ABSTRACT Title of Dissertation: PALLADIUM-CATALYZED ALLYLIC- ARYLATION: MECHANISTIC STUDIES AND APPLICATION TO THE TOTAL SYNTHESIS OF (?)-7-DEOXYPANCRATISTATIN DERIVATIVES Krupa H. Shukla, Doctor of Philosophy, 2009 Directed By: Professor Philip DeShong Department of Chemistry and Biochemistry Palladium-catalyzed carbon-carbon bond formation is one of the most widely used reactions for the synthesis of biologically active substances. The DeShong group has demonstrated that hypervalent silicates can be employed for allyl-aryl carbon-carbon bond couplings in the presence of a Pd(0) catalyst. The goals of this dissertation are (1) to demonstrate application of palladium-catalyzed allylic-arylation coupling to the total synthesis of (?)-7-deoxypancratistatin and its analogues, and (2) to study the mechanism of allyl-aryl cross coupling reactions. In spite of the potent antitumor and antiviral activity of (+)-7-deoxypancratistatin, the use of this compound is limited in clinical applications because of its low natural abundance and lack of a practical scalable synthetic route. In order to test the feasibility of siloxane-based coupling in the synthesis of 7-deoxypancratistatin, a simplified analogue of (?)-7-deoxypancratistatin was synthesized. The key reaction in the synthesis involved stereoselective construction of a carbon-carbon bond between A and C rings via coupling of an aryl siloxane with an allylic carbonate. While siloxane methodology was successfully applied to the synthesis of a (?)-7-deoxypancratistatin analogue, application of this methodology to the natural product (?)-7-deoxypancratistatin proved to be a significant challenge. To understand the causes of the failure of the coupling reaction, a detailed mechanistic study was undertaken. Hammett analysis of the allyl-aryl coupling reaction demonstrated that the rate of the coupling reaction was enhanced by electron-withdrawing groups on the aryl siloxane. The positive slope of the Hammett plot indicated a charged transition state in which negative charge on the aryl ring was stabilized inductively. Furthermore, this study provided useful information regarding the nature of ligands on the palladium. Based on this study, a new family of Pd(0) olefin catalysts was developed. These catalysts were found to be highly efficient and formed carbon-carbon bond even at ambient temperature. Novel Pd(0) olefin complexes were successfully employed in the synthesis of (?)-7-deoxypancratistatin. The key coupling reaction of allylic carbonate with aryl siloxane produced Hudlicky's intermediate, thus constituting formal total synthesis of the actual product. Though the reaction required higher catalytic loading and proceeded in moderate yields, the ability of the reaction to work at ambient temperature is advantageous for practical synthesis of the natural product. Future studies shall aim at optimization of the key coupling reaction and application of this methodology to the synthesis of pancratistatin and related derivatives. PALLADIUM-CATALYZED ALLYLIC-ARYLATION: MECHANISTIC STUDIES AND APPLICATION TO THE TOTAL SYNTHESIS OF (?)-7-DEOXYPANCRATISTATIN DERIVATIVES By Krupa H. Shukla 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 2009 Advisory Committee: Professor Philip DeShong, Chair Professor Jeffery Davis Professor Daniel Falvey Professor Richard Payne Assistant Professor Barbara Gerratana ? Copyright by Krupa H. Shukla 2009 ii If we knew what it was we were doing, it would not be called research, would it? ? Albert Einstein iii DEDICATION ~ To My Loving Parents ~ iv ACKNOWLEDGMENTS I extend my gratitude to my advisor, Professor Philip DeShong for guiding me throughout this project. Without his encouragement and guidance it would have been impossible for me to accomplish what I have. This research experience has been one of the most rewarding and challenging experiences of my life and I am indebted to him. The entire DeShong group has been very supportive. I am grateful to all the former and current members. Special acknowledgements go to Debra Boehmler, Bridget Duvall, William McElroy, and Ju-Hee Park who have been very generous assisting and teaching me important lab skills and techniques. I thank Yiu-Fai Lam, Yinde Wang, Noel Whittaker and Yue Li for their assistance obtaining spectral data. I also thank Professor Christian Wolf and his student Hanhui Xu at Georgetown University for their help with microwave experiments. Many thanks to Yomi Okunola, An-Ni Chang and Julia Khusnutdinova, my friends at the University of Maryland. Special thanks to Julia for suggestions that improved this manuscript. I also thank Vasudha and Richa, my high school friends who inspired me to stay focused. I am grateful to my TA Carmen (2003-2004), who encouraged me to do undergraduate research and enthused me to pursue doctoral degree. Finally, I express my heartfelt gratitude to my lovely family who has always been supportive and motivating. I thank my grandparents for their blessings and my parents for their endless sacrifices and providing me with the best of everything. I am greatly thankful to my sweet brother who has always stood by me and has helped me to the best of his abilities. I warmly appreciate my husband for his abundant love, encouragement and understanding. Last, but not least, I thank God for turning my dream into reality. v TABLE OF CONTENTS LIST OF TABLES ........................................................................................................ VII LIST OF FIGURES .....................................................................................................VIII LIST OF SCHEMES ....................................................................................................... X LIST OF ABBREVIATIONS ..................................................................................... XIV CHAPTER 1 PALLADIUM-CATALYZED ALLYLIC-ARYLATION ............................................ 1 Introduction..................................................................................................................... 1 Allylic-Arylation............................................................................................................. 2 Hiyama-like Coupling................................................................................................. 3 Suzuki Coupling.......................................................................................................... 7 Stille Coupling .......................................................................................................... 11 Conclusion .................................................................................................................... 13 CHAPTER 2 TOTAL SYNTHESIS OF (?)-7-DEOXYPANCRATISTATIN ANALOGUE ......... 14 Introduction................................................................................................................... 14 Isolation and Biological Activity.............................................................................. 14 Synthetic Strategies................................................................................................... 17 (i) Michael Addition ............................................................................................. 19 (ii) Nucleophilic Addition..................................................................................... 20 (iii) Electrophilic Aromatic Substitution .............................................................. 21 (iv) SN2 and SN2? Coupling................................................................................... 22 (iv) Palladium-Catalyzed Coupling ...................................................................... 26 (v) Photocyclization.............................................................................................. 27 Research Goal ............................................................................................................... 28 Results and Discussion ................................................................................................. 30 Synthesis of Coupling Partners: Allylic Carbonate and Aryl Siloxane .................... 30 Coupling of Allylic Carbonate with Aryl Siloxane .................................................. 31 Generation of B ring and Installation of diol............................................................ 33 Conclusion .................................................................................................................... 37 Experimental Details..................................................................................................... 38 vi CHAPTER 3 FORMAL TOTAL SYNTHESIS OF (?)-7-DEOXYPANCRATISTATIN............... 52 Introduction................................................................................................................... 52 Results and Discussion ................................................................................................. 54 Synthesis of Coupling Partners: Allylic Carbonate and Aryl Siloxane .................... 54 Coupling of Allylic Carbonate with Aryl Siloxane .................................................. 56 Preliminary Attempts............................................................................................ 56 Investigation of Problems ..................................................................................... 57 Successful Coupling.............................................................................................. 63 Conclusion .................................................................................................................... 68 Experimental Details..................................................................................................... 69 CHAPTER 4 MECHANISTIC STUDIES ON PALLADIUM-CATALYZED ALLYLIC- ARYLATION .................................................................................................................. 86 Introduction................................................................................................................... 86 Transmetalation......................................................................................................... 88 Hammett Analysis..................................................................................................... 93 Hiyama Coupling.................................................................................................. 93 Suzuki Coupling.................................................................................................... 94 Stille Coupling ...................................................................................................... 96 Results and Discussion ................................................................................................. 98 Hammett Analysis..................................................................................................... 98 Role of Ligands....................................................................................................... 108 Conclusion .................................................................................................................. 119 Experimental Details................................................................................................... 120 REFERENCES.............................................................................................................. 129 vii LIST OF TABLES CHAPTER 2 Table 2.1: Reported total syntheses of the Amaryllidaceae isocarbostyrils. Adapted from ref 76. ................................................................................................. 18 CHAPTER 4 Table 4.1: Summary of Hammett studies ..................................................................... 98 Table 4.2: Role of ligands in the allyl-aryl coupling reaction .................................... 110 Table 4.3: Optimization of Pd(0)-olefin catalyzed allyl-aryl coupling reaction......... 116 viii LIST OF FIGURES CHAPTER 1 Figure 1.1: Catalysts for coupling of allylic substrates with aryl boronic acid derivatives................................................................................................... 10 CHAPTER 2 Figure 2.1: Structure of Amaryllidaceae alkaloids ....................................................... 14 Figure 2.2: Signaling pathways to apoptosis. Redrawn from ref 72............................. 15 Figure 2.3: Relative stereochemistry confirmed from correlation with 1H NMR coupling constants ...................................................................................... 36 CHAPTER 3 Figure 3.1: Pd(NBD)(MAH) and Pd(COD)(NQ) complexes....................................... 62 Figure 3.2: 1H-1H COSY of carbamate 49 (Hudlicky's intermediate).......................... 64 CHAPTER 4 Figure 4.1: Espinet's model for transmetalation in Stille reaction................................ 89 Figure 4.2: Transmetalation of alkyl boranes ............................................................... 90 Figure 4.3: Hammett analysis of the reaction of diaryl(difluoro)silanes with iodobenzene. Taken from ref 156............................................................... 94 Figure 4.4: Hammett analysis of the reaction of arylboronic acid with E-bromostilbene. Taken from ref 157. .................................................................................... 95 Figure 4.5: Palladacycles used in Hammett analysis of arylboronic acid with aryl bromides ..................................................................................................... 96 Figure 4.6: Hammett analysis of Stille coupling reaction in absence of LiCl. Taken from ref 165. ............................................................................................... 97 Figure 4.7: Hammett analysis of Stille coupling reaction in presence of LiCl. Taken from ref 165. ............................................................................................... 97 ix Figure 4.8: 19F NMR spectra of silicate formation (a) TBAF in THF at 29 ?C (b) 19F NMR spectrum of silicate complexes resulting from 1:1 mixture of TBAF and Triethoxyphenylsilane at 29 ?C. Insert is 19F signal at ? -121 after cooling to -28 ?C. ..................................................................................... 101 Figure 4.9: Summary of relative rates of coupling reactions with siloxane derivatives .................................................................................................................. 103 Figure 4.10: Hammett analysis of allyl-aryl coupling reaction .................................. 104 Figure 4.11: Donor-Acceptor model for transition-metal-olefin complexes. Redrawn from ref 190. ............................................................................................. 112 Figure 4.12: Various alkenyl ligands.......................................................................... 113 Figure 4.13: Pd(NBD)(MAH), Pd(COD)(MAH) and Pd(COD)(TCNE) complexes. 114 Figure 4.14: Pd(COD)(NQ), Pd(COD)(BQ), Pd(COD)(DQ), Pd2(NBE)2(BQ)2 complexes ................................................................................................. 117 Figure 4.15: 29Si NMR spectrum of silicate formation at -28 ?C. .............................. 125 Figure 4.16: Effect of temperature on silicate formation (19F NMR spectrum). Mixture of TBAF and phenyltriethoxysilane (a) 10 min, at rt. (b) 2 h 25 min, at -30 ?C. (c) 2 h 40 min, at rt. (d) 4 h 55 min, at -30 ?C. ............................ 127 Figure 4.17: Effect of TBAF concentration on silicate formation (19F NMR spectrum) (a) 0.5 equiv. TBAF, 1.0 equiv. siloxane. (b) 1.0 equiv. TBAF, 1.0 equiv. siloxane. (c) 1.5 equiv. TBAF, 1.0 equiv. siloxane. (d) 2.0 equiv. TBAF, 1.0 equiv. siloxane.................................................................................... 128 x LIST OF SCHEMES CHAPTER 1 Scheme 1.1...................................................................................................................... 1 Scheme 1.2...................................................................................................................... 2 Scheme 1.3...................................................................................................................... 3 Scheme 1.4...................................................................................................................... 4 Scheme 1.5...................................................................................................................... 4 Scheme 1.6...................................................................................................................... 5 Scheme 1.7...................................................................................................................... 6 Scheme 1.8...................................................................................................................... 7 Scheme 1.9...................................................................................................................... 7 Scheme 1.10.................................................................................................................... 8 Scheme 1.11.................................................................................................................... 8 Scheme 1.12.................................................................................................................... 9 Scheme 1.13.................................................................................................................... 9 Scheme 1.14.................................................................................................................... 9 Scheme 1.15.................................................................................................................. 11 Scheme 1.16.................................................................................................................. 12 Scheme 1.17.................................................................................................................. 12 Scheme 1.18.................................................................................................................. 13 CHAPTER 2 Scheme 2.1.................................................................................................................... 19 Scheme 2.2.................................................................................................................... 20 Scheme 2.3.................................................................................................................... 21 xi Scheme 2.4.................................................................................................................... 22 Scheme 2.5.................................................................................................................... 23 Scheme 2.6.................................................................................................................... 24 Scheme 2.7.................................................................................................................... 25 Scheme 2.8.................................................................................................................... 25 Scheme 2.9.................................................................................................................... 26 Scheme 2.10.................................................................................................................. 27 Scheme 2.11.................................................................................................................. 27 Scheme 2.12.................................................................................................................. 28 Scheme 2.13.................................................................................................................. 28 Scheme 2.14.................................................................................................................. 29 Scheme 2.15.................................................................................................................. 30 Scheme 2.16.................................................................................................................. 31 Scheme 2.17.................................................................................................................. 31 Scheme 2.18.................................................................................................................. 32 Scheme 2.19.................................................................................................................. 34 Scheme 2.20.................................................................................................................. 34 Scheme 2.21.................................................................................................................. 35 CHAPTER 3 Scheme 3.1.................................................................................................................... 52 Scheme 3.2.................................................................................................................... 52 Scheme 3.3.................................................................................................................... 53 Scheme 3.4.................................................................................................................... 53 Scheme 3.5.................................................................................................................... 54 Scheme 3.6.................................................................................................................... 55 xii Scheme 3.7.................................................................................................................... 55 Scheme 3.8.................................................................................................................... 56 Scheme 3.9.................................................................................................................... 57 Scheme 3.10.................................................................................................................. 58 Scheme 3.11.................................................................................................................. 59 Scheme 3.12.................................................................................................................. 60 Scheme 3.13.................................................................................................................. 60 Scheme 3.14.................................................................................................................. 61 Scheme 3.15.................................................................................................................. 63 Scheme 3.16.................................................................................................................. 64 Scheme 3.17.................................................................................................................. 66 Scheme 3.18.................................................................................................................. 66 Scheme 3.19.................................................................................................................. 67 Scheme 3.20.................................................................................................................. 67 CHAPTER 4 Scheme 4.1.................................................................................................................... 86 Scheme 4.2.................................................................................................................... 87 Scheme 4.3.................................................................................................................... 88 Scheme 4.4.................................................................................................................... 90 Scheme 4.5.................................................................................................................... 91 Scheme 4.6.................................................................................................................... 92 Scheme 4.7.................................................................................................................... 93 Scheme 4.8.................................................................................................................... 93 Scheme 4.9.................................................................................................................... 95 Scheme 4.10.................................................................................................................. 