. The Rh(I)-catalyzed process is limited by expense of, and the need to 80 synthesize the [Rh(cod)MeCN)2]BF4 catalyst, but nonetheless presents a general method of silylation, encompassing electron-rich as well as electron-deficient aryl halides, including ortho-substituted arenes. The silylation of chloro arenes remains elusive. Scheme 72 Br Et3N Bu4NI 90% + H-Si(OEt)3 Si(OEt)3[Rh(cod)(MeCN)2]BF4 EtO2C I Et3N 81% + H-Si(OEt)3 Si(OEt)3[Rh(cod)(MeCN)2]BF4 Me Me EtO2C 81 Experimental General. Thin layer chromatography (TLC) was performed on 0.25 mm Merck silica- coated glass plates treated with a UV-active binder, with the compounds being visualized in one or more of the following manners: UV (254 nm), iodine, or vanillin/sulfuric acid charring. Flash chromatography was performed using thick-walled glass columns and medium-pressure silica gel (Davisil? 200-425 mesh) as described by Still.273 Flash chromatography data is reported as: (column diameter in mm, column height in cm, solvent). Infrared spectra were recorded on a Nicolet 5DXC FT-IR spectrophotometer with the samples prepared as stated. Band positions are given in reciprocal centimeters (cm-1) and relative intensities are listed as br (broad), vs (very strong), s (strong), m (medium), or w (weak). Melting points were taken in Kimax soft glass capillary tubes using a Thomas- Hoover Uni-Melt capillary melting point apparatus (Model 6406K) equipped with a calibrated thermometer. Melting points are corrected. Nuclear magnetic resonance (1H, 13C) spectra were recorded on a Bruker DRX-400 spectrometer. Chemical shifts are reported in parts per million (?) downfield from tetramethylsilane (TMS). Coupling constants (J values) are reported in Hertz (Hz), and spin multiplicities are indicated by the following symbols: s (singlet), d (doublet), t (triplet), q (quartet), m (multiplet), br s (broad singlet). Low resolution (MS) and high resolution mass spectra (HRMS) were obtained on a VG-7070E magnetic sector instrument equipped with a 486 PC-based data system. GCMS was performed on a Shimadzu QP5000MS coupled with a GC17A gas chromatograph. Gas chromatography was performed on a Hewlett Packard 5890 GC equipped with a flame ionization detector using a 25 m methyl silicon column. 82 Methylene chloride (CH2Cl2), tetraethyl orthosilicate (Si(OEt)4), tetramethyl orthosilicate (Si(OMe)4), methyl sulfoxide (DMSO), dimethyl formamide (DMF), 1-methyl-2-pyrrolidinone (NMP), toluene, pyridine, and acetonitrile (MeCN) were each distilled from calcium hydride. Tetrahydrofuran (THF), dioxane and diethyl ether (Et2O) were each distilled from sodium-benzophenone ketyl. Methanol (MeOH) and ethanol (EtOH) were stored over molecular sieves. Palladium(II) acetate (Pd(OAc)2), bis(dibenzylideneacetone)palladium (Pd(dba)2), dichlorobis(triphenylphosphine)palladium(II) (PdCl2(PPh3)2), and tetrakis(triphenylphosphine)palladium(0) (Pd(PPh3)4) were purchased from Acros and used as received. Tri-o-tolylphosphine (P(o-tol)3) was purchased from Acros and used as received. Triphenylphosphine (PPh3) was purchased from Aldrich and recrystallized from hexanes prior to use. 2-(Di-t-butylphosphino)biphenyl (120) (P(t-Bu)2(o-biphenyl)) was purchased from Strem and recrystallized from methanol (MeOH) prior to use. The ligand 2-(Dicyclohexylphosphino)biphenyl (P(cy)2(o-biphenyl)) was purchased from Strem and recrystallized from absolute ethanol (EtOH) prior to use. Tri-t-butylphosphino (P(t-Bu)3) was purchased from Strem as a 10 wt% solution in hexane and stored under argon. All other phosphines and triphenylarsine were purchased from Aldrich, stored under argon, and used as received. Aryl halides, heteroaryl halides and alkyl halides were prepared by literature procedure as noted, or purchased from Acros or Aldrich and purified using the method of Perrin274 prior to use. N-(4-Bromophenyl) acetamide275, 4-acetoxybromobenzene276, 4-acetoxyiodobenzene276, 2-iodopyridine277, 3-iodopyridine278, 2-bromoquinoline279, 3-bromoquinoline280, and 5-bromo-2-methoxypyridine281 were each prepared according to the literature procedure. In the case aryl sulfonates, the appropriate sulfonate was synthesized from the phenol immediately before use using the literature method283,284 and the crude isolated sulfonate was used without further purification. Triethoxysilane 83 (H-Si(OEt)3), and diisopropylethylamine (i-Pr2NEt), were purchased from Aldrich, stored in a dessicator, and used as received. Triisopropoxysilane (H-Si(O-i-Pr)3) and hexamethoxydisiloxane ((MeO)3Si-Si(OMe)3) were purchased from Gelest, stored in a dessicator, and used as received. Tetrabutylammonium fluoride (TBAF) was purchased from Acros as a 1.0 M solution in THF and used as received. Tetrabutylammonium triphenyldifluorosilicate (TBAT) was prepared according to the literature procedure.285 n-Butyllithium (1.6 M solution in hexane), and t-butyllithium (1.5 M solution in pentane) were purchased from Acros and used as received. The alkyllithium concentration was confirmed by titration with diphenylacetic acid using the method of Kofron.286 Magnesium turnings were cleaned as follows: enough turnings were placed in a beaker to form a thin layer. Dilute HCl solution (0.1 M) was added to just cover the magnesium turnings. The turnings were stirred until bubbling ceased and then the solution was decanted off and the process repeated three times or until the metal had a shiny mirror appearance. The turnings were then washed in the beaker three times each with water, ethanol, and diethyl ether. Finally, the turnings were dried in an oven at 120 ?C for five minutes. Glassware used in the reactions described below was dried for a minimum of 12 h in an oven at 120 ?C. All reactions were run under an atmosphere of nitrogen or argon at room temperature unless otherwise noted. All compounds were determined to be >95% pure by 1H NMR unless otherwise noted. 84 General Procedure for the Optimization of the Synthesis of 4- (Triethoxysilyl)anisole (88) Using 4-Bromomagnesium anisole (87) (Table 2). A 10 mL, 3-neck pear-shaped flask was fitted with an addition funnel, a reflux condenser, a rubber septum, and a stirbar. The flask was then charged with freshly washed magnesium turnings (134 mg, 5.50 mmol), flame-dried under vacuum, and back-filled with argon. THF (1.0 mL) was added to the magnesium turnings via syringe. The addition funnel was charged with 4-bromoanisole (86) (935 mg, 628 ?L, 5.00 mmol) in 2.0 mL THF. The reaction was initiated by addition of 5-10 drops of the bromoanisole solution to the magnesium turnings with stirring, followed by gentle heating. The rest of the bromoanisole solution was then added at such a rate that the THF maintained a moderate reflux. Upon final addition, the solution was allowed to stir at room temperature for 1 h, at which point GC analysis of a quenched aliquot of the reaction mixture indicated complete consumption of 4-bromoanisole. The 4-(bromomagnesium)anisole solution was then transferred via cannula to a second flame-dried addition funnel, to which was fitted a 50 mL round-bottom flask containing the indicated silane (1.5-3.0 equiv) and the internal standard naphthalene (64 mg, 0.50 mmol) in 10.0 ml of THF. The silane solution was cooled to the indicated temperature, and then the 4-(bromomagnesium)anisole solution was added dropwise (1 drop per second). The solution was allowed to stir at the indicated temperature for 1 h, and then at room temperature for 12 h. Progress was monitored by GC analysis of aliquots of the quenched reaction mixture. GC response factors relative to the internal naphthalene standard were determined, and the observed percentages of products were normalized accordingly. The reduced product anisole was identified by comparison of the GC retention time to that of an authentic sample. Diarylated products were identified by GCMS; triarylated products were not observed. Aliquots of the crude, concentrated 4-(trichlorosilyl)anisole were converted to the siloxane by dropwise addition (1 drop per second) of the chlorosilane to EtOH/pyridine at 0 ?C. 85 MgBr Si(OR)3 Si(OR)4, THF Br OCH3 OCH3 OCH3 86 88 + 89 Mg?, I2 (cat) 87 1. SiCl4, THF 2. EtOH, pyr, 0 ?C ORTHFreflux Ar2Si(OR)2 Conditionsa Yield (%)b,c Entry Silane / Equiv T (?C) 88 89 1 Si(OMe)4 / 3.0 25 47 47 2 Si(OMe)4 / 3.0 -30 76 3 3 Si(OMe)4 / 1.5 -30 65 4 4 Si(OEt)4 / 3.0 0 72 23 5 Si(OEt)4 / 3.0 -10 70 23 6 Si(OEt)4 / 3.0 -30 83 3 7 Si(OEt)4 / 1.5 -30 76 2 8 Si(Cl)4 / 3.0d -30 70d 3 a Reactions of p-methoxy magnesium bromide 87 (1.0 equiv) with Si(OR) 4 orSiCl 4 (1.5 to 3.0 equiv) were allowed to stir at the given temperature in THF.The reaction mixture was stirred at the indicated temperature for 1 h, and then at room temperature for 12 h. b GC yields are based on amount of 4-bromoanisole (86). c The remainder of the product was of anisole. d Yield of 4-(triethoxysilyl)anisole (88). The crude, concentrated 4-(trichlorosilyl)anisole was converted to the siloxane 88 by dropwise addition of the chlorosilane to EtOH/pyridine at 0 ?C. Table 2. Optimization of the Synthesis of Arylsiloxanes Using 4-Bromomagnesium Anisole. General Procedure for Synthesis of Siloxanes from Grignard Reagents (Table 3). Unless otherwise indicated, all reactions were performed on a 5 mmol scale. A 10 mL, 3- neck pear-shaped flask was fitted with an addition funnel, a reflux condenser, a rubber septum, and a stirbar. The flask was then charged with freshly washed magnesium turnings (134 mg, 5.50 mmol), flame-dried under vacuum, and back-filled with argon. THF (1.0 mL) was added to the magnesium turnings via syringe. The addition funnel was charged with the aryl halide (5.00 mmol) in 2.0 mL THF. The reaction was initiated by addition of 5-10 drops of the aryl halide solution to the magnesium turnings with stirring, 86 followed by gentle heating. The rest of the aryl halide solution was then added at such a rate that the THF maintained a moderate reflux. Upon final addition, the solution was allowed to stir at room temperature until GC analysis of a quenched aliquot of the reaction mixture indicated complete consumption of aryl halide. The aryl magnesiumhalide solution was then transferred via cannula to a second flame-dried addition funnel, to which was fitted a 50 mL round-bottom flask containing tetraethyl orthosilicate or tetramethyl orthosilicate (15.00 mmol) in 10.0 ml of THF. The silane solution was cooled to ?30 ?C, and then the 4-arylmagnesiumhalide solution was added dropwise (1 drop per second). The solution was allowed to stir at the indicated temperature for 1 h, and then at room temperature for 12 h. The crude reaction mixture was then poured into 50 mL of pentane in a 200 mL separatory funnel. The amber solution was washed with 3 x 25 mL water to remove the excess tetraalkoxysilane, dried over MgSO4, filtered, and concentrated in vacuo. Purification of the residue by bulb-to-bulb distillation yielded the siloxane. 2-(Triethoxysilyl)anisole (Table 3, entry 1). The general procedure for synthesis of siloxanes from Grignard reagents was followed using 2-bromoanisole (935 mg, 623 ?L, 5.00 mmol), magnesium turnings (134 mg, 5.50 mmol), and tetraethyl orthosilicate (3.13 g, 3.35 mL, 15.0 mmol) in THF. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.19 g (88%) of 2-(triethoxysilyl)anisole as a colorless oil. IR (neat) 3067 (m), 2967 (s), 2922 (s, 2891 (s), 2829 (m), 2780 (w), 2762 (w), 1590 (s), 1569 (s), 1459 (s), 1241 (s) cm-1; 1H (NMR) (CDCl3) ? 1.21 (t, J = 7.2, 9H), 3.79 (s, 3H), 3.84 (q, J = 7.2, 6H), 6.83 (d, J = 8.2, 1H), 6.95 (t, J = 7.2, 1H), 7.38 (m, 1H), 7.63 (dd, J = 1.6, 7.2, 1H); 13C (NMR) (CDCl3) ? 18.2, 55.1, 58.7, 109.6, 119.2, 120.5, 132.2, 137.5, 164.3; MS (m/z) 271 (M++1, 42), 225 (100), 195 (21), 181 (36), 139 (25), 119 (21), 91 (24), 77 (14); HRMS for C13H23O4Si calcd 271.1366 (M++1), found 271.1354. Si(OEt)3 OCH3 87 MgBr Si(OR')3 R R THF Si(OR')4 Br R THF Mg? 90 9291 Entry Aryl Siloxane Yield (%)a,b Entry Aryl Siloxane Yield (%)a,b 1 Si(OEt)3 OCH3 88 6 Si(OEt)3 CH3 81 2 Si(OMe)3 OCH3 83 7 Si(OMe)3 CH3 71 3 Si(OEt)3 H3CO 84 8 Si(OEt)3 H3C 75 4 H3CO Si(OEt)3 82 9 H3C Si(OEt)3 86 5 O O Si(OEt)3 74 10 Si(OEt)3 55 a Reactions of arylmagnesium bromide 91 (1.0 equiv) with Si(OEt) 4 or Si(OMe)4 (3.0equiv) were allowed to stir at ?30 ?C in THF. The reaction mixture was stirred at ?30 ?C for 1 h, and then at room temperature for 12 h. b Yields of 92 are after distillation (purity >95%). The remainder of the product was of the reduced (dehalogenated) arene. Table 3. Synthesis of Aryl(trialkoxy)silanes Using Grignard Reagents. 2-(Trimethoxysilyl)anisole (Table 3, entry 2). The general procedure for synthesis of siloxanes from Grignard reagents was followed using 2-bromoanisole (935 mg, 623 ?L, 5.00 mmol), magnesium turnings (134 mg, 5.50 mmol), and tetramethyl orthosilicate (2.28 g, 2.21 mL, 15.0 mmol) in THF. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 948 mg (83%) of 2-(trimethoxysilyl)anisole as a colorless oil. IR (1H (NMR) (CDCl3) ? 3.61 (s, 9H), 3.80 (s, 3H), 6.87 (d, J=8.3, 1H), 6.97 (t, J=7.2, 1H), 7.42 (m, 1H), 7.59 (dd, J=1.7, 7.2, 1H); 13C NMR (CDCl3) ? 50.4, 54.9, 109.4, 117.7, 120.3, 132.2, 137.0, 164.1; MS (m/z) 228 (100), 196 (61) 167 (55), 137 (18), 121 (44), 91 (45), 58 (29); HRMS for C10H16O4Si calcd 228.0818, found 228.0809. Si(OMe)3 OCH3 88 3-(Triethoxysilyl)anisole (Table 3, entry 3). The general procedure for synthesis of siloxanes from Grignard reagents was followed using 3-bromoanisole (935 mg, 633 ?L, 5.00 mmol), magnesium turnings (134 mg, 5.50 mmol), and tetraethyl orthosilicate (3.13 g, 3.35 mL, 15.0 mmol) in THF. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.14 g (84%) of 3-(triethoxysilyl)anisole as a colorless oil. IR (neat) 2975 (s), 2927 (m), 2887 (m), 1572 (m), 1482 (w), 1410 (w), 1391 (w), 1284 (w), 1249 (m), 1234 (m), 1167 (m), 1078 (vs) cm-1; 1H (NMR) (CDCl3) ? 1.25 (t, J = 7.0, 9H), 3.81 (s, 3H), 3.87 (q, J = 7.0, 6H), 6.96- 6.98 (m, 1H), 7.21-7.33 (m, 3H); 13C (NMR) (CDCl3) ? 18.3, 55.1, 58.8, 116.1, 119.8, 127.1, 129.2, 132.4, 159.0; MS (m/z) 270 (100), 256 (12), 255 (70), 226 (28), 225 (64), 211 (23), 197 (13), 182 (11), 181 (41), 169 (37), 168 (13), 167 (29), 163 (10), 154 (16), 153 (27), 149 (22), 147 (61), 139 (20), 137 (14), 136 (55), 135 (95); HRMS for C13H22O4Si calcd 270.1287, found 270.1282. 4-(Triethoxysilyl)anisole (Table 3, entry 4). The general procedure for synthesis of siloxanes from Grignard reagents was followed using 4-bromoanisole (935 mg, 626 ?L, 5.00 mmol), magnesium turnings (134 mg, 5.50 mmol), and tetraethyl orthosilicate (3.13 g, 3.35 mL, 15.0 mmol) in THF. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.11g (82%) of 4-(triethoxysilyl)anisole as a colorless oil. IR (neat) 2975 (s), 2926 (m), 2894 (m), 1597 (s), 1505 (m), 1282 (m), 1249 (m), 1167 (s), 1127 (vs), 1104 (vs), 1080 (vs) cm-1; 1H (NMR) (CDCl3) ? 1.24 (t, J = 7.0, 9 H), 2.34 (s, 3H), 3.86 (q, J = 7.0, 6 H), 6.92 (d, J = 8.6, 2 H), 7.61 (d, J = 8.6, 2 H); 13C (NMR) (CDCl3) ? 18.1, 54.9, 58.6, 113.6, 122.0, 136.4, 161.4; MS (m/z) 270 (52), 255 (53), 225 (38), 211 (25), 181 (28), 169 (27), 149 (22), 147 (100), 135 (32); HRMS for C13H22O4Si calcd 270.1287, found 270.1261. The IR, 1H and 13C NMR data were identical to published spectral data.221 (EtO)3Si OCH3 Si(OEt)3 H3CO 89 4-(Triethoxysilyl)-1,2-(methylenedioxy)benzene (Table 3, entry 5). The general procedure for synthesis of siloxanes from Grignard reagents was followed using 4-bromo-1,2- (methylenedioxy)benzene (1.01 g, 602 ?L, 5.00 mmol), magnesium turnings (134 mg, 5.50 mmol), and tetraethyl orthosilicate (3.13 g, 3.35 mL, 15.0 mmol) in THF. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.00 g (74%) of 4-(triethoxysilyl)-1,2- (methylenedioxy)benzene as a colorless oil. IR (CCl4) 2976 (s), 2926 (m), 2886 (s), 1613 (w), 1503 (w), 1487 (m), 1422 (m), 1237 (m), 1168 (m), 1080 (s), 1045 (m) cm-1; 1H (NMR) (CDCl3) ? 1.24 (t, J = 6.8, 9H), 3.85 (q, J = 6.8, 6H), 5.95 (s, 2H), 6.86 (d, J = 7.6, 1H), 7.12 (s, 1H), 7.19 (d, J = 7.6, 1H); 13C (NMR) (CDCl3) ? 18.1, 59.9, 100.8, 108.8, 114.2, 123.9, 129.6, 147.6, 149.7; MS (m/z) 284 (100), 239 (39), 226 (28), 211 (10), 195 (14), 183 (25), 167 (18), 153 (13), 149 (24), 148 (14), 147 (75), 135 (12); HRMS for C13H20O5Si calcd 284.1080, found 284.1083. 2-(Triethoxysilyl)toluene (Table 3, entry 6). The general procedure for synthesis of siloxanes from Grignard reagents was followed using 2-bromotoluene (855 mg, 601 ?L, 5.00 mmol), magnesium turnings (134 mg, 5.50 mmol), and tetraethyl orthosilicate (3.13 g, 3.35 mL, 15.0 mmol) in THF. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.03 g (81%) of 2-(triethoxysilyl)toluene as a colorless oil. IR (CCl4) 3054, 2971, 2922, 2881, 1442, 1393, 1283, 1162, 1079 cm-1; 1H NMR (CDCl3) ? 1.26 (t, J=7.0, 9H), 2.26 (s, 3H), 3.87 (q, J=7.0, 6H), 7.17 (m, 2H), 7.32 (m, 1H), 7.74 (m, 1H); 13C NMR (CDCl3) ? 18.2, 22.4, 58.5, 124.7, 129.7, 129.9, 130.5, 136.5, 144.5, MS (m/z) 254 (48), 209 (44), 208 (19), 162 (55), 147 (100), 119 (51), 91 (50), 79 (15); HRMS for C13H22O3Si calcd 254.1349, found 254.1338. The IR, 1H and 13C NMR data were identical to published spectral data.120 2-(Trimethoxysilyl)toluene (Table 3, entry 7). The general procedure for synthesis of siloxanes from Grignard reagents was O O Si(OEt)3 Si(OEt)3 CH3 Si(OMe)3 CH3 90 followed using 2-bromotoluene (855 mg, 601 ?L, 5.00 mmol), magnesium turnings (134 mg, 5.50 mmol), and tetramethyl orthosilicate (2.28 g, 2.21 mL, 15.0 mmol) in THF. Bulb- to-bulb distillation (125 ?C, 0.5 torr) afforded 754 mg (71%) of 2-(trimethoxysilyl)toluene as a colorless oil. IR (neat) 3052 (w), 3008 (w), 2942 (s), 2840 (s), 1592 (m), 1471 (m), 1456 (m) cm -1; 1H NMR (CDCl3) ? 2.48 (s, 3H), 3.61 (s, 9H), 7.17 (m, 2H), 7.31 (m, 1H), 7.67 (d, J=7.5, 1H); 13C NMR (CDCl3) ? 22.2, 50.4, 124.7, 128.4, 129.7, 130.6, 136.3, 144.4; MS (m/z) 212 (61), 151 (16), 121 (100), 60 (73); HRMS for C10H16O3Si 212.0869, found 212.0859. 3-(Triethoxysilyl)toluene (Table 3, entry 8). The general procedure for synthesis of siloxanes from Grignard reagents was followed using 3-bromotoluene (855 mg, 607 ?L, 5.00 mmol), magnesium turnings (134 mg, 5.50 mmol), and tetraethyl orthosilicate (3.13 g, 3.35 mL, 15.0 mmol) in THF. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 954 mg (75%) of 3-(triethoxysilyl)toluene as a colorless oil. IR (neat) 2975 (s), 2926 (s), 2885 (s), 1577 (w), 1480 (w), 1442 (m), 1390 (m), 1295 (w), 1225 (w), 1167 (s), 1104 (vs), 1079 (vs) cm-1; 1H (NMR) (CDCl3) ? 1.25 (t, J = 7.2, 9H), 2.36 (s, 3H), 3.87 (q, J = 7.2, 6H), 7.23- 7.29 (m, 2H), 7.46-7.48 (m, 2H); 13C (NMR) (CDCl3) ? 18.4, 21.7, 58.9, 128.0, 130.8, 131.4, 132.0, 135.6, 137.4; MS (m/z) 254 (41), 239 (6), 209 (52), 195 (6), 162 (44), 147 (100), 119 (53), 91 (42), 66 (6); HRMS for C13H22O3Si calcd 254.1338, found 254.1331. 4-(Triethoxysilyl)toluene (Table 3, entry 9). The general procedure for synthesis of siloxanes from Grignard reagents was followed using 4-bromotoluene (855 mg, 638 ?L, 5.00 mmol), magnesium turnings (134 mg, 5.50 mmol), and tetraethyl orthosilicate (3.13 g, 3.35 mL, 15.0 mmol) in THF. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.09 g (86%) of 4-(triethoxysilyl)toluene as a colorless oil. IR (CCl4) 2975 (s) , 2926 (m), 2885 (m), 1167(s), 1124 (vs), 1103 (vs), 1080 (vs) cm -1 ; 1H (NMR) (CDCl3) ? 1.24 (t, J = 7.0, 9 Si(OEt)3 H3C H3C Si(OEt)3 91 H), 2.36 (s, 3 H), 3.86 (q, J = 7.0, 6 H), 7.19 (d, J = 7.9, 2 H), 7.57(d, J = 7.9, 2 H); 13C (NMR) (CDCl3) ? 18.2, 21.5, 58.6, 127.4, 128.6, 134.8, 140.2; MS (m/z) 254 (23), 209 (38), 181 (5), 165 (16), 162 (30), 153 (19), 147 (100), 135 (17); HRMS for C13H22O3Si calcd 254.1363, found 254.1338. The IR, 1H and 13C NMR data were identical to published spectral data.221 Triethoxyphenylsilane (Table 3, entry 10). The general procedure for synthesis of siloxanes from Grignard reagents was followed using bromobenzene (785 mg, 527 ?L, 5.00 mmol), magnesium turnings (134 mg, 5.50 mmol), and tetraethyl orthosilicate (3.13 g, 3.35 mL, 15.0 mmol) in THF. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 661 mg (55%) of triethoxyphenylsilane as a colorless oil. Spectral data is reported above. IR (CCl4) 3140 (w), 2976 (s), 2928 (m), 2941 (m), 1431 (m), 1391 (m), 1129 (s), 1102 (s), 1094 (s), 1080 (s) cm-1; 1H (NMR) (CDCl3) ? 1.25 (t, J = 7.0, 9 H), 3.88 (q, J = 7.0, 6 H), 7.3-7.5 (m, 3 H), 7.6-7.8 (m, 2 H); 13C (NMR) (CDCl3) ? 18.1, 58.7, 127.8, 130.2, 131.1, 134.7; MS (m/z) 240 (16), 195 (38), 181 (13), 162 (28), 147 (100), 139 (33), 135 (33); HRMS for C12H20O3Si calcd 240.1182, found 240.1141. The IR, 1H and 13C NMR data were identical to published spectral data.221 General Procedure for the Optimization of the Synthesis of 4- (Triethoxysilyl)toluene (97) Using 4-Lithioanisole (96) (Table 4) A solution of n-BuLi (1.5 M in pentane, 3.33 mL, 5.00 mmol) was added dropwise (1 drop per second) to a stirring solution of 4-bromotoluene (97) (855 mg,615 ?L, 5.00 mmol) in Et2O (15.0 mL) at room temperature. After 1 h, the solution was cooled to ?78 ?C and added via cannula to a stirring solution of tetraethyl orthosilicate (1.5-3.0 equiv) and the internal standard biphenyl (77 mg, 0.50 mmol) in Et2O (15.0 mL) at ?78 ?C. Progress was monitored by GC analysis of aliquots of the quenched reaction mixture. GC response factors relative to the internal naphthalene standard were determined, and the observed percentages of products were normalized accordingly. The reduced product toluene was identified by Si(OEt)3 92 comparison of the GC retention time to that of an authentic sample. Polyarylated products were identified by GCMS. Li Si(OEt)3 Si(OEt)4 Br CH3 CH3 CH3 95 97 + Ar2Si(OEt)2 + Ar3Si(OEt) 98 99 n-BuLi 96 Conditionsa Yield (%)b,c Entry Equiv Si(OEt)4 Solvent T (?C) 97 98 99 1 1.5 THF 0 7 25 54 2 1.5 THF -30 74 12 1 3 1.5 Et2O -30 77 6 0 4 1.5 THF -78 81 9 0 5 1.5 Et2O -78 82 5 1 6 3.0 Et2O -78 86 3 1 a Reactions of p-tolyl lithium 96 (1.0 equiv) with Si(OEt) 4 (1.5 to 3.0 equiv) wereallowed to stir at the given temperature; the reaction was complete in 1 h. b GC yields are based on amount of 4-bromotoluene (95). c The remainder of the product was toluene. Table 4. Optimization of the Synthesis of Arylsiloxanes Using 4-Lithiotoluene. General Procedure for Synthesis of Siloxanes from Lithium Reagents Using n-BuLi (Table 5). Unless otherwise indicated, all reactions were performed on a 5 mmol scale. A solution of n-BuLi (1.6 M in pentane, 3.1 mL, 5.0 mmol) was added dropwise (1 drop per second) to a stirring solution of the aryl halide (5.00 mmol) in Et2O or THF (7.0 mL) at room temperature. After 1 h, the solution was cooled to ?78 ?C and added via cannula to a stirring solution of tetraethyl orthosilicate (1.56 g, 1.68 mL, 7.50 mmol) in Et2O or THF (7.0 mL) at ?78 ?C. After 1 h, the reaction was quenched with H2O (5 drops) at ?78 ?C and allowed to slowly warm to room temperature. The crude reaction mixture was then extracted with 3 x 50 mL Et2O. The combined organic extracts were dried over MgSO4, filtered, and concentrated in vacuo. The residue was purified by either column chromatography or by bulb-to-bulb distillation. 93 Li Si(OEt)3 R R Si(OEt)4 Br R Et2O -78 ?C n-Bu Li 90 92100 Et2O -78 ?C Entry Aryl Siloxane (92) % Yielda,b Entry Aryl Siloxane (92) % Yielda,b 1 Si(OEt)3 CH3 79 7 H3CO Si(OEt)3 67 2 Si(OEt)3 H3C 71 8 Si(OEt)3MeS 50 3 H3C Si(OEt)3 85 9 O O Si(OEt)3 30c 4 Si(OEt)3 74 10 S Si(OEt)3 50 5 Si(OEt)3 OCH3 60 11 O Si(OEt)3 22d 6 Si(OEt)3 H3CO 66 12 N Si(OEt)3 0 a Reactions of aryl lithium 100 (1.0 equiv) with Si(OEt) 4 (1.5 equiv) were allowed to stir at?78 ?C for 1 h in Et 2O. b Yields of 92 are after distillation or chromatography (purity >95%). c The major product was Ar2Si(OEt)2. d G.C. analysis indicated a 2:1:1 ratio of mono:di:tri arylalkoxysilanes. Table 5. Synthesis of Aryl(trialkoxy)silanes Using Lithium Reagents. 2-(Triethoxysilyl)toluene (Table 5, entry 1). The general procedure for synthesis of siloxanes from lithium reagents was followed using n-BuLi (1.6 M in pentane, 3.1 mL, 5.0 mmol), Si(OEt)3 CH3 94 2-bromotoluene (855 mg, 601 ?L, 5.00 mmol), and tetraethyl orthosilicate (1.56 g, 1.68 mL, 7.50 mmol) in Et2O. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.00 g (79%) of 2-(triethoxysilyl)toluene as a colorless oil. Spectral data is reported above. 3-(Triethoxysilyl)toluene (Table 5, entry 2). The general procedure for synthesis of siloxanes from lithium reagents was followed using n-BuLi (1.6 M in pentane, 3.1 mL, 5.0 mmol), 3-bromotoluene (855 mg, 607 ?L, 5.00 mmol), and tetraethyl orthosilicate (1.56 g, 1.68 mL, 7.50 mmol) in Et2O. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 903 mg (71%) of 3- (triethoxysilyl)toluene as a colorless oil. Spectral data is reported above. 