‘..-~ ‘1‘" ‘......u-- RARIES W mas..- W“ W\\\\\\\\\\\\‘\ H 31293 i i This is to certify that the dissertation entitled SYNTHETIC AND MEGiANISTIC ASPECTS OF ZIEGLER-NA'ITA POLYMERIZATION: REGIO- AND DIASTEREOSELECI' IVE FORMATION OF FIVE- AND SIX-MEMBERED CARBOCYCLES presented by Jonathan R. Young has been accepted towards fulfillment of the requirements for Ph. D. Chemistry degree in Dr. J. R. Stille Major professor Date 6/29/92 MSU is an Affirmative Aaron/"Equal Opportunity Institution 0-12771 .——— ._____ _—» ”hi—F LIESRAEW Ratetgan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before duo duo. DATE DUE DATE DUE DATE DUE --i T L__\ 3 D U [,1 [—_—l F J i J usu lo An Affirmative AotioNEqual Opportunity Institution C1 7:“‘7—"7 EWMHT SYNTHETIC AND MECHANISTIC ASPECTS OF ZlEGLER-NATTA POLYMERIZATION: REGIO- AND DIASTEREOSELECTIVE FORMATION OF F IVE- AND SIX-MEMBERED CARBOCYCLES By Jonathan Russell Young A DISSERTATION Submitted to _ Michigan State Universuty in partial fulfillment of the requurements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1992 {7% {W7 ABSTRACT SYNTHETIC AND MECHANISTIC ASPECTS OF ZIEGLER-NA'ITA POLYMERIZATION: REGIO- AND DIASTEREOSELECTIVE FORMATION OF FIVE- AND SIX- MEMBERED CARBOCYCLES By Jonathan Russell Young An ongoing project in our group involves the development and application of novel ring-forming reactions that feature intramolecular addition of reactive carbon intermediates to unactivated olefins. Although ring-forming methods utilizing this synthetic design are known, none of these methods were able to combine high regio- and stereoselective qualities that spanned applicability to synthesis of both five- and six-membered rings. Our objective was to develop a general, complimentary ring-tanning method that exhibited high selectivities for a variety of ring sizes. We focused our attention on the use of organometallic intermediates as a means of directing both the regio- and stereochemistry. The organometallic complexes chosen were those of alkyltitanocene chlorides, which are well known for undergoing efficient and stereoselective tat-olefin polymerization. Our initial goal was to ascertain the regioselectivity of olefin insertion into the carbon-titanium bond for the unsubstituted 5—hexen-1-yl- and 6-hepten—1- yltitanocene chlorides. These alkyltitanocene derivatives were stable for weeks and exhibited no tendency toward olefin insertion. However, activation with EtAlCIz drastically altered the electronic environment surrounding the metal and the reactivity of these complexes towards olefin insertion. As a consequence of the Lewis acid additive, facile olefin insertion resulted with exclusive exo specificity in each case, providing methylcyclopentane (5-exo-trig, 95% yield) and methylcyclohexane (6-exo-trig, 75% yield) after hydrolysis. We next turned our attention to the stereochemical consequence of the carbon-carbon bond formation with respect to an alkyl substituent on the tether. In the case of cyclopentane formation, selectivity ranged from 97:3 to >99:1 for 2-, 3-, and 4-methyl-5-hexen-1-yltitanocene chlorides. The stereocontrol exhibited for the six-exo-trig cyclizations were more complex, due to the greater flexibility of the alkyl tether. Cyclization of 4- and 5-methyI-6-hepten-1-yltitanocene chlorides resulted in >98:2 diastereomeric ratio of dimethylcyclohexanes. However, when a methyl group was [3 or y to the metal, cyclization was less successful (82:18, and 50:50 trans:cis, respectively). Yet with an isopropyl group in these positions, enhanced olefin facial selectivity was observed (92:8, transzcis and 74:26, cisztrans, respectively). In the course of deveIOping this synthetic methodology, we noticed that olefin insertion was more facile for a certain substitution pattern. The substrates that appeared to be more reactive had a common structural feature, an alkyl substituent on the carbon [3 to titanium. We have developed a series of competitive rate experiments to determine if a- or B—hydrogen activation plays a role in the cyclization process. At this stage of the project, optimization of the experimental conditions is in progress. To Mom, Dad, Nancy, and Karl Mom and Dad, thank you for always being there when I needed you. You two have been terrific parents and have always been my biggest supporters. ' From every hockey, baseball, or football games, to drama and band concerts, you were always there. I can never adequately express what your love and support has meant to me. Nancy, thank you for your patience and allowing me to fulfill my dreams during the past years. I know these years have been difficult for us, but let's be happy we survived them, together. I am sorry that we didn't always have as much time with each other as we wanted, but remember, brighter days and opportunities await us in Madison. Let's have some fun as a family. Karl, thank you for all the happiness you constantly bring into our lives. I want you to never be afraid to chase your dreams. ACKNOWLEDGMENTS I would first of all like to thank the anonymous person in the Natural Resources office that studied my application so thoroughly during the months of February to late October of 1987. I made it here anyway! I will be forever indebted to Dr. Stille (that is John, now) for all that he taught me. I will always miss our long conversations in the morning over a cup of ‘ coffee, weightlifting and running at the crack of dawn, and the enthusiasm that we shared together in pursuit of common goals. Chemistry aside, your friendship is something very special to me. Thanks for all that you have done and given to me. Please stay in touch in the years to come. I want to wish the best to all of my friends in the Stille group. Greg and Art, good luck at Stanford and say hi to Barry for me. Polymervanan, thanks for the midnight brainstorming, it was a lot of fun. Words cannot express the wonderful friend and person I met at MSU. Nancy, you provided a wonderful spark to not only me, but to the whole group. All the help and support you gave me over the last two years meant so much to me. I clearly could not have survived this dissertation without your proofreading skills or your daily positive attitude and kind words of encouragement. Thank you for all you have done and most of all, for believing in me. My morning donut will certainly not be the same anymore but it will remind be of countless special mornings we shared. How will I cool my coffee without you? You are the best friend I could ever have asked for and I will miss you very much. TABLE OF CONTENTS LIST OF TABLES ..................................................................................................... ix LIST OF SCHEMES ................................................................................................ xi LIST OF FIGURES ................................................................................................... xiii LIST OF ABBREVIATIONS ..................................................................................... xiv INTRODUCTION 1. Outline of the Synthetic Problem ............................................................ 1 2. Background on the Synthetic Aspects of Current Cyclization Methodology .............................................................................................. 4 A. Carbomtionic Cyclizations ........................................................... 5 B. Free Radical Cyclizations .............................................................. 8 C. Lithium-Mediated Cyclizations ..................................................... 17 D. Magnesium-Mediated Cyclizations ............................................. 22 E. Aluminum-Mediated Cyclizations ................................................ 24 F. Palladium-Mediated Cyclizations ................................................ 25 3. Background for the Mechanism of Zregler-Natta Polymerization ..... 26 CHAPTER 1. SYNTHESIS OF CYCLIZATION PRECURSOR SUBSTRATES AND STANDARDS 1. Introduction ................................................................................................. 34 2. Synthesis of Five-Membered Ring Precursor Substrates ................. 35 3. Synthesis of Six-Membered Ring Precursor Substrates ................... 37 4. Synthesis of Competitive Rate Substrates ........................................... 4O 5. Conclusion ................................................................................................. 43 vi 6. Experimental .............................................................................................. 44 CHAPTER 2. DIASTEREOSELECTIVE CYCLOPENTANE FORMATION MEDIATED BY TITANOCENE DICHLORIDE 1. Introduction ................................................................................................. 75 2. General Features and Limitations for the Synthesis of Methyl- Substituted 5-Hexen-1-yltitanocene Chlorides ................................... 78 3. Synthesis and Cyclization of Methyl-Substituted 5-Hexen-1- yltitanocene Chlorides ............................................................................. 80 4. Comparison of the Stereoselective Outcome of Various Methods. 82 5. Optimization and Cyclization of 133: Formation of Bicyclo[4.3.0]nonane ................................................................................ 84 6. Conclusion ................................................................................................. 86 7. Experimental .............................................................................................. 87 CHAPTER 3. DIASTEREOSELECTIVE CYCLOHEXANE FORMATION MEDIATED BY TITANOCENE DICHLORIDE 1. Introduction ................................................................................................. 91 2. Optimization Studies for the Cyclization of 6-Hepten-1-yltitanocene Chlorides .................................................................................................... 92 3. Potential Explanations for the Observed Stereoselectivities ............ 102 4. Conchsion ................................................................................................. 104 5. Experimental .............................................................................................. 106 CHAPTER 4. MECHANISTIC STUDIES OF ZlEGLER-NATI'A POLYMERIZATION: PROBING FOR a- OR B- HYDROGEN ACTIVATION 1. Introduction ................................................................................................. 113 2. Preliminary findings for the 2-Alkyl-5-hexen-1-yltitanocene Chlorides .................................................................................................... 116 3. Optimization Studies for the Competitive Cyclization of 2-AlkyI-6- hepten-l-yltitanocene Chlorides with MezAlCl, Mng, and MAO.... 118 4. Conclusion ................................................................................................. 133 vii 5. Experimental .............................................................................................. 134 CONCLUSION ...................................................................................................... 138 LIST OF REFERENCES .................................................................................... 139 viii Table 1. Table 2. Table 3. Table 4. Table 5. Table 6. Table 7. Table 8. Table 9. Table 10. Table 1 1. Table 12. Table 13. Table 14. Table 15. Table 16. LIST OF TABLES Regioselectivity of Atom Transfer Reaction ................................... 14 Diastereomer Ratios of Dimethylcyclopentane Products via Trtanium and Free Radical Methods ......................................... 79 Comparison of Diastereoselectivities for Different Methods ........... 84 Optimization of the Cyclization Conditions for 133b ....................... 85 Diastereomer Ratios of Cyclohexane Products via Titanium and Free RadicalMethods ............................................... 94 Optimization of the Cyclization Conditions for 155b ....................... 96 Optimization of the Cyclization Conditions for 164b ....................... 97 Optimization of the Cyclization Conditions for 160b ....................... 98 Optimization of the Cyclization Conditions for 137b ....................... 99 Diastereomer Ratios of Cyclohexane Products Performed Under the Optimized Conditions .................................................... 101 Cyclization of Unpurified 175b Using Mng or MAC ...................... 119 Cyclization of Purified 168b and 175b Using M9X2 ....................... 121 Cyclization of Alkyltitanocene Chlorides Promoted with MezAICI .......................................................................................... 123 Cyclization of Alkyltitanocene Chlorides Promoted with MAO ............................................................................................... 124 Competitive Cyclization of 8-Hydrogen vs 8-Hydrogen (168b vs 175b) Promoted Under the Optimized Conditions for MAO ....................................................... 127 Competitive Cyclization of 8-Hydrogen vs B-Deuteron (168b vs 178b) Promoted Underthe Optimized Conditions for MAC ....................................................... 128 ix Table 1 7. Competitive Cyclization of BDeuteron vs B-Hydrogen (171b vs 175b) and or,or-Dideuteron vs B—Hydrogen (184b vs 175b) Promoted Under the Optimized Conditions for MAO ....................................................... 129 Scheme 1. Scheme 2. Scheme 3 Scheme 4. Scheme 5. Scheme 6. Scheme 7. Scheme 8. Scheme 9. Scheme 10. Scheme 1 1. Scheme 12. Scheme 13. Scheme 14. Scheme 15. Scheme 16. Scheme 17. Scheme 18. Scheme 19. Scheme 20. LIST OF SCHEMES Regioselectivity of Carbocationic Cyclizations ........................... 6 Stereoselectivity of Carbocationic Cyclizations .......................... 6 Regioselectivity of Free Radical Cyclizations ............................. 10 Stereoselectivity of Free Radical Cyclizations ........................... 11 A Possible Propagation Sequence for the Atom Transfer Reaction ....................................................................... 12 Evidence for Proton Abstraction ................................................. 18 Regioselectivity of Alkyllithium Cyclizations ............................... 20 Stereoselectivity of Alkyllithium Cyclizations .............................. 20 Stereoselectivity of Vinyllithium Cyclizations .............................. 21 Stereoselectivity of Aluminum Cyclizations ................................ 24 The Cossee-Arlmann Mechanism for Chain Propagation .......... 28 The Green-Rooney Mechanism for Chain Prepagation ............. 29 Grubbs' Test for or-Hydrogen Activation ..................................... 32 Bercaw's Test for (l-HYdfOQSII Activation ................................... 33 Retrosynthetic Analysis of the T‘rtanocene Dichloride-Mediated Carbocyclization ......................................................................... 34 Synthetic Route to 125 ............................................................... 35 Synthetic Route to 129 ............................................................... 36 Synthetic Route to 133 ............................................................... 37 Synthetic Route to 137 ............................................................... 38 Synthetic Route to 144 and 151 ................................................. 39 xi Scheme 21. Scheme 22. Scheme 23. Scheme 24. SCherne 25. Scheme 26. Scheme 27. SCheme 28. Scheme 29. Scheme 30. Scheme 31. Scheme 32. Scheme 33. Synthetic Route to 155 and 160 ................................................. 39 Synthetic Route to 164 ............................................................... 40 Synthetic Route to Competitive Rate Substrates ....................... 42 Synthetic Route to Cyclohexane Standards ............................... 43 Synthetic Route to 196 and 197 ................................................. 43 Preparation of 5-Hexen-1-yltitanocene Chloride (198) ............... 76 Potential Pathways of Cyclization and Decomposition of198 ......................................................................................... 76 Proposed Transition States for Five-Membered Ring Formation ................................................................................... 82 Comparison of Titanium-Mediated Cyclization to Free Radical Cyclization for 133a ....................................................... 86 Possible Transition States for Six-Membered Ring Formation ................................................................................... 104 Mechanistic Possibilities for Ziegler-Natta Polymerization: or- and B—Hydrogen Interactions ........................ 115 Competitive Rate Experiments Designed to Test the Mechanistic Possibilities ............................................................ 116 Cyclization of 5-Hexen-1-yltitanocene Chlorides Performed Under the Transmetallation Conditions .................... 117 xii LIST OF FIGURES Figure 1. a-Agostic Interaction for 107 ............................................................ 30 Figure 2. B-Agostic Interaction for 108 ............................................................ 30 Figure 3. B-Agostic Interaction for 109 ............................................................ 30 Figure 4. Competitive Cyclization Between 168b and 175b ........................... 131 Figure 5. Competitive Cyclization Between 168b and 178b ........................... 131 Figure 6. Competitive Cyclization Between 171b and 175b ........................... 132 Figure 7. Competitive Cyclization Between 184b and 175b ........................... 132 xiii AIBN Bu n-BuLi tBuU CeHa CP mCPBA DiBAI DCE DIPA DMAP DMF DMSO Et EtzO LAD LAH LDA MAO Me LIST OF ABBREVIATIONS Azobisisobutyronitrile Butyl n-Butyllithium fen-Butyllithium Benzene Cyclopentadienyl meta-Chloroperbenzoic acid Deuteron (2H) Diisobutylaluminum hydride 1 ,2-Dichloroethane Diisopropylamine 4-Dimethylaminopyridine N,N-Dimethylformamide Methyl sulfoxide Ethyl Diethyl ether Lithium aluminum deuteride Lithium aluminum hydride Lithium diisopropylamide Methaluminoxane (O-Al(Me))n Methyl xiv Ms NBS Ni[COD]2 Ph PMDTA Pr iPr TBAF TBDMS TEA THF TMEDA SET Methanesulfonyl N-Bromosuccinimide Nickel cyclooctadiene Phenyl N, N, N', N", N"-Pentamethyldiethylenetriamine Propyl lsopropyl Tetrabutylammonium fluoride fert-Butyldimethylsilyl Triethylamine Tetrahydrofuran N,N,N',N'-Tetramethylethylenediamine Single Electron Transfer INTRODUCTION 1 . Outline of the Synthetic Problem Carbocycles of the quinane and steroid families have presented chemists with a number of interesting synthetic challenges. The efforts exerted toward the preparation of these medicinally useful compounds have benefited chemists and non-chemists alike. Upon examination of the complex architectural framework of the quinanes and steroids, a common theme woven throughout these families of compounds was the repetition of five- and six- membered ring subunits, fused together in a variety of regio- and stereochemical patterns. To effectively launch an assault on quinane or steroid targets, a strategic plan must be devised for the efficient construction of the carbocyclic skeleton. This plan should address, not only the preparation of five- and six-membered rings but to also incorporate the numerous stereoconters as well. An inspired plan would be one that could achieve both of these goals simultaneously, thus minimizing the number of manipulations. Although many valuable ring-forming methods exist for achieving any one of these goals, a single, general method applicable to five- and six-membered rings with good control over regio- and stereochemistry would be useful. Our interest in preparing compounds of the above variety has compelled us to develop a general, yet selective, ring-forming method capable of fulfilling the aforementioned criteria. Our efforts have concentrated on 2 developing organotransition metal reagents to direct selective, intramolecular carbon-carbon bond construction.‘ The interest in organometallic methods has been building ever since Grignard discovered the landmark reaction between alkyl halides and magnesium. The crucial observation was the novel change in the polarity of carbon (umpolung), and the useful reversal in the chemical reactivity of carbon. Grignard's efforts blossomed into an exciting new approach to carbon-carbon bond formation. Early in the development of organometallic chemistry, chemists were laying the groundwork, reacting various metals together with different organic substrates in an apparent random fashion. What resulted from this ”hit or miss" approach was a firm foundation for understanding the interaction and subsequent reaction of organic functional groups with metal complexes.2 As a result, very predictable methods for engineering carbon-carbon bond construction utilizing organometallic reagents have evolved. The sophistication of of organometallic reagents for carbon-carbon bond formation lies in the ability to fine tune the reactivity, and thus selectivity, of the metal complex by varying the steric and electronic environment surrounding the metal. This can be accomplished by judicious choice of the metal, its oxidation state, and the surrounding ligands. One of the most useful and exploited organometallic processes is the reaction of alkyl-metal complexes with olefins. Certain electron deficient metal complexes, which have a vacant coordination site available, can form a 1:- complex with olefins and can often be induced to insert the olefin into the metal- carbon bond. In many cases, the carbon—carbon bond could be formed in a highly regio- and stereoselective manner. A good example is that of Ziegler- Natta polymerization.3 3 In the early 1950's, Ziegler4 and Narta5 independently discovered that titanium salts, in combination with alkyl aluminums, facilitated the polymerization of ethylene. The active polymerization catalyst was suggested to be an alkyltitanium species, formed by alkyl ligand exchange from aluminum to titanium. This proposal was later substantiated after the combined action of ethyltitanocene chloride6 and ethylaluminum dichloride was shown to polymerize ethylene.7 Several reports have indicated that Ziegler-Natta chemistry should be well suited to ring forming reactions. In 1958, Marvel reported that 1,5- hexadiene was polymerized with TrCl4 and an alkyl aluminum co-catalyst (eqn. 1).821 The polymer contained repeating units of methylenecyclopentanes, connected in a 1,3 relationship. Similarly, 1,6-heptadiene resulted in a polymer of methylenecyclohexanes under the same conditions (eqn. 2). Thus, each intermolecular olefin insertion was followed by an intramolecular insertion resulting in carbocycle formation. At this point, the stereochemical details of this reaction have not been determined. More recently, Waymouth has shown that a zirconocene catalyst, together with methaluminoxane (MAO), polymerized 1,5- hexadiene in the same fashion as in eqn. 1, yet polymerization of 1,6- heptadiene yielded only low molecular weight oligomers.8b There have been a number of paramount reports in which monomer formation was realized as well. It has been demonstrated that 5-hexen-1-ylmetal complexes (Metal a Ti, Zr, Sc, Y)9 could be induced to isomerize to methylcyclopentylmetal products. Specifically, Grubbs described that EtAIClg successfully catalyzed the intramolecular olefin insertion of 5-hexen-1-yltitanocene chloride to provide, after hydrolysis, high yields of methylcyclopentane (Scheme 13).93 In a similar fashion, Schwartz observed that when transmetallation of 5 to aluminum was attempted, methylcyclopentane was produced (eqn. 3).9b These results provide 4 convincing evidence to the viability of cyclopentane and cyclohexane formation via organometallic intermediates, yet leave many questions unanswered concerning the stereochemical details of the newly formed asymmetric center. T'Cl4('B )3Al Me R W ' ~ . . . e (1) 1 Rs Hor iBu M “C“ “BMW.” (2) 3 n-R Hor iBu Me 2) AICIa AN 1 ) szszIC') _W ZGCch fl. (3) 1 5 6 2. Background on the Synthetic Aspects of Current Cyclization Methodology A number of ring-forming methods have been creatively designed in which a reactive carbon intermediate added across an unactivated olefin. The preparative utility of intramolecular addition of carbon centered radicals, anions, and cations, as well as covalently bound transition metals, to distal double bonds has been demonstrated. However, each of these methods had certain inherent limitations that diminished their applicability to general situations. An overview of various ring forming methods will be provided, addressing such issues as regioselectivity, stereoselectivity, generality to different ring sizes, and the ability to functionalize the final product. This discussion will begin by examining both carbocationic and anionic, as well as free radical approaches to carbocyclization. Lastly, organometallic carbocyclizations, in which themetal 5 was covalently bound to carbon, will be described for metals such as magnesium, aluminum, and palladium. A. Carbocationic Cyclizations The field of carbocationic cyclization was very active during 1960-75 because of its resemblance to nature's way of producing complex ring systems, such as steroids.10 Both the regio- and stereochemistry were well defined and predictable. Generally, cyclization proceeded in an endo manner to form the larger ring, and a more stable carbocation (Scheme 1).10 For example, cyclization of 7 formed a cyclohexane (9) rather than methylcyclopentane. However, five-membered rings have been formed when the substrate contained an appropriately placed substituent that could direct exo cyclization by producing a more stable carbocation. Thus, cyclization of 10 produced the product of exo addition (11), possibly due to the steric restriction of the endo pathway. The stereochemistry was found to be dependent on the stereochemistry of the olefins (Scheme 2).10 Thus, a polyolefinated substrate that possessed trans olefins situated in a 1,5 relationship (see 13) were found to undergo cyclization to produce trans fused six-membered rings (14). Similarily, cyclization of substrates with the corresponding cis arrangement (see 15). resulted cis ring junctions (16). 6 Scheme 1. Regioselectivity of Carbocationic Cycllzations (+) O endo addition 0 B—elimination I » +) :7 7 8 9 M0 M0 M9 (+)MO M9 MO I? ’) exo addition 6 B-elimination “° 1 O 1 1 1 2 Scheme 2. Stereoselectlvlty of Carbocationic Cycllzatlons H (+) H ‘ \ H I 13 14 H H 16M Two other important design elements were crucial to cationic cyclization. Specifically, these two elements were a carbocation initiator and terminator. Treatment of a polyolefinated compound with a protic acid was found to be a poor method for selective carbocation formation due to the propensity of the acid to protonate indiscriminately, and to isomerize olefins.1o However, the advent of Lewis acids in combination with either an acetal, epoxide, or an allylic alcohol, led to clean formation of the desired carbocation. The main function of the terminator was to direct the size of the final ring formed. Common terminators that have been used were alkyl, allyl, benzyl, or 8-silyl groups. Alkyl groups proved useful for directing the size of the final ring, however, unselective B—elimination resulted in a number of alkene isomers (see 12, Scheme 1). Allyl 7 and benzyl terminators alleviated this problem, but difficulty encountered in attempting to remove these unnatural steroid entities overshadowed this small success. Silyl groups were able to overcome both of these difficulties, owing to their well-known propensity to stabilize B-carbocations and to undergo B-silyl elimination. Thus, the strategic location of the silyl group served to selectively direct ring closure to form a B-carbocation, followed by rapid B-silyl elimination providing a single structural isomer. Me Me Me SnCl4 END"). -80 °C 70% Yleld ’ OH 17 . Johnson and co-workers were the first to apply the above principles for the complete formation of the steroid skeleton in a single tandem cyclization (eqn. 4).") Thus, treatment of allylic alcohol 17 with tin chloride resulted in the formation of three new rings in better than 70% yield (18). More importantly, five new chiral centers were produced that possessed the correct relative configuration of 16,17-dehydroprogesterone. More recently, Burke has used this methodology, equipped with a dioxane initiator and a vinyl silane terminator, as the key step in the synthesis of the decalin ring systems for mevinic acid (eqn. 5),“3 nagilactones,11b (+)-dihydrocompactin and (+)- compactin,11c (+)-fragolide,11d and (-)-perinporin B."d Another attractive feature of carbocationic cyclization was the ability to affect enantioselective bond formation.12 The simplicity with which chirality was introduced into the cyclization precursor was a major advantage. Thus, an 8 achiral alkenal was made chiral by trivial transformation to an acetal using a chiral diol (21b, eqn. 6). Cyclization of 21b resulted in the normal ratio of diastereomeric isomers (see cyclization of 21a) but with an excellent enantiomeric excess. The opposite absolute configuration was also available by using the other enantiomer of the diol. m R HO(CH2)3(_) n = H O O 5 M0 Me Ter. -78 °C _ v , (5) 74% YIBId 5 ii TMS 1 9 2 0 Me 4. > z : o ; fr 0),, Hoc,H,o HOCzH‘O a ‘3" 21a; R -H 228;90(1:1) 23a;10(1:1) 21b; R . CH3 . 