96 xiii Scheme 4.11.................................................................................................................. 99 Scheme 4.12................................................................................................................ 100 Scheme 4.13................................................................................................................ 105 Scheme 4.14................................................................................................................ 118 Scheme 4.15................................................................................................................ 126 xiv LIST OF ABBREVIATIONS Ac acetyl acac acetylacetonate aq. aqueous Ar aryl Bn benzyl BQ 1,4-benzoquinone Bu butyl Bz benzoyl calcd calculated COD 1,5-cyclooctadiene COSY correlation spectroscopy Cy cyclohexyl CBz carboxybenzyl dba dibenzylideneacetone DBU 1,8-diazabicyclo[5.4.0]undec-7-ene DQ duroquinone DCN 1,4-dicyanonaphthalene DIPHOS 1,2-bis(diphenylphosphino)ethane DMAP 4-dimethylaminopyridine DMF N,N-dimethylformamide DMSO dimethyl sulfoxide E ester EDG electron-donating group ee enantiomeric excess EI electron ionization equiv. equivalent(s) Et ethyl Et2O diethyl ether ESI electrospray ionization EWG electron-withdrawing group FAB fast atom bombardment FT fourier transform Fu furyl GC gas chromatography xv h hour(s) HMPA hexamethylphosphoramide HPLC high performance liquid chromatography HRMS high resolution mass spectrometry HSQC heteronuclear single quantum coherence Hz Hertz i-Pr isopropyl IR infrared J coupling constant L ligand LRMS low resolution mass spectrometry m meta M+ molecular ion m/z mass-to-charge ratio MAH maleic anhydride m-CPBA meta-chloroperoxybenzoic acid Me methyl MeCN acetonitrile MHz Megahertz min minute MOM methoxymethyl mp melting point MS mass spectrometry NBD norbornadiene NBE norbornene NMO N-methylmorpholine-N-oxide NMP N-methyl-2-pyrrolidone NMR nuclear magnetic resonance NQ 1,4-naphthoquinone Nu nucleophile o ortho OAc acetate OBz benzoate p para PEG poly(ethylene glycol) Ph phenyl PMB p-methoxybenzyl xvi Rf retention factor SAR structure activity relationship satd saturated rt room temperature t-Bu tertiary butyl TBAF tetrabutylammonium fluoride TBAB tetrabutylammonium bromide TBAT tetrabutylammonium triphenyldifluorosilicate TBS t-butyldimethylsilyl TCNE tetracyanoethylene Tf trifluoromethanesulfonyl TFA trifluoroacetic acid TFP tri-2-furylphosphine THF tetrahydrofuran TLC thin layer chromatography TIPS triisopropylsilyl TMS trimethylsilyl Ts tosyl UV ultra violet 1 CHAPTER 1 PALLADIUM-CATALYZED ALLYLIC-ARYLATION INTRODUCTION Palladium-catalyzed carbon-carbon bond formation is one of the most synthetically important and versatile reactions in the chemist's repertoire.1,2 The two most commonly used metal derivatives used in coupling include boron (Suzuki coupling)3 and tin (Stille coupling)4 (Scheme 1.1). The limitations associated with boron reagents include synthesis of the boronic acid derivatives, homocoupling and incompatibility with Lewis basic functions. While Stille coupling offers functional group tolerance, it is limited in pharmaceutical applications due to the toxicity of tin reagents and the removal of the trace tin byproducts. Scheme 1.1 I OMe Ph?SnBu3 Ph?B(OH)2 OMe Ph Stille Coupling Suzuki Coupling Pd(0) Ph?Si(OEt)3 Hiyama-like Coupling In contrast, silicon-based coupling (Hiyama coupling)5-9 offers a viable alternative to other coupling technologies because of the low cost, low toxicity and chemical stability of silicon-derived compounds.10,11 Hiyama demonstrated that the coupling of aryl fluorosilanes with aryl iodides and bromides proceeded in moderate to good yields. 2 Recently, the DeShong group developed siloxane methodology, which utilizes hypervalent siloxane derivatives in the presence of a palladium catalyst to form an aryl-aryl carbon bond.12-16 Subsequently, an efficient cross-coupling of aryl bromides and chlorides with aryl siloxanes using the palladium/imidazolium catalytic system was reported by Lee.17 Aryl siloxanes have also been used in aqueous systems by Wolf.18 More recently, Clarke reported the first microwave accelerated Hiyama-like coupling of aryl siloxanes with aryl halides.19 Not only are siloxanes less toxic than Stille reagents, they are also quite stable, easily prepared and purified.20,21 Recently, the use of aryl silanols and aryl silanolates as efficient coupling partners has emerged as reported by Denmark.22-25 ALLYLIC-ARYLATION In addition to aryl-aryl cross-coupling reaction, organometallic reagents of silicon, boron and tin can be used to construct the allyl-aryl bond (Scheme 1.2). In most cases, allyl halides or allyl carboxylates, prepared from the corresponding alcohols, are used as allylating agents. Scheme 1.2 Ph?SnBu3 Ph?B(OH)2 Ph Stille Coupling Suzuki Coupling Pd(0) Ph?Si(OEt)3 Hiyama-like Coupling OR 3 Organometallic reagents are unstabilized nucleophiles (pKa > 25) and follow the general mechanism outlined in the Scheme 1.3.26 Allyl benzoate 1 reacts with Pd(0) to give the ?-allyl Pd intermediate 2, with inversion of configuration. Next, the aryl group is transferred from the organometallic species 3 onto the Pd via transmetalation to form the Pd(II) intermediate 4. Subsequent reductive elimination from the same face as the palladium generates 5 with overall inversion at the reaction center. Note that the ?-allyl Pd intermediate 4 is a meso compound. Therefore, reductive elimination could also result in formation of ent-5. Scheme 1.3 BzO Pd OBz 2 Pd(0) 1 Ph Pd Ph 5 Ph-M 3 4 Hiyama-like Coupling Palladium-catalyzed coupling of allylic carbonate with aryl and vinyl fluorosilanes was reported by Hiyama and Hatanaka (Scheme 1.4).27 This reaction proceeded without fluoride ion activation; however the reaction suffered from poor 4 yields. Moreover, the synthesis of fluorosilanes involves multiple steps and fluorosilanes are hydrolytically unstable. Subsequently, Hiyama examined the coupling reaction of allylic and benzylic carbonate with an organo[2-(hydroxymethyl)phenyl]dimethylsilane in the absence of any activator (Scheme 1.5).28 Upon treatment with a mild base (K2CO3), the proximal hydroxyl group gets converted to an alkoxide and coordinates to the nearby silicon atom to produce a five-membered penta-coordinated silicate species. In comparison to fluorosilanes, this tetraorganosilicon reagent is highly stable. Scheme 1.4 OCO2Et Pd2(dba)3?CHCl3 PPh3, 60 ?C, 40% SiEtF2 Scheme 1.5 Pd2(dba)3, P(2-thienyl)3 OCO2t-Bu Me 2Si OH CuOAc, 70 ?C, 75% In an effort to extend the viability of silicon-based coupling, the DeShong group developed the palladium-catalyzed allylic-arylation methodology as depicted in the Scheme 1.6.29-34 Phenyltriethoxysilane 6 when treated with tetrabutylammonium fluoride (TBAF), generated in situ hypervalent species 7. This hypervalent reagent 7 reacted with allylic benzoate 8 to give cyclohexene 9. This methodology has been shown to efficiently transfer a wide variety of aryl groups to allylic esters in excellent yields (70-95%).32 5 In more complex systems such as allylic benzoate 10, there are regiochemical as well as a stereochemical issues (Scheme 1.6). The reaction of aryl siloxane 6 with allylic carbonate 10 forms two regioisomers 11 and 12 with an overall inversion of stereochemistry. The origin of regioselectivity and stereoselectivity can be understood from the mechanism shown in the Scheme 1.7. When allylic benzoate 10 reacts with palladium, the benzoate group is displaced by palladium from the opposite face (inversion) to form ?-allyl intermediate 13. Next, the aryl group is transferred from the hypervalent silicate species 7 onto Pd via transmetalation to give ?-allyl intermediate 14. The resulting ?-allyl intermediate 14 is ?unsymmetrical? since palladium is pushed away from the methyl group to minimize steric interactions. Reductive elimination therefore occurs predominantly on C1 from the same face (retention) as palladium to obtain 11 as the major product with a net inversion of configuration at the allylic carbon. Scheme 1.6 OBz OBz Me Me Ph Me Ph Ph Major Minor Ph?Si(OEt)3 TBAF Ph Si OEt F OEt OEt 6 7 8 Pd(0) 9 10 11 12 95% 6 Scheme 1.7 Ln OBz Me MePd Ln MePdPhMeH H H Pd LnPh Me Ph Me Ph 1 3 Major Minor Pd(0) 1 3 10 13 1 3 1 3 14 11 12 Ph Si OEt F OEt OEt 7 Recently, Pd(0) nanoparticle (generated in situ) catalyzed cross-coupling of allyl acetates and aryl and vinyl siloxane has been reported (Scheme 1.8).35 It is postulated that Pd(II) is reduced to Pd(0) by allyl acetate and TBAB (tetrabutylammonium bromide) stabilizes Pd(0) nanoparticles. The reaction is applicable to a variety of unactivated and activated allyl acetates (Baylis-Hillman acetate adducts) and organosiloxanes. This reaction is highly regioselective providing straight chain olefins through coupling from the less substituted carbon. Moreover, the nanoparticles can be recovered after reaction and remain appreciably active through three catalytic cycles. However, this reaction has not been applied to cyclic allylic substrates. Activated allyl acetates (Baylis-Hillman acetate adducts derived from methyl acrylate, acrylonitrile, acyclic and cyclic ?,?-unsaturated ketones) have also shown to couple with aryl siloxane in poly(ethylene glycol) (PEG) by Kabalka.36 7 Scheme 1.8 CO2Me OAc Si(OMe)3 PdCl2, TBAB TBAF, THF, 65 ?C, 75% OAc CO2Me Si(OMe)3 PdCl2, TBAB TBAF, THF, 65 ?C, 80% Suzuki Coupling The coupling of sodium tetraphenylborate with allylic acetate (Scheme 1.9) was demonstrated by Fiaud and Legros.37 However, the reaction was limited to the transfer of phenyl groups. Scheme 1.9 NaBPh4 Pd(dba)2, PPh3 60 ?C, 80% OAc Ph A breakthrough in this field of boron-based coupling was achieved by Moreno-Ma?as who reported coupling of aryl boronic acids with allyl bromides in refluxing benzene using Pd(dba)2 as catalyst (58-91% yield)38 and by Hayashi who coupled aryl boronic acids with allyl acetates using a resin-supported palladium catalyst under basic conditions in water (45-99% yield).39 A practical and stereoselective synthesis of C-arylglycosides by coupling of arylboronic acid and peracetylated glycols in the presence of catalytic Pd(OAc)2 was reported by Maddaford (Scheme 1.10).40 In this case, the regioselectivity is governed by 8 the electronic effect of the oxygen functionality and is limited to this specific substrate apparently since the galacto analogue did not undergo coupling. Scheme 1.10 PhB(OH)2 Pd(OAc)2, MeCN rt, 82% O OAc AcO AcO OAcO AcO Ph Balme and co-workers have developed a novel catalytic system [PdCl2(TFP)2] (TFP ? tri-2-furylphosphine) of allyl acetates with a variety of aryl boronic acids in the presence of fluoride (Scheme 1.11).41 The reaction displays excellent regioselectivity and stereoselectivity as evident by the formation of a single coupling product with E-geometry. The reaction also works well with cyclic allyl acetates. [PdCl2(TFP)2] has been shown to also couple pinacol aryl and vinyl boronates to allyl acetates in moderate to good yields by Ortar (Scheme 1.12).42 Scheme 1.11 OAc B(OH)2 [PdCl2(TFP)2], KF MeOH, rt, 86% B(OH)2 [PdCl2(TFP)2], KF MeOH, rt, 78% OAc 9 Scheme 1.12 OAc B [PdCl2(TFP)2], KF MeOH, rt, 82% OO Mino and co-workers reported palladium-catalyzed coupling of allylic acetates with boronic acids at room temperature in the presence of Pd(OAc)2 and phosphine-free hydrazone ligand to produce allyl benzene derivatives in good yields (Scheme 1.13).43 Scheme 1.13 OAc B(OH)2 + NN NN R R R RR = (CH 2)5 Pd(OAc)2, K2CO3 DMF-H2O, rt, 94% Scheme 1.14 Me B(OH)2OAc Bu MeBu N N Pd OAc SbF6 ?/? 100:0, E/Z>20 60 ?C, 65%, 97% ee 1, 2-dichloroethane Air-tolerable allyl-aryl coupling reaction in the presence of palladium catalyst was reported by Sawamura recently.44 The reaction of optically active allyl acetates with phenylboronic acid in the presence of Pd(OAc)2, 1,10-phenanthroline and AgSbF6 gave 10 coupling product in good yields. The reaction takes place with excellent ? to ? chirality transfer and syn-selectivity (Scheme 1.14). Some of the other catalysts developed for coupling of allylic substrates with arylboron derivatives are shown in Figure 1.1.45-47 N-Heterocyclic carbene palladium complex I is able to catalyze reaction of activated allyl chlorides with arylboronic acid at room temperature.45 Despite the presence of ?-hydrogens in the allylic substrate, the coupling of allylic substrates proceeded to give excellent yields of coupling products. N?jera introduced a thermally stable catalyst, di(2-pyridyl)methylamine-based Pd(II) complex II which can function in aqueous conditions.46 The coupling reaction of allylic substrates (allylic chlorides, allylic acetates, allylic carbonates) with arylboronic acids occurred in refluxing water or at room temperature in aqueous acetone to provide coupling products in good yields. Recently, N?jera and co-workers cross-coupled allyl chlorides with potassium aryltrifluoroborates using an oxime-derived palladacycle III in aqueous acetone at room temperature or 50 ?C.47 In comparison to toxic and pyrophoric phosphines, the coupling protocol employed by these catalysts (Figure 1.1) is user-friendly and environmentally benign. N N Pd (OAc)2 N NHCONHCy N Pd ClCl HO Pd N Me OH 2 I II III Figure 1.1: Catalysts for coupling of allylic substrates with aryl boronic acid derivatives Recently, cross-coupling of allylic alcohol with aryl and vinyl boronic acids in organic as well as aqueous solvents have emerged.48-51 Also, microwave assisted 11 allylic-arylations have been reported.52,53 Apart from palladium-catalyzed allylic-arylation using boron reagents, nickel-catalyzed54,55 and rhodium-catalyzed56 coupling reactions are also known in the literature. Stille Coupling Stille performed palladium-catalyzed reaction of allylic acetates with aryl and vinyl stannanes (Scheme 1.15).57 However, the reaction gave poor yields when used for cyclic allyl acetates. Later, Echavarren reported palladium-catalyzed cross-coupling reaction of allylic carbonates with organostannanes.58 Scheme 1.15 CO2Me OAc PhSnMe3 CO2Me Ph Pd(dba)2, PPh3 55 ?C, 47% In an effort to improve the yield of these couplings, the effect of polarity of the solvent on stereoselective coupling reaction of allyl chlorides and aryl stannane was reported by Kurosawa (Scheme 1.16).59 In the presence of weakly coordinating solvents such as benzene, acetone, CH2Cl2 or THF retention of configuration 15 was observed. However, nearly complete inversion 16 was observed in coordinating solvents such as DMSO or MeCN indicating a change of mechanism in the presence of strongly donor ligands on the metal. 12 Fairlamb has reported a new bromosuccinimido-Pd catalyst 17 synthesized from Pd2(dba)3?CHCl3 for the coupling of allylic and benzylic bromides with vinyl stannanes (Scheme 1.17).60 Scheme 1.16 CO2Me Cl CO2Me Ph Pd(0) CO2Me Ph SnBu3 96 4 0 100 Benzene Acetonitrile 15 16 Scheme 1.17 Pd PPh3 PPh3 Br N O OBu3Sn CO2Et Br 17 62% CO2Et Due to high tolerance towards most functional groups, allyl-aryl Stille coupling reaction has been employed in the construction of a variety of ring systems in highly functionalized molecules.61,62 Although the Stille coupling is tolerated by many functional groups, the reaction suffers from several limitations. The tin (IV) derivatives are toxic, the removal of tin byproducts is difficult and the coupling exhibits moderate stereoselectivities.29 For example, the coupling of allylic benzoate 18 with hypervalent silicate (TBAT) resulted in the stereoselective formation of regioisomers 19 and 20. Moreover, high regioselectivity was observed, evident by the formation of the major 13 regioisomer 19. However, coupling of allylic benzoate 18 using Stille reaction conditions gave mixture of several products (Scheme 1.18). Scheme 1.18 O Me Ph O Me Ph 19 20 O Me Ph O Me Ph 19 20 O Me Ph O Me Ph 21 22 O OBz Me TBAT PhSnMe3 Pd(0), 65% Pd(0), 65% 19:20 = 15:1 18 LiCl 19:20:21:22 = 15:4:1:2 CONCLUSION Palladium-catalyzed cross-couplings utilizing hypervalent silicate anions have several advantages compared to Stille and Suzuki coupling reactions. These include mild reaction conditions, stability, low toxicity and ease of preparation of silicon reagents. Moreover, allyl-aryl cross-coupling reaction involving siloxane methodology results in stereospecific arylation with net inversion of configuration. The succeeding chapters will demonstrate how palladium-catalyzed allylic-arylation methodology involving hypervalent silicates has been applied to the synthesis of natural product 7-deoxypancratistatin and its derivatives. 14 CHAPTER 2 TOTAL SYNTHESIS OF (?)-7-DEOXYPANCRATISTATIN ANALOGUE INTRODUCTION O O NH O HO OH OH OH R 23, pancratistatin, R = OH 24, 7-deoxypancratistatin, R = H A B C 7 O O NH O OH OH OH R 25, narciclasine, R = OH 26, 7-deoxynarciclasine, R = H A B C 7 O O N OH A B HO C 27, lycorine Figure 2.1: Structure of Amaryllidaceae alkaloids Isolation and Biological Activity Amaryllidaceae alkaloids (Figure 2.1) have been recognized for a long time because of their medicinal value. Among various alkaloids isolated from Amaryllidaceae species, (+)-pancratistatin (23) shows the most promising biological activity. Pancratistatin was first isolated and extracted from the Hawaiian daffodil bulb Pancratium littorale (also known as Hymenocallis littorale) in a low yield of 0.014% dry weight.63 Since, many related analogues have been isolated including the less toxic analogue, (+)-7-deoxypancratistatin (24).64 This deoxygenated analogue (24) was isolated from the roots of the bulbs Haemanthus kalbreyeri and is eight to thirty-two times less toxic65 and about ten fold less potent than pancratistatin (23).66 The U.S. National Cancer Institute identified the potent anticancer activity of (+)-pancratistatin (23) in early 1980s.67,68 Pancratistatin (23) was shown to inhibit growth 15 of numerous cell lines including leukemia and ovarian sarcoma. Recently, Pandey and co-workers have shown that pancratistatin selectively induces apoptosis in various types of cancer cell lines (breast, colon, prostrate, neuroblastoma, melanoma and leukemia) at micro molar concentrations.69-71 Apoptosis (programmed cell death) can be activated through intrinsic or extrinsic pathway (Figure 2.2).71,72 In intrinsic pathway, disruption of mitochondrial membrane followed by release of cytochrome c activates caspase-3. The extrinsic pathway is initiated by receptor/ligand binding that ultimately leads to an activation of caspase-3. Caspase-3 activates deoxyribonuclease to cause DNA fragmentation and consequently apoptosis. Conventional anticancer therapies (chemotherapy and radiotherapy) trigger intrinsic pathway by inducing DNA damage. However, these treatments carry risk of DNA damage and mutations in non-cancerous cells. Cytochrome c APAF1+ Caspase 3 Caspase 9Caspase 8 Cell Death IAP BCL2 Apoptotic stimulus (chemotherapy, UV) Death Ligand De ath Re ce pto r BID Mitochondrion Intrinsic Extrinsic Figure 2.2: Signaling pathways to apoptosis. Redrawn from ref 72. 16 To understand the selectivity of (+)-pancratistatin (23), Pandey and co-workers studied the mechanism of the action of pancratistatin in human leukemia cell line.71 These studies suggested a possible interaction between pancratistatin, caspase-3 and Fas receptor within the plasma membrane to induce apoptosis. It is postulated that high expression of Fas receptors or the presence of caspase-3 in the plasma membrane in leukemia cells might be responsible for the selective targeting of cancer cells. Additionally, an early increase in caspase-3 activity and intact mitochondrial membrane potential upon treatment with pancratistatin indicated involvement of an extrinsic pathway in apoptosis. Interestingly, DNA fragmentation did not occur prior to caspase-3 activation, indicating that pancratistatin's target is non-genomic. Apart from potent antitumor activity, (+)-pancratistatin (23) and (+)-7-deoxypancratistatin (24) also possess antiviral activity.65 Antiviral activity was observed against RNA flaviviruses (Japanese encephalitis, yellow fever and dengue) and bunya viruses (Punta Toro and Rift Valley viruses). In spite of its interesting biological profile, (+)-pancratistatin (23) and (+)-7-deoxypancratistatin (24) have found limited clinical application because of their low natural abundance and lack of practical synthetic route. Therefore, there is a need to design a practical scalable route for the preparation of multigram quantities of antitumor alkaloids (23) and (24). Several unnatural derivatives and analogues have been subjected to SAR (structure activity relationship) to identify the pharmacore. However, none of these are as potent as natural products (+)-pancratistatin (23) and (+)-7-deoxypancratistatin (24), indicating the necessity of the entire structure.73-76 17 Synthetic Strategies In addition to interesting biological activity, (+)-pancratistatin (23) and (+)-7-deoxypancratisatin (24) natural products have drawn considerable attention because of their structural complexity. The structure of (23) and (24) involves a highly functionalized cyclohexyl ring (C ring) with six contiguous stereocenters coupled through a carbon-carbon bond to a flat aromatic (A ring), with a trans-fused lactam forming the B ring of the molecule (Figure 2.1). There are numerous reports towards the total syntheses of this compounds (Table 2.1).73-93 The first total synthesis of racemic pancratistatin (23) was reported by Danishefsky in 198977, followed by the first asymmetric synthesis of pancratistatin by Hudlicky in 1995.78 Moreover, Pettit has developed synthesis of (+)-pancratistatin from (+)-narciclasine because of higher natural abundance of narciclasine in plant extracts.84 There have been attempts to synthesize simplified analogues and derivatives of pancratistatin as well.73-75 Despite these efforts, none of these approaches are suitable for commercially viable synthesis of pancratistatin (23) and 7-deoxypancratistatin (24). The majority of the synthetic routes that have been reported to date are too long or low yielding for practical preparations of the natural product (Table 2.1). Besides a short 10-step relay synthesis of (+)-pancratistatin from (+)-narciclasine by Pettit, only three other syntheses (Hudlicky, Madsen, Li) involve less than 15 steps (a number suggested by Hudlicky for synthetic practical applications). Though the syntheses of Hudlicky (entries 2, 11 and 12) and Madsen (entry 13) are comparatively short, they are rather low yielding. The shortest synthesis so far is that reported by Li's group (entry 9). The synthesis proceeds in 13 steps with 9% overall yield, however, uses expensive (+)-pinitol 18 as the starting material. Based on these criteria, none of the reported strategies are efficient for practical applications. Entry Isocarbostyril Year Author no. of stepsa Yield (%) 1 (?)-pancratistatin 1989 Danishefsky 27 0.16 2 (+)-pancratistatin 1995 Hudlicky 14 2 3 (+)-pancratistatin 1995 Trostb 19 8 4 (+)-pancratistatin 1997 Haseltine 24 0.97 5 (+)-pancratistatin 1998 Magnus 22 1.2 6 (+)-pancratistatin 2000 Rigby 23 0.35 7 (+)-pancratistatin 2001 Pettitc 10 3.6 8 (?)-pancratistatin 2002 Kim 21 4 9 (+)-pancratistatin 2006 Li 13 9 10 (+)-7-deoxypancratistatin 1995 Keck 22 4 11 (+)-7-deoxypancratistatin 1995 Hudlicky 12 2.6 12 (+)-7-deoxypancratistatin 1995 Hudlicky 10 3 13 (+)-7-deoxypancratistatin 1996 Chida 29 0.03 14 (+)-7-deoxypancratistatin 1998 Keck 15 12 15 (+)-7-deoxypancratistatin 2000 Plumet 21 3 16 (+)-7-deoxypancratistatin 2006 Madsen 13 1.4 17 (+)-7-deoxypancratistatin 2006 Madsen 15 4.3 18 (?)-7-deoxypancratistatin 2006 Padwa 23 3 a Number of steps and overall yield from commercially available starting material for the longest linear sequence. One-pot procedures are counted as one step. b Full procedures have not been disclosed. c Relay synthesis from (+)-narciclasine. Table 2.1: Reported total syntheses of the Amaryllidaceae isocarbostyrils. Adapted from ref 76. The most common strategy towards synthesis of pancratistatin (23) and 7-deoxypancratistatin (24) involve formation of a bond between aromatic ring (A ring) and more or less functionalized cyclohexane (C ring) ring. The approaches used for coupling of A and C rings include Michael addition (Plumet), nucleophilic addition (Magnus), electrophilic aromatic substitution (Haseltine), SN2/SN2' coupling reaction 19 (Hudlicky, Trost and Li), palladium-catalyzed coupling reaction (Chida) and photocyclization (Rigby). These synthetic strategies are discussed below. As mentioned earlier, these strategies are not viable for practical production of pancratistatin (23) and 7-deoxypancratistatin (24) (vide infra). Alternatively, there are less common approaches where the ring C is constructed after linkage of the aromatic ring to suitable precursor of ring C as reported by Danishfeky77, Kim85, Keck89, 90, Madsen92 and Padwa.93 (i) Michael Addition Heathcock was the first to construct ABC network based on the coupling of aryl ring with cyclohexyl ring.94 Aromatic anion 28 underwent 1,4-addition with nitrocyclohexene to give a mixture of the cis and trans aryl nitrocyclohexanes 29 (Scheme 2.1). The cis/trans mixture was equilibrated to obtain the trans isomer, which was used to generate lactam 30 in three subsequent steps. Scheme 2.1 OTBS CONEt2 O2N 1. 