4-(Triethoxysilyl)toluene (Table 5, entry 3). The general procedure for synthesis of siloxanes from lithium reagents was followed using n-BuLi (1.6 M in pentane, 3.1 mL, 5.0 mmol), 4-bromotoluene (855 mg, 638 ?L, 5.00 mmol), and tetraethyl orthosilicate (1.56 g, 1.68 mL, 7.50 mmol) in Et2O. Bulb-to- bulb distillation (125 ?C, 0.5 torr) afforded 1.08 g (81%) of 4-(triethoxysilyl)toluene as a colorless oil. Spectral data is reported above. Triethoxyphenylsilane (Table 5, entry 4). The general procedure for synthesis of siloxanes from lithium reagents was followed using n-BuLi (1.6 M in pentane, 3.1 mL, 5.0 mmol), bromobenzene (785 mg, 527 ?L, 5.00 mmol), and tetraethyl orthosilicate (1.56 g, 1.68 mL, 7.50 mmol) in Et2O. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 889 mg (74%) of triethoxyphenylsilane as a colorless oil. Spectral data is reported above. 2-(Triethoxysilyl)anisole (Table 5, entry 5). The general procedure for synthesis of siloxanes from lithium reagents was followed using n-BuLi (1.6 M in pentane, 3.1 mL, 5.0 mmol), Si(OEt)3 OCH3 Si(OEt)3 H3C H3C Si(OEt)3 Si(OEt)3 95 2-bromoanisole (935 mg, 623 ?L, 5.00 mmol), and tetraethyl orthosilicate (1.56 g, 1.68 mL, 7.50 mmol) in Et2O. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 811 mg (60%) of 2-(triethoxysilyl)anisole as a colorless oil. Spectral data is reported above. 3-(Triethoxysilyl)anisole (Table 5, entry 6). The general procedure for synthesis of siloxanes from lithium reagents was followed using n-BuLi (1.6 M in pentane, 3.1 mL, 5.0 mmol), 3-bromoanisole (935 mg, 633 ?L, 5.00 mmol), and tetraethyl orthosilicate (1.56 g, 1.68 mL, 7.50 mmol) in Et2O. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 892 mg (66%) of 3-(triethoxysilyl)anisole as a colorless oil. Spectral data is reported above. 4-(Triethoxysilyl)anisole (Table 5, entry 7). The general procedure for synthesis of siloxanes from lithium reagents was followed using n-BuLi (1.6 M in pentane, 3.1 mL, 5.0 mmol), 4-bromoanisole (935 mg, 626 ?L, 5.00 mmol), and tetraethyl orthosilicate (1.56 g, 1.68 mL, 7.50 mmol) in Et2O. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 906 mg (67%) of 4-(triethoxysilyl)anisole as a colorless oil. Spectral data is reported above. 4-(Triethoxysilyl)thioanisole (Table 5, entry 8). The following experiment was performed by Correia.228 The general procedure for synthesis of siloxanes from lithium reagents was followed using n-BuLi (1.6 M in pentane, 6.1 mL, 9.8 mmol), 4-bromothioanisole (1.99 g, 9.80 mmol), and tetraethyl orthosilicate (6.13 g, 6.56 mL, 29.4 mmol) in Et2O (20 mL). Extraction and flash chromatography (30 mm, 15 cm, 33% CH2Cl2/hexanes) afforded 1.40 g (50%) of 4-(triethoxysilyl)thioanisole as a colorless oil. TLC Rf = 0.25 (33% CH2Cl2/hexanes); IR (CCl4) 2976 (m), 2925 (m), 1585 (m), 1487 (m), 1440 (m), 1103 (s), 1080 (s) cm-1; 1H NMR (CDCl3) ? 1.24 (t, J = 6.8, 9H), 2.48 (s, 3H), 3.86 (q, J = 6.8, 6H), 7.25 (d, J = 7.9, 2H), 7.58 (d, J = 7.9, 2H); 13C NMR (CDCl3) ? 15.0, 18.2, 58.7, 125.3, Si(OEt)3 H3CO (EtO)3Si OCH3 Si(OEt)3MeS 96 126.7, 135.1, 141.4; MS (m/z) 287 (M++1, 22), 286 (100), 241 (19), 227 (17), 195 (19), 147 (49), 124 (17), 119 (17); HRMS for C13H22O3SSi calcd 286.1059, found 286.1063. 4-(Triethoxysilyl)-1,2-(methylenedioxy)benzene (Table 5, entry 9). The general procedure for synthesis of siloxanes from lithium reagents was followed using n-BuLi (1.6 M in pentane, 3.1 mL, 5.0 mmol), 4-bromo-1,2-(methylenedioxy)benzene (1.01 g, 602 ?L, 5.00 mmol), and tetraethyl orthosilicate (1.56 g, 1.68 mL, 7.50 mmol) in Et2O. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 427 mg (30%) of 4-(triethoxysilyl)-1,2-(methylenedioxy)benzene as a colorless oil. Spectral data is reported above. 3-(Triethoxysilyl)thiophene (Table 5, entry 10). The following experiment was performed by Correia.228 The general procedure for synthesis of siloxanes from lithium reagents was followed using n-BuLi (1.6 M in pentane, 7.50 mL, 12.0 mmol), 3-bromothiophene (1.96 g, 1.12 mL, 12.0 mmol), and tetraethyl orthosilicate (7.50 g, 8.05 mL, 36.0 mmol) in Et2O (20 mL). Extraction and flash chromatography (30 mm, 15 cm, 17% CH2Cl2/hexanes) gave 1.48 g (50%) of 3-(triethoxysilyl)thiophene as a colorless oil. TLC Rf = 0.25 (17% CH2Cl2/hexanes); IR (CCl4) 2975 (m), 2926 (m), 1553 (m), 1542 (m), 1103 (s), 1081 (s) cm-1; 1H NMR (CDCl3) ? 1.24 (t, J = 7.2, 9H), 3.87 (q, J = 7.2, 6H), 7.29 (d, J = 4.8, 1H), 7.40 (dd, J = 4.8, 2.0, 1H), 7.74 (d, J = 2.0, 1H); 13C NMR (CDCl3) ? 18.2, 58.7, 125.7, 131.7, 131.8, 135.5; MS (m/z) 247 (M++1, 17), 246 (100), 202 (30), 201 (46), 158 (51), 145 (30), 135 (73); HRMS for C10H18O3SSi calcd 246.0746, found 246.0743. 2-furyltriethoxysilane (Table 5, entry 11). The following experiment was performed by Ahn.228 A solution of t-BuLi (1.50 M in pentane, 16.00 mL, 24.00 mmol) was added dropwise (1 drop per second) to a O O Si(OEt)3 O Si(OEt)3 S Si(OEt)3 97 stirring solution of the furan (1.36 g, 1.46 mL, 20.0 mmol) and TMEDA (2.32 g, 3.02 mL, 20.0 mmol) in Et2O (40.0 mL) at 0 ?C. After 2 h, the solution was cooled to ?78 ?C and tetraethyl orthosilicate (6.25 g, 6.69 mL, 30.0 mmol) was added dropwise (1 drop per second). After 1 h, the reaction was quenched with H2O (1 mL) at ?78 ?C and allowed to slowly warm to room temperature. The crude reaction mixture was then extracted with 2 x 200 mL Et2O. The combined organic extracts were dried over MgSO4, filtered, and concentrated in vacuo. Purification of the residue by flash chromatography (30 mm, 15 cm, 10% EtOAc/hexanes) gave 1.01 g (22%) of 2-furyltriethoxysilane as a colorless oil. TLC Rf = 0.71 (10% EtOAc/hexane); IR (CCl4) 2974 (s), 2926 (s), 2891 (s), 1455 (m), 1390 (m), 1169 (s), 1100 (s), 1079 (s), 965 (m); 1H (NMR) (CDCl3) ? 1.23 (t, J = 7.0, 9H), 3.87 (q, J = 7.0, 6H), 6.39 (dd, J = 1.3, 3.3, 1H), 6.87 (d, J = 3.3, 1H), 7.65 (d, J = 1.3, 1H); 13C (NMR) (CDCl 3) ? 18.1, 59.0, 109.3, 123.2, 147.4, 151.3; MS (m/z) 231 (M ++1, 5), 230 (M+, 28), 215 (100), 203 (14), 185 (28), 147 (80), 119 (52), 113 (55), 79 (47), 63 (27); HRMS for C10H18O4Si calcd 230.0979, found 230.0974. 5-Triethoxysilyl-2-methoxypyridine (102). The following experiment was performed by Handy.111 The general procedure for synthesis of siloxanes from lithium reagents was followed using n-BuLi (1.4 M in pentane, 6.5 mL, 8.8 mmol), 5-bromo-2-methoxypyridine (101) (1.66 g, 1.14 mL, 8.83 mmol), and tetraethyl orthosilicate (2.76 g, 2.95 mL, 13.3 mmol) in Et2O (20 mL). Extraction and flash chromatography (9:1 hexanes/EtOAc) afforded 527 mg (22%) of 102 as a colorless oil. TLC Rf =0.28 (9:1 hexanes/EtOAc); IR (CCl4) 2976 (s), 2927 (m), 2887 (m), 1589 (s), 1491 (w), 1442 (m), 1390 (w), 1356 (m), 1286 (s), 1082 (vs) cm-1; 1H (NMR) (CD3CN) ? 1.20 (t, J = 7.0, 9H), 3.83 (q, J = 7.0, 6H), 3.87 (s, 3H), 6.73-6.75 (m, 1H), 7.76-7.78 (m, 1H), 8.32-8.35 (m, 1H); 13C (NMR) (CD3CN) ? 18.6, 53.9, 59.5, 111.5, 130.8, 118.2, 145.5, 154.3, 166.5; MS (m/z) 272 (100), 240 (13), 226 (31), 170 (11), 163 (11), 136 (15), 119 (6), 91 (5), 79 (14); HRMS for C12H22ONSi calcd 272.1318, found 272.1311. N Si(OEt)3 MeO 98 5-(Triethoxysilyl)indole (110). The following experiment was performed by Handy.111 To a solution of KH (35% dispersion in mineral oil, 292 mg, 2.55 mmol) in THF (5.0 mL) was added dropwise (1 drop per second) a solution of 5-bromoindole (498 mg, 2.54 mmol) in THF (5.0 mL). The reaction mixture was stirred for 15 min then cooled to ?78 ?C. A solution of t-BuLi (1.7 M in pentane, 3.0 mL, 5.0 mmol) was then added via cannula. A white precipitate immediately formed. The mixture was stirred for 10 min, followed by dropwise (1 drop per second) addition of a solution of tetraethyl orthosilicate (1.06 g, 1.13 mL, 5.08 mmol) in THF (2 mL). The reaction mixture was stirred for 30 min at -78 ?C, and then allowed to slowly warm to room temperature. The reaction mixture was poured into 10 mL ice water, then extracted with ether (3 ? 15 mL). The combined organic extracts were dried over MgSO4, then concentrated in vacuo. Column chromatography (4:1 hexanes/EtOAc) afforded 497 mg (70%) of 110 as a pale yellow oil; TLC Rf =0.32 (4:1 hexanes/EtOAc); IR (CCl4) 3489 (s), 2976 (s), 2926 (m), 1885 (m) cm-1; 1H NMR (CDCl3) ? 8.38 (br s, 1H), 8.03 (s, 1H), 7.48 (d, J=8.1 Hz, 1H), 7.39 (d, J=8.1 Hz, 1H), 7.16 (t, J=2.6 Hz, 1H), 6.56 (s, 1H), 3.89 (q, J=7.0, 6H), 1.26 (t, J=7.0, 9H); 13C NMR (CDCl3) ? 137.2, 128.5, 127.7, 127.6, 124.2, 120.0, 110.9, 102.7. 58.6, 18.2; MS (m/z) 280 (70), 234 (42), 206 (14), 190 (21), 163 (100), 144 (78), 117 (58); HRMS for C14H21NO3Si calcd 280.1369, found 280.1381. 5-(Triethoxysilyl)-1-methyl-indole (111). The following experiment was performed by Correia.228 A solution of t-BuLi (1.50 M in pentane, 3.76 mL, 5.64 mmol) was added dropwise (1 drop per second) to a stirring solution of 5-bromo-1-methylindole (987 mg, 4.70 mmol) in Et2O (10.0 mL) at -78 ?C. After 15 min, the solution was added dropwise (1 drop per second) via cannula to a solution of tetraethyl orthosilicate (6.25 g, 6.69 mL, 30.0 mmol) in 5 mL of Et2O cooled to ?78 ?C. After 1 h, the reaction was allowed to slowly warm to room temperature. The reaction mixture was quenched by the addition of 50 mL of water. The N Me Si(OEt)3 NH Si(OEt)3 99 crude reaction mixture was then extracted with 4 x 50 mL Et2O. The combined organic extracts were dried over Na2SO4, filtered, and concentrated in vacuo. Purification of the residue by flash chromatography (30 mm, 15 cm, 33% CH2Cl2/hexane) gave 840 mg (60%) of 111 as a yellow oil. TLC Rf = 0.37 (33% CH2Cl2/hexane); IR (CCl4) 2975 (s), 2923 (m), 2882 (m), 1611 (m), 1556 (s), 1518 (m), 1480 (m), 1104 (s), 1080 (s) cm-1; 1H NMR (CDCl3) ? 1.25 (t, J = 7.2, 9H), 3.79 (s, 3H), 3.88 (q, J = 7.2, 6H), 6.52 (d, J = 2.8, 1H), 7.05 (d, J = 2.8, 1H), 7.36 (d, J = 7.9, 1H), 7.52 (d, J = 7.9, 1H), 8.00 (s, 1H); 13C NMR (CDCl3) ? 18.2, 32.6, 58.5, 101.3, 108.9, 119.4, 127.3, 128.2, 128.6, 128.8, 137.9; MS (m/z) 294 (M++1, 29), 293 (M+, 100), 278 (10), 131 (23). HRMS for C15H23NO3Si calcd 293.1447, found 293.1439. General Procedure for the Optimization of the Silylation of 4-Bromoanisole (86) (Table 6). All siloxanes were synthesized using modifications of the procedure reported by Masuda.221 4-Bromoanisole (86)(187 mg, 125 ?L, 1.00 mmol, 1.0 equiv), and base (3.00 mmol, 3.0 equiv) were added to a stirring solution of Pd(dba)2 (29 mg, 0.05 mmol, 5 mol %), the phosphine (0.10 mmol, 10 mol%), and the internal standard naphthalene (128 mg, 1.0 mmol, 1.0 equiv) in 4 mL of solvent under an atmosphere of argon. Triethoxysilane (246 mg, 277 ?L, 1.50 mmol, 1.5 equiv) was added, causing bubbling, formation of yellow foam, and darkening of the reaction mixture. The reaction mixture was allowed to stir at room temperature for 12 h. Progress was monitored by GC analysis of aliquots of the quenched reaction mixture. GC response factors relative to the internal naphthalene standard were determined, and the observed percentages of products were normalized accordingly. The reduced product anisole 119 was identified by comparison of the GC retention time to that of an authentic sample. 100 Br OCH3 H Si(OEt)3 Si(OEt)3 OCH3 H OCH3 + + Pd(dba)2 ligand base solvent 86 118 119 Conditionsa Yield (%)b,c Ligand Base Solvent Entry (10 mol %) (3 mmol) (4 mL) 118 119 1 P(o-tol)3 i-Pr2NEt NMP 21 79 2 P(o-tol)3 i-Pr2NEt DMF 15 70 3 none i-Pr2NEt DMF 14 86 4 PPh3 i-Pr2NEt DMF 0 5 5 dppf i-Pr2NEt NMP 0 8 6 P(t-Bu)3 i-Pr2NEt NMP 59 41 7 P(t-Bu)3 i-Pr2NEt DMF 45 55 8 P[(2,4,6-OMe)Ph]3 i-Pr2NEt NMP 8 92 9 P(cy)2(o-biphenyl) i-Pr2NEt NMP 17 83 10 P(t-Bu)2(o-biphenyl) i-Pr2NEt NMP 75 25 11 P(t-Bu)2(o-biphenyl) i-Pr2NEt DMF 36 64 12 P(t-Bu)2(o-biphenyl) i-Pr2NEt THF 6 54 13 P(t-Bu)2(o-biphenyl) i-Pr2NEt CH3CN 6 93 14 P(t-Bu)2(o-biphenyl) i-Pr2NEt toluene 3 19 15 P(t-Bu)2(o-biphenyl) none NMP 0 4 16 P(t-Bu)2(o-biphenyl) Et3N NMP 55 45 17 P(t-Bu)2(o-biphenyl) pyridine NMP 0 5 18 P(t-Bu)2(o-biphenyl) 2,6-lutidine NMP 32 26 19 P(t-Bu)2(o-biphenyl) DBU NMP 0 57 20 P(t-Bu)2(o-biphenyl) KOAc NMP 22 78 21 P(t-Bu)2(o-biphenyl) Cs2CO3 NMP 3 97 a Reactions of 4-bromoanisole (86) (1.0 equiv) with H-Si(OEt) 3 (1.5 equiv) wereallowed to stir at room temperature for 12 h in 4 mL of solvent by using Pd(dba) 2 (5 mol%), phosphine (10 mol %), and base (3.0 equiv). b GC yields are based on amount of 4-bromoanisole. c Remaining percentage was unreacted starting material. Table 6. Optimization of the Silylation of 4-Bromoanisole. General Procedure for the Silylation of Aryl Halides (Tables 7-10) All siloxanes were synthesized using a modification of the procedure reported by Masuda.221 The indicated aryl halide (5.00 mmol, 1.0 equiv), and i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol, 3.0 equiv) were added to a stirring solution of Pd(dba)2 (86 mg, 0.15 mmol, 3 mol %), and 101 P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol, 6 mol%) in 20 mL of NMP under an atmosphere of argon. Triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol, 1.5 equiv) was added, causing bubbling, formation of yellow foam, and darkening of the reaction mixture. Unless otherwise noted, the reaction was performed at the given temperature until GC analysis indicated that the starting material had been consumed. The reaction mixture was extracted with 5 x 100 mL pentane, without the addition of water or aqueous solutions, which led to intractable emulsions and potential polymerization; note that NMP and pentane are immiscible. The combined pentane extracts were washed with 3 x 100 mL water to remove NMP, dried over MgSO4, filtered, and concentrated in vacuo. Purification of the residue by bulb-to-bulb distillation yielded the siloxane. 4-(Triethoxysilyl)anisole (Table 7, entry 1). The general procedure for silylation was followed using 4-bromoanisole (935 mg, 626 ?L, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 919 mg (68 %) of 4-(triethoxysilyl)anisole as a colorless oil. Spectral data is reported above. 4-(Triethoxysilyl)toluene (Table 7, entry 2). The general procedure for silylation was followed using 4-bromotoluene (855 mg, 638 ?L, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 547 mg (43%) of 4-(triethoxysilyl)toluene as a colorless oil. Spectral data is reported above. (EtO)3Si OCH3 (EtO)3Si CH3 102 Br H Si(OEt)3 Si(OEt)3 H R R R + +Pd(dba)2 P(t-Bu)2(o-biphenyl)(120) i-Pr2NEt NMP 90 92 121 Entry Aryl Bromide (90) % Yielda-c (92) Entry Aryl Bromide (90) % Yielda-c (92) 1 Br OCH3 68 6 Br OAc 30 2 Br CH3 43 7 Br OH 28d,e 3 O O Br 44 8 Br 5 4 Br NHAc 33 9 Br O 0 5 Br NH2 38d,e 10 Br Cl 0 a Reactions of arylbromide (90) (1.0 equiv) with H-Si(OEt) 3 (1.5 equiv) wereallowed to stir at 60 ?C for 12 h in NMP by using 3 mol % Pd(dba) 2,(t-Bu) 2P(o-biphenyl)(120) (6 mol %), and i-Pr2NEt (3 equiv). b Isolated yield of purified (>95%) product. c Unless otherwise indicated, remaining percentage was reduced starting material (121). d Reaction allowed to stir at room temperature. e Reaction stopped at 2 h. Table 7. Silylation of Aryl Bromides. 4-(Triethoxysilyl)-1,2-(methylenedioxy)benzene (Table 7, entry 3). The general procedure for silylation was followed using 4-(triethoxysilyl)-1,2-(methylenedioxy)benzene (1.01 g, 602 ?L, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 626 mg (44%) of O O Si(OEt)3 103 4-(triethoxysilyl)-1,2-(methylenedioxy)benzene as a colorless oil. Spectral data is reported above. 4-(Triethoxysilyl)acetanilide (Table 7, entry 4). The general procedure for silylation was followed using 4-bromoacetanilide (1.07 g, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 491 mg (33%) 4-(triethoxysilyl)acetanilide as a colorless oil. IR (neat) 3307 (br), 2975 (s), 1670 (s), 1593 (s), 1524 (s), 1392 (s), 1318 (s), 1291 (s), 1166 (s), 1072 (vs) cm-1; 1H (NMR) (CDCl3) ? 1.24 (t, J = 6.8, 9 H), 2.18 (s, 3 H), 3.85 (q, J = 6.8, 6 H), 7.52 (d, J = 8.3, 2 H), 7.64 (d, J = 8.3, 2 H); 13C (NMR) (CDCl3) ? 18.2, 24.5, 58.7, 119.1, 128.6, 135.7, 140.1, 169.0; MS (m/z) 297 (10), 252 (68), 205 (100), 176 (31), 162 (15), 147 (48); HRMS for C14H23O4NSi calcd 297.1396, found 297.1399. The IR, 1H and 13C NMR data were identical to published spectral data.221 4-(Triethoxysilyl)aniline (Table 7, entry 5). The general procedure for silylation was followed using 4-bromoaniline (860 mg, 573 ?L, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was stirred at 25 ?C for 2 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 485 mg (38%) of 4-(triethoxysilyl)aniline as a colorless oil. IR (neat) 3469 (m), 3372 (s), 3223 (w), 2974 (s), 2926 (s), 2885 (s), 1624 (s), 1601 (s), 1509 (m), 1390 (m), 1295 (m), 1166 (s), 1074 (vs) cm-1; 1H (NMR) (CDCl3) ? 1.23 (t, J = 7.0, 9H), 3.79 (s, 2H), 3.84 (q, J = 7.0, 6H), 6.67 (d, J = 8.3, 2H), 7.46 (d, J = 8.3, 2H); 13C (NMR) (CDCl3) ? 18.3, 28.6, 114.4, 118.4, (EtO)3Si NHAc (EtO)3Si NH2 104 (EtO)3Si OAc 136.3, 148.6; MS (m/z) 255 (100), 210 (48), 153 (11), 147 (34), 136 (9), 135 (3); HRMS for C12H21O3NSi calcd 255.1291, found 255.1292. 4-Acetoxy(triethoxysilyl)benzene (Table 7, entry 6). The general procedure for silylation was followed using 4-acetoxybromobenzene (1.01 g, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 448 mg (30%) of 4-acetoxy(triethoxysilyl)benzene as a colorless oil following purification. IR (neat) 2976 (s), 2887 (s), 1767 (s), 1593 (m), 1499 (w), 1392 (m) cm-1; 1H (NMR) (CDCl3) ? 1.22 (t, J = 6.8, 9 H), 2.28 (s, 3 H), 3.84 (q, J = 6.8, 6 H), 7.09 (d, J = 8.3, 2 H), 7.67 (d, J = 8.3, 2 H); 13C (NMR) (CDCl3) ? 18.1, 21.0, 58.7, 121.0, 128.6, 136.1, 152.5, 169.0; MS (m/z) 298 (3), 256 (100), 210 (51), 183 (17), 155 (27), 147 (62); HRMS for C14H22O5Si calcd 298.1237, found 298.1238. The IR, 1H and 13C NMR data were identical to published spectral data.221 4-(Triethoxysilyl)phenol (Table 7, entry 7). The general procedure for silylation was followed using 4-bromophenol (865 mg, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was stirred at 25 ?C for 2 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 359 mg (28%) of 4-(triethoxysilyl)phenol as a colorless oil following purification. IR (CCl4) 3606 (s), 3356 (vs), 2976 (s), 2926 (s), 2885 (s), 1602 (m), 1582 (w), 1506 (m), 1390 (w), 1261 (w), 1167 (m), 1125 (s), 1107 (s), 1080 (s) cm-1; 1H (NMR) (CDCl3) ? 1.24 (t, J = 7.2, 9H), 3.87 (q, J = 7.2, 6H), 6.23 (s, 1H), 6.85 (d, J = 8.3, 2H), 7.55 (d, J = 8.3, 2H); 13C (NMR) (CDCl3) ? 18.3, 59.0, 115.4, 121.2, 137.0, 158.1; MS (m/z) 256 (100), 241 (13), 220 (12), (EtO)3Si OH 105 211 (43), 210 (40), 183 (13), 167 (11), 163 (17), 155 (14), 147 (55), 119 (18); HRMS for C12H20O4Si calcd 256.1131, found 256.1135. Triethoxyphenylsilane (Table 7, entry 8). The general procedure for silylation was followed using bromobenzene (785 mg, 527 ?L, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to- bulb distillation (125 ?C, 0.5 torr) afforded 60 mg (5%) of triethoxyphenylsilane as a colorless oil following purification. Spectral data is reported above. H Si(OEt)3Ar-Br Ar-Si(OEt)3 Ar-H+ Pd(dba)2 P(t-Bu)2(o-biphenyl)(120) i-Pr2NEt NMP122 123 124 + Entry Heteroaryl Bromide (122) % Yielda-c (123) Entry Heteroaryl Bromide (122) % Yielda-c (123) 1 S Br 43 4 N Br 0 2 N Br 0d 5 N Br 0 3 N Br 0 6 N Br MeO 57 a Reactions of heteroarylbromide (122) (1.0 equiv) with H-Si(OEt) 3 (1.5 equiv) were allowed to stir at 60 ?C for 12 h in NMP by using 3 mol % Pd(dba)2, (t-Bu)2P(o-biphenyl)(120) (6 mol %), and i-Pr2NEt (3 equiv). b Isolated yield of purified (>95%) product. c Unless otherwise noted, the remaining percentage was reduced starting material (124). d Trace amounts of 2,2?-dipyridyl were observed by G.C. analysis. Table 8. Silylation of Heteroaryl Bromides. (EtO)3Si 106 2-(Triethoxysilyl)thiophene (Table 8, entry 1). The general procedure for silylation was followed using 2-bromothiophene (815 mg, 484 ?L, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 530 mg (43%) of 2-(triethoxysilyl)thiophene as a colorless oil. IR (neat) 2976 (s), 2888 (s), 1695 (w), 1501 (w), 1442 (s), 1407 (s), 1391(m), 1296 (m), 1216 (m), 1167 (s) cm-1; 1H (NMR) (CDCl3) ?1.25 (t, J = 7.0, 9 H), 3.90 (q, J = 7.0, 6 H), 7.22 (dd, J = 3.3, 4.5, 1 H), 7.48 (d, J = 3.3, 1 H), 7.66 (d, J = 4.5, 1 H); 13C (NMR) (CDCl3) ? 19.4, 60.3, 129.4, 130.6, 133.2, 138.1; MS (m/z) 246 (52), 213 (6), 201 (34), 187 (7), 167 (7), 158 (13), 147 (100), 135 (34); HRMS for C10H18O3SiS calcd 246.0765, found 246.0746. The IR, 1H and 13C NMR data were identical to published spectral data.221 5-(Triethoxysilyl)-2-methoxypyridine (102) (Table 8, entry 6). The general procedure for silylation was followed using 5-bromo-2-methoxypyridine (940 mg, 647 ?L, 5.00 mmol), i- Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 773 mg (57%) of 5-(triethoxysilyl)-2-methoxypyridine as a colorless oil following purification. Spectral data is reported above. N Si(OEt)3 H3CO S Si(OEt)3 107 Br H Si(OEt)3 Si(OEt)3 H R R R + +Pd(dba)2 P(t-Bu)2(o-biphenyl)(120) i-Pr2NEt NMP 90 92 121 Entry Aryl Bromide (90) % Yield (92)a-c Entry Aryl Bromide (90) % Yield (92)a-c 1 Br OCH3 68 4 Br CH3 43 2 Br OCH3 25 5 Br CH3 10 3 Br H3CO 0 6 Br H3C 0 a Reactions of arylbromide (90) (1.0 equiv) with H-Si(OEt) 3 (1.5 equiv) were allowed to stir at 60 ?C for 12 h in NMP using 3 mol % Pd(dba)2, (t-Bu)2P(o-biphenyl) (120) (6 mol %), and i-Pr2NEt (3 equiv). b Yields are isolated yields of purified (>95%) product. c Remaining yield was of reduced starting material (121). Table 9. Silylation of Aryl Bromides: Effect of Substituent Position. 3-(Triethoxysilyl)anisole (Table 9, entry 2). The general procedure for silylation was followed using 3-bromoanisole (935 mg, 633 ?L, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 338 mg (25%) of 3-(triethoxysilyl)anisole as a colorless oil. Spectral data is reported above. 3-(Triethoxysilyl)toluene (Table 3, entry 5). The general procedure for silylation was followed using 3-bromotoluene (855 mg, 607 ?L, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (EtO)3Si OCH3 (EtO)3Si CH3 108 (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 127 mg (10%) of 3-(triethoxysilyl)toluene as a colorless oil. Spectral data is reported above. H Si(OEt)3Ar-I Ar-Si(OEt)3 Ar-H+ Pd(dba)2 P(t-Bu)2(o-biphenyl)(120) i-Pr2NEt NMP128 123 124 + Entry Aryl Iodide (128) % Yield (123)a-c Entry Aryl Iodide (128) % Yield (123)a-c 1 I OCH3 86 10 I NHAc 68e 2 I OCH3 52 11 I NH2 77e 3 I H3CO 10 12 I OAc 75e 4 I CH3 80 13 I OH 70d 6 I CH3 65 14 I O 24 7 I H3C 0d 15 N I 0d,g 8 I Br 68e,f 16 N I 10 9 I Cl 75e 17 S I 92 a Reactions of aryliodide (128) (1.0 equiv) with H-Si(OEt) 3 (1.5 equiv) were allowed to stirat 60 ?C for 12 h in NMP using Pd(dba) 2 (3 mol %), (t-Bu)2P(o-biphenyl) (120) (6 mol %),and i-Pr 2NEt (3 equiv). b Yields shown in parentheses are GC yields; all other yields are isolated yields of purified (>95%) product. c Unless noted otherwise, the remaining percentage was of reduced starting material (124). d Yield confirmed by GCMS. e Reaction allowed to stir at room temperature. f Reaction stopped at 2 h. g Dipyridyl was isolated in 15% yield. Table 10. Palladium-Catalyzed Silylation of Aryl Iodides. 109 4-(Triethoxysilyl)anisole (Table 10, entry 1). The general procedure for silylation was followed using 4-iodoanisole (1.17 g, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.16 g (86%) of 4-(triethoxysilyl)anisole as a colorless oil. Spectral data is reported above. 3-(Triethoxysilyl)anisole (Table 10, entry 2). The general procedure for silylation was followed using 3-iodoanisole (1.17 g, 596 ?L, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 703 mg (52%) of 3-(triethoxysilyl)anisole as a colorless oil. Spectral data is reported above. 2-(Triethoxysilyl)anisole (Table 10, entry 3). The general procedure for silylation was followed using 2-iodoanisole (1.17 g, 650 ?L, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb- to-bulb distillation (125 ?C, 0.5 torr) afforded 135 mg (10%) of 2-(triethoxysilyl)anisole as a colorless oil. Spectral data is reported above. 4-(Triethoxysilyl)toluene (Table 10, entry 4). The general procedure for silylation was followed using 4-iodotoluene (1.09 g, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. (EtO)3Si H3CO (EtO)3Si OCH3 (EtO)3Si OCH3 (EtO)3Si CH3 110 Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.02 g (80%) of 4-(triethoxysilyl)toluene as a colorless oil. Spectral data is reported above. 3-(Triethoxysilyl)toluene (Table 10, entry 6). The general procedure for silylation was followed using 3-iodotoluene (1.09 g, 642 ?L, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 827 mg (65%) of 3-(triethoxysilyl)toluene as a colorless oil. Spectral data is reported above. 4-Bromo(triethoxysilyl)benzene (Table 10, entry 8). The general procedure for silylation was followed using 1-bromo-4-iodobenzene (1.42 g, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was stirred at 25 ?C for 2 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.09 g (68%) of 4-bromo(triethoxysilyl)benzene as a colorless oil. IR (neat) 2975 (s), 2926 (s), 2887 (s), 1575 (m), 1481 (m), 1442 (m), 1390 (m), 1379 (m), 1295 (w), 1165 (s), 1073 (vs) cm-1; 1H (NMR) (CDCl3) ? 1.24 (t, J = 7.2, 9H), 3.86 (q, J = 7.2, 6H), 7.50- 7.55 (m, 4H); 13C (NMR) (CDCl3) ? 18.4, 59.0, 125.5, 130.1, 132.3, 136.6; MS (m/z) 320 (5), 318 (5), 275 (26), 273 (22), 239 (16), 231 (11), 219 (11), 217 (12), 201 (10), 195 (33), 163 (12), 162 (18), 149 (11), 148 (21), 147 (100), 137 (10), 135 (32); HRMS for C12H19O379BrSi calcd 318.0287, found 318.0295. 