22b; 90 (93:7) 23: 10 (94:6) 8. Free Radical Cyclizations Free radical cyclization has enjoyed extensive success in the application to natural product synthesis.13 Far and away, the best attributes of this method were the mild conditions, functional group tolerance, and wide array of radical precursors. For example, remote functional groups such as alcohols, amines, ketones, esters, ethers and chlorides were stable to the conditions. Precursors to carbon radicals that have been exploited include bromides, iodides, alkynes, 9 alkyl and aryl sulfides, alkyl and aryl selenides, and nitro groups. In terms of stereoselectivity, the most successful examples of free radical cyclization have usually contained a structural feature, such as a ring, that restricted the freedom of rotation of the alkene or the tether. This limitation was a result of the minimal alkene facial selectivity, or lack of conformational control, exhibited by an alkyl radical. Curran and Fraser-Reid have utilized this approach to solve quinane structural problems. Curran has prepared a number of angular,”a linear,14b and propellane14c triquinane natural products using a tandem cyclization approach, as illustrated for an intermediate to crinnipelin (eqn. 7).143 In this case, Curran took advantage of the kinetic propensity of the sequential alkyl radicals to react from the same face of the cyclopentane ring, thus controlling the relative stereochemistry of the ring fusion. Fraser-Reid has controlled the absolute stereochemistry by performing similar serial cyclizations (eqn. 8),153 in which the quinane precursor was secured to a chiral carbohydrate template. In this way, enantiomerically pure quinanes were produced. Me Buaan. AIBN Benzene, 80 °C> 24:; x . CHzl 51% Yleld 25:; v . H, H 24b; X-COSOPII 62% Yleld 25b; v .. o 1) 803$nH 2) Silica Gel 73% Yleld (8) In the simplest case, radical cyclization was effective for selective carbon- carbon bond formation (Scheme 3).13 For example, cyclization of 28 provided 10 a 98:2 ratio of exo:endo products. However, the regioselectivity of ring closure was dramatically reduced when a geminally disubstituted olefin (30) was employed. Thus, the important terpene element, the gem-dimethyl group, was not accessible via this route. Moreover, extention to six-membered rings via exo cyclization (33) has been a troublesome operation for entropic reasons, providing only 11% cyclization contaminated by a trace of the endo product as well.16 In this case, the lifetime of the radical was too short for the significantly slower cyclization to occur prior to hydrogen-atom abstraction. For the sake of clarity, the rates of 5-exo-trig cyclization, six-exo-trig cyclization, and hydrogen atom transfer are, (k=2x105, 25 °C),17 (k=5x103, 25 °C),‘7 (k=6x106, 25 °C),18 respectively, which help explain the different levels of success for the two reactions. Scheme 3. Regioselectivity of Free Radical Cyclizations War m"—”—~°" G O 23 D (98:2)29 M0 MD M ”arr-W, 6 O 9' car. ,30 31 (40:60) 32 craftsmen Hm~+ 0+. M.. N\/\/ 34(89) 32 (10) 35(1) Substrates which possessed an alkyl group on the tether suffered a lack of stereoselectivity for radical ring closure (Scheme 4).19 The diastereoselectivities for bromides 36a-c ranged from 2:1 to 4:1, with the major product arising by cyclization from the more stable chair conformation.20 10 a 98:2 ratio of exo:endo products. However. the regioselectivity of ring closure was dramatically reduced when a geminally disubstituted olefin (30) was employed. Thus, the important terpene element, the gem-dimethyl group, was not accessible via this route. Moreover, extention to six-membered rings via exo cyclization (33) has been a troublesome operation for entropic reasons, providing only 11% cyclization contaminated by a trace of the endo product as well.16 In this case, the lifetime of the radical was too short for the significantly slower cyclization to occur prior to hydrogen-atom abstraction. For the sake of clarity, the rates of 5-exo-trig cyclization, six-exo-trig cyclization, and hydrogen atom transfer are, (k=2x105, 25 °C),17 (k=5x103, 25 °C),17 (k=6x105, 25 °C),18 respectively, which help explain the different levels of success for the two reactions. Scheme 3. Regioselectivity of Free Radical Cyclizatlons A/2V\a.c,-.-.,—-> BEL-“B" b O 6 (98:2)29 Me Ms M, W “‘5‘, g .30 31 (40:60) 32 arBufirHAreN HmM+ 0’". M0. W3\/\/ 34 (89) 32 (10) 35 (1) Substrates which possessed an alkyl group on the father suffered a lack of stereoselectivity for radical ring closure (Scheme 4).19 The diastereoselectivities for bromides 36a-c ranged from 2:1 to 4:1, with the major product arising by cyclization from the more stable chair conformation.” 11 Scheme 4. Stereoselectlvlty of Free Radical Cyclizations R2 R‘ 51 Bu3$nH,AIBN "3 We. 06H6,95% Yleld M“ n2 W ....in2 > + at n3 . 0" R3 36 37 F" 38 B‘ 32 32 e. Me H H 17 as b: H Me H 75 25 c H H Me 35 65 In their current form, radical cyclizations have practical and environmental limitations as well. In order to promote intramolecular events over hydrogen abstraction or dimerization, concentrations of <0.01 M were required. Industrial application of this technique would be costly and cumbersome in terms of solvent requirements, recovery or disposal of solvent, and the handling and disposal of the toxic tin byproducts. The continued efforts of many researchers have increased the level of sophistication and applicability of free radical processes. As a result, ”yesterday's limitations” are becoming obsolete. One major problem of the tin- hydride method was that substituents surrounding the radical had little effect on the rate of hydrogen-atom abstraction. Consequently, most alkyl radicals had similar lifetimes with respect to hydrogen-atom abstraction. For relatively slow cyclizations, such as 6-exo-trig cylizations, hydrogen abstraction becomes competitive with cyclization. Although not recognized at the time, a major synthetic breakthrough happened in 1948 when Kharasch discovered what is called today the ”atom-transfer reaction".21 This reaction delivers a C-X (X=halogen) bond across a double or a triple bond. The attractive feature of this reaction was that it was devoid of a hydrogen atom donor. Consequently, longer lifetimes of the incipient radicals were permitted without interception by a hydrogen-atom donor, which in turn allowed for slower inter- or intramolecular 12 events to occur. Finally, predictable and efficient five-, six-, or seven-membered ring formation at a reasonable concentration (0.3 M) was at last a viable option through radical methodology.22 Scheme 5. A Possible Propagation Sequence for the Atom Transfer Reaction I TMS I COzMa BngSnSnBu, + Ir 0.3 M. cfil_la 002m TMS 55% Yield 3 9 4 3 l l F ms ms ms I - - K, I/ . 002Me 002m I .vrl _ 4 o 4 1 4 2 "‘5‘ Curran has developed a number of new strategies based on the atom transfer principle. For example, a formal [3+2] annulation procedure was designed that began by treating alkyliodide 39 with hexabutylditin and light in the presence of an activated olefin (Scheme 5).23 The initial radical formed (40) added to the olefin in a Michael fashion which produced 41, followed by an intramolecular 5-exo-dig cyclization. The resulting vinyl radical 42, which was less stable than the preceding ones, underwent iodine atom transfer from the starting iodide which produced 43 and completed the chain process. The key ingredients to the successful design of this procedure were the delicate balance between the relative strength of the carbon-iodide bond, and the reactivity of the carbon radicals along the way. An illustrative example was the 13 divergent behavior of 44 with 4-methyI-2-pentene (eqn. 9) compared to that shown in Scheme 5. In this case, a two-step sequence was necessitated. In the first step, the intermolecular atom transfer addition across the olefin could be thermally promoted in the absence of the tin reagent, owing to the more labile carbon-iodide bond, which provided a single regioisomer. This reaction produced no trace of cyclized products, only the acyclic secondary alkyl iodide. The greater reactivity of the 2° alkyl radical compared to the stabilized radical 41 coupled with the weaker carbon-iodide bond, resulted in premature iodide atom transfer. Treatment of the 2° iodide with BU3SI‘IH yielded a single regio- and stereoisomer, methylenecyclopentane 45. I Me Me Me 1) 80 °C- CGHB .-"“k + 2) Bu35nH . Me | -» <9) NC | 59% Yield NC no M0 NC M0 44 45 14 Table 1. Regioselectivity of Atom Transfer Reactions Products Substrate EXO ENDO Yleid (cis/francs) (cis/trans) 455; R . OM 47a (53/140) 483 (SI/4) 83 46b; R . Ph 478 (21/54) 48b(8/17) 68 o o o i a fi ,1 E l a R I 498; R s H 508 (<3) 513 (>97) 56 498; R . Me 508 (25) 518 L75) 73 M020 1 MeOzc r M6020 a G (5*; C1 I 5289 . H 532 (48/32) 543 (15/5) 68 5171: - Me 538(50/50) 548(0) 60 I E E 9’ n2 55 iii R‘ 5.0028 2 553(90 ) 57 (10) 86 553 R1 - H: R - H 8 55bR1-H;R2-Me £1300) “‘2’00 g: 55cH1-Me;n2-H ‘ 0°) 57“ I e ,. e R R‘ 5.6028 555 (60) we (40) 66 532 R1 - H- R2 - H 1 '. 2 59b (10) 60b (0) 62 58c 111- Me; RZ-H 15 The results in Table 1 established guidelines for application of atom transfer reactions to larger rings.22 Treatment of 463 under the atom transfer conditions resulted in excellent regiocontrol (93:7, 47a:48a) favoring 5-exo over 6-endo cyclization. A reversal in the mode of cyclization could be brought about by placing the carbonyl on the interior of the forming ring (see 49). in this case, endo cyclization was the dominant pathway (3:97, 50a:51a). Calculations have shown that the product of this cyclization, like all other intramolecular free radical reactions, was the kinetic product. The transition state for this cyclization resembled that of the [3,3] reaction. Cyclization of 52a resulted in more endo cyclization than for the cyclization of 46. This mode could be suppressed by incorporation of a terminal alkyl group on the olefin (52b). Curran has also developed atom-transfer cyclizations based on iodomalonates. Cyclization of 553 and 58a demonstrated that for a terminal alkene, irrespective of tether length, a mixture of exo and endo products were formed. Each of these modes could be controlled and amplified by situating an alkyl group on the interior (see 558 and 58b) or on the exterior (see 55c and 58c) of the olefin. Although these examples extend the applicability of radical cyclization, control of simple diastereoselectivity remains a hurdle to overcome. Boger25 has recently described a new modification for free radical cyclization. which featured an acyi radical. Acyl radicals were shown to effectively participate in tandem intramolecular reactions as well as intramolecular-intermolecular and intermolecular-intramolecular sequences. Treatment of 61 (eqn. 10) with the standard tin hydride conditions preceded with initial 6-endo-trig cyclization, followed by a rapid 6-exo-dig cyclization. The initial radical ring closure was consistent with Curran's results for a substrate that possessed a carbonyl on the interior of the forming ring. However, Boger maintained that endo cyclization was merely a result of the kinetic deceleration 16 of the 5-exo mode due to the flanking methyl group on the olefin. Although the regiocontrol was excellent in this case, the minimal stereoselectivity at the ring junction was a serious concern. Attempted epimerization of 62 resulted in only a slight increase for the trans fusion. The 5-exo-trig cyclization of 63 (eqn. 11) was very stereoselective, as expected for a rigid system. A more impressive result was the efficiency and stereocontrol of the subsequent intermolecular Michael addition of the resultant 2° radical; only a single stereoisomer of 64 was produced. Treatment of 65 (eqn. 12) under the radical conditions, illustrated that the acyl radical bypassed the 4-exo and the 5-endo intramolecular pathways to participate in a regioselective addition to methyl acrylate. The stabilized radical that resulted underwent a 6-exo-trig cyclization to produce 66 (1.1 :1) to complete this formal [4+2] annulation sequence. Buaan AIBN SePh 62; 58:42 (cisnrans) 0" Bu38nI-I, AIBN . * II/ 53% Yield ’ '1 0 SePh 63 C02 Me Bu38nI-I, AIBN SOP" + olr 33% YIOId (12) P" COzMe Ph 65 66;1.1:1 (transzcis) (11) he lo: he: 17 A number of new and exciting topics in free radical chemistry include macrocyclizations26 and "chiral radicals".27 Using their respective methods. Boger, Curran, and Porter have have produced macrocycles with ring sizes as large as 20 members in surprisingly high yields. Chiral auxiliaries have shown promise for induction of enantioselective bond formation for both intramolecular cyclizations, and for the acyclic stereocontrolled addition of radicals to electron deficient olefins. All of these new developments in free radical chemistry have greatly expanded the repertoire of bond forming reactions available to the synthetic chemist. C. Lithium-Mediated Cyclizations Cyclization methods based on alkyllithiums have received considerable attention on a number of fronts, including synthetic and mechanistic fronts. Product analysis for the isomerization of 5-hexen-1-yllithium has been used as a probe for determining whether cyclization proceeds via an anionic mechanism28 or by single electron transfer.29 If nothing else, the debate which ensued between Bailey (anionic) and Ashby (single electron transfer) has made for some interesting reading. Suffice it to say that there is compelling evidence to support each claim28 depending on the alkyl halide (1°, 2°, or 3°), the nature of the halogen (I, Br, Cl), and the method used to generate the alkyllithium. A number of features of this method are synthetically attractive. First, the requisite alkyllithium could be generated by numerous methods, including (1) lithium- halogen exchange.30a (2) lithium-tin exchange,3°b (3) Shapiro degradation of tosylhydrazones,3°¢ and (4) electrochemical reduction of carbon-halogen”d or carbon-carbon bonds.3°° Secondly, cyclization of a variety of substituted 5- hexen-l-yl systems demonstrated complete regiospecificity,30a exo addition, 18 along with excellent and predictable stereoselectivities.31 Finally, the post- cyclization alkyllithium could be functionalized with a variety of electrophiles, providing useful functionality for further elaboration.28-3°a-3°° Scheme 6. Evidence for Proton Abstraction are—w 8 =8- 67 _ 68 J 69 70 e:X-H 100 0 eX-H 9 0 bX-D 91 0 eX-H 9 91 89% Yleld Treatment of iodide 67 with 2.2 equiv of t-BuLi at -78 °C led to clean generation of the alkyllithium 68 (Scheme 6).303 This species was stable indefinitely at ~78 °C. Also, upon quenching the reaction mixture at -78 °C with CH30H, no detectable amount of methylcyclopentane was observed. However, quenching the reaction mixture with CH3OD, resulted in a 95% yield of 693-b and only 91% deuterium incorporation (698). Thus, only an 87% yield of the alkyllithium was actually present. These results outline the fundamental limitation of this method, extreme reactivity of the alkyllithium, in which the most obvious consequences were reduced cyclization yields and eventual separation problems. Two possible explanations have been set forth for the low yields of alkyllithium formation: (1) the organolithium generated abstracted a proton from the t-butyl iodide formed in situ, competitively with the second equivalent of t-BuLi and/or (2) proton abstraction from the diethyl ether occurred.28 In either case, the extreme reactivity demonstrated by the 19 alkyllithium minimized the potential functional groups that could be present. Diminishing the amount of diethyl ether with an equal increase in the pentane concentration has been shown recently to be marginally effective in reducing the amount of substrate reduction.32 Withstanding these drawbacks, warming 68 to ambient temperature caused rapid and selective methylcyclopentane formation. The activation parameters for this isomerization have been determined to be AH = 11.8 i 0.5 kcal/mol and AS = -30 i 2 eu, which reflects the importance of temperature.303 Treatment of 72 under the same conditions, produced only 6% methylcyclohexane (Scheme 7). However, with the use of a complexant such as TMEDA, the extent of 6-exo—trig cyclization was increased to 68%. The role of the Lewis base served to lengthen, and thus weaken, the carbon-lithium bond. This, in turn, accelerated cyclization to the point of becoming competitive with proton abstraction. Substrates which possessed geminal substitution on the olefin (71) also required the use of Lewis bases to accelerate the sluggish reaction. The result was the regioselective formation of 31. It is important to note the marked difference in selectivity between lithium cyclization of 71, and the analogous cyclization of 30 by radical methods (Scheme 3). Acyclic compounds with vicinal substitution on the olefin failed to cyclize under any conditions. 20 Scheme 7. Regioselectivity of Alkyllithium Cyclizations 1)tBuLi 78°C 6 W 2m 0 ' l 8996de 6 7 (100.0) 29 Me l)tBuLi, -78°C M 2 TMEDA. 25 °c M6; gM. 35W ' ' m Yield 7 1 31 (100: 0) Me /\/\/\/ I D—l'———>B”U 78°C / 2)TNEDA, 25 °c 72 34(32) 32 (68) 35 (0) Cyclization of 73a-c provided the opportunity to probe the stereochemical aspects of the reaction (Scheme 8).31 Cyclization in the absence of any additives produced excellent diastereoselectivities in all cases. If PMDTA was present in the reaction, a noticeable increase in stereoselectivity was achieved. The enhanced selectivity was attributed to the lower temperatures needed for cyclization in the presence of PMDTA. Scheme 8. Stereoselectlvlty of Alkyllithium Cyclizations (5.... 32 1)iBuLiI 48$ W's) 38°C M8 H2 M. “1le 1 R R3 “en: 73 3 8 Adam— IIIIII 3‘ 3’ 6° none/PMDTA 79/81 0: Me Me Me 8/8 92/97 none/PMDTA 98788 b: H H H 90/94 10/6 none/PMDTA 83173 C: H H H 9/3 91I97 Chamberlin has examined the Stereoselectivities of a related process, in which a vinyllithium species was cyclized (Scheme 9).306 Treatment of tosylhydrazones 74a-c with t-BuLi (2.2 equiv) served to generate vinyllithium 21 species 75a-c. Upon warming these compounds to ambient temperature, high yields and stereoselectivities were noted. As with other 5-exo-trig cyclizations, the relative configuration of the newly formed asymmetric center could be predicted by invoking a chair transition state, wherein the alkyl groups occupied pseudo equatorial positions. Scheme 9. Stereoselectlvlty of Vinyllithium Cyclizations - R3 _ Ah RR2—_> tBUU, ‘78 °c 21) 5°C Mane. «88R2 M“ . R2 TsHNN‘Ji 2) H' . R‘ ”H1 7 4 7 5 7 6 7 7 5‘ 132 133 188 a: Mo H H 70 17 33 b: H Me H 6 5 91 9 C: H H Me 70 9 91 Bailey has performed calculations for the transition state of the lithium cyclization.31 Molecular mechanics and ab initio calculations arrived at an energy minimum where the acyclic alkenyllithium was in a pseudo chair conformation, not a linear array of atoms, and that this conformation represented an intermediate. These calculations revealed that contributing to the stabilization of this intermediate was the coordination of the lithium to the carbon-carbon 1: bond, which was first suggested by Oliver in 1966.33 This stabilization would be maximized when the carbon-lithium bond was syn- coplanar to the olefin. This interaction between lithium and the olefin could contribute to the alkene facial selectivity observed for lithium mediated cyclizations. These calculations were qualitatively consistent with experimental facts and help to ascertain the discrepancy in stereocontrol with its radical counterpart (a radical would not be expected to interact with a 1: bond). Substrates 73b and 73c illustrate that cyclization was under kinetic control. 22 Both substrates gave rise to 1,3-dimethylcycl0pentanes, yet were highly selective for different isomers. Further proof was evident in the tandem cyclization of 78 (eqn. 13),34 which produced the thermodynamically less stable trans-[3.3.0] bicyclooctane (the cis isomer is 6.4 kcal/mol more stable). In contrast, radical cyclization of the corresponding bromide produced multiple products in which the trans isomer was present as <4% of the mixture?5 I 1)tBuI.i, -78°C Me“. 2 TMED 25°C ' 3) CH30H H5. (13) 87% YIIId 78 Finally, cyclization of 80 (eqn. 14) produced in high yields the [4.3.3] propellane skeleton.36 More importantly, it serves to further illustrate the adaptability of this method to complex ring patterns and that the final alkyllithium was sufficiently stable to take part in useful, high yielding intermolecular events with C02, aldehydes, and activated alkyl halides. I 1) IBULI. -78 “C E 3114504 0 °c . e:E*-CHgOI-I e:E-H;81% Yield hrs-cg b: E.co,H;6ox Yield 1: E‘- PhCHO c: E . PhCHzoi-i; 30% Yield d: 5* - CHch-ICI-IzBr a: E . CHZCHCHZ; 02% Yield D. Magnesium-Mediated Cyclizations Grignard reagents have shown little tendency toward addition to olefins in intermolecular reactions. However, when 28 was added to magnesium in diethyl ether or THF and the products hydrolyzed, the resulting mixture 23 consisted of a 95:5 ratio of 82 and 83 respectively (eqn. 15).37 Formation of 83 was suggested to arise from radical intermediates generated by single electron transfer (SET) during the formation of the Grignard reagent. When the reaction mixture was heated in excess of 100 °C. complete conversion to 83 resulted}38 Richey studied the rates of reaction and the simple diastereoselectivity of some substituted 5-hexen-1-yl and 6-hepten-1-ylmagnesium bromides. As expected, five-membered ring formation was faster than that of the six-membered ring (by a factor of 2800). The diastereoselectivity was modest, favoring cyclization from the more stable chair conformation. Utimoto incorporated a trimethylsilyl group on the terminal carbon of the olefin to accelerate the cyclization and reduce the necessary reaction temperatures (eqn. 16).39 Thus, treatment of 84 with 1.5 equivalents of magnesium in THF at 67 °C to provided, upon quenching, 866 as the only product. Moreover, when cyclization was complete, trapping the final Grignard intermediate could be achieved with suitable electrophiles. - 8- M9 r Me Br MgBr —.W Cr + ——+:;:::i°°° O (.5, 23 82 (95:5) 83 6(100) ms ' D H ms Mo _. ”3% _.2 E’ J 1 ,67°C H 04 85 86 E*-Hzo e: R-H;81%Yleld E‘ -¢\.Br 8: R - Aliyi;m Yield 24 E. Aluminum-Mediated Cyclizations Alkyl aluminum hydrides have received limited attention as a method to convert 1,5-dienes into methylcyclopentanes. Although, when 1,5-hexadiene (1) was treated with 2.1 equivalents of iBuzAlH at 60 °C for 16 hours (eqn. 17), methylcyclopentane (6) accounted for >98% of the products formed, after hydrolysis, with the balance of material being n-hexane.4° However, when 1,6- heptadiene was treated under the same conditions, less than 1% of the product mixture was methylcyclohexane with the major component n-heptane (>99%). < (1)iBuzAlH. 60°C: [GE](2)H*GO, O’Me (17) 1 Scheme 10. Stereoselectivity of Aluminum Cyclizations Me Me (2er MUM \ mswmo-c + as 30000) 67 MO \ AIEI: 33 88 Me Me Me EtzAl Me _9o _ i%:0 7512M - Me‘. . Me (ZIH’ + WM _. Q... U... . 92 388 (8:92) 55 t. i) (1)Et,Al-I,60'C Me 9‘ iiwm L. 93 _ 388 (97.3) ' 37 25 In an attempt to answer some of the questions about the stereoregulating forces present in the Ziegler-Natta polymerization of olefins, Stefani investigated the diastereoselectivity for the aluminum-mediated cyclizations of 88 and 91 (Scheme 10).“0 Reaction of 88 and 91 with Et2AlH produced two structural isomers in each case (89.90 and 92,93 respectively), due to the indiscriminate addition of the alkylaluminum hydride to the dienes. Cyclization and analysis of the reaction mixtures revealed important selectivity information. For all cases, only exo insertion was observed, even for the geminally substituted olefin, 89. The most important feature of the reaction was the ability of the alkylaluminum to control the relative configuration of the newly formed asymmetric center. Selectivities for the cyclization of 90, 92, and 93, ranged from 92:8 to 97:3. These results. which paralleled those of lithium cyclizations, reflect the ability of both lithium and aluminum to coordinate the olefin and thus control the facial selectivity of olefin insertion and the equatorial deployment of methyl groups. F. Palladium-Mediated Cyclizations Transition metals are particularly well-suited for this type of bond construction because of their ability to insert olefins into carbon-metal bonds. Palladium remains a convenient metal to use (despite the expensive cost) because of the variety of ways it can be introduced into the cyclization substrate and because catalytic processes are possible. Palladium can be bound to carbon by either transmetallation or through oxidative addition into a carbon- heteroatom bond. Overman has recently reported tandem cyclizations using palladium, generating polycyclic molecules with spiro (eqn. 13)41a or fused rings (eqn. 19).41b The catalytic cycle for these cyclizations consisted of 26 oxidative addition of Pd(0) (generated in situ) into the carbon-triflate or the carbon-iodide bond. sequential intramolecular olefin insertions, B-hydride elimination, and reductive elimination of H-X (X=heteroatom) from palladium which regenerated the Pd(0) catalyst. Olefin insertion usually followed in an exo manner to yield the smaller ring, unless some dominating steric factor directed endo insertion. The major limitation of this methodolgy was that the initial carbon-palladium bond must either contain an aryl or a vinyl carbon to prevent B—hydride elimination prior to cyclization. OTI n \ 0.1 equiv Pd(OAc)2 0.4 equiv PPha (1 8) : m-1;n-1 72% Yield 0 9 5 I m-1; n-2 70% VIC“! I [II-2; I'll-1 53% VI.“ I I'D-2; "-2 50% VI.“ 94 GOOD Me 0.1 equiv Pd(OAc)2 l 0.4 equiv Part, 87% Yield (1 9) 96 ' 97;H.a:8-1.3:1 3. Background for the Mechanism of Ziegler-Matte Polymerization Ziegler-Nana polymerization remains one of the most important industrial processes. Due to the vivid imaginations of a number of polymer chemists, extensive knowlege exists on how to control catalyst activity and the tacticity (stereochemistry) of the polymer, be it isotactic42 or syndiotactic.43 However, very little is actually known about the fundamental steps that comprise the 27 mechanism. Central to the controversy was the interaction of the olefin with the metal, mode of insertion, and how these factors influenced the stereochemistry of each insertion. Of the plethora of mechanisms that have come and gone, two have received serious consideration, the first by Cossee44 and the other by the combined efforts of Green and Rooney.45 Proof of the credibility of these proposals can be seen by the enduring efforts to substantiate and/or disprove these mechanistic possibilities. Cossee suggested, in 1960, a direct insertion mechanism, the details of which are shown in Scheme 11.44 Compound 98 (shown only for the monomer, yet the dashed titanium-chloride bond indicates the point of attachment of another monomer) was suggested to result from the treatment of TiCla with an alkyl aluminum compound. Reaction of 98 with ethylene generated complex 99 wherein the olefin interacted with the available 0 orbitals of the titanium to form a 1: complex. Cossee reasoned that 99 was a credible structure based on a similar finding by Chatt and Duncanson,46 in which a crystal structure revealed that a platinum complex bound an olefin in this manner. The 6 bond between the olefin and titanium was formed through the interaction of the it electrons of the olefin with the dx2-y2 orbitals of the metal.““:“7 Further stabilization could be provided by interaction of the dxy orbital and the antibonding orbitals of the olefin. Propagation was proposed to occur via a four centered transition state, 100, which yielded 101 after insertion.48 28 Scheme 11. The Cossee-Arlmann Mechanism for Chain Propagation p- - P - R n a... or 0H2 l CH i I: H cruyL/ir || —_ cr-7lr/C | 2 —» 08711-151" 2 or | CH2 or | CH2 or | "CH2 or or Cl 9 8 _ 9 9 _ _ 1 00 R\TH2 CI-7 lie-CH2 bonding C. or _( 101 Green and Rooney discounted mechanisms such as Cossee's, which involved direct insertion.45 They did so on the grounds that isolable alkyl- metal-olefin complexes (metal=W or Mo) could not be induced to insert the alkyl group. On the other hand, similar hydrogen-metal-olefin complexes readily inserted the hydrogen into the olefin. It has also been shown that certain olefin polymerization catalysts could participate in ring-opening metathesis reactions. All of these facts were combined to give rise to their mechanistic proposal (Scheme 12), which differed from all others by the presence of a metallocyclobutane intermediate (105). Thus. a-hydrogen elimination of 102 (the net result being oxidation of the metal) produced metallocarbene 103 which was proposed to complex the olefin to give rise to 104. A [2+2] reorganization of the it electrons could form the metallocyclobutane 105 as the key intermediate. The propagation sequence was completed by a 1,2-hydride shift resulting in 106 (and return to the initial oxidation state of the metal). 29 Scheme 12. The Green-Rooney Mechanism for Chain Propagation r 7 CH2 r - H II H K H CH2 ‘7: . ‘1“ CH2 [Tl] H—[Ti] H—mljl 1 0 2 _ 1 0 3 .. _ 1 041 “Hz— H r H ' $13 ‘__ [1"] HI | 1 06 _ 1 05 ] The lack of conclusive experimental evidence for these mechanisms has prompted a number of different investigations. One path toward obtaining such evidence involved synthesizing isolable metal complexes that resembled proposed intermediates, and the second, which will be addressed shortly, involved monitoring the stereochemical details of a single olefin insertion to probe the microstructure of these intermediates. There have recently been a number of reports of alkyl-metal complexes in which hydrogens on the alkyl chain were suggestively close to the metal center, implying a bonding interaction to the metal. Representative examples of these "agostic hydrogen interactions” are shown in Figures 1-3, in which the only difference was whether an a- or a B-hydrogen participated. Green suggested that the tar-hydrogen interaction in Figure 1 lent support and resembled his propagation mechanism.49 in contrast. B-hydrogen interaction (Figures 250a and 350D) was proposed to represent a termination intermediate. namely that of B-hydride elimination. A similar claim has been set forth for a zirconocene polymerization catalyst.42a 30 “32 ’CH3 < 5 or c2 \ s C1 I \ 1 e C /+{‘~- H‘\ / [CH3 ‘93P / l H :Zr ___P..0\\\CH3 H30 I Cl cr H706“);- CH3 H706 CH3 Figure 1. a-Agostic Interaction "007° 2- 51'A92051i0 Interaction for 107: Ti-C1-H - 70°; Ti-H . 2.03 A. for 108: Zr-C C - 85°: Zr-H - 2.16 A. ”39 CH3 I Figure 3. B-Agostic interaction for 109: 1"i-c‘-c2 . 86°; T-H - 2.29 A. Two major problems were apparent with the foundation of the mechanism of Green and Rooney. First, the fact that the complexes of tungsten and molybdenum did not insert an alkyl group should have not been surprising at all. These complexes were electron rich, 18 electron complexes, which is in direct contrast to typical, active Ziegler-Nana catalysts (primarily 14-16 electron complexes). Secondly, a more serious problem was illustrated in Figures 2 and 3. These figures demonstrated that when 01- and B-hydrogens were present, as in the growing polymer chain, B-agostic interactions dominate over or- interactions. Unless 01- and B-hydrogen interactions are in equilibrium in solution, it is unclear, based on the experimental facts and their conjecture, how polyolefins would be produced. In fact, B-hydrogen interactions have been suggested to play a role in polymerization,43-509851 although the cursory role has not explicitly been determined. 