2. HOAc O O OTBS CONEt2 O O O2N OH O O NH O 28 29 30 70% Plumet devised synthesis of (+)-7-deoxypancratistatin (24) via ring opening of vinyl sulfone 32 (Scheme 2.2).91 1,4-addition of aromatic anion 31 to vinyl sulfone 32 gave cyclohexenol 33 with the correct configuration at C10b. Epoxidation of cyclohexenol 33, followed by oxirane opening with concomitant intramolecular 20 lactonization completed the synthesis of (+)-7-deoxypancratistatin (24). Plumet's synthesis is fairly long (21 steps) as well as low yielding (3%). Scheme 2.2 O O O OPhO2S O O O O OPhO2S OH96% 31 33 32 10b -78 ?C 24 10a O O NH O HO OH OH OHA B C Branchaud and Friestad proposed a palladium-catalyzed intramolecular 1,4-addition of aryl iodide on an chiral enone to construct bond between A and C rings.95,96 However, this route has not been extended to pancratistatin (23) and 7-deoxypancratistatin (24) to date. (ii) Nucleophilic Addition Magnus reported the total synthesis of (+)-pancratistatin (23) via intermolecular coupling of an aromatic anion 34 with the cyclohexanone derivative 35 to form 10a-10b carbon-carbon bond (Scheme 2.3).82 Nucleophilic addition of anion 34 to the cyclohexanone 35 gave alcohol 36. Benzylic alcohol 36 was converted to ketone 37 in three steps. The treatment of ketone 37 with the chiral amide 38 in the presence of LiCl gave asymmetric lithium enolate, which was trapped with TIPSOTf to form silyl enol ether 39. Further manipulations completed the synthesis of (+)-pancratistatin (23) (22 steps) in a low yield of 1.2%. 21 Scheme 2.3 O O OMe O O O O O OMe OTIPS O O OMe OH85%34 35 36 39 10a 10b 10a 10b -78 ?C O O OMe O 37 10a 10b Ph NLi Ph Me Me LiCl, TIPSOTf -78 ?C, 95% 383 Steps O O O O NH O HO OH OH OH OH A B C 7 23 (iii) Electrophilic Aromatic Substitution Haseltine accomplished a formal total synthesis of (+)-pancratistatin (23) via Danishfesky's intermediate 42.81 The key bond between A and C rings was constructed using an intramolecular electrophilic aromatic substitution (Scheme 2.4). Triflation of alcohol 40 in the presence of 2,6-lutidine gave the expected cyclized product 41. Several protection and deprotection steps provided Danishfesky's lactone intermediate 42 constituting a formal total synthesis (24 steps, 0.97% yield). It is interesting to note that when a donor substituent (OBn) is present on the aromatic ring (on carbon 7), the desired cyclized adduct was obtained in very low yield (8%). Instead, a major regioisomer arising from an undesired rearrangement was observed. Electrophilic aromatic substitution has also been applied to construct 7-deoxynarciclasine (26) analogue via epi-selenonium ion.97 22 Scheme 2.4 O O O Tf2O, LutidineO O OH O O O O O 73% 7 40 41 O O O OBn OH 42 OBn O Danishefsky's Intermediate 9 Steps (+)-pancratistatin 23 (iv) SN2 and SN2? Coupling Hudlicky's Approach The first asymmetric synthesis of (+)-pancratistatin (23) by Hudlicky involved opening of aziridine ring 44 by aromatic cuprate reagent 43 (Scheme 2.5).78 The regioselective SN2 ring opening of aziridine 44 led to inversion at C10b, establishing the desired stereochemistry. Subsequent key steps included installation of trans-diol and the formation of lactam to complete the synthesis of (23). However, the presence of benzamide moiety led to lengthy manipulations to affect the lactamization and the functionalization of the double bond. To overcome problems associated with the manipulations of benzamide, the amide was introduced in (+)-7-deoxypancratistatin (24) as the last step after the coupling reaction (Scheme 2.6) in the second-generation synthesis.80,87 However, the coupling of cuprate 46 with aziridine 44 produced compound 45 in lower yield. The low yield might be due to the instability of the organolithium compound and the corresponding cuprate compared to the compound 43. 23 Scheme 2.5 OTBS CONMe2 O O CuCNLi2 2 O O N Ts OTBS O O CONMe2 O O NHTs-78 ?C to rt BF3?Et2O 75% 44 45 43 10b O O NH O HO OH OH OH OH A B C10a 23 To save the number of functional group interconversion steps, aziridine 48 was prepared where the amide carbon was also used as the protecting group of aziridine (Scheme 2.6).87 The ring opening of aziridine 48 by cuprate 46 produced carbamate 49. Similar synthetic sequence (using enantiomer of aziridine 48) was used to synthesize an enantiomer of 7-deoxypancratistatin.98 Hudlicky has also made considerable efforts towards the synthesis of analogues of pancratistatin (23) and 7-deoxypancratistatin (24) using a similar approach (Scheme 2.6).73-75 Though short (10 steps), the synthesis by Hudlicky is rather low yielding (3%) and arduous. The key steps that reduce the yield are aziridination and aziridine ring-opening. The aziridine ring-opening requires extremely low temperature, thus making this synthesis unsuitable for practical applications. Furthermore, both coupling partners 46 and 48 are moderately unstable, entailing a difficulty in their preparation. 24 Scheme 2.6 O O CuCNLi2 2 O O N Ts O O O O NHTs-78 ?C to rt BF3?Et2O 32% 44 4746 10b O O O O NHE 49 10bO O CuCNLi2 2 O O N E -78 to -30 ?C BF3?Et2O 34% 4846 2 Steps E=CO2Me O O NH O HO OH OH OHA B C 24 10a 10a Trost's Approach In the asymmetric synthesis of (+)-pancratistatin (23) by Trost, carbon-carbon bond between A and C rings was constructed by allylic substitution of the carbonate 51 via SN2? chemistry (Scheme 2.7).79 Reaction of allylic carbonate 51 with mixed cuprate 50 resulted in inversion of configuration to provide allylic-arylated system 52. This type of mixed cuprate 50 is unstable and significant optimization was required for this step. Due to the difficulty associated with purification of the azide, azide 52 was not isolated. Additionally, full procedures have not been disclosed for this synthesis. 25 Scheme 2.7 OCO2Me N3 O O O O CuCN OMe O O O O N3 OMe MgBr 51 52 50 23 0 ?C O O NH O HO OH OH OH OH A B C Li's Approach Li's synthetic approach towards (+)-pancratistatin (23) is outlined in the Scheme 2.8. Intramolecular nucleophilic opening of cyclic sulfate 53 using aryl cerium reagent constructed the key bond to generate 54.86 In comparison to organomagnesium and organolithium reagents, organocerium reagents are much milder to avoid side reactions. Though the synthesis represents the shortest (13 steps) and the most high yielding (9%) route, the starting material (+)-pinitol is relatively expensive ($1,027 per 10 g).75 Scheme 2.8 N OOR O O R O O OMe OS O O O t-BuLi, CeCl3 NR HO3SO OOR O O OMe O O (+)-pancratistatin72% Br 5453 R=MOM -78 ?C to rt 23 26 (iv) Palladium-Catalyzed Coupling Heck Reaction Ogawa synthesized (+)-7-deoxypancratistatin88 from the key intermediate 56 produced in the synthesis of (+)-7-deoxynarciclasine.99 Intramolecular Heck reaction of aryl bromide 55 generated intermediate 56 (Scheme 2.9). Interestingly this Heck reaction proceeds via anti elimination instead of generally observed syn elimination. The desired stereochemistry at C10b for the synthesis of (+)-7-deoxypancratistatin (24) was achieved via hydrogenation of alkene 56. The synthetic route towards (24) is the longest reported to date (29 steps) and is also extremely low yielding (0.03%). Scheme 2.9 NPMB O O O NPMB O O O (+)-7-deoxypancratistatin OR OR Br OPMB OR OR OPMB TlOAc, 140 ?C, 68% (+)-7-deoxynarciclasine Pd(OAc)2, DIPHOS 5655 26 24 R=MOM 10b 10a Suzuki Coupling Hudlicky employed palladium-catalyzed Suzuki coupling to construct the bond between A and C rings in narciclasine (25) (Scheme 2.10).98 The reaction of arylboronic acid 57 with vinyl bromide 58 formed alkene 59. Analogous approach has also been used by Banwell to construct an enantiomer of narciclasine (25).100 27 Scheme 2.10 O 5857 O O B(OH)2 OMe (+)-narciclasine NCO 2Me Br O O O O OMe O NCO 2Me Br O O Br Pd(PPh3)4 Na2CO3, 30% 59 25 Recently, Pandey and co-workers formed epi-7-deoxypancratistatin 63 via cross-coupling of arylboronic acid 60 with iodo enone 61 (Scheme 2.11).101 This was followed by intramolecular aza-Michael addition to generate cis-fused lactam ring in 63. Scheme 2.11 O O NH OH OH OH HO O O O B(OH)2 NHCBz O OR OR OR I Pd(0) 84% O OR OR OR NHCBz O O 61 60 62 63 R=MOM (v) Photocyclization Rigby's synthetic approach is based on a hydrogen bond controlled aryl enamide photocyclization (Scheme 2.12).83 Irradiation of enamide 64 at 254 nm in benzene established desired stereochemistry at C10b in compound 65 in 30% yield (60% based on recovered starting material). Further functionalization of C ring completed the synthesis 28 of (+)-pancratistatin (23). This synthesis is fairly long (23 steps) and enormously low yielding (0.35%) as well. Scheme 2.12 NPMB OTBSO OO H O O h?, PhH NPMB OTBSO OO H O O 30% 64 65 10b 10a O O NH O HO OH OH OH OH A B C 23 Analogous photocyclization approach was adapted in the synthesis of 7-deoxypancratistatin analogue 68 by Pandey (Scheme 2.13).102 Irradiation of carbamate 66 at 280 nm gave the cyclized product 67 as single isomer. Scheme 2.13 O O NH OH OHHO O 66 68 O O NCO2Me OR OR 67 O O O NCO2Me OR ORRO h?, DCN 68% R=TBS RESEARCH GOAL The purpose of this work is to design an efficient synthetic route to antitumor alkaloids pancratistatin (23) and 7-deoxypancratistatin (24) using palladium-catalyzed allylic-arylation coupling methodology, thus providing multi-gram quantities of these compounds for clinical evaluations. 29 In order to test the feasibility of aryl-allyl coupling, (?)-7-deoxypancratistatin analogue 69 was first synthesized. The retrosynthetic approach to the synthesis of 7-deoxypancratistatin analogue 69 is outlined in the Scheme 2.14. The trans diol in 69 will be generated from the alkene 70 via opening of epoxide. The B ring of the alkene 70 will be obtained from the carbamate 71 by Friedel-Crafts cyclization. The allyl-aryl bond between the A and C rings in carbamate 71 will be constructed from palladium-mediated coupling of the aryl siloxane 72 and allylic carbonate 73. The synthetic route to aryl siloxane 72 and allyl carbonate 73 had already been established in the DeShong group.34 Also, the synthesis of carbamate 71 had been accomplished by the coupling reaction between aryl siloxane 72 and allylic carbonate 73. However, the subsequent steps to form the B ring and installation of trans diol had not been optimized. Provided with carbamate 71, the goal of my research project was to optimize these subsequent steps and complete the synthesis of 7-deoxypancratistatin analogue 69. Scheme 2.14 O O Si(OEt)3 + NHCO2Et OCO2Et O O NH O HO OH O O NHCO2Et O O NH O A C B A C A C BA C 69 70 71 72 73 7-deoxypancratistatin analogue 30 RESULTS AND DISCUSSION Synthesis of Coupling Partners: Allylic Carbonate and Aryl Siloxane Allylic carbonate 73 was synthesized as outlined in the Scheme 2.15. The synthesis of carbonate 73 commenced from the commercially available cyclohexadiene 74. Diels-Alder reaction of cyclohexadiene 74 with the acyl nitroso dienophile (generated in situ) formed hydroxamate 75.103 The cleavage of N-O bond in hydroxamate 75 using molybdenum hexacarbonyl yielded allylic alcohol 76.104 Finally, acylation of 76 with ethyl chloroformate afforded allylic carbonate 73. Aryl siloxane 72 was synthesized as shown in the Scheme 2.16.20 The commercially available aryl bromide 77 underwent a Grignard reaction to generate the organomagnesium species which was quenched by tetraethylorthosilicate (Si(OEt)4) to form aryl siloxane 72. Scheme 2.15 ClCO2Et, pyridine O N CO2Et NHCO2Et OH OCO2Et NHCO2Et 74 75 76 73 EtOCONHOH, N+ Bu4IO4- Mo(CO)6, reflux , 82 ?C CHCl3/DMF, 71% MeCN/H2O, 62% CH2Cl2, rt, 83% 31 Scheme 2.16 Si(OEt)3O O O O Br 61% 77 72 (ii) Si(OEt)4, THF, -78 ?C to rt (i) Mg(0), THF, 75 ?C Coupling of Allylic Carbonate with Aryl Siloxane The coupling of allylic carbonate 73 with electron-rich aryl siloxane 72 in the presence of TBAF and Pd(dba)2 catalyst gave the allylic arylated coupling products 73 and 79, respectively, as 1:1.6 ratio of diastereomers in 81% yield (Scheme 2.17). Scheme 2.17 NHCO2Et OCO2Et O O Si(OEt)3 O O NHCO2Et Pd LnAr H NHCO2Et HH O O NHCO2Et 1 3 1 3 + 1 3 1 373 78 72 71 79 71:79 = 1.0:1.6 Pd(dba)2, TBAF 65 ?C, THF 81% The low regioselectivity in the products is consistent with the previously proposed model for regioselectivity in cyclohexenyl systems. Using the model developed in our lab, the pi-allyl Pd complex 78 derived from allylic carbonate 73 formed on the face opposite from the departing carbonate. This ?symmetrical? Pd-complex 78 does not have substituents on the face of the ?-complex which has Pd attached. Accordingly, subsequent reductive elimination of the aryl group from silicon is equally probable at the 1 and 3 positions and therefore, a modest regioselectivity was observed. The role of 32 electronic factors on the regioselectivity of the coupling reaction was unknown. However, the results from Szab?105 indicated that the coupling reaction would favor carbamate 79, as observed in the experiment. Surprisingly, when benzyl carbamate (CBz) analogue of 73 was used as the coupling partner with the aryl siloxane 72, the yield of coupling products (CBz analogues of 71/79) decreased (33-50%) dramatically. Scheme 2.18 NHCO2Et OCO2Et Si(OEt)3 NHCO2Et NHCO2Et + 73 67%, 80:81 = 1.0:1.7 Pd2dba3?CHCl3, TBAF 55 ?C, THF SnBu3 B(OH)2 80 81 PdCl2TFP2, KF rt, MeOH 80%, 80:81 = 1.0:3.0 51%, 80:81 = 1.0:1.0 Pd2(dba)3, AsPh3, LiCl 50 ?C, NMP Similar allyl-aryl coupling reaction can also be accomplished via Stille29 (tin) and Suzuki41 (boron) reaction (Scheme 2.18). Both Stille and Suzuki reaction worked well to produce regioisomers 80 and 81. However, the undesired coupling product 81 was obtained in much greater amount in the Suzuki reaction compared to Stille and Hiyama coupling reaction. 33 Generation of B ring and Installation of diol With carbamate 71 in hand, the next goal was to construct the B ring via Friedel-Crafts acylation. When both ethyl carbamate 71 and CBz analogue of 71 were subjected to Bischler-Napieralski cyclization using DMAP/Tf2O, the yield of lactam was very low (15-20%). Therefore, an alternative route was investigated. The mixture of regioisomeric carbamates 71 and 79 was subjected to Friedel-Crafts acylation using P2O5 and POCl3/Me3SiOSiMe3 (1:1) (Scheme 2.19).106 The cyclization reaction was completely regioselective where only regioisomer 71 underwent cyclization to form the lactam 20 in good yield (55-65%). Attempts to optimize the reaction conditions indicated that reaction worked well as indicated in the literature. However, when mixture of CBz analogues 71/79 was subjected to similar reaction conditions, the lactam 70 was obtained in low yield (15%) comparable to that using Tf2O/DMAP reagent. We decided to change the sequence of steps by forming the epoxide first and then perform the cyclization (Scheme 2.19). While the epoxidation of alkenes 71/79 proceeded in moderate yield to give epoxides 82/83, the cyclization of 82/83 using POCl3/P2O5 showed decomposition. This route was abandoned therefore. Having optimized the yield of lactam 70, the next goal was to install trans diol via epoxidation. Previous work in DeShong group had shown that the epoxidation of the double bond using m-CPBA107 gave a mixture of diastereomeric epoxides 84 in low yield (15-30%) (Scheme 2.20). Attempts to improve the yield of epoxides using more reactive epoxidation reagents led to extensive decomposition of the alkene. The subsequent step of epoxide opening of 84 gave poor yield of diol 69 as well. 34 Scheme 2.19 O O NHCO2Et O O O O NH O O NHCO2Et O NHCO2Et O O 71 79 + POCl3/Me3SiOSiMe3 P2O5, 85 ?C, 56% 70 82 83 m-CPBA, NaHCO3 40 ?C, 51% POCl3/Me3SiOSiMe3 P2O5, 85 ?C O O NH O O Scheme 2.20 O O NH O O O O NH O O O O NH HO OH 51% 15-30% i) H2O2, HCO2H, rt ii) NaOH, 90 ?C < 30% 70 84 69 m-CPBA, NaHCO3, rt A more direct method of introducing the trans diol functionality was used for the synthesis of diol 69 (Scheme 2.20). According to this one-pot trans-dihydroxylation, the 35 reaction of alkene 70 with hydrogen peroxide in formic acid generates diastereomeric epoxides 84a/84b (Scheme 2.21).108 Under acidic conditions (formic acid), the epoxides are protonated and then opened by formic acid to generate a mixture of alcohol formates 85a/85b. Because of trans-diaxial opening of epoxide to avoid intermediate boat conformation, only the desired trans relationship will be installed. Therefore, hydrolysis of mono-formates provides the desired diol 69 with a single stereoconfiguration. In principle, the stereoselectivity in the epoxidation reaction was of no consequence since the diaxial opening of each epoxide should be regiospecific and give only trans-diol 69. Scheme 2.21 O O NH O O O NH O HO OH 1. H2O2, HCOOH , rt 2. NaOH , 90 ?C O O NH O O O NH O O O H H O H H O H H Nu Nu O O NH O OH OHCO O O NH O OCHO HO NaOH Nu = HCOOH H2O2, HCOOH 70 69 84a 84b 85a 85b 36 Much optimization was required for the isolation of diol 69 because of difficulty while working with such a polar compound, which is poorly soluble in most organic solvents (ethyl acetate, acetone, methanol). A variety of normal or reverse-phase methods led to an extensive loss of product. The most efficient purification method ultimately involved an extraction-precipitation regime, which gave trans diol in 51% yield (Scheme 2.20). Attempts to acylate hydroxyl groups to improve the solubilty of diol 69 in organic solvents failed, presumably because of sterical interference caused by peri hydrogen on C10 of aromatic A ring (Figure 2.1). The relative stereochemistry of diol 69 was confirmed by the COSY experiment and by correlation to the published 1H NMR data of the benzoate of the 7-deoxypancratistatin derivative 86 (Figure 2.3).109 The coupling constant of 2 Hz between H1 and H10b of 69 is diagnostic of a cis relationship between the two protons. Protons H4a and H10b of 69 exhibit a 13 Hz coupling constant which is indicative of a trans diaxial relationship. These coupling constant values are in accord with the coupling observed (J = 2 Hz) for protons H1 and H10b and (J = 13 Hz) for protons H4a and H10b respectively in 86. NH OH H4a O O O H H10b OH H1 NR OR H4a O O O H1 H10b OR OR ORH J 1, 10b = 2 Hz J 4a, 10b = 13 Hz J 1, 10b = 2 Hz J 4a, 10b = 13 Hz 69 86 Figure 2.3: Relative stereochemistry confirmed from correlation with 1H NMR coupling constants 37 CONCLUSION The synthesis of (?)-7-deoxypancratistatin analogue has been accomplished via palladium-catalyzed allylic-arylation. The key reaction in the synthesis involves the stereoselective formation of a carbon-carbon bond between the A and C rings by coupling of aryl siloxane with allylic carbonate. Subsequent steps include generation of B ring by Friedel-Crafts acylation and installation of the trans diol. The key coupling reaction is advantageous compared to Stille reaction because it avoids the use of toxic tin reagents. Additionally, the coupling is superior to the Suzuki reaction which results in formation of undesired regioisomer predominantly (Scheme 2.18). Our palladium-catalyzed allylic-arylation coupling methodology is also superior to allyl-aryl coupling by Hudlicky and Trost because of the ease of preparation of coupling partners, aryl siloxane and allyl carbonate as well as trouble-free coupling reaction. 38 EXPERIMENTAL DETAILS General Methods All reactions were run under an atmosphere of argon unless otherwise noted. Glassware used in the reactions was dried for a minimum of 12 h in an oven at 120 ?C or flame dried prior to use. Tetrahydrofuran was distilled from sodium/benzophenone ketyl, while methylene chloride, pyridine, methanol and N-methyl-2-pyrrolidone were distilled from calcium hydride. Phosphorous oxychloride was distilled from P2O5. Thin-layer chromatography (TLC) was performed on 0.25 mm silica gel coated plates treated with a UV-active binder with compounds being identified by one or more of the following methods: UV (254 nm), vanillin/sulfuric acid charring, or KMnO4 charring. Flash chromatography was performed using glass columns and medium pressure silica gel (Sorbent Technologies, 45-70 ?). Infrared spectra were recorded on a Nicolet 560 FT-IR spectrophotometer. Samples used for obtaining infrared spectra were either dissolved in carbon tetrachloride or taken neat. IR band positions are reported in reciprocal centimeters (cm-1) and relative intensities are listed as br (broad), s (strong), m (medium), or w (weak). Nuclear magnetic resonance (1H, 13C NMR) spectra were recorded on a 400 MHz spectrometer. Chemical shifts are reported in parts per million (?) and coupling (J values) are reported in hertz (Hz). Spin multiplicities are indicated by the following symbols: s (singlet), d (doublet), t (triplet), q (quartet), m (multiplet), br s (broad singlet), br d (broad doublet). Low resolution mass spectrometry (LRMS) and high resolution mass spectrometry (HRMS) were obtained on a JEOL SX-02A instrument. 39 Aryl Siloxane 72 1. Mg(0), 75 ?C 2. Si(OEt)4, rtO O Br Si(OEt)3 O O 61% 7277 To 811 mg (33.8 mmol, 2.00 equiv.) of washed magnesium turnings was added 10.0 mL of anhydrous THF. After heating the reaction mixture at 75 ?C, 2.00 mL (16.9 mmol, 1.00 equiv.) of the aryl bromide 77 was added dropwise under argon and reaction was allowed to stir for 45 min. This was followed by addition of 10.0 mL of anhydrous THF to the reaction mixture. After 4 h, the Grignard mixture was cannulated into 9.55 mL (42.3 mmol, 2.50 equiv.) of Si(OEt)4 in 20.0 mL anhydrous THF. The reaction was allowed to stir for 3 h under argon at room temperature. The black solution was extracted with 3 ? 200 mL Et2O and the organic layers were washed with 200 mL H2O. The combined organic extracts were dried over MgSO4, filtered and concentrated in vacuo to give a brown oil. Purification by column chromatography (2% EtOAc/98% hexane, Rf = 0.26) yielded 2.94 g (61%) of aryl siloxane 72 as a colorless oil: IR (CCl4) 2970 (s), 2926 (m), 2885 (m), 2780 (w), 2739 (w), 1611 (w), 1479 (s), 1234 (s), 1169 (s), 1095 (s) cm-1; 1H NMR (400 MHz, CDCl3) ? 7.16 (dd, J = 1, 8 Hz, 1H), 7.09 (d, J = 1 Hz, 1H), 6.84 (d, J = 8 Hz, 1H), 5.93 (s, 2H), 3.83 (q, J = 7 Hz, 6H), 1.20 (t, J = 7 Hz, 9H); 13C NMR (100 MHz, CDCl3) ? 149.3, 147.3, 129.2, 123.6, 113.8, 108.5, 100.5, 58.6, 18.1; LRMS (FAB) 284.4 (M+, 100), 283.4 (42), 239 (80), 161 (65); HRMS (FAB) calcd 284.1080 (M+), found 284.1080. 40 Hydroxamate 75 EtOCONHOH, N+Bu4IO4- O N CO2Et CHCl3/DMF, 0 ?C to rt 71% 7574 To 6.29 g (14.5 mmol, 1.40 equiv.) of NBu4+IO4- (tetrabutylammonium periodate) under argon was added chloroform (12.0 mL) and DMF (4.00 mL). The reaction mixture was cooled to 0 ?C and then 0.990 mL (104 mmol, 1.00 equiv.) of cyclohexadiene 74 was added. Finally, 1.09 g (104 mmol, 1.00 equiv.) of hydroxamic acid (EtOCONHOH) dissolved in chloroform (6.00 mL) and DMF (2.00 mL) was added via addition funnel. The reaction mixture was stirred at 0 ?C to room temperature for 17 h. The product was extracted with Et2O (4 ? 100 mL), washed with H2O (100 mL), dried over MgSO4, concentrated in vacuo to give the crude hydroxamate 75 as a brown oil. Flash column chromatography on silica gel (50% EtOAc/50% hexane, Rf = 0.49) gave 1.36 g (71%) of hydroxamate 75 as a light orange oil: IR (CCl4) 3062 (w), 2984 (m), 2936 (m), 2862 (w), 1706 (s), 1380 (s), 1268 (s), 1081 (s) cm-1; 1H NMR (400 MHz, CDCl3) ? 6.55-6.46 (m, 2H) , 4.76 (m, 1H), 4.70 (m, 1H), 4.18-4.07 (m, 2H), 2.17-2.04 (m, 2H), 1.45 (qt, J = 2, 12 Hz, 1H), 1.33 (t, J = 12 Hz, 1H), 1.20 (dt, J = 2, 7 Hz, 3H); 13C NMR (100 MHz, CDCl3) ? 158.3, 132.0, 131.6, 71.0, 62.2, 49.9, 23.4, 20.6, 14.5; EI mass spectrum, m/z (relative intensity) 183 (M+, 98), 105 (72), 79 (92), 77 (80), 67 (65), 29 (76), 27 (65); HRMS (ESI) calcd for C9H14NO3 (M+H)+ 184.0974, found 184.0974. 41 Alcohol-Carbamate 76 O N CO2Et Mo(CO)6 NHCO2Et OH CH3CN/H2O, reflux 62% 7675 To 7.71 g (42.1 mmol, 1.00 equiv.) of the hydroxamate 75 dissolved in 240 mL acetonitrile and 15 mL distilled water was added 12.2 g (46.3 mmol, 1.10 equiv.) of molybdenum hexacarbonyl (Mo(CO)6). The reaction mixture was refluxed at 82 ?C for 66 h. The black-brown reaction mixture was vacuum filtered through Celite and the filtrate was concentrated in vacuo to give crude alcohol-carbamate 76 as a yellow oil. Flash column chromatography on silica gel (75% EtOAc/25% hexane, Rf = 0.33) gave 4.81 g (62%) of carbamate 76 as a yellow oil: IR (CCl4) 3619 (m), 3450-3100 (br s), 3028 (m), 2981 (s), 2940 (s), 1713 (s), 1502 (s), 1316 (s), 1231 (s), 1067 (s) cm-1; 1H NMR (400 MHz, CDCl3) ? 