4-Chloro(triethoxysilyl)benzene (Table 10, entry 9). The general procedure for silylation was followed using 1-chloro-4-iodobenzene (1.19 g, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 (EtO)3Si Br (EtO)3Si Cl (EtO)3Si CH3 111 (EtO)3Si OAc mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was stirred at 25 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.03 g (75%) of 4-chloro(triethoxysilyl)benzene as a colorless oil. IR (neat) 2975, 2888, 2360, 1699, 1583, 1557, 1485, 1443, 1389, 1296, 1261 cm-1; 1H (NMR) (CDCl3) ? 1.17 (t, J = 7.0, 9 H), 3.79 (q, J = 7.0, 6 H), 7.28 (d, J = 8.5, 2 H), 7.54 (d, J = 8.5, 2 H); 13C (NMR) (CDCl3) ? 18.1, 58.8, 128.1, 129.5, 136.1, 136.7; MS (m/z) 274 (3), 239 (27), 229 (44), 185 (15), 173 (25), 162 (31), 147 (100), 135 (9); HRMS for C12H19O3ClSi calcd 274.0792, found 274.0792. The IR, 1H and 13C NMR data were identical to published spectral data.221 4-(Triethoxysilyl)acetanilide (Table 10, entry 10). The general procedure for silylation was followed using 4-iodoacetanilide (1.31 g, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was stirred at 25 ?C for 12h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.01 g (68%) of 4-(triethoxysilyl)acetanilide as a colorless oil. Spectral data is reported above. 4-(Triethoxysilyl)aniline (Table 10, entry 11). The general procedure for silylation was followed using 4-iodoaniline (1.10 g, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was stirred at 25 ?C for 12 h. Bulb- to-bulb distillation (125 ?C, 0.5 torr) afforded 983 mg (77%) of 4-(triethoxysilyl)aniline as a colorless oil. Spectral data is reported above. 4-Acetoxy(triethoxysilyl)benzene (Table 10, entry 12). The general procedure for silylation was followed using (EtO)3Si NHAc (EtO)3Si NH2 112 4-acetoxyiodobenzene (1.31 g, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was stirred at 25 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.12 g (75%) of 4-acetoxy(triethoxysilyl) benzene as a colorless oil. Spectral data is reported above. 4-(Triethoxysilyl)phenol (Table 10, entry 13). The general procedure for silylation was followed using 4-iodophenol (1.10 g, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was stirred at 25 ?C for 12 h. Bulb- to-bulb distillation (125 ?C, 0.5 torr) afforded 897 mg (70%) of 4-(triethoxysilyl)phenol as a colorless oil. Spectral data is reported above. 4-(Triethoxysilyl)acetophenone (Table 10, entry 14). The general procedure for silylation was followed using 4-iodoacetophenone (1.23 g, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 339 mg (24%) of 4-(triethoxysilyl)phenol as a colorless oil. IR (neat) 2975 (s), 2927 (s), 2890 (s), 1728 (m), 1689 (s), 1497 (w), 1443 (m), 1390 (m), 1360 (m), 1263 (s), 1167 (s), 1079 (vs) cm-1; 1H (NMR) (CDCl3) ? 1.26 (t, J = 7.0, 9H), 2.61 (s, 3H), 3.89 (q, J = 7.0, 6H), 7.79 (d, J = 8.0, 2H), 7.95 (d, J = 8.0, 2H); 13C (NMR) (CDCl3) ? 18.3, 26.8, 59.0, 127.4, 135.2, 137.4, 138.4, 198.5; MS (m/z) 282 (8), 268 (22), 267 (100), 239 (15), 238 (67), 237 (57), 223 (16), 209 (18), 193 (13), 181 (20), 165 (19), 163 (24), 147 (16), 138 (12), 135 (9); HRMS for C14H22O4Si calcd 282.1287, found 282.1288. (EtO)3Si O (EtO)3Si OH 113 2-(Triethoxysilyl)pyridine (Table 10, entry 15). The general procedure for silylation was followed using 2-iodopyridine (532 ?L, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. By GCMS of the crude reaction mixture, the reaction yielded pyridine as the major product, and none of the desired silylated product. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 50 mg (15%) of 2,2'-dipyridyl (130) following purification. For 2,2?-dipyridyl: IR (neat) 3069 (m), 3053 (m), 3011 (w), 1584 (s), 1562 (s), 1467 (s), 1421 (s), 1255 (m), 1147 (m) cm-1; 1H (NMR) (CDCl3) ? 7.12 (t, J = 7.8, 2 H), 7.66 (t, J = 7.8, 2 H), 8.50 (d, J = 7.8, 2 H), 8.59 (d, J = 7.8, 2 H); 13C (NMR) (CDCl3) ? 121.0, 123.7, 136.8, 149.0, 156.0. The IR, 1H and 13C NMR data were identical to an authentic sample. 3-(Triethoxysilyl)pyridine (Table 10, entry 16). The general procedure for silylation was followed using 3-iodopyridine (1.03 g, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb-to- bulb distillation (125 ?C, 0.5 torr) afforded 121 mg (10%) of 3-(triethoxysilyl)pyridine as a colorless oil. 1H (NMR) (CDCl3) ? 1.26 (t, J = 6.8, 9H), 3.90 (q, J = 6.8, 6H), 7.2-7.3 (m, 1H), 7.95 (d, J = 5.9, 1H), 8.65 (d, J = 4.9, 1H), 8.83 (s, 1H); 13C (NMR) (CDCl3) ? 18.0, 58.8, 123.1, 126.6, 142.4, 151.1, 155.0; MS (m/z) 241 (54), 240 (100), 226 (18), 212 (27), 196 (88), 147 (29), 182 (25); HRMS for C11H19O3NSi calcd 241.1134, found 241.1112. The IR, 1H and 13C NMR data were identical to published spectral data.221 In our laboratories, using the conditions given by Masuda,221 we obtained a maximum isolated yield of 10% (Masuda reported a 56% yield). N Si(OEt)3 N Si(OEt)3 114 2-(Triethoxysilyl)thiophene (Table 10, entry 17). The general procedure for silylation was followed using 2-iodothiophene (552 ?L, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triethoxysilane (1.23 g, 1.38 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 60 ?C for 12 h. Bulb- to-bulb distillation (125 ?C, 0.5 torr) afforded 1.13 g (92%) of 2-(triethoxysilyl)thiophene as a colorless oil. Spectral data is reported above. 4-(Triisopropoxysilyl)anisole. The general procedure for silylation was followed using 4-iodoanisole (1.17 g, 5.00 mmol), i-Pr2NEt (1.94 g, 2.61 mL, 15.0 mmol), Pd(dba)2 (86 mg, 0.15 mmol), P(t-Bu)2(o-biphenyl) (120) (90 mg, 0.30 mmol), and triisopropoxysilane (1.55 g, 1.50 mL, 7.50 mmol) in 20 mL of NMP. The reaction was heated at 80 ?C for 12 h. Bulb-to-bulb distillation (125 ?C, 0.5 torr) afforded 1.17 g (75%) of 4-(triisoproxysilyl)anisole as a colorless oil. IR (neat) 2972 (s), 2934 (m), 2896 (m), 1597 (s), 1505 (m), 1381 (m), 1369 (m), 1281 (m), 1252 (m), 1174 (s), 1124 (vs) cm-1; 1H (NMR) (CDCl3) 1.20 (d, J = 6.1, 18 H), 3.81 (s, 3 H), 4.25 (septet, J = 6.1, 3 H), 6.90 (d, J = 8.7, 2 H), 7.61 (d, J = 8.7, 2 H) ? ; 13C (NMR) (CDCl3) 25.5, 54.9, 65.3. 113.4, 123.8, 136.6, 161.1 ? ; MS (m/z) 314 (17), 299 (9), 253 (72), 211 (15), 205 (38), 187 (15), 167 (100), 147 (15), 121 (20); HRMS for C16H28O4Si calcd 312.1757, found 312.1772. General Procedure for the Homocoupling of 2-Halopyridines under Palladium Catalysis. The indicated 2-halopyridine (1.00 mmol, 1.0 equiv), and i-Pr2NEt (388 mg, 523 ?L, 3.00 mmol) were added to a stirring solution of Pd(dba)2 (29 mg, 0.05 mmol), P(t-Bu)2(o-biphenyl) (120) (30 mg, 0.1 mmol), and the internal standard naphthalene (128 mg, 1.0 mmol, 1.0 equiv) in 4 mL of NMP under an atmosphere of argon. The reaction was heated at 60 ?C for 12 h. Progress was monitored by GC analysis of aliquots of the quenched reaction mixture. GC response factors relative to the internal naphthalene S Si(OEt)3 (i-PrO)3Si OCH3 115 standard were determined, and the observed percentages of products were normalized accordingly. The homocoupled product 2,2?-dipyridyl was identified by comparison of the GC retention time to that an authentic sample. 2,2?-Dipyridyl (130). The general procedure for homocoupling was followed using 2-iodopyridine (129)(205 mg, 106 ?L, 1.00 mmol), i-Pr2NEt (388 mg, 523 ?L, 3.00 mmol), Pd(dba)2 (29 mg, 0.05 mmol), and P(t-Bu)2(o-biphenyl) (120) (30 mg, 0.1 mmol), in 4 mL of NMP at 60 ?C for 12 h gave 99% 130 by G.C. analysis. 2,2?-Dipyridyl (130). The general procedure for homocoupling was followed using 2-bromopyridine (42)(158 mg, 95 ?L, 1.00 mmol), i-Pr2NEt (388 mg, 523 ?L, 3.00 mmol), Pd(dba)2 (29 mg, 0.05 mmol), and P(t-Bu)2(o-biphenyl) (120) (30 mg, 0.1 mmol), in 4 mL of NMP at 60 ?C for 12 h gave 2% 130 and 25% pyridine by G.C. analysis. The remaining percentage was unreacted starting material. General Procedure for the Optimization of the Silylation of Aryl Triflates (131) (Table 11). The indicated aryl triflate (1.00 mmol, 1.0 equiv), and base (3.00 mmol, 3.0 equiv) were added to a stirring solution of Pd(dba)2 (29 mg, 0.05 mmol, 5 mol %), the phosphine (0.10 mmol, 10 mol%), and the internal standard naphthalene (128 mg, 1.0 mmol, 1.0 equiv) in 4 mL of solvent under an atmosphere of argon. Triethoxysilane (246 mg, 277 ?L, 1.50 mmol, 1.5 equiv) was added, causing bubbling, formation of yellow foam, and darkening of the reaction mixture. The reaction mixture was allowed to stir at room temperature for 12 h. Progress was monitored by GC analysis of aliquots of the quenched reaction mixture. GC response factors relative to the internal naphthalene standard were determined, and the observed percentages of products were normalized N N N N 116 accordingly. The reduced product 121 was identified by comparison of the GC retention time to that of an authentic sample. OTf R H Si(OEt)3 Si(OEt)3 R H R + + Pd(dba)2 ligand base solvent 131 92 121 Conditionsa Yield (%)b,c Entry R Ligand Base Solvent 92 121 1 OMe P(t-Bu)2(o-biphenyl) i-Pr2NEt NMP 3 90 2 OMe P(t-Bu)2(o-biphenyl) none NMP 0 5 3 OMe P(t-Bu)2(o-biphenyl) i-Pr2NEt dioxane 2 96 4 OMe P(t-Bu)2(o-biphenyl) none dioxane 0 7 5 OMe PPh3 i-Pr2NEt NMP 0 5 6 OMe PPh3 i-Pr2NEt dioxane 0 24 7 OMe P(cy)2(o-biphenyl) i-Pr2NEt NMP 1 32 8 OMe P(cy)2(o-biphenyl) i-Pr2NEt dioxane 1 99 9 NO2 P(t-Bu)2(o-biphenyl) i-Pr2NEt NMP 0 100 10 NO2 P(t-Bu)2(o-biphenyl) none NMP 0 22 11 NO2 P(t-Bu)2(o-biphenyl) i-Pr2NEt dioxane 0 100 12 NO2 P(t-Bu)2(o-biphenyl) none dioxane 0 15 13 NO2 PPh3 i-Pr2NEt NMP 0 100 14 NO2 PPh3 i-Pr2NEt dioxane 0 100 15 NO2 P(cy)2(o-biphenyl) i-Pr2NEt NMP 0 100 16 NO2 P(cy)2(o-biphenyl) i-Pr2NEt dioxane 0 100 a Reactions of aryltriflate (131) (1.0 mmol) with H-Si(OEt) 3 (1.5 mmol) were performed atroom temperature for 12h in 4 mL of solvent by using Pd(dba) 2 (5 mol %), phosphine (10mol %), and base (3 mmol). b GC yields are based on amount of aryltriflate (131). c Remaining percentage was unreacted starting material. Table 11. Silylation of Aryltriflate Derivatives. General Procedure for the Optimization of the Silylation of Aryl Halides or Acyl Chlorides Using Hexamethoxydisilane (Table 12 and Scheme 71). The indicated aryl halide or acyl chloride (1.00 mmol, 1.0 equiv), and base (3.00 mmol, 3.0 equiv) were added to a stirring solution of palladium catalyst (0.05 mmol, 5 mol %), the phosphine 117 (0.10 mmol, 10 mol%), and the internal standard naphthalene (128 mg, 1.0 mmol, 1.0 equiv) in 4 mL of solvent under an atmosphere of argon. Hexamethoxydisilane (364 mg, 332 ?L, 1.50 mmol, 1.5 equiv) was added, causing darkening of the reaction mixture. The reaction mixture was allowed to stir at the given temperature for 12 to 48 h. Progress was monitored by GC analysis of aliquots of the quenched reaction mixture. GC response factors relative to the internal naphthalene standard were determined, and the observed percentages of products were normalized accordingly. The reduced product 121 and methylbenzoate were identified by comparison of the GC retention time to that of an authentic sample. X (MeO)3Si Si(OMe)3 Si(OMe)3 H R R R + + Pd(0) ligand base solvent 139 92 121 Substrate Conditionsa Yield (%)b,c Entry X R Ligand Pd catalyst / base solvent T (?C) 92 121 1 I OMe P(t-Bu)2(o-biphenyl) Pd(dba)2 / i-Pr2NEt NMP 25 0 0 2 I OMe P(o-tol)3 Pd(dba)2?CHCl3/i-Pr2NEt NMP 25 0 2 3 Br OMe P(t-Bu)2(o-biphenyl) Pd(dba)2 / i-Pr2NEt NMP 25 0 0 4 Cl OMe P(t-Bu)2(o-biphenyl) Pd(dba)2 / i-Pr2NEt NMP 25 0 0 5 Br Ac P(t-Bu)2(o-biphenyl) Pd(dba)2 / i-Pr2NEt NMP 100 0 100 6 Br Ac P(t-Bu)2(o-biphenyl) Pd(dba)2 / i-Pr2NEt DMPU 100 0 100 7 I Ac P(t-Bu)2(o-biphenyl) Pd(dba)2 / KF (aq) DMPU 75 0 100 8 Br Ac P(t-Bu)2(o-biphenyl) Pd(dba)2 / KF (aq) DMPU 75 0 100 9 Br OMe P(t-Bu)2(o-biphenyl) Pd(dba)2 / K2CO3 (aq) DMPU 75 0 0 10 I OMe Pd(PPh3)4 / TBAF NMP 25 0 0 11 I Me Pd(PPh3)4 / TBAF DMF 25 0 3 12 I Me Pd(PPh3)4 / TBAF HMPA 25 0 3 13d,e Br H Pd(PPh3)4 toluene 140 27 10 a Reactions of aryl halide 139 (1.0 equiv) with (MeO) 3Si-Si(OMe)3 (2.0 equiv) wereperformed using the indicated temperature and solvent for 12h-48h by using the given palladium catalyst (5 mol %), phosphine (10 mol %), and base (3 equiv). b GC yields are based on amount of aryl halide 139. c Unless otherwise noted, remaining percentage was unreacted starting material (139). d Reaction was performed in a sealed tube for 72h. e Biphenyl was observed in 7% yield. Table 12. Silylation of Arylhalides Using Hexamethoxydisilane. 118 Chapter 2. Trimethylsilyl Cyanide As A Cyanide Source For Nucleophilic Substitution Introduction Fluoride Ion Activation of Silicon Bonds The study of pentacoordinate and hexacoordinate organosilicon derivatives is approximately a century old. Hypercoordinate organosilicon compounds have been studied since the early 19th century when Dilthey reported the first known hypercoordinate organosilicates, six-coordinate silicon diketonates.1-4 To date, a wide variety of penta- and hexacoordinate organosilicon compounds have been synthesized and characterized; surprisingly, despite the nonexistence of naturally-occurring organosilicates,5,6 a handful of silicates which have no carbon analogs are biologically active.7-13 The synthesis and structure of hypercoordinate silicates most notably have been studied and reviewed by Voronkov,1,14-19 Frye12,20-33 and Lukevics.34-41 In turn, Corriu42-46 has investigated and extensively reviewed silicate reactivity and reaction mechanisms. Hypercoordinate silicates first became an area of active research in the DeShong laboratories with the synthesis of tetrabutylammonium triphenyldifluorosilicate (TBAT, 1, Scheme 1).47-51 TBAT was initially developed as a source of nucleophilic fluoride ion for SN2-type displacements48 and for Si-C bond cleavage.47 Stable, crystalline, non-basic, and non-hygroscopic, pentacoordinate silicate 1 is an excellent fluoride surrogate when compared to alkali metal fluorides or tetraalkylammonium fluorides (Scheme 1, eq. a). More recently, TBAT has been developed as a highly effective phenylation reagent in palladium-catalyzed cross coupling reactions, offering a silicon-based alternative to Suzuki (boron) and Stille (tin) reagents (Scheme 1, eq. b).50 119 Scheme 1 Si MeO I MeO Ph Br 1, TBAT Pd(0) (a) (b) Ph PhPh F F Ph FPh Bu4N+ Relative to other fluoride sources, TBAT was shown to yield less alkene elimination by-products in standard nucleophilic displacements of primary and secondary alkyl halides and sulfonates. Most notably, when compared to the popular organic- soluble reagent tetrabutylammonium fluoride (Bu4N+ F-, TBAF), TBAT is far less basic. For example in Scheme 2, whereas TBAT achieved nearly quantitative fluoride displacement of a secondary alkyl tosylate, TBAF provoked elimination leading to alkene by-products. Scheme 2 OTs TBAT, 1 MeCN, reflux F + OTs TBAF MeCN, 25 ?C F + 98% 2% 58% 32% 24 h 1 h It is also noteworthy that reactions with TBAT proceeded much more slowly than the corresponding reaction with anhydrous TBAF. Based on the observed rate difference as well as NMR spectroscopic studies, it was concluded that TBAT does not operate by a simple in situ disproportionation to TBAF (Scheme 3) and fluorotriphenylsilane (2). Instead, the hypercoordinate silicate itself must directly transfer a softer, less basic fluoride nucleophile to the substrate. 120 Scheme 3 Bu4N+F- + Ph3Si-F 2 Si 1, TBAT Ph PhPh F F Bu4N+ Given the effectiveness of TBAT as a practical reagent for the delivery of fluoride anion, extension of this strategy to deliver other nucleophiles by generating the reactive hypercoordinate silicate in situ was investigated (Scheme 4). In this chapter, it is reported that trimethylsilyl cyanide (Me3SiCN, 3) and trimethylsilyl azide (Me3SiN3, 4) underwent reaction with tetrabutylammonium fluoride (TBAF) to generate the respective hypercoordinate trimethylfluorosilicate (5 or 6) in situ. Analogous with the observations using TBAT, silicates 5 and 6 were extremely reactive alternates of cyanide and azide anion, respectively, for SN2 displacement (vide infra).52-54 Scheme 4 Si F Br R3Si X TBAF X X Bu4N+ 1: R = Ph, X = F (TBAT) 5: R = Me, X = CN 6: R = Me, X= N3 R RR Ph Ph 2: R = Ph, X = F 3: R = Me, X = CN 4: R = Me, X= N3 The chemistry described in Scheme 4 is one of the characteristic reactions of tetracoordinate silanes: the activation of Si-H, Si-O, Si-C, Si-N, or Si-M (M = Si, Ge, Sn) bonds by addition of a nucleophile, typically fluoride ion. The Si-X bonds of tetracoordinate silanes normally are not labile in the absence of nucleophilic activation. Pentacoordinate silicates such as 5 and 6 are widely accepted as the reactive intermediates in most reactions at silicon, and many new isolable hypercoordinate silicates have been synthesized as probes of the mechanism of the reactions of silanes.43 Chap. 1,44 121 In his dissertation on the development of TBAT as a fluorinating reagent, Pilcher presented the properties, reactivity, and methods of analysis of hypercoordinate fluorosilicates.55 The aim of this chapter is twofold: to review the literature precedence which supports the existence of hypercoordinate intermediates such as 5 and 6, and to present pertinent examples of fluoride-mediated reactions of silanes. The following features are discussed: 1. The expansion of coordination at silicon. Formation of pentacoordinate silicon has been proposed as the fundamental step of most transformations of silicon compounds.43 Chap. 1,44 Examples of isolable pentacoordinate silicates are presented as analogues of the key pentacoordinate reaction intermediates 5 and 6. In addition, analytical methods for the characterization of hypercoordinate silicates are discussed. 2. Mechanisms of nucleophilic substitutions at silicon. The process depicted in Scheme 4 involves nucleophilic attack on silicon by fluoride ion, and displacement of the group X (cyanide or azide). Nucleophilic substitutions at silicon take place via the hypercoordinate intermediate, such as 5 and 6, and not through an SN1 or SN2-type process. 3. Pertinent examples of fluoride ion activation of silicon bonds in organic synthesis, encompassing desilylation (removal of protecting groups), reduction, and group transfer reactions (allylation, arylation, acylation, etc.). Particular attention will be paid to experimental evidence for the participation of hypercoordinate silicate intermediates. Finally, the results of studies in the DeShong lab on the use of organosilicon compounds under nucleophile-activated conditions are presented, and discussed. 122 Atomic and Molecular Properties of Tetracoordinate and Hypercoordinate Organosilicates Silicon compounds are often compared to their carbon-based analogues: silicon is in the same period of the periodic table as carbon; intuitively, their chemistries should be similar. Silanes, like alkanes, are neutral, tetracoordinate species. However, beyond sharing a valence number of four, the properties of carbon and silicon compounds are surprisingly different. On the basis of this periodic proximity of silicon to carbon, the first researchers of silicon chemistry correctly postulated that silicon compounds should undergo similar chemical transformations, most notably nucleophilic substitution.56 However, fundamental differences in the atomic and molecular natures of silicon and carbon cause a divergence in their respective mechanistic pathways to nucleophilic substitution. The larger atomic size, lower electronegativity, and the availability of low energy d orbitals endow silicon with enhanced reactivity, and metallic properties. The most notable contrast between silicon and carbon is that silicon can expand beyond its normal coordination number of four, becoming what is termed hypercoordinate. Silicates are a class of anionic compounds featuring a silicon atom with five (7 and 8) or six (9) ligands as opposed to the normal silane coordination number of four. N F Si F F F NSiN S N N 9TAS-F, 8 F F MeMeMeSi 7 Ph PhPh OMe OMe K+ 18-crown-6 2- The ability of silicon to become hypercoordinate is due to a number of steric and electronic factors (Table 1). With a significantly larger covalent radius than carbon, silicon can satisfy the steric demands of penta- and hexacoordination.55,57 Addition of an anion to tetracoordinate silicon is a lower energy process than for carbon, as indicated by the higher silicon electron affinity.58 Conversely, silicon has a lower ionization potential than carbon, allowing ligands to share the negative charge on, and thereby stabilize a penta- or hexacoordinate complex.59 123 Si Atomic Property C 1.06 ? atomic radius60 0.66 ? 1.17 ? covalent radius61 0.77 ? 3.35 eV for F3Si? electron affinity58 1.85 eV for CF3C? 8.15 eV ionization potential59 11.26 eV 1.74 Allred-Rochow electronegativity56,62 2.50 Table 1. Comparison of Atomic Properties of Si and C. The relatively low silicon electronegativity has two major consequences: (1) silicon is often polarized in the opposite sense to carbon (i.e., Ph3Si?+-H?- vs. Ph3C???H?+); and (2) the Si-X and Si-O bonds are highly polarized?much more so than C-X and C-O bonds?rendering silicon more electrophilic and susceptible to nucleophilic attack.56 Using ab initio methods and experimentally determined heats of reaction to compare the structures and bond energies of silyl (H3Si-X) and methyl (H3C-X) derivatives (X = Li, BeH, BH2, CH3, NH2, OH, F, and Na), Schleyer demonstrated that the relative Si-X and C-X bond energies are directly proportional to the electronegativity of the group X.63 For reference, Table 2 lists the relative electronegativities of the atoms to which silicon typically forms bonds. In addition, Schleyer described the Si-X covalent bond as more ionic (Si?+-X?-) in nature than the predominantly covalent C-X bond.64 As indicated by the bond dissociation energies (Table 3), silicon forms remarkably strong bonds to F and O, and somewhat weak bonds to Si, and H. In contrast, carbon forms strong C-C and C-H bonds, and relatively weak C-O and C?X bonds. The weak Si-H bond explains in part why hydrosilanes are good reducing agents, in stark contrast to alkanes.56 Also, formation of the strong Si-F bond is often the thermodynamic driving force of reactions at silicon; fluoride is typically the unrivaled catalyst for the cleavage of other Si-X bonds.42 Note that, although the Si?Cl and Si-F bond strengths are greater than the C-Cl and C-F bond strengths, the Si-X bonds are kinetically quite reactive (readily broken), and more labile than the equivalent C-X bond. This discrepancy between kinetic reactivity and thermodynamic bond strength in substitution reactions at silicon in comparison to carbon is attributed in part to the attraction of nucleophiles to the larger partial positive charge on Si.43,56,61 124 Atom Electronegativity Atom Electronegativity F 4.10 S 2.44 CN 3.60 H 2.20 O 3.50 As 2.20 OH 3.50 P 2.06 N 3.07 Ge 2.02 NH2 3.05 B 2.01 OCH3 2.90 BH2 1.90 Cl 2.83 Si 1.74 Br 2.74 BeH 1.50 C 2.50 Al 1.47 Se 2.48 Na 1.01 CH3 2.45 Li 0.97 Table 2. Atomic and Group Allred-Rochow Electronegativities.56,62 Bond Energy (kJ mol-1)61,65 Bond Length (?)56 Bond to C Si C66 Si67,68 F 485 582 1.39 1.6 Cl 327 391 1.78 2.05 Br 285 310 1.94 2.21 I 213 234 2.14 2.44 O 336 368 1.41 1.63 N 335 400 1.47 1.74 C 356 373 1.53 1.87 Si 250-335 210-250 1.87 2.34 Table 3. Approximate Average Bond Dissociation Energies and Bond Lengths for Tetravalent C and Si. 125 X-Ray crystallographic studies of stable, solid silicates show that tetracoordinate, pentacoordinate and hexacoordinate silicon compounds adopt the tetrahedral, trigonal bipyramidal (TBP), and octahedral geometries, respectively (Figure 1).1 Most known silicon compounds are tetrahedral (3sp3 hybridized).56 Tetrahedral silanes exhibit high configurational stability, albeit lower than carbon,44 and chiral silanes have been synthesized predominantly for use as stereochemical probes of the mechanism of nucleophilic displacement at Si.33,44-46,69-71 Si eq Si eqeq ap ap Sieqeq eqeq ap ap tetrahedral sp3 ? =109.5? octahedral sp3d2 ? (eq-Si-ap) = 90? ? (eq-Si-eq) = 90? 