31 The role of B-agostic hydrogens is complex. Brookhart has shown that a cationic ethylcobalt(lll) complex adopts a B-agostic structure.51 Treatment of this cobalt complex with three equivalents of ethylene in an NMR tube resulted in sequential oligomers of ethylene. However, the first insertion was markedly slower than subsequent insertions. Brookhart attributed this rate difference to the unfavorable disturbance of the B-hydrogen interaction (9.2 kcal/mol), which was contributing electronic stabilization to the electron deficient metal center. After the initial insertion, other agostic interactions were observed by NMR analysis. Independent synthesis and NMR analysis of a terminally bound cationic butylcobalt(lll) complex, with a distinguishable [3-hydrogen agostic interaction, provided direct correlation to that observed after the first insertion of ethylene. Based on these results, Brookhart contended that alkyl migration was the rate limiting step, and the agostic interaction played an undetermined subordinate role. Bercaw has observed similar results for a scandium complex.48 For example, ethylene insertion was found to be faster for those alkylscandium complexes that did not contain a B-agostic hydrogen than those that did. Bercaw rationalized this phenomenon in which he proposed that B- agostic interactions [represented ground state structures, and or-agostic interactions represented transition states for chain propagation. Jordan has suggested that the role of B-agostic interactions may be to influence the conformation of the 8-carbon, and thus play a role in the stereoselectivity of or- olefin insertion!”a Grubbs designed a clever experiment to test the Green and Rooney mechanism.9a To simplify the analysis, a system was choosen that would allow for only a single insertion to take place (Scheme 13). Central to the design was the the incorporation of an tar-deuteron atom and an tar-hydrogen (110). Grubbs reasoned that if or-activation was important, there would be an unequal 32 partitioning between the diastereomeric transition states (111a and 111b) based on the relative strengths of the CH and C-D bonds. Thus, if ot-activation was pivotal in the insertion process, an excess of the trans product should be produced. Experimentally, Grubbs determined that, upon activation of 110 with EtAlClz, a 1:1 ratio of diastereomers was produced, which provided convincing evidence against the Green and Rooney mechanism. Grubbs proceeded to point out that "...polymerization is the chemistry of carbon-metal single bonds and metathesis is the chemistry of carbon-metal double bonds."9a Furthermore, Grubbs suggested that these findings supported the Cossee mechanism. Scheme 13. Grubbs' Test for a-Hydrogen Activation {o /LH \ ‘/ \ 092T ~ 0927' . H D 1) EtAlCl2, ‘3?“ -—- ‘13,“? W “W2Cl '1 00 C ‘s‘ 5 9‘ '4' 1 1 0 8,1426, Et’A'CCl Oi Cl 1 1 1 a 1 1 1 b trans (favored) cis (disfavored) J 1) olefin insertion 2) H” Me cis/trans - 1.00 3: 0.05 QC 1 1 2 In contrast to the findings of Grubbs, Bercaw recently has shown that a scandium complex (isoeiectronic with the Grubbs' complex) did in fact show an o-agostic assistance.9° Treatment of 113 with a scandium-hydride complex resulted in a regioselective production of 114 (Scheme 14). Cyclization of this complex provided an excess of the trans product. Moreover. the deuterium 33 isotope effect measured correlated in a normal enthalpic manner with respect to temperature. In order to insure that these results were not a consequence of stories, based on the olefin geometry and the size of a deuterium atom versus a hydrogen atom, the cis olefins (116) were reacted under the same conditions. Consistent with the trans olefins, a deuterium isotope effect of the same magnitude was observed. Bercaw also noted or-activation for the cyclization of 6-hepten-1-ylscandium complexes generated from 117. These results, coupled with Grubbs' findings, suggest that not all olefin polymerization catalysts proceeded through the same mechanistic pathways. Scheme 14. Bercaw's Test for a-Hydrogen Activation - H H _ 0. l D o. J H o OpScI-I(PM63) , . = ‘ -_ WV‘ 0 Hz, 4 atm OpSc 095,. 1 1 3 H 1149 1148 L trans (favored) cis (disfavored) d Me Me Me CHzD Me\ M. trans/cis - 1.26 1; 0.03 (-10 °C) 0 /Si m - Op trans/cis - 1.19 :l: 0.04 (25 °C) Me trans/cis - 1.07 :I: 0.03 (120 °C) Me Me 1 1 5 M0 . 0 CH2” OpScHlPMOo) 0 M ”2 4 atm trans/cis . 1.20 i: 0.04 (25 °C) > D 116 115 D D OPSGHIPMOa) CHZD WVV H2, 4 atm . o ,, D ois/trans - 1.21 i 0.05 (25 C) 117 118 CHAPTER 1. SYNTHESIS OF CYCLIZATION PRECURSOR SUBSTRATES AND STANDARDS 1 . Introduction The following chapters discuss the regio- and stereoselective features, along with mechanistic considerations for the transformation of 120 to 119 (Scheme 15). The antithetic sequence indicates alkenyltitanocene chlorides 120 as common intermediates in each reaction. Substrates analogous to 120. differing in tether length and location of an alkyl group (R), were prepared by existing methodology6 via Grignard reagent 122, which could be prepared from alkyl bromides 121. The present chapter will describe the synthetic routes for the preparation of the cyclization precursors (121) to be used in our various studies. Scheme 15. Retrosynthetlc Analysls of the Tltanocene Dichloride-Medlated Carbocyclization opzcrn g: = Tlefil =>ifmm =gV/B' R R R R 120 122 121 1 1 9 n-3.4 R-akyl 34 35 2. Synthesis of Five-Membered Ring Precursor Substrates The synthetic sequence followed for the preparation of compound 125, needed for our mechanistic study, is shown in Scheme 16. Beginning from the known a-methyl ester,19 reaction with LDA at -78 °C generated the ester enolate. In order to remove acidic protons from the reaction mixture (i.e. DIPA), 1.25 eq nBuLi was added. The reaction was quenched with an excess of 020, which provided the a-methyl-a-deuterioester (123) in modest yield (64% yield). Subsequent treatment of 123 with LAH provided alcohol 124 (91% yield), which was converted to bromide 125 (84% yield) by the combined action of NBS and PPh3.52 Scheme 16. Synthetic Route to 125 o o b MOE-5.1K" More: 1115* W011 M0 M9 D M. D 123 124 6 W / e A/MX‘ , 1 2 5 (a) LDA. THF. -73 '0; neuu. -70 °C; op. (b) LAH. Ego. NaOH. Hp. (c) NBS. Plan, cruel, Compound 129, to be used in a competitive rate experiment with 125, was prepared in four linear steps beginning from the commercially available diethyl propylmalonate (Scheme 17). Enolate formation of the malonate derivative was accomplished with NaH in DMF and then subsequently coupled with 4-bromo-1-butene to provide 126 in 96% yield.19 Deethoxycarboxylation was encouraged by the action of LiCl and H20 in refluxing DMSO. which resulted in the formation of ester 127 (91% yield).19 Conversion of 127 into 129 was brought about by adjustment of the oxidation state of 127 with LAH to 36 the alcohol level (128, 94% yield), followed by functional group manipulation to give bromide 129 (82% yield). Scheme 17. Synthetic Route to 129 E E a H b E Y '56—’16 / W / Pr We /\/\l': 5,0025, 126 127 new» /\/\/\ou%> We: W Pr 128 129 (a) NaH, DMF; 4-Bromo-1-butene. (b) ucu. H20, DMSO, 135 °c. (c) LAH, Ego; NaOi-l, H2044 Nesmpha. cw.» An ongoing project in our group deals with the synthesis of multiple ring compounds. Cyclization of 1 33 (Scheme 18) would provide the bicyclo[4.3.0]nonane, which is common to many families of natural products. Synthesis of 133 required six linear steps beginning with radical bromination of cyclohexene, accomplished with N88 and peroxides (130; 60% yield).53 Numerous literature methods for allylic coupling were attempted to prepare 131.54 After many sophisticated organometallic routes failed to give sufficeint quanities of 131, recourse to a simple Grignard coupling was pursued. Thus, when allylmagnesium bromide55 was reacted with 130, the key carbon—carbon bond formation was secured (131; 60% yield). Next, a chemo- and regioselective oxidation of the terminal olefin was needed. Through the use of a bulky boranefi‘"6 this turned out to be a trivial transformation. Addition of 131 to a THF solution of dicyclohexylborane, followed by oxidation with peroxides,57 resulted in a single structural isomer (132). To facilitate the chromatographic separation of 132 from the reagent byproduct cyclohexanol, the 1° alcohol was selectively protected as the TBDMS ether.58 Deprotection of the silyl other with 37 TBAF, followed by chromatography, provided 132 in pure form (52% yield). Conversion of 132 to 133 was best accomplished by transforming the alcohol to the mesylate followed by displacement with LiBr. This sequence provided a modest yield of 133 (70% yield).59 Scheme 18. Synthetic Route to 133 O a O b 55% 60% Br 130 131 6,.0 52% Br ‘ OH cu mm Cu 133 132 (a) NBS, CCI4. (b) Allylmagneslum bromide, THF. (c) Dicyclohexylborohydride, THF; H202, NaOH. (d) TBDMSCl, DMAP, CHZClz; chromatographic separation. (e) TBAF; chromatographic separation. (f) MsCl, CHZClz; LiBr. THF. 3. Synthesis of Slx-Membered Ring Precursor Substrates The synthetic route for the parent substrate (137) is illustrated in Scheme 19. Alkylation of diethyl malonate with 5—bromo-1-pentene furnished 134 (83% yield), which was subsequently treated with LiCl and H20 in refluxing DMSO to provide ester 135 (86% yield). Reduction of the ester followed by bromination of the resulting alcohol (136, 92% yield), completed the synthesis of bromide 137 (89% yield). 38 Scheme 19. Synthetic Route to 137 E E E -——> W —?——> c sass / E sees ME 92% 134 135 5.0025: d AAA/OH __> MS! / sees / 136 137 (a) NaH, DMF; 5—Bromo-1-pentene. (b) UCI. H20, DMSO, 180 °C. (c) LAH, E120; NaOH, H20. (d) NBS, PPha, CH2C'2. Scheme 20 illustrates the synthesis of the allylic- (144) and the homoallyic (151) methyl-7-bromo-1-heptenes. The key bromides 140 and 147, needed for coupling with diethylmalonate, were prepared from esters 138 and 145, which were accessible by employing the ortho-ester Claisen methodology. For example, crotyl alcohol was heated with triethyl orthoacetate and an acid catalyst to provide 138 in 65% yield.60 Treatment of 138 with LAH (139, 69% yield) followed by bromination gave 140 (75% yield). In a similar fashion, allyl alcohol was condensed with triethyl Orthopropionate and rearranged to ester 145 (79% yield), which was subsequently reduced (146, 78% yield) and brominated (147, 64% yield). Following the usual sequence of reactions, 140 and 147 were converted to the key cyclization substrates, 144 and 151, in 56% and 53% overall yield. Moreover, Scheme 21 illustrates the synthesis of 155 and 160, prepared in the usual fashion, which were synthesized in 57% and 55% overall yield, respectively. 39 Scheme 20. Synthetic Route to 144 and 151 R2 R2 32 R‘ OH L b CH C B W / 0025: —> / ——> / r + R‘ R‘ 31 W08): 130; uses; R'aMe, R’sH 1393914; R‘uMe, R241 140; 75%; R'-Me, az-H 145; m; R'sH, R2=Me 145; me; R'sH, Rz-Me 147; 04%; R'sH, stMe R2 E R2 R2 d o 9 OH R‘ R‘ R1 141; 95%: R‘IMO, RzzH 1‘2; 30%; R‘ aMe, R2311 143; 85%; R‘sm, Rat-H 14s; R‘af-l, stm 149; we (2 steps); R‘sH, R’sMe 150; 93%; 8‘41, RZ-Me R2 ——D- f R1 144; uses; Hum. stH 151; sees; R‘aH, R2=Me (a) (HachCO-fi, 145 °c. (b) LAH, Ego; NaOH, H20. (c) was, Pang CH20l2. (d) Diethyl malonate, NaH, DMF. (e) ucr. Hp, 01490, 100 °C. Scheme 21. Synthetic Route to 155 and 160 1 E E R d EVE - EYE . ,, W“, I We nz R2 5.0025: R‘ R'uH L» 152; ms; R‘.H, Fla-Me 153; ms; nun, 1:11.149 “new; R1-iPr -°—> 157:95%;R‘-iPr,R2-H 150;m;n‘.ipr.n2.u R‘ at . OH f -°-> W —>- WB' n2 32 154; ease; R'sl-l, 93-149 155; sex; R'cl-f, 92.1w; 159; ease; R'a-iPr, nz-H 100; e494; R'siPr, 122.14 (a) NaH, DW; 2-Bnomopropene. (b) Nal-l, DMF; s-Bromo-Htexene. (c) NaH, DMF;5-Bromo-1-pentene. (d) UCl, H20. 01490, 185 ‘C. (e) LAH, Efgo; NaOH. H20. (0 N88. PPM, CHZClz. 40 Scheme 22. Synthetic Route to 164 CH O O O CH C: fi” ¢ %, E ) 36%, WW iPr iPr iPr iP 1 6 1 1 6 2 1 6 3 125:. WBr iPr 1 64 (a) Jones' reagent. Ergo. (b) mCPBA, 0142c:2 (c) DiBAl, toluene, -73 (*0. (d) Ph3PCH2; Hgot. (e) NBS, PPhg, cuzcr, Jones' oxidation61 of 4-isopropylcyclohexanol (Scheme 22) provided ketone 161 (92% yield), which was further oxidized using mCPBA62 to give caprolactone 162 (82% yield). The reductive-olefination sequence, 162 to 163, proved to be troublesome, however, after a number of attempts and chromatographic purification, 163 was in hand albeit in only 22% yield. Lactone 161 was reduced with DiBAI to the lactol at -78 °C, and olefinated with the methyl Wittig reagent, which secured the alkenol.63 Conversion of 163 to bromide 164 was performed with NBS and PPha without further difficulties (89% yield). 4. Synthesis of Competitive Rate Substrates Our interest in the elucidation of the mechanism of Ziegler-Natta polymerization has compelled us to design an experiment to test whether on- and/or B-activation contribute to the mechanism. Substrates 168, 171, 175, 178, 182, 184, and 186 (Scheme 23) were chosen to test our hypothesis.64 Upon examination of Scheme 23, three unique, yet similar, linear sequences 41 were envisioned in which three intermediates (166, 173, and 180) were common to all. Reaction of the enolate from the appropriate malonate (R: ethyl, propyl, or butyl) with 5-bromo-1-pentene in DMF provided 165, 172, and 179, which were independently carried on to the respective esters. At this juncture, the three sequences diverged. The aforementioned esters were reduced and brominated to provide 168, 175, and 182, respectively. The second sequence involved reduction of esters 166 and 173 with LAD followed by bromination, which furnished a,or-dideuterio bromides 184 and 186. Finally, conversion of esters 166 and 173 to the B-deuterio esters 169 and 176 was accomplished, as previously described and carried on to bromides 171 and 178 in the usual fashion. Schemes 24 and 25 illustrate the sequences performed to prepare the expected products of cyclization for the appr0priate substrates. In Scheme 24, cyclohexenone was reacted at low temperatures with the appropriate Grignard reagent in the presence of a catalytic amount of a copper salt.55v66 The resultant B-alkyl ketones were methylenated via Corey's procedure.67 Finally, hydrogenation of the exocylic olefins provided a mixture of the cis- and trans- dialkylcyclohexanesfi.8 In a similar manner, ketone 161 was methylenated and hydrogenated to provide 197. 42 Scheme 23. Synthetic Route to Competitive Rate Substrates E E a We Y -» / . R 105; ease; RsEt 172; 99%; RsPr E43025! 170; 04%; 11.80 R W9 175; 83%; R-Pr R R Br MO”; M o o o o 183; 97%; R28 184; 88%; Flea 185; 73%; Rst 186; 91%; R-Pr ). R R 0 b We 1.,W\> 180; 88%; R-Et 189; 78%; RsEt 173; 84%: RsPr 176; 61%; RzPr 180; 88%; R-Bu l‘ . l; . d W +— / OH 107; 95%; R-Et 170; 92%; R-Et 174:0596; R-Pr 177; 91%; R-Pr 181; 90%; RsBu ‘d R 0 MW 1 71:91 %; RsE! 173:77%; R-Pr (a) NaH, DMF; 5-Bromo-1-pentene. (b) ucr, Hp, ouso, 100 °c. (c) LAH, Exp; NaOH, Hp. (d) NBS. Pm, chi, (0) LAD, Exp; NaOf-l. Hp. (r) LDA, THF, -73 -c; nBuLi, .73 .c; op. 43 Scheme 24. Synthetic Route to Cyclohexane Standards {heeded 187; 82%; RsEt 188; 71%; R28 189; R-Et 190; 57%; RsPr 191; 71%; R=Pr 192; R-Pr 193; 78%; R: iPr 194; 89%; R: iPr 195; R: iPr (a) CulP(nBu)3, RMgBr, Et-p; Hp’. (b) Ph3PCH2, DMSO. (c) H2, P02 Scheme 25. Synthetic Route to 196 and 197 0 Me a b iPr iPr iPr 1 6 2 1 9 6 1 9 7 (a) Ph3PCl-12, DMSO. (b) H2. P02, MeOH. 5. Conclusion The preparation of the substrates needed for the diastereoselective five- membered ring formation (Chapter 2), diastereoselective six-membered ring formation (Chapter3), and our mechanistic studies (Chapter 4) was described. In all cases, recourse to established literature methods was utilized. 44 6. Experimental General Methods. 6-Bromo-3-methyl-1-hexene (363), 6-bromo-4-methyl- 1-hexene (36b), 6-bromo-5-methyl-1-hexene (36c) and 6-bromo-6-methyl-1- hexene (36d) were kindly provided by Dr. J. R. Stille. Diethyl 2-butyl-2-(4- pentenyl)-propanedioate (179), 2-butyl-6-heptenoate (180), 2-butyl-6-hepten-1-ol (181), and 7-bromo-6-butyl-1-heptene (182) were kindly donated by Nancy S. Barta. 7-Bromo-6-methylhept-1-ene was provided by Steve Gabel. The following cyclization standards were purchased from either Aldrich Chemical Co. or Wiley Organics: cis— and trans-1,2- and 1,3-dimethylcyclopentane, methylcyclopentane, cis- and trans-12, 1.3-, and 1,4-dimethylcyclohexane, cyclohexane, methylcyclohexane, cycloheptane, methylcycloheptane, and bicyclo [4.3.0] nonane. MezAlCl and EtAlClz were purchased from Aldrich Chemical Co. from which stock solutions were prepared. Tetrahydrofuran (T HF), diethyl ether (Et20), toluene, and benzene, when used as solvents, were distilled from sodium/benzophenone prior to use. Hexane was stirred over sulfuric acid for five days, then washed sequentially with water, saturated NaHCOa, dried over CaCl2, and the supernatant was distilled from sodium/benzophenone/tetraglyme. Dichloromethane (CH2Cl2) was distilled from CaH2 prior to use. Anhydrous dimethylsulfoxide and N,N-dimethylformamide were purchased from Aldrich. All reactions were conducted under nitrogen or argon atmospheres. NMR spectra were obtained on a Varian Gemini 300 or a VXR-300 instrument with CDCl3 as solvent. Signals are reported in units of ppm relative to the residual chloroform peak. Multiplicities are given as s (singlet), d (doublet), t (triplet), q (quartet), quint (quintet), sext (sextet), sept (septet), oct (octet), rn (multiplet), or b (broad). 13C NMR resonances for carbons containing a carbon- 45 deuterium were not observed due to the quadrapolar effect caused by the deuterium atom. Infrared spectra (IR) were obtained on a Nicolet FT -IR 42 spectrometer using NaCl plates with thin films of organic compounds and reported in cm'1. Analytical gas chromatography (GC) was performed on a Perkin Elmer 8500 instrument with a 50 m RSLZOO column (5% methyl phenyl silicone equivalent to SE-54 or BBS). All GO yields described were calculated by comparison to an internal standard and corrected for detector response. Ethyl 2-methyl-2-deuterlo-S-hexenoate (123). A solution of LDA was prepared by adding n-BuLi (2.29 M in hexanes, 10.1 mL, 23.1 mmol) to a 0 °C solution of DIPA (2.55 g, 25.2 mmol) in THF (30 mL). Stirring was continued for 30 minutes and the mixture was cooled to -78 °C. Ethyl 2-methyl-5-hexenoate (3.28 g, 21 mmol) was added dropwise to the LDA solution and stirred for 50 minutes. n-BuLi (2.29 M in hexanes, 13.80 mL, 31.5 mmol) was added to the reaction mixture at -78 °C (to deprotonate DIPA), stirred for 30 minutes, quenched with 020 (10 mL), and warmed to ambient temperature. The aqueous layer was extracted with diethyl ether (4 x 20 mL). The combined diethyl extracts were concentrated and redissolved in diethyl ether (60 mL). The diethyl other solution was washed with 1M HCI (2 x 30 mL), saturated NaHCCa (2 x 60 mL), saturated NaCl (60 mL), dried over MgSO4, filtered, and conentrated to an oil. The crude residue was distilled under reduced pressure (bp22 70-80 °C) and 2.10 g of 123 were obtained (64% yield). 1H NMR (300 MHz) (CDCl3) 81.11 (s, 3H), 1.22 (t, J . 7 Hz, 3H), 1.30-1.52 (m, 1H), 1.67-1.79 (m, 1H), 2.03 (tq, J- 1, 7 Hz, 2H), 4.10 (q, J- 7 Hz, 2H), 4.93 (ddt, J: 10, 2, 1 Hz, 1H), 4.98 (dq, J: 17, 2 Hz, 1H), 5.75 (ddt, J . 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3) 8 14.22, 16.88, 31.32, 32.74, 60.11, 114.96, 137.92, 176.00. General Procedure For The Reduction 01 Esters- 2-DeuterIo-2-methyl- 5-hexen-1-ol (124). A 100 mL flask, equipped with a dropping funnel, was loaded 46 with LAH (0.36 g, 9.36 mmol), suspended with diethyl ether (35 mL), and cooled to 0 °C. A solution of 123 (2.10 g, 13.38 mmol) in diethyl ether (40 mL) was added to the LAH suspension over 30 minutes. Stirring was continued at 0 °C for 1 hour, after which the dropping funnel was replaced with a reflux condenser and the reaction was heated to reflux for 3 hours. The reaction mixture was cooled to 0 °C and quenched by the slow sequential addition of H20 (0.36 mL), 15% NaOH (0.36 mL), H20 (1.08 mL) and stirred for 8 hours. The white precipitate was removed by filtration through a pad of silica gel on top of a glass frit. Concentration of the solution by rotary evaporation followed by distillation of the final residue (bp13 70-80 °C) provided 1.40 g of 124 (91% yield); 1H NMR (300 MHz) (CDCI3) 5 0.88 (s, 3H), 1.08-1.22 (m, 1H), 1.40-1.58 (m, 1H), 1.74 (bs, 1H), 1.88-2.18 (m, 2H), 3.38 (d, J: 10 Hz, 1H), 3.46 (d, J= 10 Hz, 1H), 4.91 (dq, J= 10,2 Hz, 1H), 4.97 (dq, J: 17, 2 Hz, 1H), 5.77 (ddt, J: 17, 1o, 7 Hz, 1H);13C NMR (75 MHz) (00013) a 16.26, 31.09, 32.15, 67.99, 114.33, 138.87. General Procedure For The Bromlnatlon Of Alcohols— 6-Bromo-5- deuterlo-S-methyI-1-hexene (125). Triphenylphosphine (3.15 g, 12.0 mmol) and alcohol 124 (1.15 g, 10.0 mmol) were combined and diluted with CHzClz (20 mL), and cooled to 0 °C. NBS (2.14 g, 12.0 mmol) was added over 2 hours by means of a solid addition funnel, stirred for 1 hour at 0 °C, and an additional 3 hours at ambient temperature. The reaction mixture was concentrated to a slurry by rotary evaporation, and the resulting solids were dissolved in a minimum of CH2CI2. To this solution was added 50 mL petroleum ether (bp 30-65 °C) at ambient temperature, stirred for 10 minutes, and cooled to 0 °C for an additional 10 minutes. The supernatant was filtered via cannula, with a piece of filter paper attached to the end of the cannula, using a positive pressure of argon. The solids were washed sequentially with petroleum ether (2 x 25 mL) at 0 °C as before. This filtration sequence was performed twice more. The organic fractions were 47 combined and concentrated by rotary evaporation to approximately 15 mL and then filtered through a plug of basic alumina. The remaining solvent was removed by rotary evaporation and the residue was purified by distillation (bp4o 65-75 °C) to provide 1.51 g of bromide 125 (84% yield); IR (Film) 3079, 2965, 2930, 2260, 1642 cm"; 1H NMR (300 MHz) (00013) 51.00 (s, 3H), 1.30 (m, 1H), 1.50 (m, 1H), 1.95-2.15 (m, 2H), 3.31 (d, J= 10 Hz, 1H), 3.37 (d, J= 10 Hz, 1H), 4.95 (ddt, J: 10, 2, 1 Hz, 1H), 5.01 (dq, J=17, 2 Hz, 1H),5.77(ddt,J=17, 10, 7 Hz, 1H);13C NMR (75 MHz) (CDCI3) 518.50, 31.01, 33.81, 41.18, 114.79, 138.29. General Procedure For Alkylation Of Malonate Derivatives- Diethyl 2- (3-butenyl)-2-(propyI)-propanedioate (126). To a suspension of NaH (1.83 g, 76.39 mmol) in DMF (70 mL) was added diethyl propylmalonate (14.45 g, 71.4 mmol) over 15 minutes by means of a pressure-equalized dropping funnel. After the addition was complete, the reaction was stirred for 2 hours. 4-Bromo-1- butene (10.79 g, 80.0 mmol) was added over 15 minutes, stirred for 30 minutes, and then heated to 65 °C (generally 12-24 hours). When the alkylation was complete, the reaction mixture was cooled to 0 °C, diluted with diethyl ether (80 mL), and quenched with H20 (100 mL). The aqueous layer was extracted with diethyl ether (3 x 80 mL). The combined extracts were washed with H20 (2 x 50 mL), saturated NaCl (60 mL), and dried over M9804. The organic solution was filtered through silica on a glass frit, eluted with diethyl ether, and the solvent was evaporated. Distillation under reduced pressure (bpd 70-80 °C) provided 17.58 g of 126 as a colorless oil (96% yield); IR (film) 3081 , 2967, 2876, 1732, 1644 cm"; 1H NMR (300 MHz) (CDCI3) 5 0.88 (t, J =- 7 Hz, 3H), 1.07-1.20 (m, 2H), 1.20 (t, J a: 7 Hz, 6H), 1.78-1.98 (m, 6H), 4.14 (q, J a 7 Hz, 4H), 4.88-5.03 (m, 2H), 5.65- 5.82 (m, 1H); 13C NMR (75 MHz) (CDCl3) 514.04, 14.34, 17.34, 28.36, 31.53. 34.53, 57.27, 60.95, 114.83, 137.70, 171.68. 48 General Procedure For Deethoxycarboxylation of Malonate Derivatives- Ethyl 2-propyI-5-hexenoate (127). In a 250 mL flask were combined the following: 126 (17.0 g, 66.4 mmol), LiCl (5.35 9, 126.2 mmol), H20 (1.20 g, 66.4 mmol), and DMSO (130 mL). The reaction mixture was heated to 185 °C until the malonate was consumed (generally 12-15 hours). After cooling to 0 °C, the reaction was diluted with H20 (125 mL) and diethyl ether (85 mL). The aqueous phase was extracted with diethyl ether (5 x 75 mL) and the organic fractions were combined and washed with H20 (75 mL), saturated NaHCOa (75 mL), and then saturated NaCl (75 mL). The organics were dried over M9804, filtered through a pad of silica gel on top a glass frit, concentrated, and distilled under reduced pressure (bp15 75-85 °C) which provided 11.10 g of 127 as a colorless oil (91% yield); IR (Film) 3079, 2961, 2938, 2874, 1736, 1642 cm"; 1H NMR (300 MHz) (CDCI3) 5 0.87 (t, J: 7 Hz, 3H), 1.23 (t, J- 7 Hz, 3H), 1.25-1.76 (m, 6H), 1.95206 (m, 2H), 2.33 (tt, J= 5, 9 Hz, 1H), 4.11 (q, J: 7 Hz, 2H), 4.84- 5.02 (m, 2H), 5.75 (ddt, J . 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 5 13.96, 14.31, 20.53, 31.58, 34.60, 44.83, 59.99, 114.93, 137.98, 176.30. Two 130 NMR resonances coincided in the spectrum. 2-Propyl-5-hexen-1-ol (128). Compound 127 (5.52 g, 30 mmol) was dissolved in diethyl ether (80 mL) and added dropwise to a 0 °C suspension of LAH (0.80 g, 21 mmol) in diethyl ether (70 mL). Following the usual workup, distillation of the crude residue (bpz7 85-95 °C) provided 4.01 g of 128 (94% yield); IR (Film) 3341, 3079, 2959, 2930, 2874, 1642 cm";1H NMR (300 MHz) (CDCI3) 50.87 (t, J- 7 Hz, 3H), 1.15-1.50 (m, 7H), 1.56 (bs, 1H), 2.05 (tq, J- 1, 7 Hz, 2H), 3.50 (m, 2H), 4.91 (ddt, J=10, 2, 1 Hz, 1H), 4.98 (dq, J: 17, 2 Hz, 1H), 5.78 (ddt, J . 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3) 514.40, 19.91, 30.09, 31.06, 33.09, 39.66, 65.34, 114.32, 139.01. 49 6-Bromo-5-propyl-1-hexene (129). NBS (4.27 g, 24 mmol) was added slowly to a 0 °C solution of 128 (2.84 g, 20 mmol) and PPh3 (6.29 g, 24 mmol) in CH2CI2 (30 mL) according to the general procedure. After a typical workup and distillation, 3.35 g of bromide 129 (82% yield) were recovered (bp19 75-80 °C); IR (Film) 3079, 2959. 2932, 2874, 2859, 1642 cm'1;1H NMR (300 MHz) (coma) 5 0.89 (t, J a 7 Hz, 3H), 1.20-1.56 (rn, 6H), 1.58-1.70 (m, 1H), 1.96-2.16 (m, 2H), 3.42 (dd, J- 5, 11 Hz, 1H), 3.46 (dd, J = 5, 11 Hz, 1H), 4.94 (ddt, J: 10, 2, 1 Hz, 1H), 5.01 (dq, J: 17, 2 Hz, 1H), 5.78 (ddt, J = 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3)514.17, 19.65, 30.74, 31.67, 34.67, 38.44, 39.27, 114.75, 138.39. 3-Bromo-1-cyclohexene (130). A 1-L flask was loaded with NBS (39.14 g, 220 mmol), cyclohexane (16.4 g, 200 mmol), and a catalytic amount of benzoyl peroxide in CCI4 (500 mL) and the mixture was brought to reflux for 4 hours. The reaction mixture was filtered, solvent evaporated, and the crude residue was distilled ([5940 80 °C) to provide 18.13 g of 130 (60% yield); 1H NMR (300 MHz) (CDCI3) 5 1.60-2.30 (m, 6H), 4.70-4.90 (m, 1H), 5.75-5.83 (m, 1H), 5.85-5.93 (m, 1H); 13c NMR (75 MHz) (CDCI3) 518.47, 24.61, 32.65, 48.97, 128.84, 131.03. 3-(2-propenyl)-1-cyclohexene (131). Allylmagnesium bromide was prepared from allyl bromide (24.20 g, 200 mmol) and Mg (19.45 g, 800 mmol) in B20 (180 mL) as previously reported.55 The resultant Grignard solution was determined to have a concentration of 1.07 M by titration with sec-butyl alcohol. The Grignard solution (100 mL, 107 mmol) was transferred to a 250 mL 3-necked flask equipped with a dropping funnel and a reflux condenser. A solution of 130 (7.00 g, 43.5 mmol) in THF (50 mL) was added dropwise to the Grignard solution over 60 minutes at ambient temperature and subsequently heated to reflux for 2-3 hours. After cooling to 0 °C, the solution was quenched with aqueous H2S04 (20 mL H2804 in 50 mL H20). The aqueous layer was extracted with Et20 (4 x 75 mL). The diethyl ether extracts were combined and washed with NaCl (2 x 100 50 mL), dried over Na2804, filtered through silica gel, and concentrated. The crude residue was distilled under reduced pressure (bpgo 70-80 °C) to provide 3.10 g of 131 (60% yield); IR (Film) 3077, 3019, 2928, 1668, 1642 cm"; 1H NMR (300 MHz) (CDCI3) 51.14-1.28 (m, 1H), 1.42-1.58 (m, 1H), 1.64-1.80 (m, 2H), 1.90- 2.00 (m, 2H), 2.00-2.07 (m, 2H), 2.05-2.20 (m, 1H), 4.95-5.05 (m, 2H), 5.52-5.70 (m, 1H), 5.62-5.60 (m, 1H), 5.77 (ddt, J = 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 521.41, 25.27, 28.79, 34.99, 40.64, 115.71, 127.20, 131.38, 137.25. 3-(3-Hydroxypropyl)-1-cyclohexene (132). Dicyclohexyl-borohyride (3.94 g, 22.1 mmol) was loaded into a 250 mL flask, equipped with a dropping funnel, under an atmosphere of nitrogen and diluted with THF (60 mL). Diene 131 (2.70 g, 22.1 mmol) was added dropwise to the solution, contained in a flask immersed in an ambient temperature water bath, and stirred for 5 hours during which time the diene was completely consumed. Oxidation of the trialkylborane was accomplished at 0 °C by the slow sequential addition of 3 M NaOH (7.4 mL) and 30% H202 (7.5 mL). The mixture was heated to 67 °C for 2 hours to facilitate complete oxidation. The solution was cooled to ambient temperature and saturated with K2003. The aqueous layer was extracted with diethyl ether (2 x 25 mL) and the organic fractions were combined and washed with H20 (25 mL), dried over M9804, filtered, and concentrated to an oil. The crude residue contained the expected mixture of cylohexanol and 132. Separation of this mixture was found to be a difficult task; however, selective conversion of the 1° alcohol to the TBDMS ether allowed for trivial separation. To this end, the mixture of alcohols was dissolved in CH2CI2 (75 mL) followed by the addition of DMAP (0.14 g, 1.10 mmol) and TEA (3.35 g, 33.15 mmol). Successive amounts of TBDMSCI were added until complete consumption of the 1° alcohol was achieved. The mixture was washed with H20 (20 mL), NH4CI (20 mL), dried over M9804, filtered, and concentrated to an oil. 51 The TBDMS ether was separated from cyclohexanol by washing the oil down a 6" silica gel column eluted with low-boiling petroleum ether (35-60°C). Removal of the solvent left behind an oil that contained the TBDMS ether of 132 and only a trace amount of the TBDMS ether of cyclohexanol; 13C NMR (75 MHz) (CDCI3) 5 -5.25, 21.52, 25.37, 25.91, 25.98, 29.07, 30.26, 32.43, 34.96, 63.53, 126.83, 132.14. Unmasking of the 1° alcohol was achieved by the addition of the silyl ethers to a 250 mL flask which contained TBAF (60 mL, 1M) in THF (60 mL) at 0 °C. The reaction mixture was diluted with diethyl other (150 mL) and 1M HCI (60 mL). The aqueous layer was extracted with diethyl ether (3 x 100 mL). The combined organic fractions were washed with H20 (25 mL), NaCl (2 x 50 mL), dried over M9804, filtered, and concentrated to an oil. Alcohol 132 was separated from the cyclohexyI-TBDMS ether by filtration of the oil down a 6" column of silica gel, eluted first with petroleum ether (which removed the TBDMS ether) and finally with diethyl ether (which removed the desired alcohol). After evaporation of the diethyl ether, the crude residue was distilled under reduced pressure (bp13 100- 115 °C) which provided 1.60 g of 132 (52% yield for 4 steps); IR (Film) 3339, 3017, 2934, 2859, 1670 cm"; 1H NMR (300 MHz) (CDCla) 51.10-1.88 (m, 9H), 1.90-1.96 (m, 2H), 1.98-2.10 (m, 1H), 3.61 (t, J . 