5.86 (d, J = 10 Hz, 1H), 5.73 (dd, J = 2, 10 Hz, 1H), 4.67 (br s, 1H), 4.16-4.07 (m, 4H), 1.88-1.80 (m, 2H), 1.72-1.63 (m, 2H), 1.53 (br s, 1H), 1.22 (t, J = 7 Hz, 3H); 13C NMR (100 MHz, CDCl3) ? 156.0, 132.5, 130.9, 64.5, 60.8, 46.2, 28.9, 25.6, 14.6; EI mass spectrum, m/z (relative intensity) 185 (M+, 4), 157 (100), 141 (92), 96 (76), 68 (87), 55 (99), 39.5 (99.5); HRMS calcd for C9H15NO3 (M+) 185.1053, found 185.1052. 42 Carbonate-Carbamate 73 ClCO2Et NHCO2Et OH OCO2Et NHCO2Et pyridine/ CH2Cl2 rt, 83% 76 73 To 2.84 g (15.4 mmol, 1.00 equiv.) of alcohol-carbamate 76 in 35.0 mL anhydrous CH2Cl2 and 1.85 mL (23.0 mmol, 1.50 equiv.) anhydrous pyridine was added 2.28 mL (23.0 mmol, 1.50 equiv.) of ethyl chloroformate dropwise via syringe under argon. The reaction was allowed to stir at room temperature for 5 days. The reaction mixture was extracted with CH2Cl2 (3 ? 50 mL), washed with H2O (50 mL), dried over MgSO4 and concentrated in vacuo. Flash chromatography on silica gel (30% EtOAc/70% hexane, Rf = 0.43 afforded 3.28 g (83%) of the carbonate-carbamate 73 as a yellow oil: IR (CCl4) 3450 (m), 3042 (w), 2987 (w), 1747 (s), 1727 (s), 1502 (s), 1265 (s) cm-1; 1H NMR (400 MHz, CDCl3) ? 5.86 (m, 2H), 5.03 (m, 1H), 4.64 (br s, 1H), 4.19-4.14 (m, 3H), 4.08 (q, J = 7 Hz, 2H), 1.88 (br s, 3H), 1.67-1.63 (m, 1H), 1.28 (t, J = 7 Hz, 3H), 1.21 (t, J = 7 Hz, 3H); 13C NMR (100 MHz, CDCl3) ? 156.2, 155.0, 134.4, 128.3, 127.7, 70.3, 64.2, 61.2, 46.6, 26.2, 25.6, 14.9, 14.5; FAB mass spectrum, m/z (relative intensity) 258 ((M+H)+, 2), 168 (100), 90 (40), 62 (46); HRMS (FAB) calcd for C12H19O5N (M+H)+ 258.1332, found 258.1342. 43 Carbamates 80 and 81 NHCO2Et OCO2Et NHCO2Et NHCO2Et + 73 M 80 81 PdLn M = Si, Sn, B Hiyama-like Coupling The coupling reaction of carbonate 73 with PhSi(OEt)3 had been performed previously by Bogaczyk in the DeShong group.31 To a solution of 100 mg (0.389 mmol) of allyl carbonate 73 and 20 mg (0.020 mmol) Pd2(dba)3?CHCl3 in 10 mL dry THF was added 190 ?L (0.778 mmol) PhSi(OEt)3, followed by 778 ?L (0.778 mmol, 1.0 M solution in THF) TBAF. The solution was degassed via a single freeze-pump-thaw cycle and then heated to 55 ?C. The reaction mixture was quenched after 48 h with 40 mL H2O. The layers were separated and the aqueous phase was extracted with 3 ? 40 mL Et2O. The combined organic layers were dried over Na2SO4 and concentrated in vacuo. Purification of the residue by flash chromatography (20 mm, 20 cm, 5% EtOAc/95% hexane) gave 48 mg (51%) of carbamates 80:81 as a white solid. The ratio of carbamates 80 to 81 was 1.0:1.0. The pure regioisomers were separated using preparative HPLC (3:1 hexane:EtOAc). Carbamate 80: TLC Rf = 0.35 (3:1 hexane:EtOAc); IR (CCl4) 3447 (w), 3029 (w), 2933 (w), 1729 (s), 1552 (s) cm-1; 1H NMR (CDCl3) ? 7.21-7.29 (m, 5H), 5.89 (ddd, J = 2, 4, 10 Hz, 1H), 5.62 (ddd, J = 2, 4, 10 Hz, 1H), 4.80 (br s, 1H), 4.01 (q, J = 7 Hz, 2H), 3.78-3.83 (m, 1H), 3.29-3.33 (m, 1H), 2.14-2.24 (m, 2H), 1.87-1.92 (m, 1H), 1.58-1.62 (m, 1H), 1.16 (t, J = 7 Hz, 3H); 13C NMR (CDCl3) ? 156.0, 142.6, 128.4 44 (2C), 128.3, 127.9, 126.7, 60.6, 52.6, 48.0, 25.5, 23.0, 14.5; FAB mass spectrum m/z (relative intensity) 246 ((M + H), 17), 102 (100), 91 (25); HRMS (FAB) calcd for C15H20O2N (M+H) 246.1494, found 246.1491. Carbamate 81: TLC Rf = 0.35 (3:1 hexane:EtOAc); IR (CCl4) 3454 (w), 3026 (w), 2943 (w), 1728 (s), 1500 (s) cm-1; 1H NMR (CDCl3) ? 7.16-7.31 (m, 5H), 5.82 (d, J = 10 Hz, 1H), 5.77 (d, J = 10 Hz, 1H), 4.63 (br s, 1H), 4.32-4.34 (m, 1H), 4.12 (q, J = 7 Hz, 2H), 3.36-3.38 (m, 1H), 2.07-2.09 (m, 2H), 1.58-1.63 (m, 1H), 1.47-1.52 (m, 1H), 1.25 (t, J = 7 Hz, 3H); 13C NMR (CDCl3) ? 156.0, 145.2, 133.1, 129.6, 128.4, 127.5, 126.3, 60.7, 47.0, 41.8, 31.0, 29.6, 14.6; FAB mass spectrum m/z (relative intensity) 246 ((M + H), 17), 157 (66), 91 (100); HRMS (FAB) calcd for C15H20O2N (M+H) 246.1494, found 246.1491. Suzuki Coupling To 60.2 mg (0.234 mmol, 1.00 equiv.) of allyl carbonate 73 under argon was added 13.6 mg (0.0234 mmol, 0.100 equiv.) PdCl2TFP2. This was followed by addition of 57.1 mg (0.468 mmol, 2.00 equiv.) PhB(OH)2, 57.4 mg (0.936 mmol, 4.00 equiv.) KF and 4.00 mL dry MeOH. The reaction mixture was stirred at room temperature for 24 h. The product was extracted with Et2O (5 ? 20 mL), washed with H2O (20 mL), dried over MgSO4, concentrated in vacuo to give the crude as a yellow oil. Flash column chromatography on silica gel (10% EtOAc/90% hexane, Rf = 0.14) gave 46.0 mg (80%) of carbamates 80 and 81, respectively, in a 1.0:3.0 ratio as a yellow solid. Stille Coupling To 70.4 mg (0.274 mmol, 1.00 equiv.) carbonate 73 and 201 mg (0.548 mmol, 2.00 equiv.) PhSnBu3 in 10 mL dry NMP was added 25.2 mg (0.0822 mmol, 3.00 equiv.) AsPh3, 75.3 mg (0.0822 mmol, 3.00 equiv.) Pd2(dba)3 and 69.7 mg (1.64 mmol, 6.00 45 equiv.) LiCl. The reaction mixture was stirred at 50 ?C for 24 h. The product was extracted with Et2O (5 ? 20 mL), washed with H2O (20 mL), dried over MgSO4, concentrated in vacuo to give the crude as a yellow oil. Flash column chromatography on silica gel (20% EtOAc/80% hexane, Rf = 0.33) gave 45.0 mg (67%) of carbamates 80 and 81, respectively, in a 1.0:1.7 ratio as a yellow solid. Carbamates 71 and 79 NHCO2Et OCO2Et O O Si(OEt)3 O O NHCO2Et O O NHCO2Et + 73 72 71 7971:79 = 1.0:1.6 Pd(dba)2, TBAF 65 ?C, THF 81% To 338 mg (1.19 mmol, 2.00 equiv.) of aryl siloxane 72 nd 153 mg (0.595 mmol, 1.00 equiv.) of carbonate-carbamate 73 dissolved in 15.0 mL anhydrous THF was added 1.19 mL (1.19 mmol, 2.00 equiv.) TBAF under argon. This was followed by addition of 68.4 mg (0.119 mmol, 0.100 equiv.) Pd(dba)2. The reaction mixture was subjected to one freeze pump thaw cycle and then heated at 65 ?C for 24 h. The reaction was then quenched by addition of 30 mL H2O. The product was extracted with 3 ? 25 and washed with H2O. The combined organic extracts were dried over MgSO4, filtered and concentrated in vacuo to give a brown oil. Flash column chromatography on silica gel (15% EtOAc/85% hexane, Rf = 0.20) gave 140 mg (81%) of carbamates 71 and 79, respectively, in a 1.0:1.6 ratio as a yellow solid. The two regioisomers were inseparable using column chromatography and were thus carried as mixture through the next step. 46 The regioisomers however, are separable by HPLC.34 A small amount of the mixture was separated on preparative HPLC (25% EtOAc/75% hexane) for spectral analysis. Carbamate 71: mp 111-113 ?C; IR (CCl4) 3443 (w), 2930 (w), 2858 (w), 1727 (m), 1550 (s), 1502 (m), 1485 (m), 1251 (m) cm-1; 1H NMR (400 MHz, CDCl3 ? 6.73-6.67 (m, 3H), 5.90 (s, 2H), 5.87 (dd, J = 2, 10 Hz, 1H), 5.58 (dd, J = 2, 10 Hz, 1H) 4.72 (br s, 1H), 4.03 (q, J = 7 Hz, 2H), 3.72 (br s, 1H), 3.22 (s, 1H), 2.25-2.10 (m, 2H), 1.91-1.88 (m, 1H), 1.58 (sextet, J = 7 Hz, 1H) 1.18 (t, J = 7 Hz, 3H); 13C NMR (100 MHz, CDCl3) ? 156.0, 147.6, 146.2, 136.5, 128.0, 127.9, 121.5, 108.7, 108.0, 100.9, 60.6, 52.7, 47.6, 25.5, 22.9, 14.5; FAB mass spectrum m/z (relative intensity) 290 ((M+H)+, 21), 201 (100), 174 (31), 135 (81), 73 (90); HRMS (FAB) calcd for C16H20O4N (M+H)+ 290.1392, found 290.1379. Carbamate 79: mp 90-92 ?C; IR (CCl4) 3446 (w), 3028 (w), 2943(w), 1727 (s), 1547 (s), 1489 (s) cm-1; 1H NMR (400 MHz, CDCl3) ? 6.72 (d, J = 8 Hz, 1H), 6.64-6.59 (m, 2H), 5.90 (m, 2H), 5.75 (m, 2H), 4.63 (br s, 1H), 4.29 (br s, 1H), 4.10 (q, J = 7 Hz, 2H), 3.29-3.28 (m, 1H), 2.07-2.01 (m, 2H), 1.54-1.41 (m, 2H), 1.23 (t, J = 7 Hz, 3H); 13C NMR (100 MHz, CDCl3) ? 156.1, 147.6, 146.0, 139.2, 133.2, 129.6, 120.4, 108.1, 108.0, 100.8, 60.7, 46.9, 41.4, 31.1, 29.5, 14.6; FAB mass spectrum m/z (relative intensity) 290 ((M+H)+, 46), 201 (100), 174 (24), 135 (60), 73 (48); HRMS (FAB) calcd for C16H20O4N (M+H)+ 290.1392, found 290.1390. 47 Lactam 70 O O NHCO2Et O O O NH O NHCO2Et O + POCl3, P2O5 Me3SiOSiMe3 71 79 70 85 ?C, 56% To 294 mg (1.02 mmol, 1.00 equiv.) of coupling product mixture 71/79 was added P2O5 (3.13 g, 21.6 equiv.) and then 5.00 mL (23.0 equiv.) hexamethyldisiloxane under argon. This was followed by dropwise addition of POCl3 (5.00 mL) via syringe. The reaction mixture was stirred at 85 ?C for 19 h. The purple reaction mixture was quenched with 100 mL ice water and was stirred for 5.5 h at room temperature. The product was extracted with EtOAc (7 ? 75 mL), dried over MgSO4, concentrated in vacuo to give crude as a yellow solid. Flash column chromatography on silica gel (75% EtOAc/25% hexane, Rf = 0.23) gave 48 mg (56%, based on the amount of carbamate 71 in the original mixture) of alkene 70 as a white solid: mp >300 ?C; IR (Neat) 3185 (w), 2920 (m), 1665 (s), 1612 (m), 1451 (s), 1254 (s), 1036 (w) cm-1; 1H NMR (400 MHz, CDCl3) ? 7.55 (s, 1H), 6.85 (s, 1H), 6.09-6.07 (m, 2H 1H), 6.00 (s, 2H), 5.88-5.86 (m, 1H), 3.46-3.44 (m, 1H), 2.26 (m, 2 H), 1.94-1.93 (m, 1H), 1.86-1.82 (m, 1H); 13C NMR (100 MHz, CDCl3) ? 166.0, 151.2, 146.4, 136.6, 128.8, 123.7, 122.7, 108.7, 103.6, 101.6, 53.5, 41.1, 28.1, 24.5; HRMS (FAB) calcd for C14H13NO3 244.0973, found 244.0974. 48 Epoxides 82 and 83 O O NHCO2Et O O O NHCO2Et O NHCO2Et O O71 + 82 83 m-CPBA, NaHCO3 CH2Cl2, 40 ?C, 51% O NHCO2Et O 79 + To 209 mg (0.723 mmol, 1.00 equiv.) of mixture of coupling products 71 and 79 dissolved in anhydrous CH2Cl2 was added 243 mg (2.89 mmol, 4.00 equiv.) of sodium bicarbonate (NaHCO3), followed by addition of 375 mg (2.17 mmol, 3.00 equiv.) of m-CPBA. The reaction mixture was stirred at 40 ?C for 25 h and was quenched with 25.0 mL of saturated NaHCO3 and 25 mL of saturated Na2S2O3 (sodium thiosulfate) solution. The product was extracted with 4 ? 50 mL EtOAc and washed with 50 mL water. The combine organic layers were dried over MgSO4 and concentrated in vacuo to give crude as a yellow solid. Flash chromatography on silica gel (30% EtOAc/70% hexane) yielded 41 mg of epoxide 82 (Rf = 0.25) and 72 mg of epoxide 83 (Rf = 0.34) as white solids (51% combined yield): Epoxide 82: mp 178-181 ?C; IR (Neat) 3314 (m), 2979 (w), 2921 (w), 2845 (w), 1684 (s), 1540 (s), 1473 (s), 1282 (m), 1238 (s), 1051 (s) cm-1; 1H NMR (400 MHz, CDCl3) ? 6.77-6.69 (m, 3H), 5.93 (s, 2H), 4.87 (br s, 1H), 3.99 (q, J = 7 Hz, 2H), 3.60 (br s, 1H), 3.31 (m, 1H), 3.17 (d, J = 4 Hz, 1H), 2.96 (d, J = 7 Hz, 1H), 2.21- 2.16 (m, 1H), 2.06-2.00 (m, 1H), 1.69-1.65 (m, 1H), 1.46-1.38 (m, 1H) , 1.15 (t, J = 7 Hz, 3H); 13C NMR (100 MHz, CDCl3) ? 155.7, 147.9, 146.6, 134.3, 121.4, 108.5, 108.4, 101.0, 60.6, 56.0, 52.4, 51.9, 46.8, 22.6, 22.4, 14.5. HRMS (ESI) calcd for C16H20O5N (M+H)+ 306.1342, found 306.1310. Epoxide 83: mp 157-160 ?C; IR (Neat) 3295 (m), 49 2988 (w), 2936 (w), 2859 (w), 1675 (s), 1531 (s), 1488 (s), 1440 (m), 1243 (s), 1042 (s) cm-1; 1H NMR (400 MHz, CDCl3) ? 6.76-6.65 (m, 3H), 5.93 (s, 2H), 4.91 (d, J = 8 Hz, 1H), 4.15-4.10 (m, 3H), 3.36 (m, 1H), 3.25 (d, J = 4 Hz, 1H), 2.97 (dd, J = 6, 4 Hz, 1H), 1.87-1.81 (m, 1H), 1.72-1.67 (m, 1H), 1.37-1.31 (m, 2H), 1.24 (t, J = 7 Hz, 3H); 13C NMR (100 MHz, CDCl3) ? 156.1, 147.8, 146.2, 137.1, 120.6, 108.4, 108.1, 101.0, 60.9, 59.1, 55.0, 48.0, 40.2, 30.0, 24.6, 14.6. HRMS (ESI) calcd for C16H20O5N (M+H)+ 306.1342, found 306.1310. Epoxides 84a, 84b O O NH O O O NH O O O O NH O O rt 70 84a/b 84a/b m-CPBA NaHCO3, CH2Cl2 To a solution of the lactam 70 (59 mg, 0.24 mmol, 1.0 equiv.) in dichloromethane (2.0 mL), was added sodium bicarbonate (41 mg, 0.48 mmol, 2.0 equiv.), followed by purified m-CPBA (63 mg, 0.36 mmol, 1.5 equiv.). The two phase reaction mixture was vigorously stirred for 24 h at room temperature, before diluting with water (2.0 mL) and stirring for another 30 min. The two phases were separated, and then extracted with dichloromethane (3 ? 5 mL). The combined organics were washed with satd. aq. sodium bicarbonate (2 ? 3 mL), dried over MgSO4 and concentrated in vacuo. Flash chromatography on silica gel (gradient elution 75%-100% EtOAc/hexane) gave faster eluting 84 (3 mg, 5%) and slower eluting 84 (6 mg, 9%), both as white solids. Faster eluting 84: TLC Rf = 0.37 (80% EtOAc/20% hexane); mp >300 ?C; IR (CCl4) 2957 (w), 2930 (s), 2851 (m), 1720 50 (s), 1360 (w), 1217 (w), 909 (m) cm-1; 1H NMR (400 MHz, CDCl3) ? 7.56 (s, 1H), 7.00 (s, 1H), 6.03 (s, 2H), 3.54 (d, J = 4 Hz, 1H), 3.44 (br s, 1H), 3.29 (dd, J = 2, 2 Hz, 1H), 3.23 (dd, J = 4, 12 Hz, 1H), 3.00 (d, J = 12 Hz, 1H), 1.95-1.90 (m, 1H), 1.65-1.59 (m, 3H); HRMS (FAB) calcd for C14H13NO4 260.0924 (M+H), found 260.0923. Slower eluting 84: TLC Rf = 0.21 (80% EtOAc/20% hexane); mp >300 ?C; 1H NMR (400 MHz, CDCl3) ? 7.51 (s, 1H), 6.97 (s, 1H), 6.02 (s, 2H), 5.97 (br s, 1H), 3.72 (s, 1H), 3.51 (dt, J = 3, 12 Hz, 1H), 3.36 (t, J = 3 Hz, 1H), 3.11 (d, J = 12 Hz, 1H), 2.20-2.16 (m, 1H), 2.07- 2.02 (m, 1H), 1.69-1.59 (m, 2H). Diol 69 O O O NH O O O NH HO OH 1. H2O2, HCO2H, rt 2. 1M NaOH, 90 ?C 70 69 51% To a solution of alkene 70 (68 mg, 0.28 mmol, 1.0 equiv.) in formic acid (1.7 mL) was added 0.15 mL of 30% aqueous hydrogen peroxide. This yellow solution was stirred for 14 h at room temperature and then volatile material was removed on rotary evaporator to obtain 90 mg white solid. Aqueous sodium hydroxide (1 M, 0.35 mL, pH = 9) and 2.0 mL MeOH was then added and reaction mixture was heated at 90 ?C for 3 h. The reaction mixture was cooled and solvents were evaporated to obtain a light brown solid. Washing the crude product with copious amount of hot ethyl acetate (30 ? 30 mL) gave 60 mg light yellow solid. An attempt to recrystallize the crude solid using acetone/methanol failed because of poor solubility of the product in most organic solvents as well as water. 51 Instead, white solid precipitated when the acetone/methanol solution was cooled at 0 ?C for a week. The solution was vacuum filtered to obtain product as a white solid. The filtrate was cooled at 0 ?C and the process was repeated two additional times to obtain 40 mg (51%) of diol 69 as a white solid: mp 232-234 ?C; IR (Neat) 3356-3199 (br s), 2938 (w), 2892 (w), 1629 (s), 1601 (s), 1469 (s), 1261 (s), 1040 (s) cm-1; 1H NMR (400 MHz, DMSO) ? 7.64 (s, 1H), 7.28 (s, 1H), 6.89 (s, 1H), 6.05 (d, J = 2 Hz, 2H), 4.86 (d, J = 5 Hz, 1H), 4.81 (d, J = 4 Hz, 1H), 4.21 (m, 1H), 3.76 (m, 1H), 3.48 (ddd, J = 8, 8, 12 Hz, 1H), 2.85 (dd, J = 2, 13 Hz, 1H), 1.79-1.74 (m, 1H), 1.65-1.60 (m, 1H), 1.53 (dd, J = 2, 13 Hz, 1H); 13C NMR (100 MHz, DMSO) ? 163.8, 150.2, 145.6, 136.1, 124.3, 106.7, 104.9, 101.3, 68.2, 67.3, 49.0, 40.6, 25.5, 25.3; HRMS (FAB) calcd for C14H15NO5 (M+H)+ 278.1029, found 278.1028. 52 CHAPTER 3 FORMAL TOTAL SYNTHESIS OF (?)-7-DEOXYPANCRATISTATIN INTRODUCTION We have successfully applied palladium-catalyzed allylic-arylation to the synthesis of (?)-7-deoxypancratistatin analogue (69) (see Chapter 2) by coupling an aryl siloxane 72 with allylic carbonate 73 (Scheme 3.1).110 The reaction proceeds via formation of a ?symmetrical? ?-allyl palladium complex 78 (Scheme 3.2). In this instance, reductive elimination from complex 78 is equally probable onto either carbon 1 and 3, resulting in formation of two regioisomers 71 and 79. Scheme 3.1 O O Si(OEt)3 + NHCO2Et OCO2Et O O NH O HO OH O O NHCO2Et A C A C BA C 697172 73 Pd(0) TBAF Scheme 3.2 NHCO2Et OCO2Et O O Si(OEt)3 O O NHCO2Et Pd LnAr H NHCO2Et HH O O NHCO2Et 1 3 1 3 + 1 3 1 373 78 72 71 79 71:79 = 1.0:1.6 Pd(dba)2, TBAF 65 ?C, THF 81% 53 Having demonstrated that this coupling reaction would occur, the next goal was to extend siloxane methodology to the synthesis of (?)-7-deoxypancratistatin (24). As shown in the retrosynthesis presented in Scheme 3.3, the coupling of aryl siloxane 72 with allylic carbonate 87 would form the desired bond between A and C rings to produce carbamate 49. In comparison to allyl carbonate 73 (Scheme 3.1), allyl carbonate 87 has large isopropylidene protected diol portion. We anticipate the coupling of allylic carbonate 87 will result in formation of ?unsymmetrical? ?-allyl palladium complex 88 (Scheme 3.4). Due to steric bulk provided by isopropylidene group, palladium would predominantly reside toward carbon 3 in complex 88. Subsequently, reductive elimination would preferentially result in formation of carbamate 49, the desired regioisomer required for the synthesis of (24). Scheme 3.3 O O Si(OEt)3 + NHE OCO2Et O O O O NH O HO OH OH OH O O NHE O O A CA C BA C 24 49 72 87 E = Ester Scheme 3.4 O O Si(OEt)3 + NHE OCO2Et O O O O NHE O O 87 Pd(0) TBAF 72 49 O O NHE PdPh Ln 88 1 3 1 3 1 3 54 RESULTS AND DISCUSSION Synthesis of Coupling Partners: Allylic Carbonate and Aryl Siloxane Allylic carbonate 87 was synthesized using strategy similar to that used for carbonate 73 (see Chapter 2) from 1,4-cyclohexadiene. Diene 93 was synthesized from commercially available 1,4-cyclohexadiene (89) using Yang's procedure (Scheme 3.5).111 Dibromination of diene 89 at low temperature afforded dibromoalkene 90, which was dihydroxylated to obtain cis-diol 91. Protection of diol 91 with 2,2-dimethoxypropane gave acetonide 92. Subsequent dehydrobromination of 92 with DBU provided diene 93. Scheme 3.5 Br Br O O 89 90 93 Br2, CHCl3 -78 ?C, 82% OsO4, NMO acetone, rt, 93% Br Br OH OH 91 p-TsOH, CH2Cl2, rt, 82% OMe DBU, benzene reflux, 49% Br Br O O 92 MeO Diene 93 was used to prepare allyl carbonate 87 (Scheme 3.6) using the methodology developed for the model system (see Chapter 2). Diels-Alder reaction of diene 93 with acyl nitroso dienophile (generated in situ) provided racemic hydroxamate 94.103 Reduction of N-O bond to generated allylic alcohol 95104 and subsequent protection of alcohol with chloroformate provided allylic carbonate 87. 55 Scheme 3.6 OH NHE O O OCO2Et NHE O O N O O O ClCO2Et, pyridine CH2Cl2, rt NBu4+IO4- Mo(CO)6, MeCN/H2O reflux O O 93 94 E = CO2Me, 31% 95 87 E H NHO E CHCl3/DMF, 0 ?C to rt E = CO2Et, 57% E = CO2Me, 79% E = CO2Et, 67% E = CO2Me, 86% E = CO2Et, 93% Aryl siloxane 72 was synthesized as shown in the Scheme 3.7.20 The commercially available aryl bromide 77 underwent a Grignard reaction to generate the organomagnesium species which was quenched by tetraethylorthosilicate (Si(OEt)4) to form aryl siloxane 72. Scheme 3.7 Si(OEt)3O O O O Br 61% 77 72 (ii) Si(OEt)4, THF, -78 ?C to rt (i) Mg(0), THF, 75 ?C 56 Coupling of Allylic Carbonate with Aryl Siloxane Preliminary Attempts It was anticipated that palladium-catalyzed coupling of allylic carbonate 87 with aryl siloxane 72 would yield the coupling product 49 with high regioselectivity based on the formation of ?unsymmetrical? ?-allyl palladium complex 88 (Scheme 3.8). However, repeated attempts to couple allylic carbonate 87 and aryl siloxane 72 were unsuccessful and no traces of carbamate 49 were detected. The use of a more active catalyst such as ?-allyl palladium chloride dimer did not form carbamate 49, but interestingly gave arene ether 96.34 Scheme 3.8 O O Si(OEt)3OCO 2Et NHE O O O O NHE O O O O O NHE O O (allylPdCl)2, PPh3 TBAF, THF, 60 ?C 10 % 72 O O NHE PdPh Ln 88 96 1 3 1 3 87 49 E = CO2Et The plausible mechanism for the formation of arene ether 96 is outlined in the Scheme 3.9. Upon reaction of palladium with allylic carbonate 87, the carbonate rearranged to form carbonate 97, which underwent hydrolysis under the coupling conditions to give allylic alcohol 98. Alcohol 98 reacted with aryl siloxane 72 in the presence of palladium to generate arene ether 96. This mechanism is supported by the isolation of allylic alcohol 98 (18-32%) under various reaction conditions. The highly 57 regioselective formation of ether 96 is in accordance with our proposed model in Scheme 3.4. The regiochemistry of compound 96 was confirmed using 1H-1H COSY experiment. The crystal structure of arene ether 96 further verified regiochemistry, stereochemistry as well as ether functionality.34 Since the coupling of siloxane 72 and carbonate 87 failed under all conditions, it was decided that the coupling of carbonate 87 using arylboronic acid (Suzuki coupling) and aryl stannane (Stille coupling) would be investigated. While traces of coupled product (7% yield) were detected in the Stille coupling, no coupling was seen in the Suzuki reaction. The coupling of carbonate with aryl siloxane in the presence of Pd(dba)2 was unsuccessful under microwave irradiation as well. Scheme 3.9 OCO2Et NHE O O O O Si(OEt)3 NHE O O EtO2CO NHE O O HO TBAF O O O O O NHE Pd(0) hydrolysis Pd(0) 97 98 72 96 87 E = CO2Et Investigation of Problems The failure of allylic carbonate 87 to undergo allylic-arylation reaction can be attributed to the steric bulk of isopropylidene group that blocks the ? face of the alkene, 58 the face on which the palladium metal must reside (Scheme 3.8). We wanted to know if the steric bulk is blocking the formation of ?-allyl intermediate 99 or if it is blocking the subsequent transmetalation step (Scheme 3.10). To test if the ?-allyl intermediate 99 was formed, we decided to couple malonate anion (a soft nucleophile) with allylic carbonate 87. Since the Tsuji-Trost coupling26 of malonate anion with allylic carbonate 87 will occur via the same ?-allyl intermediate 99, the success of this coupling reaction will infer that formation of ?-allyl intermediate 99 is possible. Scheme 3.10 NHE O OPd Ph NHE O OPd Ln pi-allyl intermediate transmetalation Ln O O NHE PdPh Ln 88 O O NHE PdLn 99 Coupling of malonate anion 100 was first attempted with the model allylic carbonate 73 (Scheme 3.11). Unfortunately, Pd(dba)2 and Pd2(dba)3?CHCl3 catalysts which were used for the coupling of carbonate 73 with aryl siloxane 72 were ineffective (see Chapter 2), and showed predominance of starting material 73 even with stoichiometric catalyst loading. However, the Pd(OAc)2/PPh3 system gave diester 101 in 56% yield. The success of this coupling reaction was not surprising. Since we had already demonstrated that coupling of aryl siloxane 72 with allylic carbonate 73 proceeds via ?-allyl intermediate 102 to form carbamates 71/79 (Scheme 3.11), it was expected that 59 coupling of carbonate 73 with malonate anion 100 would proceed through the same ?-allyl intermediate 102 and generate diester 101 with equal feasibility. The observance of single regioisomer (101) in the coupling was consistent with the model developed in our group (Scheme 3.12). After the formation of ?-allyl adduct 102, the malonate anion 100 attacked from the face opposite to the palladium (the same face as the leaving group) to retain the stereochemistry. However, because the carbamate moiety (NHE) is on the same face of attacking malonate anion 100, the nucleophile attacks at C1, further away from the carbamate moiety: resulting in the formation of regioisomer 101 with an overall retention of stereochemistry. Scheme 3.11 NHE OE O O NHE NHE NHE EE O O Si(OEt)3 O NHE O Pd Ln E E + Pd(0) 56%73 (E = CO2Et) 102 100 71 79 101 81% 72 TBAF Having accomplished the coupling of malonate anion 100 with model allylic carbonate 73 (Scheme 3.11), we investigated whether the ?-allyl intermediate 99 (Scheme 3.10) is formed in the coupling of more complex allylic carbonate 87 with malonate system. The coupling of allylic carbonate 87 with malonate anion 100 60 proceeded in a good yield to give regioisomeric diesters 102 and 103 in 4.8:1.0 ratio (Scheme 3.13). The ratio of two regioisomers was determined from the integration of methyl groups of isopropylidene moiety. The major regioisomer was identified as diester 102 using a 1H-1H COSY experiment (see Experimental Section for details of the COSY analysis). Scheme 3.12 NHE EEPd Ln H NHE HH NHE Pd Ln 1 31 3 favored disfavored 102 E E 100 101 NHE OE 1 3 73 (E = CO2Et) Pd(0) Scheme 3.13 OE NHE O O E E NHE E E O O NHE O O E E Pd(OAc)2, PPh3 THF, 65 ?C, 70% 102:103 = 4.8:1.0 87 (E = CO2Et) 102 103100 The regioselectivity of the malonate coupling reaction with allyl carbonate 102 is rationalized in Scheme 3.14. It was anticipated that the reaction of carbonate 87 will result in formation of "unsymmetrical" ?-allyl palladium complex 99 as previously discussed. Since palladium resides more towards C1 due to steric compression with the 61 isopropylidene group, malonate anion 100 was expected to attack at C3 to produce diester 103 as the major regioisomer. Furthermore, since the nucleophile attacks from the bottom face, attack further away from the carbamate moiety (NHE) would be preferred. However, diester 102 was observed as the predominant regioisomer. The formation of diester 102 suggests preferential attack of malonate anion 100 at C1 termini. This is attributed to possible palladium-oxygen coordination with the isopropylidene moiety which pushes palladium away from C1, preferring attack of nucleophile at C1. Scheme 3.14 O O NHE Pd Ln 99 1 3 E E 100 O O NHE PdLn 99 3 NHE E E O O 103 1 Predicted E E 100 Observed 1 NHE O O E E 102 3 Major 1 3 The success of this coupling reaction proves that the ?