2 trigonal bipyramidal sp3d ? (eq-Si-ap) = 90? ? (eq-Si-eq) = 120? Figure 1. Hybridization of Tetra-, Penta-, and Hexacoordinate Silicon. Pentacoordinate complexes are trigonal bipyramidal (TBP) (3sp3d hybridized) with distinct apical (ap) and equatorial (eq) ligands (Figure 1); the apical bonds (pd-like) are longer and weaker than the equatorial bonds (sp2-like).1,72 ?Bent?s Rule? dictates that electronegative groups prefer to occupy the apical sites.73-75 Electron-withdrawing groups (EWG) tend to decrease the apical silicon-ligand bond lengths and thus increase the overall stability of the complex. Unless a ligand in the complex bears a neutralizing positive charge, the complex bears a negative charge. Octahedral silicates bear a negative two charge, again, unless cationic ligands are present to neutralize the overall charge (as in compounds 9 and 11). A large number of hexacoordinate silicates have been synthesized and isolated, two of which are shown below. The triscatecholate silicate 10 was isolated by Corriu, and shown to be a useful starting material for the preparation of alkylsilanes.76 Stacked arrays of phthalocyanine derivative 11 are electrically conductive.77-81 126 R SiOO OO O O 2- 2Na+ 10 SiN NNN N N N N R' 11 The results of calculation and experimentation indicate that the negative charge lies on the ligands, with the most electronegative ligands bearing the brunt of the charge.82 Remarkably, as a silicon complex goes from neutral (tetracoordinate) to anionic (penta-, and then hexacoordinate), the silicon center becomes increasingly positive, and electrophilic.1 In turn, the ligands become more negative, and nucleophilic. For example, the CNDO/2 calculated charges on the central Si atom of SiF4 and SiF5- are +1.214 and +2.028, respectively; in turn, the charges on the fluorine ligands are -0.304 (sp3), and - 0.399(eq)/-0.415(ap), respectively.83 The net effect of hypercoordination on the reactivity of a silicate is enhanced ligand Lewis basicity, and increased silicon Lewis acidity.1 Corriu illustrated the ?snowball effect? of increasing electrophilicity at silicon with increasing coordination number by demonstrating that the rate of conversion from tetra- (12) to pentacoordinate (13) is slower than the analogous conversion from penta- (14) to hexacoordinate (15) for a series of catecholate derivatives (Figure 2).84 The researchers attributed the observed kpent99:1) R = OBz (anti : syn = 96:4) 36 37 TBAF, THF As with the fluoride-mediated reactions of silyl enol ethers, the stereochemical outcome of the hydrosilation of aldehydes and ketones supports the proposed mechanism (Scheme 21). For instance, the anti-Felkin-Anh stereoselectivity favoring formation of diol 39 in the intramolecular reduction of ?-dimethylsiloxy ketones 38 can be explained by the influence of a chair-like intermediate on the facial selectivity of the reduction.176 Scheme 21 Ph O TAS-F, 8 Ph OSiMe2H H Si F O O Me Me Ph H H H Ph F Si H O O Me Me Ph H H H Ph Ph Ph OHOH Ph Ph OHOH 39 40 38 39:40 = 3:1 THF 68% Fluoride Activation of the Si-C Bond The fluoride-promoted desilylation of organosilicates containing the C-SiR3 bond is a general method for C-C bond formation via the transfer to carbon electrophiles of allyl,177 acyl,178 alkynyl,179,180 propargyl,181,182 benzyl,183,184 oxiranyl,185 and other stabilized carbanions.35,42,133,159 As described in the first chapter, organosilicates also participate in a 142 variety of interesting fluoride-promoted transition metal catalyzed C-C bond forming reactions; however, these are beyond the scope of this chapter.42 In particular, fluoride activation of the Si-allyl bond has been extensively used in organic synthesis for the allylation of aldehydes, ketones, and ?,?-unsaturated compounds (Michael addition).56 The regioselectivity of the addition reaction of trimethylallylsilanes (42 and 43) to aldehydes (41) is controlled by steric crowding and the electronic distribution at each allyl terminus, and mixtures of ?- and ?-alkylated products (44 and 45, respectively) are observed (Scheme 22).177 Isomeric trialkyl silanes 42 and 43 yield comparable ?- and ?- alkylated product ratios (44 : 45). Scheme 22 i-Pr H O SiMe3 SiMe3 TBAF TBAF i-Pr OH i-Pr OH + 41 44 45 42 43 44:45 = 1.5:1 The observed regiochemistry is consistent with a mechanism involving generation of the free allyl anion.177 However, formation of a truly ?free? allyl anion is unlikely: the free allyl anion is basic enough to abstract a proton from the tetrabutylammonium cation of TBAF, yet no proton abstraction is observed.42,186 Neither extreme?a free anion or a silicate-based cyclic transition state?is plausible based on experimental observation. A hybrid of these two extremes has been proposed, wherein a non-basic hypercoordinate silicon complex 46 acts simply as an allyl transfer reagent, and the silicon center exacts no influence on the regiochemical outcome (Scheme 23).177 Scheme 23 SiMe3 TBAF Si F Me MeMe i-PrCHOSi F Me MeMe - FSiMe3 i-Pr OH ?? ?? 46 143 In contrast to the trimethylallylsilane analogs, the more Lewis acidic trifluoro- and trialkoxyallylsilanes more readily form a hexacoordinate intermediate.42 As with the reactions of silyl enol ethers, allylation takes place with a high degree of regio- and stereoselectivity as dictated by the nucleophilicity of the fluoride activator. Irrespective of the nature of the fluoride species, the mechanism begins with the formation of the pentacoordinate fluorosilicate 47 (Scheme 24) When an extremely nucleophilic fluoride donor (one that is reactive enough to cleave the relatively strong Si-C bond, such as phosphazenium fluoride, PZ+F-) is employed, fluoride rapidly attacks the pentacoordinate intermediate 47, displacing a free allyl anion 49 and generating SiF5- (eq. b); as illustrated in the mechanism, 2 equiv of PZ+F- are required.187 In agreement with the reactivity of free allyl anions generated by other means, the allyl anion released from the silicate reacts with primary alkyl halides.187 When a less reactive fluoride source is employed (in this case TBAF is sufficiently weak), the substrate coordinates to pentacoordinate intermediate 47 (eq. a). A hexacoordinate species 48 is formed which ultimately controls the stereo- and regiochemical (exclusively ?-alkylation) outcome of the allylation by inducing a 6- membered cyclic transition state.42 Scheme 24 SiF3 Si F F FFF CH2 Si F F F O F H R R OSiF3 PZ+47 (a) 48 49 PZ+F- RCHO TBAF - F RCHO R O (b) 2 F- SiF5- - SiF5- For example, a high degree of stereospecificity was observed for the CsF-induced addition of (E)- and (Z)-crotyltrifluorosilanes to aldehydes (Scheme 25): erythro isomers 54 were obtained from (E)-crotyltrifluorosilane 50; and threo isomers 55 were obtained from (Z)-crotyltrifluorosilane 51.188 All yields were moderate to excellent (68-96%). 144 Scheme 25 Si FF F F Me O H R SiF3 R OH R OHSiF3 Si FF F FO H RMe H H + RCHO 54 erythro 55 threo + RCHO R = Ph, 50 (92%): 99/1 erythro/threo 51 (96%): 1/99 erythro/threo R = octyl, 50 (96%): 99/1 erythro/threo 51 (89%): 2/98 erythro/threo R = i-Pr, 50 (68%): 99/1 erythro/threo 51 (90%): 10/90 erythro/threo 50 51 52 53 CsF THF CsF THF ? ? The use of fluoride for the activation of Si-X bonds (especially, X = H, O, and C) is an effective and well-studied synthetic strategy. Particularly noteworthy is the formation of C-C bonds with a high level of regio- and/or stereocontrol. Notoriously problematic nucleophiles, such as enolates, or carbon nucleophiles that can not be easily generated by other means (such as deprotonation) are accessible when delivered as the silicate.133 To date, this approach has been exploited for the delivery of a wide range of nucleophiles with a reactivity that is markedly different from the free nucleophile, including the fluoride ion activation of Si-N and Si-M (M = Si, Ge, Sn) bonds.35,42,133 Ubiquitously, the experimentally observed difference in reactivity and product distribution between the silicate-delivered vs. the ?free? nucleophile is rationalized by the formation of a hypercoordinate silicon reaction intermediate. 145 Results and Discussion Background Nucleophilic substitution is one of the most prized and useful transformations in the chemical arsenal. At the practical level, however, nucleophilic displacements often require relatively harsh reaction conditions with high reaction temperatures, polar solvents such as DMSO, HMPA or DMF, and a large excess of the nucleophile if high yields of product are to be achieved. Recently, the DeShong research group reported the use of tetrabutylammonium triphenyldifluorosilicate (Ph3SiF2- Bu4N+, TBAT, 1) as a source of nucleophilic fluoride for SN2 displacement of primary and secondary alkyl substrates.47 As illustrated in Scheme 26, TBAT is an excellent fluoride surrogate when compared to alkali metal fluorides or tetraalkylammonium fluorides because TBAT is relatively non-basic, as attested to by the reduced amount of elimination products. In addition, TBAT is a more practical reagent to handle, and store: TBAT is crystalline, soluble in a wide range of organic solvents, and is non-hygroscopic. The anhydrous nature of TBAT is its greatest advantage over other fluoride sources: the presence of hydroxide in traditional fluorinating reagents, from the reaction of fluoride anion and water, causes unwanted side reactions, most notably elimination or hydroxylation. The displacement of a secondary tosylate is particularly illustrative (Scheme 26, eq. b): 98% yield of the desired fluoroalkane was achieved with TBAT (1, 6 equiv) in refluxing MeCN after 24 h; only 58% yield was obtained with TBAF (2 equiv) in THF at room temperature after 1 h, with the remaining yield being the alkene elimination products.48,55 146 Scheme 26 TBAT, 1 (4 equiv) MeCN, 82 ?C, 24 h 85 % THF, 25 ?C, 1h 48 % Br O O O O F F TBAF (2 equiv) OTs F THF, 25 ?C, 1h 58 % TBAF (2 equiv) TBAT, 1 (6 equiv) MeCN, 82 ?C, 24 h 98 % O O O O OMs TBAT, 1 (6 equiv) MeCN, 82 ?C, 24 h 73 % DMAC, 100 ?C, 2.5h 55 % KF (saturated) Admittedly, this technology does have some drawbacks. The fact that 4 to 6 equivalents of TBAT in MeCN are required to obtain reasonable reaction rates is wasteful and causes purification problems. These failings are partly ameliorated by the fact that unreacted TBAT can be precipitated and filtered from the reaction mixture and recycled. Also, subsequent optimization showed that excellent results are obtained with just two equivalents of TBAT in refluxing dioxane.55 In an effort to enhance the reactivity of TBAT, Pilcher and Loezos studied the performance of a series of TBAT analogues, wherein the phenyl ligands were systematically replaced by methyl or fluoro substituents.55,189 The new fluorosilicates were tested in a standard displacement reaction (Table 5). The observed trend can be summarized as follows: as a strongly electron-withdrawing fluoride ligand of SiF5- Bu4N+ (59, entry 10) is replaced by a weaker EWG (phenyl), or a weak EDG (methyl), the 147 complex becomes more reactive. The increase in reactivity parallels the increase in the length of the Si-F bond, and the concomitant bond weakening.55,189 4 equiv fluorosilicate MeCN, 85 ?C + 9Br9 F9 Entry # Fluorosilicate Time (h) Fluoro % Alkene % 1 (Me2N)3S+ Me3SiF2- (TAS-F, 8) 0.5 40 60 2 Bu4N+ Me2PhSiF2-, 56 2 84 16 3 Bu4N+ Me2SiF3-, 57 5 85 15 4 Bu4N+ MePh2SiF2-, 58 6 92 8 5 Bu4N+ Ph3SiF2- (TBAT, 1) 24 85 15 6 Bu4N+ MePhSiF3-, 59 24 22 11 7 Bu4N+ Ph2SiF3-, 60 24 22 7 8 Bu4N+ MeSiF4-, 61 24 0 0 9 Bu4N+ PhSiF4-, 62 24 0 0 10 Bu4N+ SiF5-, 63 24 0 0 Table 5. Reactivity of a Series of Fluorosilicate Complexes in the Fluorination of 1-Bromododecane at 85 ?C.55,189 Although Bu4N+ Me2PhSiF2- (56), Bu4N+ Me2SiF3- (57), and Bu4N+ MePh2SiF2- (58) (Table 5, entries 2-4), offer a better reactivity profile than TBAT (1, entry 5), with shorter reaction times and comparable yields, TBAT remains the reagent of choice due a number of factors: (1) ease of handling and of synthesis from commercially available starting materials; and (2) the crystalline nature allows for the synthesis of a pure anhydrous reagent. TBAT is one of the few isolable solid silicates depicted in Table 5: simple exchange of methyl for phenyl not only impacts the reactivity of TBAT, but also greatly destabilizes the complex; even exchange of fluoro for phenyl yields a less stable silicate.55,189 Conveniently, once the tetracoordinate silane had been synthesized, the unstable TBAT analogs 56-58 were easily generated in situ by treatment with TBAF (Scheme 27, eqs. a-c); TBAT itself can be generated without isolation by treatment of Ph3SiF (2) with TBAF (eq. d). 148 Scheme 27 aq. HF Bu 4N+ [Me2PhSiF2]-, 56TBAF Bu4N+ [Me2SiF3]-, 57TBAF aq. HF Bu 4N+ [MePh2SiF2]-, 58TBAFaq. HF Bu4N+ [Ph3SiF2]-, 1TBAFPh3SiOH Ph3SiF, 2 MePh2SiCl MePh2SiF Me2PhSiCl Me2PhSiF Me2SiCl2 Me2SiF2SbF3 (a) (b) (c) (d) Having demonstrated that hypercoordinate silicate derivatives are practical reagents for the delivery of fluoride anion,47,48,189 extension of this strategy to deliver other nucleophiles was investigated. The goal was to focus on developing a method in which the hypercoordinate silicon reagent is made in situ from either readily available or easily synthesized tetracoordinate silane starting materials. The attention focused on the use of substituted trimethylsilyl derivatives (Me3SiX), many of which are commercially available (Scheme 28). Scheme 28 Me Si MeMe F X Me3Si X TBAF Bu4N+ 5: R = Me, X = CN 6: R = Me, X= N3 3: X = CN 4: X= N3 Ph Ph XBr The initial experiments involved Me3SiCN (3) and Me3SiN3 (4), the synthetic utility of cyanide and azide anion in nucleophilic displacements already being recognized.190 Organonitriles are important synthetic intermediates for the preparation of nitrogen- containing compounds such as amines and amides, and are readily converted to carboxylic acids, esters, or aldehydes (Scheme 29). Because of the polarization of the carbon-nitrogen triple bond, nitriles undergo a variety of reactions with electrophiles on nitrogen and nucleophiles on carbon. In addition, the acidic ?-proton can be removed by strong base.190 The nitrile moiety itself is present in a number of pharmaceutically useful natural products, and in industrially important synthetic materials.191 149 Scheme 29 COOHR NCR NCR R1 NC R Cl NC R EtO2C R COOH R COOR1 R C R O NH2NH O R CH2NH2 R CHO R C R NMgX R C R O R1 C R NH R1 R R1 NH2 NCR base R1X CCl4 Cl-CO2Et 1. t-Bu3SnCH2I 2. LiMe O2 H3O+ or OH- R1OH H+ BF3 HOAcH2SO4 LiAlH4 LiAlH(OEt)3 R1MgX H3O+ NH3LiNH 3 150 The DeShong group was interested in synthesizing glycosyl azide derivatives as precursors to N-linked or N-substituted glycoconjugates. The chemistry and biology of N-linked glycosides is a topic of intense interest due to the significant role that these compounds have in biological processes.192 Synthesis of the 1- or 6-glycosyl azide offers a facile route to the sugar amide (including glycoproteins), amine, or azanucleoside (Scheme 30).190 Scheme 30 O N3 O NH O NH2 O N R' O R'R' NN R' R' R'-CO2H PR3 H2/PdRO RO RO RO O X RO N3 - The most common method of introducing the nitrile or azide moiety is by nucleophilic displacement.190,193 Nucleophilic displacement is inherently regiospecific and stereoselective, making this approach particularly valuable to the synthetic chemist. Alkali metal cyanide or azide has traditionally been used for this purpose; however, these reagents require vigorous reaction conditions, a large excess of the reagent, and long reaction times due in part to their limited solubility in organic solvents.193 To overcome the poor solubility, phase transfer catalysts are typically employed,194 most notably crown ethers;147,195,196 however, the high cost and relative toxicity197-200 makes their use less attractive. Adsorption of alkali salts onto a solid support has been reported,201-206 however the addition of water is required which limits the use of these reagents to hydrolytically stable substrates; in addition, lengthy reaction times were reported, due to poor reaction kinetics. Other methods employ toxic reagents or activating agents such as tin(IV)207,208 or 151 HMPA.209-211 Traditional alkali cyanide or azide displacement reactions require polar aprotic solvents, such as DMSO or DMF, which can be problematic to remove. The more recently developed tetraalkylammonium salts are advantageous over their alkali metal salt analogs because they are soluble in THF and acetonitrile, and reactions are fast at or below room temperature.193,212 Unfortunately, along with displaying enhanced nucleophilicity relative to alkali metal salts, these reagents are also strongly basic, promoting elimination from the substrate. Another problem is that these salts are hygroscopic, or only available as the hydrate; dehydration often results in decomposition via attack of the nucleophile on the tetraalkylammonium cation (Scheme 31).144 Water in the presence of the nucleophile generates hydroxide that may act as a nucleophile, forming unwanted hydroxy products in addition to the desired cyano- or azido products, or a base, forming elimination products. Scheme 31 X Et3N:HX H2C=CH2 +H NEt3 Initial experiments to determine the viability of using in situ generated silicates Me3Si(CN)F- Bu4N+ (5) and Me3Si(N3)F- Bu4N+ (6) as sources of nucleophilic cyanide or azide, respectively, employed benzyl bromide (64) as the substrate, in order to avoid the possibility of elimination by-products (Scheme 32). The reaction conditions were analogous to those used for TBAT displacement reactions. Benzyl bromide (60) was treated with 2 equiv Me3SiCN (3) or Me3SiN3 (4), and 2 equiv TBAF in acetonitrile and brought to reflux to generate the proposed reactive silicate intermediates 5 or 6, respectively. In the case of Me3SiCN, phenylacetonitrile (65) was formed quantitatively in less than 5 min (Scheme 32, eq. a); in the case of Me3SiN3, benzyl azide (66) was formed quantitatively in 8 h (eq. b). No reaction occurred in the absence of TBAF. 152 Scheme 32 Br CN Br N3 Br Br CN N3 NaN3c Me3SiN3 (4), TBAF MeCN, 25 ?C, 12h 97 % KCN, 18-crown-6 Me3SiCN (3), TBAF NaCN, calix[4]areneb Me3SiCN (3), TBAF MeCN, 82 ?C, <5 min 95 % CH3CN, 25 ?C, <5 min 95 % MeCN, 82 ?C, 24h 100 % Me3SiN3 (4), TBAF MeCN, 82 ?C, 8h 95 % NaN3, hemina Benzene, 60oC, 6h 92 % H2O, 60 ?C, 2h 83 % DMF, 25 ?C, 12h 92 % (a) (b) (c) (d) 64 64 67 67 65 66 68 69 (a) Ref.212 (b) Ref.194 (c) Ref213 In order to gauge the basicity of the reaction conditions, phenethyl bromide (67), which is prone to base-catalyzed elimination to afford styrene, was employed as the substrate (Scheme 32, eqs. c and d). Under identical reaction conditions, quantitative yields of the respective cyano-(68) and azido-(69) products were obtained. Comparison of the reaction conditions for nucleophilic displacement with those typically employed demonstrates the effectiveness of the silicate strategy. Displacements employing alkali 153 metal salts in polar solvents consistently required lengthy reaction times, and afforded more elimination products than the silicate method (vide supra), attesting to the relative non-basicity of the silicate methodology. Concurrent with these studies, Takaya and co-workers published a preliminary report in which alkyl azide derivatives were prepared using the in situ-generated hypercoordinate azidosilicate 6.214 Having shown in our laboratories that the simple substrates depicted in Scheme 32 undergo facile cyanide and azide displacements using silicates 5 and 6, the goal of the research described below was to determine the generality, and to optimize the reaction conditions for the use of silicate 5 as a source of nucleophilic cyanide. Soli, et al. from the DeShong research group extended the azide methodology to the synthesis of glycosyl azide derivatives.52,54 Cyanide Displacements Utilizing Hypercoordinate Cyanosilicate 5 The major advantage of the silicate method with the substrates shown in Scheme 32 above was the rate of cyanide displacement in acetonitrile. The silicate-based displacements were extremely rapid, whereas the displacements using alkali metal salts of cyanide were sluggish (typically >12h)(Scheme 32, eqs. a and c). For example, the silicate-based displacement of benzyl bromide 64 occurred nearly instantaneously in refluxing acetonitrile, compared with the crown ether procedure that required 24 hours under identical conditions (eq. a). Results for the cyanide displacement with a variety of primary and secondary substrates bearing halide or sulfonate leaving groups are summarized in Table 6. 154 Si MeMeMe CN F Bu4N+ Me3SiCN TBAF R X R CN MeCN 53 Entry Substrate R-X Product R-CN Temp (?C) Time (h) Yielda.b (%) 1 (C6H5)CH2Br 64 (C6H5)CH2CN 65 82 25 0.1 1 95 (>95) 2 (C6H5)CH2Cl 70 (C6H5)CH2CN 65 82 25 2 72 (>95) (>95) 3 CH3(CH2)11I 71 CH3(CH2)11CN 72 82 25 0.1 6 95 (95) 4 CH3(CH2)11Br 73 CH3(CH2)11CN 72 82 25 2 36 95 (>95) 5 CH3(CH2)11Cl 74 CH3(CH2)11CN 72 82 25 3 96 95 (33)c 6 CH3(CH2)11OMs 75 CH3(CH2)11CN 72 82 0.1 95 7 (C6H5)CH2CH2Br 67 (C6H5)CH2CH2CN 68 82 25 0.1 32 95 (>95) 8 CH3(CH2)5CH(I)CH3 76 CH3(CH2)5CH(CN)CH3 77 82 25 1 72 (83) (68) 9 CH3(CH2)5CH(Br)CH3 78 CH3(CH2)5CH(CN)CH3 77 82 25 2 120 (82) (79) 10 CH3(CH2)5CH(OTs)CH3 79 CH3(CH2)5CH(CN)CH3 77 82 25 1 48 (76) (70) 11 (C6H5)CH(Br)CH3 80 (C6H5)CH(CN)CH3 81 82 25 1 5 92 (>95) 12 Br 82 CN 83 7:1 endo : exo 82101e 4896 (0) d 70 13 Br 84 CN 85 82 48 (<5)f 14 Cl 86 CN 87 82 101e 48 96 (0)d (0)d a The indicated substrate, Me 3SiCN (3), and TBAF (1:1.5:1.5 molar ratio) in acetonitrile wereallowed to react at the given temperature, unless otherwise noted. b Isolated yield after purification. The yield determined by G.C. analysis of the crude reaction mixture (vs. an internal standard) is reported in parentheses. c Reaction was stopped before completion. d No reaction was observed. e Reaction was performed in dioxane. f Starting material was consumed; remaining yield assumed to be cyclohexene. Table 6. Reaction of Hypercoordinate Silicate 5 with Alkyl Halides. 155 Several features of the results in Table 6 are noteworthy. Although displacement was predictably longer for benzyl chloride (70, entry 2) than for the bromo analog (64, entry 1), the chloride underwent cyanide displacement in comparable yield. Previous methods, while achieving similar yield, required longer reaction times, painstakingly dried reagents, DMSO as the solvent, or toxic 18-crown-6 as a phase transfer catalyst.147,215,216 Primary alkyl iodide (71, entry 3), bromide (73, entry 4), chloride (74, entry 5), and mesylate (75, entry 6) were efficiently converted to the corresponding nitrile, where the relative order of reactivity was as follows: I ? OMs > Br > Cl. Again, the yield of displacement product was comparable to traditional methodologies; however, these more conventional methods employed either high temperatures and lengthy reaction times, or required DMF as the solvent.203,212 Phenethyl bromide (67, entry 7) which should be more prone to elimination than the dodecyl substrates, gave no elimination products with silicate 5, again attesting to the non-basic nature of the reagent. Predictably, secondary halides and sulfonates reacted slower with competing elimination lowering the yields of nitrile. For example, silicate 5 rapidly converted 2-iodooctane (72, entry 8), bromide (74, entry 9), and tosylate (75, entry 10), to the nitrile 73 in good yield (76-83%), although traces of the alkene by-product were also observed. For secondary substrates, the silicate methodology was superior to prior methods with regard to both yield and ease of use. Regen and co-workers reported the displacement of the primary bromide in entry 4 (73, Table 6) in quantitative yield using NaCN-coated alumina in refluxing toluene for 24 hours.203 However, under the same conditions, Regen reported that secondary bromide 78 in entry 9 gave only 27% of the corresponding nitrile after 40 hours. The Bram group was able to perform the same transformation in 72% yield, but only under aqueous phase-transfer conditions.217 Finally, secondary benzylic bromide 80 (entry 11) underwent smooth conversion to the corresponding nitrile upon treatment with cyanosilicate 5 without the formation of the elimination product. Again the yield was superior to that previously reported.201 156 In refluxing acetonitrile, no appreciable amount of nitrile product formed from norbornyl bromide (82, Table 6, entry 12); however, in refluxing dioxane, displacement of the exo bromide occurred in good yield to give endo nitrile 83, albeit sluggishly. This reaction occurred with predominantly inversion of configuration, supporting our proposed SN2 mechanism for displacement at the carbon center. The 1H NMR spectrum as well as the gas chromatograph of mixtures of endo- and exo-2-norbornane carbonitrile (84) were unresolved, and therefore could not be used for determining the endo:exo ratio.218-220 Nor could the two products be separated by physical means. The 13C NMR spectrum of each stereoisomer has been reported previously. Fortunately, the 13C NMR spectrum of the mixture of stereoisomers is partially resolved at the C1, C6, and Cx (the nitrile carbon) signals. The endo/exo ratio was determined by integration of the 13C NMR spectrum: the relative areas of the carbon signals as shown above were: C6 (6.91:1.00), C1 (6.94:1.00), Cx (6.92:1.00), for an average endo/exo ratio of 6.92:1.00. 