7 Hz, 2H), 5.50-5.58 (m, 1H), 5.60-5.67 (m, 1H); 13C NMR (75 MHz) (CDCI3) 5 21.43, 25.31, 29.00, 30.11, 32.30, 34.90, 63.20, 127.05, 131.77. 3-(3-Bromopropyl)-1-cyclohexene (133). Conversion of 132 to the mesylate was achieved by the dropwise addition of methane sulfonyl chloride (1.26 g, 11 mmol) to a -10 °C solution of 132 (1.40 g, 10 mmol) and TEA (1.52 g, 11 mmol) in CH2CI2 (50 mL). The mixture was allowed to stir for 45 minutes at -10 °C and 90 minutes at ambient temperature. The mixture was diluted with CH2CI2 (50 mL) and washed with H20 (20 mL), 10% HCI (30 mL), saturated 52 NaHC03 (30 mL), and saturated NaCl (30 mL). The organic solution was dried over M9804, filtered, and concentrated to an oil. The mesylate was added to a -10 °C solution of LiBr (1.74 9, 20 mmol) in THF (30 mL). Stirring was continued for 30 minutes and the reaction was warmed to ambient temperatrure for 18 hours. After complete consumption of the mesylate, the mixture was filtered through basic alumina, eluted with diethyl ether, and concentrated to an oil. The . oil was redissolved in diethyl ether (60 mL) and sequentially washed with saturated NaHC03 (20 mL), saturated NaCl (20 mL), dried over M9804, filtered, and concentrated to an oil. The crude residue was distilled (bpio 90-100°C) and 1.40 g of 133 were recovered (70% yield); IR (Film) 3017, 2930, 2856, 1672. 1449, 1437 cm“; 1H NMR (300 MHz) (coma) 51.14-1.26 (m, 1H), 1.29-1.60 (m, 3H), 1.64-1.82 (m, 2H), 1.82-2.00 (m, 4H), 1.98-2.14 (m, 1H), 3.38 (t, J = 7 Hz, 2H), 5.47-5.56 (m, 1H), 5.61-5.70 (m, 1H); 130 NMR (75 MHz) (CDCI3) 5 21.37, 25.26, 28.89, 30.29, 34.14, 34.48, 34.78, 127.37, 131.31. Diethyl 2-(pent-4-enyl)-propanedloate (134). The following reagents were combined in the usual fashion: diethylmalonate (10.75 g, 67.2 mmol), NaH (1.69 9, 70.6 mmol) in DMF (60 mL), and 5-bromo-1-pentene (12.0 9, 80.6 mmol). Coupling was achieved by heating the mixture at 65 °C for 24 hours. After workup and distillation under reduced pressure (bps 112-115 °C), 12.71 9 of 134 were recovered (83% yield): IR (Film) 2982, 2938, 1755, 1734, 1642, 1460, 1446 cm"; 1H NMR (300 MHz) (CDCI3) 51.23 (t, J =- 7 Hz, 6H), 1.40 (tt, J . 8, 7 Hz, 2H), 1.88 (q, J - 8 Hz, 2H), 2.05 (q. J a 7 Hz, 2H), 3.29 (t, J a 8 Hz, 1H), 4.16 (q, J a 7 Hz, 4H), 4.90-5.15 (m, 2H), 5.75 (ddt, J = 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3) 5 14.05, 26.54, 28.17, 33.24, 51.91, 61.23, 114.95, 137.92, 169.42. Ethyl-6-heptenoate (135). Compound 134 (8.75 9, 38.4 mmol) was combined with LiCl (3.09 9, 72.9 mmol) and H20 (0.69 g, 38.4 mmol) in DMSO (80 mL) and heated to 185 °C for 18 hours. After the usual workup, 5.16 9 of 53 ester 135 were obtained (86% yield) after distillation (bp22 72-75 °C); IR (Film) 3079, 2980, 2936, 2865, 1738, 1642, 1462, 1446, 1420 cm“; 1H NMR (300 MHz) (CDCI3) 5 1.22 (t, J: 7 Hz, 3H), 1.39 (quint, J: 7 Hz, 2H), 1.61 (quint, J: 7 Hz, 2H), 2.03 (q, J: 7 Hz, 2H), 2.27 (t, J = 7 Hz, 2H), 4.09 (q, J: 7 Hz, 2H), 4.88-5.02 (m, 2H), 5.76 (ddt, J = 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3) 514.21, 24.40, 28.33, 33.32, 34.16, 60.13, 114.60, 138.39, 173.63. 6-Hepten-1-ol (136). Compound 135 (4.70 9, 30.1 mmol) was dissolved in diethyl ether (50 mL) and added dropwise to a suspension of LAH (0.80 9, 21.1 mmol) in diethyl ether (100 mL) as described in the general procedure. Following the usual workup, distillation of the crude residue (bp27 75-85°C) provided 3.15 g of 136 (92% yield); IR (Film) 3337, 3079, 2998, 2978, 2936, 2861, 1642, 1462, 1437 cm"; 1H NMR (300 MHz) (CDCI3) 51.25-1.43 (m, 4H), 1.53 (quint, J: 7 Hz, 2H), 1.86-1.96 (bs, 1H), 2.02 (q, J : 7 Hz, 2H), 3.58 (t, J : 6 Hz, 2H), 4.82-5.01 (m, 2H), 5.76 (ddt, J : 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3) 5 25.17, 28.62, 32.51, 33.64, 62.75, 114.29, 138.79. 7-Bromo-1-heptene (137). The following reagents were combined the according to the general procedure: alcohol 136 (1.50 9, 13.2 mmol), PPh3 (4.14 9, 15.8 mmol), NBS (2.81 9, 15.8 mmol), and CH2CI2.(20 mL). After distillation (bp44 77-81 °C), 2.08 9 of bromide 137 were recovered as a colorless oil (89% yield); IR (Film) 3077, 3000, 2977, 2961, 2938, 2859, 1642, 1460, 1439 cm“; 1H NMR (300 MHz) (CDCI3) 5 1.32-1.52 (m, 4H), 1.78-1.90 (m, 2H), 1.98-2.10 (m, 2H), 3.38 (t, J: 7 Hz, 2H), 4.90-5.02 (m, 2H), 5.78 (ddt, J: 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (cool-3) 8 27.62, 28.01, 32.66, 33.50, 33.75, 114.57, 138.55. Ethyl 3-Methyl-4-Pentenoate (138). In a 100 mL 3-necked flask equipped with a thermometer and a vigreux column with distillation head, were added crotyl alcohol (10 g, 139 mmol), triethyl orthoacetate (67.65 9, 417 mmol), and a catalytic amount of trimethyl acetic acid. This mixture was slowly heated to 145 54 °C which was maintained until a quanitatlve amount of ethanol was collected. After cooling the reaction to ambient temperature, H20 (5.5 g, 306 mmol) was added and stirred for 4-6 hours. Heating was resumed to 110 °C until all of the ethanol and ethyl acetate had been collected, at which time the mixture was heated to distill the ester (138; bp7so 152-154 °C) which provided 11.71 9 of 138 (63% yield); IR (Film) 2980, 2936, 2874, 1738, 1642, 1462, 1420 cm"; 1H NMR (300 MHz) (CDCI3)51.03 (d, J : 7 Hz, 3H), 1.22 (t, J : 7 Hz, 3H), 2.22 (dd, J a 15, 7 Hz, 1H), 2.32 (dd, J . 15, 7 Hz, 1H), 2.65 (m, 1H), 4.10 (q, J = 7 Hz, 2H), 4.93 (dt, J: 10, 2 Hz, 1H), 4.99 (dt, J: 17, 2 Hz, 1H), 5.74 (ddd, J: 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCl3) 514.25, 19.68, 34.42, 41.34, 60.22, 113.27, 142.49, 172.55. 3-Methyl-4-penten-1-ol (139). A solution of 138 (11.0 9, 77.5 mmol, 220 mL Et20) was added to a suspension of LAH (2.06 9, 54.2 mmol) in B20 (170 mL). After the usual workup, 139 was distilled (bp7so 146-148 °C) and 5.35 9 were collected (69% yield); IR (Film) 3339, 3081, 2963, 2932, 2876, 1640, 1454, 1420 cm"; 1H NMR (300 MHz) (CDCI3) 51.00 (d, J = 7 Hz, 3H), 1.42-1.48 (bs, 1H), 1.50-1.60 (m, 2H), 2.28 (sept, J = 7 Hz, 1H), 3.63 (t, J: 7 Hz, 2H), 4.92 (dt, J a 10, 2 Hz, 1H), 4.98 (dt, J:17, 2 Hz, 1H), 5.69 (ddd, J: 17, 10, 7 Hz, 1H);13C NMR (75 MHz) (CDCI3) 5 20.38, 34.89, 39.27, 61.21, 113.10, 144.23. 1-Bromo-3-methyl-4-pentene (140). The following reagents were combined according to the general procedure: 139 (5.15 g, 51.5 mmol), NBS (10.99 g, 61.8 mmol), PPh3 (16.19 9, 61.8 mmol), and CH2CI2 (65 mL). After distillation (bp75 72-74 °C), 6.29 g of bromide 140 were recovered as a colorless oil (75% yield); IR (Film) 3079, 2969, 2930, 2869, 1642, 1455, 1435, 1418 cm"; 1H NMR (300 MHz) (CDCI3)51.01 (d, J : 7 Hz, 3H), 1.76-1.86 (m, 2H), 2.35 (sept, J: 7 Hz, 2H), 3.28-3.45 (m, 2H), 4.94-5.07 (m, 2H), 5.61 (ddd, J: 18, 10, 7 55 Hz, 1H); 13C NMR (75 MHz) (CDCI3)819.91, 31.92, 36.54, 39.21, 114.15, 142.59. Dlethyl 2-(3-methyI-4-pentenyI)-propanedIoate (141). Diethyl malonate (6.48 9, 40.5 mmol) was combined with NaH (1.01 9, 42.4 mmol) in DMF (32 mL) and subsequently reacted with 140 (5.70 g, 34.6 mmol) as described in the general procedure. After the usual workup, the residue was distilled under reduced pressure (bpd 80-90 °C) which provided 8.02 g of 141 (95% yield); IR (Film) 3079, 2980, 2965, 2938, 2911, 1752, 1734, 1642, 1464 cm"; 1H NMR (300 MHz) (coma) 8 0.97 (d, J .. 7 Hz, 3H), 1.23 (t, J =- 7 Hz, 6H), 1.19-1.35 (m, 2H), 1.75-1.95 (m, 2H), 2.00-2.20 (m, 1H), 3.26 (t, J = 7 Hz, 1H), 4.10-4.22 (m, 4H), 4.86-4.99 (m, 2H), 5.63 (ddd, J: 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3) 5 14.05, 20.09, 26.52, 33.98, 37.57, 52.05, 61.25, 113.18, 143.76, 169.47, 169.50. Ethyl 5-methyI-6-heptenoate (142). The following were combined and reacted according to the usual procedure: 141 (7.40 g, 30.6 mmol), LiCl (2.53 g, 59.7 mmol), H20 (0.57 g, 31.4 mmol) and DMSO (65 mL). After workup and distillation (bp15 80-85 °C), 4.12 9 of ester 142 were recovered as a colorless oil (80% yield); IR (Film) 3079, 2963, 2934, 2911, 2872, 1740, 1642, 1456, 1420 cm' 1; 1H NMR (300 MHz) (CDCla) 8 0.96 (d, J = 7 Hz, 3H), 122 (t, J . 7 Hz, 3H), 1.22-1.36 (m, 2H), 1.50-1.70 (m, 2H), 2.10 (sept, J : 7 Hz, 1H), 2.25 (t, J : 7 Hz, 2H), 4.09 (q, J - 7 Hz, 2H), 4.84-4.99 (m, 2H), 5.65 (ddd, J a 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 514.23, 20.13, 22.70, 34.39, 35.96, 37.56, 60.18, 112.80, 144.26, 173.77. 5-MethyI-6-hepten-1-ol (143). Following the usual procedure, ester 142 (3.73 g, 21.9 mmol) was reduced with LAH (0.63 9, 16.5 mmol) in 820 (120 mL). After workup and distillation (bpao 90-100 °C), 2.41 9 of the alcohol were collected (86% yield); IR (Film) 3337, 3079, 2934, 2865, 1640, 1458, 1437, 1420 cm‘1;1H NMR (300 MHz) (00013) 8 0.95 (d, J :7Hz, 3H), 1.15-1.40 (m, 4H), 1.40-1.60 (m, 56 2H), 1.5-1.8 (bs, 1H), 2.10 (sept, J = 7 Hz, 1H), 3.59 (t, J = 7 Hz, 2H), 4.82-4.98 (m, 2H), 5.65 (ddd, J = 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCl3) 8 20.17, 23.38, 32.82, 36.35, 37.77, 62.95, 112.50, 144.66. 7-Bromo-3-methyl-1-heptene (144). According to the usual bromination procedure, alcohol 143 (2.0 g, 15.6 mmol), NBS (3.34 g, 18.8 mmol), and PPh3 (4.91 9, 18.8 mmol) were combined and reacted in CH2CI2 (20 mL). After a typical workup, 2.54 g of bromide 144 were recovered (85% yield) after distillation (0914 68-70 °C); IR (Film) 3079, 2963, 2936, 2865, 1642, 1456, 1439, 1431, 1420 cm"; 1H NMR (300 MHz) (CDCI3) 8 0.97 (d, J = 7 Hz, 3H), 1.20-1.50 (m, 4H), 1.83 (quint, J : 7 Hz, 2H), 2.10 (sept, J : 7 Hz, 1H), 3.38 (t, J : 7 Hz, 2H), 4.90 (ddd, J: 10, 2, 1 Hz, 1H), 4.94 (ddd, J: 17, 2, 1 Hz, 1H), 5.66 (ddd, J: 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 5 20.18, 25.84, 32.88, 33.89, 35.65, 37.64, 112.72, 144.39. Ethyl 2-methyI-4-pentenoate (145). The ortho ester Claisen rearrangement was performed as described for 138. Allyl alcohol (9.28 9, 160 mmol), triethyl orthopropionate (84.61 9, 480 mmol), and trimethyl acetic acid (1.64 9, 16 mmol) were combined and heated to 125 °C. After the reaction was quenched as in the usual fashion, 17.85 g of ester 145 were collected by distillation at atmospheric pressure (bp7eo 154 °C, 79% yield); IR (Film) 3079, 2980, 2940,2911, 1736, 1644, 1462 cm"; 1H NMR (300 MHz) (CDCI3) 51.12 (d, J : 7 Hz, 3H), 1.22 (t, J : 7 Hz, 3H), 2.08-2.22 (m, 1H), 2.32-2.54 (m, 2H), 4.10 (q, J : 7 Hz, 2H), 4.95-5.04 (m, 2H), 5.72 (ddt, J : 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3)514.22, 16.51, 37.78, 39.21 , 60.20, 116.73, 135.49, 176.10. 2-Methyl-4-penten-1-ol (146). Ester 145 (14.97 9, 105.4 mmol) was reduced with LAH (2.91 9, 76.7 mmol) in B20 (530 mL) by the usual procedure. After atmospheric distillation (bp 146 °C), 8.24 9 of 145 was obtained as a colorless oil (78% yield); IR (Film) 3341, 3079, 2974, 2959, 2919, 2876, 1642, 57 1458, 1441, 1416 cm"; 1H NMR (300 MHz) (CDCI3) 8 0.90 (d, J:7 Hz, 3H), 1.51 (bs, 1H), 1.71 (oct, J: 7 Hz, 1H), 1.86-1.98 (m, 1H), 2.09-2.20 (m, 1H), 3.43 (dd. J: 6, 11 Hz, 1H), 3.49 (dd, J: 6, 11 Hz, 1H), 4.85-5.05 (m, 2H), 5.78 (ddt, J: 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3) 5 16.34, 35.59, 37.80, 67.86, 116.04. 136.95. 5-Bromo-4-methyI-1-pentene (147). According to the usual bromination procedure, alcohol 146 (7.59 9, 75.9 mmol) was reacted with NBS (16.65 9, 93.6 mmol) and PPh3 (24.52 g, 93.6 mmol) in CH2CI2 (100 mL). Distillation of the crude residue at reduced pressure (bpyo 76-82 °C) furnished 7.85 g of 147 (64% yield); IR (Film) 3079, 2965,2930, 2913, 2874, 1642, 1458, 1437 cm";1H NMR (300 MHz) (CDCI3) 51.01 (d, J: 7 Hz, 3H), 1.87 (oct, J: 7 Hz, 1H), 1.96-2.07 (m, 1H), 2.13-2.25 (m, 1H), 3.31 (dd, J : 6, 10 Hz, 1H), 3.37 (dd, J : 6, 10 Hz, 1H), 5.00-5.10 (m, 2H), 5.72 (ddt, J: 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3) 8 18.54, 35.01, 39.06, 40.62, 116.98, 135.76. Dlethyl 2-(2-methyl-4-pentenyl)-propanedIoate (148). The sodium enolate of diethyl malonate, prepared by reacting diethyl malonate (7.89 g, 49.3 mmol) and NaH (1.18 g, 49.3 mmol) in DMF (40 mL), was coupled with 147 (7.24 g, 44.8 mmol) in the usual manner. After workup and distillation (bpd 80-90 °C), 9.40 g of diester 148 was isolated (87% yield). This compound was carried on without characterization. Ethyl 4-methyl-6-heptenoate (149). By following the general procedure, 148 (9.40 9, 38.8 mmol) was deethoxycarboxylated by the combined action of LiCl (3.13 g, 73.8 mmol) and H20 (0.73 g, 40.7 mmol) in DMSO (80 mL) at 185 °C. After the usual workup, the crude residue was distilled under reduced pressure (bp"; 80-85 °C) which provided 4.95 9 as a colorless oil (66% yield for 2 steps): IR (Film) 3079, 2978, 2961, 2932, 2874, 1740, 1642, 1462, 1449 cm“; 1H NMR (300 MHz) (00013) 8 0.87 (d, J : 7 Hz, 3H), 1.23 (t, J : 7 Hz, 3H), 1.37-1.57 (m, 58 2H), 1.61-1.74 (m, 1H), 1.82-1.95 (m, 1H), 1.98-2.10 (m, 1H), 2.19-2.37 (m; 2H), 4.10 (q, J : 7 Hz, 2H), 4.95-5.05 (m, 2H), 5.74 (ddt, J : 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3)514.20, 19.02, 31.38, 32.11, 32.33, 41.00, 60.20, 115.94, 136.97, 173.95. 4-Methyl-6-hepten-1-ol (150). A solution of ester 149 (4.40 g, 25.9 mmol) was added to a 0 °C suspension of LAH (0.69 g, 18.1 mmol) in diethyl other (130 mL). The reaction mixture was purified in the usual manner which yielded 3.08 g (93 % yield ) of 150 after distillation (bpao 90-95 °C); IR (Film) 3335, 3077, 2955, 2934, 2872, 1642, 1458, 1441, 1416 cm"; 1H NMR (300 MHz) (CDCI3) 5 0.85 (d, J: 7 Hz, 3H), 1.04-1.22 (m, 1H), 1.28-1.64 (m, 4H), 1.73 (bs, 1H), 1.80-1.92 (m, 1H), 1.98-2.09 (m, 1H), 3.58 (t, J: 6 Hz, 2H), 4.92-5.00 (m, 2H), 5.74 (ddt, J: 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 5 19.36, 30.28, 32.38, 32.59, 41.31. 63.28, 1 15.65, 137.42. 7-Bromo-4-methyl-1-heptene (151). Alcohol 150 (1.46 g, 11.7 mmol), NBS (2.50 9, 14.1 mmol), and PPh3 (3.68 g, 14.1 mmol) were combined in CH2CI2 (15 mL) in the usual manner. After workup and distillation (bp14 68-70 °C), 1.92 g of bromide 23 (86% yield) was obtained; IR (film) 3079, 3002, 2961, 2913, 2874, 2855, 1642, 1460, 1439 cm"; 1H NMR (300 MHz) (CDCI3) 8 0.87 (d, J: 7 Hz, 3 H), 1.14-1.32 (m, 1H), 1.36-1.60 (m, 2H), 1.72-1.96 (m, 3H), 1.97-2.12 (m, 1H), 3.37 (t, J : 7 Hz, 2H), 4.93-5.03 (m, 2H), 5.66-5.83 (m, 1H); 13C NMR (75 MHz) (CDCI3) 519.34, 30.52, 32.20, 34.16, 34.95, 41.20, 115.89, 137.14. Dlethyl 2-(1-methyI-4-pentenyl)-propanedIoate (152). The following reagents were combined as described in the general procedure: diethyl malonate (12.97 9, 81 mmol), NaH (1.97 9, 82.5 mmol), 5-bromo-1-hexene (12.00 9, 73.7 mmol) and DMF (50 mL). After a typical workup procedure, the crude residue was distilled (bpd 80-90 °C) and 15.86 9 of 152 was recovered (81% yield); IR (Film) 3079, 2982, 2938, 2878, 1736, 1642, 1466, 1449 cm"; 1H NMR (300 MHz) 59 (00013) 8 0.97 (d, J : 7 Hz, 3H), 1.18-1.35 (m, 1H), 1.24 (t, J : 7 Hz, 6H), 1.42- 1.58 (m, 1H), 1.92-2.35 (m, 3H), 3.22 (d, J : 8 Hz, 1H), 4.17 (q, J : 7 Hz, 4H), 4.93 (ddt, J: 10, 2, 1 Hz, 1H), 4.99 (dq, J: 17, 2 Hz, 1H), 5.76 (ddt, J: 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 5 14.11, 16.76, 31.04, 32.81, 33.50, 57.66, 61.06, 61.12, 114.73, 138.21, 168.72, 168.86. Ethyl 3-methyI-6-heptenoate (153). The following reagents were combined and heated to 185 °C for 15 hours: 152 (14.75 9, 62 mmol), LiCl (4.99 9, 117.8 mmol), H20 (1.12 9, 62 mmol), and DMSO (115 mL). After the usual workup, distillation of the crude residue under reduced pressure (bp12 86-90 °C) furnished 8.54 g of 153 (81% yield); IR (Film) 3079, 2979, 2932, 2876, 2853, 1738, 1642, 1460, 1418 cm'1;‘H NMR (300 MHz) (CDCI3) 8 0.92 (d, J : 6 Hz, 3H), 1.18-1.48 (m, 2H), 1.24 (t, J : 7 Hz, 3H), 1.88-2.14 (m, 4H), 2.27 (dd, J : 6, 14 Hz, 1H), 4.10 (q, J : 7 Hz, 2H), 4.86-5.04 (m, 2H), 5.77 (ddt, J: 17, 10, 7 Hz. 1H); 13C NMR (75 MHz) (CDCI3) 514.26, 19.51, 29.86, 31.13, 35.82, 41.79, 60.07, 114.44, 138.60, 173.11. 3-MethyI-6-hepten-1-ol (154). Ester 153 (7.75 9, 46.5 mmol) was reduced with LAH (1.23 g, 32.5 mmol) in 320 (235 mL) following the general procedure. The alcohol was purified by distillation (bp25 90-95 °C) and 5.74 9 (98% yield) of 154 was obtained as a colorless oil; IR (Film) 3333, 3079, 2961, 2928, 2874, 1642, 1458 cm"; 1H NMR (300 MHz) (CDCI3) 8 0.87 (d, J: 7 Hz, 3H), 1.14-1.30 (m, 1H), 1.26-1.40 (m, 2H), 1.46-1.65 (m, 3H), 1.92-2.14 (m, 2H), 3.55-3.72 (m, 2H), 4.90 (ddt, J: 10, 2, 1 Hz, 1H), 4.97 (dq, J: 17,2 Hz, 1H), 5.77 (ddt, J: 17, 10, 6 Hz, 1H); 13C NMR (75 MHz) (CDCI3)519.41, 29.00, 31.18, 36.23, 39.79, 61.02, 114.17, 139.03. 7-Bromo-5-methyI-1-heptene (155). According to the general procedure, NBS (5.00 g, 28.1 mmol) was combined with a solution of 154 (3.00 g, 23.4 mmol) and PPh3 (7.37 g, 28.1 mmol) in CH2CI2 (30 mL). The usual workup was 60 performed, and the crude residue was purified by distillation (bp17 78-80 °C) which provided 5.74 g of bromide 155 (88% yield); IR (Film) 3079, 2965, 2930, 2872, 1642, 1458 cm"; 1H NMR (300 MHz) (CDCI3) 5 0.89 (d, J: 6 Hz, 3H), 1.22 (dddd, 1H), 1.41 (dddd, 1H), 1.56-1.72 (m, 2H), 1.78-1.94 (m, 1H), 1.94-2.15 (m, 2H), 3.33-3.48 (m, 2H), 4.93 (ddt, J: 10, 2, 1 Hz, 1H), 4.99 (dq, J: 17, 2 Hz, 1H), 5.78 (ddt, J : 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 518.76, 31.05. 31.18, 31.91, 35.64, 39.94, 114.42, 138.74. Dlethyl 2-(1-methylethyI)-propanedloate (156). The following reagents were combined and reacted as described in the general procedure; diethyl malonate (10.40 g, 65 mmol), NaH (1.71 9, 71.5 mmol), 2-bromopropane (10.30 g, 84.50 mmol) and DMF (55 mL). After workup and distillation (bp<1 50-55°C). 12.46 9 of 156 was recovered (95% yield); IR (Film) 2971, 2942, 2909, 1736, 1468, 1448 cm"; 1H NMR (300 MHz) (CDCI3) 5 0.94 (d, J: 7 Hz, 6H), 1.21 (t, J: 7 Hz, 6H), 2.25-2.42 (m, 1H), 3.05 (d, J : 9 Hz, 1H), 4.13 (q, J: 7 Hz, 4H); 13C NMR (75 MHz) (00013) 514.03, 20.28, 28.64, 59.02, 60.99, 168.75. Dlethyl 2-(1-methylethyl)-2-(4-pentenyl)-propanedIoate (157). The following reagents were combined and reacted as described in the general procedure; 156 (13.00 9, 64.40 mmol), NaH (1.69 9, 70.8 mmol), 5-bromopentene (11.99 g, 84.50 mmol) and DMF (60 mL). After workup and distillation (bpz1 80- 85 °C), 16.22 g of 29 was recovered (93% yield); 1H NMR (300 MHz) (CDCI3) 5 0.94 (d, J : 7 Hz, 6 H), 1.23 (t, J : 7 Hz, 6H), 1.18-1.35 (m, 2H), 1.78-1.90 (m, 2H), 2.02 (tq, J : 1, 7 Hz, 2H), 2.29 (sept, J: 7 Hz, 1H), 4.16 (q, J : 7 Hz, 4 H), 4.92 (ddt, J: 10, 2, 1 Hz, 1H), 4.97 (dq, J: 17,2 Hz, 1H), 5.74 (ddt, J: 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3) 514.12, 18.55, 23.92, 31.87, 33.19, 33.97, 60.59, 61.57, 114.78, 138.19, 171.17. Ethyl 2-(1-methylethyI)-6-heptenoate (158). Diester 157 (15.79 g, 58.5 mmol) was treated with LiCI (4.71 9, 111.2 mmol) and H20 (1.05 g, 58.5 mmol) in 61 DMSO (120 mL) and heated to 185 °C for 15 hours. After a typical workup, the crude residue was distilled (bp2o 105-110 °C) and 8.08 9 of ester 158 was recovered (70% yield); IR (Film) 2963, 2940, 2874, 1734, 1642, 1462, 1447 cm’1; 1H NMR (300 MHz) (CDCI3) 5 0.87 (d, J : 7 Hz, 3H), 0.89 (d, J: 7 Hz, 3H), 1.22 (t, J : 7 Hz, 3H), 1.25-1.39 (m, 2H), 1.40-1.64 (m, 2H), 1.81 (oct, J : 7 Hz, 1H), 1.952.10 (m, 3H), 4.10 (q, J: 7 Hz, 2H), 4.86-5.00 (m, 2H), 5.74 (ddt, J: 17, 10, 7 Hz, 1H); ‘30 NMR (75 MHz) (CDCI3) 5 14.34, 20.19, 20.47, 26.98, 29.09, 30.63. 33.63, 52.63, 59.80, 114.54, 138.51, 175.80. 2-(1-MethylethyI)-6-hepten-1-ol (159). Ester 158 (7.66 g, 39.9 mmol) was reduced with LAH (1.06 9, 27.9 mmol) in B20 (200 mL) by the usual procedure. The alcohol was purified by distillation (bp1o 90-105 °C) and 5.62 g (93% yield) of 159 was obtained as a colorless oil; IR (Film) 3343, 3079, 2959, 2932, 2874, 1642, 1464, 1443, 1416 cm'1;1H NMR (300 MHz) (CDCI3) 8 0.85 (d, J : 7 Hz, 3H), 0.87 (d, J : 7 Hz, 3H), 1.15-1.50 (m, 6H), 1.70-1.88 (m, 1H), 1.96-2.07 (m, 2H), 3.44-3.60 (m, 2H), 4.85-5.05 (m, 2H), 5.78 (ddt, J: 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (c0013) 519.18, 19.75, 27.13, 27.18, 27.83, 34.11, 46.46, 63.65. 114.35, 138.86. 7-Bromo-6-(1-methylethyI)-1-heptene (160). According to the general procedure, NBS (4.119, 23.1 mmol) was combined with a solution of 159 (3.00 g, 19.2 mmol) and PPh3 (6.05 9, 23.1 mmol) in CH2CI2 (25 mL). The usual workup was performed and the ande residue was purified by distillation (bpg 90-95 °C) which provided 3.94 9 of bromide 160 (94% yield); IR (film) 3079, 2963, 2934, 2872, 1642, 1462, 1441, 1416 cm"; 1H NMR (300 MHz) (CDCI3) 8 0.88 (d, J : 7 Hz, 3H), 0.89 (d, J : 7 Hz, 3H), 1.25-1.50 (m, 5H), 1.79 (oct, J : 7 Hz, 1H), 1.95- 2.10 (m, 2H), 3.44 (m, 2H), 4.94 (ddt, J : 10, 2, 1 Hz, 1H), 5.00 (dq, J : 17, 2 Hz, 1H), 5.79 (ddt, J: 17, 10,7 Hz, 1H); 13C NMR (75 MHz) (CDCla) 519.17, 19.76, 26.65, 28.78, 29.19, 33.85, 37.41, 45.98, 114.57, 138.62. 62 4-(1-MethylethyI)-cylohexanone (1 61). To a solution of 4-(1-methylethyl)- cyclohexanol (12 9, 85.7 mmol) in diethyl ether (135 mL) was added 48 mL of a chromic acid solution. The two layers were separated, and the aqueous phase was extracted with diethyl ether (3 x 100 mL). The organic fractions were combined and washed with saturated NaHC03 (100 mL), saturated NaCl (100 mL), dried over M9804, filtered through a pad of silica gel, eluted with diethyl ether, and the solvent was removed by rotary evaporation. Distillation at reduced pressure (bpzo 80-90 °C) provided 10.51 g of 161 (89% yield); IR (Film) 2961, 2932, 2872, 1721 cm";1H NMR (300 MHz) (00013) 8 0.87 (d, J : 7 Hz, 6H), 1.30-1.60 (m, 4H), 1.90-2.02 (m, 2H), 2.19-2.40 (m, 4H); 13C NMR (75 MHz) (c0013) 819.89, 29.56, 31.69, 40.99, 42.44, 212.42. 4-(1-MethylethyI)-caprolactone (162). To an ambient temperature solution of ketone 33 (9.0 9, 64.3 mmol), NaHC03 (9.44 9, 112.5 mmol) in CH2CI2 (425 mL) was added mCPBA (22.83 9, 112.5 mmol) over 2 hours, and the resultant solution was stirred for an additional 48 hours. After complete consumption of the ketone, the solution was washed with a 1N NaHC03 (270 mL), saturated NaCI (135 mL), and the solvent was removed by rotary evaporation. Lactone 34 was distilled under reduced pressure (bp2o 145-150 °C) using a Kugelrohr apparatus to provide 8.22 9 of 162 (82% yield); IR (Film) 3019, 2963, 2934, 2874, 1734, 1719 cm"; 1H NMR (300 MHz) (00013) 8 0.84 (d, J: 7 Hz, 6H), 1.28-1.64 (m, 4H), 1.75-1.95 (m, 2H), 2.48-2.70 (m, 2H), 4.11 (dd, J: 9. 2 Hz, 1H), 4.21-4.32 (m, 1H); 13C NMR (75 MHz) (CDCIa) 519.18, 19.29, 25.60, 31.93, 32.47, 33.31, 46.48, 68.42, 176.14. 3-(1-methylethyl)-6-hepten-1-ol (163). Methyltriphenylphosphonium bromide (10.31 9, 28.85 mmol) was loaded into a Schlenk flask, which was then evacuated, purged with argon, suspended in THF (135 mL), and cooled to -20 °C. nBuLi (11.15 mL, 2.5 M solution in hexane) was added to the \Mttig salt, stirred for 63 1 hour, and cooled to -50 °C. In a separate flask, a solution of Iactone 162 (3.0 9, 19.23 mmol) in toluene (100 mL) was cooled to -78 °C followed by the addition of a DiBAI solution (23.1 mL, 1.0 M in toluene) over 10 minutes and stirred for 1 hour. The lactol was warmed to 0 °C and transferred, via cannula, to the preformed ylide. The reaction temperature was maintained for 10 minutes at -50 °C, warmed to 0 °C for 30 minutes, and finally allowed to warm to ambient temperature. After 30 minutes at ambient temperature, an aqueous solution of Na2804-H20 (6.3 g) in H20 (50 mL) was added to the reaction. The two layers were separated and the aqueous phase was extracted with diethyl ether (4 x 80 mL). The organic fractions were combined and washed with saturated NaCl (2 x 50 mL) dried over M9804, and the organic solution was filtered through silica gel. The solvent was removed by rotary evaporation and the residue was purified by flash chromatography (silica gel 60 mesh, 8:2 diethyl etherzpetroleum ether; R(:0.51). The solvent was removed and the residue was distilled (bp7 92-100° C) to give 650 m9 163 (22% yield); IR (Film) 3337, 3079, 2959, 2932, 2872, 1642, 1466 cm"; 1H NMR (300 MHz) (CDCI3) 5 0.82 (d, J: 7 Hz, 3H), 0.83 (d, J: 7 Hz, 3H), 1.13-1.48 (m, 5H), 1.50-1.75 (m, 2H), 1.98-2.07 (m, 2H), 3.55-3.70 (m, 2H), 4.91 (ddt, J: 10, 2, 1 Hz, 1H), 4.98 (dq, J: 17, 2 Hz, 1H), 5.78 (ddt, J: 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 518.69, 19.15, 29.28, 30.03, 31.80, 33.61, 39.60, 61.81, 114.27, 139.14. 7-Bromo-5-(1-methylethyl)-1-heptene (164). Alcohol 163 (500 mg, 3.2 mmol), NBS (684 mg, 3.85 mmol), PPh3 (1000 mg, 3.85 mmol), in CH2CI2 (5 mL) were combined according to the general procedure. Following the usual workup, the crude residue was distilled (bp11 88-92 °C) and 620 mg were collected (89% yield); IR (Film) 3079, 2961, 2932, 2872, 1642, 1466, 1439, 1416 cm";1H NMR (300 MHz) (CDCI3) 5 0.83 (d, J: 7 Hz, 6H), 1.14-1.44 (m, 4H), 1.54-1.92 (m, 4 H), 1.94-2.10 (m, 2 H), 3.39 (m, 2H), 4.94 (ddt, J: 10, 2, 1 Hz, 1H), 4.99 (dq, J: 17, 64 2 Hz, 1H), 5.78 (ddt, J: 17, 10, 7 Hz, 1H); 13C NMR (75 MHZ) (CDCI3) 5 18.77, 18.97, 28.80, 29.36, 31.62, 32.76, 34.07, 41 .96, 114.50, 138.81. Dlethyl 2-(ethyI)-2-(4-pentenyl)-propanedIoate (165). The following reagents were combined as outlined in the general procedure: diethyl ethylmalonate (37.64 g, 200 mmol), NaH (5.03 g, 210 mmol) in DMF (200 mL), and 5-bromo-1-pentene (32.16 g, 216 mmol). Coupling was achieved by heating the mixture at 50 °C for 40 hours. After workup and distillation under reduced pressure (bpd 90-100°C), 50.08 g of 165 were recovered (98% yield); IR (Film) 2980, 2942, 2884, 1732, 1642 cm"; 1H NMR (300 M2) (CDCI3) 8 0.78 (t, J : 7 Hz, 3H), 1.21 (t, J: 7 Hz, 6H), 1.14-1.28 (m, 2H), 1.76-1.95 (m, 4H), 2.03 (tq, J: 2, 7 Hz, 2H), 4.15 (q, J: 7 Hz, 4H), 4.88-5.01 (m, 2H), 5.75 (ddt, J: 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 5 8.39, 14.09, 23.24, 25.17, 31 .06, 33.80, 57.82, 60.94, 114.89, 138.13, 171.82. Ethyl 2-ethyl-6-heptenoate (166). Compound 165 (50.0 9, 195.3 mmol) was combined with LiCl (15.73 9, 371.1 mmol) and H20 (3.51 9, 195.3 mmol) in DMSO (400 mL) and heated to 180 °C for 18 hours. After the usual workup, 31.61 9 of ester 166 were obtained (88% yield) after distillation (bp1o 105-115 °C); IR (Film) 3079, 2967, 2934, 1734, 1642 cm"; 1H NMR (300 MHz) (CDCI3) 8 0.86 (t, J: 7 Hz, 3H), 1.23 (t, J: 7 Hz, 3H), 1.27-1.68 (m, 6H), 2.02 (bq, J: 7 Hz, 2H), 2.23 (tt, J: 5, 9 Hz, 1H), 4.11 (q, J : 7 Hz, 2H), 4.87-5.02 (m, 2H), 5.75 (ddt, J : 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 511.80, 14.34, 25.48, 26.65. 31.49, 33.61, 47.18, 59.97, 114.58, 138.50, 176.29. 2-EthyI-6-hepten-1-ol (167). Compound 166 (10.12 g, 55 mmol) was dissolved in diethyl other (150 mL) and added dropwise to a 0 °C suspension of LAH (1.46 g, 38.5 mmol) in diethyl other (125 mL). Following the usual workup, distillation of the crude residue (bp22 90-100 °C) provided 7.39 g of 167 (95% yield); IR (film) 3337, 3079, 2963, 2876, 1642 cm"; 1H NMR (300 MHz) (CDCI3) 65 8 0.87 (t, J : 7 Hz, 3H), 1.15-1 .45 (m, 8H), 2.02 (bq, J : 7 Hz, 2H), 3.51 (m, 2H), 4.88-5.02 (m, 2H), 5.78 (ddt, J : 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 5 11.06, 23.28, 26.16, 29.88, 34.10, 41.86, 65.17, 114.36, 138.90. 7-Bromo-6-ethyl-1-heptene (168). NBS (4.27 g, 24 mmol) was added slowly to a 0 °C solution of 167 (2.84 9, 20 mmol) and PPh3 (6.29 9, 24 mmol) in CH2CI2 (25 mL). After a typical workup, 3.66 g of bromide 168 (89% yield) were recovered after distillation (bp24 90-100 °C); IR (Film) 3079, 2965, 2934, 2876, 1642 cm";1H NMR (300 MHz) (00013) 8 0.87 (t, J : 7 Hz, 3H), 1.28-1.46 (m, 6H), 1.48-1.60 (m, 1H), 1.98-2.10 (m, 2H), 3.42 (dd, J: 5, 10-11 Hz, 1H), 3.44 (dd, J: 5, 10-11 Hz, 1H), 4.94 (ddt, J: 10, 2, 1 Hz, 1H), 4.99 (dq, J: 17, 2 Hz, 1H), 5.79 (ddt, J: 17, 10, 7 Hz, 1H);13C NMR (75 MHz) (CDCI3)510.85, 25.12, 25.88, 31.65, 33.81, 38.93, 40.95, 114.55, 138.64. Ethyl 2-deuterlo-2-ethyI-6-heptenoate (169). According to the procedure described for 123, ester 166 (6.44 g, 35 mmol) was added dropwise to a -78 °C THF (70 mL) solution of LDA (38.5 mmol) and stirred for 50 minutes. The reaction mixture was diluted with THF (20 mL) and n-BuLi (18.35 mL, 45.5 mmol) was added, stirred for 30 minutes, and quenched with D20 (10 mL). After the usual workup, the crude residue was distilled under reduced pressure (bp21 90-100°C) and 4.91 g of 169 were recovered (78% yield); IR (Film) 3079, 2977, 2936, 1734, 1642 cm“; 1H NMR (300 MHz) (CDCI3) 8 0.86 (t, J: 7 Hz, 3H), 1.23 (t, J: 7 Hz, 3H), 1.26-1.68 (m, 6H), 2.01 (q, J : 7 Hz, 2H), 4.11 (q, J : 7 Hz, 2H), 4.92 (ddt, J a 10, 2, 1 Hz,1H), 4.96 (dq, J: 17, 2 Hz, 1H), 5.75 (ddt, J: 17, 10, 7 Hz,1H); 13C NMR (75 MHz) (CDCI3)511.76, 14.33, 25.39, 26.62, 31.40, 33.60, 59.94. 114.57, 138.49, 176.29. 2-Deuterlo-2-ethyI-6-hepten-1-ol (170). Ester 169 (4.47 g, 24.2 mmol) was dissolved in diethyl ether (75 mL) and added slowly to a 0 °C suspension of LAH (0.65 g, 17.0 mmol) in diethyl ether (50 mL) according to the general 66 procedure. Following the usual workup, distillation of the crude residue (bpzz 90- 100 °C) provided 3.17 g of 170 (92% yield); IR (Film) 3343, 3079, 2963, 2861, 2130, 1642 cm"; 1H NMR (300 MHz) (00013) 8 0.87 (t, J : 7 Hz, 3H), 1.18-1.42 (m, 7H), 2.03 (bq, J : 7 Hz, 2H), 3.52 (m, 2H), 4.94-5.02 (m, 2H), 5.79 (ddt, J : 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3) 511.05, 23.17, 26.14, 29.77. 