-allyl palladium intermediate 99 from the allylic carbonate 87 is formed during the Tsuji-Trost coupling. However, when the same reaction is conducted with aryl siloxane 72, the undesired arene ether 96 as well as rearrangement product 98 was obtained (Scheme 3.9). This suggests either transmetalation or subsequent reductive elimination must be the cause for the 62 failure of coupling reaction between aryl siloxane 72 and allylic carbonate 87 (Scheme 3.8). However, since reductive elimination of aryl systems is known to be fast (ary-aryl > alkyl-aryl > alkyl-alkyl)4, it was more likely that transmetalation was the root of the problem. In order to understand the failure of allyl-aryl coupling in the 7-deoxypancratistatin (24) synthesis (Scheme 3.8) mechanistic study on the siloxane reaction was performed. The details of this study are reported in Chapter 4. This study had indicated that the best catalyst system for allyl-aryl cross-coupling reaction would consist of palladium bonded to sterically demanding, but weakly ?-bonding ligands. This set of ligand requirements is not found in most Pd(0) catalysts since Pd(0) is stabilized typically by strong ?-donating ligand systems (phosphines). However, some Pd(0) catalysts have been prepared that have these characteristics (IV112 and V113,114, Figure 3.1) O OO Pd Pd(NBD)(MAH) Pd O O Pd(COD)(NQ) IV V Figure 3.1: Pd(NBD)(MAH) and Pd(COD)(NQ) complexes 63 Successful Coupling Pd(NBD)(MAH) IV and Pd(COD)(NQ) V were employed in the coupling reaction of complex allyl carbonate 87 and aryl siloxanes 72 and 6 (Scheme 3.15). The coupling reaction catalyzed by Pd(NBD)(MAH) IV resulted in lower yield of carbamate 104. However, the reaction in the presence of Pd(COD)(NQ) V worked reasonably well, even at ambient temperatures to give coupled products 49 and 104 in 30-40% yield. As anticipated the coupling reaction exclusively produced regioisomers 49 and 104, respectively. Carbamate 49 is the much desired regioisomer for the synthesis of natural product 7-deoxypancratistatin (24). Scheme 3.15 Si(OEt)3OCO2Et NHE O O NHE O O TBAF, rt or 55 ?C 30-40% 87 Pd(COD)(NQ) O O Si(OEt)3 72 Pd(COD)(NQ) O O NHE O O 49 104 6 Si(OEt)3 NHE O O Pd(NBD)(MAH) 104 6 TBAF, rt or 55 ?C 30-40% TBAF, rt or 55 ?C 11-19% E = CO2Me 64 Scheme 3.16 NHaE O OO O NHaE O O O O NHE PdPh Ln 88 E = CO2Me 49 1 3 1 3 O O Not observed 49' 1 3 Observed Hg HcHb Hd Hb H Figure 3.2: 1H-1H COSY of carbamate 49 (Hudlicky's intermediate) 65 The observed regiochemistry is consistent with the model proposed by DeShong (Scheme 3.16). Due to steric bulk arising from the acetonide moiety, an ?unsymmetrical? palladium complex is formed where palladium resides farther from the acetonide group. The regioselectivity of carbamate 49 was established using 1H-1H COSY (Figure 3.2). Having identified the NHa proton at ? 6.18 using HSQC spectroscopy, it was possible to locate Hb proton adjacent to the NHa. Since Hb proton is not coupled to adjacent olefinic proton, the regioisomer was determined to be 49. If product 49' (Scheme 3.16) had formed, Hb proton would show correlation with the neighboring olefinic proton. The formation of carbamate 49 via palladium-catalyzed allylic-arylation constitutes the formal total synthesis of 7-deoxypancratistatin (24). Carbamate 49 is an intermediate reported in Hudlicky's synthesis of 7-deoxypancratistatin. The spectra of 49 were identical to those reported by Hudlicky.80 Allylic-arylation using Hudlicky's approach (Scheme 3.17) involves coupling of aziridine 48 with cuprate 46 via SN2 process.80,87 The yield of the key reaction is comparable to our palladium-catalyzed allylic-arylation using siloxane methodology. However, the use of extremely low temperature (-78 to -30 ?C) in Hudlicky's procedure, limits its application for large-scale synthesis. On other hand, our allyl-aryl coupling reaction can be performed at ambient temperature (Scheme 3.15), and is suitable for large-scale industrial production. Allylic-arylation approach has also been used by Trost to synthesize (+)-pancratistatin (23) (Scheme 3.18).79 Reaction of allylic carbonate 51 with mixed cuprate 50 provide allylic-arylated system 52 via SN2? reaction. However, Trost's synthesis used complex reaction conditions and was limited to small quantities of 66 material. Moreover, preparation of mixed cuprate is arduous due to its instability, thus making this synthesis unsuitable for practical applications. Scheme 3.17 O O CuCNLi2 2 O O N E O O O O NHE BF3?Et2O, THF -78 to -30 ?C, 34%E = CO 2Me 48 49 46 O O NH O HO OH OH OHA B C 24 Scheme 3.18 OCO2Me N3 O O O O CuCN OMe O O O O N3 OMe MgBr 51 52 50 23 0 ?C O O NH O HO OH OH OH OH A B C According to Hudlicky's procedure, the key steps after the formation of carbamate include installation of trans-diol via allylic-alcohol 52 directed epoxidation and Friedel- Crafts acylation to generate B ring (Scheme 3.19). Following Hudlicky's procedure, carbamate 49 was deprotected to generate allylic alcohol 105 (not isolated). The subsequent step of epoxidation of allyl alcohol 105 was performed in benzene as reported by Hudlicky. However, the poor solubility of diol 105 in benzene led to longer reaction times and in consequence extensive decomposition. Switching to acetonitrile as solvent 67 improved solubility of the diol 105 and NMR analysis of the crude product indicated formation of epoxide 106. However, the yield of 106 has not been optimized. Nonetheless, the formation of carbamate 49 by siloxane methodology constitutes formal total synthesis of (?)-7-deoxypacratistatin (24). In future, this allylic-arylation approach shall be extended to the synthesis of pancratistatin (23) (Scheme 3.20). Scheme 3.19 O O O O NHE 49 E = CO2Me O O OH OH NHE t-BuOOH VO(acac)2 benzene, 80 ?C O O OH OH NHE O AcOH/THF H2O, 65 ?C 105 106 O O NH O HO OH OH OHBA C 24 Scheme 3.20 OCO2Et NHE O O 87 Pd(COD)(NQ) O O Si(OEt)3 72 Pd(COD)(NQ) O O NHE O O 49 O O Si(OEt)3 107 OMe O O NHE O O 108 OMe O O NH O HO OH OH OHBA C 24 O O NH O HO OH OH OHBA C 23 OH 68 CONCLUSION Palladium-catalyzed allylic-arylation involving coupling of allylic carbonate and aryl siloxane has been applied to produce (?)-7-deoxypancratistatin (24) via Hudlicky's intermediate. The key reaction involves the use of a novel Pd(0) olefin complex (Pd(COD)(NQ)) and results in stereospecific arylation to form single regioisomer. This is a first example of application of palladium-catalyzed allylic arylation to the synthesis of 7-deoxypancratistatin (24). Though the yield for the key reaction is moderate, the ability of the coupling to work at ambient temperature is an advantage compared to allylic-arylation methodology by Hudlicky and Trost. Also, due to the ease of preparation of coupling partners, aryl siloxane and allyl carbonate compared to unstable aryl cuprate and aziridine, the siloxane coupling methodology is particularly well-suited for the synthesis of these materials. We believe our synthetic approach is short as well as efficient and thus has potential of getting commercialized. Future goals aim at development of olefin based palladium catalysts to optimize the key reaction with the goal of obtaining multigram quantities of (24) and application of this methodology to the synthesis of pancratistatin (23) derivatives. 69 EXPERIMENTAL DETAILS General Methods All reactions were run under an atmosphere of argon unless otherwise noted. Glassware used in the reactions was dried for a minimum of 12 h in an oven at 120 ?C or flame dried prior to use. Tetrahydrofuran was distilled from sodium/benzophenone ketyl, while methylene chloride and pyridine were distilled from calcium hydride. Thin-layer chromatography (TLC) was performed on 0.25 mm silica gel coated plates treated with a UV-active binder with compounds being identified by one or more of the following methods: UV (254 nm), vanillin/sulfuric acid charring, or KMnO4 charring. Flash chromatography was performed using glass columns and medium pressure silica gel (Sorbent Technologies, 45-70 ?). Infrared spectra were recorded on a Nicolet 560 FT-IR spectrophotometer. Samples used for obtaining infrared spectra were either dissolved in carbon tetrachloride or taken neat. IR band positions are reported in reciprocal centimeters (cm-1) and relative intensities are listed as br (broad), s (strong), m (medium), or w (weak). Nuclear magnetic resonance (1H, 13C NMR) spectra were recorded on a 400 or 500 MHz spectrometer. Chemical shifts are reported in parts per million (?) and coupling (J values) are reported in hertz (Hz). Spin multiplicities are indicated by the following symbols: s (singlet), d (doublet), t (triplet), q (quartet), m (multiplet), br s (broad singlet), br d (broad doublet). Low resolution mass spectrometry (LRMS) and high resolution mass spectrometry (HRMS) were obtained on a JEOL SX-02A instrument. 70 Dibromoalkene 90 Br Br 89 90 -78 ?C, 82% Br2, CHCl3 The dibromoalkene 90 was prepared from 1,4-cyclohexadiene 89 according to Yang's procedure.111 To 15.0 mL (159 mmol, 1.00 equiv.) of 1,4-cylcohexadiene 89 dissolved in 35.0 mL of CHCl3 at -78 ?C was added 8.15 mL (159 mmol, 1.00 equiv.) of Br2 dropwise over 2 h. The reaction mixture was stirred for additional 2 h, and then quenched by the addition of 75.0 mL 0.1 M aqueous sodium thiosulfate (Na2S2O3) solution. The product was extracted with 5 ? 100 mL CH2Cl2 and washed with 100 mL of Na2S2O3 solution. The combined organic layers were dried over MgSO4, filtered and concentrated in vacuo to afford 31.0 g (82%) of dibromoalkene 90 as a white solid, mp 33-37 ?C (lit.111 34-37 ?C) which was used without further purification. IR (CCl4) 3037 (w), 2937 (w), 2882 (w), 2819 (w), 1660 (w), 1421 (s), 1325 (m), 1211 (s), 1187 (m), 1159 (m) cm-1; 1H NMR (400 MHz, CDCl3) ?? 5.65 (s, 2H), 4.50 (s, 2H), 3.17 (d, J = 19 Hz, 2H), 2.59 (d, J = 19 Hz, 2H); 13C NMR (100 MHz, CDCl3) ? 122.3, 48.7, 31.2; LRMS (EI+) m/z 240 (M+ 25), 161 (23), 159 (25), 79 (100), 77 (38). The spectral data was identical to the literature.111 Dibromodiol 91 Br Br Br Br OH OH 90 91 OsO4, NMO acetone, rt, 93% 71 To 16.0 g (66.9 mmol, 1.00 equiv.) of dibromoalkene 90 dissolved in 170 mL of acetone and 25.0 mL of H2O was added 11.8 g (100 mmol, 1.50 equiv.) of NMO under argon. This was followed by addition of 100 mg of OsO4. The reaction was allowed to stir for 3 days at room temperature and was then quenched by the addition of 28.5 g of (150 mmol, 2.24 equiv.) sodium metabisulfite. The product was extracted with 5 ? 200 mL CH2Cl2 and washed with 200 mL H2O. The combined organic extracts were dried over MgSO4, filtered and concentrated in vacuo to give 17.0 g (93%) of dibromodiol 91 as a creamy solid, mp 103-104 ?C (lit.111 103-105 ?C), which was used without further purification. IR (CCl4) 3491-2953 (br s), 2951 (w), 2889 (w), 1685 (w) cm-1; 1H NMR (400 MHz, pyridine - d5) ? 6.11 (br s, 2H), 4.86 (ddd, J = 12, 11, 4 Hz, 1H), 4.40 (ddd, J = 12, 11, 4 Hz, 1H), 4.26 (m, 1H), 3.98 (ddd, J = 12, 4, 3 Hz, 1H), 2.94 (ddd, J = 12, 12, 11 Hz, 1H), 2.84 (ddd, J = 15, 4, 3 Hz, 1H), 2.71 (dddd, J = 12, 4, 4, 1 Hz, 1H), 2.15 (ddd, J = 15, 12, 2 Hz, 1H); 13C NMR (100 MHz, pyridine - d5) ? 71.5, 70.9, 56.5, 55.6, 43.7, 42.1; LRMS (EI+) m/z 274 (M+, 5), 195 (90), 193 (95), 177 (78), 175 (98), 113 (28), 95 (100), 67 (94), 55 (51). The spectral data was identical to the literature.111 Acetonide 92 Br Br OH OH Br Br O O 91 92 p-TsOH, CH2Cl2, rt, 82% 2,2-dimethoxypropane To 17.0 g (62.1 mmol, 1.00 equiv.) of dibromodiol 91 in 400 mL CH2Cl2 was added 87.7 mL (621 mmol, 10.0 equiv.) of 2,2-dimethoxypropane, followed by the addition of 2.10 g (11.0 mmol, 0.177 equiv.) p-TsOH. The reaction was allowed to stir for 64 h. The 72 reaction mixture was extracted with 3 ? 300 mL CH2Cl2 and washed with 300 mL H2O. The combined organic layers were dried over MgSO4, filtered and concentrated in vacuo to obtain 15.9 g (82%) of acetonide 92 as a yellow oil, which was used without further purification. IR (CCl4) 2990 (m), 2932 (m), 2877 (m), 2827 (w), 1720 (w), 1456 (m) cm-1; 1H NMR (400 MHz, CDCl3) ? 4.39-4.44 (dt, J = 4, 8, Hz, 1H), 4.28 (q, J = 5 Hz, 1H), 4.14-4.22 (m, 2H), 2.70-2.77 (m, 2H), 2.32-2.40 (m, 1H), 2.17-2.25 (m, 1H), 1.51 (s, 3H), 1.32 (s, 3H); 13C NMR (100 MHz, CDCl3) ? 109.4, 73.0, 72.6, 51.6, 49.4, 36.5, 35.0, 28.7, 26.5. The spectral data was identical to the literature.111 Diene 93 Br Br O O O O 92 93 DBU, benzene, reflux, 49% To 16.3 g (52.0 mmol, 1.00 equiv.) of acetonide 92 in 58.0 mL benzene was added 18.7 mL (187 mmol, 3.6 equiv.) of DBU dropwise via an addition funnel. The reaction mixture was heated at reflux (80 ?C) for 24 h. The HBr was removed by vacuum filtration and the filtrate was extracted with 3 ? 500 mL pentane and washed with 500 mL H2O. The combined organic extracts were dried over MgSO4, filtered and concentrated in vacuo to afford a yellow oil. Purification by column chromatography (60% CH2Cl2/pentane; Rf = 0.22) rendered 3.87 g (49%) of diene 93 as a yellow oil. IR (CCl4) 3049 (m), 2986 (m), 2928 (m), 2862 (m), 2792 (w), 1961 (w), 1472 (w), 1441 (m), 1383 (m), 1351 (m), 1243 (m), 1216 (m) cm-1; 1H NMR (400 MHz, CDCl3) ? 5.96-5.99 (m, 2H), 5.85-5.89 (m, 2H), 4.63 (s, 2H), 1.40 (s, 3H), 1.38 (s, 3H); 13C NMR (100 MHz, 73 CDCl3) ? 126.7, 123.7, 104.5, 70.2, 27.8, 25.2. The spectral data was identical to the literature.111 Hydroxamate 94 N O O O NBu4+IO4-O O 93 94 E = CO2Me, 31% E H NHO E CHCl3/DMF, 0 ?C to rt E = CO2Et, 57% Hydroxamate (E = CO2Et): To 23.6 g (54.6 mmol, 1.40 equiv.) of NBu4+IO4- (tetrabutylammonium periodate) under argon was added chloroform (80.0 mL) and DMF (26.0 mL). The reaction mixture was cooled to 0 ?C and then 5.93 g (39.0 mmol, 1.00 equiv.) of diene 93 was added. Finally, 4.10 g (39.0 mmol, 1.00 equiv.) of hydroxamic acid (HONHCO2Et) dissolved in chloroform (40.0 mL) and DMF (13.0 mL) was added via addition funnel. The reaction mixture was stirred at 0 ?C to room temperature for 50 h. The product was extracted with Et2O (5 ? 100 mL), washed with H2O (100 mL), dried over MgSO4, concentrated in vacuo to give the crude hydroxamate 94 as a brown oil. Flash column chromatography on silica gel (30% EtOAc/70% hexane, Rf = 0.26) gave 5.72 g (57%) of hydroxamate 94 as a creamy solid, mp 126-128 ?C; IR (CCl4) 2994 (m), 2933 (m), 1717 (s), 1373 (s), 1254 (s), 1210 (s), 1081 (s) cm-1; 1H NMR (400 MHz, CDCl3) ? 6.47-6.38 (m, 2H), 5.06-5.03 (m, 1H), 4.90-4.87 (m, 1H), 4.55-4.49 (m, 2H), 4.25-4.12 (m, 2H), 1.30-1.25 (m, 9H); 13C NMR (100 MHz, CDCl3) ? 158.1, 130.6, 129.5, 111.0, 73.1, 72.6, 71.3, 62.8, 52.9, 25.6, 25.4, 14.4; EI mass spectrum, m/z 74 (relative intensity) 255 (M+, 37), 240 (100), 95 (78), 85 (50), 29 (47); HRMS (EI) calcd for C12H17NO5 (M)+ 255.1107, found 255.1112. Hydroxamate (E = CO2Me): light yellow solid. mp 146-148 ?C; IR (CCl4) 3087 (w), 2999 (m), 2985 (m), 2935 (m), 1719 (s), 1440 (s), 1379 (s), 1336 (s), 1250 (s), 1215 (s) cm-1; 1H NMR (400 MHz, CDCl3) ? 6.45-6.35 (m, 2H), 5.03-5.00 (m, 1H), 4.87-4.84 (m, 1H), 4.53-4.46 (m, 2H), 3.72 (s, 3H), 1.27 (s, 3H), 1.26 (s, 3H); 13C NMR (100 MHz, CDCl3) ? 158.4, 130.6, 129.4, 110.9, 73.0, 72.5, 71.3, 53.6, 52.8, 25.5, 25.3; HRMS (ESI) calcd for C11H16O5N (M+H)+ 242.1029, found 242.1016. Alcohol-Carbamate 95 N O O O Mo(CO)6, MeCN/H2O reflux 94 E = CO2Me, 31% E E = CO2Et, 57% OH NHE O O 95 E = CO2Me, 79% E = CO2Et, 67% Alcohol-Carbamate (E = CO2Et): To 5.72 g (22.4 mmol, 1.00 equiv.) of the hydroxamate 94 dissolved in 240 mL acetonitrile and 12.0 mL distilled water was added 7.10 g (26.9 mmol, 1.20 equiv.) of molybdenum hexacarbonyl (Mo(CO)6). The reaction mixture was refluxed at 82 ?C for 15 h. The black-brown reaction mixture was vacuum filtered through celite and the filtrate was concentrated in vacuo to give crude alcohol-arbamate 95 as a black solid. Flash column chromatography on silica gel (50% EtOAc/50% hexane, Rf = 0.44) gave 3.85 g (67%) of carbamate 95 as a white solid, mp 124-127 ?C; IR (CCl4) 3617 (w), 3451 (w), 2988 (w), 2902 (w), 1735 (s), 1545 (m), 1210 75 (m) cm-1. 1H NMR (400 MHz, CDCl3) ? 5.94-5.90 (m, 1H), 5.81-5.78 (m, 1H), 5.12 (br s, 1H), 4.24-4.18 (m, 3H), 4.14-4.09 (m, 3H), 2.42 (br s, 1H), 1.43 (s, 3H), 1.33 (s, 3H), 1.23 (t, J = 7 Hz, 3H); 13C NMR (100 MHz, d6- DMSO) ? 155.9, 132.6, 129.5, 107.9, 79.8, 75.9, 69.2, 59.7, 51.5, 27.2, 25.0, 14.6; LRMS (FAB) m/z 258 (M+1+, 42), 240 (66), 182 (70), 154 (63), 110 (100); HRMS calcd 258.1342, found 258.1341. Alcohol-Carbamate (E = CO2Me): white solid. mp 86-88 ?C; IR (CCl4) 3617 (m), 3449 (m), 3414 (w), 3049 (w), 2995 (m), 2942 (m), 2909 (m), 1730 (s), 1551 (m) cm-1. 1H NMR (400 MHz, CDCl3) ? 5.91-5.88 (m, 1H), 5.78-5.74 (m, 1H), 5.35 (br d, J = 8Hz, 1H), 4.21-4.18 (m, 3H), 4.08-4.06 (m, 1H), 3.65 (s, 3H), 3.12 (br s, 1H), 1.41 (s, 3H), 1.31 (s, 3H); 13C NMR (100 MHz, CDCl3) ? 156.6, 131.1, 129.7, 109.1, 79.2, 76.8, 68.9, 52.2, 51.1, 26.9, 24.6; HRMS (ESI) calcd for C11H18O5N (M+H)+ 244.1185, found 244.1180. Carbonate-Carbamate 87 OH NHE O O OCO2Et NHE O OClCO2Et, pyridine CH2Cl2, rt 95 87 E = CO2Me, 79% E = CO2Et, 67% E = CO2Me, 86% E = CO2Et, 93% Carbonate-Carbamate (E = CO2Et): To 1.11 g (4.32 mmol, 1.00 equiv.) of alcohol- carbamate 95 in 15 mL anhydrous CH2Cl2 and 0.522 mL (6.48 mmol, 1.50 equiv.) anhydrous pyridine was added 0.642 mL (6.48 mmol, 1.50 equiv.) of ethyl chloroformate dropwise via syringe under argon. The reaction was allowed to stir at room temperature 76 for 5 days. The reaction mixture was extracted with CH2Cl2 (5 ? 25 mL), washed with H2O (25 mL), dried over MgSO4 and concentrated in vacuo. Flash chromatography on silica gel (75% EtOAc/25% hexane, Rf = 0.75 afforded 1.32 g (93%) of the carbonate- carbamate 87 as a white solid, mp 82-84 ?C; IR (CCl4) 3446 (m), 2984 (m), 2933 (m), 2912 (m), 1754 (s), 1730 (s), 1506 (s), 1370 (s), 1254 (s), 1224 (s), 1064 (s), 1040 (s) cm- 1; 1H NMR (400 MHz, CDCl3) ? 5.93-5.86 (m, 2H), 5.11 (br s, 1H), 4.93 (br s, 1H), 4.36- 4.33 (m, 1H), 4.21 (q, J = 7 Hz, 4H), 4.12 (q, J = 7 Hz, 2H), 1.44 (s, 3H), 1.33-1.29 (m, 6H), 1.24 (t, J = 7 Hz, 3H); 13C NMR (100 MHz, CDCl3) ? 156.0, 154.3, 131.5, 127.1, 109.3, 76.1, 75.9, 73.9, 64.4, 61.1, 50.2, 26.9, 24.9, 14.6, 14.2; LRMS (FAB) m/z 330 (M+1+, 6), 154 (38), 136 (40), 73 (48); HRMS (FAB) calcd (M + 1)+ 330.1553, found 330.1553. Carbonate-Carbamate (E = CO2Me): creamy solid. mp 63-65 ?C. IR (CCl4) 3442 (m), 2992 (m), 2956 (m), 2906 (m), 1751 (s), 1733 (s), 1554 (m), 1508 (s) cm-1; 1H NMR (400 MHz, CDCl3) ? 5.86-5.80 (m, 2H), 5.21 (br s, 1H), 5.1 (s, 1H), 4.30-4.27 (m, 1H), 4.19- 4.14 (q, J = 8 Hz 4H), 1.40 (s, 3H), 1.28-1.24 (m, 6H); 13C NMR (100 MHz, CDCl3) ? 156.4, 154.2, 131.3, 127.0, 109.2, 75.9, 75.8, 73.9, 64.3, 52.1, 50.3, 26.8, 24.7, 14.1; HRMS (ESI) calcd for C14H22O7N (M+H)+ 316.1396, found 316.1383. Arene Ether 96 Si(OEt)3O O O O NHE O O O OE NHE O O (allylPdCl)2, PPh3 TBAF, THF, 60 ?C 10% 72 9687 E = CO2Et 77 To a solution of 245 mg (0.900 mmol, 2.00 equiv.) of siloxane 72 in anhydrous THF (10.0 mL) was added 154 mg (0.468 mmol, 1.00 equiv.) of carbonate 87, 8 mg (0.021 mmol, 5 mol %) of (allylPdCl)2, 16.0 mg (0.063 mmol, 15 mol %) of PPh3 and 0.900 mL (0.900 mmol, 1 M in THF, 2.00 equiv.) of TBAF. The reaction mixture was heated to 60 ?C and after 10 min, the color changed from yellow to amber. After 18 h, the reaction mixture was quenched by addition of brine (30 mL), extracted with ether (3 ? 30 mL), dried over MgSO4 and concentrated in vacuo. Flash chromatography on silica gel (10% EtOAc/90% hexane) gave 18 mg (10%) of arene ether 96 as a white solid: recrystallized from CH2Cl2/hexane, mp 126-127 ?C; TLC Rf = 0.27 (25% EtOAc/75% hexane); IR (CCl4) 3444 (m), 2963 (m), 2928 (m), 1724 (s), 1558 (m) cm-1; 1H NMR (400 MHz, CDCl3) ? 6.69 (d, J = 8 Hz, 1H), 6.49 (d, J = 2 Hz, 1H), 6.34 (dd, J = 2, 8 Hz, 1H), 6.08 (dd, J = 4, 10 Hz, 1H), 6.01 (dd, J = 4, 10 Hz, 1H), 5.95 (s, 2H), 5.01 (d, J = 6 Hz, 1H), 4.72-4.70 (m, 1H), 4.66-4.64 (m, 1H), 4.45-4.42 (m, 1H), 4.22-4.19 (m, 1H), 4.10 (q, J = 7 Hz, 2H), 1.48 (s, 3H), 1.38 (s, 3H), 1.27 (t, J = 7 Hz, 3H); 13C NMR (400 MHz, CDCl3) ? 156.9, 154.1, 148.8, 143.0, 129.2, 128.7, 108.6, 108.4, 101.8, 100.3, 76.1, 75.3, 73.0, 71.7, 52.7, 28.4, 26.2, 14.5. FAB mass spectrum m/z (relative intensity) 378, 320, 240, 180, 119, 85. X-ray crystal confirmed regiochemistry, stereochemistry as well as ether functionality.34 Diester 101 Pd(OAc)2, PPh3, NHCO2Et OCO2Et NHCO2Et CO2EtEtO2C 101 THF, 65 ?C, 56% 73 EtO2C CO2Et 100 78 Sodium hydride (33 mg of 60% dispersion in oil, 0.83 mmol, 3.0 equiv.) was washed with 3 ? 2 mL of hexane and 2 ? 3 mL of anhydrous THF. To a suspension of the oil-free sodium hydride in 2 mL THF was added 0.13 mL (0.88 mmol, 3.2 equiv.) of diethyl malonate and stirred for 10 minutes. The diethyl malonate anion 100 was then added to a solution of 71 mg (0.28 mmol, 1.0 equiv.) of allyl carbonate 73 dissolved in 1 mL anhydrous THF. This was followed by addition of 36 mg (0.14 mmol, 0.50 equiv.) triphenyl phosphine and 6.2 mg (0.028 mmol, 0.10 equiv.) of palladium acetate. The reaction mixture was allowed to stir at 65 ?C for 22 h. The solution was filtered through Celite and the filtrate was extracted with 5 ? 25 mL Et2O and washed with 25 mL H2O. The combined organic layers were dried over MgSO4, concentrated in vacuo to give crude diester 101 as a brown oil. Flash column chromatography on silica gel (20% EtOAc/80% hexane, Rf = 0.15) gave 50 mg (56%) of diester 101 as a colorless oil; IR (CCl4) 3450 (w), 2987 (m), 2936 (m), 2872 (w), 1730 (s) 1502 (m), 1214 (m) cm-1; 1H NMR (400 MHz, CDCl3) ? 5.71 (s, 2H), 4.74 (d, J = 8 Hz, 1H), 4.21-4.11 (m, 5H), 4.06 (q, J = 7 Hz, 2H), 3.26 (d, J = 8 Hz, 1H), 2.83-2.79 (m, 1H), 1.74-1.67 (m, 3H), 1.50-1.44 (m, 1H), 1.25-1.17 (m, 9H); 13C NMR (100 MHz, CDCl3) ? 168.2, 168.1, 155.7, 131.5, 128.8, 61.4, 60.7, 56.1, 44.7, 35.0, 27.7, 22.2, 14.6, 14.0; FAB mass spectrum m/z (relative intensity) 328 ((M+H), 9), 239 (28), 161 (100), 79 (38); HRMS (FAB) calcd for C16H26O6N (M+H)+ 328.1760, found 328.1773. The regiochemistry of carbamate 101 was established using 1H-1H COSY (400 MHz, CDCl3) (see page 79). NHa proton identified using HSQC spectroscopy is coupled only to Hb proton. Since, Hb proton is coupled to vinyl proton; the regioisomer was determined to be 101. (If 101' had formed, Hb would not couple to vinyl proton.) 79 NHaCO2Et CO2EtEtO2C 101 Hb NHaCO2Et 101' Hb EtO2C EtO2C Only Observed Not Observed COSY of Diester 101 80 Diesters 102 and 103 OE NHE O O E E NHE E E O O NHE O O E E Pd(OAc)2, PPh3 THF, 65 ?C, 70% 102:103 = 4.8:1.0 87 102 103100 E = CO2Et Sodium hydride (37.9 mg of 60% dispersion in oil, 0.948 mmol, 3.00 equiv.) was washed with 3 ? 3 mL of hexane and 3 ? 3 mL of anhydrous THF. To a suspension of the oil-free sodium hydride in 2 mL THF was added 0.152 mL (1.01 mmol, 3.20 equiv.) of diethyl malonate and stirred for 10 minutes. The diethyl malonate anion 100 was then added to a solution of 104 mg (0.316 mmol, 1.00 equiv.) of allyl carbonate 87 dissolved in 2 mL anhydrous THF. This was followed by addition of 41.4 mg (0.158 mmol, 0.500 equiv.) triphenyl phosphine and 7.09 mg (0.0316 mmol, 0.100 equiv.) of palladium acetate. The reaction mixture was allowed to stir at 65 ?C for 24 h. The solution was filtered through celite and the filtrate was extracted with 5 ? 25 mL Et2O and washed with 25 mL H2O. The combined organic layers were dried over MgSO4, concentrated in vacuo to give crude as brown oil. Flash column chromatography on silica gel (20% EtOAc/80% hexane, Rf = 0.29) gave 88 mg (70%) of diesters 102 and 103 (102:103 = 4.8:1.0) as light yellow oil; The ratio of regioisomers was determined from the 1H NMR. The major regioisomer was established to be 102 based on COSY (400 MHz, CDCl3). Diesters 102 and 103 : IR (CCl4) 3439 (w), 2984 (m), 2936 (w), 2913 (w), 1737 (s) 1506 (s), 1373 (m), 1220 (s) cm-1; Diester 102: 1H NMR (400 MHz, CDCl3) 5.82-5.72 (m, 2H), 4.77 (br s, 1H), 4.52 (s, 1H), 4.28 (s, 1H), 4.24-4.13 (m, 5H), 4.06 (q, J = 7 Hz, 2H), 3.45 (s, 2H), 81 1.39 (s, 3H), 1.34 (s, 3H), 1.27-1.19 (m, 9H); Diesters 102 and 103: FAB mass spectrum m/z (relative intensity) 532 ((M+Cs)+, 75), 342 (32), 179 (22), 133 (100); HRMS (FAB) calcd for (M+Li)+ 406.2053, found 406.2056. 1H NMR spectrum of diesters 102 and 103 indicated predominantly one regioisomer. The ratio of two regioisomers was determined from the integration of methyl groups of isopropylidene moiety. Diester 102 was established as the major regioisomer by COSY (400 MHz, CDCl3). Ha proton (NHa proton identified using HSQC spectroscopy) is coupled only to Hb proton. However, Hb proton is not coupled to vinyl proton (which is the case only for regioisomer 102). Therefore, the predominant regioisomer was determined to be diester 102. COSY of diesters 102 and 103 82 Carbamate 104 Si(OEt)3 NHE O OOCO2Et NHE O O TBAF, THF, 55 ?C 30-40% 87 Pd(COD)(NQ) 1046 Carbamate (E = CO2Me): To 97.0 mg (0.404 mmol, 2.00 equiv.) of aryl siloxane 6 and 66.5 mg (0.202 mmol, 1.00 equiv.) of carbonate-carbamate 87 dissolved in 4.00 mL anhydrous THF was added 0.404 mL (0.404 mmol, 2.00 equiv.) TBAF under argon. This was followed by addition of 37.6 mg (0.101 mmol, 0.500 equiv.) Pd(COD)(NQ). The reaction mixture was heated at 55 ?C for 24 h. The reaction was then quenched by addition of 25.0 mL H2O. The product was extracted with 3 ? 