1 2 34 7 6 5 CNx 83 As expected, the cyclohexyl halides were poor substrates for the displacement reaction even with the silicate derivative. Cyclohexyl bromide (84, Table 6, entry 13) failed to give the displacement product, and gave the alkene instead, mirroring the results of previous researchers with more basic reagents.147,196,209 (-)-Menthyl chloride (86, entry 14), was similarly unreactive, presumably due to steric hindrance. However, it is noteworthy that the reagent did not induce elimination under these reaction conditions. Cyanide is an ambident nucleophile and is known to react as a carbon nucleophile to give nitriles (R-CN), or as a nitrogen nucleophile to yield isocyanides (R-NC).221 Traditionally, NaCN or KCN has been used to form nitriles, and AgCN has been used to form isocyanides.190 It has been rationalized that alkali metal cyanide dissociates to give "free" cyanide ion which attacks with its more basic carbon terminus, whereas AgCN does 157 not entirely dissociate, leaving Ag+ complexed to the carbon of the nucleophile, so that only nitrogen is available for nucleophilic attack.222 We hypothesized that the silicon could function analogously, so that C- complexation to silicon223 would leave the cyano nitrogen available to act as a nucleophile, as in Scheme 33. However, no traces of isonitrile products were detected in the reaction mixtures by 13C-NMR or IR. In addition, isonitriles have a characteristic stench even at low concentration; no foul odor was detected in the crude product. Scheme 33 Me Si MeMe F C N Br NCX :5 Exploration of the Use of Catalytic Amounts of Fluoride Having shown that a stoichiometric amount of TBAF induces trimethylsilyl cyanide to transfer cyanide nucleophile to a variety of primary and secondary substrates, studies were initiated to explore if TBAF could be used catalytically. As shown in Table 7, it was found that the displacement of benzyl bromide is not catalytic in fluoride. More interestingly, the reaction is not even stoichiometric in TBAF: note that 1.0 equiv of TBAF relative to the substrate resulted in 89% conversion to the nitrile product. This is consistent with observations that when less than 1.2 equiv of TBAF to 1.2 equiv of Me3SiCN (3) relative to 1.0 equiv of the substrate were employed, the reaction did not always go to completion. This is attributable to a number of factors, including: (1) hydrolysis of Me3SiCN by the water present in commercial grade TBAF (5% H2O), resulting in formation of Me3SiOH; and (2) the consumption of TBAF by reaction with the Me3Si-F by-product, resulting in the formation of Me3SiF2- Bu4N+. As described below, NMR studies confirm the formation of both R3SiOH and R3SiF2- Bu4N+ in mixtures of R3SiCN and TBAF. 158 Br CNMe3SiCN (3) / TBAF MeCN, 1h, 25 ?C 64 65 Entry TBAF (mol %) Yielda,b,c (%) 1 100 89 2 50 42 3 25 15 a Benzyl bromide (64, 1 equiv), Me 3SiCN (3, 1.5 equiv), and thegiven amount of TBAF in MeCN were allowed to react at room temperature. b Reactions were complete by 1 h. c Yield of 65 determined by G.C. analysis of the crude reaction mixture (vs. an internal standard). Table 7. Reaction of Benzyl Bromide (64) with Trimethylsilyl Cyanide (3) and Varying Amounts of TBAF. As depicted in Scheme 34, in order for the reaction to be substoichiometric in TBAF, the leaving group anion Bu4N+Br- must itself become the activating catalyst by attacking Me3SiCN to form a hypercoordinate bromo cyanosilicate reactive intermediate 88. In effect, TBAF would act as an initiator, not as a catalyst. Fluoride catalysis is unlikely, because regeneration of fluoride would require the cleavage of the strong Si?F bond and formation of a much weaker Si?Br bond. Scheme 34 CN Si Br Me MeMeMe3SiCN Ph BrPh CN Me3Si-CN 3 CN Si F Me MeMeF - Ph Br Ph CN Me3Si-F Me3Si-Br Bu4N+ Br- 885 + + 3 159 With the aim of testing the ability of bromide or other typical leaving group anions to activate Me3SiCN, 1 equiv of benzyl bromide was treated with 1.5 equiv of Me3SiCN and 1.5 equiv of a series of TBAX salts (X = Br, Cl, I, OTF). The results are depicted in Table 8. As anticipated, bromide, chloride, iodide and triflate ion (Table 8, entries 2-5) were not sufficiently nucleophilic to activate Me3SiCN. When a more basic oxygen nucleophile was employed (tetrabutylammonium acetate, entry 6), phenylacetonitrile was formed, however the reaction conversion was poor. These results are explained by the relative Si?X and Si?O bond strengths, and indicate that there is a strong thermodynamic driving force behind the formation of the Si?F bond. The Si?F bond energy (582 kJ?mol-1) is significantly greater than the Si?O, Si?Cl, Si?Br, or Si?I bond energies (368, 391, 310, and 234 kJ?mol?1, respectively). Br CNMe3SiCN (3) / X- Bu4N+ 64 65 MeCN reflux Entry Nucleophile Time (h) Yield (%)a,b 1 F- Bu4N+ 0.1 >95 2 Br- Bu4N+ 24 0 3 Cl- Bu4N+ 24 0 4 I- Bu4N+ 24 0 5 TfO- Bu4N+ 24 0 6 AcO- Bu4N+ 0.1 45c a Benzyl bromide (64) (1 equiv), Me 3SiCN (1.5 equiv), and the indicatednucleophile (1.5 equiv) in MeCN were allowed to react at reflux. b The yield of 65 was determined by G.C. analysis of the crude reaction mixture (vs. an internal standard). c Reaction was allowed to proceed for 24 h, but no additional progress was detected. Table 8. Reaction of Benzyl Bromide (64) and Trimethylsilyl Cyanide (3) with Tetrabutylammonium Nucleophiles. 160 Investigation of Alternative Fluoride Sources As illustrated in Table 8, fluoride is the unrivaled species for the activation of Me3SiCN. In an effort to fully optimize this methodology, a number of other fluoride sources were surveyed in a standard displacement reaction. It was believed that the hydroxide contaminant in TBAF, and/or the inherent basicity of TBAF could contribute to the formation of elimination products, and that other less basic fluoride sources could reduce alkene formation. In addition, alkali metal salts of fluoride are significantly less expensive than TBAF. In order to test the ability of other common fluoride sources to activate Me3SiCN, 1 equiv of benzyl bromide was treated with 1.5 equiv of Me3SiCN and 1.5 equiv of a series of fluoride salts (KF, CsF, and TBAT, 1). The order of fluoride reactivity toward the primary substrate was as follows: TBAF ? TBAT >> KF/18-crown-6 > KF ? CsF. Under heterogeneous conditions, the reaction was sluggish, and generally failed to go to completion (Table 9, entries 2 and 4). Addition of a phase transfer catalyst to KF increased the overall yield and shortened the reaction time (entry 3). Br CNMe3SiCN (3) / F 64 65 MeCN reflux Entry Fluoride Salt Time (h) Yield (%)a,b 1 Bu4N+ F- >1 100 2 KF 72 80c 3 KF / 18-crown-6d 48 100 4 CsF 48 75c 5 TBAT, 1 >1 100 a Benzyl bromide (64) (1.0 equiv), Me 3SiCN (3) (1.5 equiv), and the indicated fluoride salt(1.5 equiv) in MeCN were allowed to react at reflux. b The yield of 65 was determined by G.C. analysis of the crude reaction mixture (vs. an internal standard). c Reaction was allowed to proceed for and additional 24 hours, but no further progress was detected. d Catalytic in 18-crown-6 (20 mol %). Table 9. Reaction of Benzyl Bromide (64) and Trimethylsilyl Cyanide (3) with Alternative Fluoride Sources. 161 TBAT and TBAF both rapidly induced the reaction (Table 9, entries 1 and 5). While both reagents performed similarly in the above displacement reaction with benzyl bromide, TBAT potentially could outperform TBAF in the reaction with secondary substrates. As described previously, TBAT is both less basic and more anhydrous than TBAF, and in theory should suppress elimination. However, effectively no reactivity difference was observed in the displacement of secondary alkyl substrates, regardless of the leaving group (Br, I, or OTs)(Table 10). X CN 77 Me3SiCN (3) / F MeCN reflux Entry Substrate Fluoride Source Time (h) Yielda,b,c (%) 1 CH3(CH2)5CH(I)CH3 76 TBAF TBAT, 1 1 1 83 85 2 CH3(CH2)5CH(Br)CH3 78 TBAF TBAT, 1 2 3 82 86 3 CH3(CH2)5CH(OTs)CH3 79 TBAF TBAT, 1 1 1 76 74 a The yield of 77 was determined by G.C. analysis of the crude reaction mixture (vs. an internal standard). b All reactions went to completion; the remaining yield was 2-octene. c The remaining yield was 2-octene. Table 10. Reaction of Secondary Alkyl Halides with Trimethylsilyl Cyanide (3) and TBAF or TBAT (1). Proposed Mechanism Two feasible mechanistic pathways for the Me3SiCN/TBAF cyanide displacement reaction are depicted in Scheme 35. Based on standard silicate chemistry (vide supra), either path begins with attack of fluoride on silicon to form the hypercoordinate cyano fluorosilicate 5. One potential mechanism subsequently involves direct transfer of cyanide from silicate 5, without formation of a free cyanide nucleophile (route A). The second proposed mechanism entails the overall disproportionation of intermediate 5 to 162 tetrabutylammonium cyanide (Bu4N+CN-, 90) and Me3SiF (89); Bu4N+CN- would then act as the cyanide source (route B). Scheme 35 Me3SiCN CN Si F Me MeMeTBAF 53 Bu4N+ - Me3SiF 89 Bu4N+CN- 90 R Br R CN R Br R CN route A route B The cyanide anion delivered by silicate 5 should have a different reactivity than its disproportionation product, Bu4N+CN-. In order to determine a reactivity difference, and to ultimately determine the reaction mechanism, commercially available Bu4N+CN- was purchased, and its reactivity compared to the Me3SiCN/TBAF reagent. The results are summarized in Table 11. As seen in Table 11, Bu4N+CN- reacted 1.5 to 9 times faster than the cyanosilicate reagent 5 with a variety of primary and secondary alkyl halides and sulfonates to form the corresponding nitrile in good yield. The product yields were identical for both methods. As demonstrated by entries 1 through 4, the order of reactivity for both systems with primary alkyl halides was I ? OTs > Br > Cl. In each case, Bu4N+CN- reacted more rapidly than the silicate system. Entry 3 is particularly illustrative: where the Me3SiCN/TBAF reaction only reached 33% conversion after 96 hours, the Bu4N+CN- reaction was complete within 36 hours. The secondary substrates in entries 5-8 showed a less dramatic difference in reaction time, and comparable product yields. Finally, norbornyl bromide (82, entry 9) was equally as unreactive with Bu4N+CN- as with Me3SiCN/TBAF in refluxing acetonitrile. In contrast, both the silicate and the Bu4N+CN- methods furnished norbornane carbonitrile (83) in refluxing dioxane. 163 R X R CN Me3SiCN (3) / TBAF or Bu4N+CN- (90) MeCN Entry Substrate Product Reagent Temp (?C) Time (h) Yielda,b (%) 1 CH3(CH2)11I 71 CH3(CH2)11CN 72 3/TBAF Bu4N+CN- 25 25 6 1 95 95 2 CH3(CH2)11Br 73 CH3(CH2)11CN 72 3/TBAF Bu4N+CN- 25 25 36 3 >95 >95 3 CH3(CH2)11Cl 74 CH3(CH2)11CN 72 3/TBAF Bu4N+CN- 25 25 96 36 33c >95 4 (C6H5)CH2CH2Br 67 (C6H5)CH2CH2CN 68 3/TBAF Bu4N+CN- 25 25 32 2 >95 >95 5 CH3(CH2)5CH(I)CH3 76 CH3(CH2)5CH(CN)CH3 77 3/TBAF Bu4N+CN- 25 25 72 24 68 71 6 CH3(CH2)5CH(Br)CH3 78 CH3(CH2)5CH(CN)CH3 77 3/TBAF Bu4N+CN- 25 25 120 96 79 77 7 CH3(CH2)5CH(OTs)CH3 79 CH3(CH2)5CH(CN)CH3 77 3/TBAF Bu4N+CN- 25 25 48 24 70 74 8 (C6H5)CH(Br)CH3 80 (C6H5)CH(CN)CH3 81 3/TBAF Bu4N+CN- 25 25 5 3 >95 >95 9 Br 82 CN 83 7:1 endo : exo 3/TBAF Bu4N+CN- 82 101e 82 101e 48 96 48 72 0d 70 0d 60 a Where 3/TBAF is the reagent, the indicated substrate, Me 3SiCN (3), and TBAF (1:1.5:1.5molar ratio) in acetonitrile were allowed to react at the given temperature, unless otherwise noted. Where Bu4N+CN- (90) is the reagent, the indicated substrate, and Bu4N+CN- (1:1.5 molar ratio) in acetonitrile were allowed to react at the given temperature. b Yield determined by G.C. analysis of the crude reaction mixture (vs. an internal standard). c Reaction was stopped before completion. d No reaction was observed. e Reaction was performed in dioxane. Table 11. Comparison of Nitrile Synthesis Using Me3SiCN/TBAF or Bu4N+CN-. 164 The observed reactivity difference between the two methods suggests that the displacement reaction does not involve a simple disproportionation of cyanosilicate 5 to form Bu4N+CN- (90) in situ (Scheme 36, eq. a). Scheme 36 Me Si MeMe CN F 5 Bu4N+ CN-Bu4N+ + Me3Si-F Ph Si PhPh F F TBAT, 1 Bu4N+ F -Bu4N+ + Ph3Si-F (a) (b) 89 2 90 Prior research with the analogous fluorosilicate TBAT (1) mirrors these results. Experimental evidence supported the conclusion that TBAT does not simply disproportionate to generate TBAF in situ (Scheme 36 above, eq. b), but rather directly delivers a softer, less basic fluoride nucleophile. For example, Pilcher demonstrated that the hypercoordinate fluorosilicate TBAT performed fluoride displacements more slowly, and gave less elimination than its disproportionation product, TBAF (Scheme 37).48,55 Scheme 37 OTs TBAT, 1 MeCN, reflux F + OTs TBAF MeCN, 25 ?C F + 98% 2% 58% 32% 24 h 1 h Having determined that Bu4N+CN- was not the reactive intermediate, investigations into a mechanism involving hypercoordinate cyanosilicate 5 as the direct source of cyanide were undertaken. In addition, the participation of other silicon species such as dicyano trimethylsilicate (91), or difluoro trimethylsilicate (92) was investigated. 165 CN Si CN Me MeMe Bu4N+ F Si F Me MeMe Bu4N+ 91 92 F Si CN Me MeMe Bu4N+ 5 Chris Handy of the DeShong group endeavored to obtain spectroscopic evidence supporting the existence of hypercoordinate cyanosilicate intermediates such as 5, 91 and 92.224 Triphenylsilyl cyanide (Ph3SiCN, 93, Scheme 38) was chosen for study rather than trimethylsilyl cyanide (3): given that hypercoordinate silicates are inductively stabilized by aryl substituents (vide supra), it was expected that upon treatment with 1 equiv of TBAF, a greater amount of the more stable hypercoordinate intermediate 94 would form (Scheme 38). Also, in contrast to the non-isolable trimethylsilyl silicates 91 and 92, authentic samples of the triphenylsilyl silicate analogs (95 and 1, respectively) were readily prepared and characterized. Scheme 38 Ph3Si TBAF 93 CN Si PhPh F F 1, TBAT Ph Bu4N+ Si PhPh CN CN 95 Ph Bu4N + Si PhPh CN F 94 Ph Bu4N + For reference, Table 12 summarizes the 29Si NMR spectral data for all authentic samples prepared by Handy. It was expected that any four- or five-coordinate silicate species formed in the reaction would be detected the ranges 45 to ?115 ppm or ?70 to ?200 ppm, respectively (vide infra).93 166 Compound 29Si (ppm) JSi-F (Hz) Ph3SiCN, 93 -28.2, sb - Ph3SiF, 2 -4.0, d 282 [Ph3SiF2]-[NBu4]+, 1 -106.3, t 251 [Ph3Si(CN)2]-[NBu4]+, 95 -118.0 - Ph3SiOH, 96 -16.9, s - Ph3Si-O-SiPh3, 97 -18.2, s - a Value for the phenyl carbon directly attached to silicon. b Spectrum in THF. c Value for the nitrile carbon. Table 12. 29Si-NMR Spectral Data for Authentic Samples Used In Mechanistic Studies (Room Temperature, CDCl3).47-49,55 Treatment of Ph3SiCN (93) with 1 equiv of TBAF led to the immediate disappearance of the starting material signal and formation of hydrolyzed silane [Ph3SiOH (96) and its ether Ph3SiOSiPh3 (97)]; at room temperature, no other peaks were observed in the 29Si NMR spectrum (Figure 6, spectrum b). The silicon hydrosylates 96 and 97 can be attributed to the reaction of the water present in TBAF with the starting silane 93 and/or silicate derivatives 94 and 95. Handy conclusively demonstrated that the source of these silicon hydrosylates was likely to be cyano fluorosilicate 94 and/or dicyanosilicate 95. Treatment of Ph3SiCN (93) with water did not result in formation of either silanol 96 or its silyl ether 97. In contrast, an authentic sample of dicyanosilicate 95 underwent immediate conversion to hydrolyzed silanes 96 and 97. It was concluded that silicate formation was a prerequisite for the hydrolysis process. In order to probe the existence of a rapid equilibrium, Handy performed the same NMR experiment at -30?C. As before, treatment of Ph3SiCN (93) with 1 equiv of TBAF led to the disappearance of the starting material signal and formation of hydrolyzed silane; however at reduced temperature, a peak corresponding to Ph3SiF2- (TBAT, 1) also was observed (Figure 6, spectrum c). In the 13C NMR spectrum, neither Bu4N+CN- (90) nor the 167 starting materials Ph3SiCN (93) and TBAF were observed at any temperature, implying the rapid and possibly irreversible formation of intermediate 94 at the reaction outset. 29Si spectrum with inverse gated proton decoupling, recorded at 99 MHz on a Bruker AMX 500 spectrometer (D1= 10 sec, P1= 9.00 ?sec, no spinning). a: Ph3SiCN (93) in THF, 25 ?C, 98 scans; b: Ph3SiCN (93) + 1 equiv TBAF in THF, 25 ?C, 250 scans; c: Ph3SiCN (93) + 1 equiv TBAF in THF, -30 ?C, 400 scans. The broad peak appearing in the range -90 to -120 ppm in all 29Si spectra is present in all samples including solvent blanks, and is likely due to the use of silicon-based adhesives in the assembly of the probe. (Reprinted by permission of C. J. Handy) Figure 6. Triphenylsilyl Cyanide (93) and TBAF (1 equiv), 29Si Spectrum. Formation of TBAT in 1:1 mixtures of Ph3SiCN and TBAF, without the formation of Bu4N+CN- or Ph3SiF would imply a mechanism involving disproportionation of two molecules of the hypercoordinate cyano fluorosilicate 94 to dicyanosilicate 95 and TBAT (1) (Scheme 39). The absence of signals corresponding to a number of silicates, most especially the key intermediates cyano fluorosilicate 94 and dicyanosilicate 95, does not disprove their existence, but rather suggests that these hypercoordinate silicates are significantly less stable than TBAT, and may exist fleetingly or undergo rapid hydrolysis 168 to the silanol Ph3SiOH (91). Ultimately, the process is driven at least in part by formation of the highly stable difluorosilicate 1. In addition, the fact that dicyanosilicate 90 can be prepared and observed by IR and NMR spectroscopy, whereas the mixed cyano fluorosilicate 94 cannot suggests that the equilibrium is further driven by formation of symmetrical dicyanosilicate 95. Scheme 39 +2 Si Ph Ph CN F 94 Ph Bu4N+ Si PhPh CN CN 95 Ph Bu4N+ Si PhPh F F 1, TBAT Ph Bu4N+ In summary, no direct evidence of production of hypercoordinate intermediate 94 was provided by the NMR studies performed by Handy, however the spectroscopic data strongly points away from a simple in situ generation of Bu4N+CN-. The NMR study indirectly points to a process involving rapid disproportionation of unstable cyano fluorosilicate intermediate 94 to the more stable dicyano and difluoro silicates (95 and 1, respectively, Scheme 39, above). In theory when this reaction is performed in the presence of a displacement substrate, either hypercoordinate cyanosilicate 94 or 95 could function as the source of nucleophilic cyanide (Scheme 40). However the disproportionation of 94 to dicyanosilicate 95 and TBAT is very rapid by NMR spectroscopy, in contrast to the rate of displacement of alkyl halides. Thus, silicate 95 is strongly implicated as the reactive intermediate in this process. Scheme 40 94 TBAT, 1 fast slowslow slow2 Si Ph Ph CN F Ph Si PhPh CN CN 95 Ph2 R X R X R CNR CN 169 Scheme 41 depicts a plausible mechanism for the Me3SiCN/TBAF displacement reaction that takes into account the following conclusions based on the NMR data, experimental observations, and literature precedence: (1) the mechanism involves as its first step the transient formation of the hypercoordinate cyano fluorosilicate 5 followed by rapid disproportionation to dicyanosilicate Me3Si(CN)2- Bu4N+ (91) and Me3SiF2- Bu4N+ (92); (2) Bu4N+CN- (90) at no time is generated in significant quantity; (3) TBAF is rapidly consumed at the reaction outset; (4) the leaving group nucleophile does not participate in the reaction sequence; and (5) only slightly more than 1 equiv of the Me3SiCN/TBAF mixture is required for reaction completion, most likely due to water contamination. Scheme 41 CN Si F Me MeMe F Si CN Me MeMe Bu4N+ F- + fast 5 5 Me Si MeMe CN CN91 Me Si Me Me F F92 92 Me3SiCN 3 R X R CN Me3SiCN, 3 Bu4N+Br- The above proposed mechanism begins with rapid formation and disproportionation of cyano fluorosilicate intermediate 5 to generate Me3SiF2- (92) and dicyanosilicate 91. Hypercoordinate silicate 91 transfers cyanide nucleophile to the substrate to form alkylnitrile, and to regenerate Me3SiCN (3). The Me3SiCN thus produced undergoes rapid fluoride activation by Me3SiF2- (92) to form a second molecule of cyano fluorosilicate 5, which again disproportionates to Me3SiF2- (92) and a second molecule of dicyanosilicate 91. A notable feature of the above proposed mechanism is Me3SiF2- (92) is required to function as a fluoride source; otherwise 2 equiv of TBAF to 1 equiv of Me3SiCN would be required for reaction completion. Silicate Me3SiF2- Bu4N+ (92) should be an excellent fluorinating species in analogy to the powerful fluorinating reagent Me3SiF2- (Me3N)S+ (TAS-F, 8). 170 F Si F Me Me Me N S N N TASF, 8 Bu4N+ 92 F Si F Me Me Me Conclusions Treatment of trimethylsilyl cyanide with TBAF resulted in the in situ generation of a hypercoordinate cyanosilicate. The reactive silicate has been shown to be highly effective as a nucleophilic cyanide donor under extremely mild conditions in contrast to traditional cyanide reagents. Primary and secondary alkyl halides and sulfonates undergo rapid and efficient cyanide displacement in the absence of phase transfer catalysts with the silicate methodology. The Me3SiCN/TBAF system is significantly less reactive and less basic than TBA-CN, therefore the mechanism of reaction most like involves the in situ generation of a hypercoordinate cyanosilicate, rather than disproportionation of Me3SiCN and TBAF to form TBA-CN in situ. Epilogue Since the publication of the above method for azide and cyanide displacements via hypercoordinate silicate intermediates in 1999,52-54 a number of papers have appeared in the literature225 which report using our technique, primarily for the stereocontrolled synthesis of azido sugars,226-233 and for the nucleophilic ring opening of aziridines.227,234 Hou reported the ring-opening reactions of aziridines 98 with trimethylsilyl compounds triggered by TBAF to give the corresponding products 99 and 100 regioselectively in excellent yield (Scheme 42).234 TBAF functions to simply initiate the reaction: the amide leaving group is itself nucleophilic enough to attack silicon, and cleave the Me3SiX bond. 171 Scheme 42 NHTs R NHTsR XR X NHTs Me3SiX (X = N3, CN) TBAF (5 mol %) + R = H, X = N3, >99% 99 R = H, X = CN, 95% 99 R = CH3, X = N3, 63% (>99 : 1) (99:100) R = CH3, X = CN, 79% 99 98 99 100THF Aggarwal demonstrated a potential limitation of the Me3SiX/TBAF strategy: although this method is relatively mild, highly base-sensitive substrates can react with TBAF to yield undesired elimination products.227 As illustrated in Scheme 43, eq. a, attempts by Aggarwal to isolate the ring-opened nitrile 102 were thwarted by formation of alkene 103. Control experiments by Aggarwal implicated TBAF as the cause of elimination (Scheme 43, eq. b): exposure of an authentic sample of the ring-opened product 104 to TBAF promoted elimination. Regrettably, Aggarwal did not try substituting the less basic TBAT for TBAF as the fluoride source, to preempt unwanted elimination. Scheme 43 N Ph SiMe3 Ts TsHN Nu SiMe3 Ph Nu101 104 TBAF Ph Nu 105 (a) (b) TsHN CN SiMe3 Ph Ph CNTBAF (5 mol %) THF 102 103 Me3SiCN, 3 Preliminary reports indicate that mannose and glucose derivatives are poor substrates for cyanide displacement by hypercoordinate cyanosilicate 5 (Scheme 44, eq. a). This is in stark contrast to azide displacements by hypercoordinate azidosilicate 6: Soli demonstrated the synthesis of glycosyl azides from various glycosyl donors in high yields with excellent regio- and stereocontrol (Scheme 44, eq. b). 172 Scheme 44 O X RO F Si CN Me MeMe Bu4N+5 X O CNRO O N3 ROF Si N3 Me MeMe Bu4N+6 (a) (b) Prompted by the interest of the DeShong research group in methods for the synthesis of mannosyl glycoconjugates, the author and Handy performed investigative experiments which indicated that 6-O-tosyl and 1-?-bromo per-O-acetylated mannose derivatives (106 and 107, respectively) were poor substrates for cyanide displacement. Deacetylation and eventual decomposition of the starting material were observed, even for the primary tosyl sugar (106). O AcO AcO AcO OAc Br 107 O TsO AcO AcO OAc OAc 106 The sensitivity of sugar substrates has been confirmed by Gervay-Hague, using 1-?-iodo per-O-benzylated mannose derivative 108 (Scheme 45).235 When subjected to standard Me3SiCN/TBAF conditions, the more robust benzyl protecting groups of sugar 108 remained intact, however none of the desired cyano sugar 109 was obtained (Scheme 45, eq. a). Instead, the 1,2-elimination compound 110 was formed quantitatively (Scheme 45, eq. b). 173 Scheme 45 (a) (b) O BnO BnO BnO OBn I Me3SiCN Me3SiCN X O BnO BnO BnO OBn CN O BnO BnO BnO BnO 108 109 110 TBAF THF TBAF THF The method described herein for the in situ formation of reactive hypercoordinate fluorosilicates is potentially quite general; the Me3SiCN/TBAF methodology is a prototype for the delivery of other anions. For example, trimethylsilyl isocyanate (Me3Si(NCO)) is commercially available, and upon activation with TBAF has the potential of delivering the synthetically valuable190 isocyanate nucleophile via hypercoordinate silicate 111. Me Si Me Me OCN F Bu4N+ 111 Me Si Me Me F3C F Bu4N+ 113 Me Si Me Me F Bu4N+ 112 Preliminary studies by Poli indicate that cyclopentadienyl fluorosilicate 112 is a useful source of cyclopentadienyl anion for the synthesis of organomolybdenum complexes where other reagents fail; compound 112 is generated in situ by treatment of 5-(trimethylsilyl)-1,3-cyclopentadiene with TBAF. 236 Lastly, the utilization of Me3SiCF3/TBAF for the transfer of trifluoromethyl anion has recently been reported; the researchers attributed the relative mildness and low basicity of this method to the formation of hypercoordinate silicate intermediate 113.237-241 The scope of this methodology continues to be more precisely defined outside of the DeShong laboratories, as new reports of its application to complex natural product 174 synthesis appear in the literature.225,229,242 Although no further optimization of this technique is planned, the DeShong group does continue to employ this reaction for the synthesis of glycosyl azide derivatives.243 175 Experimental General. Thin layer chromatography (TLC) was performed on 0.25 mm Merck silica- coated glass plates treated with a UV-active binder, with the compounds being visualized in one or more of the following manners: UV (254 nm), iodine, or vanillin/sulfuric acid charring. Flash chromatography was performed using thick-walled glass columns and medium-pressure silica gel (Davisil? 