34.12, 65.15, 114.39, 138.92. A 7-Bromo-6-deuterlo-6-ethyl-1-heptene (171). NBS (3.43 g, 19.3 mmol) was added slowly to a 0 °C solution of 170 (2.28 g, 16 mmol) and PPh3 (5.05 g, 19.3 mmol) in CH2CI2 (25 mL) according to the general procedure. After a typical workup, 2.99 g of bromide 171 (91% yield) were recovered after distillation (bp19 80-90 °C); IR (Film) 3079, 2965, 2934, 2876, 2114, 1642 cm"; 1H NMR (300 MHz) (CDCI3) 5 0.87 (t, J : 7 Hz, 3H), 1.30-1.48 (m, 6H), 1.95-2.10 (m, 2H), 3.40 (d, J: 11 Hz, 1H), 3.44 (d, J: 11 Hz, 1H), 4.94 (ddt, J :10, 2, 1 Hz, 1H), 4.99 (dq, J : 17, 2 Hz, 1H), 5.79 (ddt, J : 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 510.80, 25.00, 25.85, 31.53, 33.82, 38.82, 114.55, 138.65. Dlethyl 2-(4-pentenyl)-2-(propyl)-propanedloate (172). The following reagents were combined as described in the general procedure: diethyl propylmalonate (40.45 g, 200 mmol), NaH (5.03 g, 210 mmol) in DMF (200 mL), and 5-bromo-1-pentene (32.16 g, 216 mmol). Coupling was achieved by heating the mixture at 50 °C for 40 hours. After workup and distillation under reduced pressure (bp<1 95-110 °C), 53.6 g of 172 were recovered (99 % yield); IR (Film) 3079, 2965, 2876, 1734, 1642 cm“; 1H NMR (300 MHz) (CDCI3) 8 0.89 (t, J : 7 Hz, 3H), 1.07-1.27 (m, 4H), 1.21 (t, J: 7 Hz, 6H), 1.77-1.89 (m, 4H), 2.02 (bq, J: 7 Hz, 2H), 4.14 (q, J : 7 Hz, 4H), 4.89-5.02 (m, 2H), 5.74 (ddt, J : 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (c0013) 8 14.09, 14.40, 17.36, 23.33, 31.71, 33.81, 34.46, 57.46, 60.94, 1 14.88, 138.14, 171.89. 67 Ethyl 2-propyI-6-heptenoate (173). Compound 172 (53.0 9, 196.3 mmol) was combined with LiCl (15.81 g, 373 mmol) and H20 (3.53 9, 196.3 mmol) in DMSO (400 mL) and heated to 180 °C for 18 hours. After the usual workup, 32.49 g of ester 173 were obtained (84% yield) after distillation (bpd 60-70 °C); IR (Film) 3079, 2959, 2938, 1734, 1642 cm"; 1H NMR (300 MHz) (CDCI3) 8 0.87 (t, J : 7 Hz, 3H), 1.19-1.66 (m, 8H), 1.23 (t, J: 7 Hz, 3H), 2.02 (bq, J: 7 Hz, 2H), 2.31 (tt, J: 5, 9 Hz, 1H), 4.11 (q, J: 7 Hz, 2H), 4.88-5.02 (m, 2H), 5.76 (ddt, J: 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3)514.00, 14.33, 20.60, 26.66, 31.90, 33.60, 34.67, 45.36, 59.98, 114.58, 138.51, 176.47. 2-PropyI-6-hepten-1-ol (174). Compound 173 (10.89 g, 55 mmol) was dissolved in diethyl ether (150 mL) and added dropwise to a 0 °C suspension of LAH (1.46 g, 38.5 mmol) in diethyl other (125 mL) as described in the general procedure. Following the usual workup, distillation of the crude residue (bp2o 90- 100 °C) provided 8.16 g of 174 (95% yield); IR (Film) 3337, 3079, 2930, 2870, 1642 cm“; 1H NMR (300 MHz) (CDCI3) 5 0.87 (t, J : 7 Hz, 3H), 1.12-1.52 (m, 9H), 1.53 (bs, 1H), 1.94-2.06 (m, 2H), 3.44-3.54 (m, 2H), 4.86502 (m, 2H), 5.78 (ddt, J : 17, 10, 7 Hz, 1H);13C NMR (75 MHz) (00013) 514.40, 19.96, 26.14, 30.36, 33.16, 34.10, 40.14, 65.53, 114.34, 138.89. 7-Bromo-6-propyl-1-heptene (175). N88 (4.27 g, 24 mmol) was added slowly to a 0 °C solution of 174 (3.12 g, 20 mmol) and PPh3 (6.29 g, 24 mmol) in CH2CI2 (25 mL) according to the general procedure. After the typical workup, 3.63 9 of bromide 175 (83% yield) were recovered after distillation (bp21 100-110 °C); IR (film) 3079, 2959, 2860, 1642 cm"; 1H NMR (300 MHz) (CDCI3) 5 0.89 (t, J: 7 Hz, 3H), 1.18-1.50 (m, 8H), 1.50-1.66 (m, 1H), 1.95-2.10 (m, 2H), 3.43 (d, J : 5 Hz, 2H), 4.93 (ddt, J: 10, 2, 1 Hz, 1H), 4.99 (dq, J: 17, 2 Hz, 1H), 5.79 (ddt, J: 17, 10, 7 Hz, 1H); 13c NMR (75 MHz) (00013) 514.17, 19.69, 25.83, 32.00, 33.82, 34.79, 39.13, 39.45, 114.54, 138.65. 68 Ethyl 2-deuterlo-2-propyl-6-heptenoate (176). According to the procedure described for 123, ester 173 (6.88 g, 35 mmol) was added dropwise to a -78 °C THF (70 mL) solution of LDA (38.5 mmol) and stirred for 50 minutes. The reaction mixture was diluted with THF (20 mL), n-BuLi (18.35 mL, 45.5 mmol) was added, stirred for 30 minutes, and quenched with D20 (10 mL). After a typical workup, the crude residue was distilled under reduced pressure (bp21 95- 105 °C) which provided 4.19 g of 176 (61% yield); IR (Film) 3079, 2961, 2934, 2874, 2170, 1736, 1642 cm"; 1H NMR (300 MHz) (00013) 8 0.87 (t, J : 7 Hz, 3H), 1.18-1.48 (m, 8H), 1.23 (t, J: 7 Hz, 3H), 1.50-1.64 (m, 2H), 2.02 (bq, J : 7 Hz, 2H), 4.11 (q, J : 7 Hz, 2H), 4.87502 (m, 2H), 5.75 (ddt, J : 17, 10, 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 513.98, 14.33, 20.56, 26.63, 31.80, 33.59. 34.57, 59.95, 114.57, 138.49, 176.46. 2-DeuterIo-2-propyI-6-hepten-1-ol (177). Compound 176 (3.87 9, 19.6 mmol) was dissolved in diethyl ether (60 mL) and added dropwise to a 0 °C suspension of LAH (0.52 g, 13.72 mmol) in diethyl ether (40 mL) as described in the general procedure. Following the usual workup, distillation of the crude residue (bp20 90-100 °C) provided 2.78 g of 177 (91% yield); IR (Film) 3335, 3079, 2959. 2128, 1642 cm'1;1H NMR (300 MHz) (00013) 8 0.87 (t, J : 7 Hz, 3H), 1.15-1.44 (m, 9H), 2.02 (tq, J: 2, 7 Hz, 2H), 3.50 (s, 2H), 4.91 (ddt, J: 10, 2, 1 Hz, 1H), 4.97 (dq, J : 17, 2 Hz, 1H), 5.78 (ddt, 17, 1o, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3)514.43, 19.94, 26.11, 30.24, 33.05, 34.12, 65.51, 114.36. 138.91. 7-Bromo-6-deuterlo-6-propyI-1-heptene (178). NBS (3.43 g, 19.3 mmol) was slowly added to a 0 °C solution of 177 (2.28 g, 14.5 mmol) and PPha (5.06 g, 19.3 mmol) in CH2CI2 (25 mL) according to the usual procedure. After a typical workup, 3.63 g of bromide 178 (77% yield) were recovered after distillation (bp21 95-105 °C); IR (Film) 3079, 2961, 2932, 2128, 1642 cm"; 1H NMR (300 MHz) 69 (CDCI3) 8 0.89 (t, J : 7 Hz, 3H). 1.20-1.45 (m, 8H), 1.98-2.08 (m, 2H), 3.42 (s, 2H), 4.94 (ddt, J: 10, 2, 1 Hz, 1H), 4.99 (dq, J: 17, 2 Hz, 1H), 5.79 (ddt, J: 17, 10, 7 Hz, 1H); 130 NMR (75 MHz) (CDCI3) 514.18, 19.65, 25.79, 31.88, 33.82, 34.68, 39.33, 114.54, 138.65. 1,1-Dldeuterlo-2-ethyl-6-hepten-1-oI (183). Compound 166 (2.76 g, 15.0 mmol) was dissolved in diethyl ether (25 mL) and added dropwise to a 0 °C suspension of LAD (0.44 g, 10.5 mmol) in diethyl ether (50 mL) as described in the general procedure. Following the usual workup, distillation of the crude residue (bp2o 95-105 °C) provided 2.09 g of 183 (97% yield); IR (Film) 3345, 3079, 2965,2934, 2863, 2197, 2093, 1642 cm";1H NMR (300 MHz) (00013) 8 0.87 (t, J : 7 Hz, 3H), 1.20-1.50 (m, 8H), 2.03 (tq, J: 1, 7 Hz, 2H), 4.92 (ddt, J: 10, 2, 1 Hz, 1H), 4.98 (dq, J: 17, 2 Hz, 1H), 5.79 (ddt, J: 17, 10, 7 Hz, 1H);13C NMR (75 MHz) (c0013) 8 11.07, 23.24, 26.16, 29.84, 34.11, 41.66, 114.37, 138.91. 7-Bromo-7,7-dldeuterio-S-ethyI-1-heptene (184). N88 (2.22 g, 12.5 mmol) was slowly added to a 0 °C solution of 183 (1 .50 g, 10.4 mmol) and PPh3 (3.28 g, 12.5 mmol) in CH2CI2 (15 mL) according to the usual procedure. After a typical workup, 1.86 g of bromide 184 (88% yield) were recovered after distillation (bp19 80-90 00); IR (Film) cm“; 1H NMR (300 MHz) (CDCI3) 8 0.87 (t, J : 7 Hz, 3H), 1.28-1.46 (m, 6H), 1.46-1.58 (m, 1H), 1.98-2.10 (m, 2H), 4.94 (ddt, J: 10, 2, 1 Hz, 1H), 4.99(dq, J: 17,2 Hz, 1H), 5.79 (ddt, J: 17, 10, 7 Hz, 1H);13C NMR (75 MHz) (CDCIa) 510.86, 25.07, 25.88, 31.60, 33.82, 40.71 , 114.56, 138.65. 1,1-DIdeuterIo-2-propyl-6-hepten-1-oI (185). Compound 173 (2.97 g, 15.0 mmol) was dissolved in diethyl ether (25 mL) and added dropwise to a 0 °C suspension of LAD (0.44 g, 10.5 mmol) in diethyl ether (50 mL) as described in the general procedure. Following the usual workup, distillation of the crude residue (bp22 105-115 °C) provided 1.72 g of 185 (73% yield); IR (Film) 3339, 70 3079. 2959, 2930. 2872. 2197, 2093. 1642, 1460, 1416 cm"; 1H NMR (300 MHz) (CDCI3) 8 0.87 (t, J: 7 Hz, 3H). 1.17-1.49 (m, 10H), 2.02 (bq, J: 7 Hz. 2H), 4.87- 5.02 (m. 2H), 5.78 (ddt, J : 17, 10. 7 Hz, 1H); 13C NMR (75 MHz) (CDCI3) 5 14.39, 19.94, 26.10. 30.27, 33.08, 34.08, 39.92, 114.32. 138.88. 7-Bromo-6-deuterio-6-propyI-1-heptene (186). NBS (1.92 9, 10.8 mmol) was slowly added to a 0 °c solution of 185 (1.42 g. 9.0 mmol) and PPh3'(2.83 g, 10.8 mmol) in CH2CI2 (15 mL) according to the usual procedure. After a typical workup, 1.82 g of bromide 186 (91% yield) were recovered after distillation (bp22 100-110 °C); IR (Film) 3079, 2959, 2930. 2872, 2861, 1642. 1458, 1416 cm"; 1H NMR (00013) 8 0.89 (t, J : 7 Hz, 3H), 1 .18-1.50 (m, 8H), 1.52-1.65 (m, 1H). 1.97- 2.01 (m, 2H), 4.90-5.04 (m. 2H), 5.79 (ddt, J : 17, 10, 7 Hz, 1H); 13c NMR (CDCI3) 5 14.19, 19.70, 25.83, 31.95. 33.83, 34.74. 38.90. 11455, 138.66. 3-Ethylcyclohexanone (187). In a 100 mL Schlenk flask, equipped with a dropping funnel, were loaded magnesium turnings (7.78 g, 320 mmol). THF (60 mL), and a small crystal of iodine. and the flask was heated to 45 °C. Ethyl bromide (8.71 g, 80 mmol) was added dropwise over 2 hours and stirred an additional 4 hours. A separate 250 mL flask was loaded with n-BuaPCul (1.99 g, 5.08 mmol). diluted with THF (10 mL). and cooled to -45 °C. The Grignard solution was transferred via cannula to the copper salt over 5 minutes and stirred for an additional 20 minutes. Cyclohexenone (3.70 g, x mmol) was added dropwise over 15 minutes, stirred for 90 minutes, and the solution was allowed to warm to -20 °C until all cyclohexenone was consumed. The reaction mixture was quenched with saturated NH4CI (8 mL) and diluted with petroleum other (500 mL) The organic layer was washed with NH40H (3 x 100 mL; 6 mL NH40H in 294 mL H20), NaCI (40 mL), dried over M9804, filtered through silica gel, concentrated to an oil, and distilled (bpzo 80-100 °C) which provided 3.96 g of 37 (82% yield); IR (Fllm) 2963. 2934. 2876. 1715 cm"; 1H NMR (300 MHz) (CDCI3) 5 0.85 (t. J : 7 71 Hz, 3H), 1.18-1.42 (m, 3H), 1.48-1.70 (m. 2H). 1.78-2.05 (m. 3H), 2.15-2.42 (m, 3H); 130 NMR (75 MHz) (00013) 811.10, 25.22. 29.23, 30.82, 40.70, 41 .43, 47.78. 212.07. 3-Ethylmethylenecyclohexane (188). NaH (0.39 g. 16.5 mmol) was loaded into a 250 mL flask under an atmosphere of nitrogen and suspended in DMSO (45 mL). The suspension was heated to 75 °C for 45 minutes or until no further H2 was evolved. A separate flask was loaded methyltriphenylphosphonium bromide (5.90 g, 16.5 mmol). diluted with DMSO (30 mL), and heated to 50 °C until the solution was homogeneous. The Wittig salt was added to the dimsyl anion at ambient temperature and stirred for 30 minutes, during which time the solution became orange-red. Ketone 187 was added at room temperature and the heated to 55 °C. After the ketone was consumed, the reaction mixture was quenched with H20 (25 mL) and extracted with petroleum ether (4 x 60 mL). The organic layers were combined and washed with (1:1) DM80:H20 (65 mL). NaCI (130 mL), dried over M9804, filtered through neutral alumina, and concentrated. The crude residue was distilled (bp51 60-65 °C) which provided 1.14 g of 188 was recovered (71% yield); IR (Film) 3073, 2963. 2932. 2876, 2859, 1651 cm"; 1H NMR (300 MHz) (CDCI3) 5 0.86 (t. J: 7 Hz. 3H). 0.90-1.10 (m. 1H). 1.18-1.38 (m, 4H). 1.58-1.82 (m. 3H), 1.84-2.00 (m, 1H). 2.18-2.36 (m. 2H), 4.58 (bs, 2H); 13C NMR (75 MHz) (CDCI3)511.42, 27.13, 29.45, 32.21, 35.21, 40.95, 41.41, 106.66, 149.73. 1-EthyI-3-methylcyclohexane (189). In 50 mL flask were loaded Pt02 (25 mg). 188 (0.75 g. 6 mmol). and methanol (15 mL). The flask was filled and purged with H2 three times and stirred under an atmosphere of H2. The consumption of the olefin was monitored by NMR. GC analysis determined the ratio of isomers to be 1.96:1 (trans:cis); 13C NMR (75 MHz) (CDCI3) 511.43, 72 11.97. 20.76, 20.87, 23.00, 26.37, 27.23. 27.28. 31.20. 31.30. 32.58, 32.73, 33.89, 34.23. 35.44, 38.93. 39.50. 41.99. 3-Propylcyclohexanone (190). The following reagents were combined according to the procedure described for 187: propyl bromide (9.83 g. 80 mmol), Mg (7.78 g, 320 mmol) and THF (60 mL); n-BU3PCuI (1.99 g, 5.08 mmol) in THF (1 0 mL). and cyclohexenone (3.75 g. 39.1 mmol). The crude residue was distilled (bP<1 50-60 °C) and 3.13 g 190 was obtained (57% yield); IR (Film) 2959, 2872, 1713 cm";1H NMR (300 MHz) (CDCla) 8 0.83 (1. J : 7 Hz, 3H). 1.16-1.38 (m, 5H), 1.50-1.90 (m, 3H), 1.92-2.06 (m. 2H), 2.14-2.42 (m, 3H); 13c NMR (75 MHz) 0003) 5 14.04. 19.67, 25.25, 31.23. 38.75, 38.77. 41 .45, 48.12. 212.09. 3-Propylmethylenecyclohexane (191). The following reagents were combined as described for 188; NaH (0.39 g. 16.5 mmol) in DMSO (45 mL); methyltriphenylphosphonium bromide (5.90 g, 16.5 mmol) in DMSO (30 mL). and 190 (1.93 g. 13.8 mmol). After workup and distillation (bps; 70-80 °C). 1.32 g 191 was recovered (71% yield); IR (Film) 3071. 2980. 1651 om";1H NMR (300 MHz) (00%) 8 0.86 (1. J : 7 Hz. 3H), 0.95-1.12 (m. 1H), 1.10-1.42 (m, 6H), 1.50-1.82 (m. 3H), 1.85-2.02 (m. 1H). 2.18-2.35 (m. 2H). 4.60 (bs, 2H);13C NMR (75 MHz) (CDCI3)514.33, 19.93. 27.16. 32.59. 35.21. 38.89, 39.12. 41 .79, 106.65. 149.76. 1-MethyI-3-propylcylohexane (192). In a 25 mL flask were loaded Pt02 (25 mg), 191 (0.82 g, 6 mmol). and diethyl ether (15 mL). The hydrogenation was carried out as in 189. GO analysis determined the ratio of isomers to be 1.71:1 (trans:cis); 13C NMR (75 MHz) (CDCI3)514.33, 15.22, 19.89, 20.46, 20.75, 20.88. 22.95, 26.38. 27.22, 31.64, 32.05. 32.72, 32.98, 33.89, 35.42, 36.92, 37.40. 39.25. 39.97, 42.40. 3-(1-methylethyl)-cyclohexanone (193). The following reagents were combined according to the procedure described for 187: 2-bromopropane (7.37 g. 60 mmol), Mg (5.83 g. 240 mmol) in THF (60 mL); n-BuaPCuI (1.99 g, 5.08 73 mmol) in THF (10 mL); and cyclohexenone (3.53 g, 36.8 mmol). The crude residue was distilled (bp1o 80 °C) and 4.00 g 193 was obtained (78% yield); IR (Film) 2961, 2872, 1715, 1466. 1449. 1426 cm";1H NMR (300 MHz) (CDCI3) 8 0.85 (d. J : 6 Hz, 3H), 0.86 (d, J : 6 Hz, 3H). 1.25-1.40 (m. 1H), 1.45-1.62 (m, 3H), 1.78-1.85 (m. 1H). 1.95210 (m, 2H), 2.15238 (m. 3H); 130 NMR (75 MHz) (CDCI3) 5 19.25, 19.48, 25.47. 28.27. 32.42, 41 .43. 45.26. 45.35. 212.49. 3-(1-methylethyI)-methylenecyclohexane (194). The following reagents were combined as described for 188; NaH (0.72 g, 30 mmol) in DMSO (75 mL); methyltriphenylphosphonium iodide (12.12 g, 30 mmol) in DMSO (50 mL); and 193 (3.5 g, 25 mmol). After workup and distillation (bpss 60-70 °C), 3.08 g 194 were recovered (89% yield); IR (Film) 3073. 2959, 2932. 2874, 2839, 1651. 1464, 1447. 1433 cm“; 1H NMR (300 MHz) (00%) 8 0.861 (d, J: 7 Hz, 3H), 0.864 (d, J: 7 Hz, 3H). 1.00-1.35 (m. 3H), 1.43 (oct. J: 7 Hz, 1H), 1.65-2.00 (m, 4H). 2.18- 2.35 (m, 2H). 4.56-4.60 (m, 2H); 130 NMR (75 MHz) (CDCI3) 519.72. 19.74, 27.50, 29.37. 32.59, 35.23, 38.72. 45.74, 106.65, 150.17. I 3-(1-MethylethyI)-1-methylcyclohexane (195). A 25 mL flask was loaded with Pt02 (50 mg). 1 94 (0.35 g, 2.5 mmol). and methanol (5 mL). The hydrogenation was carried out as in 189. G0 analysis determined the ratio of isomers to be 2.42:1 (trans:cis). 4-(1-MethylethyI)-methylenecyclohexane (196). The following reagents were combined as described for 188; NaH (0.41 g, 17.1 mmol) in DMSO (17 mL); methyltriphenylphosphonium iodide (6.92 9, 17.1 mmol) in DMSO (40 mL); and 161 (1.94 g, 14.3 mmol). After workup and distillation (bpsa 60-70 °C). 1.64 g 196 was recovered (86% yield); IR (Film) 3073, 2963, 2874. 1651 om";1H NMR (300 MHz) (CDCI3) 5 0.85 (d. J: 7 Hz. 6H), 0.98-1.25 (m, 3H). 1.42 (oct. J : 7 Hz, 1H). 1.70-1.84 (m, 2H). 1.90-2.05 (m. 2H). 2.22-2.35 (m, 2H). 4.56 (t, J : 2 Hz. 2H); 130 NMR (75 MHz) (CDCI3) 519.84, 31.10. 32.39. 34.94, 43.66, 106.24, 150.31. 74 4-(1-Methylethyl)-1-methylcyclohexane (197). A 25 mL flask was loaded with Pt02 (50 m9). 196 (0.35 g. 2.5 mmol), and methanol (5 mL). The hydrogenation was carried out as in 189. GO analysis determined the ratio of isomers to be 2.23:1 (trans:cis). CHAPTER 2. DIASTEREOSELECTIVE CYCLOPENTANE FORMATION MEDIATED BY TITANOCENE DICHLORIDE 1 . Introduction Our initial study began with the cyclization of a substrate utilized by Grubbs in an experiment previously described.9a 0ur intentions were to optimize the different steps in the linear sequence (Scheme 15). more accurately probe the regiocontrol of the olefin insertion, and determine the extent of B- elimination. Preparation of hex-5-en-1-yltitanocene chloride (198) utilized established procedures6 in which bromide 28 was metallated by treatment with magnesium in THF (82. Scheme 26). Upon quenching an aliquot of 82 with an HCI/diethyl ether solution, the product mixture was determined to consist of a 95:5 ratio of 1-hexene (69) and methylcyclopentane (6). Furthermore, this ratio did not change upon continued heating. Transfer of the alkenyl ligand to titanium was accomplished with TiCp20l2 in CH2Cl2, which provided 198 in quanitatlve yield after purification.69 It is noteworthy that the ratio of 69 to 6 did not change during the transmetallation process. 75 76 Scheme 26. Preparation of 5-Hexen-1-yltltanocene Chloride (1 98) X M: “092012 Br 31". x + &. noppl 20 02;X:Mper 83;X:MgBr 190; 100% YIeId 69;X:H 6; X:H Scheme 27 represents the potential pathways that complex 198 could undergo. Treatment of 198 with a Lewis acid could result in simple destructive B-elimination to 1,5-hexadiene if the Lewis acid was too strong or the reaction temperature too high. 0n the other hand, if the conditions were too mild, 198 could be recovered unchanged. Upon fine tuning of the reaction conditions, productive bond formation could proceed in either an exo (five-membered ring formation) or an endo mode (six-membered ring formation). After cyclization. the alkyltitanocene complexes could presumably undergo B-elimination or, if stable to the reaction conditions, provide either methylcyclopentane or cyclohexane upon hydrolysis. Scheme 27. Potentlal Pathways of Cyclization and Decomposition of 198 noppl H 9) - 4...... _..,, <5 1.4m... Nv4—MEWW [mezc'lfi WH 196 1...: m8. Trcppl H 0 8+1 Elirrlinulon 6 13’ G 77 Treatment of a toluene solution of 198 with EtAlCl2 (0.5-1.0 equiv) at -78 °C resulted in an immediate darkening of the mixture.73 This has been attributed to the complexation of the aluminum catalyst to the chloride on the titanium. The role of the aluminum was to either weaken or sever the titanium-chloride bond. The net result was the development of substantial positive charge on titanium, polarization the carbon-titanium bond. and to create an available coordination site for olefin complexation. Hydrolysis of the reaction mixture after 30-60 minutes revealed that complete consumption of 198 had occurred with concomitant formation of a single regioisomer. that of exo insertion. Yields of 85- 95% (from 198) were common for this insertion.1b Furthermore, no detectable amount of B-elimination was observed under the reaction conditions. We have since reported a thorough investigation on the regiospecificity of 5-exo-trig olefin insertions with respect to structural differences of the olefin.13 A variety of vicinally and geminally substituted olefins. including systems where the olefin was either endo- or exo-cyclic in a ring and simple acyclic systems, were prepared and cyclized. In general. to promote complete cyclization, longer reaction times, higher temperatures, and higher catalyst concentrations were required. However, these factors had no effect on the yields or the regiospecificity of the reaction, the product of exo insertion was exclusively observed. A major consequence of the regiocontrol exhibited by this organometallic approach was the ease with which quaternary centers were produced. including geminal dimethyl and angular methyl groups, features which have proved troublesome by other methods. In the present chapter, we report our findings regarding the effect a methyl group on the newly formed asymmetric center. In essence. could a respectable level of diastereomeric excess be achieved? In order to probe this question. we Prepared and cyclized substrates analogous to hex-5-en-1-yltitanocene 78 chlorides in which a methyl group was situated at the various positions on the tether. The titanium metal and its ligands were anticipated to have a strong influence on the stereochemistry of the product. First, as with other metals, titanium should be able to complex the olefin, thus restricting all four atoms into a syn-coplanar relationship, and thereby controlling the olefin geometry during bond formation. Secondly, in contrast to other ionic or radical methods, the ligands on the metal could have a pronounced effect on the geometrical deployment of the olefin and/or the alkyl group. and therefore exert more influence on the diastereomeric conformations in the transition state. 2. General Features and Limitations for the Synthesls of Methyl- Substituted 5-Hexen-1-yltltanocene Chlorides Scheme 26 illustrates our chosen method for the preparation of the requisite alkenyltitanocene complexes. As already mentioned. reduction of the alkyl bromide with magnesium produced a small amount of cyclopentylmethyl magnesium bromide. This phenomena has been well documented by Ashby and others and has been explained in terms of single electron transfer (SET).37 During this process. transient radical intermediates, thought to be formed in the reaction, underwent a rapid 5-exo-trig cyclization. This process presented no major problems for our model substrate since both radical cyclization and our titanocene-mediated method produced the same compound. However, for the substrates of interest in this chapter. SET was a menacing dilemma. Compounds 368-d cyclized to a much greater extent (12-21%) than did 27 (T able 2, Method A), presumably due to the Thorpe-lngold substituent effect.70 In contrast to 28, 368-d produced a pair of diastereomers in a relatively indiscriminant fashion. which paralleled those for radical cyclization (Table 2, Method D). These shortcomings prompted a search for an alternative method of 79 Grignard formation. Rieke has developed a variety of methods to generate “active metals” which have been shown to metallate allylic and benzylic halides without dimerization.71 We had hoped that rapid generation of the Grignard reagent at low temperatures with ”active magnesium" may diminish cyclization at this stage. However, much to our dismay. we found only a marginal decrease in cyclization and a substantial decrease in the yield (Table 2. Method E). Table 2. Diastereomer Ratlos of Dlmethylcyclopentane . Products via Titanium and Free Radical Methods . Wins Oyizann Precursor Method' YIOIdb Uncyclized 1.2 (trans:cis) 1,3 (trans:cis) A 100 m 10 2 War 8 100 87 10 3 M. c in o 92 8 (35.) D 90 O 79 21 E 76 91.5 6.5 2 J” A 98 83 3 14 B m 5 3 12 ’ 3' c 93 0 3 97 ‘3‘”) o 100 14 19 67 A 100 5 9 6 W9 8 64 77 14 9 Ma 0 94 0 92 8 (366) o m o s 41 MD A a 79 7 12° N B 72 16 2 55° 9' - o as 17 19 61 ~ (3“) E 81 2 7° ‘MetItodA:(1)w,THF.45°C. MethodB:(1) MethodA; (2)1'ioppl,.c+lpl, MethodC:(1) Method A; (2)MethodB; (3) 1.0 equiv. 614161,. reasons. -78 '0; (4) Hcl. Modo:AlBN,eu,SnH, C.H.,w°c. MethodE: MgCI,K.THF.0°C. hReference”. °Belenceofmehndooneiehdofhree pueblo p—hydrogon elimination products. Ligand transfer from magnesium to titanium provided two interesting results. Typically, no further cyclization occurred during the transmetallation Process. However, an 8% increase was observed for 360 (Table 2. Method B). When Grignard formation of 36c was performed in diethyl ether, as opposed to THF, followed by transmetallation, 85-100% cyclization took place. Presumably 80 the less Lewis basic solvent, diethyl ether, did not complex the M9X2 as tightly as THF. thereby allowing M9X2 to promote the olefin insertion. Interestingly, M9X2 was not an effective catalyst for the other substrates under these conditions. We are currently investigating the nature of this apparent rate enhancement for substrate 366. It is noteworthy that the 01.8, and y carbons to the titanium resemble the growing polyproylene chain in Ziegler-Natta polymerization, a system in which M9X2 efficiently and stereoselectively catalyzed the polymerization of propylene. Secondly, ligand transfer of 36d from magnesium to titanium resulted in almost complete cyclization. even in THF (Table 2. Method B). The similarity of the diastereomeric ratio formed to that observed for radical methods, suggest that radical ring closure occurred. not olefin insertion. It is not clear if the radical was produced by SET or by homolytic cleavage of the sterically crowded 2° carbon-titanium bond. 3. Synthesis and Cycllzatlon of Methyl-Substltuted 5-Hexen-1- yltitanocene Chlorides1b Incorporation of a methyl group on the tether was found to have a significant effect on the stereochemical outcome of carbon-carbon bond formation. Treatment of the titanium complex of 36a (Method 0, Table 2) with EtAlCl2, resulted in a 92:8 ratio (trans:cis) of 1.2-dimethylcyclopentanes. Taking into account the amount of cis isomer already present, a 94:6 diastereoselectivity during the Ziegler-Natta olefin insertion was realized. Similarily, 36b underwent the Lewis acid assisted insertion which produced a 97:3 (cisztrans) diastereomeric mixture of 1,3-dimethylcyclopentanes. Thus. the actual diastereoselectivity was > 99:1. Finally. cyclization of the titanium complex of 366 resulted in a high trans selectivity (92:8) and > 99:1 in the EtAlCl2 step. The fact that titanium complexes of 36b and 360 yielded the same product selectivity 81 for different isomers, negates the possibility for product equilibration under the reaction conditions. The stereoselectivity can be explained in terms of a psuedo-chair transition state (Scheme 28). Treatment of the organometallic complex with EtAlCl2, resulted in complexation of the Lewis acidic aluminum to the chloride atom on the titanium.7a Although the extent of bond breaking of the titanium- .chloride bond is unclear. the net result was distribution of substantial positive charge on titanium, increasing the electrophilicity of the metal, and weakening of the carbon-titanium bond. During olefin insertion. a four-centered transition state has been proposed in which the carbon-carbon double bond was assumed to be syn-coplanar with the carbon-titanium bond. Substantial stabilization would result from such an orientation based on orbital considerations previously described. Conformations 199-202 could accommodate such an interaction. However. in order to minimize eclipsing interactions. 199 and 202 would be strongly disfavored. Consequently, the diastereomeric ratio observed resulted from an unequal partitioning between 200 (major) and 201 (minor). Conformation 201 would be disfavored. with respect to 200. since the methyl groups were axially deployed. Additional destabilization of an axial methyl group could arise from steric interactions with the cyclopentadienyl ligands on the metal. 82 Scheme 28. Proposed Transition States for Flve-Membered Rlng Formation \‘mppl at +1" m ~— H 4. Comparlson of the Stereoselective Outcome of Varlous Methods With the variety of methods available to facilitate cyclization of the 5- hexen-1-yl system (Table 3), each having different steric and electronic characteristics at the activated carbon, some qualitative generalizations can be made about the factors that contribute to high stereoselectivities. Factors that deserve consideration include the nature of the reactive intermediate. ligands on the intermediate, and the temperature at which the reaction was conducted. In the case of radical cyclization (80 °C). selectivity drops off sharply as the methyl group was further removed from the olefin. Beckwith17 and Houk2° agree. based on independent calculations. that the major product was formed through a chair transition state in which the methyl group was in a pseudo equatorial position. As a point of divergency, Houk suggested that the conformation leading to the minor product resembled a boat transition state with the methyl group in an equatorial position. not a chair conformation with an axial methyl 83 group. In support of this. Houks calculations more accurately predict the relative ratio of products. The difference in energy between the two transition states (estimated to be 0.5 kcal/mol) was attributed to the closer proximity of the C3 methylene (relative to the radical center) to the olefin. Examination of lithium cyclizations (~30 to 25 °C) revealed that excellent stereocontrol was realized for each substrate. In order to show that this drastic difference was not merely related to temperature, aluminum cyclizations are noted (67 °C). As in the case of lithium. excellent control over the newly formed asymmetric center was produced, despite the higher temperatures. Assuming Houk's model was an accurate description of the minor isomer, incorporation of the 1: complexing ability of the metal(s) into the transition state would increase the nonbonding interactions with the 03 methylene and in turn, suppress the boat transition state. It can be seen by comparison of aluminum- and lithium-mediated cyclizations. that the ligands on aluminum had a subordinate role in the stereochemical outcome of bond formation. The minimal effect that the alkyl ligands on aluminum had on the stereoselectivity was likely a consequence of the sp2 hybridization of aluminum, which directs the ligands away from the alkyl tether. Finally, cyclization mediated by titanium demonstrated the combined effects of it complexation. low temperatures (-78 °C), and proper geometry of the ligands for increased interactions with the alkyl tether. Thus. maximum stereocontrol was achieved. 84 Table 3. Comparison of Dlastereoselectlvltles for Different Methods Cyclization Dimethycyclopentanes Precursor M 1,2 (trans :cis) 1.3 (trans:cis) WM Radical 79:21 Me U 97:3 AlEtz 97:3 TICpZCI 94:6 M. Radical 19:67 W Li 6:94 M AIEIZ 8192 T10p20l 1 :99 MM Radical 59:41 Me U 97:3 A1512 96:4 TiCp20I 99:1 5. Optimization and Cyclization of 133: Formation of Bicyclo[4. 3. 0]nonane Methodology directed toward the efficient formation of multiple ring products is an important transformation. With the advent of new reaction conditions to facilitate olefin insertion for our titanium approach, we were compelled to reinvestigate the cyclization of 13313-72 In this substrate, 5-exo- trig insertion into the vicinally substituted olefin presented new challenges compared to the other substrates in this chapter. The primary obstacle to be surmounted was bringing the sterically congested olefin into the titanocene wedge. Table 4 illustrates some represenative conditions attempted for cyclization of 133b. As previously shown for other vicinal olefins.“) EtAlCl2 (entry 1) was ineffective at -78 °C. Since EtAlCl2 was incompatible with 133b at higher temperatures, recourse to other Lewis acids was necessary. Me2AlCl was marginally successful in toluene (entry 2), providing a 54% recovery of material (combined yield of acyclic and cyclic). By simply changing the reaction medium 85 to CH2CI2. reaction times were diminished and conversions were enhanced (entry 3). Table 4. Optimizatlon of the Cyclization Conditions for 133b x a H OJ -. UV)" 133.; x . Br Yield” 133b; X 3 “C920. H 1) EtAlCl2, (2 equiv.) Toluene, -78 °C 93 86 1 4 2) Me2AICI (1.5 equiv.) Toluene, ~30 to -10°C 54 23 77 3) MezAlCl (1.5 equiv.) 