25 ml CH2Cl2 and washed with H2O. The combined organic extracts were dried over MgSO4, filtered and concentrated in vacuo to give a brown oil. Flash column chromatography on silica gel (20% EtOAc/80% hexane, Rf = 0.15) gave 24 mg (38%) of carbamate 104 as pale yellow solid, mp 148-150 ?C; IR (CCl4) 3469 (w), 3448 (w), 2987 (w), 1730 (s), 1507 (s), 1246(s), 1221 (s) cm-1; 1H NMR (400 MHz, (CD3)2CO) ? 7.30-7.27 (m, 2H), 7.22-7.18 (m, 3H), 6.22 (br d, J = 8 Hz, 1H), 5.99-5.95 (m, 1H), 5.87 (d, J = 10 Hz, 1H), 4.68 (t, J = 5 Hz, 1H), 4.26 (m, 1H), 3.69 (q, J = 10 Hz, 1H), 3.51 (br d, J = 10 Hz, 1H), 3.38 (s, 3H), 1.45 (s, 3H), 1.33 (s, 3H); 13C NMR (100 MHz, (CD3)2CO) ? 157.3, 142.7, 136.8, 129.3, 129.1, 127.5, 124.8, 109.8, 77.9, 73.3, 56.7, 51.6, 47.4, 28.7, 26.5; HRMS (ESI) calcd for C17H22O4N (M+H)+ 304.1549, found 304.1563. Carbamate (E = CO2Et): light yellow solid. mp 109-112 ?C. IR (CCl4) 3469 (w), 3452 (w), 3032 (w), 2987 (m), 2929 (w), 2876 (w), 1726 (s), 1548 (s), 1511 (s), 1381 (m), 83 1246 (s), 1221(s) cm-1; 1H NMR (400 MHz, (CD3)2CO) ? 7.30-7.26 (m, 2H), 7.22-7.18 (m, 3H), 6.23 (br d, J = 8 Hz, 1H), 5.99-5.96 (m, 1H), 5.87 (d, J = 12 Hz, 1H), 4.68 (t, J = 4 Hz, 1H), 4.28 (dd, J = 4 Hz, J = 8 Hz, 1H), 3.81 (q, J = 4 Hz, 2H), 3.69 (dd, J = 8 Hz, J = 12 Hz, 1H), 3.53 (br d, J = 8 Hz, 1H), 1.45 (s, 3H), 1.33 (s, 3H), 1.01 (t, J = 8 Hz, 3H); 13C NMR (100 MHz, (CD3)2CO) ? 156.9, 142.8, 136.8, 129.3, 129.1, 127.5, 124.9, 109.8, 77.9, 73.4, 60.3, 56.7, 47.5, 28.7, 26.5, 15.0; HRMS (ESI) calcd for C18H24O4N (M+H)+ 318.1705, found 318.1674. The regiochemistry of carbamate 104 was established using 1H-1H COSY (400 MHz, (CD3)2CO) in a manner similar to carbamate 49 (page 64, Figure 3.2) COSY of carbamate 104 84 Carbamate 49 (Hudlicky's Intermediate) Si(OEt)3O O NHE O OOCO2Et NHE O O TBAF, THF, 55 ?C 30-40% 72 87 O O Pd(COD)(NQ) 49 Carbamate (E = CO2Me): To 462 mg (1.63 mmol, 2.00 equiv.) of aryl siloxane 72 and 256 mg (0.813 mmol, 1.00 equiv.) of carbonate-carbamate 87 dissolved in 15.0 mL anhydrous THF was added 1.63 mL (1.63 mmol, 2.00 equiv.) TBAF under argon. This was followed by addition of 152 mg (0.407 mmol, 0.500 equiv.) Pd(COD)(NQ). The reaction mixture was heated at 55 ?C for 24 h. The reaction was then quenched by addition of 30.0 mL H2O. The product was extracted with 3 ? 50 ml CH2Cl2 and washed with H2O. The combined organic extracts were dried over MgSO4, filtered and concentrated in vacuo to give brown oil. Flash column chromatography on silica gel (30% EtOAc/70% hexane, Rf = 0.17) gave 98 mg (35%) of carbamate 49 as pale yellow solid, mp 177-179 ?C ((lit.80 190-191 ?C); IR (CCl4) 3467 (w), 3442 (w), 3042 (w), 2995 (w), 2881 (w), 1730 (s), 1504 (s), 1486 (s), 1250 (s) cm-1; 1H NMR (500 MHz, (CD3)2CO) ? 6.75 (d, J = 8 Hz, 1H), 6.68-6.65 (m, 2H), 6.18 (br d, J = 8 Hz, 1H), 5.97- 5.93 (m, 3H), 5.85 (d, J = 10 Hz, 1H), 4.66 (t, J = 5 Hz, 1H), 4.25-4.22 (m, 1H), 3.62 (q, J = 10 Hz, 1H), 3.46 (br d, J = 10 Hz, 1H), 3.42 (s, 3H), 1.45 (s, 3H), 1.33 (s, 3H); 13C NMR (125 MHz, (CD3)2CO) ? 157.4, 148.6, 147.4, 136.9, 136.6, 124.8, 122.5, 109.9, 109.4, 108.8, 101.9, 77.9, 73.3, 56.9, 51.6, 47.1, 28.7, 26.6; HRMS (ESI) calcd for C18H22O6N (M+H)+ 348.1447, found 348.1440. 1H and 13C NMR spectra in CDCl3 are 85 identical with that reported by Hudlicky.80 The regiochemistry of carbamate 49 was established using 1H-1H COSY (500 MHz, (CD3)2CO), provided on page 64, Figure 3.2. Carbamate (E = CO2Et): light yellow solid. mp 151-153 ?C; IR (CCl4) 3467 (w), 3446 (w), 3042 (w), 2995 (w), 2935 (w), 2874 (w), 1726 (s), 1547 (s), 1511 (s), 1486 (s), 1250 (s) cm-1; 1H NMR (500 MHz, (CD3)2CO) ? 6.75 (d, J = 8 Hz, 1H), 6.68-6.65 (m, 2H), 6.11 (br d, J = 8 Hz, 1H), 5.97-5.94 (m, 3H), 5.86 (d, J = 10 Hz, 1H), 4.66 (t, J = 5 Hz, 1H), 4.25-4.22 (m, 1H), 3.86 (q, J = 5 Hz, 2H), 3.62 (q, J = 10 Hz, 1H), 3.46 (br d, J = 10 Hz, 1H), 1.45 (s, 3H), 1.33 (s, 3H), 1.05 (t, J = 5 Hz, 3H); 13C NMR (125 MHz, (CD3)2CO) ? 157.0, 148.7, 147.4, 136.8, 136.7, 124.9, 122.6, 109.9, 109.5, 108.8, 102.0, 77.9, 73.4, 60.5, 57.0, 47.3, 28.8, 26.6, 15.0; HRMS (ESI) calcd for C19H24O6N (M+H)+ 362.1604, found 362.1563. 86 CHAPTER 4 MECHANISTIC STUDIES ON PALLADIUM-CATALYZED ALLYLIC-ARYLATION INTRODUCTION The DeShong group has developed palladium-catalyzed arylation of allylic esters via coupling of aryl siloxanes (see Chapter 1).29-34 This methodology has been shown to be both highly regio- and stereoselective and has been successfully applied to the synthesis of a (?)-7-deoxypancratistatin analogue 69 (Scheme 4.1) (see Chapter 2). However, repeated attempts to effect coupling between carbonate 87 and siloxane 72 for the synthesis of 7-deoxypancratistatin (24) were unsuccessful and no trace of the desired adduct 49 was detected. Scheme 4.1 OE NHE OE NHE O O O O Si(OEt)3 NHE O O NHE O O O O O NH HO O O OH OH OH O O NH O HO OH Pd(0), TBAF 24 69 72 73 71 49 Pd(0), TBAF 7-deoxypancratistatin analogue 7-deoxypancratistatin 87 87 It was hypothesized initially that failure of allyl carbonate 87 to undergo arylation was the result of the steric bulk of the isopropylidene group that blocked the ?-face of the alkene and prevented formation of the requisite ?-allyl intermediate 99 (Scheme 4.2). However, the successful coupling of allylic carbonate 87 with malonate anion 100 demonstrated that formation of ?-allyl intermediate 99 had been produced since it underwent Tsuji-Trost coupling. This result led us to infer that either transmetalation or the subsequent reductive elimination step, latter steps in the mechanism, must be responsible for failure of the coupling reaction between aryl siloxane 72 and allyl carbonate 87 (Scheme 4.3). Since both transmetalation and reductive elimination involve transfer of an aryl group during the catalytic cycle, we anticipated that altering substituents on the aryl ring would affect the relative rates of these processes. Accordingly, we chose to investigate the role of substituents on aryl siloxane in controlling the rate of the coupling reaction via Hammett analysis. Scheme 4.2 OO O NHE O O O NHE OE E E NHE E E O O NHE O O E E O O Si(OEt)3 F 70% 100 109 102 103 102:103 = 4.8:1.0 Pd(0) 87 (E = CO2Et) 49 O O NHE PdLn 99 88 Scheme 4.3 O O NHE O O NHE OO O NHE O Si OEt F OEt OEt O O O O NHE OE Pd Ln Pd LnO O Pd(0) transmetalation reductive elimination substitution/pi ? allyl formation O O NHE PdLn 87 9988 49 109 Transmetalation Transmetalation of organometallic compounds with transition metal complexes is one of the key steps in carbon-carbon bond formation. However, mechanistic details of the transmetalation process are not well understood for most of these catalytic processes. Among many useful coupling variants, the Stille reaction is the most extensively studied system with regard to the mechanism of transmetalation. One of the earliest studies is that of palladium-catalyzed coupling of benzoyl chloride with benzyltin reagents by Stille.115,116 Subsequently, Farina reported a kinetic analysis of the effect of palladium ligands on the Stille reaction117 and Hartwig investigated the mechanism of transmetalation of tin amides and tin thiolates.118 More recently, Amatore and Jutand have studied the mechanism of Stille reaction in the presence of AsPh3 ligated palladium 89 catalyst, and confirmed Farina's proposal that AsPh3 increased the efficiency of the Stille reaction compared to PPh3.119 An extensive kinetic study of the transmetalation reaction in Stille coupling has been carried out recently by Espinet, who has proposed an associative transmetalation model for the key step and also investigated the nature of the transition state (Figure 4.1).120-123 Studies on internal-coordination driven transmetalation are also known in literature.124-126 Transmetalation studies on palladium-catalyzed cross couplings of alkynyl stannanes with aryl iodides and that with metal-halides have been reported by Crociani127 and Lo Sterzo128, respectively. Additionally, Wendt129 and Clarke130 have reported studies on transmetalation of organostannanes and organozincs, respectively, with platinum complexes. More recently, theoretical calculations on transmetalation of Stille reaction have appeared.131-135 X Sn R2 PdLn R1 Sn R 2 ?+ PdLn X R1? + ?- SE2 (cyclic) SE2 (open) Figure 4.1: Espinet's model for transmetalation in Stille reaction Transmetalation in the Suzuki-Miyaura and Hiyama protocols have been studied less extensively. Transmetalation processes for cross-coupling of organoboron compounds in alkaline solution have been studied by Miyaura.136 Three possible pathways for transmetalation process have been proposed (Scheme 4.4). In Path A, the addition of sodium hydroxide generates tetravalent boronate anion 110 which enhances nucleophilicity of organic group and thus accelerates transmetalation. Alternatively, ligand exchange between R-Pd-X and a base R?O generates oxo palladium(II) complex 90 111 in situ (Path B). The high bascity of the Pd-OR species as well as the high oxophilicity of boron results in enhanced reactivity of the oxo palladium complex. On the other hand, reaction of allyl acetate can proceed under neutral conditions since oxidative addition directly yields ?-allylpalladium acetate complex 111 (Path C). These three pathways are highly dependent on the combination of bases and organoboron reagents, as well as organic electrophiles. Scheme 4.4 R-Pd-X Path A B OH OH OHR' R-Pd-R' Path B R''O R-Pd-OR'' 110 111 Path C R-OR'' + Pd(0) R'-B(OH)2 (HO)2B R' R''O Pd R R'L2Pd O B H SE2 (cyclic) 112 Figure 4.2: Transmetalation of alkyl boranes Woerpel and Soderquist studied transmetalation of primary alkyl borane derivatives to palladium and proposed a hydroxo bridged SE2 (cyclic) transition state 112 (Figure 4.2 ).137,138 More recently, theoretical studies on mechanism of Suzuki-Miyaura 91 reaction catalyzed by diphosphine palladium complexes have been reported. According to these studies arylboronic acid is activated by an external base, which attacks the palladium center as an boronate anion.139-144 The use of organosilanes rather than boranes as cross-coupling partners has been less studied and mechanistic knowledge for Hiyma coupling is scarce. Because the Si-C bond is less polarized than the corresponding B-C bond, Hiyama introduced use of a nucleophilic fluoride source to polarize Si-C bond via formation of a reactive pentacoordinate silicate and enhance transmetalation.145-148 Hiyama showed that the stereochemistry of transmetalation can be influenced by the reaction temperature and the solvent used.145,147 For example, the cross-coupling reaction of aryl triflates 113 with chiral alkylsilanes 114 (Scheme 4.5) in THF at low temperatures proceeded with retention of configuration, whereas reaction at higher temperatures or in polar solvents (HMPA) resulted in inversion of configuration. Retention of configuration is attributed to fluorine-bridged SE2 (cyclic) transition state 115 in the transmetalation (Scheme 4.6). At higher temperature or in polar solvents, a fluorine-silicon bridge is cleaved resulting in SE2 (open) transition state 116 and thus inversion of configuration. Stereochemistry in cross-coupling of allyltrifluorosilanes with aryl triflates was shown also to be influenced by temperature and fluoride source.148 Scheme 4.5 X F3Si Me X = H, OMe OSO2CF3 Y Y = H, 3-CHO, 4-COMe Pd(PPh3)4 TBAF Y X Me 50-100 ?C 113 114 92 Scheme 4.6 F4Si Ph HMe Pd(Ar)LnF F4Si Ph HMe Pd(Ar)(F)Ln Ln(Ar)Pd Ph HMe Ph Pd(Ar)Ln HMe Ar Ph HMe Ph Ar HMe SE2(cyclic) SE2(open) Inv retention inversion 115 116 Recently, Sakaki reported a theoretical study of the transmetalation between palldium(II)-vinyl complex and vinyl silane.149 The study indicated a very large activation barrier for transmetalation process in the absence of fluoride anion. In the presence of fluoride anion, transmetalation is accelerated by the formation of a very strong Si-F bond and the stabilization of the transition state by hypervalent Si center induced by the fluoride anion. Denmark has also performed mechanistic studies on organosilanes via kinetic analysis under both fluoride-mediated and fluoride-free conditions.150,151 In an effort to study transmetalation process, stable transmetalation intermediates in Stille, Suzuki and Hiyama cross-coupling have also been isolated.152-155 93 Hammett Analysis Hiyama Coupling Mechanistic studies of transmetalation of an aryl group in metal-catalyzed couplings of organosilanes using Hammett analysis have been reported previously (Scheme 4.7). For example, Hatanaka and Hiyama have shown that electron-donating groups (EDG) enhance the rate of transmetalation of diarylfluorosilicates with an aryl- palladium complex (Figure 4.3).156 The negative slope is indicative of the electrophilic character of the transmetalation. The presence of EDG on diarylfluorosilicate 117 increases the nucleophilicity of the aryl-silicon bonds, which aids in electrophilic attack of arylpalladium(II) complex 118 via a Si-Pd binulclear intermediate 119 formed by a fluoride bridge (Scheme 4.8). Scheme 4.7 SiF2R I R (?3-C3H5PdCl)2 R = OMe, Me, H, F, CF3 Scheme 4.8 Ar Si Ar Si F F Ar' PdIIFLn PdLAr' F SiAr F Ar' PdIILn Ar Ar'-Ar PdoLn 117 118 119 94 Figure 4.3: Hammett analysis of the reaction of diaryl(difluoro)silanes with iodobenzene. Taken from ref 156. Suzuki Coupling The reactivity of various arylboronic acid for the coupling of (E)-bromostilbene in the presence of Pd(OAc)2/PPh3 was evaluated by competitive experiments (Scheme 4.9).157 Arylbornic acids containing an electron-donating group (EDG) in the para position were found to be more reactive (? = -0.71) (Figure 4.4). EDG increases nucleophilicity of the aryl group, promoting transfer of aryl group to the electron deficient palladium. This result is analogous to the results from the Pd(OAc)2/PPh3 catalyzed cross-coupling reaction of arylboronic acid with vinyl bromide, generated in situ from 1,2-dibromoethane (? = -1.26).158 On the other hand, nickel-catalyzed Suzuki coupling of an arylboronic acid and an aryl tosylate gave the opposite result: electron- withdrawing groups (EWG) on the aryl boronic acid facilitated transmetalation and gave a slope of ? = 0.81.159 95 Scheme 4.9 B(OH)2 B(OH)2 R R = OMe, Me, Cl, CF3 Br Pd(OAc)2/PPh3, KOH R Figure 4.4: Hammett analysis of the reaction of arylboronic acid with E-bromostilbene. Taken from ref 157. Additionally, Hammett analysis has also been employed to study the mechanism of the reaction of aryl bromides and arylboronic acid catalyzed by palladacycles (Figure 4.5). It was found that aryl bromides bearing EWG accelerated the rate of reaction and gave correlation values of ? = 0.48, ? = 0.66, and ? = 0.99 for palladacycles VI160, VII160 and VIII161 respectively. Transmetalation processes have also been studied for the cross- coupling of phenyl boronates with propargylic carbonates (? = 0.73)162, ?-selective cross- coupling of potassium allyltrifluoroborates with aryl bromides (? = -1.1)163, and 1,4- addition of arylboronic acids to enones (? = -0.54).164 96 PPh 2 Pd N Cl i-Pr 2 PPh2 Pd N Cl PCy3 i-Pr VI VII Pd N VIII F3OCO 2 Pr-i Pr-i Figure 4.5: Palladacycles used in Hammett analysis of arylboronic acid with aryl bromides Stille Coupling Farina studied the electronic influence of aryl stannane on the transmetalation step by competitive experiment of vinyl triflate with various para-substituted aryl stannanes (Scheme 4.10).165 In the absence of lithium chloride, EDG on the aryl stannane accelerated transmetalation (? = -0.89), indicating development of positive charge in the transition state (Figure 4.6). However, in the presence of LiCl, the linear relationship could not be obtained (Figure 4.7). This indicated two mechanistically different pathways for the transmetalation process in the presence of salt. Farina's results are in contrast to that by Stille employing acid chlorides and benzylic stannanes (? = 1.2).116 Scheme 4.10 SnBu3 SnBu3 R R = NMe2, OMe, Cl, CF3 Pd2dba3, AsPh3 (LiCl) t-Bu R OTs t-Bu t-Bu 97 Figure 4.6: Hammett analysis of Stille coupling reaction in absence of LiCl. Taken from ref 165. Figure 4.7: Hammett analysis of Stille coupling reaction in presence of LiCl. Taken from ref 165. In summary (Table 4.1), Hammett studies of cross-coupling reactions have indicated that transmetalation is a complex process in which the rates of coupling are strongly influenced by substituents on the ring as well as the catalyst and/or ligand. In light of the failure of our aryl-allyl coupling reaction in the 7-deoxypancratistatin (24) synthesis (Scheme 4.1), we chose to investigate the mechanism of the siloxane-based coupling reaction in detail. The goal of this project was to perform a Hammett study of the coupling reaction utilizing palladium-catalyzed siloxane derivatives. The study reported below is the first mechanistic investigation of an allyl-aryl coupling process involving silicon-based reagents. 98 Organometallic Partner Electrophilic Partner Catalyst ? SiF2R I (?3-C3H5PdCl)2 B(OH)2R Br Pd(OAc)2/PPh3 SnBu3R Pd2dba3, AsPh3t-BuTsO TsOB(OH)2R -1.5 -0.71 0.81 -0.89 1. 2. 3. 4. NiCl2(PCy3)2 Table 4.1: Summary of Hammett studies RESULTS AND DISCUSSION Hammett Analysis The proposed mechanism for the allyl-aryl coupling reaction is summarized in Scheme 4.11. The relative rates for each individual step of the coupling will be discussed below. Cyclohexenyl carbonate (120) was chosen as the coupling partner for the siloxane study because it was known to undergo facile allyl-aryl coupling with siloxane derivatives under established protocols. There is significant literature precedent for the rapid and reversible formation of ?-allyl palladium complexes from allylic derivatives with both phosphine and dibenzylideneacetone (dba) ligands.166,167 (The reversibility of 99 the reaction is diminished with carbonates since the carbonate anion decomposes to carbon dioxide and alkoxide under the typical coupling conditions).166 Scheme 4.11 Pd LnOCO2Et Pd R Ln Si(OEt)3 R F Pd Si(OEt)3F R Ln Si(OEt)3 R TBAF Pd(0) transmetalation ?+ ??? ? 120 123 124 125 126 122 fast fast R121 However, it was important to conclusively demonstrate that this step was not rate- determining in this coupling system, and that result can be inferred from the subsequent Hammett analysis reported below. If formation of the ?-allyl intermediate 123 were rate-determining, then a correlation of rates with Hammett parameters on the siloxane moiety would not be observed because the siloxane would not appear in the rate equation for formation of the ?-allyl complex. Accordingly, the Hammett correlation observed (vide infra) is consistent with the formation of the ?-allyl palladium complex being a fast process. Formation of the hypercoordinate silicate 122 from the reaction of fluoride anion (TBAF) with the siloxane derivative 121 is not rate-determining either. This was demonstrated by 19F NMR spectroscopy of the reaction of TBAF and phenylsiloxane derivatives 121. In the key experiment, a hypercoordinate silicate species (121, R = H) 100 was generated in situ by reaction with fluoride source (TBAF) (Scheme 4.12 and Figure 4.8). The formation of hypercoordinate silicates was observed using 19F NMR where the fluorine signal of TBAF (? -114)168 disappeared rapidly (10 min) at room temperature on addition of phenylsiloxane to give two new fluorine signals: a sharp singlet at ? -121, and a broad resonance centered at ca. ? -127, respectively, as shown in Figure 4.8. The signals at ? -121 and ? -127, respectively, are consistent with chemical shifts of hypercoordinate fluorosilicate species such as 122 and 127, respectively reported by this and other groups.169-177 Upon cooling (-29 ?C), the 19F signal at ? -121 sharpened, showing silicon satellites (JSi-F = 207 Hz). Additonally, 29Si NMR also indicated coupling of 207 Hz at ? -129 confirming the formation of hypervalent silicate species (see Experimental Section, Figure 4.15). Scheme 4.12 Si(OEt)3 R FSi(OEt)3 R TBAF Si(OEt)2 R F OEt Si(OEt)3 R F 122 F 2 127 101 Figure 4.8: 19F NMR spectra of silicate formation (a) TBAF in THF at 29 ?C (b) 19F NMR spectrum of silicate complexes resulting from 1:1 mixture of TBAF and Triethoxyphenylsilane at 29 ?C. Insert is 19F signal at ? -121 after cooling to -28 ?C. The broad signals in the 19F NMR spectrum are consistent with formation of hypercoordinate complex(es) that undergo dynamic processes including ligand exchange and pseudorotation. The equilibrium between the various silicates is dependent on stoichiometry of fluoride:siloxane and other electronic factors. The 19F NMR spectra are also temperature dependent, indicating a dynamic process (see Experimental Section). Preliminary studies of silicate formation indicated that the electronic effects of the groups attached to the aryl ring had an effect on the relative quantities of each component, but additional studies have to be undertaken to determine the relative importance of each silicate to the overall coupling reaction. Nonetheless, the conclusion drawn from this JSi-F = 207 Hz -28 ?C a) TBAF at 29 ?C b) TBAF and Triethoxyphenylsilane (1:1) at 29 ?C 102 study is that at room temperature, the siloxane reacted with fluoride ion to provide hypercoordinate silicates rapidly, much more rapidly than coupling occurred. Analogously, when TBAF (2 equiv.) was added to a mixture of phenylsiloxane (1 equiv.) and its p-methoxy congener (1 equiv.), the signal for TBAF rapidly disappeared and was replaced by a series of resonances indicative of hypercoordinate silicate formation. This experiment conclusively demonstrated that formation of the silicate from the reaction of fluoride ion and the siloxane derivatives was fast and could not be the rate-determining step in the coupling reaction. We had proposed initially that either transmetalation of silicate 122 to ?-allyl palladium complex 123 or subsequent reductive elimination was the rate-determining step in the coupling reaction (Scheme 4.11). If this assumption were correct, then the rate of the allyl-aryl coupling should be influenced by the electronic characteristics of the substituent present on the aryl ring of siloxane 121, and substituents in the para-position would stabilize (or destabilize) transition state 124 resulting in a rate enhancement (or diminution). Competition experiments between phenyltriethoxysilane 6 and para- substituted aryl siloxanes 121 (R = OMe, Me, Cl, CO2Et) were performed and the results are summarized in Figure 4.9. The relative rate of transmetalation was enhanced by electron-withdrawing groups (EWG). Electron-withdrawing groups are better at stabilizing the developing negative charge on the ipso-carbon in transition state 124 through inductive effects. 103 OCO2Et Si(OEt)3 H HSi(OEt) 3 R R Si(OEt)3 OMe Si(OEt)3 Me Si(OEt)3 H Si(OEt)3 Cl Si(OEt)3 CO2Et Pd(dba)2, 55 ?C THF, TBAF 0.42 0.72 1 3.4 4.2 Relative rate of coupling reaction 6 121 9 126 120 Figure 4.9: Summary of relative rates of coupling reactions with siloxane derivatives More importantly, the excellent correlation observed in the Hammett analysis is consistent with the proposal that the rate-determining step is either the transmetalation or reductive elimination reaction. As was noted above, if ?-allyl formation were the slow step, then no difference in the relative rates would have been observed since the substituents on the aryl ring would not be able to manifest their influence in the rate determining step of the coupling reaction. Having established the nature of the substituent effect for the siloxane coupling protocol, the magnitude of the stabilization that occurred in the transition state for the coupling was determined from the Hammett correlation. Hammett plots were obtained by plotting log (k/k0) against substituent parameters ?p, ?-, and ?+.178 The plot of log (k/k0) 104 against ?p gave the best linear correlation to the experimental data (Figure 4.10) with slope ? = 1.4. The linear regression with ?p value indicated that an inductive effect was chiefly responsible for stabilization of the transition state. The positive slope (? = 1.4) of the line indicated that coupling was sensitive to the electronic effects of substituents and that a significant amoung of negative charge was to be found on the aromatic ring in the transition state 124 (Scheme 4.11). Contrast the magnitude of the ? value (+1.4) with the values obtained in studies of borates and stannanes, respectively (described above) where |?