200-425 mesh) as described by Still.244 Flash chromatography data is reported as: (column diameter in mm, column height in cm, solvent). Infrared spectra were recorded on a Nicolet 5DXC FT-IR spectrophotometer with the samples prepared as stated. Band positions are given in reciprocal centimeters (cm-1) and relative intensities are listed as br (broad), vs (very strong), s (strong), m (medium), or w (weak). Melting points were taken in Kimax soft glass capillary tubes using a Thomas- Hoover Uni-Melt capillary melting point apparatus (Model 6406K) equipped with a calibrated thermometer. Melting points are corrected. Nuclear magnetic resonance (1H, 13C) spectra were recorded on a Bruker DRX-400 spectrometer. Chemical shifts are reported in parts per million (?) downfield from tetramethylsilane (TMS). Coupling constants (J values) are reported in Hertz (Hz), and spin multiplicities are indicated by the following symbols: s (singlet), d (doublet), t (triplet), q (quartet), m (multiplet), br s (broad singlet). Low resolution (MS) and high resolution mass spectra (HRMS) were obtained on a VG-7070E magnetic sector instrument equipped with a 486 PC-based data system. GCMS was performed on a Shimadzu QP5000MS coupled with a GC17A gas chromatograph. Gas chromatography was performed on a Hewlett Packard 5890 GC equipped with a flame ionization detector using a 25 m methyl silicon column. 176 Methyl sulfoxide (DMSO), and acetonitrile (MeCN) were each distilled from calcium hydride. Dioxane was distilled from sodium-benzophenone ketyl. All reagents and alkyl halides were purchased from Acros or Aldrich and purified using the method of Perrin245 prior to use, or prepared by literature procedure as noted. The substrate 2-iodooctane246 was prepared according to the literature procedure. In the case of alkyl sulfonates 75 and 79, the appropriate sulfonate was synthesized from the alcohol immediately before use using the literature method247,248 and the crude isolated sulfonate was used without further purification. Tetrabutylammonium fluoride (TBAF) was purchased from Acros as a 1.0 M solution in THF and used as received. Tetrabutylammonium triphenyldifluorosilicate (TBAT) was prepared according to the literature procedure.48 Glassware used in the reactions described below was dried for a minimum of 12 h in an oven at 120 ?C. All reactions were run under an atmosphere of nitrogen or argon at room temperature unless otherwise noted. All compounds were determined to be >95% pure by 1H NMR unless otherwise noted. General Procedure for Synthesis of Nitriles Using Trimethylsilyl Cyanide (Me3SiCN, 3) and Tetrabutylammonium Fluoride (TBAF) (Table 6). Trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAF (1 M in THF, 1.5 mL, 1.5 mmol) were added to a stirring solution of the alkyl halide (1.00 mmol, 1 equiv) in 10 mL solvent (MeCN or dioxane) under an atmosphere of nitrogen. Unless otherwise noted, the reaction was performed at the given temperature until GC analysis indicated that the starting material had been consumed. The reaction mixture was concentrated in vacuo, and the resulting syrup was purified by flash chromatography. 177 Si MeMeMe CN F Bu4N+ Me3SiCN TBAF R X R CN MeCN 53 Entry Substrate R-X Product R-CN Temp (?C) Time (h) Yielda.b (%) 1 (C6H5)CH2Br 64 (C6H5)CH2CN 65 82 25 0.1 1 95 (>95) 2 (C6H5)CH2Cl 70 (C6H5)CH2CN 65 82 25 2 72 (>95) (>95) 3 CH3(CH2)11I 71 CH3(CH2)11CN 72 82 25 0.1 6 95 (95) 4 CH3(CH2)11Br 73 CH3(CH2)11CN 72 82 25 2 36 95 (>95) 5 CH3(CH2)11Cl 74 CH3(CH2)11CN 72 82 25 3 96 95 (33)c 6 CH3(CH2)11OMs 75 CH3(CH2)11CN 72 82 0.1 95 7 (C6H5)CH2CH2Br 67 (C6H5)CH2CH2CN 68 82 25 0.1 32 95 (>95) 8 CH3(CH2)5CH(I)CH3 76 CH3(CH2)5CH(CN)CH3 77 82 25 1 72 (83) (68) 9 CH3(CH2)5CH(Br)CH3 78 CH3(CH2)5CH(CN)CH3 77 82 25 2 120 (82) (79) 10 CH3(CH2)5CH(OTs)CH3 79 CH3(CH2)5CH(CN)CH3 77 82 25 1 48 (76) (70) 11 (C6H5)CH(Br)CH3 80 (C6H5)CH(CN)CH3 81 82 25 1 5 92 (>95) 12 Br 82 CN 83 7:1 endo : exo 82101e 4896 (0) d 70 13 Br 84 CN 85 82 48 (<5)f 14 Cl 86 CN 87 82 101e 48 96 (0)d (0)d a The indicated substrate, Me 3SiCN (3), and TBAF (1:1.5:1.5 molar ratio) in acetonitrile wereallowed to react at the given temperature, unless otherwise noted. b Isolated yield after purification. The yield determined by G.C. analysis of the crude reaction mixture (vs. an internal standard) is reported in parentheses. c Reaction was stopped before completion. d No reaction was observed. e Reaction was performed in dioxane. f Starting material was consumed; remaining yield assumed to be cyclohexene. Table 6. Reaction of Hypercoordinate Silicate 5 with Alkyl Halides. 178 CN Phenylacetonitrile (65) (Table 6, entry 1). The above general procedure for the synthesis of nitriles using trimethylsilyl cyanide and TBAF was followed using benzyl bromide (64) (171 mg, 119?L, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAF (1.5 mL, 1.5 mmol) in 10 mL of MeCN. The reaction was heated at reflux for 5 min. Flash chromatography (25 mm, 16 cm, 10% CH2Cl2/pentane) afforded 111 mg (95%) of 65 as a colorless oil. TLC Rf = 0.40 (30% CH2Cl2/pentane). IR (thin film) 3092 (s), 3069 (s), 3035 (s), 2912 (w), 2254 (m), 1604 (m), 1549 (m), 1497 (vs), 1416 (vs) cm-1; 1H NMR (CDCl3) ? 3.65 (s, 2H), 7.23- 7.35 (m, 5H); 13C NMR (CDCl3) ? 23.4, 118.0, 127.8, 127.9, 129.0, 130.0. The IR, 1H and 13C NMR data were identical to an authentic sample purchased from Aldrich, as well as published spectral data.218,249 Phenylacetonitrile (65) (Table 6, entry 2). The above general procedure for the synthesis of nitriles using trimethylsilyl cyanide and TBAF was followed using benzyl chloride (70) (127 mg, 115?L, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAF (1.5 mL, 1.5 mmol) in 10 mL of MeCN. The reaction was heated at reflux for 2 h. Flash chromatography (25 mm, 16 cm, 10% CH2Cl2/pentane) afforded 114 mg (97%) of 65 as a colorless oil. Spectral data is reported above. Tridecanenitrile (72) (Table 6, entry 5). The above general procedure for the synthesis of nitriles using trimethylsilyl cyanide and TBAF was followed using 1-chlorododecane (74) (205 mg, 236 ?L, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAF (1.5 mL, 1.5 mmol) in 10 mL of MeCN. The reaction was heated at reflux for 2 h. Flash chromatography (25 mm, 16 cm, 5% CH2Cl2/pentane) afforded 186 mg (95%) of 72. Spectral data is reported above. CN CN 179 CN Tridecanenitrile (72) (Table 6, entry 6). The above general procedure for the synthesis of nitriles using trimethylsilyl cyanide and TBAF was followed using methanesulfonic acid dodecyl ester (75) (264 mg, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAF (1.5 mL, 1.5 mmol) in 10 mL of MeCN. The reaction was heated at reflux for 5 min. Flash chromatography (25 mm, 16 cm, 5% CH2Cl2/pentane) afforded 186 mg (95%) of 72. Spectral data is reported above. Hydrocinnamonitrile (68) (Table 6, entry 7). The above general procedure for the synthesis of nitriles using trimethylsilyl cyanide and TBAF was followed using phenethyl bromide (67) (185 mg, 137 ?L, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAF (1.5 mL, 1.5 mmol) in 10 mL of MeCN. The reaction was heated at reflux for 5 min. Flash chromatography (25 mm, 16 cm, 10% CH2Cl2/pentane) afforded 125 mg (95%) of 68 as a colorless oil. TLC Rf = 0.35 (30% CH2Cl2/pentane). IR (thin film) 3090 (w), 3067 (m), 3032 (s), 2935 (m), 2869 (w), 2250 (w), 1606 (w), 1497 (s), 1455 (s), 1426 (s) cm-1; 1H NMR (CDCl3) ? 2.50 (t, J = 7.4, 2H), 2.85 (t, J = 7.4, 2H), 7.17-7.31 (m, 5H); 13C NMR (CDCl3) ? 19.1, 31.3, 119.2, 127.0, 128.2, 128.7, 138.1. The IR, 1H and 13C NMR data were identical to an authentic sample, as well as published spectral data.218,250 2-Methyloctanenitrile (77) (Table 6, entry 8). The above general procedure for the synthesis of nitriles using trimethylsilyl cyanide and TBAF was followed using 2-iodooctane (76) (240 mg, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAF (1.5 mL, 1.5 mmol) in 10 mL of MeCN. The reaction was heated at reflux for 1 h. Flash chromatography (25 mm, 16 cm, 10% CH2Cl2/pentane) afforded 116 mg (83%) of 77 as a colorless oil. TLC Rf = 0.65 (30% CH2Cl2/pentane). IR (thin film) 2949 (vs), 2855 (vs), 2240 (w), 1461 (m), 1451 (m), 1390 (m) cm-1; 1H NMR (CDCl3) ? 0.89 (t, J = 6.8, 3H), 1.22-1.40 (m, 8H), 1.31 (d, J = 7.0, 3H), 1.44-1.67 (m, 2H), 2.52-2.67 (m, 1H); 13C NMR (CDCl3) ? 13.8, 17.9, 22.4, 25.3, 26.8, 28.6, 31.4, 33.9, 122.9. The IR, 1H and 13C NMR data were identical to published spectral data.220,251,252 CN CN 180 CN CN CN 2-Methyloctanenitrile (77) (Table 6, entry 9). The above general procedure for the synthesis of nitriles using trimethylsilyl cyanide and TBAF was followed using 2-bromooctane (78) (193 mg, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAF (1.5 mL, 1.5 mmol) in 10 mL of MeCN. The reaction was heated at reflux for 1 h. Flash chromatography (25 mm, 16 cm, 10% CH2Cl2/pentane) afforded 114 mg (82%) of 77 as a colorless oil. Spectral data is reported above. 2-Methyloctanenitrile (77) (Table 6, entry 10). The above general procedure for the synthesis of nitriles using trimethylsilyl cyanide and TBAF was followed using p-toluenesulfonic acid 1-methyl-heptyl ester (79) (284 mg, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAF (1.5 mL, 1.5 mmol) in 10 mL of MeCN. The reaction was heated at reflux for 1 h. Flash chromatography (25 mm, 16 cm, 10% CH2Cl2/pentane) afforded 114 mg (82%) of 77 as a colorless oil. Spectral data is reported above. 2-Phenylpropionitrile (81) (Table 6, entry 11). The above general procedure for the synthesis of nitriles using trimethylsilyl cyanide and TBAF was followed using 1-(bromoethyl)benzene (70) (185 mg, 136 ?L, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAF (1.5 mL, 1.5 mmol) in 10 mL of MeCN. The reaction was heated at reflux for 5 min. Flash chromatography (25 mm, 16 cm, 10% CH2Cl2/pentane) afforded 121 mg (92%) of 81 as a colorless oil. TLC Rf = 0.45 (30% CH2Cl2/pentane). IR (thin film) 3091 (m), 3068 (m), 3033 (s), 2987 (s), 2939 (m), 2878 (m), 2244 (m), 1453 (s) cm-1; 1H NMR (CDCl3) ? 1.56 (d, J = 7.3, 3H), 3.84 (q, J = 7.3, 1H), 7.28-7.34 (m, 5H); 13C NMR (CDCl3) ? 21.4, 31.1, 121.6, 126.7, 128.0, 129.1, 137.1. The IR, 1H and 13C NMR data were identical to an authentic sample, as well as published spectral data.218,250 181 endo/exo-2-Norbornane Carbonitrile (83) (Table 6, entry 12). The above general procedure for the synthesis of nitriles using trimethylsilyl cyanide and TBAF was followed using exo-2-bromonorbornane (82) (175 mg, 128 ?L, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAF (1.5 mL, 1.5 mmol) in 10 mL of dioxane. The reaction was heated at reflux for 96 h. Flash chromatography (25 mm, 16 cm, 10% CH2Cl2/pentane) afforded 85 mg (70%) of endo/exo-83 as a colorless oil. TLC Rf = 0.50 (30% CH2Cl2/pentane). IR (CCl4) 2970 (b), 2876 (s), 2238 (m), 1456 (m) cm-1; 1H NMR (CDCl3) ? 1.1-1.9 (m, 7.4H), 1.9-2.1 (m, 0.9H), 2.3-2.4 (m, 1.2H), 2.5 (s, 0.7H), 2.6 (s, 0.2H), 2.6-2.7 (m, 0.6H); 13C NMR (CDCl3) assigned as endo: ? 25.1 (C6), 29.2 (C5), 30.2 (C2), 35.5 (C3), 36.6 (C4), 38.9 (C7), 40.0 (C1), 123.0 (Cx); assigned as exo: ? 28.5, (C6), 28.6 (C5), 31.2 (C2), 36.1 (C3), 36.2 (C4), 37.3 (C7), 41.9 (C1), 123.7 (Cx). The IR, 1H and 13C NMR data were identical, except for the endo/exo ratio, to an authentic sample, as well as published spectral data.218,219,249 The endo/exo ratio was determined by integration of the 13C NMR spectrum: the relative areas of the carbon signals as shown above were: C6 (6.91:1.00), C1 (6.94:100), Cx (6.92:1.00), for an average endo/exo ratio of 6.92:1.0. Cyclohexanecarbonitrile (85) (Table 6, entry 13). The above general procedure for the synthesis of nitriles using trimethylsilyl cyanide and TBAF was followed using cyclohexyl bromide (84) (163 mg, 124 ?L, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAF (1.5 mL, 1.5 mmol) in 10 mL of MeCN. The reaction was heated at reflux for 48 h. GC analysis indicated complete consumption of the starting material, and formation of a trace amount (<5%) of cyclohexanecarbonitrile (85). Cyclohexanecarbonitrile (85) was identified by comparison of the GC retention time of an authentic sample. The remaining yield was assumed to be cyclohexene, which was neither observed by GC nor isolated, due to volatility. 1 2 34 7 6 5 CNx CN 182 General Procedure For the Reaction of Benzyl Bromide (64) with Trimethylsilyl Cyanide (3) and Varying Amounts of Tetrabutylammonium Fluoride (TBAF) (Table 7). All reactions were performed on a 1.0 mmol scale. Trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol), and the indicated amount of TBAF were added to a stirring solution of benzyl bromide (64) (171 mg, 119?L, 1.00 mmol) and the internal standard naphthalene (128 mg, 1.0 mmol, 1.0 equiv) in 10 mL MeCN under an atmosphere of nitrogen. The reaction was stirred at room temperature for 1 h, when GC analysis indicated that all reactions had ceased to progress. Reaction progress was monitored by GC analysis of aliquots of the quenched reaction mixture. GC response factors relative to the internal naphthalene standard were determined, and the observed percentages of products were normalized accordingly. Br CNMe3SiCN (3) / TBAF MeCN, 1h, 25 ?C 64 65 Entry TBAF (mol %) Yielda,b,c (%) 1 100 89 2 50 42 3 25 15 a Benzyl bromide (64, 1 equiv), Me 3SiCN (3, 1.5 equiv), and thegiven amount of TBAF in MeCN were allowed to react at room temperature. b Reactions were complete by 1 h. c Yield of 65 determined by G.C. analysis of the crude reaction mixture (vs. an internal standard). Table 7. Reaction of Benzyl Bromide (64) with Me3SiCN (3) and Varying Amounts of TBAF. 183 General Procedure for the Reaction of Benzyl Bromide (64) and Trimethylsilyl Cyanide (3) with Tetrabutylammonium Nucleophiles (Table 8). All reactions were performed on a 1.0 mmol scale. Trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol), and the indicated tetrabutylammonium salt (1.5 mmol, 1.5 equiv) were added to a stirring solution of benzyl bromide (64) (119?L, 1.00 mmol) and the internal standard naphthalene (128 mg, 1.0 mmol, 1.0 equiv) in 10 mL MeCN under an atmosphere of nitrogen. The reaction was stirred at reflux until GC analysis indicated that the reaction had ceased to progress. Reaction progress was monitored by GC analysis of aliquots of the quenched reaction mixture. GC response factors relative to the internal naphthalene standard were determined, and the observed percentages of products were normalized accordingly. Br CNMe3SiCN (3) / X- Bu4N+ 64 65 MeCN reflux Entry Nucleophile Time (h) Yield (%)a,b 1 F- Bu4N+ 0.1 >95 2 Br- Bu4N+ 24 0 3 Cl- Bu4N+ 24 0 4 I- Bu4N+ 24 0 5 TfO- Bu4N+ 24 0 6 AcO- Bu4N+ 0.1 45c a Benzyl bromide (64) (1 equiv), Me 3SiCN (1.5 equiv), and the indicatednucleophile (1.5 equiv) in MeCN were allowed to react at reflux. b The yield of 65 was determined by G.C. analysis of the crude reaction mixture (vs. an internal standard). c Reaction was allowed to proceed for 24 h, but no additional progress was detected. Table 8. Reaction of Benzyl Bromide (64) and Trimethylsilyl Cyanide (3) with Tetrabutylammonium Nucleophiles. 184 General Procedure for the Reaction of Benzyl Bromide (64) and Trimethylsilyl Cyanide (3) with Alternative Fluoride Sources (Table 9). All reactions were performed on a 1.0 mmol scale. Trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol), and the indicated fluoride salt (1.5 mmol, 1.5 equiv) were added to a stirring solution of benzyl bromide (64) (119?L, 1.00 mmol) and the internal standard naphthalene (128 mg, 1.0 mmol, 1.0 equiv) in 10 mL MeCN under an atmosphere of nitrogen. The reaction was stirred at reflux until GC analysis indicated that the reaction had ceased to progress. Reaction progress was monitored by GC analysis of aliquots of the quenched reaction mixture. GC response factors relative to the internal naphthalene standard were determined, and the observed percentages of products were normalized accordingly. Br CNMe3SiCN (3) / F 64 65 MeCN reflux Entry Fluoride Salt Time (h) Yield (%)a,b 1 Bu4N+ F- >1 100 2 KF 72 80c 3 KF / 18-crown-6d 48 100 4 CsF 48 75c 5 TBAT, 1 >1 100 a Benzyl bromide (64) (1.0 equiv), Me 3SiCN (3) (1.5 equiv), and the indicated fluoride salt(1.5 equiv) in MeCN were allowed to react at reflux. b The yield of 65 was determined by G.C. analysis of the crude reaction mixture (vs. an internal standard). c Reaction was allowed to proceed for and additional 24 hours, but no further progress was detected. d Catalytic in 18-crown-6 (20 mol %). Table 9. Reaction of Benzyl Bromide (64) and Trimethylsilyl Cyanide (3) with Alternative Fluoride Sources. 185 General Procedure for the Reaction of Secondary Alkyl Halides with Trimethylsilyl Cyanide (3) and TBAT (1) (Table 10). All reactions were performed on a 1.0 mmol scale. Trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol), and TBAT (1) (874 mg, 1.5 mmol) were added to a stirring solution of alkyl halide (1.00 mmol, 1 equiv) and the internal standard naphthalene (128 mg, 1.0 mmol, 1.0 equiv) in 10 mL MeCN under an atmosphere of nitrogen. The reaction was performed at 82 ?C until GC analysis indicated that the reaction had ceased to progress. Reaction progress was monitored by GC analysis of aliquots of the quenched reaction mixture. GC response factors relative to the internal naphthalene standard were determined, and the observed percentages of products were normalized accordingly. The elimination product 2-octene was identified by comparison of the GC retention time to the value of that of an authentic sample X CN 77 Me3SiCN (3) / F MeCN reflux Entry Substrate Fluoride Source Time (h) Yielda,b,c (%) 1 CH3(CH2)5CH(I)CH3 76 TBAF TBAT, 1 1 1 83 85 2 CH3(CH2)5CH(Br)CH3 78 TBAF TBAT, 1 2 3 82 86 3 CH3(CH2)5CH(OTs)CH3 79 TBAF TBAT, 1 1 1 76 74 a The yield of 77 was determined by G.C. analysis of the crude reaction mixture (vs. an internal standard). b All reactions went to completion; the remaining yield was 2-octene. c The remaining yield was 2-octene. Table 10. Reaction of Secondary Alkyl Halides with Trimethylsilyl Cyanide (3) and TBAF or TBAT (1). 186 2-Methyloctanenitrile (77) (Table 10, entry 1). The above general procedure for the reaction of secondary alkyl halides with trimethylsilyl cyanide and TBAT was followed using 2-iodooctane (76) (240 mg, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAT (874 mg, 1.50 mmol) in 10 mL of MeCN. The reaction was heated at reflux for 1 h to yield 85% of 77 by GC analysis. 2-Methyloctanenitrile (77) (Table 10, entry 2). The above general procedure for the reaction of secondary alkyl halides with trimethylsilyl cyanide and TBAT was followed using 2-bromooctane (78) (193 mg, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAT (874 mg, 1.50 mmol) in 10 mL of MeCN. The reaction was heated at reflux for 3 h to yield 86% of 77 by GC analysis. 2-Methyloctanenitrile (77) (Table 10, entry 3). The above general procedure for the reaction of secondary alkyl halides with trimethylsilyl cyanide and TBAT was followed using toluene-4-sulfonic acid 1-methyl-heptyl ester (79) (284 mg, 1.00 mmol), trimethylsilyl cyanide (149 mg, 200 ?L, 1.50 mmol) and TBAT (874 mg, 1.50 mmol) in 10 mL of MeCN. The reaction was heated at reflux for 1 h to yield 74% of 77 by GC analysis. General Procedure for Synthesis of Nitriles Using Tetrabutylammonium Cyanide (90) (Table 11). All reactions were performed on a 1.0 mmol scale. The indicated alkyl halide (1.00 mmol, 1 equiv) was added to a stirring solution of tetrabutylammonium cyanide (90) (403 mg, 1.50 mmol) and the internal standard naphthalene (128 mg, 1.0 mmol, 1.0 equiv) in 10 mL MeCN under an atmosphere of nitrogen. The reaction was stirred at the given temperature until GC analysis indicated that the reaction had ceased to progress. Reaction progress was monitored by GC analysis of aliquots of the quenched reaction mixture. GC response factors relative to the internal naphthalene standard were determined, and the observed percentages of products were normalized accordingly. CN CN CN 187 R X R CN Me3SiCN (3) / TBAF or Bu4N+CN- (90) MeCN Entry Substrate Product Reagent Temp (?C) Time (h) Yielda,b (%) 1 CH3(CH2)11I 71 CH3(CH2)11CN 72 3/TBAF Bu4N+CN- 25 25 6 1 95 95 2 CH3(CH2)11Br 73 CH3(CH2)11CN 72 3/TBAF Bu4N+CN- 25 25 36 3 >95 >95 3 CH3(CH2)11Cl 74 CH3(CH2)11CN 72 3/TBAF Bu4N+CN- 25 25 96 36 33c >95 4 (C6H5)CH2CH2Br 67 (C6H5)CH2CH2CN 68 3/TBAF Bu4N+CN- 25 25 32 2 >95 >95 5 CH3(CH2)5CH(I)CH3 76 CH3(CH2)5CH(CN)CH3 77 3/TBAF Bu4N+CN- 25 25 72 24 68 71 6 CH3(CH2)5CH(Br)CH3 78 CH3(CH2)5CH(CN)CH3 77 3/TBAF Bu4N+CN- 25 25 120 96 79 77 7 CH3(CH2)5CH(OTs)CH3 79 CH3(CH2)5CH(CN)CH3 77 3/TBAF Bu4N+CN- 25 25 48 24 70 74 8 (C6H5)CH(Br)CH3 80 (C6H5)CH(CN)CH3 81 3/TBAF Bu4N+CN- 25 25 5 3 >95 >95 9 Br 82 CN 83 7:1 endo : exo 3/TBAF Bu4N+CN- 82 101e 82 101e 48 96 48 72 0d 70 0d 60 a Where 3/TBAF is the reagent, the indicated substrate, Me 3SiCN (3), and TBAF(1:1.5:1.5 molar ratio) in acetonitrile were allowed to react at the given temperature, unless otherwise noted. Where Bu4N+CN- (90) is the reagent, the indicated substrate, and Bu4N+CN- (1:1.5 molar ratio) in acetonitrile were allowed to react at the given temperature. b Yield determined by G.C. analysis of the crude reaction mixture (vs. an internal standard). c Reaction was stopped before completion. d No reaction was observed. e Reaction was performed in dioxane. Table 11. Comparison of Nitrile Synthesis Using Me3SiCN/TBAF or Bu4N+CN-. 188 CN CN CN Tridecanenitrile (72) (Table 11, entry 1). The above general procedure for the synthesis of nitriles using tetrabutylammonium cyanide was followed using 1-iodododecane (71) (296 mg, 247 ?L, 1.00 mmol), and tetrabutylammonium cyanide (404 mg, 1.50 mmol) in 10 mL of MeCN. The reaction was stirred at room temperature for 1 h to yield 95% of 72 by GC analysis. Tridecanenitrile (72) (Table 11, entry 2). The above general procedure for the synthesis of nitriles using tetrabutylammonium cyanide was followed using 1-bromododecane (73) (249 mg, 240 ?L, 1.00 mmol), and tetrabutylammonium cyanide (404 mg, 1.50 mmol) in 10 mL of MeCN. The reaction was stirred at room temperature for 3 h to yield 97% of 72 by GC analysis. Tridecanenitrile (72) (Table 11, entry 3). The above general procedure for the synthesis of nitriles using tetrabutylammonium cyanide was followed using 1-chlorododecane (74) (205 mg, 236 ?L, 1.00 mmol), and tetrabutylammonium cyanide (404 mg, 1.50 mmol) in 10 mL of MeCN. The reaction was stirred at room temperature for 36 h to yield 97% of 72 by GC analysis. Hydrocinnamonitrile (68) (Table 11, entry 4). The above general procedure for the synthesis of nitriles using tetrabutylammonium cyanide was followed using phenethylbromide (67) (185 mg, 137 ?L, 1.00 mmol), and tetrabutylammonium cyanide (404 mg, 1.50 mmol) in 10 mL of MeCN. The reaction was stirred at room temperature for 2 h to yield 100% of 68 by GC analysis. CN 189 CN CN CN CN 2-Methyloctanenitrile (77) (Table 11, entry 5). The above general procedure for the synthesis of nitriles using tetrabutylammonium cyanide was followed using 2-iodooctane (76) (240 mg, 1.00 mmol), and tetrabutylammonium cyanide (404 mg, 1.50 mmol) in 10 mL of MeCN. The reaction was stirred at room temperature for 24 h to yield 71% of 77 by GC analysis. 2-Methyloctanenitrile (77) (Table 11, entry 6). The above general procedure for the synthesis of nitriles using tetrabutylammonium cyanide was followed using 2-bromooctane (78) (193 mg, 1.00 mmol), and tetrabutylammonium cyanide (404 mg, 1.50 mmol) in 10 mL of MeCN. The reaction was stirred at room temperature for 96 h to yield 77% of 77 by GC analysis. 2-Methyloctanenitrile (77) (Table 11, entry 7). The above general procedure for the synthesis of nitriles using tetrabutylammonium cyanide was followed using toluene-4-sulfonic acid 1-methyl-heptyl ester (79) (284 mg, 1.00 mmol), and tetrabutylammonium cyanide (404 mg, 1.50 mmol) in 10 mL of MeCN. The reaction was stirred at room temperature for 24 h to yield 74% of 77 by GC analysis. 2-Phenylpropionitrile (81) (Table 11, entry 8). The above general procedure for the synthesis of nitriles using tetrabutylammonium cyanide was followed using (1-bromoethyl)benzene (80) (185 mg, 136 ?L, 1.00 mmol), and tetrabutylammonium cyanide (404 mg, 1.50 mmol) in 10 mL of MeCN. The reaction was stirred at room temperature for 3 h to yield 99% of 81 by GC analysis. 