0120': 9° °0 59 10 90 O O o Int-g: ‘(1)133a, Mg, THF, 45 1'0; (2)1'le20I2, 50% yield. ”Reference 69. Scheme 29 depicts the comparison of radical and titanium-mediated cyclization of 133. Conversion of 133 to 133b was accomplished in the usual fashion, which resulted in a 50% yield after purification. Further optimization for the isolation of this complex is warranted. Treatment of 133b under the optimized conditions resulted in the exclusive formation of the cis isomer in 82% yield. Similarily, treatment of 133 under the tin-hydride conditions produced the same stereoselectivity. Apparently, the highly rigid nature of this 5-hexen-1-yl analog precludes attack from the opposite face. 86 Scheme 29. Comparison of Titanium-Mediated Cyclization to Free Radical Cyclization for 1338 H H X 8u3SnH. AIBN. Me 8.4.808 O\/l .. 99% YIOId I 5 H 8 133a; X:Br I: 1) Mg. THF 0 2) T160202. CH 50% Yield X MezAICl (2 equiv) CHZCIZ, 9 91 0 -50 to -30 °C. 90% Yleld 1330; X-TiszCI 6. Conclusion The results described in this chapter demonstrate a powerful new tool for stereoselective cyclopentane formation. Several general features about the efficiency and selectivity of the Ziegler-Natta insertion deserve mentioning. Activation of the hex-5-en-1-yltitanocene complexes with 0.5-1.0 equivalents of EtAICI2 facilitated complete cyclization within 30 minutes at -78 °C. Olefin insertion was completely regiospecific (exo cyclization) and the conditions did not produce detectable amounts of B-hydrogen elimination products. The elegance of this methodology was inherent in the ability of the titanium and its ligands to control the selectivity of carbon-carbon bond formation. In all cases. >94:6 diastereoselectivity was observed for the intramolecular olefin insertion with >90% yields achieved. 87 7. Experimental Radical Cyclization of 6-Bromo-3-methyl-1-hexene (368). A Schlenk flask was loaded with AIBN (10 mg), evacuated, purged with argon. and then benzene (10 mL) was added. Bromide 368 (88.5 mg, 0.5 mmol) and Buaan (174.6 mg. 0.6 mmol) were added to the flask and the reaction mixture was heated to 80 °C for 8 hours. Analysis of the reaction mixture using gas chromatographic techniques69 revealed that a 90% yield of cis- and trans-1,2- dimethylcyclopentanes (21 :79, respectively) was obtained (see Table 2. Method D). Radical Cyclization of 6-Bromo-4-methyI-1-hexene (36b). A Schlenk flask was loaded with AIBN (10 mg), evacuated. purged with argon, and then benzene (10 mL) was added. Bromide 36b (88.5 mg, 0.5 mmol) and Bu38nH (174.6 mg, 0.6 mmol) were added to the flask and the reaction mixture was heated to 80 °C for 8 hours. Analysis of the reaction mixture (see Table 2, Method D)69 showed the reaction resulted in a 100% yield of a mixture of 4- methyl-t-hexene (14%) and cis- and trans-1,3-dimethylcyclopentanes (86%. 7822. respectively). Radical Cyclization of 6-Bromo-5-methyI-1-hexene (360). A Schlenk flask was loaded with AIBN (10 mg), evacuated, purged with argon. and then benzene (10 mL) was added. Bromide 36c (88.5 mg, 0.5 mmol) and Buaan (174.6 mg, 0.6 mmol) were added to the flask and the reaction mixture was heated to 80 °C for 8 hours. Analysis of the reaction mixture using gas chromatographic techniques69 revealed that an 88% yield of cis- and trans-1,3- dimethylcyclopentanes (41 :59, respectively) was obtained (see Table 2. Method D). 88 Radical Cyclization of 6-Bromo-6-methyI-1-hexene (36d). A Schlenk flask was loaded with AIBN (10 mg), evacuated, purged with argon, and then benzene (10 mL) was added. Bromide 36d (88.5 mg, 0.5 mmol) and Buaan (174.6 mg, 0.6 mmol) were added to the flask and the reaction mixture was heated to 80 °C for 8 hours. Analysis of the reaction mixture (see Table 2, Method 0)69 showed the reaction resulted in a combined yield of 85% of the following mixture: 1-heptene (17%) and cis- and trans-1.3-dimethylcyclopentanes (83%, 77:23, respectively). Grignard Formation of 6-Bromo-3-methyI-1-hexene (368) Using Rieke Magnesium. Potassium metal (70.4 mg. 2.18 mmol) and anhydrous MgCl2 (95.21 mg, 1.00 mmol) were loaded into a Schlenk flask under an atmosphere of nitrogen and THF (4 mL) was added. The mixture was heated to reflux for 3 hours, cooled to 0 °C, and 368 (88.45 mg, 0.5 mmol) was added. Analysis of the final mixture (see Table 2, Method E). after quenching with 1.0 M HCI/E120 (3 mL). showed that a combined yield of 76% of the following mixture was obtained: 3-methyI-1-hexene (92%) and cis- and trans-1.2-dimethylcyclopentanes (8%. 2:6, respectively). Grignard Formation of 6-Bromo-6-methyI-1-hexene (36d) Uslng Rieke Magnesium. Potassium metal (70.4 mg. 2.18 mmol) and anhydrous MgCl2 (95.21 mg. 1.00 mmol) were loaded into a Schlenk flask under an atmosphere of nitrogen and THF (4 mL) was added. The mixture was heated to reflux for 3 hours, cooled to 0 °C, and 36d (88.45 mg. 0.5 mmol) was added. Analysis of the final mixture (see Table 2. Method E). after quenching with 1.0 M HCL/Et20 (3 mL). showed that a combined yield of 76% of the following mixture was obtained: 1-heptene (81%), cis- and trans-1.2-dimethylcyclopentanes (9%, 2:7, respectively), and a mixture of acyclic B-hydrogen elimination products (10%). 89 General Procedure for ZIegIer—Natta Cyclization: 6-Bromo-3-methyI-1- hexene (368). A Schlenk flask was charged with M9 (195 mg. 8.0 mmol) and heated under vacuum with a heatgun for 30 minutes. After cooling to ambient temperature, the flask was purged with argon, THF (2 mL) was added. and the flask was then placed in a 45 °C oil bath. Bromide 368 (353.8 mg. 2.0 mmol) was added with a gas-tight syringe over 6 hours. followed by stirring for an additional 4 hours. A separate Schlenk flask was loaded with Tle2Cl2 (598 mg, 2.4 mmol), evacuated and purged with argon, the Tle20l2 was suspended in CH2CI2 (8 mL), and the flask was cooled to -45 °C. Transmetallation of the organic ligand was accomplished by addition of the Grignard intermediate to the titanocene dichloride solution via cannula (an additional 1 mL THF was used to rinse the Mg). stirred for 30 minutes at -45 °C. and warmed to ambient temperature for an additional 5 hours. After transmetallation. the mixture was concentrated to approximately 4 mL followed by purification of the organometallic complex by filtration to remove the excess TiCp20l2 and the Lewis acidic Grignard salts. Purification was accomplished by taking the solution up in toluene (5 mL) and hexane (5 mL), stirring for 10 minutes at room temperature. and then cooling to 0 °C for 10 minutes. The entire mixture was transferred via cannula into another Schlenk flask (the flask was previously evacuated and filled with argon as before) that was equipped with a glass frit and the solution was pulled through the frit under a slight vacuum. The solids were rinsed with additional toluene (3 x 10 mL) and filtered. Concentration of the reaction mixture in vacuo provided the alkyltitanocene complex as a red slurry. which was diluted with toluene (14 mL) and cooled to -78 °C. The precise composition of the reaction mixture. consisting of cyclic and acyclic alkyltitanocene complexes, is illustrated in Table 2 (Method B). A 1.8 M solution of EtAlCl2 (2 mmol) was added to the alkyltitanocene solution over 10 minutes. the reaction mixture was stirred for 60 minutes at -78 °C, and finally 90 quenched with 1.0 M HCI/Et20 (5 mL). Analysis of the final reaction mixture (Table 2. Method C) showed that a 93% overall yield from the bromide of cis- and trans-1.2-dimethylcyclopentanes (8:92. respectively) was obtained. Ziegler-Natta Cyclization of 6-Bromo-4-methyI-1-hexene (36b). Bromide 36b (353.8 mg. 2.00 mmol) was cyclized according to the general procedure described for 368. The precise composition of the reaction mixture prior to addition of EtAlCl2, consisting of cyclic and acyclic alkyltitanocene complexes. is illustrated in Table 2 (Method 8). Analysis of the final reaction mixture, after hydrolysis (Table 2. Method C), showed that a 78% overall yield from 36b to cis- and trans-1,3-dimethylcyclopentanes (97:3. respectively) was obtained. ZIegler-Natta Cyclization of 6-Bromo-5-methyl-1-hexene (36c). Bromide 36c (353.8 mg, 2.00 mmol) was cyclized according to the general procedure described for 368. The precise composition of the reaction mixture prior to addition of EtAlCl2, consisting of cyclic and acyclic alkyltitanocene complexes. is illustrated in Table 2 (Method B). Analysis of the final reaction mixture, after hydrolysis (Table 2. Method C). showed that a 61% overall yield from 36c to cis- and trans-1,3-dimethylcyclopentanes (8:92, respectively) was obtained. Ziegler-Natta Cyclization of 6-Bromo-6-methyI-1-hexene (36d). Bromide 36d (353.8 mg. 2.00 mmol) was converted to the titanocene derivative as described in the general procedure. Analysis of the reaction mixture after transmetallation (Table 2, Method 8) showed a complex mixture of products was obtained in 72% overall yield from bromide 36d, which consisted of the following: cis-1.2-dimethylcyclopentane (55%). trans-1.2-dimethylcyclopentane (22%). 1- heptene (16%), 1.6-heptadiene (3%), and 1,5-heptadiene (4%, mixture of isomers). CHAPTER 3. DIASTEREOSELECTIVE CYCLOHEXANE FORMATION MEDIATED BY TITANOCENE DICHLORIDE 1 . Introduction Probing the limits of our carbocycle forming methodology. compelled us to investigate the regio- and stereochemical preferences for the cyclization of substituted 6-hepten-1-yltitanocene chlorides, leading to dialkylcyclohexanes. The potential impact of such a study would be significant, if successful. since few methods exist for such a transformation. We were concerned that the entropic factors that denied access to 6-exo—trig cyclizations for other methods may diminish the success of our pursuits. Our primary concerns were the possible repercussions posed by the increased tether length. The increased flexibility of the alkyl tether could provide access to the endo insertion pathway. or at least diminish the diastereomeric preference during carbon-carbon bond formation. Based on the intramolecular-intermolecular polymerization of 1,6-heptadiene, Marvel has provided convincing evidence that endo insertion was a minor consideration.8a The polymer which was obtained consisted of 1.3- dialkylcyclohexane monomer units, although the stereochemical details for this polymerization are undetermined. In our intramolecular Ziegler-Natta carbocyclization study. we found that the optimum conditions for cyclization were highly dependent on the structure of the substrate. However. by judicious choice 91 92 of solvent, Lewis acid, temperature, and concentration, each substrate could be efficiently cyclized to yield methylcyclohexane products. 2. Optimization Studies for the Cyclization of 6-Hepten-1- yltitanocene Chlorides Synthesis of the various 6-hepten-1-yltitanocene complexes paralleled the method discussed previously with yields ranging from 75-100% (Method B, Table 5). A number of differences between the 5-hexen-1-yl and 6-hepten-1-yl systems are noted. First, although not surprising, no significant amount of cyclization resulting from SET could be detected during Grignard formation. This result was consistent with the fact that 6-exo-trig radical cyclizations were much slower (k:5.4 x 103 s“, 25 °C) than the corresponding 5-exo-trig (k:2.3 x 105 5' i. 25 °C) cyclization.17 This fact was emphasized by the low conversions (11- 30%) obtained for the 7-bromo-1-heptene substrates when treated under the tin- hydride conditions (Table 5, Method A) used for the preparation of five- membered rings (Chapter 2). In most cases. transfer of the alkenyl ligand from magnesium to titanium did not cause further cyclization. However. in the case of 160, 30-35% cyclization was observed during transmetallation when performed in the usual solvent (CH2CI2). The extent of ligand cyclization could be controlled (10-20%) by use of toluene as a solvent. This cyclization phenomena paralleled that of 366, as both compounds contained an alkyl group (3 to the metal. These peculiarities will be the subject of Chapter 4. Finally, more forceful conditions were required to promote these cyclizations, which were found to be substrate dependent. For the sake of comparison. reaction times and Lewis acid concentrations increased from 30 minutes/1.0 equiv EtAICI2 to promote five- membered ring formation, to >150 minutes/>20 equiv EtAlCl2 required for six- membered ring formation. What follows is a description of the various reaction 93 conditions used in order to achieve optimum conversion, yield, diastereoselectivity, and minimization of B-hydrogen elimination for each substrate. 0ur optimization study was initiated by attempting the cyclization of the 6- hepten-1-yltitanocene chlorides using the conditions developed for cyclopentane formation (Table 5). Much to our delight, the first substrate chosen (203b) produced an 89% yield of cyclohexane products (Method C); no evidence of methylcycloheptane was observed. The reaction mixture consisted of a 76:17 (trans:cis) ratio of 1.3-dimethylcyclohexanes plus 5% 3-methyl- methylenecyclohexane (the product of B-hydrogen elimination). The amount of B-hydrogen elimination was dependent on the temperature at which the reaction was quenched, but even at -78 °C, approximately 5% B-hydrogen elimination was routinely observed. Treatment of 144b under the conditions of Method 0 resulted in complete conversion (85% yield) to 1,2-dimethylcyclohexanes with excellent stereocontrol (>99:1. transzcis). In this case. no evidence of B- elimination was seen. Similarly, activation of 151b with EtAICI2 (Method C) produced a 79% yield of 1,3-dimethylcyclohexanes in >98:2 selectivity favoring the cis isomer. At this juncture. the only modification required of the previous cyclization conditions was an increase in reaction times. However. treatment of 155b under these conditions produced only 25% cyclization. Because the uncyclized material coincided with the solvent peak by gas chromatographic analysis. we could not determine if the low yield of cyclohexane products was due to incomplete cyclization or some unknown side reaction. Moreover, low conversions were also observed for the cyclizations of 160b and 137b (56% and 27%, respectively). 94 Table 5. Diastereomer Ratios of Cyclohexane Products via Titanium and Free Radical Methods Me R‘ | R‘ I H R‘ R1 Method C —> + + R2 R‘ R2 R‘ R2 R‘ R2 R‘ R3 R3 R3 R3 Cyclization 6 Methyl- Dialkylcvclohexane Products Precursor Method‘ Yieldb Uncyclized 07mm 1.2 (to) 1.3 (to) 1.4 (to) x . W A 84‘ 84 8:5 144.; x . a. B 100 m 1:1 1448; x . rrcopl C m 0 99:1 Me X A 90‘ 81 7:10 1818; x . Br 8 100 100 0.0 181b; x .. Troppl c 79 o 2:98 x AM A 7541 a 20:" M’ 8 1884: x - or 95 9 1 :1 1588; x .mppl C 25 e 7129 Me NW" ‘ 9°“ 8° 79 9888; x . Br 3 100 m 22 2088; x . Troppl C 89' 2 75;; 7 iPr x A a 70 1020 1808; x - 8r 3 E 64 1323 1888; x mom 0 95° 41 3128 x 1978; x . Br 8 73 m 1 1m: X “le9 O 51 a 52 'MethodA: AIBN, Bu,SnH. 0.11.. 8O 110. Mehod 8: (1) Mg. THF, 45 “c. (2) rrcppl, 011,01, Melted c: (1) Method 8, (2) 1.2 equiv. EtAlCl2, Toluene, -78 -c. (3) Hovefher. °nelerence 89. “Numbers represent molar percent d1141.81yleyeloheprane (314) was produced. ° 8-Hydropen eirninetion (51011.) was prowoed 'Cvdohom (1%) was produced 95 Entropic factors were certainly responsible for the slower cyclizations and possibly the decreased yields for certain 6-exo-trig cyclization substrates. The greater flexibility of the alkyl tether apparently diminished the ability of the olefin to become properly aligned with the carbon-titanium bond. This being the case, intermolecular insertions may supersede intramolecular insertions under these conditions. Since the transition state geometry was unclear. the non-bonding interactions experienced by the alkyl group may be more detrimental for the 6- exo-trig cyclization, thus retarding cyclization. The following conditions were systematically varied in order to attenuate intra- versus intermolecular events: (1) concentration (0.01-0.06 M), (2) different solvents of greater polarity, and (3) investigation of higher temperatures. Polar solvents were investigated in an attempt to increase the solubility of the organometallic complex. and to aid in the stabilization of polar intermediates.” Both of these factors could help to increase the rate of cyclization. Higher temperatures were explored to surmount these entr0pic and/or steric activation barriers. Table 6 illustrates a number of conditions used to facilitate the cyclization of 155b. When temperatures higher than -78 °C were used, Me2AlCI was the Lewis acid employed since EtAlCl2 was shown to destroy the starting organometallic complex under these conditions. Entry 1 demonstrated that. in less than 140 minutes at -30 °C, three cyclohexane products were accounted for in 93% yield . The excessive amount of B-hydrogen elimination (34%) made it impossible to determine the selectivity. However. achieving complete conversion for this substrate was gratifying enough at that moment. In an attempt to decrease the amount of methylenecyclohexane produced. cyclization was conducted -50 °C. However, the reaction was substantially slower at this temperatute (26% cyclization. entry 3). When CH2CI2 was used as the solvent under otherwise identical conditions (entry 2), a significant enhancement in the 96 conversion was observed (76% yield) with a concomitant decrease in the (3- hydrogen elimination product (compare entries 1 and 3). Cyclization with Me2AlCI at -78 °C could not be induced in CH2CI2. However. encouraging results were realized with EtAlCl2 under these conditions (entry 4). Although the yield of cyclized material was only 55%. 20% uncyclized material was still present. An increase in reaction time or Lewis acid concentration may help drive this reaction to completion. Under these conditions, selectivity was not observed. Table 6. Optimization of the Cyclization Conditions for 155b Me Me 0leer I a 5 time b ; ¢ % Yield“ 1 55b Me (min) uncyclized M9 M9 M9 (cisnrans) 1) MezAICI (1.5 equiv.). 0 0.98 0.02 0.00 0.00 Toluene, -30 °C, Molarity . 994 M 140 d 0.31 0.28 0.34 m 2) MezAlCl (1.5 equiv.). o 0.98 0.02 0.00 0.00 CI-I20I2 -50 °C. Molarity _ 0.04 M 130 d 0.36 0.33 0.08 76 (50:50) 3) MezAICl (1.5 equiv.), 0 0.98 0.02 0.00 0.00 Toluene, -50 °C. m . 0,04 14 180 d 0.12 0.10 0.04 26 (5050) 4) EtAlClz (1.8 equiv.). 0 0.98 0.02 0.00 0.00 -78 '0. m . 0.04 M 205 0.20 0.25 0.27 0.03 55 (52:48) '(‘I‘fll- W-THF.45’C.(2)TICfiCl-2.OH20I,°Nunbenrepreeentrrillimolrratioeeceledb1for time-0nd peroenhgflorlhefimliime.°Referenoes9.‘Cen‘tbedennninedbeceusethepeekformieprmooincidee withlheeolventpeek. Due to the lack of selectivity encountered for cyclization of 155b, we were interested in seeing what effect a larger alkyl group would have on the diastereoselectivity of bond formation. Table 7 illustrates the results for 164b. As before. excellent conversion occurred at -30 °C with Me2AlCI (entry 1) yet 8- 97 hydrogen elimination (22%) was the fate of a substantial portion of the cyclohexylmethyltitanocene intermediates under the reaction conditions. The longer reaction time in this case was due to the concentration of the reaction. Similar yields were obtained under the conditions of entry 2 with the usual decrease in B-hydrogen elimination. As opposed to the substrate with a corresponding methyl group in this position (155b), 164b underwent facile and efficient cyclization in toluene at -78 °C (entry 4). More importantly, a 74:26 (cisnrans) ratio of 1-(1-methylethyl)-4-methylcyclohexanes was produced. Table 7. Optimization of the Cyclization Conditions for 164b I a A.“ M“ 0mm 5 time $ 9" Yb”: 1 54b 5p, (min) uncyclized” 5p, 3p, iPr ( clearans ) 1) MezAlCl (1.0 equiv.), 0 0.97 0.02 0.01 0.00 Toluene. -30 °C. mm, _ 0.02 M 390 0.10 0.28 0.27 0.22 n 2) MGM. (1.5 mun). 0 0.94 0.04 0.02 0.00 9 & °C. who}, _ 0.05 M 120 0.03 0.25 0.37 0.15 77 3) EWC'z “-5 “WW-)- 0 0.94 0.04 0.02 0.00 0112012, -78 °C. Molan’ty - 0.05 M 150 0.14 0.18 0.50 0.02 69 (74:26) 4) EtAlCl2 (1.5 equiv.), T 0100110, -78 .0. 0 0.97 0.02 0.01 0.00 Molar")! - 0-04 M 150 0.00 0.18 0.52 0.12 82 (7426) ‘(1)1848, w,THF.45°c.(2)T16ppl,.chl,.°Nm18enmpreeanmiundcmecescmdu1for fime-oend peroentegeforthefineltlrne.°ReIefenoe89. The conditions developed for 155b were applied to 160b with the results shown in Table 8. Entry 1 shows that high conversions were possible with a bulky isopropyl group near the metal center. A 79% yield of three cyclohexane products was observed with 35% B-hydrogen elimination as before. Changing 98 solvents and decreasing the reaction temperature (entry 2) resulted in a 68% yield and a marked decrease in B-hydrogen elimination. Finally. cyclization in CH2CI2 at -78 °C (entry 3) resulted in low conversions (<55%). but the diastereoselectivity was >99:1 during the Ziegler-Natta insertion reaction. Table 8. Optimization of the Cyclization Conditions for 160b a Me Me,“ CIC'iPzTl time ‘9 ¢p % Yieldc ' iPr iPr iPr 1 60b iPr (min) uncyclizedb (cisnrans ) 1) MezAlCl (1.5 equiv.), 0 0.93 0.05 0.02 0.00 Toluene, -30 °C, Molar", .. 0.07 M 180 0.04 0.15 0.30 0.34 79 2) MezAlCl (1.5 equiv.), 0 0.69 0.20 0.11 0.00 0112012. -50 °C. Mommy .. 0.04 M 300 0.07 0.29 0.27 0.11 8 3 EtAICI 1.5 uiv. )Tomnifqat‘é' ) 0 0.69 0.20 0.11 0.00 MolarIty - 0.04 M 150 0,23 0.20 0.31 0.05 84 (199) '(1) 1808. My. THF. 45 °c. (2) noppl, Toluene. ”Numbers represenf milirnoler ratios scaled b 1 for time-0 and percentage Iorthefinel time. ‘Referenoe 69. Finally, Table 9 contains the results for the cyclization of 137b. Since alkyl substituents were absent on the tether. this substrate was expected to cyclize even more slowly than the other 6-hepten-1-yl substrates. Under the conditions for the 5-exo-trig cyclization, only 27% cyclization occurred along with 24% uncyclized material (Table 5). Once again, entry 1 showed that high yields and conversions were possible at higher temperatures but a large amount of 8- elimination rendered these conditions useless. A decrease in the concentration (0.044 M vs 0.124 M) resulted in a marked enhancement in conversion. yet only a 56% yield of methylcyclohexane was produced. An increase in the polarity of "‘9 reaction medium (entries 4 and 5) revealed that moderate yields but good °°nversions were possible at -78 °C. To combat the intermolecular insertions. 99 which were probably responsible for the low yields, substantially lower concentrations were warranted. Table 9. Optimization of the Cyclization Conditions for 137b I a MC Cl 11 °"’ 15 c 65 (min) uncyclizedb 1%. Yielo° 137b 1) Me c1(1.5equiv.), Tolficlm’ .30 °C 0 0.99 0.01 0.00 WISH” I- 0.05 M . 125 0.04 0.71 0.19 90 2 EtAIC 1.5 uiv. ) “my”? I 0 0.99 0.01 0.00 Molarlty' . 0.12 M 210 0.24 0.27 0.00 41 3) EtAICI (2.0 equiv.), Tom": .78 °c 0 0.98 0.02 0.00 Molarity . 0.04 M 120 0.00 0.56 0.00 56 4) EtAlClz (15 equiv.) -78 °C 0 0.98 0.02 0.00 Toluene:CH2Cl2 (1:1) Molarity- 0.03 M 150 0.04 0.64 0.00 64 5 12 EtAI 12 uiv.. ) 0.432, .78¢93( .q ) 0.98 0.02 0.00 Molarity . 0.04 M 105 0.07 0.52 0.00 52 °(1) 13711, My. THF, 45 -c (2)1'10ppl,. 011201, °Humbert: representmiliimolar rallossceled 81 IortIme-Oend percentage for the final time. cReference 69. - With optimal conditions obtained for each substrate, attention was focused on performing these cyclizations on a synthetically more useful scale.“ Synthesis of the requisite 6-hepten-1yltitanocene complexes was accomplished without incident (Table 10. Method A). Activation of 137b with EtAlCl2 under high dilution conditions (0.01 M, CH20I2. -78 °C), resulted in exclusive exo cIfolization to provide methylcyclohexane in modest yield (76% yield). CYcIization of 144b and 151b were efficiently achieved under much more c=0hcentrated conditions (0.12 M. toluene, -78 °C) and provided essentially a 100 single stereoisomer in each case, 99:1 and 3:97 (trans:cis), respectively. Although the y-methyl group of 155b had no bearing on the selectivity of bond formation (50:50). the increased steric demands of the isopropyl group of 164b significantly enhanced the production of the cis diastereomer (trans:cis, 23:77), with a paralleled increase in yield (91%). Cyclization of 203b and 160b were unique compared to the others, as B-elimination seemed to be a more accessible pathway (9% in each case). Compound 203b showed a useful preference for the trans isomer (81 :19). Metallation of 1608 with magnesium provided a high yield of the Grignard intermediate with no evidence of cyclization. However, transfer of the alkenyl ligand to titanium resulted in 17% cyclization, even with THF present. The most interesting observation was that M9X2 was selective for the opposite isomer (trans:cis, 1:2) than produced with an aluminum cocatalyst. Cyclization of 160b provided three cyclohexane products (71 :18:9). If the extraordinary amount of cyclization that transpired during transmetallation was factored out, a 92:8 preference for the trans isomer was produced during the Ziegler-Natta olefin insertion process. 101 Table 10. Diastereomer Ratios of Cyclohexane Products Performed Under the Optimized Conditions Me R‘ I Ft1 I H R1 R1 Memod A and B + + ———_—> n2 n4 82 n‘ R2 R‘ 1:12 81 R3 R3 R3 R3 Cyclizan’on MEW- Dialkylcyolohexanes Precursor MethodI Yield" Uncyclized‘= cyclohexane 1,2 (czt) 1.3 (on) 1.4 (on) x M 1379 x . Br A 73 m 1 1378 x . Trcppi e 76 3 97 NWT Me 1445 x . Br A e m 29 1448 x .TICppl B G <1 99:1 Ms AMX 1519 x . Br A 82 97 03 1518 x : Troppl B 91 <1 3:97 WK Me 15511 x . Br A as 97 2,, 1558 x : Troppl 8" 63 <1 50:50 Wx iPr 1548 x . Br 8' 8) 97 2:1 1848 x .Trcppl 8 91 3 23:77 Mi IM" / 2088 X - Br A 97 8 2:2 2088 x . Trcppi 8 72° 2 74:17 iPr ' M" 1809 x . Br A 74 a 6:11 108 x .‘ricppl 8° 88° 2 71 :18 ‘MefhodA: (1) Bromide, Mg.THF,45°C. (2) TiCp-filz. 011,01, W8: (1)1.75-295 equiv. 5111101,, Toluene.-78 -c.(3) HCIfether. °neference 69. °N1snoersrepresentrsficsscsletoloo. °ch12msusedss aeohnnt'Coddnotbedebrmined. Findyieldoelwlabdfromthebromide. 'Toluenewnmedesasoivent. °8-I-Iyd'ogen elimination (9%) w. observed. 102 These results illustrate for the first time a useful method to affect 6-exo-trig cyclizations into unactivated alkenes with synthetically satisfying selectivities. In addition, a number of useful pieces of information have been acquired about the Ziegler-Natta process by examination of the alkene facial selectivity of a single insertion. A methyl group was shown to have a pronounced effect during the syn-coplanar insertion for 1448 and 1518. In contrast, the presence of the methyl group yto the metal (155b) had no bearing at all on the stereogenicity of the new asymmetric center. These results paralleled those of the polymerization of racemic or optically active of-olefins.73 For example. 3-methyl-1-pentene and 4-methyl-1-hexene induced a high degree of stereoselection during polymerization. yet 5-methyl-1-heptene had little effect on the facial selectivity. The stereochemical outcome for the cyclization of 203b, provided evidence that the polymerization of 1.6-heptadiene contained primarily a trans 1,3 relationship. Finally, the stereoselective divergence between the different cocatalysts. M9X2 and EtAlCl2, suggested that the catalyst-cocatalyst interaction is more sophisticated than simple generation of [Cp2TlR+]. 3. Potential Explanations for the Observed Stereoselectivities By examination of molecular models. three possible transition states for Cyclization seemed reasonable. Scheme 30 illustrates these possibilities: chair (204), boat (206). and twist-chair (208). Each transition state allowed for the alkyl groups to be directed in psuedo-equatorial positions to minimize unfavorable non-bonding interactions. The chair transition state was less likely Since the desirable syn-coplanar alignment of the olefin and the carbon-titanium bond was precluded. The boat and twist-chair conformations were able to accommodate this transition state requirement. Severe eclipsing interactions 103 between the substituents at the B and 7 positions (relative to titanium) may tend to disfavor the boat transition state. Examination of the products formed from the substrates with various substitution patterns on the alkyl tether, however. did not allow for differentiation between the boat and the twist-chair. Both conformations accurately predicted the major product for substitution at the allylic and homoallylic sites. The apparent similarity between the steric environment surrounding the 8 and 7 positions suggested that a methyl substituent would have similar effects in both cases. However, y-methyl substitution produced no selectivity. Differences in selectivity between the B- and y-methyl compounds was attributed to the increased steric interactions at the B-position due to the closer proximity to the congested metal center. The steric environment created by the cyclopentadienyl ligands on titanium and the Lewis acid-chloride-titanium complex strongly influenced the axial or equatorial orientation of neighboring alkyl groups. 104 Scheme 30. Possible Transition States for Slx-Membered Ring Formation H RI H olezn H EtAlClz H | Fl“ “2 n4 H H 205 H R‘ “a 112 R1 EtAlCI2 H —-—> H H ole21'1 TWIST-CHAIR 208 209 4. Conclusion The results in this chapter describe a powerful new method for the construction of six-membered rings. This method represents the most successful procedure to affect 6-exo-trig cyclization from an sp3 carbon with an unactivated alkene. The need for a syn-coplanar orientation of the alkene and the carbon- titanium bond had a significant impact on the regio- and stereoselectivity. Olefin insertion was found to proceed with complete exo specificity, which resulted in the formation of a new asymmetric center. In most cases, an alkyl group induced a high degree of alkene facial selectivity, providing synthetically useful diastereomeric excesses. Although a single general procedure was not 105 realized. a variety of new conditions were developed along the way which will provide additional avenues to affect the desired bond construction. This method furnishes the synthetic chemist with another useful tool for the construction of five- and six-membered rings with the high degree of stereocontrol demanded by today's complex synthetic targets. 