| = 0.7-1.1 (Table 4.1). The positive slope of ? = 1.4 is in sharp contrast with the studies reported by Hiyama (? = -1.5)156, Monteiro (? = -0.71)157 and Farina (? = -0.89)165 using silicon, boron and stannane derivatives, respectively, in the presence of palladium catalyst, where EWG retarded the rates of coupling reaction (Table 4.1). COOEtCl H OMe Me ? = 1.4 R2 = 0.95 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 ?p lo g( k/k o) Figure 4.10: Hammett analysis of allyl-aryl coupling reaction It is noteworthy to reiterate the significance of the Hammett correlation: the observed substituent effects are consistent with formation of the ?-allyl complex being a fast and reversible reaction as was noted above and either the transmetalation or the 105 reductive elimination is rate-determining (Scheme 4.11). If formation of ?-allyl complex were rate-determining, then the Hammett plot should have had a slope of zero because the aryl siloxane was not involved in that step of the mechanistic cycle. It is worth emphasizing that the term ?rate-determining step? in the discussion above is not precise. Rate-determining and product-determining have been used interchangably, although this is not valid in a strict interpretation. We know that formation of pi-allyl intermediate 123 is fast and reversible step (Scheme 4.13). Assuming transmetalation and reductive elimination is irreversible, if reductive elimination is slow, the product-determining step is transmetalation since it is the first irreversible step. This may not be the rate-determining step, but it is the step that leads to selective formation of one coupling product over the alternative. On the other hand, if transmetalation is slower than reductive elimination, transmetalation is still product-determining as well as rate- determining. In this study, relative rates of irreversible steps were measured and not the absolute rates. Nonetheless, regardless of which step is slow (rate-determining), transmetalation is the product determining step. Scheme 4.13 Si(OEt)3 F Ar1 Si(OEt)3 F Ar2 Ar1 Ar2 transmetalation reductive elimination Pd Ln 123 Pd Ar1 Ln Pd Ar2 Ln transmetalation reductive elimination 106 From the data provided it is not possible to unambiguously determine which of these two steps is rate-determining. For this coupling reaction, however, we propose that the rate-determining step is transmetalation, rather than reductive elimination, based on several lines of circumstantial evidence. First, since no coupled product 49 was obtained in the 7-deoxypancratistatin (24) synthesis, it is reasonable to assume that the transmetalation had not occurred to give palladium complex 88 (Scheme 4.3). Although the rationale for this conclusion is complex, the analysis is important for the mechanistic study in question. Formation of pi-allyl complex 99 (Scheme 4.3) was fast and reversible; thus the experimentally observed rearrangement of the cyclohexenyl carbonate was observed as a byproduct in this coupling (see Chapter 3). If the rate-determining step was transmetalation, then a slow transmetalation reaction would provide greater opportunity for rearrangement and decomposition of the cyclohexenyl starting material without leading to coupled product. If, on the other hand, reductive elimination were the rate-determining step, coupled product, even if only trace amounts, would have been obtained since it is unlikely that transmetalation was reversible. Reductive elimination must result in formation of coupled product or reduced arene (via beta-hydride elimination followed by reductive elimination). No trace of either coupled product or reduced arene was observed under these conditions. A second piece of evidence supporting the conclusion that transmetalation was rate-determining was derived from the study of Kurosawa on the reductive elimination of diorganopalladium complexes.179,180 Kurosawa was able to prepare allyl-aryl palladium complexes (utilizing an alternative methodology) and measured the rate of reductive 107 elimination at 0 ?C. If extrapolated to 55 ?C, as in our coupling protocol, the rates of reductive elimination would be much faster than the rate of coupling observed, thus suggesting that transmetalation, and not reductive elimination, is rate-determining step for the allyl-aryl coupling reaction reported in our study. Admittedly, the evidence is circumstantial, since the Kurosawa system involved different ligands on the metal center than our coupling protocol. Even more significant is that Kurosawa demonstrated that electron-deficient alkenes promoted, not retarded, the reductive elimination in his system. This observation is particularly germane to our coupling system in which the electron-deficient dba (dibenzylideneacetone) is a ligand on the catalyst that functions for this coupling. Assuming that the silicate system behaves analogously to the Kurosawa analog, we would anticipate that the reductive elimination step would be facilitated by the presence of electron-deficient ligand as was observed by Kurosawa.179-181 As has been noted by Denmark in his mechanistic studies of analogous silanol- based aryl-aryl coupling reactions, there is no unambiguous evidence for the intermediacy of a diorganopalladium (II) complex that undergoes reductive elimination to produce the product (Scheme 4.11).151 In the allyl-aryl system described herein, it is possible that ?-allyl complex 123 reacted with the activated silicate 122 via a substitution reaction to yield the product directly. Nonetheless, one would observe that this step would be rate-determining. Further support for our hypothesis that the rate-determining step in the coupling reaction is transmetalation, and not reductive elimination comes from Hartwig's lab. Hartwig has reported several mechanistic studies of palladium-catalyzed coupling 108 reactions in which reductive elimination is the rate determining step.182-186 In each of these studies, however, an isolable palladium (II) complex was prepared and then decomposed thermally via reductive elimination to provide product. Hammett analyses of these systems have shown significant substituent effects, but the results are not as straightforward as is observed in our system. For C-S bond formation, the substituent effect was qualitatively similar to those observed in our study, namely faster rates with electron-withdrawing substituents, but there was no correlation with Hammett sigma values. They were able to correlate the rates of reductive elimination only with a mixed Hammett value that included both inductive and resonance contributions.184 Analogous situations were observed for C-N and C-C coupling reactions, respectively.185,186 Role of Ligands The studies summarized above have established that Hammett analysis is an excellent method for gathering mechanistic information regarding the transition state (124) for the transmetalation step of the mechanism (Scheme 4.11). Furthermore, this Hammett methodology can be employed for investigating the roles of catalyst-ligand combinations in the coupling reaction. Typically, the development of an "optimized" catalyst system for a coupling reaction involves the empirical development of conditions and reagents using various ligand-metal ratios and is based on yield or turnover of product. The actual role that the ligand (Ln in 124, Scheme 4.11) plays in transmetalation cannot be assessed except in a qualitative sense: the reaction yield was high or low. Adding ligand or substituting a new metal may change the yield of the reaction, but does not provide precise mechanistic information about the rate-determining step in the 109 catalytic process. However, once the relative rates for various substituents had been measured under standardized conditions, we were able to extend our Hammett study to include various catalyst-ligand combinations. In particular, we were interested in determining whether the rate of transmetalation could be enhanced by changing ligands on the palladium. Ligands with different electronic and steric properties187 might be able to stabilize the transition state 124 (Scheme 4.11) differently and hence affect the rate of transmetalation. By measuring the relative rates between two aryl siloxanes, it should be possible to investigate the role that electronic factors play in the coupling reaction. As shown in the Table 4.2, entry 1, the best catalyst for the coupling of siloxanes and cyclohexenyl carbonate (120) is Pd(dba)2. Changing the catalyst to either Pd2(dba)3 or Pd2(dba)3?CHCl3 resulted in a considerable decrease in the yield of coupled product. Why the yield is diminished is less clear since all three of these Pd-complexes are thought to behave comparably as Pd(0) sources. On the other hand, the relative rates of p-anisoylsiloxane/phenylsiloxane (121:6) with all three catalyst systems, it was observed that the ratio of 0.42 ? 0.02 was maintained. The result clearly demonstrated that the rate- determining step in the coupling reaction was identical with all three catalysts systems and that another step in the mechanism must be responsible for the diminished yields of product. The effect of various added ligands on the coupling reaction was evaluated and the results are summarized in the Table 4.2. 110 OCO2Et Si(OEt)3 H H Si(OEt)3 OMe OMe Pd(0), 55 ?C Ligand THF, TBAF 120 9 1261216 Entry Pd (0) Source (10 mol %) Ligand (20 mol %) Relative Rate Yield (%)a 1 Pd(dba)2 - 0.42 80 2 Pd2(dba)3 - 0.42 63 3 Pd2(dba)3?CHCl3 - 0.40 61 4 Pd(dba)2 AsPh3 0.39 53 5 Pd(dba)2 PPh2(C6F5) 0.42 32 6 Pd(dba)2 PCy3 0.42 55 7 Pd(dba)2 P(o-tolyl)3 0.43 48 8 Pd(dba)2 PPh3 0.25 <20b 9 Pd(dba)2 I 0.33 <20b 10 Pd(dba)2 II - NA 11 Pd(dba)2 III - NA 12 Pd(dba)2 P(2-Fu)3 - NA 13 Pd(dba)2 - 0.45 79c 14 Pd(dba)2 - 0.39 64d 15 Pd(OAc)2 PCy3 0.39 <20b 16 Pd(OAc)2 AsPh3 0.44 <20b 17 Pd(OAc)2 PPh3 - NA 18 Pd(OAc)2 P(2-Fu)3 - NA 19 Pd2(dba-4,4'-OMe)3 - 0.39 55 20 Pd(dba-4,4'-CF3)2?H2O - 0.41 51 P t-Bu t-Bu P Cy Cy P P 128 129 130 a Yield determined by gas chromatography using a standard. b Yield determined by column chromatography. c DMF was used as a solvent. d Dioxane was used as a solvent. Note: relative rates and yields in entry 1 are averages of three runs, entries 2-9, 15-16 and 19-20 are averages of two runs. P(2-Fu)3 is tri-2-furylphosphine (TFP). Table 4.2: Role of ligands in the allyl-aryl coupling reaction 111 Strongly electron-donating ligands (PCy3 and P(o-tolyl)3, entries 6 and 7) with cone angles significantly greater than triphenylphosphine, gave higher yields of coupled products than triphenylphosphine, respectively, but did not alter the relative rate of arylated products. The larger cone angle of these phosphines would be expected to provide a ?-allyl complex with a low coordination number due to steric bulk, thus facilitating transfer of aryl group from the fluoride-activated siloxane. More electron- withdrawing ligands (entries 5 and 12) significantly reduced the yields of coupled product also, but had no effect on the relative rates. AsPh3 (entry 4) gave a better yield of coupled products when compared with PPh3 (entry 8). Since, the As-Pd bond is longer than the P-Pd bond; the As-Pd complex might be experiencing less steric hindrance from phenyl groups. With stannane derivatives, Farina had also shown that AsPh3 ligand dissociates more readily from palladium intermediates compared to PPh3, thus enhancing the rate of transmetalation (>103) compared to PPh3.117 Large deviations from the relative rate ratios were observed for the addition of ligands such as PPh3 and the Buchwald ligand 128188,189 (entries 8 and 9). Unfortunately, the yields of coupling products in these systems were so low that it is inappropriate to draw meaningful conclusions about the coupling reactions. The study of ligand effects on allyl-aryl coupling reaction revealed that Pd(dba)2 in absence of any phosphine ligands is the best catalytic system (Table 4.2, entry 1). It is known that electronic and steric properties of phosphine ligands can be tuned to achieve desired catalytic reactivity. We wondered if the same idea can be applied to design metal-olefin catalysts for our allyl-aryl coupling reaction. As seen in the Figure 4.11, the metal-olefin bond is described as a combination of ?-donation from a filled olefin 112 ?-orbital to an empty metal orbital and ?-back donation from a filled metal orbital to an empty olefin ?* orbital.190,191 For an electron rich d10-configuration such as Pd0, ?-back donation from palladium center to olefin is important. Stronger ?-back donation results in stronger palladium-olefin bond, thus increasing the stability of palladium complex, but reducing the lability of the olefin ligand. Both the electronic and steric nature of olefins plays an important role in determining the binding affinity to palladium. Electron deficient alkenes are bound more tightly to palladium because of increased ?-back- donation. Additionally, strained olefins such as norbornene possess high binding affinity to palladium because of relief of ring strain upon carbon rehybridization and reduction of steric hindrance.190 C C M ? donation C C M pi pi* pi donation Figure 4.11: Donor-Acceptor model for transition-metal-olefin complexes. Redrawn from ref 190. The catalytic activity as well as stability of metal-olefin complex can be tuned by adjusting the electronics and sterics of the olefin ligand. It is believed that for most transmetalations an open coordination site is required.190 We envisioned that an olefin ligand lacking an electron withdrawing group that can readily dissociate from palladium would facilitate transmetalation of the aryl group from silicon to palladium. However, such olefins may also reduce the stability of palladium complex. For example, palladium 113 complexes such as Pd(norbornene)3, Pd(COD)2 and Pd(ethene)3 are stable only at extremely low temperatures.192,193 First, effect of electronic nature of dba (dibenzylideneacetone) ligands (Figure 4.12) on the relative rate of allyl-aryl coupling reaction was studied (Table 4.2, entries 19 and 20). It was anticipated that electron donating groups (OMe) on the dba ligand will destabilize ?-back donation compared to more electron withdrawing groups (H, CF3). This would promote dissocation of dba from the palladium and faciliate transmetaltion. Accordingly, [Pd2(dba-4,4'-OMe)3] and [Pd(dba-4,4'-CF3)2?H2O] catalysts were prepared according to Fairlamb's procedure194 and a competition experiment of allyl carbonate 120 was performed. The change of electronic character of the dba ligands, respectively, did not change the relative rate; however the yield of coupled product decreased (entries 19 and 20, Table 4.2). In contrast to our coupling results, the Suzuki coupling of aryl chlorides with arylboronic acids performed by Fairlamb showed that electron donating groups on Pd(dba)2 increased the rate of the cross-coupling reaction, relatively to unsubstituted dba (131, R = H).194 Very recently, Fairlamb has reported palladium(0) complexes containing thienyl analogues (132, 133) of dibenzylideneacetone (dba) and evaluated their reactivity in oxidation addition reaction with iodobenzene.195 O R Rdba R = OMe, H, CF3 131 O 132 S O S S 133 Figure 4.12: Various alkenyl ligands 114 Pd2(dba)3 serve as an important precursor for the synthesis of zerovalent palladium complexes bearing additional ligands such as phosphines, nitrogen, sulfur as well as olefins.112-114,196-208 The stability of these complexes is governed by subtle interplay between electron-donating and electron-withdawing properties of the ligand. Our intial attempts involved preparation of mixed olefin complexes according to the procedure reported by Itoh and co-workers.112 Itoh prepared mixed olefin complexes of Pd(0) by a ligand substitution of Pd2dba3?CHCl3. By an appropriate combination of electron-donating and electron-withdrawing olefin ligands it was possible to isolate three- coordinate mixed olefin complexes (IV, IX, X, Figure 4.13). Catalysts IV, IX and Xwere prepared and used in the aryl-allyl coupling reaction (Table 4.3, entries 1-3). O OO Pd Pd CN CNNC NC Pd(NBD)(MAH) Pd(COD)(TCNE) Pd O OO Pd(COD)(MAH) IV XIX Figure 4.13: Pd(NBD)(MAH), Pd(COD)(MAH) and Pd(COD)(TCNE) complexes As seen in Table 4.3, replacing olefin MAH (maleic anhydride) with TCNE (tetracyanoethylene), but retaining COD (cyclooctadiene) as diene, reduced the yield of coupled product dramatically (entries 2 and 3). On other hand, by subsituting COD (cyclooctadiene) with NBD (norbornadiene) but retaining MAH improved yield (entries 2 and 4). Interestingly, the catalytic activity of Pd(NBD)(MAH) IV and Pd(COD)(MAH) 115 IX does not correlate to the strain energy of NBD and COD. One would anticpate NBD having larger strain energy compared to COD to bind tighter to the palladium due to the relief of strain and thus be less labile. However, higher yields with Pd(NBD)(MAH) IV compared to Pd(COD)(MAH) IX suggested NBD to be more labile olefin. This result is constistent with that of Orchad and Weiss who showed despite higher strain energy, NBD bings less strongly to metal (copper or silver) compared to COD and pointed to significance of steric factors.209,210 It is important to note that both Pd(NBD)(MAH) IV and Pd(COD)(MAH) IX catalyzed allyl-aryl coupling reaction even at ambient temperature, which is not the case with Pd(dba)2. This suggests that these catalysts are far more reactive than Pd(dba)2. However they produce lower yields compared to Pd(dba)2 (Table 4.3, entry 1). This can be attributed to instability of these complexes in solution compared to Pd(dba)2. During coupling reaction precipitation of palladium black is observed within 2-3 hrs when using Pd(NBD)(MAH) IV and Pd(COD)(MAH) XI, but this is not the case with Pd(dba)2. While the preparation of Pd(NBD)(MAH) IV was easy, the preparation of Pd(COD)(MAH) XI and Pd(COD)(TCNE) X was tedious and required careful handling due to their instability (precipitation of palladium black) when exposed to air and ambient temperature. This was also the case with catalysts Pd(cyclopentene)(MAH)2 and Pd(norbornene)MAH2 which could not be isolated due to extensive decomposition to palladium black. 116 OCO2Et Si(OEt)3 Pd(0) TBAF 120 6 9 Entry Pd(0) (25 mol%) TBAF equiv. Siloxane equiv. Yield (%) 1 Pd(dba)2 2.0 2.0 70 2 Pd(COD)(MAH) 2.0 2.0 38 3 Pd(COD)(TCNE) 2.0 2.0 17 4 Pd(NBD)(MAH)a 2.0 2.0 54 5 Pd(COD)(NQ) 2.0 2.0 48 6 Pd(COD)(DQ) 2.0 2.0 23 7 Pd(COD)(NQ) 2.0 2.0 41b 8 Pd(COD)(NQ) 8.0 8.0 45 9 Pd(COD)(BQ) 2.0 2.0 20 10 Pd(NBE)2(BQ)2 2.0 2.0 21 All reactions were performed at 55 ?C for 24 h. a 12 mol% of catalyst used. b Reaction in entry 6 was performed in DMF. Table 4.3: Optimization of Pd(0)-olefin catalyzed allyl-aryl coupling reaction The instability of mixed olefin complexes prepared by Itoh limited their application in the allyl-aryl coupling reaction. We next chose to examine Pd(0)-triolefin complexes containing quinones prepared from Pd2dba3?CHCl3 (Figure 4.14).113,114 These catalysts are stable at ambient temperature and therefore easy to prepare. When employed in the allyl-aryl coupling reaction, Pd(COD)NQ V gave superior yields in comparison to palladium complexes formed using Pd(COD)BQ XI and Pd(COD)DQ XII, indicating a strong electronic influence on the catalyst's performance. It is interesting to note that subtle changes such as inclusion of aromatic ring (NQ = naphthoquinone) in the catalyst dramatically improve the yield compared to BQ (benzoquinone) and DQ (duroquinone). 117 Optimization of reaction conditions using Pd(COD)(NQ) V showed no improvements in yield. Dinuclear Pd(0) complex possessing BQ as a bridging ligand and NBE (norbornene) as a monodentate ligand was also prepared114, however, this catalyst XIII resulted in poor yield (Table 3, entry 10). Pd O O O O Pd O O Pd Pd2(NBE)2(BQ)2Pd(COD)(NQ) Pd(COD)(DQ) Pd Pd(COD)(BQ) V XI XII XIII OO Pd OO Figure 4.14: Pd(COD)(NQ), Pd(COD)(BQ), Pd(COD)(DQ), Pd2(NBE)2(BQ)2 complexes From the results summarized in Table 4.3, Pd(NBD)(MAH) IV and Pd(COD)(NQ) V were identified as appropriate cataysts for allyl-aryl coupling reaction using cyclohexenyl carbonate as substrate 120. Next, these catalysts were applied in the coupling of more complex carbonate 87 with aryl siloxane 6 and siloxane 72. (Scheme 4.14). It was observed that even stoichiometric amounts of Pd(NBD)(MAH) IV gave worse yields (11-13%) compared to Pd(COD)(NQ) V for the coupling reaction (Scheme 4.13). Thus, Pd(COD)(NQ) V is the best catalyst as of now for the reaction shown in the Scheme 4.14. 118 Scheme 4.14 Si(OEt)3OCO2Et NHE O O NHE O O TBAF, rt or 55 ?C 30-40% 87 Pd(COD)(NQ) O O Si(OEt)3 72 Pd(COD)(NQ) O O NHE O O 49 104 6 Si(OEt)3 NHE O O Pd(NBD)(MAH) 104 6 TBAF, rt or 55 ?C 30-40% TBAF, rt or 55 ?C 11-19% E = CO2Me Though Pd(dba)2 works well (70% yield, Table 4.3) with the simple allyl-aryl coupling reaction it does not function in the coupling of complex allyl carbonate 87 with aryl siloxane 72. This might be because electron withdrawing dba disssociates less readily from the palladium compared to COD, hindering transfer of aryl group to the palladium center. However, Pd(COD)(NQ) results in lower yields compared to Pd(dba)2 in case of cyclohexenyl carbonate (Table 4.3) presumably because of its decompostion to palladium black. The coupling product 49 is the desired compound for the synthesis of natural product 7-deoxypancratistatin (24) (see Chapter 3). The moderate yield of the coupling reaction is one of the drawbacks of this reaction. Future goals would aim at application of different kinds olefin-based catalysts in palladium-catalyzed allylic arylation to optimize the key reaction. 119 CONCLUSION Mechanistic studies on coupling reaction of allyl carbonate and aryl siloxane has offered several useful insights; 1. On the basis of Hammett analaysis of allyl-aryl coupling of para-substituted siloxane derivatives with cylcohexenyl carbonate, the rate-determining step of the coupling reaction was identified as either transmetalation or reductive elimination. Furthermore, it was observed that the rate of coupling reaction was enhanced by electron-withdrawing substituents, indicating deveopment of negative charge on the aromatic ring in the transition state of the rate-determining step. 2. Competition studies as a function of ligand type revealed that electronic as well as the steric nature of phosphine ligands on the metal site dramatically affect yield of coupled product, but rarely affect the relative rates of the coupling reaction. This idea was used to explore steric and electronic nature of olefin-based palladium catalyst on allyl-aryl coupling reaction. 3. A new family of catalysts for the siloxane coupling process that overcomes the limitations have been developed. These catalyst are more reactive than Pd(dba)2 which is apparent from their ability to catalyze the reaction at ambient temperature. 4. Tri-olefin-based Pd(0) catalyst, Pd(COD)(NQ) has been succesfully applied in the coupling reaction of complex allyl carbonate 87 with aryl siloxane 72 (Scheme 4.14) to produce much needed coupling product 49 (Hudlicky's intermediate) for the synthesis of natural product 7-deoxypancratistatin (24). 120 EXPERIMENTAL DETAILS General Methods All reactions were run under an atmosphere of argon unless otherwise noted. Glassware used in the reactions was dried for a minimum of 12 h in an oven at 120 ?C. Tetrahydrofuran was distilled from sodium/benzophenone ketyl, while methylene chloride, pyridine, dimethylformamide and dioxane were distilled from calcium hydride. PPh3, PCy3, and 2-(dicyclohexylphosphino)biphenyl were recrystallized from hexanes prior to use. P(o-tolyl)3 and P(2-Fu)3 were recrystallized from ethanol. AsPh3, PPh2(C6F5) and 2-(di-t-butylphosphino)biphenyl were used as received. [Pd2(dba-4,4'-OMe)3] and [Pd(dba-4,4'-CF3)2?H2O] catalysts were prepared using the procedure reported by Fairlamb.194 Pd(COD)(TCNE)112, Pd(COD)(MAH)112, Pd(NBD)(MAH)112 Pd(COD)(NQ)113, Pd(COD)(BQ)113, Pd(COD)(DQ)113, Pd2(NBE)2(BQ)2114 were prepared as reported in the literature. 19F NMR and 29Si NMR were recorded on a high field 500 MHz NMR spectrometer. 19F and 29Si chemical shifts are referenced to external standard TFA and TMS respectively. Gas chromatography was performed on a Hewlett Packard 5890 GC equipped with a flame ionization detector using a 25m methyl silicon column. 121 Allyl carbonate 120 OH ClCO2Et, pyridine CH2Cl2, rt, 64% OCO2Et 120 To 2.09 g (21.3 mmol, 1.00 equiv.) of commercially available 2-cyclohexen-1-ol in 20.0 mL anhydrous CH2Cl2 and 2.57 mL (31.9 mmol, 1.50 equiv.) anhydrous pyridine was added 3.17 mL (31.9 mmol, 1.50 equiv.) of ethyl chloroformate dropwise via syringe under argon. The reaction was allowed to stir at room temperature for 7 days. The reaction mixture was extracted with CH2Cl2 (3 ? 50 mL), washed with H2O (50 mL), dried over MgSO4 and concentrated in vacuo. Flash chromatography on silica gel (5% EtOAc/95% hexane, Rf = 0.51) afforded 2.32 g (64%) of the allyl carbonate 120 as a colorless oil; IR (CCl4) 3042 (w), 2981 (w), 2947 (m), 2875 (w), 2838 (w), 1737 (s), 1373 (s), 1265 (s), 1017 (s) cm-1; 1H NMR (400 MHz, CDCl3) ? 