190 1 2 34 7 6 5 CNx endo/exo-2-Norbornanecarbonitrile (83) (Table 11, entry 9). The above general procedure for the synthesis of nitriles using tetrabutylammonium cyanide was followed using exo-2-bromonorbornane (82) (175 mg, 128 ?L, 1.00 mmol), and tetrabutylammonium cyanide (404 mg, 1.50 mmol) in 10 mL of dioxane. The reaction was stirred at reflux for 72 h to yield 60% of endo/exo- 83 by GC analysis. 191 References CHAPTER ONE 1. Hatanaka, Y.; Goda, K.-i.; Okahara, Y. Tetrahedron 1994, 50, 8301-8316. 2. Tonks, L.; Williams, J. M. J. J. Chem. Soc., Perkin Trans. 1 1998, 3637-3652. 3. Nicolaou, K. C.; Chakraborty, T. K.; Piscopio, A. D.; Minowa, N.; Bertinato, P. J. J. Am. Chem. Soc. 1993, 115, 4419-4420. 4. Kalivretenos, A.; Stille, J. K.; Hegedus, L. S. J. Org. Chem. 1991, 56, 2883-2894. 5. Gyorkos, A. C.; Stille, J. K.; Hegedus, L. S. J. Am. Chem. Soc. 1990, 112, 8465- 8472. 6. Smith III, A. B.; Wiu, Y.; Jones, D. R.; Kobayashi, K. J. Am. Chem. Soc. 1995, 117, 12011-12012. 7. Su, D.-S.; Meng, D.; Bertinato, P.; Balog, A.; Sorensen, E. J.; Danishefsky, S. J.; Zheng, Y.-H.; Chou, T.-C.; He, L.; Horwitz, S. B. Angew. Chem., Int. Ed. Engl. 1997, 36, 757-759. 8. Hatanaka, Y.; Hiyama, T. Synlett 1991, 845-853. 9. Tamao, K.; Kakui, T.; Akita, M.; Iwahara, T.; Kanatani, R.; Yoshida, J.; Kumada, M. Tetrahedron 1983, 39, 983-990. 10. Metal-Catalyzed Cross-Coupling Reactions; Diederich, F.; Stang, P. J., Eds.; Wiley-VCH: Weinheim, Germany, 1998. 11. Heck, R. F. Palladium Reagents in Organic Synthesis; Academic Press: New York, 1985. 12. Tsuji, J. Palladium Reagents and Catalysts. Innovations in Organic Synthesis; John Wiley & Sons: New York, 1995. 13. Trost, B. M.; Verhoeven, T. R. In Comprehensive Organometallic Chemistry; Wilkinson, G., Stone, F. G. A., Abel, E. W., Eds.; Pergamon: Oxford, 1982; Vol. 8, p 799-938. 14. Tamao, K. In Comprehensive Organic Synthesis; Trost, B. M., Fleming, I., Eds.; Pergamon: Oxford, 1991; Vol. 3, p 435-480. 15. Knight, D. W. In Comprehensive Organic Synthesis; Trost, B. M., Fleming, I., Eds.; Pergamon: Oxford, 1991; Vol. 3, p 481-578. 16. Casado, A. L.; Espinet, P. J. Am. Chem. Soc. 1998, 120, 8978-8985. 17. Casado, A. L.; Espinet, P. Organometallics 1998, 17, 954-959. 192 18. Stille, J. K.; Echavarren, A. M.; Williams, R. M.; Hendrix, J. A. Org. Synth. 1993, 71, 97-106. 19. Farina, V.; Krishnamurthy, V.; Scott, W. J. In Organic Reactions; Paquette, L. A., Ed.; John Wiley & Sons, Inc.: New York, 1997; Vol. 50. 20. Miyaura, N.; Suzuki, A. Chem. Rev. 1995, 95, 2457-2483. 21. Mowery, M. E.; DeShong, P. J. Org. Chem. 1999, 64, 1684-1688. 22. Mowery, M. E.; DeShong, P. J. Org. Chem. 1999, 64, 3266-3270. 23. Mowery, M. E.; DeShong, P. Org. Lett. 1999, 1, 2137-2140. 24. Handy, C. J.; Mowery, M. E.; DeShong, P. Pure Appl. Chem. 2000, 72, 1655- 1658. 25. Correia, R.; DeShong, P. J. Org. Chem. 2001, 66, 7159-7165. 26. Hiyama, T. In Metal-Catalyzed Cross-Coupling Reactions; Diederich, F., Stang, P. J., Eds.; Wiley-VCH: Weinheim, Germany, 1998. 27. Hiyama, T.; Hatanaka, Y. Pure Appl. Chem. 1994, 66, 1471-1478. 28. Hirabayashi, E.; Mori, A.; Kawashima, J.; Suguro, M.; Nishihara, Y.; Hiyama, T. J. Org. Chem. 2000, 65, 5342-5349. 29. Gouda, K.; Hagiwara, E.; Hatanaka, Y.; Hiyama, T. J. Org. Chem. 1996, 61, 7232- 7233. 30. Hagiwara, E.; Gouda, K.; Hatanaka, Y.; Hiyama, T. Tetrahedron Lett. 1997, 38, 439-442. 31. Hirabayashi, K.; Kawashima, J.; Nishihara, Y.; Mori, A.; Hiyama, T. Org. Lett. 1999, 1, 299-301. 32. Homsi, F.; Nozaki, K.; Hiyama, T. Tetrahedron Lett. 2000, 41, 5869-5872. 33. Homsi, F.; Hosoi, K.; Nozaki, K.; Hiyama, T. J. Organomet. Chem. 2001, 624, 208- 216. 34. Tamao, K.; Kobayashi, K.; Ito, Y. Tetrahedron Lett. 1989, 30, 6051-6054. 35. Ito, H.; Hosomi, A. J. Synth. Org. Chem. Jpn. 2000, 58, 274-284. 36. Denmark, S. E.; Wu, Z. Org. Lett. 1999, 1, 1495-1498. 37. Denmark, S. E.; Choi, J. Y. J. Am. Chem. Soc. 1999, 121, 5821-5822. 38. Denmark, S. E.; Neuville, L. Org. Lett. 2000, 2, 3221-3224. 39. Denmark, S. E.; Wehrli, D.; Choi, J. Y. Org. Lett. 2000, 2, 2491-2494. 40. Denmark, S. E.; Wang, Z. G. Synthesis 2000, 999-1003. 41. Denmark, S. E.; Wehrli, D. Org. Lett. 2000, 2, 565-568. 193 42. Denmark, S. E.; Wang, Z. G. J. Organomet. Chem. 2001, 624, 372-375. 43. Denmark, S. E.; Wang, Z. G. Org. Lett. 2001, 3, 1073-1076. 44. Denmark, S. E.; Pan, W. T. Org. Lett. 2001, 3, 61-64. 45. Horn, K. A. Chem. Rev. 1995, 95, 1317-1350. 46. Chuit, C.; Corriu, R. P. J.; Reye, C.; Young, J. C. Chem. Rev. 1993, 93, 1371- 1448. 47. Lee, H. M.; Nolan, S. P. Org. Lett. 2000, 2, 2053-2055. 48. Rossi, R.; Carpita, A.; Messeri, T. Gazz. Chim. Ital. 1992, 122, 65-72. 49. Tamao, K. J. Synth. Org. Chem. Jpn. 1990, 48, 457-462. 50. Babudri, F.; Farinola, G. M.; Naso, F.; Panessa, D. J. Org. Chem. 2000, 65, 1554- 1557. 51. Groenendaal, L.; Frechet, J. M. J. J. Org. Chem. 1998, 63, 5675-5679. 52. Hirabayashi, K.; Mori, A. J. Synth. Org. Chem. Jpn. 2000, 58, 926-933. 53. Mateo, C.; FernandezRivas, C.; Echavarren, A. M.; Cardenas, D. J. Organometallics 1997, 16, 1997-1999. 54. Mateo, C.; Fernandez-Rivas, C.; Cardenas, D. J.; Echavarren, A. M. Organometallics 1998, 17, 3661-3669. 55. Franz, A. K.; Woerpel, K. A. Acc. Chem. Res. 2000, 33, 813-820. 56. Hirabayashi, K.; Kondo, T.; Toriyama, F.; Nishihara, Y.; Mori, A. Bull. Chem. Soc. Jpn. 2000, 73, 749-750. 57. Ito, H.; Sensui, H.-o.; Arimoto, K.; Miura, K.; Hosomi, A. Chem. Lett. 1997, 639- 640. 58. Lam, P. Y. S.; Deudon, S.; Averill, K. M.; Li, R. H.; He, M. Y.; DeShong, P.; Clark, C. G. J. Am. Chem. Soc. 2000, 122, 7600-7601. 59. Lam, P. Y. S.; Deudon, S.; Hauptman, E.; Clark, C. G. Tetrahedron Lett. 2001, 42, 2427-2429. 60. Chuit, C.; Corriu, R. J. P.; Reye, C. In Chemistry of Hypercoordinate Compounds; Akiba, K.-Y., Ed.; Wiley-VCH: New York, 1999, p 81-146. 61. Gordon, M. S.; Carroll, M. T.; Davis, L. P.; Burggraf, L. W. J. Phys. Chem. 1990, 94, 8125-8128. 62. Deiters, J. A.; Holmes, R. R. J. Am. Chem. Soc. 1990, 112, 7197-7202. 63. Hosomi, A. In Reviews on Heteroatom Chemistry; Oae, S., Ed.; MYU-Printed in Japan: Tokyo, 1992; Vol. 7, p 214-228. 194 64. Hosomi, A.; Kohra, S.; Tominaga, Y. J. Chem. Soc., Chem. Commun. 1987, 20, 1517-1518. 65. Hosomi, A.; Kohra, S.; Tominaga, Y. Chem. Pharm. Bull. 1988, 36, 4622-4625. 66. Hojo, M.; Murakami, C.; Aihara, H.; Komori, E.-i.; Kohra, S.; Tominaga, Y.; Hosomi, A. Bull. Soc. Chim. Fr. 1995, 132, 499-508. 67. Mowery, M. E. Doctoral Dissertation, University of Maryland, College Park, 2000. 68. Seganish, W. M.; DeShong, P. J. Org. Chem. 2004, 69, 1137-1143. 69. Mehrotra, R. C. In Chemistry and Technology of Silicon and Tin; Kumar Das, V. G., Weng, N. G., Gielen, M., Eds.; Oxford University Press: New York, 1992, p 93-109. 70. Freeman, G. G. Silicones: An Introduction to Their Chemistry and Applications; The Chapel River Press Ltd.: Andover, Hants, England, 1962. 71. Ritter, K. Synthesis 1993, 735-762. 72. Riggleman, S. L.; DeShong, P. J. Org. Chem. 2003, 68, 8106-8109. 73. Pestunovich, V.; Kirppichenko, S.; Voronkov, M. G. In The Chemistry of Organic Silicon Compounds; Rappoport, Z., Apeloig, Y., Eds.; Wiley: Chichester, UK, 1998; Vol. 2, Part 2, Chap. 24, p 1447-1538. 74. Voronkov, M. G. Pure Appl. Chem. 1966, 13, 35-59. 75. Voronkov, M. G.; Yarosh, O. G.; Shchukina, L. V.; Tsetlina, E. O.; Tandura, S. N.; Korotaeva, I. M. Zh. Obshch. Khim. 1979, 49, 614-617. 76. Voronkov, M. G. Top. Curr. Chem. 1979, 84, 77-135. 77. Frye, C. L.; Vincent, G. A.; Finzel, W. A. J. Am. Chem. Soc. 1971, 93, 6805-6811. 78. Verkade, J. G. Acc. Chem. Res. 1993, 26, 483-489. 79. Iwamiya, J. H.; Maciel, G. E. J. Am. Chem. Soc. 1993, 115, 6835-6842. 80. Goodman, L. S.; Hardman, J. G.; Limbird, L. E.; Gilman, A. G. Goodman and Gilman's The Pharmacological Basis of Therapeutics; McGraw-Hill: New York, 2001. 81. Mahboobi, S.; Pongratz, H.; Hufsky, H.; Hockemeyer, J.; Frieser, M.; Lyssenko, A.; Paper, D. H.; Burgermeister, J.; Bohmer, F. D.; Fiebig, H. H.; Burger, A. M.; Baasner, S.; Beckers, T. J. Med. Chem. 2001, 44, 4535-4553. 82. Fitzgerald, T. J. Biochem. Pharmacol. 1976, 25, 1383-1387. 83. Rao, K. V.; Cullen, W. P. Antibiotics Annual 1959-1960, 950-953. 84. Rao, K. V.; Biemann, R.; Woodward, R. B. J. Am. Chem. Soc. 1963, 85, 2532- 2533. 195 85. Balitz, D. M.; Bush, J. A.; Bradner, W. T.; Doyle, T. W.; O'Herron, F. A.; Nettleton, D. E. J. Antibiot. 1982, 35, 259-265. 86. Doyle, T. W.; Balitz, D. M.; Grulich, R. E.; Nettleton, D. E. Tetrahedron Lett. 1981, 22, 4595-4598. 87. Graening, T.; Friedrichsen, W.; Lex, J.; Schmalz, H. G. Angew. Chem., Int. Ed. Engl. 2002, 41, 1524-1526. 88. Lee, J. C.; Jin, S. J.; Cha, J. K. J. Org. Chem. 1998, 63, 2804-2805. 89. Schreiber, J.; Threlfall, T.; Eschenmoser, A.; Schudel, P.; Leimgruber, W.; Pesaro, M. Helv. Chim. Acta 1961, 44, 540-597. 90. Van Tamelen, E.; Spencer, T. A.; Orvis, R. L.; Allen, D. S. Tetrahedron 1961, 14, 8-34. 91. Scott, A. I.; McCapra, F.; Buchanan, R. L.; Day, A. C.; Young, D. W. Tetrahedron 1965, 21, 3605-3631. 92. Martel, J.; Toromanoff, E.; Huynh, C. J. Org. Chem. 1965, 30, 1752-1759. 93. Kato, M.; Kido, F.; Wu, M. D.; Yoshikos.A Bull. Chem. Soc. Jpn. 1974, 47, 1516- 1521. 94. Boger, D. L.; Brotherton, C. E. J. Am. Chem. Soc. 1986, 108, 6713-6719. 95. Nakamura, T.; Murase, Y.; Endo, Y.; Hayashi, R. Chem. Pharm. Bull. 1962, 10, 281-290. 96. Evans, D. A.; Tanis, S. P.; Hart, D. J. J. Am. Chem. Soc. 1981, 103, 5813-5821. 97. Banwell, M. G. Pure Appl. Chem. 1996, 68, 539-542. 98. Banwell, M. G.; Cameron, J. M.; Collis, M. P.; Crisp, G. T.; Gable, R. W.; Hamel, E.; Lambert, J. N.; Mackay, M. F.; Reum, M. E.; Scoble, J. A. Aust. J. Chem. 1991, 44, 705-728. 99. Weinreb, S. M.; Basha, F. Z.; Hibino, S.; Khatri, N. A.; Kim, D.; Pye, W. E.; Wu, T.-T. J. Am. Chem. Soc. 1982, 104, 536-544. 100. Kende, A. S.; Lorah, D. P.; Boatman, R. J. J. Am. Chem. Soc. 1981, 103, 1271- 1273. 101. Boger, D. L.; Panek, J. S. J. Am. Chem. Soc. 1985, 107, 5745-5754. 102. Behforouz, M.; Haddad, J.; Cai, W.; Arnold, M. B.; Mohammadi, F.; Sousa, A. C.; Horn, M. A. J. Org. Chem. 1996, 61, 6552-6555. 103. Molina, P.; Murcia, F.; Fresneda, P. M. Tetrahedron Lett. 1994, 35, 1453-1456. 104. Behforouz, M.; Gu, Z.; Cai, W.; Horn, M. A.; Ahmadian, M. J. Org. Chem. 1993, 58, 7089-7091. 105. Ciafolini, M. A.; Bishop, M. J. J. Chem. Soc., Chem. Commun. 1993, 18, 1463- 1464. 196 106. Rao, A. V. R.; Chavan, S. P.; Sivadasan, L. Tetrahedron 1986, 42, 5065-5071. 107. Boger, D. L.; Duff, S. R.; Panek, J. S.; Yasuda, M. J. Org. Chem. 1985, 50, 5790- 5795. 108. Hibino, S.; Okazaki, M.; Ichikawa, M.; Sato, K.; Ishizu, T. Heterocycles 1985, 23, 261-264. 109. Kende, A. S.; Ebetino, F. H. Tetrahedron Lett. 1984, 25, 923-926. 110. Banwell, M. G.; Cameron, J. M.; Collis, M. P.; Gravatt, G. L. Aust. J. Chem. 1997, 50, 395-407. 111. Handy, C. J. Doctoral Dissertation, University of Maryland, College Park, 2002. 112. McElroy, W. T.; DeShong, P. Org. Lett. 2003, 5, 4779-4782. 113. Godard, A.; Marsais, F.; Ple, N.; Trecort, F.; Turck, A.; Queguener, G. Heterocycles 1995, 40, 1055-1091. 114. Tamao, K.; Miyaura, N. Top. Curr. Chem. 2002, 219, 1-9. 115. Schlosser, M. In Organometallics in Synthesis A Manual; 2nd ed.; Schlosser, M., Ed.; John Wiley and Sons: New York, NY, 2002, p 1-352. 116. Shibata, K.; Miyazawa, K.; Goto, Y. J. Chem. Soc., Chem. Commun. 1997, 14, 1309-1310. 117. Mix, A.; Berlekamp, U. H.; Stammler, H.-G.; Neumann, B.; Jutzi, P. J. Organomet. Chem. 1996, 521, 177-184. 118. Shea, K. J.; Loy, D. A.; Webster, O. J. Am. Chem. Soc. 1992, 114, 6700-6710. 119. Corriu, R. J. P.; Kpoton, A.; Poirier, M.; Royo, G.; Saxce, A. d.; Young, J. C. J. Organomet. Chem. 1990, 395, 1-26. 120. Selin, T. G.; West, R. J. Am. Chem. Soc. 1962, 84, 1856-1859. 121. Chappelow, C. C.; Elliot, R. L.; Goodwin, J. T. J. Org. Chem. 1960, 25, 435-439. 122. Gilman, H.; Brannen, C. J. Am. Chem. Soc. 1951, 73, 4640-4641. 123. Brown, H. C.; Cole, T. E. Organometallics 1983, 2, 1316-1319. 124. Wakefield, B. J. Organomagnesium Methods in Organic Synthesis; Academic Press, Inc.: San Diego, CA, 1995. 125. Raston, C. L.; Salem, G. In The Chemistry of the Metal-Carbon Bond; Hartley, F. R., Ed.; John Wiley and Sons: Chichester, 1987; Vol. 4, p 159-306. 126. Wardell, J. L. In The Chemistry of the Metal-Carbon Bond; Hartley, F. R., Ed.; John Wiley and Sons: Chichester, 1987; Vol. 4, p 1-158. 127. Wakefield, B. J. The Chemistry of Organolithium Compounds; 1st ed.; Pergamon Press: Oxford, 1974. 197 128. Gilman, H.; Morton, J. W., Jr. In Organic Reactions; Adams, R., Ed.; John Wiley & Sons, Inc.: New York, 1954; Vol. 8. 129. Jones, R. G.; Gilman, H. In Organic Reactions; Adams, R., Ed.; John Wiley & Sons, Inc.: New York, 1951; Vol. 6. 130. Blomberg, C. The Barbier Reaction and Related One-Step Processes; 1st ed.; Springer-Verlag: Heidelberg, 1993. 131. Snieckus, V. Chem. Rev. 1990, 90, 879-933. 132. Beak, P.; Kerrick, S. T.; Gallagher, D. J. J. Am. Chem. Soc. 1993, 115, 10628- 10636. 133. Beak, P.; Meyers, A. I. Acc. Chem. Res. 1986, 19, 356-363. 134. Gallagher, D. J.; Beak, P. J. Org. Chem. 1995, 60, 7092-7093. 135. Betz, J.; Bauer, W. J. Am. Chem. Soc. 2002, 124, 8699-8706. 136. Kremer, T.; Junge, M.; Schleyer, P. V. Organometallics 1996, 15, 3345-3359. 137. Han, J. W.; Son, S. U.; Chung, Y. K. J. Org. Chem. 1997, 62, 8264-8267. 138. Stille, J. K. Angew. Chem., Int. Ed. Engl. 1986, 25, 508-523. 139. Hassan, J.; Sevignon, M.; Gozzi, C.; Schulz, E.; Lemaire, M. Chem. Rev. 2002, 102, 1359-1469. 140. Sessler, J. L.; Sathiosatham, M.; Doerr, K.; Lynch, V.; Abboud, K. A. Angew. Chem., Int. Ed. Engl. 2000, 39, 1300-1303. 141. Davies, A. G.; Smith, P. J. In Comprehensive Organometallic Chemistry; Wilkinson, G., Stone, F. G. A., Abel, E. W., Eds.; Pergamon Press: Oxford, 1982; Vol. 2, p 519, Chapt. 11. 142. Pereyre, M.; Quintard, J.-P.; Rahm, A. Tin in Organic Synthesis; Butterworths: London, 1987. 143. Fugami, K.; Kosugi, M. Top. Curr. Chem. 2002, 219, 87-130. 144. Azizian, H.; Eaborn, C.; Pidcock, A. J. Organomet. Chem. 1981, 215, 49-58. 145. Murata, M.; Watanabe, S.; Masuda, Y. Synlett 2000, 1043-1045. 146. Rossi, R. A.; Pierini, A. B.; Penenory, A. B. Chem. Rev. 2003, 103, 71-167. 147. Gielen, M. Rev. Silicon Germanium Tin Lead Compd. 1981, 5, 5-152. 148. Gielen, M.; de Poorter, B. Rev. Silicon Germanium Tin Lead Compd. 1977, 3, 9- 144. 149. C?rsico, E. F.; Rossi, R. A. Synlett 2000, 230-232. 150. Jousseaume, B. J. Chem. Soc., Chem. Commun. 1984, 1452-1453. 198 151. Yammal, C. C.; Podesta, J. C.; Rossi, R. A. J. Org. Chem. 1992, 57, 5720-5725. 152. Jousseaume, B.; Villeneuve, P. Tetrahedron 1989, 45, 1145-1154. 153. Martin, A. R.; Yang, Y. Acta Chem. Scand. 1993, 47, 221-230. 154. Miyaura, N. Top. Curr. Chem. 2002, 219, 11-59. 155. Matteson, D. S. In The Chemistry of the Metal-Carbon Bond; Hartley, F. R., Ed.; John Wiley and Sons: Chichester, 1987; Vol. 4, p 307-402. 156. Letsinger, R. L.; Dandegaonker, S. H. J. Am. Chem. Soc. 1959, 81, 498-501. 157. Brindley, P. B.; Gerrard, W.; Lappert, M. F. J. Chem. Soc. 1955, 2956-2958. 158. Prim, D.; Campagne, J. M.; Joseph, D.; Andrioletti, B. Tetrahedron 2002, 58, 2041- 2075. 159. Marshall, J. A. Chem. Rev. 2000, 100, 3163-3185. 160. Ishiyama, T.; Miyaura, N. J. Organomet. Chem. 2000, 611, 392-402. 161. Ishiyama, T.; Murata, M.; Miyaura, N. J. Org. Chem. 1995, 60, 7508-7510. 162. Zhu, L.; Duquette, J.; Zhang, M. B. J. Org. Chem. 2003, 68, 3729-3732. 163. Ishiyama, T.; Ishida, K.; Miyaura, N. Tetrahedron 2001, 57, 9813-9816. 164. Furstner, A.; Seidel, G. Org. Lett. 2002, 4, 541-543. 165. Ishiyama, T.; Itoh, Y.; Kitano, T.; Miyaura, N. Tetrahedron Lett. 1997, 38, 3447- 3450. 166. Abraham, M. H.; Grellier, P. L. In The Chemistry of the Metal-Carbon Bond; Hartley, F. R., Patai, S., Eds.; Wiley: New York, 1985; Vol. 2. 167. Willis, D. M.; Strongin, R. M. Tetrahedron Lett. 2000, 41, 8683-8686. 168. Murata, M.; Watanabe, S.; Masuda, Y. J. Org. Chem. 1997, 62, 6458-6459. 169. Murata, M.; Oyama, T.; Watanabe, S.; Masuda, Y. J. Org. Chem. 2000, 65, 164- 168. 170. Baudoin, O.; Guenard, D.; Gueritte, F. J. Org. Chem. 2000, 65, 9268-9271. 171. Doux, M.; Mezailles, N.; Melaimi, M.; Ricard, L.; Le Floch, P. Chem. Commun. 2002, 1566-1567. 172. Euzenat, L.; Horhant, D.; Ribourdouille, Y.; Duriez, C.; Alcaraz, G.; Vaultier, M. Chem. Commun. 2003, 2280-2281. 173. Song, Y. L.; Morin, C. Synlett 2001, 2, 266-268. 174. Ishiyama, T.; Miyaura, N. J. Organomet. Chem. 2003, 680, 3-11. 199 175. Davies, M. W.; Johnson, C. N.; Harrity, J. P. A. J. Org. Chem. 2001, 66, 3525- 3532. 176. Takagi, J.; Sato, K.; Hartwig, J. F.; Ishiyama, T.; Miyaura, N. Tetrahedron Lett. 2002, 43, 5649-5651. 177. Hatanaka, Y.; Fukushima, S.; Hiyama, T. Chem. Lett. 1989, 1711-1714. 178. Hiyama, T.; Shirakawa, E. Top. Curr. Chem. 2002, 219, 62-85. 179. Hatanaka, Y.; Fukushima, S.; Hiyama, T. Heterocycles 1990, 30, 303-306. 180. Hatanaka, Y.; Hiyama, T. Tetrahedron Lett. 1990, 31, 2719-2722. 181. Hatanaka, Y.; Fukushima, S.; Hiyama, T. Tetrahedron 1992, 48, 2113-2126. 182. Damrauer, R.; Danahey, S. E. Organometallics 1986, 5, 1490-1494. 183. Yen, V. K.; Papouskova, Z.; Schraml, J.; Chvalovsky, V. Collect. Czech. Chem. Commun. 1973, 38, 3167-3175. 184. Farooq, O. J. Organomet. Chem. 2000, 613, 239-243. 185. Corriu, R. J. P.; Fernandez, J. M.; Guerin, C. J. Organomet. Chem. 1980, 192, 347-352. 186. Ohmura, H.; Matsuhashi, H.; Tanaka, M.; Kuroboshi, M.; Hiyama, T.; Hatanaka, Y.; Goda, K.-i. J. Organomet. Chem. 1995, 499, 167-71. 187. Yarosh, O. G.; Burnashova, T. D.; Voronkov, M. G. Bull. Acad. Sci. USSR Div. Chem. Sci. (Engl. Transl.) 1987, 36, 1030-1031. 188. Marans, N. S.; Sommer, L. H.; Whitmore, F. C. J. Am. Chem. Soc. 1951, 73, 5127- 5130. 189. Boyer, J.; Breliere, C.; Carre, F.; Corriu, R. J. P.; Kpoton, A.; Poirier, M.; Royo, G.; Young, J. C. J. Chem. Soc., Dalton Trans. 1989, 43-51. 190. Kunai, A.; Sakurai, T.; Toyoda, E.; Ishikawa, M. Organometallics 1996, 15, 2478- 2482. 191. Smit, C. N.; Bickelhaupt, F. Organometallics 1987, 6, 1156-1163. 192. Unno, M.; Masuda, H.; Matsumoto, H. Bull. Chem. Soc. Jpn. 1998, 71, 2449- 2458. 193. Jutzi, P.; Kanne, D.; Hursthouse, M.; Howes, A. J. Chem. Ber. 1988, 121, 1299- 1305. 194. Nakadaira, Y.; Oharu, K.; Sakurai, H. J. Organomet. Chem. 1986, 309, 247-256. 195. Hyde, J. F.; DeLong, R. C. J. Am. Chem. Soc. 1941, 63, 1194-1196. 196. Rich, J. D. J. Am. Chem. Soc. 1989, 111, 5886-5893. 200 197. Matsumoto, H.; Nagashima, S.; Yoshihiro, K.; Nagai, Y. J. Organomet. Chem. 1975, 85, C1-C3. 198. Cros, S.; Bennetau, B.; Dunogues, J.; Babin, P. J. Organomet. Chem. 1994, 468, 69-74. 199. Qiu, H.; Feng, S.; Chen, J.; Du, Z. Phosphorus, Sulfur Silicon Relat. Elem. 1990, 47, 325-326. 200. Chadha, R. K.; Drake, J. E.; Neo, M. K. H. J. Organomet. Chem. 1984, 277, 47- 52. 201. Bochkarev, V. N.; Turetskaya, R. A.; Troitskaya, N. N.; Luskina, B. M.; Savushkina, V. I.; Korablina, T. P.; Dzvonar, V. G.; Chernysheva, N. I. Zh. Obshch. Khim. 1983, 53, 824-827. 202. Bogovtseva, L. P.; Vainshtein, B. I.; Krasnova, T. L.; Rogachevskii, V. L.; Chernyshev, E. A. Zh. Obshch. Khim. 1978, 48, 1370-1373. 203. Rochow, E. G.; Gilliam, W. F. J. Am. Chem. Soc. 1945, 67, 1772-1773. 204. Brook, M. A. Silicon in Organic, Organometallic, and Polymer Chemistry; John Wiley & Sons, Inc.: New York, 2000. 205. Minge, O.; Mitzel, N. W.; Schmidbaur, H. Organometallics 2002, 21, 680-684. 206. Rosenberg, S. D.; Walburn, J. J.; Ramsden, H. E. J. Org. Chem. 1957, 22, 1606- 1607. 207. Lewis, D. W. J. Org. Chem. 1958, 23, 1893-1895. 208. Breed, L. W.; Haggerty, W. J. J. Org. Chem. 1960, 25, 126-128. 209. Breslow, R.; Heyer, D. J. Am. Chem. Soc. 1982, 104, 2045-2046. 210. Rubinsztajn, S.; Zeldin, M.; Fife, W. K. Synth. React. Inorg. Met.-Org. Chem. 1990, 20, 495-502. 211. Kuhn, N.; Laarz, W.; Schulten, M.; Boese, R.; Blaser, D.; Brett, W. A.; Stellberg, P. Z. Naturforsch., B: Chem. Sci. 1992, 47, 1233-1240. 212. Auner, N.; Probst, R.; Hahn, F.; Herdtweck, E. J. Organomet. Chem. 1993, 459, 25-41. 213. Belzner, J.; Schar, D.; Herbst-Irmer, R.; Kneisel, B. O.; Noltemeyer, M. Tetrahedron 1998, 54, 8481-8500. 214. Corriu, R. J. P.; Mix, A.; Lanneau, F. G. J. Organomet. Chem. 1998, 570, 183-193. 215. Cerveau, G.; Chappellet, S.; Corriu, R. J. P.; Dabiens, B. J. Organomet. Chem. 2001, 626, 92-99. 216. Bezombes, J. P.; Chuit, C.; Corriu, R. J. P.; Reye, C. J. Mater. Chem. 1998, 8, 1749-1759. 217. Post, H. W.; Norton, H. M. J. Org. Chem. 1942, 7, 528-533. 201 218. Corriu, R. J. P.; Granier, M.; Lanneau, G. F. J. Organomet. Chem. 1998, 562, 79- 88. 219. Tacke, R.; Bentlage-Felten, A.; Linoh, H.; Magda, S. Z. Naturforsch., B: Anorg. Chem. Org. Chem. 1986, 41B, 649-654. 220. Frohn, H. J.; Giesen, M.; Klose, A.; Lewin, A.; Bardin, V. V. J. Organomet. Chem. 1996, 506, 155-64. 221. Murata, M.; Suzuki, K.; Watanabe, S.; Masuda, Y. J. Org. Chem. 1997, 62, 8569- 8571. 222. Colvin, E. W. In The Chemistry of the Metal-Carbon Bond; Hartley, F. R., Ed.; John Wiley and Sons: Chichester, 1987; Vol. 4, p 539-621. 223. Petrov, A. D.; Mironov, B. F.; Ponomarenko, V. A.; Chernyshev, E. A. Synthesis of Organosilicon Monomers; Consultants Bureau: New York, 1964. 224. Jacob, S.; Popall, M.; Froehlich, L.; Kahlenberg, F.; Olma, K.; Fraunhofer- Gesellschaft zur Foerderung der Angewandten Forschung e.V., Germany: Denmark Patent 10159859, 2001, see Chem Abstr. 2003, 138, 899915. 225. Pak, J. J.; Greaves, J.; McCord, D. J.; Shea, K. J. Organometallics 2002, 21, 3552-3561. 226. Kano, N.; Komatsu, F.; Kawashima, T. J. Am. Chem. Soc. 2001, 123, 10778- 10779. 227. Riedmiller, F.; Jockisch, A.; Schmidbaur, H. Organometallics 1998, 17, 4444-4453. 228. Ahn, C.; Manoso, A. S.; Correia, R.; Soheili, A.; DeShong, P. unpublished results. 229. Correia, R. Doctoral Dissertation, University of Maryland, College Park, 2002. 230. French, H. E.; Sears, K. J. Am. Chem. Soc. 1951, 73, 469-470. 231. Fischer, F. C.; Havinga, E. Recl. Trav. Chim. Pays-Bas 1965, 84, 439-439. 232. Taylor, E. W.; Nikam, S. S.; Lambert, G.; Martin, A. R.; Nelson, D. L. Mol. Pharmacol. 1988, 34, 42-53. 233. Corriu, R. J. P.; Gu?rin, C.; Moreau, J. J. E. Top. Stereochem. 1984, 15, 43-198. 234. Wolfe, J. P.; Buchwald, S. L. Angew. Chem., Int. Ed. Engl. 1999, 38, 2413-2415. 235. Wolfe, J. P.; Singer, R. A.; Yang, B. H.; Buchwald, S. L. J. Am. Chem. Soc. 1999, 121, 9550-9561. 236. Yang, B. H.; Buchwald, S. L. J. Organomet. Chem. 1999, 576, 125-146. 237. Old, D. W.; Wolfe, J. P.; Buchwald, S. L. J. Am. Chem. Soc. 1998, 120, 9722- 9723. 238. Aranyos, A.; Old, D. W.; Kiyomori, A.; Wolfe, J. P.; Sadighi, J. P.; Buchwald, S. L. J. Am. Chem. Soc. 1999, 121, 4369-4378. 202 239. Littke, A. F.; Fu, G. C. Angew. Chem., Int. Ed. Engl. 1998, 37, 3387-3388. 240. Utimoto, K.; Otake, Y.; Yoshino, H.; Kuwahara, E.; Oshima, K.; Matsubara, S. Bull. Chem. Soc. Jpn. 2001, 74, 753-754. 241. Haebich, D.; Effenberger, F. Synthesis 1978, 755-756. 242. Eaborn, C.; Walton, D. R. M.; Young, D. J. J. Chem. Soc. B 1969, 15-20. 243. Benkeser, R. A.; Hickner, R. A.; Hoke, D. I. J. Am. Chem. Soc. 1958, 80, 2279- 2282. 244. Benkeser, R. A.; Krysiak, H. R. J. Am. Chem. Soc. 1954, 76, 6353-6357. 245. Brook, M. A.; Henry, C. Tetrahedron 1996, 52, 861-868. 246. Effenberger, F.; Spiegler, W. Chem. Ber. 1985, 118, 3872-3899. 247. Seconi, G.; Taddei, M.; Eaborn, C.; Stamper, J. G. J. Chem. Soc., Perkin Trans. 2 1982, 643-646. 248. Szele, I. Helv. Chim. Acta 1981, 64, 2733-2737. 249. Penalva, V.; Hassan, J.; Lavenot, L.; Gozzi, C.; Lemaire, M. Tetrahedron Lett. 1998, 39, 2559-2560. 250. Hassan, J.; Penalva, V.; Lavenot, L.; Gozzi, C.; Lemaire, M. Tetrahedron 1998, 54, 13793-13804. 251. Torii, S.; Tanaka, H.; Morisaki, K. Tetrahedron Lett. 1985, 26, 1655-1658. 252. Kotsuki, H.; Datta, P. K.; Hayakawa, H.; Suenaga, H. Synthesis 1995, 1348- 1350. 253. Echavarren, A. M.; Stille, J. K. J. Am. Chem. Soc. 1988, 110, 1557-1565. 254. Baillargeon, V. P.; Stille, J. K. J. Am. Chem. Soc. 1986, 108, 452-461. 255. Boukherroub, R.; Chatgilialoglu, C.; Manuel, G. Organometallics 1996, 15, 1508- 1510. 256. Benkeser, R. A.; Gaul, J. M.; Smith, W. E. J. Am. Chem. Soc. 1969, 91, 3666- 3667. 257. Hatanaka, Y.; Hiyama, T. Tetrahedron Lett. 1987, 28, 4715-4718. 258. Kunai, A.; Sakurai, T.; Toyoda, E.; Ishikawa, M.; Yamamoto, Y. Organometallics 1994, 13, 3233-3236. 259. Atwell, W.; Bokerman, G.; Dow Corning Corp.: U.S. Patent 3772347, 1971, see Chem Abstr. 1973, 44, 66560. 260. Atwell, W.; Bokerman, G.; Dow Corning Corp.: U.S. Patent 3746732, 1971, see Chem Abstr. 1973, 79, 66585. 203 261. Matsumoto, H.; Yoshihiro, K.; Nagashima, S.; Watanabe, H.; Nagai, Y. J. Organomet. Chem. 1977, 128, 409-413. 262. Matsumoto, H.; Yako, T.; Nagashima, S.; Motegi, T.; Nagai, Y. J. Organomet. Chem. 1978, 148, 97-106. 263. Matsumoto, H.; Shono, K.; Nagai, Y. J. Organomet. Chem. 1981, 208, 145-152. 264. Eaborn, C.; Griffiths, R. W.; Pidcock, A. J. Organomet. Chem. 1982, 225, 331-341. 265. Calas, R.; Dunogues, J.; Deleris, G.; Duffaut, N. J. Organomet. Chem. 1982, 225, 117-130. 266. Babin, P.; Bennetau, B.; Theurig, M.; Dunogues, J. J. Organomet. Chem. 1993, 446, 135-138. 267. Goossen, L. J.; Ferwanah, A. R. S. Synlett 2000, 1801-1803. 268. Viciu, M. S.; Grasa, G. A.; Nolan, S. P. Organometallics 2001, 20, 3607-3612. 269. Rich, J. D. Organometallics 1989, 8, 2609-2615. 270. Rich, J. D.; Krafft, T. E. Organometallics 1990, 9, 2040-2045. 