106 5. Experimental Radical Cyclization of 7-Bromo-1-heptene (1378). The following reagents were combined and heated to 80 °C in a sealed tube for 10 hours: bromide 1378 (88.5 mg. 0.5 mmol), AIBN (10 mg). Buaan (174.6 mg. 0.6 mmol), and benzene (10 mL). Analysis of the reaction mixture (see Table 5. Method A)69 indicated that a combined yield of 96% was obtained which consisted of the following compounds: 1-heptene (89%), methylcyclohexane (10%), and cycloheptane (1%). Radical Cyclization of 7-Bromo-3-methyI-1-heptene (1448). The following reagents were combined and heated to 80 °C in a sealed tube for 10 hours: bromide 1448 (95.5 mg, 0.5 mmol), AIBN (10 mg), Buaan (174.6 mg, 0.6 mmol), and benzene (10 mL). Analysis of the reaction mixture (see Table 5, Method ill)69 indicated that a combined yield of 84% was obtained which consisted of the following compounds: 3-methyI-1-heptene (84%). cis- and trans- 1.2-dimethylcyclohexanes (13%. 5:8, respectively), and methylcycloheptane (3%). Radical Cyclization of 7-Bromo-4-methyl-1-heptene (1518). The following reagents were combined and heated to 80 °C in a sealed tube for 10 hours: bromide 1518 (95.5 mg, 0.5 mmol), AIBN (10 mg). Bu38nH (174.6 mg. 0.6 mmol). and benzene (10 mL). Analysis of the reaction mixture (see Table 5, Method A)69 indicated that a combined yield of 90% was obtained which consisted of the following compounds: 4-methyl-1-heptene (81%). cis- and trans- 1,3-dimethylcyclohexanes (17%. 10:7. respectively). and methylcycloheptane (2%). Radical Cyclization of 7-Bromo-5-methyI-1-heptene (1558). The following reagents were combined and heated to 80 °C in a sealed tube for 10 hours: bromide 1558 (95.5 mg, 0.5 mmol), AIBN (10 m9). Bu38nH (174.6 mg, 0.6 107 mmol). and benzene (10 mL). Analysis of the reaction mixture (see Table 5, Method A)69 indicated a combined yield of 75% was obtained which consisted of the following compounds: 5-methyl-1-heptene (66%). cis- and trans-1,4- dimethylcyclohexanes (31%, 11:20. respectively). and methylcycloheptane (3%). Radical Cyclization of 7-Bromo-6-methyl-1-heptene (2038). The following reagents were combined and heated to 80 °C in a sealed tube for 10 hours: bromide 2038 (95.5 mg. 0.5 mmol), AIBN (10 mg), Buaan (174.6 mg. 0.6 mmol), and benzene (10 mL). Analysis of the reaction mixture (see Table 5. Method A)69 indicated a combined yield of 84% was obtained which consisted of the following compounds: 6-methyl-1-heptene (83%), cis- and trans-1.3- dimethylcyclohexanes (16%, 9:7, respectively), and methylcycloheptane (1%). Radical Cyclization of 7-Bromo-6-(1-methylethyI)-1-heptene (1608). The following reagents were combined and heated to 80 °C in a sealed tube for 10 hours: bromide 1608 (110 mg. 0.5 mmol). AIBN (10 mg). Buaan (174.6 mg, 0.6 mmol). and benzene (10 mL). Analysis of the reaction mixture (see Table 5, Method A)69 indicated a combined yield of 88% was obtained which consisted of the following compounds: 6,7-dimethyl-1-octene (70%), cis- and trans-t-(t- methylethyI)-3-methylcyclohexanes (30%, 20:10. respectively). General Procedure for ZIegler-Natta Cyclization for 7-Bromo-1- heptene Substrates: A Schlenk flask was charged with M9 (195 mg, 8 mmol) and flame-dried under vacuum. After cooling to ambient temperature. the flask was purged with argon. THF (2 mL) was added. and the flask was then placed in a 45 °C oil bath. The bromide (2 mmol) was added to the magnesium with a gas- tight syringe over 6 hours, followed by stirring for an additional 4 hours. A separate Schlenk flask was loaded with Tle2Cl2 (598 mg, 2.4 mmol). evacuated and purged with argon, the Tle2Cl2 was suspended CH2CI2 (8 mL). and the flask was cooled to -45 °C. Transmetallation was accomplished by adding the Grignard 108 reagent to the TiCp2Cl2 suspension via cannula (an additional 1 mL THF was used to rinse the Mg). stirring for 30 minutes at -45 °C, and warming to ambient temperature where the reaction mixture remained for an additional 5 hours. After transmetallation was complete. the mixture was concentrated to approximately 4 mL, taken up in toluene (5 mL) and hexane (5 mL), stirred for 10 minutes at room temperature, and then cooled to 0 °C for 10 minutes. The entire mixture was transferred via cannula and filtered under argon through a glass frit. The solids were rinsed with additional toluene (3 x 10 mL) and filtered. Removal of the solvents in vacuo left behind the organotitanocene chloride as a red slurry. Toluene was then added and the solution was cooled to -78 °C. In general. EtAlCI2 was added to the alkyltitanocene chloride solution by three sequential additions over a period of 90 minutes. When the reaction was complete. 1.0 M HCI/Et20 (10 mL) was added at -78 °C to quench the reaction. General Procedure for the Optimization of Cyclization Conditions. Preparation of the alkyltitanocene chloride complex was the same as described in the general procedure unless otherwise stated. Typically. 2 mL of the alkyltitanocene chloride complex was added to a flame-dried flask, diluted to the desired molarity, and cooled to the appropriate temperature. The Lewis acid to be used was added over 10-15 minutes and the progress of the reaction was monitored by gas chromatography. For optimization results for specific substrates, see the following tables: Table 6 (155b). Table 7 (1648), Table 8 (1608), and Table 9 (137b). Ziegler-Matte Cyclization of 7-Bromo-1-heptene (137a). Bromide 1378 (353.8 mg, 2.0 mmol) was converted to alkyltitanocene 137b in 73% yield following the general procedure. The product mixture prior to addition of the cocatalyst consisted of 1-heptene and methylcyclohexane (99:1), respectively (see Table 10, Method A). The titanium residue, after filtration and evaporation of 109 solvents, was diluted with toluene (200 mL). cooled to -78 °C, and 1.0 M EtAlCl2 (2.25 eq) was added sequentially as follows: 1.0 equiv. (t:0 min). 0.75 equiv. (t:45 min). and 0.50 equiv. (t:90 min). Analysis of the final reaction mixture after hydrolysis (Table 10, Method B) showed that a 76% yield for the Lewis acid assisted insertion (56% yield from the bromide) was obtained which consisted of the following compounds: methylcyclohexane (97%). heptene (3%). and methlyenecyclohexane (0%). Ziegler-Matte Cyclization of 7-Bromo-3-methyI-1-heptene (1448). Bromide 1448 (381.8 mg. 2.0 mmol) was converted to alkyltitanocene 144b according to the general procedure. The product mixture prior to addition of the cocatalyst consisted of 3-methyl-1-heptene and cis- and trans-1.2- dimethyloyclohexanes (98:2:0), respectively (see Table 10. Method A). The titanium residue, after filtration and evaporation of solvents. was diluted with toluene (15 mL), cooled to -78 °C, and 1.0 M EtAICI2 (2.25 eq) was added sequentially as follows: 1.0 equiv. (t:0 min). 0.75 equiv. (t:45 min). and 0.50 equiv. (t:90 min). Analysis of the final reaction mixture (Table 10, Method B), after hydrolysis, showed that an 89% yield overall from the bromide was obtained which consisted of the following compounds: trans-1.2-dimethylcyclohexane (98%), cis-1,2-dimethylcyclohexane (1 %). and 2-methyl-1-methlyenecyclohexane (<1%). ZIegIer-Natta Cyclization of 7-Bromo-4-methyI-1-heptene (151a). Bromide 1518 (353.8 mg, 2.0 mmol) was converted to alkyltitanocene 151b in 82% yield following the general procedure. The product mixture prior to addition of the cocatalyst consisted of 4-methyl-1-heptene and cis- and trans-1,3- dimethylcyclohexanes (97:0:3), respectively (see Table 10. Method A). The titanium residue. after filtration and evaporation of solvents, was diluted diluted toluene (15 mL), cooled to -78 °C. and 1.0 M EtAICI2 (2.25 eq) was added 110 sequentially as follows: 1.0 equiv. (1:0 min), 0.75 equiv. (1:45 min), and 0.50 equiv. (t:90 min). Analysis of the final reaction mixture after hydrolysis (Table 10. Method B) showed that a 91% yield for the Lewis acid assisted insertion (75% yield from the bromide) was obtained which consisted of the following compounds: trans-1.3-dimethylcyclohexane (3%), cis-1,3-dimethylcyclohexane (97%), and 3-methyl-1-methlyenecyclohexane (<1%). Ziegler-Natta Cyclization of 7-Bromo-5-methyI-1-heptene (1558). Bromide 1558 (353.8 mg. 2.0 mmol) was converted to alkyltitanocene 155b in 65% yield following the general procedure. The product mixture prior to addition of the cocatalyst consisted of 5-methyl-1-heptene and cis- and trans-1,4- dimethylcyclohexanes (97:2:1), respectively (see Table 10. Method A). The titanium residue, after filtration and evaporation of solvents, was diluted with CH2CI2 (45 mL). cooled to -78 °C, and 1.0 M EtAlCI2 (2.25 eq) was added sequentially as follows: 1.0 equiv. (i=0 min), 0.75 equiv. (t:45 min). and 0.50 equiv. (1:90 min). Analysis of the final reaction mixture after hydrolysis (Table 10, Method B) showed that a 63% yield for the Lewis acid assisted insertion (41% yield from the bromide) was obtained which consisted of the following compounds: trans -1,4-dimethylcyclohexane (50%). and cis-1 .4- dimethylcyclohexane (50%). No detectable amount of 4-methyI-1- methlyenecyclohexane or 5-methylhept-1-ene was observed. Ziegler-Nana Cyclization of 7-Bromo-5-(1-methylethyI)-1-heptene (1648). Bromide 1648 (218.9 mg, 1.0 mmol) was converted to alkyltitanocene 164b in 50% yield following the general procedure. The product mixture prior to addition of the cocatalyst consisted of 5-ethyl-6-methyl-1-heptene and cis- and trans-1,4-isopropylcyclohexanes (97:2:1), respectively (see Table 10, Mehtod A). The titanium residue, after filtration and evaporation of solvents, was diluted with toluene (12 mL). cooled to -78 °C, and 1.0 M EtAlCl2 (2.25 eq) was added 111 sequentially as follows: 1.0 equiv. (t:0 min). 0.75 equiv. (t:45 min), and 0.50 equiv. (t:90 min). Analysis of the final reaction mixture after hydrolysis (Table 10, Method B) showed that a 91% yield for the Lewis acid assisted insertion (47% yield from the bromide) was obtained which consisted of the following compounds: cis-1-(1-methylethyl)-4-methylcyclohexane (74%). trans-1-(1- methylethyl)-4-methylcyclohexane (22%), and 5-ethyl-6-methyl-1-heptene (3%). No detectable amount of 4-(1-methylethyl)-1-methlyenecyclohexane was observed. ZIegIer-Natta Cyclization of 7-Bromo-6-methyI-1-heptene (203a). Bromide 2038 (381.8 mg, 2.0 mmol) was converted to alkyltitanocene 2038 in 97% yield following the general procedure. The product mixture prior to addition of the cocatalyst consisted of 6-methyl-1-heptene and cis- and trans-1,3- dimethylcyclohexanes (96:2:2), respectively (see Table 10, Method A). The titanium residue, after filtration and evaporation of solvents. was diluted with toluene (12 mL). cooled to -78 °C. and 1.0 M EtAlCl2 (2.25 eq) was added sequentially as follows: 1.0 equiv. (t:0 min), 0.75 equiv. (t:45 min), and 0.50 equiv. (1:90 min). Analysis of the final reaction mixture after hydrolysis (Table 10, Method B) showed that a 72% yield for the Lewis acid assisted insertion (63% yield from the bromide) was obtained which consisted of the following compounds: trans -1,3-dimethylcyclohexane (74%). cis-1.3-dimethylcyclohexane (17%). 3-methyI-1-methlyenecyclohexane (7%), and 6-methylhept-1-ene (2%). ZIegIor-Natta Cyclization of 7-Bromo-6-(1-methylethyI)-1-heptene (1608). Bromide 1608 (437.8 mg. 2.0 mmol) was converted to alkyltitanocene 160b in 74% yield following the general procedure except that the transmetallation was performed in toluene. The product mixture prior to addition of the cocatalyst consisted of 6,7-dimethyl-1-octene and cis- and trans-1-(1-methylethy)-3- methylcyclohexanes (83:11z6), respectively (see Table 10. Method A). The 112 titanium residue after filtration and evaporation was diluted CH2CI2 (30 mL), cooled to -78 °C, and 1.0 M EtAlCl2 (2.25 eq) was added sequentially as follows: 1.0 equiv. (i=0 min). 0.50 equiv. (t:120 min). and 0.25 equiv. (1:240 min). Analysis of the final reaction mixture after hydrolysis (Table 10. Method B) showed that a 88% yield for the Lewis acid assisted insertion (63% yield from the bromide) was obtained which consisted of the following compounds: trans-1-(1- methylethy)-3-methylcyclohexane (70%). cis-1-(1-methylethy)-3- methylcyclohexane (18%), 3-(1-methylethyl)-1-methylenecyclohexane (9%). and 6,7-dimethyl-1-octene (2%). CHAPTER 4. MECHANISTIC STUDIES OF ZIEGLER-NATTA POLYMERIZATION: PROBING FOR a- OR B-HYDROGEN ACTIVATION 1. Introduction During the development of our titanocene mediated ring-forming methodology. we observed peculiar behavior for both the 5-hexen-1-yl- and 6- hepten-1-yltitanocene chloride complexes when an alkyl group was in the )3 position with respect to the metal. For instance, reaction of a Grignard solution of 368 in diethyl other with titanocene dichloride resulted in nearly complete cyclization in less than one hour (eqn. 20). To fulfill our synthetic objectives at the time, we found that olefin insertion during transmetallation could be suppressed if the reaction was run in the presence of THF. This result was presumably due to the higher basicity of THF. which allowed for stronger complexation with M9X2 in comparison to the corresponding diethyl ether complex. Substrates 368-b did not cyclize to any extent under these conditions. In an analogous situation, the M9X2 salts catalyzed the cyclization of 160b to an extent of 35%, even with THF present (eqn. 21). The most interesting aspect of this latter example was that these conditions promoted olefin insertion with an opposite stereochemical preference (cisrtrans, 2:1) compared to that promoted with EtAlCl2 (cisrtrans. 8:92). These results have launched us into an exploration of the mechanism of Ziegler-Natta polymerization. 113 114 231985881. M I p; . o uone M e W Br (3) HCI 2 W H + 90, Me > (20) 364: Me 15:85 (trans:cis, 75:10) iPr (1) M9. 5120 II" M" iPr /\/\)\/B (2)1‘leZCl2. Tom“ NVk/ U / I (3) HCI ’ / H + (21) 160 65:35 (trans:cisJ 1 :24) The apparent rate enhancement, observed only with an alkyl group present in the position 8 to the metal. suggested the possible intervention of an electronic effect. Possible explanations for these observations included either 01- or 8- hydrogen activation (Scheme 31 ).74 During both a-hydrogen activation (210) and B-hydrogen activation (211). the cocatalyst presumably complexes the chloride atom on titanium, causing an increase in the positive charge on titanium and a simultaneous polarization, and thus weakening, of the carbon-titanium bond. Based on Figures 1-3 (see Introduction). there are precedents that either 01- or 8- hydrogens could interact with the metal center. The nature of this interaction is unclear, but assuming the interaction would be to stabilize the metal center, a partial negative charge on hydrogen would be necessary. Also. the possible interaction of the cocatalyst with either the 01- or B-hydrogen cannot be ruled out. The net result would be the development of positive charge on either the tit-carbon (210) or the 8-carbon (211) depending on the mechanism. The role of the B—alkyl group could be to stabilize the development of the positive charge by either hyperconjugation (210) or through induction (211). Both 01- or B-hydrogen activation could reasonably account for an increased rate of cyclization. 115 Scheme 31 . Mechanistic Possibilities for ZIegIer-Natta Polymerization: 01- and [ii-Hydrogen Interactions a”’ 5- 5+ 5+ R CPzTI ‘ 4" := := ‘/ c“j4 CI I e I d ‘ M1... 210 211 To test these potential pathways, we designed a series of competitive rate experiments (Scheme 32). Our idea resembled aspects of independent 9° in which the relative strengths experiments reported by Grubbsga and Bercaw, of carbon-hydrogen and carbon-deuterium bonds were central to the design. The major difference in our approach was how the deuterium isotope effect was to be measured. Both Grubbs and Bercaw used the stereochemistry of the cyclized products as an instrument for measuring an isotope effect. Our intentions were to compare the relative rates of olefin insertion of two similar alkenyltitanocene complexes in which one contained a hydrogen atom (01 or 8) and the other a corresponding deuterium atom (01 or 8). Ideally, the most informative reaction would be to directly monitor the relative rates of cyclization between 2128 and 2138 where R1:R2. However, we intended to monitor the cyclization progress using gas chromatographic techniques which would make this approach impossible. Alternatively, an indirect method for obtaining the same information is outlined in Scheme 32 in which R1:R2. First, cyclization of a mixture of 2125 and 2138 (R‘s-R2) will determine the minor role of the B-alkyl substituents on the relative rates of olefin insertion. By graphing the conversion versus time for each substrate, the relative rate of cyclization can be obtained from the ratio of the slopes [km-slope(2128/213a)]. Secondly. 212b and 2138 will be treated under the identical conditions and the resultant conversions graphed in a similar fashion. 116 In this way, kH/ko can be obtained by the following ratio: [slope(212a/213a)]/[slope(2128/213a)]. Finally, the competitive cyclization of 2128 and 213b provides a control for the second cyclization. A similar series of competitive reactions will also be performed in which a,01-dihydrogen substrates will be compared to or,0:-dideuteron substrates. These reactions outlined above could provide insight into the fundamental processes of Ziegler-Natta polymerization. Scheme 32. Competitive Rate Experiments Designed to Test the Mechanistic Possibilities ClC 11E: ) ClC nu (”Mead Mega, Mag 92 H R‘+ 92 H R2—> (-—>2)H. H2 2128 213a ClezTif: ) ni- ClezTif: P 82,—. (1)Lewisacid M°§_RH NIKE—Ra D H (2)H 212b 2138 CIC Tin CIC 1; : >2 (1)LewisacidMe‘21q:—W mxé— H D 2128 2138 2152 2. Preliminary Findings for the 2-AIkyl-5-hexen-1-yltltanocene Chlorides Our initial investigation involved competitive rate experiments in which cyclization was to be facilitated under the transmetallation conditions (Scheme 33). Two major problems were associated with this system. First of all, quenching an aliquot of the Grignard intermediates revealed that 40% of the reaction mixture from 129 were cyclopentane products compared to 20% for 36c. Secondly. transmetallation of these Grignard species to 'I”le2Cl2 resulted in rapid (<20 minutes) cyclization upon warming to ambient temperature. When colder temperatures were maintained throughout the transmetallation process. 117 conversion did not occur to as great an extent. One major problem, SET cyclization during Grignard formation. could be avoided if THF was used as a solvent. but this also inhibited the effectiveness of M9X2 as a Lewis acid and decreased the amount of cyclization. The problems associated with this system suggested that extraction of the pertinent rate information would be difficult. Specifically, the time window for monitoring the reaction progress was too small. Scheme 33. Cyclization of 5-Hexen-1-yltltanocene Chlorides Performed Under the Transmetallation Conditions + ——> + + 9' R1 Me Pr M0 Me Pr Pr 36c : RI-H 129 (1) Mg, Etp 80 20 62 38 (2) TICQCIZ. Toluene 25 75 13 87 -45 to 25 °C 125 :R‘:D 129 (1) M9. 5120 80 20 62 38 (2)1'iCmCI2, Toluene 80 20 62 38 -45 °C In order to avoid the rapid and uncontrollable cyclization discussed above. we opted to pursue 7-bromo-1-heptene substrates. We have demonstrated that these substrates cyclized to a lesser extent during Grignard formation, and the corresponding titanocene derivatives undement 6-axo-trig insertion more slowly than the corresponding 6-bromo-1-hexene compounds. Therefore, this modification has the potential for overcoming both of the problems associated with the cyclization of 6-bromo-1-hexene substrates. which should increase the window of observation with a simultaneous increase in the accuracy and precision of the data. 118 3. Optimizatlon Studies for the Competitive Cyclization of 2-AIkyI-6-hepten- 1-yltltanocene Chlorides with MezAICI, M9X2, and MAO Our first task was to develop reproducible cyclization conditions using the aforementioned Lewis acids that would provide high yields of cyclic products. concurrent with a large window of time (90-180 minutes) to view the linear conversion to cyclic products. In order to graph conversion versus time and interpret the results with a high degree of certainty, we needed to accumulate enough data points (5-8) to minimize the error. These considerations are legitimate since an isotope effect in a similar system was reported to be kH/ko:1.2.9c The variables that were systematically changed include the Lewis acid. equivalents of Lewis acid. solvent, concentration. and temperature. We chose to investigate three Lewis acids for these competitive rate experiments. Me2AlCI. M9X2, and MAO. The interesting reversal in stereochemistry using M9X2 as Opposed to Me2AlCI or EtAlCl2, suggesting a different mechanism, coupled with the propensity of this catalyst to provide isotactic polypropylene made M9X2 ideal for our investigation. Although the structure of this catalyst is unclear. MAO also furnished isotactic polypropylene which compelled us to study its behavior. 119 Table 11. Cyclization of Unpurifled 175b Using M9X2 or MAO a Me I noppl I H “.11 H Pr time I'I pr p H (min) P' (cis/trans) r % Yieldc 1) ngz (x.c1 or Br, 0 96 371 o 1-0 equiv-l. Toluene. 420 41 39/8 4 d 45 °c. Molarity - 0.02 M 2) MgCI2(8.0 equiv.) + o 96 371 0 )lflolxz (X-Sls or 82'" c 210 44 24720 5 93 o uone, to 5 ° . Molar” _ 0.02 M. 590 14 33721 14 82 3) MgCl2(8.0 equiv.) + 0 96 3/1 0 M9X2 (X-CI or Br). 150 10 37716 23 86 T°'U°_fl°- ~45 I0 50 °C- 410 10 27715 24 76 Molarity . 0.02 M, 4) MAO (2.0 equiv.), + 0 96 371 0 M9X2 “'31:,” 322°C 80 46 25724 5 100 O UORO, 10 . Molamy _ 0.02 M. 330 14 46/19 18 97 I(1) 1758 . Mg. THF. 45 °C, (2)11CQCI2, Toluene. I’Numbers represent millimolar ratios scaled to 100 for time-0 and percentage for the final time. °Reference 69. 6No internal standard was used in this reaction. Our initial efforts for promoting the cyclization of 175b saw a return to magnesium dihalides, either with the M9X2 formed in situ during transmetallation or with an excess of the salts (Table 11). The alkyltitanocene chloride 175b illustrated in Table 11 was prepared by the addition of a diethyl other solution of the Grignard reagent to TiCp2Cl2 in toluene. After transmetallation was complete, the alkyltitanocene complex was not purified from the magnesium salts, but rather, the diethyl ether was removed in vacuo and the mixture diluted with CH2CI2. Entry 1 illustrates the effect of one equivalent of M9X2 on the cyclization process. After 4 hours at 25 °C, no cyclization was detected. However, heating this mixture to 45 °C led to 51% cyclization after 7 hours, at which time the reaction ceased. These conditions most notably produced a 5:1 ratio of 120 diastereomers selective for the cis isomer. In an attempt to increase the amount of conversion, excess MgCl2 (8 equiv.) was employed at 25 °C (entry 2) and 50 °C (entry 3). As expected. higher temperatures facilitated conversion, yet product loss resulted after prolonged exposure to the reaction conditions. Interestingly, the use of excess M9X2 caused a dramatic decrease in selectivity and also an increase in B-hydrogen elimination. Finally. entry 4 demonstrates the behavior of MAO (2 equiv.) in combination with M9X2 (1 equiv.). The desirable features of this reaction were the high yield, near complete conversion, and the extended period of time required for conversion. Although these conditions appeared ideal for our study. the slow steady conversion proved difficult to repeat at this temperature. A limitation imposed by avoiding purification of the alkyltitanocene complexes before introduction of the cyclization catalyst was that occasionally the effect of concentrating the alkyltitanocene complexes in the presence of M9X2 caused nearly complete cyclization. In order to alleviate this problem. all reactions henceforth were performed with a "magnesium free“ alkyltitanocene solution, which was combined with a solution (or suspension in the case of M902 and MgBr2) of the Lewis acid. Table 12. Cyclization of Purified 1688 and 175b Uslng M9X2 121 a; rcppi (I/F:Cp:01 8:2-O- .I'E R‘ Eta“ time "p % Yeld° (min) 8' (cis/trans)R (cis/trans) Et/Pr 1) MgCl2(8.0 equiv.). 0 87 875 89 774 1.2-DOE. 0 °C. 490 74 1 878 74 1576 100/95 Molarity I 0.02 M RI-H; Fri-H 2) MgCl2(8.0 equiv.). 0 87 875 89 774 1,2-DCE. 45 °C. 120 35 38720 44 33715 93788 Molarity . 0.02 M, R‘.H; #:H 3) MgCl2(8.0 equiv.). 0 94 472 95 372 Toluene, 25 °C, 180 70 976 75 1076 85791 Molarity - 0.02 M, 255 52 16710 54 16710 78780 R‘IH; #IH 4) MgCI2(10.0 equiv.). o 84 1076 83 1077 Toluene:CI-§CI2 (3:1 ). 490 50 22713 46 18712 85776 25 “C. Molarity I 0.02 M, RI-H; 1?an I5) MgCI2(10.0 equiv.). 0 82 1177 83 1077 Toluene:CI-I2CI2:Et20 60 65 1479 66 14710 88790 (251.005). 25 °C, 420 45 20713 46 18713 78777 Molarity I 0.02 M. R‘IH; RZIH I6) MgBr2 (2.0 equiv.). 0 92 573 92 573 ToluenezEt20 (>955). 420 62 1379 60 14710 84784 25 °C, Molarity - 0.02 M. H‘sH; RZ-H 7) MgCI2(2.0 equiv.). 0 92 573 92 573 MgBr2 (2.0 equiv.). 90 78 1176 75 976 95790 Tdugng;thC|2 230 46 22,20 51 16/14 88/81 (15:19). 25 11c, 350 22 27729 31 23714 78768 Molarity - 0.02 M, R‘IH; RQIH ‘(1)1888 8nd1788 . Mg, THF. 45 °C. (21'IszCI,.Tquene."Number8 represent millimoler r8808 scaled to 100 for time-Omdperoentageforihefinel IIme.°Referen0869. 122 Using the ”magnesium free" alkyltitanocene solution, we were persistent in our quest to optimize M9X2 promoted cyclizations. The solvent 1,2- dichloroethane has been shown to be an effective solvent for olefin polymerization. Entries 1 and 2 (Table 12) illustrate the combined effects of MgCI2 (8 equiv.) in 1.2-dichloroethane at either 0 °C or 45 °C. Although no cyclization was observed at 0 °C, promising results were achieved at higher temperatures. yet the reaction ceased after two hours. Entries 3-7 represented systematic variations in the reaction conditions in an attempt to reproduce the results obtained in Table 11, in which the organometallic complex was not isolated. During those particular cyclizations, toluene and CH2CI2 were present, however, the exact amount of diethyl ether and the precise composition of M9X2 (X:Br or Cl) was not clear. Cyclization of the mixture of alkyltitanocene complexes with MgCI2 (8 equiv.) in toluene (entry 3) or in a toluene:CH2Cl2 mixture (3:1, entry 4) produced the same overall conversion and comparable yields. Since MgCI2 was virtually insoluble in the previous solvent systems, the need for a small amount of diethyl ether seemed necessary to solubilize the catalyst. To explore this avenue, entries 5 and 6 utilized varying amounts of diethyl other but, to our dismay, no significant improvements were observed. Slightly improved conversion was finally realized when a 1:1 mixture of MgCl2zMgBr2 (entry 7) was used. However, the irreproducibility of the M9X2 catalysis prompted us to pursue the other Lewis acids. We have found in previous investigations that Me2AlCI was an efficient catalyst for cyclohexane formation. During the optimization of 1648. we observed a linear conversion to 197 when Me2AlCI (1.0 equiv.) in toluene (0.01 M) was reacted with 1648 at -35 °C. Since these conditions appeared to provide results that fulfilled our initial criteria. we treated a mixture of the B-ethyl- and B-propyl- titanocene complexes under these conditions (T able 13). In entry 1, a slight 123 variation in the previous conditions were used which resulted in complete cyclization in under 30 minutes. Using the precise conditions as in 164b, only a marginal improvement in time was realized. We determined that accurate analysis of the reaction progress under more dilute conditions (<0.01 M) was not reliable. Although an extensive study of lesser equivalents of Me2AlCI was not completed, we decided to concentrate all efforts on MAO. Table 13. Cyclization of Alkyltitanocene Chlorides Promoted with Me2AlCIa drdr 8120- d: 0 time "I :1 0792 98 Y'Ieldc (min) 8' (cis/trans) (cis/trans) EI/Pr 1) Me2AlCI (1.5 equiv.). 0 83 1176 83 1176 Touene, -50 to -35 °C, 30 8 28758 7 29/55 94791 Molarity . 0.015 M RI-H; RZ-D 2) Me2AlCI (1.0 equiv.). 0 75 18” 83 1077 Toluene, ~50 to -35 °C, 20 35 21735 52 19729 917100 mlarity . 0.01 M, 55 7 28753 12 32156 887100 RI-H; 8°.H '(1)1608 8nd 1788 mm end175e ,Mg. THF. 45 °c, (2)1'le20l2, Toluene. I’lvurnbers represent millimolsr fatioeoceledb 100 fortime-Oendperoentage forthefinal time. °Referlena69. Tables 14 and 15 represent preliminary results using MAO. The information contained in these tables was separated based on the order of addition of the reagents. and the form (solid or solution) in which MAO was handled. In Table 14, MAO was weighed under an inert atmosphere prior to each reaction, diluted with toluene and cooled to -78 °C. followed by addition of a solution of the alkyltitanocene complexes. In contrast. the results in Table 15 were obtained by the addition of a 0.5 M solution of MAO (in toluene) to a -78 °C solution of the alkyltitanocene complexes. In general, the latter approach was more reliable since the precise amount of catalyst was more easily measured (in the former case, 10-15 mg of MAO were required for each reaction). 124 Table 14. Cyclization of Alkyltitanocene Chlorides Promoted with MAO l l a I I“ l I“ " fiche; Tle-fii H H GB + dFPr - dpEt GE: (I/PPr OH - 6 time 'R' "IR-2 % Yield 8‘ 9’ (min) RI (cis/trans) R2 (cisnrans) EUP' 1) MAO (1.5 equiv.). 25 .c 0 79 1279 79 1279 CH-plz. -78 to 25 °c, 360 70 1279 79 1279 917100 Molarity - 0.035 M 660 36 27718 40 29716 81785 R‘ID; R2.” 2) MAO (1.5 equiv.). 0 88 874 88 775 Toluene, :7810 25 °C, 100 11 28750 15 27745 89787 Molarity - 0.01 M, R‘IH; #IH 3) MAO (0.75 equiv.). 0 94 472 95 372 Toluene, -78 to 25 °C. 25 5 26764 7 25759 95791 Molarity I 0.02 M. R‘II‘I; FfIH 4) MAO (3.0 equiv.). 0 83 1176 83 1176 CH2CI2, -78 to -55 °C. 65 6 28754 7 29751 88787 Molarity - 0.02 M. RI-H; 82.0 °(1)188e snd175e or1888 md1788 «1718 «M758, Mg. THF,45°C, (2)110th Toluene. °Numbers moment millimolarratiosscaledb 100 fortime-Oandperoenugefortheflnal tine.