5.97-5.93 (m, 1H), 5.77- 5.73 (m, 1H), 5.10-5.09 (m, 1H), 4.16 (q, J = 7 Hz, 2H), 2.05 (m, 1H), 2.05-1.98 (m, 1H), 1.88-1.80 (m, 3H), 1.62 (m, 1H), 1.28 (t, J = 7 Hz, 3H); 13C NMR (100 MHz, CDCl3) ? 154.8, 133.2, 125.0, 71.5, 63.6, 28.2, 24.8, 18.5, 14.2. The spectral data (1H NMR) were identical to that reported in the literature.211 122 General procedure for competition experiments for allyl-aryl coupling reaction (Table 4.2) OCO2Et Si(OEt)3 H H Si(OEt)3 R R Pd(dba)2, 55 ?C THF, TBAF 6 9 126 120 R = OMe, Me, Cl, CO2Et 121 To 121 mg (0.712 mmol, 1.00 equiv.) of allyl carbonate 120, 342 mg (1.42 mmol, 2.00 equiv.) of aryl siloxane 6 and 383 mg (1.42 mmol, 2.00 equiv.) of p-anisoylsiloxane 121 (R = OMe) dissolved in 4.00 mL of anhydrous THF was added 40.9 mg (0.0712 mmol, 0.100 equiv.) of Pd(dba)2 under an atmosphere of argon. This was followed by addition of 2.84 mL (2.84 mmol, 4.00 equiv.) of 1 M TBAF solution in THF and the reaction mixture was stirred at 55 ?C for 24 h. The product was extracted with 5 ? 20 mL Et2O and washed with 20 mL H2O. The combined organic layers were dried over MgSO4 and concentrated in vacuo to give coupling products 9 (R = H) and 126 (R = OMe). The crude product was filtered through a short silica plug. The relative quantity of 0.44 for p-anisoylsiloxane 121 (R = OMe) was determined from the amount of methoxyphenylcyclohexene 126 (R = OMe) obtained relative to phenylcyclohexene 9 using GC (gas chromatography). The same experimental procedure was used to determine relative rates of different aryl siloxanes 121. Moreover, effects of different catalysts, ligands and solvents on the relative rate of transmetalation were examined 123 (Table 4.2), using analogous competition experiments, where yields were determined by GC using standard unless otherwise noted. Aryl siloxanes 121 Si(OEt)3R 121 R = OMe, Me, Cl, CO2Et p-anisoylsiloxane (R = OMe), p-tolylsiloxane (R = Me), and p-chlorophenylsiloxane (R = Cl) were prepared from commercially available p-bromoansiole, p-bromotoluene and p-chloroiodobenzene respectively according to the procedure previously reported by DeShong and Manoso.20,212 p-Carboethoxyphenylsiloxane (R = CO2Et) was prepared from ethyl-4-iodobenozate by Masuda's procedure.213 The 1H NMR spectral data of aryl siloxanes 121 (R = OMe212, R = Me212, R = Cl20, R = CO2Et213) were identical to that reported in the literature. Alkene (9, 126) R 126 R = OMe, Me, Cl, CO2Et 9 R = H The spectral data of phenylcyclohexene214 9, methoxyphenylcyclohexene.32,215 (126, R = OMe), methylphenylcyclohexene32,33 (126, R = Me), chlorophenylcyclohexene32,33 (126, R = Cl) and carboethoxyphenylcyclohexene33 126 (R = CO2Et) matched to those previously reported in the DeShong group. 124 General procedure for allyl-aryl coupling using Tri-olefin-based Pd(0) catalysts (Table 4.3) OCO2Et Si(OEt)3 Pd(0) TBAF 120 6 9 To 76.5 mg (0.450 mmol, 1.00 equiv.) of allyl carbonate 120, 216 mg (0.900 mmol, 2.00 equiv.) of aryl siloxane 6 in 4.00 mL of anhydrous THF was added 41.9 mg (0.113 mmol, 0.250 equiv.) of Pd(COD)(NQ) under an atmosphere of argon. This was followed by addition of 0.900 mL (0.900 mmol, 2.00 equiv.) of 1 M TBAF solution in THF and the reaction mixture was stirred at 55 ?C for 24 h. The product was extracted with 5 ? 25 mL Et2O and washed with 25 mL H2O. The combined organic layers were dried over MgSO4 and concentrated in vacuo to give crude cyclohexene 9. Purification by column chromatography (100% pentane; Rf = 0.65) rendered 37.0 mg (52%) of cyclohexene 9 as a colorless oil. 1H NMR (400 MHz, CDCl3) ? 7.31-7.27 (m, 2H), 7.22-7.18 (m, 3H), 5.89-5.86 (m, 1H), 5.72-5.69 (m, 1H), 3.41-3.38 (m, 1H), 2.08-1.99 (m, 3H), 1.74-1.71 (m, 1H), 1.61-1.53 (m, 2H); The NMR spectrum was identical to the literature.214 125 Formation of hypercoordinate silicates Triethoxyphenylsilane (0.18 mmol, 43 mg) was mixed with 1 M TBAF (0.18 mmol, 0.18 mL) in 0.50 mL THF and 19F NMR spectrum was recorded after 10 minutes at 29 ?C. 19F spectrum indicated two major resonances at ? -121 and -127 and two minor resonances at ? -113 and -128 (see Figure 1, (b)). After 1 h, TBAF and siloxane mixture was cooled to -28 ?C. Upon cooling, the 19F signal at ? -121 sharpened, showing silicon satellites (JSi-F = 207 Hz). 19F NMR spectrum of hypercoordinate silicates is provided in Chapter 4, Figure 4.8. While 19F NMR was obtained at 0.36 mM concentration, 19Si NMR required higher concentration (3 ? 0.36 mM) and longer time (18 h). The chemical shift of triplet at ? -129 ppm indicates hypercoordinate silicate anion, JSi-F = 207 Hz. Figure 4.15: 29Si NMR spectrum of silicate formation at -28 ?C. JSi-F = 207 Hz 126 Based on 19F and 29Si NMR, we have tentatively assigned conformationally stable bis-fluorosilicate derivative 135 or 136 as the hypercoordinate arylfluorosilicate which arises from conformationally mobile mono-fluorosilicate 134 (Scheme 4.15). The 19F and 29Si NMR spectra and coupling constant are consistent with that of TBAT, previously characterized by the DeShong group. Additionally, it was observed that bis-fluorosilicate when reacted with allyl carbonate 120, gave coupled product in 60% yield. Scheme 4.15 Si(OEt)3 F Si(OEt)3 F Si OEt OEt F F Si F F OEt OEt 60% Si Ph Ph F F TBAT 29Si shift = ? -106 19F shift = ? -96 J Si-F = 252 Hz Si(OEt)2 F OEt F 136 135 120 134 Pd(dba)2, 55 ?C OCO2Et F 127 Figure 4.16: Effect of temperature on silicate formation (19F NMR spectrum). Mixture of TBAF and phenyltriethoxysilane (a) 10 min, at rt. (b) 2 h 25 min, at -30 ?C. (c) 2 h 40 min, at rt. (d) 4 h 55 min, at -30 ?C. 128 Figure 4.17: Effect of TBAF concentration on silicate formation (19F NMR spectrum) (a) 0.5 equiv. TBAF, 1.0 equiv. siloxane. (b) 1.0 equiv. TBAF, 1.0 equiv. siloxane. (c) 1.5 equiv. TBAF, 1.0 equiv. siloxane. (d) 2.0 equiv. TBAF, 1.0 equiv. siloxane. 129 REFERENCES 1. Diederich, F.; Stang, P. J., Eds. Metal-catalyzed cross-Coupling Reactions. Wiley-VCH: New York, 1998. 2. Tsuji, J., Eds. Palladium Reagents and Catalysts. Innovations in Organic Synthesis; John Wiley & Sons: New York, 1995. 3. Stille, J. K. Pure Appl. Chem. 1985, 57, 1771-1780. 4. Miyaura, N.; Suzuki, A. Chem Rev. 1995, 95, 2457-2483. 5. Hiyama, T. J. Org. Chem. 1988, 53, 918-920. 6. Hiyama, T.; E. Shirakawa, E. Top. Curr. Chem. 2002, 219, 61-85. 7. Hatanaka, Y.; Goda, K.; Okahara, Y. Tetrahedron 1994, 50, 8301-8316. 8. Hiyama, T.; Hatanaka, Y. Pure Appl. Chem. 1994, 66, 1471-1478. 9. Horn, K. A. Chem. Rev. 1995, 95, 1317-1350. 10. Hiyama, T. J. Organomet. Chem. 2002, 653, 58-61. 11. Denmark, S. E.; Sweis, R. F. Acc. Chem. Res. 2002, 35, 835-846. 12. Mowery, M. E.; DeShong, P. J. Org. Chem. 1999, 64, 3266-3270. 13. Mowery, M. E.; DeShong, P. J. Org. Chem. 1999, 64, 1684-1688. 14. Mowery, M. E.; DeShong, P. Org. Lett. 1999, 1, 2137-2140. 15. DeShong, P.; Handy, C. J.; Mowery, M. E. Pure Appl. Chem. 2000, 72, 1655- 1658. 16. Seganish, W. M.; DeShong, P. J. Org. Chem. 2004, 69, 1137-1143. 17. Lee, H. M.; Nolan, S. P. Org. Lett. 2000, 2, 2053-2055. 18. Wolf, C.; Lerebours, R. Org. Lett. 2004, 6, 1147-1150. 19. Clarke, M. L. Adv. Synth. Catal. 2005, 347, 303-307. 20. Manoso, A. S.; DeShong, P. J. Org. Chem. 2001, 66, 7449-7455. 21. Riggleman, S.; DeShong, P. J. Org. Chem. 2003, 68, 8106-8109. 130 22. Denmark, S. E.; Sweis, R. F. J. Am. Chem. Soc. 2001, 123, 6439-6440. 23. Denmark, S. E.; Ober, M. H. Adv. Synth. Catal. 2004, 346, 1703-1714. 24. Denmark, S. E.; Smith, R. C.; Chang, W. T.; Muhuhi, J. M. J. Am. Chem. Soc. 2009, 131, 3104-3118. 25. Denmark, S. E.; Regens, C. S. Acc. Chem. Res. 2008, 41, 1486-1499. 26. Trost, B. M.; Van Vranken, D. L. Chem. Rev. 1996, 96, 395-422. 27. Hiyama, T.; Hatanaka, Y.; Mori, A.; Matsuhashi, H.; Asai, S.; Hirabayashi, K. Bull. Chem. Soc. Jpn. 1997, 70, 1943-1952. 28. Nakao, Y.; Ebata, S.; Chen, J.; Imanaka, H.; Hiyama, T. Chem. Lett. 2007, 36, 606-607. 29. Brescia, M. R.; DeShong, P. J. Org. Chem. 1998, 63, 3156-3157. 30. Brescia, M. R.; Shimshock, Y. C.; DeShong, P. J. Org. Chem. 1997, 62, 1257- 1263. 31. Hoke, M. E.; Brescia, M. R.; Bogacyzk, S.; DeShong, P.; King, B. W.; Crimmins, M. T. J. Org. Chem. 2002, 67, 327-335. 32. Correia, R.; DeShong, P. J. Org. Chem. 2001, 66, 7159-7165. 33. Correia, R. Ph.D. Dissertation, University of Maryland, College Park, MD, 2003. 34. Bogaczyk, S. Ph.D. Dissertation, University of Maryland, College Park, MD, 2002. 35. Dey, R.; Chattopadhyay, K.; Ranu, B. C. J. Org. Chem. 2008, 73, 9461-9464. 36. Kabalka, G. W.; Dong, G.; Venkataiah, B.; Chen, C. J. Org. Chem. 2005, 70, 9207-9210. 37. Legros, J. Y.; Fiaud, J. C. Tetrahedron Lett. 1990, 31, 7453-7456. 38. Moreno-Ma?as, M.; Pajuelo, F.; Pleixats, R. J. Org. Chem. 1995, 60, 2396- 2397. 39. Uozumi, Y.; Danjo, H.; Hayashi, T. J. Org. Chem. 1999, 64, 3384-3388. 40. Ramnauth, J.; Poulin, O.; Rakhit, S.; Maddaford, S. P. Org. Lett. 2001, 3, 2013-2015. 41. Bouyssi, D.; Gerusz, V.; Balme, G. Eur. J. Org. Chem. 2002, 2445-2448. 131 42. Ortar, G. Tetrahedron Lett. 2003, 43, 4311-4314. 43. Mino, T.; Kajiwara, K.; Shirae, Y.; Sakamoto, M.; Fujita, T. Synlett 2008, 2711-2715. 44. Ohmiya, H.; Makida, Y.; Tananka, T.; Sawamura, M. J. Am. Chem. Soc. 2008, 130, 17276-17277. 45. Singh, R.; Viciu, M. S.; Kramareva, N.; Navarro, O.; Nolan, S. P. Org. Lett. 2005, 7, 1829-1832. 46. N?jera, C.; Gil-Molt?, J.; Karlstr?m, S. Adv. Synth. Catal. 2004, 346, 1798- 1811. 47. Alacid, E.; N?jera, C. Org. Lett. 2008, 10, 5011-5014. 48. Manabe, K.; Nakada, K.; Aoyama, N.; Kobayashi, S. Adv. Synth. Catal. 2005, 347, 1499-1503. 49. Kayaki, Y.; Koda, T.; Ikariya, T. Eur. J. Org. Chem. 2004, 4989-4993. 50. Tsukamoto, H.; Sato, M.; Kondo, Y. Chem. Commun. 2004, 1200-1201. 51. Kabalka, G. W.; Dong, G.; Venkataiah, B.; Org. Lett. 2003, 5, 893-895. 52. Kabalka, G. W.; Dadush, E.; Al-Masum, M. Tetrahedron Lett. 2006, 47, 7459-7461. 53. Kabalka, G. W.; Al-Masum, M. Org. Lett. 2006, 8, 11-13. 54. Chung, K.-G.; Miyake, Y.; Uemura, S. J. Chem. Soc.; Perkin Trans. 1, 2000, 15-18. 55. Hansen, A. L.; Ebran, J.-P.; G?gsig, T. M.; Skrydstrup, T. J. Org. Chem. 2007, 72, 6464-6472. 56. Menard, F.; Chapman, T. M.; Dockendorff, C.; Lautens, Mark. Org. Lett. 2006, 8, 4569-4572. 57. Stille, J. K.; Hegedus, L. S.; Del Valle, L. J. Org. Chem. 1990, 55, 3019-3023. 58. Casta?o, A. M.; Echavarren, A. M. Tetrahedron Lett. 1996, 37, 6587-6590. 59. Kurosawa, H.; Kajimaru, H.; Ogoshi, S.; Yoneda, H.; Miki, K.; Kasai, N.; Murai, S.; Ikeda, I. J. Am. Chem. Soc. 1992, 114, 8417-8424. 60. Crawforth, C. M.; Burling, S.; Fairlamb, I. J. S.; Taylor, R. J. K.; Whitwood, A. C. Chem. Commun. 2003, 2194-2195. 132 61. Shipe, W. D.; Sorensen, E. J. Org. Lett. 2002, 4, 2063-2066. 62. Nicolaou, K. C.; Koftis, T. V.; Vyskocil, S.; Petrovic, G.; Tang, W.; Frederick, M. O.; Chen, D. Y.-K.; Li, Y.; Ling, T.; Yamada, Y. M. A. J. Am. Chem. Soc. 2006, 128, 2859-2872. 63. Pettit, G. R.; Gaddamidi, V.; Cragg, G. M.; Herald, D. L.; Sagawa, Y. J. Chem. Soc, Chem. Commun. 1984, 1693-1694. 64. Ghosal, S.; Singh, S.; Kumar, Y.; Srivastava, R. S. Phytochemistry 1989, 28, 611-613. 65. Gabrielsen, B.; Monath, T. P.; Huggins, J. W.; Kefauver, D. F.; Pettit, G. R.; Groszek, G.; Hollingshead, M.; Kirsi, J. J.; Shannon, W. M.; Schubert, E. M.; DaRe, J.; Ugarkar, B.; Ussery, M. A.; Phelan, M. J. J. Nat. Prod. 1992, 55, 1569-1581. 66. Hudlicky, T.; Moser, M.; Banfield, S. C.; Rinner, U.; Chapuis, J.-C.; Pettit, G. R. Can. J. Chem. 2006, 84, 1313-1337. 67. Pettit, G. R.; Gaddamidi, V.; Herald, D. L.; Singh, S. B.; Cragg, G. M.; Schmidt, J. M.; Boettner, F. E.; Williams, M.; Sagawa, Y. J. Nat. Prod. 1986, 49, 995-1002. 68. Pettit, G. R.; Pettit III, G. R.; Bachaus, R. A.; Boyd, M. R.; Meerow, A. W. J. Nat. Prod. 1993, 56, 1682-1687. 69. Pandey, S.; Kekre, J.; Naderi, J.; McNulty, J. Artif. Cells, Blood Substitutes, Immobilization Biotechnol. 2005, 33, 279-295. 70. Kekre, N.; Griffin, C.; McNulty, J.; Pandey, S. Cancer Chem Pharmacol. 2005, 56, 29-38. 71. Griffin, C.; McNulty, J.; Hamm, C.; Pandey, S. In Cell Apoptosis Research Trends; Zhang, C. V., Ed.; Nova Science Publishers, Inc. 2007, pp 93-109. 72. Andersen, M. H.; Becker, J. C.; Straten, P. Nat. Rev. Drug Discov. 2005, 4, 399-409. 73. Rinner, U.; Hudlicky, T. Synlett. 2005, 3, 365-387 and references therein. 74. Chapleur, Y.; Chr?tien, F.; Ahmed, U.; Khaldi, M. Curr. Org. Synth. 2006, 3, 341-378 and references therein. 75. Manpadi, M.; Kornienko, A. Org. Prep. Proced. Int. 2008, 40, 107-161 and references therein. 133 76. Kornieno, A.; Evidente, A. Chem. Rev. 2008, 108, 1982-2014 and references therein. 77. Danishefsky, S.; Lee, J. Y. J. Am. Chem. Soc. 1989, 111, 4829-4837. 78. Tian, X.; Hudlicky, T.; K?nigsberger, K. J. Am. Chem. Soc. 1995, 117, 3643- 3644. 79. Trost, B. M.; Pulley, S. R. J. Am. Chem. Soc. 1995, 117, 10143-10144. 80. Hudlicky, T.; Tian, X.; K?nigsberger, K.; Maurya, R.; Rouden, J.; Boreas, F. J. Am. Chem. Soc. 1996, 118, 10752-10765. 81. Doyle, T. J.; Hendrix, M.; VanDerveer, D.; Javanmard, S.; Haseltine, J. Tetrahedron 1997, 53, 11153-11170. 82. Magnus, P.; Sebhat, I. K. Tetrahedron 1998, 54, 15509-15524. 83. Rigby, J. H.; Maharoof, U. S. M.; Mateo, M. E. J. Am. Chem. Soc. 2000, 122, 6624-6628. 84. Pettit, G. R.; Melody, N.; Herald, D. L. J. Org. Chem. 2001, 66, 2583-2587. 85. Ko, H.; Kim, E.; Park, J. E.; Kim, D.; Kim, S. J. Org. Chem. 2004, 69, 112- 121. 86. Li, M.; Wu, A.; Zhou, P. Tetrahedron Lett. 2006, 47, 3707-3710. 87. Tian, X.; Maurya, R.; K?nigsberger, K.; Hudlicky, T. Synlett. 1995, 11, 1125- 1126. 88. Chida, N.; Jitsuoka, M.; Yamamoto, Y.; Ohtsuka, M.; Ogawa, S. Heterocycles, 1996, 43, 1385-1389. 89. Keck, G. E.; McHardy, S. F.; Murry, J. A. J. Org. Chem. 1999, 64, 4465- 4476. 90. Keck, G. E.; Wager, T. T.; McHardy, S. F. J. Org. Chem. 1998, 63, 9164- 9165. 91. Ace?a, J. L.; Arjona, O.; Le?n, M. L.; Plumet, J. Org. Lett. 2000, 2, 3683- 3686. 92. H?kansson, A. E.; Palmelund, A.; Holm, H.; Madsen, R. Chem. Eur. J. 2006, 12, 3243-3253. 93. Padwa, A.; Zhang, H. J. Org. Chem. 2007, 72, 2570-2582. 134 94. Lopes, R. S. C.; Lopes, C. C.; Heathcock, C. H. Tetrahedron Lett. 1992, 33, 6775-6778. 95. Friestad, G. K.; Branchaud, B. P. Tetrahedron Lett. 1997, 38, 5933-5936. 96. Grubb, L. M.; Dowdy, A. L.; Blanchette, H. S.; Friestad, G. K.; Branchaud, B. P. Tetrahedron Lett. 1999, 40, 2691-2694. 97. Ibn Ahmed, S.; Chr?tien, F.; Chapleur, Y.; Hajjaj, N. Heterocycl. Commun. 1997, 3, 135-138. 98. Hudlicky, T.; Rinner, U.; Gonzalez, D.; Akg?n, H.; Schilling, S.; Siengalewicz, P.; Martinot, T. A.; Pettit, G. R. J. Org. Chem. 2002, 67, 8726- 8743. 99. Chida, N.; Ohtsuka, M.; Ogawa, S. Tetrahedron Lett. 1991, 32, 4525-4528. 100. Banwell, M. G.; Cowden, C. J.; Gable, R. W. J. Chem. Soc. Perkin Trans. 1 1994, 3515-3518. 101. Pandey, G.; Balakrishnan, M.; Swaroop, P. S. Eur. J. Org. Chem. 2008, 5839- 5847. 102. Pandey, G.; Murugan, A.; Balakrishnan, M. Chem. Commun. 2002, 624-625. 103. Jenkins, N. E.; Ware, R. W.; Atkinson, R. N.; King, S. B. Synth.Commun. 2000, 30, 947-953. 104. Tranmer, G. K.; Tam, W. Org. Lett. 2002, 4, 4101-4104. 105. Jonasson, C.; Kritikos, M.; B?ckvall, J.-E.; Szab?, K. Chem. Eur. J. 2000, 6, 432-436 and references therein. 106. Chern, M.-S.; Li, W.-R. Tetrahedron Lett. 2004, 45, 8323-8326. 107. Castillo, P.; Rodriguez-Ubis, J. C.; Rodriguez, F. Synthesis 1986, 839-840. 108. Kelly, T. R. J. Org. Chem. 1972, 37, 3393-3397. 109. Paulsen, H.; Stubbe, M. Liebigs Ann. Chem. 1983, 535-556. 110. Shukla, K. H.; Boehmler, D. J.; Bogacyzk, S.; Duvall, B. R.; Peterson, W. A.; McElroy, W. T.; DeShong, P. Org. Lett. 2006, 8, 4183-4186. 111. Yang, N.-C.; Chen, M.-J.; Chen, P. J. Am. Chem. Soc. 1984, 106, 7310-7315. 112. Itoh, K.; Ueda, F.; Hirai, K.; Ishii, Y. Chem. Lett. 1977, 877-880. 135 113. Hiramatsu, M.; Shiozaki, K.; Fujinami, T.; Sakai, S. J. Organomet. Chem. 1983, 246, 203-211. 114. Yamamoto, Y.; Ohno, T.; Itoh, K. Organometallics 2003, 22, 2267-2272. 115. Labadie, J. W.; Stille, J. K. J. Am. Chem. Soc. 1983, 105, 669-670. 116. Labadie, J. W.; Stille, J. K. J. Am. Chem. Soc. 1983, 105, 6129-6137. 117. Farina, V.; Krishnan, B. J. Am. Chem. Soc. 1991, 113, 9585-9595. 118. Louie, J.; Hartwig, J. F. J. Am. Chem. Soc. 1995, 117, 11598-11599. 119. Amatore, C.; Bahsoun, A. A.; Jutand, A.; Meyer, G.; Ntepe, A. N.; Ricard, L. J. Am. Chem. Soc. 2003, 125, 4212-4222. 120. Casado, A. L.; Espinet, P. J. Am. Chem. Soc. 1998, 120, 8978-8985. 121. Casado, A. L.; Espinet, P.; Gallego, A. M. J. Am. Chem. Soc. 2000, 122, 11771-11782. 122. Espinet, P.; Echavarren, A. M. Angew. Chem. Int. Ed. 2004, 43, 4704-4734. 123. Casares, J. A.; Espinet, P.; Salas, G. Chem. Eur. J. 2002, 8, 4843-4853. 124. Itami, K.; Kamei, T.; Yoshida, J. J. Am. Chem. Soc. 2001, 123, 8773-8779. 125. Crisp, G. T.; Gebauer, M. G. Tetrahedron Lett. 1995, 36, 3389-3392. 126. Kakusawa, N.; Yamaguchi, K.; Kurita, J. J. Organomet. Chem. 2005, 690, 2956-2966. 127. Crociani, B.; Antonaroli, S.; Canovese, L.; Uguagliati, P.; Visentin, F. Eur. J. Inorg. Chem. 2004, 732-742. 128. Antonella, R.; Lo Sterzo, C. J. Organomet. Chem. 2002, 653, 177-194. 129. Nilsson, P.; Puxty, G.; Wendt, O. F. Organometallics 2006, 25, 1285-1292. 130. Clarke, M. L.; Heydt, M. Organometallics 2005, 24, 6475-6478. 131. ?lvarez, R.; Faza, O. N.; de Lera, ?. R.; C?rdenas, D. J. Adv. Synth. Catal. 2007, 349, 887-906. 132. ?lvarez, R.; Perez, M.; Faza, O. N.; de Lera, ?. R. Organometallics 2008, 27, 3378-3389. 133. Ariafard, A.; Lin, Z.; Fairlamb, I. J. S. Organometallics 2006, 25, 5788-5794. 136 134. Napolitano, E.; Farina, V.; Persico, M. Organometallics 2003, 22, 4030-4037. 135. Nova, A.; Ujaque, G.; Maseras, F.; Lled?s, A.; Espinet, P. J. Am. Chem. Soc. 2006, 128, 14571-14578. 136. Miyaura, N. J. Organomet. Chem. 2002, 653, 54-57. 137. Ridgway, B. H.; Woerpel, K. A. J. Org. Chem. 1998, 63, 458-460. 138. Matos, K.; Soderquist, J. A. J. Org. Chem. 1998, 63, 461-470. 139. Braga, A. A. C.; Morgon, N. H.; Ujaque, G.; Lled?s, A.; Maseras, F. J. Organomet. Chem. 2006, 691, 4459-4466. 140. Sicre, C.; Braga, A. A. C.; Maseras, F.; Cid, M. M. Tetrahedron 2008, 64, 7437-7443. 141. Braga, A. A. C.; Ujaque, G.; Maseras, F. Organometallics 2006, 25, 3647- 3658. 142. Braga, A. A. C.; Morgon, N. H.; Ujaque, G.; Maseras, F. J. Am. Chem. Soc. 2005, 127, 9298-9307. 143. Goossen, L. J.; Koley, D.; Hermann, H. L.; Thiel, W. Organometallics 2006, 25, 54-67. 144. Sumimoto, M.; Iwane, N.; Takahama, T.; Sakaki, S. J. Am. Chem. Soc. 2004, 126, 10457-10471. 145. Hatanaka, Y.; Hiyama, T. J. Org. Chem. 1988, 53, 918-920. 146. Hatanaka, Y.; Hiyama, T. J. Am. Chem. Soc. 1990, 112, 793-794. 147. Hatanaka, Y.; Hiyama, T. Synlett 1991, 845-853. 148. Hiyama, T.; Hatanaka, Y. Pure Appl. Chem. 1994, 66, 1471-1478. 149. Sugiyama, A.; Ohnishi, Y.; Nakaoka, M.; Nakao, Y.; Sato, H.; Sakaki, S.; Nakao, Y.; Hiyama, T. J. Am. Chem. Soc. 2008, 130, 12975-12985. 150. Denmark, S. E.; Sweis, R. F.; Wehrli, D. J. Am. Chem. Soc. 2004, 126, 4865- 4875. 151. Denmark, S. E.; Sweis, R. F. J. Am. Chem. Soc. 2004, 126, 4876-4882. 152. Mateo, C.; Ferna?ndez-Rivas, C.; Ca?rdenas, D. J.; Echavarren, A. M. Organometallics 1998, 17, 3661-3669. 137 153. Cotter, W. D.; Barbour, L.; McNamara, K. L.; Hechter, R.; Lachicotte, R. J. J. Am. Chem. Soc. 1998, 120, 11016-11017. 154. Suzaki, Y.; Yagyu, T.; Osakada, K. J. Organomet. Chem. 2007, 692, 326- 342. 155. Pantcheva, I.; Osakada, K. Organometallics 2006, 25, 1735-1741. 156. Hatanaka, Y.; Goda, K.; Okahara, Y.; Hiyama, T. Tetrahedron 1994, 50, 8301-8316. 157. Nunes, C. M.; Monteiro, A. L. J. Braz. Chem. Soc. 2007, 18, 1443-1447. 158. Lando, V. R.; Monteiro, A. L. Org. Lett. 2003, 5, 2891-2894. 159. Zim, D.; Lando, V. R.; Dupont, J.; Monteiro, A. L. Org. Lett. 2001, 3, 3049- 3051. 160. Liang, L.-C.; Chien, P.-S.; Huang, M.-H. Organometallics 2005, 24, 353-357. 161. Weissman, H.; Milstein, D. Chem. Commun. 1999, 1901-1902. 162. Moriya, T.; Miyaura, N.; Suzuki, A. Synlett 1994, 149-151. 163. Yamamoto, Y.; Takada, S.; Miyaura, N. Organometallics 2009, 28, 152-160. 164. Nishikata, T.; Yamamoto, Y.; Miyaura, N. Organometallics 2004, 23, 4317- 4324. 165. Farina, V.; Krishnan, B.; Marshall, D. R.; Roth, G. P. J. Org. Chem. 1993, 58, 5434-5444. 166. Amatore, C.; Gamez, S.; Jutand, A.; Meyer, G.; Moreno-Ma?as, M.; Morral, L.; Pleixats, R. Chem. Eur. J. 2000, 6, 3372-3376. 167. Amatore, C.; Jutand, A.; Meyer, G.; Mottier, L. Chem. Eur. J. 1999, 5, 466- 473. 168. Sharma, R. K.; Fry, J. L. J. Org. Chem. 1983, 48, 2112-2114. 169. Klanberg, F.; Muetterties, E. L. Inorg. Chem. 1968, 7, 155-160. 170. Marat, R. K.; Janzen, A. F. Can. J. Chem. 1977, 55, 3845-3849. 171. Damrauer, R.; Danahey, S. E. Organometallics 1986, 5, 1490-1494. 172. Damrauer, R.; O?Connell, B.; Danahey, S. E.; Simon, R. Organometallics 1989, 8, 1167-1171. 138 173. Johnson, S. E.; Day, R. O.; Holmes, R. R. Inorg. Chem. 1989, 28, 3182-3189. 174. Johnson, S. E.; Payne, J. S.; Day, R. O.; Holmes, J. M.; Holmes, R. R. Inorg. Chem. 1989, 28, 3190-3198. 175. Tamao, K.; Hayashi, T.; Ito, Y. Organometallics 1992, 11, 182-191. 176. Handy, C. J.; Lam, Y.; DeShong, P. J. Org. Chem. 2000, 65, 3542-3543. 177. Kost, D.; Kalikhman, I. In The Chemistry of Organic Silicon Compounds; Rappoport, Z., Apeloig, Y., Eds.; John Wiley and Sons Ltd.: Chichester, 1998, Vol. 2, pp 1339-1445. 178. Hansch, C.; Leo, A.; Taft, R. W. Chem. Rev. 1991, 97, 165-195. 179. Kurosawa, H.; Emoto, M.; Urabe, A.; Miki, K.; Kasai, N. J. Am. Chem. Soc. 1985, 107, 8253-8254. 180. Kurosawa, H.; Emoto, M.; Ohnishi, H.; Miki, K.; Kasai, N.; Tatsumi, K.; Nakamura, A. J. Am. Chem. Soc. 1987, 109, 6333-6340. 181. Kurosawa, H.; Kajimaru, H.; Miyoshi, M.; Ohnishi, H.; Ikeda, I. J. Mol. Catal. 1992, 74, 481-488. 182. Hartwig, J. F. Inorg. Chem. 2007, 46, 1936-1947. 183. Hartwig, J. F. Acc. Chem. Res. 1998, 31, 852-860. 184. Mann, G.; Baranano, D.; Hartwig, J. F.; Rheingold, A. L.; Guzei, I. A. J. Am. Chem. Soc. 1998, 120, 9205-9219. 185. Driver, M. S.; Hartwig, J. F. J. Am. Chem. Soc. 1997, 119, 8232-8245. 186. Culkin, D. A.; Hartwig, J. F. Organometallics 2004, 23, 3398-3416. 187. Tolman, C. A. Chem. Rev. 1977, 77, 313-348. 188. Wolfe, J. P.; Tomori, H.; Sadighi, J. P.; Yin, J.; Buchwald, S. L. J. Org. Chem. 2000, 65, 1158-1174. 189. Wolfe, J. P.; Singer, R. A.; Yang, B. H.; Buchwald, S. L. J. Am. Chem. Soc. 1999, 121, 9550-9561. 190. Johnson, J. B.; Rovis. T. Angew. Chem. Int. Ed. 2008, 47, 840-871. 191. Fairlamb, I. J. S. Org. Biomol. Chem. 2008, 6, 3645-3656. 192. Schwalbe, M.; Walther, D.; Schreer, H.; Langer, J.; Gorls, H. J. Organomet. Chem. 2006, 691, 4868-4873. 139 193. Green, M.; Howard, J. A. K.; Spencer, J. L.; Stone, F. G. A. J. Chem. Soc., Dalton Trans. 1977, 271-277. 194. Fairlamb, I. J. S.; Kapdi, A. R.; Lee, A. F. Org. Lett. 2004, 6, 4435-4438. 195. Sehnal, P.; Taghzouti, H.; Fairlamb, I. J. S.; Jutand. A.; Lee, A. F.; Whitwood, A. C. Organometallics 2009, 28, 824-829. 196. Sprengers, J. W.; Wassenaar, J.; Clement, N. D.; Cavell, K. J.; Elsevier, C. J. Angew. Chem. Int. Ed. 2005, 44, 2026-2029. 197. Clement, N. D.; Cavell, K. J.; Ooi, L. Organometallics 2006, 25, 4155-4165. 198. Selvakumar, K.; Zapf, A.; Spannenberg, A.; Beller, M. Chem. Eur. J. 2002, 8, 3901-3906. 199. Cort?s, J.; Moreno-Ma?as, M.; Pleixats, R. Eur. J. Inorg. Chem. 2000, 239- 243. 200. Moreno-Ma?as, M.; Pleixats, R.; Sebasti?n, R. M.; Vallribera, A. R. J. Organomet. Chem. 2004, 689, 3669-3684. 201. van Asselt, R.; Elsevier, C. J. Tetrahedron 1994, 50, 323-334. (PhZn, PhMg) 202. van Asselt, R.; Elsevier, C. J.; Smeets, W. J. J.; Spek, A. L. Inorg. Chem. 1994, 33, 1521-1531. 203. Klein, R.; Witte, P.; van Belzen, R.; Fraanje, J.; Goubitz, K.; Numan, M.; Schenk, H.; Ernsting, J. M. Elsevier, C. J. Eur. J. Inorg. Chem. 1998, 319- 330. 204. Grasa, G. A.; Hillier, A. C.; Nolan, S. P. Org. Lett. 2001, 3, 1077-1080. 205. Kluwer, A. M.; Elsevier, C. J.; B?hl, M.; Lutz, M.; Spek, A. L. Angew. Chem. Int. Ed. 2003, 42, 3501-3504. 206. Krause, J.; Haack, K.-J.; Cestaric, G.; Goddard, R.; Porschke, K.-R. J. Am. Chem. Soc. 1999, 121, 9807-9823. 207. Grundl, M. A.; Kennedy-Smith, J. J.; Trauner, D. Organometallics 2005, 24, 2831-2833. 208. Scrivanti, A.; Beghetto, V.; Matteoli, U.; Antonaroli, S.; Marini, A.; Crociani, B. Tetrahedron 2005, 61, 9752-9758. 209. Rencken, I.; Orchard, S. W.; Inorg. Chem. 1986, 25, 1972-1976. 210. Muhs, M. A.; Weiss, F. T. J. Am. Chem. Soc. 1962, 84, 4697-4705. 140 211. Genet, J. P.; Juge, S.; Achi, S.; Mallart, S.; Ruiz Montes, J.; Levif, G. Tetrahedron 1988, 44, 5263-5275. 212. Manoso, A. S.; Ahn, C.; Soheili, A.; Handy, C. J.; Correia, R.; Seganish, W. M.; DeShong, P. J. Org. Chem. 2004, 69, 8305-8314. 213. Masuda, Y.; Murata, M.; Ishikura, M.; Nagata, M.; Watanbe, S. Org. Lett. 2002, 4, 1843-1845. 214. Mowery, M. E.; DeShong, P. J. Org. Chem. 1999, 64, 1684-1688. 215. Tseng, C. C.; Paisley, S. D.; Goering, H. L. J. Org. Chem. 1986, 51, 2884- 2891.