271. Manoso, A. S.; DeShong, P. J. Org. Chem. 2001, 66, 7449-7455. 272. Murata, M.; Ishikura, M.; Nagata, M.; Watanabe, S.; Masuda, Y. Org. Lett. 2002, 4, 1843-1845. 273. Still, W. C.; Kahn, M.; Mitra, A. J. Org. Chem. 1978, 42, 2923-2924. 274. Perrin, D. D.; Armarego, W. L. F.; Perrin, D. R. Purification of Laboratory Chemicals; 2 ed.; Pergamon: New York, 1980. 275. Bjornestedt, R.; Zhong, G.; Lerner, R. A.; Barbas, C. F., III J. Am. Chem. Soc. 1996, 118, 11720-11724. 276. Davies, S. G.; Mobbs, B. E. J. Chem. Soc., Perkin Trans. 1 1987, 2597-2604. 277. Corcoran, R. C.; Bang, S. H. Tetrahedron Lett. 1990, 31, 6757-6758. 278. Coudret, C. Synth. Commun. 1996, 26, 3543-3547. 279. Young, T. E.; Amstutz, E. D. J. Am. Chem. Soc. 1951, 73, 4773-4775. 280. Balicki, R.; Kozlowska, M.; Sobotka, W. Bull. Pol. Acad. Sci.-Chem. 1986, 34, 281- 287. 281. Kunishima, M.; Friedman, J. E.; Rokita, S. E. J. Am. Chem. Soc. 1999, 121, 4722- 4723. 282. Olah, G. A.; Narang, S. C.; Balaram Gupta, B. G.; Malhotra, R. J. Org. Chem. 1979, 44, 1247-1251. 283. Schleyer, P. v. R. In Reagents for Organic Synthesis; Fieser, L. F., Fieser, M., Eds.; John Wiley and Sons: New York, 1967; Vol. 1, p 1179-1184. 204 284. Wawzonek, S.; Klimstra, P. D.; Kallio, R. E. J. Org. Chem. 1960, 25, 621-623. 285. Pilcher, A. S.; Ammon, H. L.; DeShong, P. J. Am. Chem. Soc. 1995, 117, 5166- 5167. 286. Kofron, W. G.; Baclawski, L. M. J. Org. Chem. 1976, 41, 1879-1880. CHAPTER TWO 1. Tandura, S. N.; Voronkov, M. G.; Alekseev, N. V. Top. Curr. Chem. 1986, 99-189. 2. Dilthey, W. Chem. Ber. 1903, 36, 923, 1595, 3207. 3. Dilthey, W. Liebigs Ann. Chem. 1906, 344, 300-300. 4. Rosenheim, A. V. A. Chem. Ber. 1903, 36, 1833-1833. 5. Armitage, D. A. In Comprehensive Organometallic Chemistry II: A Review of the Literature 1982-1994; Wilkinson, G., Stone, F. G. A., Abel, E. W., Davies, A. G., Eds.; Pergamon: Oxford, 1982; Vol. 2, Chap. 1. 6. Armitage, D. A. In Comprehensive Organometallic Chemistry; Wilkinson, G., Stone, F. G. A., Abel, E. W., Eds.; Pergamon: Oxford, 1995; Vol. 2, Chap. 9. 7. Voronkov, M. G. In Biochemistry of Silicon and Related Problems; Bendz, G., Lindqvist, I., Eds.; Plenum: New York, 1978, p 395-591. 8. Voronkov, M. G. Top. Curr. Chem. 1979, 84, 77-135. 9. Pestunovich, V.; Kirppichenko, S.; Voronkov, M. G. In The Chemistry of Organic Silicon Compounds; Rappoport, Z., Apeloig, Y., Eds.; Wiley: Chichester, UK, 1998; Vol. 2, Part 2, Chap. 24, p 1447-1538. 10. Voronkov, M. G. Pure Appl. Chem. 1966, 13, 35-59. 11. Voronkov, M. G.; Yarosh, O. G.; Shchukina, L. V.; Tsetlina, E. O.; Tandura, S. N.; Korotaeva, I. M. Zh. Obshch. Khim. 1979, 49, 614-617. 12. Frye, C. L.; Vincent, G. A.; Finzel, W. A. J. Am. Chem. Soc. 1971, 93, 6805-6811. 13. Verkade, J. G. Acc. Chem. Res. 1993, 26, 483-489. 14. Shklover, V. E.; Struchkov, Y. T.; Voronkov, M. G. Russ. Chem. Rev. 1989, 58, 211-240. 15. Voronkov, M. G.; Brodskaya, E. I.; Belyaeva, V. V.; Lazareva, N. F. Russ. Chem. Bull. 2001, 50, 757-763. 16. Voronkov, M. G. Bull. Acad. Sci. USSR Div. Chem. Sci. (Engl. Transl.) 1991, 40, 2319-2334. 205 17. Voronkov, M. G.; Pukhnarevich, V. B. Bull. Acad. Sci. USSR Div. Chem. Sci. (Engl. Transl.) 1982, 31, 939-957. 18. Voronkov, M. G.; Baryshok, V. P. J. Organomet. Chem. 1982, 239, 199-249. 19. Voronkov, M. G.; Dyakov, V. M.; Kirpichenko, S. V. J. Organomet. Chem. 1982, 233, 1-147. 20. Chu, H. K.; Johnson, M. D.; Frye, C. L. J. Organomet. Chem. 1984, 271, 327-336. 21. Sommer, L. H.; Frye, C. L.; Korte, W. D. J. Am. Chem. Soc. 1972, 94, 3463-3469. 22. Sommer, L. H.; Parker, G. A.; Lloyd, N. C.; Frye, C. L.; Michael, K. W. J. Am. Chem. Soc. 1967, 89, 857-861. 23. Frye, C. L.; Vincent, G. A.; Hauschil, G. I. J. Am. Chem. Soc. 1966, 88, 2727- 2730. 24. Sommer, L. H.; Frye, C. L.; Parker, G. A. J. Am. Chem. Soc. 1964, 86, 3280-3283. 25. Sommer, L. H.; Frye, C. L.; Parker, G. A. J. Am. Chem. Soc. 1964, 86, 3276-3279. 26. Sommer, L. H.; Michael, K. W.; Frye, C. L.; Parker, G. A. J. Am. Chem. Soc. 1964, 86, 3271-3276. 27. Sommer, L. H.; Frye, C. L.; Citron, J. D. J. Am. Chem. Soc. 1964, 86, 5684-5685. 28. Frye, C. L. J. Am. Chem. Soc. 1964, 86, 3170-3170. 29. Sommer, L. H.; Okaya, Y.; Parker, G. A.; Michael, K. W.; Musolf, M. C.; Rodewald, P. G.; Frye, C. L.; Pepinsky, R. J. Am. Chem. Soc. 1961, 83, 2210-2212. 30. Frye, C. L.; Hall, J. A.; Vogel, G. E. J. Am. Chem. Soc. 1961, 83, 996-997. 31. Sommer, L. H.; Frye, C. L. J. Am. Chem. Soc. 1960, 82, 3796-3796. 32. Sommer, L. H.; Frye, C. L. J. Am. Chem. Soc. 1960, 82, 4118-4119. 33. Sommer, L. H.; Frye, C. L. J. Am. Chem. Soc. 1959, 81, 1013-1013. 34. Lukevics, E.; Pudove, O.; Sturkovich, R. Molecular Structure of Organosilicon Compounds; Ellis Horwood Limited: Chichester, 1989. 35. Abele, E.; Lukevics, E. Main Group Met. Chem. 2001, 24, 315-350. 36. Lukevics, E.; Dzintara, M. J. Organomet. Chem. 1985, 295, 265-315. 37. Lukevics, E.; Pudova, O. Main Group Met. Chem. 1998, 21, 123-183. 38. Lukevics, E.; Pudova, O. Main Group Met. Chem. 1998, 21, 649-727. 39. Lukevics, E.; Pudova, O. Main Group Met. Chem. 1999, 22, 385-403. 40. Lukevics, E.; Pudova, O. Main Group Met. Chem. 2000, 23, 207-231. 41. Lukevics, E.; Pudova, O. A. Khim. Geterotsikl. Soedin. 1996, 1605-1646. 206 42. Chuit, C.; Corriu, R. P. J.; Reye, C.; Young, J. C. Chem. Rev. 1993, 93, 1371- 1448. 43. Corriu, R. J. P.; Young, J. C. In The Silicon-Heteroatom Bond; Patai, S., Rappoport, Z., Eds.; John Wiley & Sons: New York, 1991, Chap. 1. 44. Corriu, R. J. P.; Gu?rin, C.; Moreau, J. J. E. Top. Stereochem. 1984, 15, 43-198. 45. Corriu, R. J. P.; Gu?rin, C. Adv. Organomet. Chem. 1982, 20, 265-312. 46. Corriu, R. J. P.; Gu?rin, C. J. Organomet. Chem. 1980, 198, 231-320. 47. Pilcher, A. S.; DeShong, P. J. Org. Chem. 1996, 61, 6901-6905. 48. Pilcher, A. S.; Ammon, H. L.; DeShong, P. J. Am. Chem. Soc. 1995, 117, 5166- 5167. 49. Handy, C. J.; Lam, Y. F.; DeShong, P. J. Org. Chem. 2000, 65, 3542-3543. 50. Mowery, M. E.; DeShong, P. J. Org. Chem. 1999, 64, 3266-3270. 51. Brescia, M. R.; DeShong, P. J. Org. Chem. 1998, 63, 3156-3157. 52. Soli, E. D.; Manoso, A. S.; Patterson, M. C.; DeShong, P.; Favor, D. A.; Hirschmann, R.; Smith, A. B. J. Org. Chem. 1999, 64, 3171-3177. 53. Soli, E. D.; Manoso, A. S.; Patterson, M. C.; DeShong, P.; Favor, D. A.; Hirschmann, R.; Smith, A. B. J. Org. Chem. 1999, 64, 6526-6526. 54. Soli, E. D.; DeShong, P. J. Org. Chem. 1999, 64, 9724-9726. 55. Pilcher, A. S. Doctoral Dissertation, University of Maryland, College Park, 1995. 56. Brook, M. A. Silicon in Organic, Organometallic, and Polymer Chemistry; John Wiley & Sons, Inc.: New York, 2000. 57. Sanderson, R. T. J. Am. Chem. Soc. 1983, 105, 2259-2261. 58. Day, M. C. J.; Selbin, J. In Theoretical Inorganic Chemistry; Reinhold: New York, 1969. 59. Moore, C. E. In Ionization Potentials and Ionization Limits Derived from the Analysis of Optical Spectra; NSRDS-NBS 34: National Bureau of Standards: Washington, D. C., 1969. 60. Corey, E. J. In The Chemistry of Organic Silicon Compounds; Patai, S., Rappoport, Z., Eds.; Wiley: Chichester, UK, 1989; Vol. 1, Part 1, Chap. 1, p 3-56. 61. Cotton, F. A.; Wilkinson, G. Advanced Inorganic Chemistry; 5th ed.; Wiley: New York, 1988. 62. Allred, A. L.; E., R. J. Inorg. Nucl. Chem. 1958, 5, 264-268. 63. Luke, B. T.; Pople, J. A.; Krogh-Jespersen, M.-B.; Apeloig, Y.; Chandraskhar; Schleyer, P. v. R. J. Am. Chem. Soc. 1986, 108, 260-269. 207 64. Apeloig, Y. In The Chemistry of Organic Silicon Compounds; Patai, S., Rappoport, Z., Eds.; Wiley: Chichester, UK, 1989; Vol. 1, Part 1, Chap. 2, p 57- 226. 65. Dewar, M. J. S.; Healy, E. F.; Stewart, J. J. P.; Friedheim, J.; Grady, G. Organometallics 1986, 5, 375-379. 66. Colvin, E. W. Silicon in Organic Synthesis; Butterworths: London, 1981. 67. Kaftory, M.; Kapon, M.; Botoshansky, M. In The Chemistry of Organic Silicon Compounds; Rappoport, Z., Apeloig, Y., Eds.; Wiley: Chichester, UK, 1998; Vol. 2, Part 1, Chap. 5, p 181-266. 68. Sheldrick, W. S. In The Chemistry of Organic Silicon Compounds; Patai, S., Rappoport, Z., Eds.; Wiley: Chichester, UK, 1989; Vol. 1, Part 1, Chap. 3, p 227- 304. 69. Maryanoff, C. A.; Maryanoff, B. E. In Asym. Synth.; Morrison, J. D., Ed.; Academic: New York, 1984; Vol. 4, p 355-374. 70. Djerourou, A.-H.; Blanco, L. Tetrahedron Lett. 1991, 32, 6325-6326. 71. Fukui, T.; Kawamoto, T.; Tanaka, A. Tetrahedron: Asymmetry 1994, 5, 73-82. 72. Huheey, J. E. Inorganic Chemistry; 3rd ed.; Harper and Row: New York, 1983. 73. Bent, H. A. J. Chem. Phys. 1960, 33, 1259-1260. 74. Bent, H. A. Chem. Rev. 1961, 61, 275-311. 75. Bent, H. A. J. Chem. Phys. 1960, 33, 1260-1261. 76. Boudin, A.; Cerveau, G.; Chuit, C.; Corriu, R. J. P.; Rey?, C. Organometallics 1988, 7, 1165-1171. 77. Pommier, J. C. Rev. Silicon Germanium Tin Lead Compd. 1979, 4, 91-149. 78. Zheng, J.-Y.; Konishi, K.; Aida, T. Chem. Lett. 1998, 453-454. 79. Kane, K. M.; Lorenz, C. R.; Heilman, D. M.; Lemke, F. R. Inorg. Chem. 1998, 37, 669-673. 80. Kane, K. M.; Lemke, F. R.; Peterson, J. R. Inorg. Chem. 1995, 34, 4085-4091. 81. Marks, T. J. Science 1985, 227, 881-889. 82. Deiters, J. A.; Holmes, R. R. J. Am. Chem. Soc. 1990, 112, 7197-7202. 83. Kleboth, K.; Rode, B. M. Monatsh. Chem. 1974, 105, 815-821. 84. Boudin, A.; Cerveau, G.; Chuit, C.; Corriu, R. J. P. Angew. Chem., Int. Ed. Engl. 1986, 25, 473-474. 85. Birkofer, L.; Stuhl, O. In The Chemistry of Organic Silicon Compounds; Patai, S., Rappoport, Z., Eds.; Wiley: Chichester, UK, 1989; Vol. 1, Part 1, Chap. 10, p 655-762. 208 86. Lehn, J. M.; Sauvage, J. P. J. Am. Chem. Soc. 1975, 97, 6700-6707. 87. Laine, R. M.; Blohowiak, K. Y.; Robinson, T. R.; Hoppe, M. L.; Nardi, P.; Kampf, J.; Uhm, J. Nature 1991, 353, 642-644. 88. Kansal, P.; Laine, R. M. J. Am. Ceram. Soc. 1994, 77, 875-882. 89. Iwamiya, J. H.; Maciel, G. E. J. Am. Chem. Soc. 1993, 115, 6835-6842. 90. Smith, A. L. The Analytical Chemistry of Silicones; Wiley: New York, 1991; Vol. 112. 91. Taylor, R. B.; Parbhoo, B.; Fillmore, D. M. In The Analytical Chemistry of Silicones; Wiley: New York, 1991; Vol. 112. 92. Takeuchi, Y.; Takayama, T. In The Chemistry of Organic Silicon Compounds; Rappoport, Z., Apeloig, Y., Eds.; Wiley: Chichester, U. K., 1998; Vol. 2, Part 1, Chap. 6, p 267-354. 93. Williams, E. A. In The Chemistry of Organic Silicon Compounds; Patai, S., Rappoport, Z., Eds.; Wiley: Chichester, U. K., 1989; Vol. 1, Part 1, Chap. 8, p 511-554. 94. Engelhardt, G.; Jancke, H. Polym. Bull. 1981, 5, 577-584. 95. Williams, E. A.; Cargioli, J. D. In Annual Reports on NMR Spectroscopy; Webb, G. A., Ed.; Academic: New York, 1979; Vol. 9, p 221-318. 96. Harris, R. K.; Kennedy, J. D.; McFarlane, W. In NMR and the Periodic Table; Harris, R. K., Mann, B. E., Eds.; Academic: London, 1978. 97. Blinka, T. A.; Helmer, B. J.; West, R. Adv. Organomet. Chem. 1984, 23, 193-218. 98. Macomber, R. S. A Complete Introduction to Modern NMR Spectroscopy; John Wiley and Sons: New York, 1998. 99. March, J. Advanced Organic Chemistry: Reactions, Mechanisms, and Structure; 4 ed.; Wiley: New York, 1992. 100. Bethell, D.; Gold, V. Carbonium Ions: An Introduction; Academic Press: London and New York, 1967. 101. Allen, A. D.; Charlton, J. C.; Eaborn, C.; Modena, G. J. Chem. Soc. 1957, 3668- 3670. 102. Allen, A. D.; Modena, G. J. Chem. Soc. 1957, 3671-3678. 103. Schott, G.; Bondybey, V. Chem. Ber. 1967, 100, 1773-1777. 104. Kaufmann, K. D.; Ruhlmann, K. Zh. Obshch. Khim. 1967, 7, 391-392. 105. Schowen, R. L.; Latham, K. S. J. Am. Chem. Soc. 1967, 89, 4677-4683. 106. Corriu, R. J. P.; Royo, G. J. Organomet. Chem. 1968, 14, 291-307. 107. Corriu, R. J. P.; Royo, G. Bull. Soc. Chim. Fr. 1972, 1490-1497. 209 108. Corriu, R. J. P.; Larcher, F.; Royo, G. J. Organomet. Chem. 1976, 104, 161-171. 109. Ottosson, C.-H.; Cremer, D. Organometallics 1996, 15, 5309-5320. 110. Maerker, C.; Kapp, J.; Schleyer, P. v. R. In Organosilicon Chemistry II: From Molecules to Materials; Auner, N., Weis, J., Eds.; VCH: Weinheim, 1996, p 329- 359. 111. Eaborn, C.; Mahmoud, F. M. S. J. Chem. Soc., Perkin Trans. 2 1981, 1309-1315. 112. Corriu, R. J. P.; Dabosi, G.; Martineau, M. J. Organomet. Chem. 1978, 150, 27-38. 113. Corriu, R. J. P.; Dabosi, G.; Martineau, M. J. Organomet. Chem. 1978, 154, 33-43. 114. Corriu, R. J. P.; Dabosi, G.; Martineau, M. J. Organomet. Chem. 1980, 186, 25-37. 115. Corriu, R. J. P.; Dabosi, G.; Martineau, M. J. Chem. Soc., Chem. Commun. 1979, 457-458. 116. Krusic, P. J.; Fagan, P. J.; San Fillipo, J. J. Am. Chem. Soc. 1977, 99, 250-252. 117. San Fillipo, J.; Silbermann, J.; Fagan, P. J. J. Am. Chem. Soc. 1978, 100, 4834- 4842. 118. Felkin, H.; Knowles, P. J.; Meunier, B. J. Organomet. Chem. 1978, 146, 151-167. 119. Corriu, R. J. P.; Henner, B. J. L. J. Chem. Soc., Chem. Commun. 1973, 116-117. 120. Chauviere, G.; Corriu, R. J. P.; Henner, B. J. L. J. Organomet. Chem. 1975, 86, C1-C2. 121. Corriu, R. J. P.; Henner, B. J. L. J. Organomet. Chem. 1975, 102, 407-416. 122. Yoder, C. H.; Ryan, C. M.; Martin, G. F.; Ho, P. S. J. Organomet. Chem. 1980, 190, 1-7. 123. Macharashvili, A. A.; Shklover, V. E.; Struchkov, Y. T.; Oleneva, G. I.; Kramarova, E. P.; Shipov, A. G.; Baukov, Y. I. J. Chem. Soc., Chem. Commun. 1988, 683- 685. 124. Boer, F. P.; van Remoortere, F. P. J. Am. Chem. Soc. 1970, 92, 801-807. 125. Albanov, A. I.; Gubanova, L. I.; Larin, M. F.; Pestunovich, V. A.; Voronkov, M. G. J. Organomet. Chem. 1983, 244, 5-16. 126. Onan, K. D.; McPhail, A. T.; Yoder, C. H.; Hillyard, R. W. J. Chem. Soc., Chem. Commun. 1978, 209-210. 127. Bassindale, A. R.; Borbaruah, M. J. Chem. Soc., Chem. Commun. 1991, 1499- 1501. 128. Bassindale, A. R.; Borbaruah, M. J. Chem. Soc., Chem. Commun. 1991, 1501- 1503. 210 129. Bassindale, A. R. In Progress in Organosilicon Chemistry (Proceedings of the 10th International Symposium on Organosilicon Chemistry, Poznan, Poland, 1993; Marceniec, B., Chojnowski, J., Eds.; Gordon and Breach: Basel, 1995, p 191-207. 130. Sommer, L. H. Stereochemistry, Mechanism, and Silicon; McGraw Hill: New York, 1965. 131. Corey, E. J.; Snider, B. B. J. Am. Chem. Soc. 1972, 94, 2549-2550. 132. Corey, E. J.; Venkateswarlu, A. J. Am. Chem. Soc. 1972, 94, 6190-6191. 133. Furin, G. G.; Vyazankina, O. A.; Gostevsky, B. A.; Vyazankin, N. S. Tetrahedron 1988, 44, 2675-2749. 134. Corriu, R. J. P. In Frontiers of Organosilicon Chemistry, Edinburgh, 1990; Bassindale, A. R., Gaspar, P. P., Eds.; Royal Society of Chemistry: Cambridge, 1991, p 185-296. 135. Corriu, R. J. P.; Henner, B. J. L. J. Organomet. Chem. 1974, 71, 393-397. 136. Corriu, R. J. P.; Dabosi, G.; Martineau, M. J. Chem. Soc., Chem. Commun. 1977, 649-650. 137. Carpino, L. A.; Sau, A. C. J. Chem. Soc., Chem. Commun. 1979, 514-515. 138. Noyori, R.; Nishida, I.; Sakata, J. J. Am. Chem. Soc. 1981, 103, 2106-2108. 139. Nicolaou, K. C.; Reddy, K. R.; Skokotas, G.; Sato, F.; Xiao, X.-Y.; Hwang, K.-Y. J. Am. Chem. Soc. 1993, 115, 3558-3575. 140. Prakash, C.; Saleh, S.; Blair, I. A. Tetrahedron Lett. 1994, 35, 7565-7568. 141. Corey, E. J.; Yi, K. Y. Tetrahedron Lett. 1992, 33, 2289-2290. 142. Pilcher, A. S.; Hill, D. K.; Shimshock, S. J.; Waltermire, R. E.; Deshong, P. J. Org. Chem. 1992, 57, 2492-2495. 143. Gnann, R. Z.; Wagner, R. I.; Christe, K. O.; Bau, R.; Olah, G. A. J. Am. Chem. Soc. 1997, 119, 112-115. 144. Sharma, R. K.; Fry, J. L. J. Org. Chem. 1983, 48, 2112-2114. 145. Wetter, H.; Oertle, K. Tetrahedron Lett. 1984, 26, 5515-5518. 146. Cunico, R. F.; Bedell, L. J. Org. Chem. 1980, 45, 4797-4798. 147. Cook, F. L.; Bowers, C. W.; Liotta, C. L. J. Org. Chem. 1974, 39, 3416-3418. 148. Greene, T. W.; Wuts, P. G. Protective Groups in Organic Synthesis; John Wiley & Sons: New York, 1991. 149. Nelson, T. D.; Crouch, R. D. Synthesis 1996, 1031-1069. 150. Kuwajima, I.; Nakamura, E. J. Am. Chem. Soc. 1975, 97, 3257-3258. 151. Kuwajima, I.; Nakamura, E.; Shimizu, M. J. Am. Chem. Soc. 1982, 104, 1025-1030. 211 152. Noyori, R.; Nishida, I.; Sakata, J. Tetrahedron Lett. 1980, 21, 2085-2088. 153. Noyori, R.; Nishida, I.; Sakata, J. J. Am. Chem. Soc. 1983, 105, 1598-1608. 154. Noyori, R.; Yokoyama, K.; Sakata, J.; Kuwajima, I.; Nakamura, E.; Shimizu, M. J. Am. Chem. Soc. 1977, 99, 1265-1267. 155. Nakamura, E.; Shimizu, M.; Kuwajima, I.; Sakata, J.; Yokoyama, K.; Noyori, R. J. Org. Chem. 1983, 48, 932-945. 156. Jefford, C. W.; Jaggi, D.; Boukouvalas, J. Tetrahedron Lett. 1987, 28, 4037-4040. 157. RajanBabu, T. V. J. Org. Chem. 1984, 49, 2083-2089. 158. RajanBabu, T. V.; Reddy, G. S.; Fukunaga, T. J. Am. Chem. Soc. 1985, 107, 5473-5483. 159. Corriu, R. J. P.; Young, J. C. In The Chemistry of Organic Silicon Compounds; Patai, S., Rappoport, Z., Eds.; Wiley: Chichester, UK, 1989; Vol. 1, Part 2, Chap. 20, p 1241-1288. 160. Corriu, R. J. P.; Perz, R.; Reye, C. Tetrahedron 1983, 39, 999-1009. 161. Sogah, D. Y.; Farnham, W. B. In Organosilicon and Bioorganosilicon Chemistry; Sakurai, H., Ed.; Ellis Horwood: Chichester, 1985. 162. Corriu, R. J. P.; Perz, R. Tetrahedron Lett. 1985, 26, 1311-1314. 163. Nakamura, E.; Yamago, S.; Machii, D.; Kuwajima, I. Tetrahedron Lett. 1988, 29, 2207-2210. 164. Shenvi, S.; Stille, J. K. Tetrahedron Lett. 1982, 23, 627-630. 165. Miura, K.; Nakagawa, S.; Suda, A.; Hosomi, A. Chem. Lett. 2000, 150-151. 166. Shioiri, T.; Bohsako, A.; Ando, A. Heterocycles 1996, 42, 93-97. 167. Boyer, J.; Corriu, R. J. P.; Perz, R.; Rey?, C. J. Chem. Soc., Chem. Commun. 1981, 122-123. 168. Fujita, M.; Hiyama, T. J. Am. Chem. Soc. 1984, 106, 4629-4630. 169. Deneux, M.; Akhrem, I. C.; Avetissian, D. V.; Myssoff, E. I.; Vol'pin, M. E. Bull. Soc. Chim. Fr. 1973, 2638-2642. 170. Boyer, J.; Corriu, R. J. P.; Perz, R.; Rey?, C. J. Organomet. Chem. 1978, 157, 153-162. 171. Boyer, J.; Corriu, R. J. P.; Perz, R.; Rey?, C. J. Organomet. Chem. 1979, 172, 143-152. 172. Corriu, R. J. P.; Guerin, C.; Henner, B.; Wang, Q. J. Organometallics 1991, 10, 2297-2303. 173. Ch?rest, M.; Felkin, H.; Prudent, N. Tetrahedron Lett. 1968, 2199-2204. 212 174. Anh, N. T.; Eisenstein, O. Nouv. J. Chim. 1977, 1, 61-70. 175. Schinzer, D. In Organic Syntheses. Highlights II; Waldman, H., Ed.; VCH: New York, 1995, p 123-127. 176. Fujita, M.; Hiyama, T. J. Org. Chem. 1988, 53, 5405-5415. 177. Hosomi, A.; Shirahata, A.; Sakurai, H. Tetrahedron Lett. 1978, 3043-3046. 178. Degl' Innocenti, A.; Pike, S.; Walton, D. R. M.; Seconi, G.; Ricci, A.; Fiorenza, M. J. Chem. Soc., Chem. Commun. 1980, 1201-1202. 179. Nakamura, E.; Kuwajima, I. Angew. Chem., Int. Ed. Engl. 1976, 15, 498-499. 180. Holmes, A. B.; Jenningswhite, C. L. D.; Schulthess, A. H.; Akinde, B.; Walton, D. R. M. J. Chem. Soc., Chem. Commun. 1979, 840-842. 181. Pornet, J.; Randrianoelina, B. Tetrahedron Lett. 1981, 22, 1327-1328. 182. Pornet, J. Tetrahedron Lett. 1981, 22, 455-456. 183. Ricci, A.; Degl' Innocenti, A.; Fiorenza, M.; Taddei, M.; Spartera, M. A.; Walton, D. R. M. Tetrahedron Lett. 1982, 23, 577-578. 184. Effenberger, F.; Spiegler, W. Angew. Chem., Int. Ed. Engl. 1981, 20, 265-266. 185. Chan, T. H.; Lau, P. W. K.; Li, M. P. Tetrahedron Lett. 1976, 2667-2670. 186. Majetich, G.; Casares, A.; Chapman, D.; Behnke, M. J. Org. Chem. 1986, 51, 1745-1753. 187. Schwesinger, R.; Link, R.; Thiele, G.; Rotter, H.; Honert, D.; Limbach, H. H.; M?nle, F. Angew. Chem., Int. Ed. Engl. 1991, 30, 1372-1375. 188. Kira, M.; Hino, T.; Sakurai, H. Tetrahedron Lett. 1989, 30, 1099-1102. 189. Loezos, P.; Pilcher, A. S.; DeShong, P. Unpublished results. 190. Brown, B. R. The Organic Chemistry of Aliphatic Nitrogen Compounds; Oxford University Press: New York, 1994; Vol. 28. 191. Fleming, F. F. Nat. Prod. Rep. 1999, 16, 597-606. 192. Soli, E. D. Doctoral Dissertation, University of Maryland, College Park, 1999. 193. Larock, R. C. Comprehensive Organic Transformations: A Guide to Functional Group Preparations; VCH Publishers, Inc.: New York, 1989. 194. Shimizu, S.; Kito, K.; Sasaki, Y.; Hirai, C. Chem. Commun. 1997, 1629-1630. 195. Mizuno, A.; Hamada, Y.; Shioiri, T. Synthesis 1980, 1007-1008. 196. Thoman, C. J.; Habeeb, T. D.; Huhn, M.; Korpusik, M.; Sliso, D. F. J. Org. Chem. 1989, 54, 4476-4478. 213 197. Little is known about the biological activity of crown ethers, but limited oral toxicological studies demonstrate CNS effects. Aldrich Technical Bulletin AL-115; Aldrich Chemical Company, Inc.: Milwaukee, WI. 198. Leong, B. K. J. Chem. Eng. News 1975, 53, 5-5. 199. Hendrixson, R. R.; Mack, M. P.; Palmer, R. A.; Ottolenghi, A.; Ghirardelli, R. G. Toxicol. Appl. Pharmacol. 1978, 44, 263-268. 200. Mohora, M.; Olinescu, R.; Mihailescu, I.; Dinu, V.; Luca, C. Rom. J. Biophys. 1993, 3, 53-60. 201. Saito, K.; Harada, K. Bull. Chem. Soc. Jpn. 1989, 62, 2562-2566. 202. Harrison, C. R.; Hodge, P. Synthesis 1980, 299-301. 203. Regen, S. L.; Quici, S.; Liaw, S.-J. J. Org. Chem. 1979, 44, 2029-2030. 204. Ando, A.; Kawate, T.; Ichihara, J.; Hanafusa, T. Chem. Lett. 1984, 725-728. 205. Bram, G.; Fillebeen-Khan, T.; Geraghty, N. Synth. Commun. 1980, 10, 279-289. 206. Bram, G.; Loupy, A.; Pedoussaut, M. Bull. Soc. Chim. Fr. 1985, 124-128. 207. Matsubara, K.; Mukaiyama, T. Chem. Lett. 1994, 247-250. 208. Meinjohanns, E.; Meldal, M.; Paulsen, H.; Bock, K. J. J. Chem. Soc., Perkin Trans. 1 1995, 405-415. 209. Shaw, J. E.; Hsia, D. Y.; Parries, G. S.; Sawyer, T. K. J. Org. Chem. 1978, 43, 1017-1018. 210. Ogawa, T.; Nakabayashi, S.; Shibata, S. Agric. Biol. Chem. 1983, 47, 281-285. 211. Takeda, T.; Sugiura, Y.; Ogihara, Y.; Shibata, S. Can. J. Chem. 1980, 58, 2600- 2603. 212. White, D. A.; Baizer, M. M. J. Chem. Soc., Perkin Trans. 1 1973, 2230-2236. 213. Hudlicky, T.; Edoma, M. A.; Butora, G. J. Chem. Soc., Perkin Trans. 1 1996, 17, 2187-2192. 214. Ito, M.; Koyakumaru, K.-i.; Ohta, T.; Takaya, H. Synthesis 1995, 4, 376-378. 215. Harusawa, S.; Yoneda, R.; Omori, Y.; Kurihara, T. Tetrahedron Lett. 1987, 28, 4189-4190. 216. Friedman, L.; Shechter, H. J. Org. Chem. 1960, 25, 877-879. 217. Bram, G.; Loupy, A.; Pedoussaut, M. Bull. Soc. Chim. Fr. 1986, 467-475. 218. Pouchert, C. J.; Behnke, J. The Aldrich Library of 13C and 1H FT NMR Spectra; 1st ed.; The Aldrich Chemical Company, Inc.: Milwaukee, 1993; Vol. 1. 219. Grutzner, J. B.; Jautelat, M.; Dence, J. B.; Smith, R. A.; Roberts, J. D. J. Am. Chem. Soc. 1970, 92, 7107-7120. 214 220. Okimoto, M.; Chiba, T. J. Org. Chem. 1990, 55, 1070-1076. 221. Black, T. H. Org. Prep. Proced. Int. 1989, 21, 179-217. 222. Carretero, J. C.; Ruano, J. L. G. Tetrahedron Lett. 1985, 26, 3381-3384. 223. Dixon, D. A.; Hertler, W. R.; Chase, D. B.; Farnham, W. B.; Davidson, F. Inorg. Chem. 1988, 27, 4012-4018. 224. Handy, C. J. Doctoral Dissertation, University of Maryland, College Park, 2002. 225. Walz, A. J.; Miller, M. J. Org. Lett. 2002, 4, 2047-2050. 226. van Ameijde, J.; Albada, H. B.; Liskamp, R. M. J. J. Chem. Soc., Perkin Trans. 1 2002, 1042-1049. 227. Aggarwal, V. K.; Alonso, E.; Ferrara, M.; Spey, S. E. J. Org. Chem. 2002, 67, 2335-2344. 228. Adachi, H.; Nishimura, Y.; Takeuchi, T. J. Antibiot. 2002, 55, 92-98. 229. Schweizer, F. Angew. Chem., Int. Ed. Engl. 2001, 41, 230-253. 230. Sambaiah, T.; Fanwick, P. E.; Cushman, M. Synthesis 2001, 1450-1452. 231. Suzuki, T.; Tanaka, S.; Yamada, I.; Koashi, Y.; Yamada, K.; Chida, N. Org. Lett. 2000, 2, 1137-1140. 232. Hager, C.; Miethchen, R.; Reinke, H. J. Carbohydr. Chem. 2000, 19, 997-1018. 233. Fraser-Reid, B.; Anilkumar, G.; Nair, L. G.; Olsson, L.; Martin, M. G.; Daniels, J. K. Isr. J. Chem. 2000, 40, 255-262. 234. Wu, J.; Hou, X.-L.; Dai, L.-X. J. Org. Chem. 2000, 65, 1344-1348. 235. Bhat, A. S.; Gervay-Hague, J. Org. Lett. 2001, 3, 2081-2084. 236. Quadrelli, A.; Poli, R. Unpublished results. 237. Prakash, G. K. S.; Mandal, M. J. Am. Chem. Soc. 2002, 124, 6538-6539. 238. Prakash, G. K. S.; Mandal, M.; Olah, G. A. Angew. Chem., Int. Ed. Engl. 2001, 40, 589-590. 239. Nawrot, B.; Michalak, O.; Olejniczak, S.; Wieczorek, M. W.; Lis, T.; Stec, W. J. Tetrahedron 2001, 57, 3979-3985. 240. Bainbridge, J. M.; Corr, S.; Kanai, M.; Percy, J. M. Tetrahedron Lett. 2000, 41, 971-974. 241. Lipshutz, B. H.; Chrisman, W.; Noson, K.; Papa, P.; Sclafani, J. A.; Vivian, R. W.; Keith, J. M. Tetrahedron 2000, 56, 2779-2788. 242. Suzuki, T.; Suzuki, S. T.; Yamada, I.; Koashi, Y.; Yamada, K.; Chida, N. J. Org. Chem. 2002, 67, 2874-2880. 215 243. Damkaci, F.; DeShong, P. J. Org. Chem. 2003, 125, 4408-4409. 244. Still, W. C.; Kahn, M.; Mitra, A. J. Org. Chem. 1978, 42, 2923-2924. 245. Perrin, D. D.; Armarego, W. L. F.; Perrin, D. R. Purification of Laboratory Chemicals; 2 ed.; Pergamon: New York, 1980. 246. Olah, G. A.; Narang, S. C.; Balaram Gupta, B. G.; Malhotra, R. J. Org. Chem. 1979, 44, 1247-1251. 247. Schleyer, P. v. R. In Reagents for Organic Synthesis; Fieser, L. F., Fieser, M., Eds.; John Wiley and Sons: New York, 1967; Vol. 1, p 1179-1184. 248. Wawzonek, S.; Klimstra, P. D.; Kallio, R. E. J. Org. Chem. 1960, 25, 621-623. 249. Pouchert, C. J. The Aldrich Library of FT-IR Spectra; 1st ed.; Aldrich Chemical Company, Inc.: Milwaukee, 1985; Vol. 1. 250. Pouchert, C. J. The Aldrich Library of Infrared Spectra; 3rd ed.; Aldrich Chemical Company, Inc.: Milwaukee, 1981. 251. Otera, J.; Nakazawa, K.; Sekoguchi, K.; Orita, A. Tetrahedron 1997, 53, 13633- 13640. 252. Freerksen, R. W.; Selikson, S. J.; Wroble, R. R. J. Org. Chem. 1983, 48, 4087- 4096.