°Referonoe 69. Although the experimental technique used to obtain the results in Table 14 was suspect. qualitative information about the necessary catalyst concentration and reaction temperature was available. In entry 1, no cyclization was observed at -78 °C after 6 hours with MAO (1.5 equiv.). However. allowing the reaction mixture to warm to ambient temperature overnight resulted in approximately 50% conversion. More importantly, these conditions showed a preference for the cis isomer (2:1) which paralleled the results for M9X2. If the reagents were added in the same manner as entry 1. but warmed to ambient temperature immediately (entries 2 and 3), cyclization was very rapid and the product mixture was trans selective (2:1). Entry 4 showed that with higher MAO concentrations (3 equiv.). cyclization could be promoted at -55 °C. and trans selectivity (64:36) was again 125 observed. The observed selectivity dependence on the temperature is not without precedent. Erker has recently reported that the tacticity of propylene polymerization was drastically reversed at different temperatures with MAO (isotactic at -50 °C and syndiotactic at -10 °C).75 Apparently, based on the stereochemical outcome for entries 2 and 3, cyclization must have been complete before ambient temperature was reached. The stereochemical results in entry 1 have not been duplicated. During the initial six hours of sampling the reaction, a large portion of the catalyst was probably hydrolyzed, which prevented significant cyclization from occurring prior to ambient temperature being reached. Exhaustive attempts to reproduce these results have utilized less MAO, more dilute conditions, inverse addition of reagents. and mixing of reagents at ambient temperatures. All of these attempts have been met without reward. Addition of a stock solution of MAO to the alkyltitanocene complexes provided only minor improvements for controlling the cyclization rate and the reproducibility of any set of reaction conditions (Tables 15-17). The results in Tables 15-17 demonstrated the necessary ratio of alkyltitanocene complexeszMAO to be between 1.0:0.5 to 1.0:O.65. However, comparison of the results in entries 2 and 3 (T able 15), and entries 4 and 5 (Table 15), demonstrated that under identical conditions, a marked difference in cyclization rate (or no cyclization at all, entry 3) was observed. These divergent results could be attributed to the age of the alkyltitanocene complexes or to the technique with which MAO was added. An apparent trend was that 12-24 hours after the alkyltitanocene complexes were purified. a slight increase in the MAO concentration was needed to initiate cyclization. This suggests that adventitious moisture or oxygen may play a role in retarding cyclization. Another possible scenario for the divergent results was that the requisite amount of MAO needed for cyclization never reached the solution. In order to suppress rapid cyclization, 126 MAO solution was precooled by slow addition down the side of the flask (approximately 0.25-0.50 inches above the solution) to a -78 °C solution of the alkyltitanocene complexes, and rinsed by vigorous stirring. What may have been occuring was that MAO was crystallizing on the side of the flask and failed to redissolve. Another level of frustration encountered during this study was an uncontrollable, exponential conversion to cyclic products in the initial minutes after MAO was added (see select entries in Tables 16 and 17). Moreover, during the initial exponential rise, the substrate with the B-ethyl group with respect to titanium cyclized to a far greater exent than the B-propyl substrate. Although we can merely speculate as to the cause of the difference in rate in the initial time period, we can say that these results were not isotope dependent. The net effect was to diminish the useful window of observation. 127 Table 15. Competitive Cyclization of B-l-lydrogen vs B-Hydrogen (168b vs 175b) Promoted Under the Optimized Conditions tor MAO I I a 1 M" l M“ b nopzc1 TIszCl H f H d a3 + a” time GB '47:: ELF” 07:; % Yieldc R‘ 82 (min) 9' (cis/trans) "2 (cis/trans) EUPI' 1) MAO (0.5 equiv.). 0 97 271 96 272 Toluene, ~78 to 25 °C, 20 <1 33764 9 31758 100797 Molarity II 0.02 M R‘ul-l; Rz-H 2) MAO (0.5 equiv.). 0 91 873 90 874 Toluene. ~78 to ~50 °C, 30 80 1179 84 977 100/100 Mohmy - 002 M, 150 55 1 9726 67 1371 8 10071 00 R1_H; Rag” 420 21 29750 48 20732 1007100 3) MAC (0.5 + 0.25 equiv.). 0 95 372 95 372 Toluene, ~78 to ~50 °C, 55 87 974 91 473 100798 Molarity . 0.02 M. 130 ELM; 112-H 025" 230 <1 33764 4 29759 98792 4) MAO (0.65 equiv.). 0 91 873 93 473 Toluene, ~78 to ~50 °C. 120 43 19734 59 15723 96/97 Molarity . 0,02 M, 240 5 32756 25 24744 93793 31.”; Re.” 300 3 32762 21 26750 97797 15) MAO (0.65 equiv.). 0 91 673 90 874 Toluene, ~78 to ~50 'C, 60 7 ? 48 20732 1007100 Molarity - 0.02 M. 120 <1 31789 7 28764 100799 R‘IH; R2." ‘ (1) 188. endt‘lSa . Mg, THF. 45 °C. (2)1sz02. Toluene. ° Numbers represent millimolar ratios scaled b 100 lor time-0 end percenme for the final time. °Relerence 69. 128 Table 16. Competitve Cyclization of B-Hydrogen vs B-Deuteron (168b vs 178b) Promoted Under the M”Optimized Conditlobns tor MAO a @2092“ fifiszcl g/pw Ea: :Pr 9' time “E": R. w: 1%. Yield° F" (min) 9' (cis/trans? (cis/trans) Et/Pr 1) MAO (0.65 equiv.). 0 96 371 96 371 Toluene. ~78 to -45 °C, 30 23 26751 45 20735 1007100 Molarity . 0.02 M 120 <1 33766 28 24748 1007100 R‘-H; nz-o 2) MAO (0.55 equiv.). 0 95 372 94 472 Toluene, ~78 to ~50 °C, 1 0 54 19727 70 12718 1 007100 Molarity . 0,02 M, 25 40 21739 59 15726 1007100 Ru”; 82.0 40 23 26751 48 19733 1007100 60 13 28757 23 29745 98797 3) MAO (0.50 equiv.). 0 93 473 93 473 Toluene. ~78 to ~55 °C, 10 61 15723 70 1271 8 997100 Molarity . 0.02 M. 40 25 25749 45 20/35 997100 an”, R24) 90 10 29780 21 27751 99799 4) MAO (0.4 equiv.). 0 93 473 93 473 Toluene, ~78 to 25 °C, 10 84 878 84 878 1007100 Mommy _ 0.02 M. 25 -c 105 53 25722 80 10710 1007100 an”; 92.0 145 4 38758 19 31747 100797 “(1) 1680 andt'ne . Mg. THF. 45 “C. (2"GC20l2. Toluene. E’Numbere represent millimolar ratios scaled to 100 f0 time-0 and percentage for the final time. °Fleierenoe 69. 129 Table 17. Competitive Cyclization of B—Deuteron vs B-i-iydrogen (171b vs 175b) and B,B~Dideuteron vs ma-Dlhydrogen (184b vs 175b) Promoted Under the Mt,Optlmlzed Conditions for MAO time ":1 % Yieldc (min) 8' E(cis7trans) Pr(ms/trans.) Et/Pr 1) MAC (0.50 + 0.05 equiv.). 0 97 271 97 271 Toluene. ~78 to ~55 “C. 20 81 8711 85 679 1007100 Molarity - 0.02 M 95 73 10717 80 7713 100/100 R‘.0; rte-11:18.11, H 120 64 13722 75 8716 99798 0.05 OWN: 140 47 20738 49 1 7732 99798 1 80 4 30/65 9 30761 99/100 2) MAO (0.55 equiv. ), 0 92 573 90 674 Toluene. ~78 to ~55 °C, 10 45 22733 60 16724 1007100 mlarity .. 0,02 M, 25 30 26743 47 20732 99799 R1-H; R2...“ 59.0. D 55 21 28749 35 23740 98/98 70 13 29755 20 27750 97797 3) MAO (0.55 equiv.). 0 d 7/6 92 573 Toluene. ~78 to ~55 °C. 10 d 16721 62 13720 d795 Molarity _ 0.02 M, 25 (I 23731 60 14722 (”96 an“; R’.H;a’~0. 0 55 0 33744 37 21738 d796 '(1)17ta and1750 «184a and175a . Mg. THF. 45 °C, (2TiCp20l2.Toiuene.°Numbersrepreeentmiliimolar retioesceledthOlor fine-Omdperoentageiorthefinaltime.‘neference69.‘CoUdnotbe dehmimdaoanblymebpoubanimmsohfimhmmofmeoudwmawicm. 130 Figures 4-7 illustrate a graphic view of the relative rate data for various competitive experiments. If the initial growth period is ignored and only the linear portion considered for each graph, the 01- or B- deuterium atoms had no measurable effect on the rate. However, with the minimum number of data points obtained in each linear portion of the graph, the error in these slopes was considerable. in spite of these shortcomings, an amazing correlation was observed. For the competitive cyclization of 168b and 175b (Figure 4), a relative rate of k1sablk175b=122 was determined, which demonstrated the contribution of the alkyl group to the rate of insertion. Figure 5 suggests that B-hydrogen activation does not enter into the mechanism under the conditions that employ MAO, as the relative rate was comparable to Figure 4, k153uk17ab=1.24. The best graph of the relative rate data for the control experiment, Figure 6, was not conclusive, as the erroneous graph indicates. Finally, the test for a-hydrogen activation is illustrated in Figure 7 and the relative rate calculated to be k13457k175b=1.30. These results indicate that a-hydrogen activation is likewise, not important to the mechanism for this system. 131 lOO —'_ 168b —._ 1750 % Conversion I ' I f 200 300 400 Time (min) V ' r O 100 Figure 4. Competitive Cyclization Between 168b and 175b (see Table 15, entry 4). Slopes were calculated using a staple line program and omitting the exponential portions of the graph. The relative rate was calculated to be: krel=k168b7k175b=122~ 100 168!) 178!) ll % Conversion 80 fime (min) Figure 5. Competitive Cyclization Between 168b and 178b (see Table 16, entry 2). Slopes were calculated using a sinple line program and omitting the exponential portions of the graph. The relative rate was calculated to be: krel-4‘168b7kt78b-1~24- 132 lOO 60 -1 C .0 g 40 7 —9— 171b .\° 20 - o - , - 0 '00 0.05 mmol MAO 20° added at time-125 Time (min) Figure 6. Competitive Cyclization Between 171b and 175b (see Table 17, entry 1). Slopes were calculated using a simple line program and omitting the exponential portions of the graph. Relative rate calculated to be: krel=k171b7k175b-1-60- lOC ’C 7% -—9— 184b 32 . . r r . . . 0 20 4o 60 80 Time (min) Figure 7. Competitive Cyclization Between 184b and 175b (see Table 17, entry 2). Slopes were calculated using a slrmle line program and omitting the exponential portions of the graph. The relative rate was calculated to be: kreI-k184b7k175b-1-30. 133 Athough the above results suggest that a- and B-hydrogen activation do not participate in the mechanism, the limited data points for each graph do not provide confidence for any firm conclusion. in order to more accurately assess the situation, we have recently prepared 6-butyl-7-br0mo-1~heptene (182) and plan to compare it to 175. Preliminary results demonstrated that 182b underwent cyclization more slowly than the B-propyl substrate (175b) in the initial time after MAO was added. A trend has been clearly established in which the longer B-alkyl chains gave rise to decreased cyclization in the early stages of the reaction. This new substrate should allow for the rapid completion of this study. 4. Conclusion Although these results were not conclusive as to the existence of 01- or [3- hydrogen activation, conditions have been worked out using MAO that should allow this project to be completed soon. Although, preliminary results discussed here indicate that olefin insertion promoted by MAO did not show an a- or (3- hydrogen effect. A number of interesting stereochemical observations were made based on the Lewis acid and the temperature of the reaction. it was shown that both MAO and Mng were selective for the cis isomer when cyclization was carried out at 25 °C or higher. Yet, when lower temperatures were employed with MAO. the trans isomer was preferentially obtained. The future of this mechanistic project will be directed towards the investigation of the magnesium catalysts in an attempt to understand the nature of stereochemistry obtained. For now, the secrets of Ziegler-Natta polymerization are still secure. 1 34 5. Experimental Methaluminoxane (MAO). A Schlenk flask was loaded with Al2(SO4)3(H20)(14-13) followed by the removal of oxygen under vacuum and purging with argon. The aluminum sulfate was suspended in toluene (80 mL) and the flask immersed in a 25 °C water bath. MeaAl (20 mL, 209 mmol) was diluted with toluene (20 mL) and the solution was added rapidly, via cannula, under argon pressure to the aluminum sulfate suspension. after which the reaction mixture was heated to 40 °C for 15—20 hours. The MAO was purified by filtration under argon using a Schenk filter. To accomplish this purification, the slurry was transferred to the filter via cannula. The solvent was removed in vacuo and the MAO was stored under nitrogen. For the reactions in Table 14, MAO (MW=130 g/mol) was weighed under nitrogen prior to each reaction, and diluted with the appropriate solvent. For the all other reactions employing MAO. a 0.5 M MAO solution (toluene) was used (3.25 g MAO. 25 mmol, 50 mL toluene). Magnesium leromlde. Magnesium turnings (1.46 g, 60 mmol) were loaded in a 3~necked flask equipped with a reflux condenser and a dropping funnel, and diethyl ether (30 mL) was added. An diethyl other solution of 1,2- dibromoethane (11.27 g. 60 mmol, 10 mL diethyl ether) was added slowly to the magnesium metal over 2 hours followed by heating to reflux for an additional 2 hours. Diethyl ether was removed under reduced pressure and the white precipitate was heated 150 °C under vacuum (<0.001 mm Hg) for 8-10 hours. The white solid was stored under nitrogen. General Procedure for Metal Mediated Cyclizations Using Unpurlfled Alkyltitanocene Chlorides. A Schlenk flask was charged with magnesium (583 mg, 24 mmol), flame dried under vacuum (<0.001 mm Hg), purged with argon 135 prior to the addition of diethyl ether (4 mL). Alkyl bromide 175 (438 mg, 2 mmol) was slowly added to the magnesium over 6 hours using a gas tight syringe. After the addition was complete, the reaction mixture was allowed to stir for an additional 4 hours. In a separate flask was loaded Tiszclz (548 mg. 2.2 mmol), and the flask was subsequently evacuated under reduced pressure. and purged with argon. The TICp2Cl2 was suspended in toluene (8 mL) and the flask cooled to ~50 °C. The alkylmagnesium bromide was transferred to the Tiszclz suspension via cannula and rinsed with additional diethyl ether (1 mL). The reaction temperature was maintained at ~50 °C for 30 minutes and then warmed to ambient temperature for an additional 5 hours. (A) Magnesium Dihallde Promoted Cyclization Uslng In slfu M9X2. After transmetallation was complete, the reaction mixture was brought to 45 °C (Table 11; entry 1). Cyclization was promoted using the Mng (X: Cl or Br) which was formed in situ by the reaction of the Grignard reagent and Tiszclz. (B) Magnesium Dihallde Promoted Cyclization Uslng Excess ngz or MAC. After transmetallation was complete, the diethyl ether was removed under vacuum and the reaction mixture was diluted with CH20I2 (5 mL). In a separate flask was loaded either MgCI2 (189 mg, 1.98 mmol, Table 11: entries 2 and 3) or MAO (86 mg, 0.66 mmol, Table 11; entry 4) under and atmosphere of nitrogen and diluted with toluene (8 mL). A solution of 175b (2.0 mL, 0.33 mmol) was added to a ~45 °C solution (or suspension) of the respective Lewis acids. via syringe, and that temperature maintained for 10 minutes followed by warming to the desired temperature (see Table 11). General Procedure for Metal Promoted Cyclizations Uslng Purified (M9X2 Free Solutions) Alkyltitanocene Chlorides. A Schlenk flask was charged with magnesium (780 mg, 32 mmol), flame dried under vacuum, and purged with argon. THF (4 mL) was added to the flask and heated to 45 °C. in a 136 typical competitive rate experiment. a mixture of alkyl bromides (1 mmol each) was simultaneously added to the magnesium over 6 hours using a gas tight syringe and after the addition was complete, the reaction temperature was maintained for an additional 4 hours. In a separate Schlenk flask was loaded Tiszclz (548 mg, 2.2 mmol) and the flask was evacuated and purged with argon. The TiCp2CI2 was suspended in toluene (8 mL) and cooled to ~78 °C. The Grignard reagents were transferred via cannula under argon pressure to the cold suspension of Tiszclz and rinsed with additional THF (2 mL). The transmetallation reaction was maintained at ~78 °C for 8 hours and then allowed to warm to ambient temperature for an additional 90 minutes, at which time the reaction mixture was diluted with hexane (25 mL) and cooled to 0 °C. The combined action of the hexane and cold temperatures caused most of the M9X2 and excess TlezClz salts to fall out of solution. The cold reaction mixture was transferred via cannula under argon pressure to a Schlenk filter and sequentially washed with cold hexane (3 x 7 mL). The alkyltitanocene chlorides were evaporated to dryness under vacuum (<0.001 mm Hg) to leave behind a dark red oil (or slurry). Toluene (30 mL) and an internal standard were added to the alkyltitanocene chlorides and the concentration of each substrate determined by gas chromatographic techniques. All subsequent reactions utilized this stock solution of known molarity. (A) Competitive Bate Cyclizations Promoted by ngz or Crystalline MAO. MgCI2, MgBrz (or both. see Table 12; entry 3) or MAO were loaded in a Schlenk flask under nitrogen, suspended or dissolved in the solvent employed, and cooled to low temperature (see Table 12 and 13 for more details on each reaction). A calculated volume of the alkyltitanocene chlorides was added over one minute to the Lewis acid, maintained at that temperature for 10 minutes, and finally warmed to the appropriate temperature. 137 (B) Competitive Rate Cyclizations Promoted by Me2AlCI or MAC. Solutions of Me2AlCI and MAO were prepared by dissolving the Lewis acids in an appropriate volume of toluene to prepare a 0.5 M solution of each. A calculated volume of the alkyltitanocene chloride solution was diluted to the appropriate concentration (see Table 14, 15, 16, or 17) and cooled to either ~50 °C if MegAlCl was to be used, or ~78 °C if MAO was to be used. The appropriate Lewis acid was taken up in a gas tight syringe and slowly added down the side of the flask about 0.5 inches above the solution, and the flask was vigourously swirled to rinse the sides of the flask. The initial temperature employed was maintained for 10 minutes and then warmed to the described temperature (see Table 14. 15, 16, or 17 for more details). CONCLUSION The development of a novel ring-forming method which featured an intramolecular monomeric analog of Ziegler-Nana polymerization was discussed. The ease with which unactivated olefins reacted with the carbon-titanium bond was demonstrated. The metal and its surrounding ligands played a significant role during the carbocyclization, controlling both the regio~ and the stereochemistry of bond formation for both 5~hexen~1~yl~ and 6~hepten~1~ yltitanocene chlorides. Exo insertion was the sole pathway observed in both cases. When an alkyl group was located on the alkyl tether. excellent and predictable diastereomeric excesses were achieved during carbon-carbon bond formation. in most cases. This ring-forming method represents the first procedure that did not result in a loss of selectivity or efficiency when extended to six~ membered ring synthesis. This general and versatile procedure will certainly enhance the repertoire of carbon-carbon bond forming reactions available to the synthetic chemist. 138 LIST OF REFERENCES LIST OF REFERENCES (a) Rigollier, P.; Young, J. R.; Fowley, L. A.; Stille, J. R. J. Am. Chem. Soc. 1990, 112, 9441. (D) Young. J. R.; Stille. J.R. Organometallics 1990. 9, 3022. (c) Young, J. FL; Stille, J. R. J. Am. Chem. Soc. 1992, 114, 4936. For an excellent summary of modern organometallic chemistry, see: Collman, J. P.; Hegedus, L. 5.; Norton, J. Ft; Finke. Ft. 6.; ln Principles and Applications of Organotransition Metal Chemistry: University Science Books: California, 1987. For reviews of Ziegler-Natta polymerization, see: (a) Boor, J., Jr. In Ziegler- Natta Catalysts and Polymerization; Academic Press: New York, 1979. (b) Pino, P.; Mulhaupt, R. Angew. Chem, Int. Ed. Eng/.1980, 19, 857. (c) Kaminsky. W.; Sinn. H. Transition Metals and Organometallics as Catalysts for Olefin Polymerization; Springer-Verlag: Berlin, Heidelberg, 1988, 361. Ziegler, K. Angew. Chem. 1952, 64, 323. Natta, 6.; Pine, P.; Corradin, P.; Danusso, F.; Mantica, E.; Mazzanti, G.; Moraglio. G. J. Am. Chem. Soc. 1955, 77, 1708. Waters. J. A.; Mortimer, G. A. J. Organomet. Chem. 1970, 22, 417. (a) Breslow, D. 5.; Newburg, N. R. J. Am. Chem. Soc.1959. 81, 81. (b) Ziegler, K.; Gellert, H. G.; Zosel, K.; Holzkamp, E.; Schneider, J.; Soil, M.; Kroll, W. R. Ann. Chem. 1960, 121, 629. (a) Marvel. C. S.; Stille, J. K. J. Am. Chem. Soc. 1958. 80. 1740. (b) Waymouth, B. M.; Resconi, L. J. Am. Chem. Soc. 1990, 112, 4953. 139 10. 11. 12. 13. 14. 15. 140 For cyclization of 5~hexen~1~yl metal systems using titanium (a) and zirconium (0) catalysts, see: (a) Clawson, L.; Soto, J.: Buchwald. S. L.; Steigerwald, M. L.; Grubbs, R. H. J. Am. Chem. Soc.1985, 107, 3377. (b) Carr, D. B.; Schwartz, J. J. Am. Chem. Soc. 1979, 101, 3521. For related examples using scandium (c) and yttrium (d). see: (c) Piers, W. E.; Bercaw, J. E. J. Am. Chem. Soc. 1990, 112, 9406. (d) Molander, G. A.; Hoberg, J. O. J. Am. Chem. Soc. 1992, 114, 3123. For an excellent review in the area of cationic cyclization see: Johnson, W. S. Biorg. Chem. 1976, 5, 51. (a) Burke, S. D.; Takeuchi, K.; Murtiashaw, C. W.; Liang, D. W. M. Tetrahedron Lett.1989, 30. 6299. (b) Burke. S. D.; Strickland, S. M. 5.; Organ, H. M.; Silks, L. A. III Tetrahedron Lett. 1989, 30. 6303. (c) Burke, 8. D.; Deaton, D. N. Tetrahedron Lett.1991, 32, 4651. (d) Burke. 8. D.; Shankaran. K.; Helber, M. J. Tetrahedron Left. 1989. 30, 4655. For a review on the use of chiral acetals in organic synthesis, see: Alexakis, A.; Mangeney, P. Tetrahedron 1990, 46. 477. For a comprehensive perspective on the development of free-radical processes, see: Giese, B. in Radicals in Organic Synthesis: Formation of Carbon-Carbon Bonds, Volume 5; Oxford: 1986. (a) Schwartz. 0. E.; Curran, D. P. J. Am. Chem. Soc. 1990. 112, 9272. For other examples of tandem cyclizations leading to linear and propellane triquinane natural products, respectively, see: (b) Curran, D. P.; Fiakiewicz, D. M. J. Am. Chem. Soc. 1985. 107, 1448. (c) Curran, D. P.; Jasperse. C. P. J. Am. Chem. Soc. 1990, 112, 5601. (a) Fraser-Reid, B.; Pak, H.; Dickson, J. K., Jr. J. Org. Chem. 1989, 54, 5357. For other examples of tandem cyclizations using carbohydrate templates, see: (b) Fraser-Field, B.; Dickson, J. K., Jr.; Tsang, H.; Llera, M. J. Org. Chem. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 141 1989, 54, 5350. (c) Fraser-Reid, B.; Pak, H.; Canalda. I. l. J. Org. Chem. 1990. 55. 3009. For examples of free radical cyclization of 7-bromo-1-heptene compounds, see the experimental in Chapter 3. Beckwith. A. L. J.; Schiesser, C. H. Tetrahedron 1985. 41, 3925. Chatgilialoglu, C.; lngold, K. U.; Scaiano. J. C. J. Am. Chem. 8001981. 103, 7739. Beckwith. A. L. J.; Easton, C. J.; Lawrence, T.; Serelis, A. K. Aust. J. Chem. 1 983. 36, 545. For comparison of these literature results. see the experimental in Chapter 2. For a theoretical discussion on the regio- and stereochemical details for 5- exo-trig radical ring closures. see reference 17 and: Spellmeyer, D. 0.; Houk, K. N. J. Org. Chem. 1987, 52, 959. Kharasch, M. S.; Skell, P. 8.; Fisher, P. J. J. Am. Chem. 800.1948, 70. 1055. Curran, D. P.; Chang, C. J. Org. Chem. 1989, 54, 3140. Curran, D. P.; Chen, M. J. Am. Chem. Soc. 1987, 109. 6558. Curran, D. P.; Seong, C. M.. J. Am. Chem. Soc. 1990, 112, 9401. Boger, D. L.; Mathvink, R. J. J. Am. Chem. Soc.1990, 112, 4003. (a) Porter, N. A.; Chang, V. H. T.; Magnin, D. H.; Wright, B.T. J. Am. Chem. Soc. 1988, 110, 3554, and references therein. (b) Boger, D. L.; Mathvink, R. J. J. Am. Chem. Soc. 1990, 112, 4008, see reference 24 also. (a) Curran, D. P.; Shen, W.; Zhang, J.; Heffner, T.A. J. Am. Chem. 800.1990, 112, 6738. (b) Porter, N. A.; Swann, E.; Nally, J.; McPhail, A. T. J. Am. Chem. Soc. 1990, 112, 6741. (c) Giese, B.; Zehnder. M.; Roth, M.; Zeitz, H. J. Am. Chem. Soc. 1990, 112, 6741. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 142 Bailey. W. F.; Punzalan, E. R. J. Org. Chem. 1990. 55. 5404, and references therein. (a) Ashby, E. 0.; Pham, T. N. J. Org. Chem. 1987, 52, 1291. (b) Ashby. E. C.; Pham, T. N.; Park. B. Tetrahedron Lett.1985. 26, 4691, and references therein. (a) Bailey. W. F.; Normi, T. T.; Patricia, J. J.; Wang, W. J. Am. Chem. Soc. 1987, 109. 2442. (b) Broka. C. A.; Lee, W. J.; Shen, T. J. J. Org. Chem. 1988, 53, 1336. (c) Chamberlin, A. J.; Bloom, S. H.; Cervin, L. A.; Fotsch, C. H. J. Am. Chem. Soc. 1988, 110, 4788. (d) Koppang, M. 0.; Ross. G. A.; Woolsey. N. F.; Bartak, D. E. J. Am. Chem. Soc.1986, 108, 1441. (e) Paquette, L. A.; Gilday, J. P.; Maynard, G. D. J. Org. Chem. 1989. 54, 5044. Bailey, W. F.; Khanolkar, A. D.; Gavaskar, K.; Ovaska, T. V.; Rossi, K.; Thiel, Y.; Wiberg. K. B. J. Am. Chem. Soc. 1991, 113, 5720. Bailey, W. F.; Khanolkar, A. D. J. Org. Chem. 1990. 55, 6058. Oliver, J. P.; Smart. J. B.; Emerson, M. T. J. Am. Chem. Soc. 1966, 88, 4101. Bailey, W. F.; Khanolkar, A. D. Tetrahedron Left. 1990, 31, 5993. Beckwith. A. L. J.; Phillipou, G. Serelis, A. K. Tetrahedron Lett. 1981, 22, 2811. Bailey, W. F.; Rossi, K. J. Am. Chem. 800.1989, 111, 765. Ashby. E. C.; Oswald, J. J. Org. Chem. 1988. 53. 6068. Kossa, W. 0.. Jr.; Richey, H. 6., Jr.; Rees, T. C. Tetrahedron Left. 1971, 3455. Utimoto. K.; Imi, K.; Shiragami, H.; Fujikura. S.; Nozaki, H. Tetrahedron Lett. 1985, 26, 2101. Stefani, A. Helv. Chim. Acta. 1 974, 57, 1346. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 143 (a) Carpenter, N. E.; Kucera, D. J.; Overman, L. E. J. Org. Chem. 1989, 54, 5846. (b) Ableman, M. M.; Overman, L. E. J. Am. Chem. Soc. 1988, 110, 2328. (a) Roll, W.; Brintzinger, H.; Rieger, B.; Zolk, R. Angew. Chem, Int. Ed. Engl. 1990, 29. 279. (b) Mullin, D. T.; Rausch. M. 0.; Lin, Y.; Dong, S.; Chien, J. C. W. J. Am. Chem. Soc. 1990, 112, 2030. (c) Soga, K.; Park, J. R.; Uchino, H.; Uozumi, T.; Shiono, T. Macromolecules 1989, 22, 3824. (d) Ewen, J. A. J. Am. Chem. Soc. 1984, 106, 6355. (a) Ewen, J. A.; Jones, R. L.; Razqui, A. J. Am. Chem. Soc. 1988, 110. 6255. (b) lshihara, N.; Kuramoto, M.; Uoi, M. Macromolecules 1988. 21, 3356. Cossee. P. Tetrahedron Lett. 1960. 17, 12. lvin, K. J.; Rooney, J. J.; Stewart. C. D.; Green, M. L. H.; Mahtab, R. J. Chem. Soc., Chem. Commun. 1978. 604. Chatt. J.; Duncanson, L. A. J. Chem. Soc. 1953, 2939. For a molecular orbital description of bis(cyclopentadienyl)~MLn complexes and their chemistry. see: Hoffman, R.; Lauher, J. W. J. Am. Chem. Soc. 1976, 98. 1729. For a description of the propagation and termination steps for the polymerization of ethylene with a scandium complex, see: Burger, B. J.; Thompson, M. E.; Cotter, W. D.; Bercaw, J. E. J. Am. Chem. Soc.1990, 112, 1556. Dawoodi, 2.; Green, M. L. H.; Mtetwa, V. S. B.; Prout, K. J. Chem. Soc, Chem. Commun. 1982, 1410. For a theoretical study of this structure, see: Koga. N.; Obara, S.; Morokuma, K. J. Am. Chem. Soc. 1984, 106. 4625. (a) Jordan, R. F.; Bradley, P. K.; Baenziger, N. C.; LaPointe, R. E. J. Am. Chem. 800.1990, 112, 1289. (b) Dawoodi, 2.; Green, M. L. H.; Mtetwa, V. S. B.; Prout, K. J. Chem. 800., Chem. Commun. 1982, 802. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 144 Schmidt, G. F. ; Brookhart, M. J. Am. Chem. Soc. 1985, 107, 1443. Bose. A. K.; Lal. B. Tetrahedron Lett. 1973. 3937. Hartman, G. D.; Halczenko. W.; Phillips, B. J. Org. Chem 1986, 51, 147. The following reactions were mutually unsuccessful: allylic coupling of 131, an alkyl halide, and Ni(COD)2, and Grignard or cuprate addition of 131 to ethylacrylate. For the use of Ni(COD)2 in organic synthesis, see: (a) Hegedus, L. S.; Varaprath, S. Organometallics 1982, 259. For the addition of Grignards or cuprates to acrylate derivatives, see: (b) Liu, S. H. J. Org. Chem. 1977, 42, 3209. Finally, it was found that allyl metal reagents of Mg or Cu could be successfully reacted with 131, see (c) Corey, E. J.; Posner, G. H. J. Am. Chem. Soc. 1967, 89. 3911. Org. Synthesis, Coll. Vol. 4, 1 963. 749. Brown, H. C.; Mandal, A. K.; Kulkarni J. Org. Chem. 1977. 42, 1392. Brown. H. C.; Prasad, J. V. N. V. J. Org. Chem. 1985. 50. 3002. Chaudhary, S. K.; Hernandez, O. Tetrahedron Left. 1979, 99. Howden, M. E. H.; Maercker, A.; Burdon, J.; Roberts, J. D. J. Am. Chem. Soc. 1986. 88. 1732. Gadwood, R. C.; Lett. R. M.; Wissinger, J. E. J. Am. Chem. Soc. 1986, 108. 6343. Brown, H. C.; Garg, C. F.; Liu. K. J. Org. Chem. 1971, 36, 387. Whitesell, J. K.; Wang. P. K. S.; Aguilar, D. A. J. Org. Chem. 1983, 48. 2511. Boland, W.; Ney, P.; Jaenicke, L. Synthesis 1980, 1015. The linear sequence from and including 179-182 was kindly. proficiently, and expediently performed by Nancy S. Barta. The proper gratitude cannot be adequately expressed for her efforts when l was in need. lnorg. Syn. 1963, 7, 9. 66. 67. 68. 69. 70. 71. 72. 73. 145 Danishefsky, S.; Vaughan, K.; Gadwood, R.; Tsuzuki, K. J. Am. Chem. Soc. 1981, 103, 4136. Greenwald, R.; Chaykovsky, M.; Corey, E. J. J. Org. Chem. 1963, 28, 1128. Sauvage, J. F.; Baker, R. H.; Hussey. A. S. J. Am. Chem. Soc. 1960, 82. 6090. All yields reported here for free radical or titanium-mediated cyclizations were obtained by gas chromatographic techniques compared to an internal standard and corrected for detector response. (a) Beesley, R. M.; lngold, C. K.; Thorpe. J. F. J. Chem. 800.1915, 107, 1080. (b) lngold, C. K. J. Chem. Soc. 1921, 113, 305. (c) Hammond, G. S. In Steric Effects in Organic Chemistry; Newman, M. 8., Ed.; John Wiley and Sons: New York. 1956; p. 463. Rieke, R. D.; Li, P. T. J.; Burns, T. P.; Uhm, S.T. J. Org. Chem. 1981. 46. 4323. Upon close examination of the mass spectrum fragmentation pattern of the compound reported in reference 1a, we determined that it was not 133, but rather a structural isomer. The actual compound in reference 1a was 4~(3~ bromopropyl)~1~cyclohexene. (a) Pino, P. Adv. Polym. Sci. 1965, 4, 393. (b) Pino, P.; Ciardelli. F.; Zandomeneghi, M. Ann. Rev. Phys. Chem. 1970, 21. 561. (c) lfissin, Y. V. In lsospecitic Polymerization of Olefins with Heterogeneous Ziegler-Natta Catalysts; Springer-Verlag: New York, 1985; p. 295-306. (d) Ciardelli. F.; Carlini, C.; Altomore, A.; Menconi, F.; Chien, J. C. W. In Transition Metal Catalyzed Polymerizations; Quirk, R. P., Ed.; Cambridge. University Press: Cambridge, 1988; p. 25. (e) Vizzini, J.; Ciardelli. F.; Chien, J. C. W. Macromolecules 1 992, 25, 108. 146 74. The term a-hydrogen activation is somewhat misleading and controversial. Grubbs used this term when referring to the Green-Rooney mechanism (see reference 9a). Bercaw chose the term ”or-agostic assistance” rather than "a- hydrogen activation" (see reference 9c). The use of ”hydrogen-activation” is often used to denote breaking of the carbon-hydrogen bond. When hydrogen-activation was used in this thesis. only agostic stabilization and acceleration of olefin insertion was implied. 75. Erker, G.; Fritze, C. Angew. Chem, Int. Ed. Engl. 1992, 31, 199.