.IA 1.3.).- o- a I vulhzfi. | I O a ‘ w . l .5 v I. . y . 102. [Not . O 4 O I ll”. LA .Ah I A. . I In” vl‘ U ' J\ I I IV \ 7 . fill. )I.‘ Ill " A 0‘ Ilth II I .9 ‘ yet-ll“! “out... 5. WWII-(PH ‘ {flu-hung il- .A . Q . .II n \ . m. . v. o v "Hollovt- Lug“ . {‘11:} l ’{Xfl.l. . A: r n. u .c l 10 1 I ‘ . v..." v. . Ir’.‘ . vi 1, H . a I i I I I . III-‘OAJI‘IIIO "Tt‘J-Lrlg‘n In 41‘ I a JIIJI-II o oblHkHI'IIO .\_|’ . . o . VLNHHMHV «I: ’u». . . :‘ltllu \ - - . 9111!”! Jun»: If“: t '17}! E ‘ K -- - 1:11.}... riggrcf!f! .... 5.3... -. .. - 1!} HENRY Mews-'31: Etate University WV ‘7 ‘ v v This is to certify that the dissertation entitled "An Approach to the Synthesis of Dodecahedrane" presented by Theresa Ann Monego has been accepted towards fulfillment of the requirements for Ph . D. degree in Chm. W Major professor Date September 10L 1982 MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 MSU LIBRARIES m. ~ RETURNING MATERIALS: Place in book drop to remove this checkout from ‘ your record. FINES will be charged if book is returned after the date stamped below. AN APPROACH TO THE SYNTHESIS OF DODECAHEDRANE BY Theresa Ann Monego A DISSERTATION Submitted to Michigan State University in partial filfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1982 ABSTRACT AN APPROACH TO THE SYNTHESIS OF DODECAHEDRANE BY Theresa Ann Monego A novel synthetic route to dodecahedrane l involving a 12 n electron reorganization in hexaene lg is proposed. A key intermediate in the synthesis of hexaene lg, gig, eggng,6-2i§(hydroxymethyl)bicyclo[3.3.0]octa-3,7-diene 4Q, has been realized. The synthesis of diol éé proceeds in .3 steps in 34% overall yield from the known gi§,ggdgf2,6- dihydroxybicyclo[3.3.0]octa-3,7-diene éé' This sequence employs a highly stereoselective chelation-controlled SNZ' ring opening of tricyclic ether Zé with "methanol dianion". Related studies in this area demonstrate that the [1,2] Wittig rearrangement rather than anionic [2,3] sigmatropic rearrangement occurs in functionalized bicyclo[3.3.0]octen- ols. Finally, model studies on the saturated analog of diol éé have served to elucidate further transformations required for the synthesis of $3. To my Mother and to the Memory of My Father and to My Husband, Peter, for Their Love and Understanding ii ACKNOWLEDGMENTS I would like to thank Professor Donald G. Farnum for his support and guidance during the pursuit of my degree. His critical insight and chemical knowledge have greatly enhanced my understanding of organic chemistry. I grate- fully acknowledge the advice of Professor Eugene LeGoff, who served as second reader, and I also thank the remainder of my committee for their interest and suggestions concern- ing this work. My stay here at Michigan State University has been rendered immeasureably more pleasant by the com- panionship of my colleagues, especially those in both past and present Farnum and LeGoff groups. I cannot express my 'gratitude nearly enough to my husband, Peter, who has made it all worthwhile. iii TABLE OF CONTENTS Chapter LIST OF TABLES. . . . . . . . . . . . . LIST OF FIGURES . . . . . . . . . . . . INTRODUCTION. . . . . . . . . . . . . . RESULTS AND DISCUSSION. . . . . . . . . 1. C10 tion Approach . . . . . . . . . 2. Sulfide Contraction/Elimination Approach. . . . . . . . . . . . 2.1. Model Studies . . . . . 2.2. The Fate of One Bis-Wittig Re- arrangement . . . . . . . 2.3. Another Wittig Rearrangement. Lactone (or Diester) Dimeriza- 2.4. The Synthesis of Cis, endo-2,6- bis(hydroxymethyl)bicyc15f3.3.0] octa-3,7-diene 3Q . . . . . . . EXPERIMENTAL O C O O C O O O O O O O O 0 General . . . . . . . . . . . . . . Bi(2-oxa-3-oxotricycloE7.2.1.0 undeca-6,9-dien-4-y1) . Oxidative Dimerization of Lactone 2g Lithium Enolate . . . . . . . . . . . . . . 2,6-Bis-methylenebicycloE3.3.0]- octane g2 . . . . . . . . . . . . . gi§,endo-2,6-bis(hydroxymethyl)- bicyEloE3.3.oTBEtane g . Hydro- boration of Diene g2 w1th Disiamyl- borane. . . . . . . . . . . . . . . iv 5'11]- Page . viii ix 16 16 20 30 43 58 80 88 88 89 91 93 Chapter Hydroboration of Diene 6% with BB 3 - THF COmp 1 ex 0 o o o o o o o o o o o o Cis,endo-2,6-bis(hydroxymethyl)- bicycloi3.3.0]octane Dimethane- sulfonate 6%. . . . . . . . . . . . . . . Cis,endo-2,6-bis(iodomethyl)bi- cycloi3.3.0]octane, éé. . . . . . . . . . Cis,endo-Z,6-bi§(bromomethyl)bi- cyclol3.3.0]octane 69. Reaction of Diol with Carbon Tetrabromide/ Tripheny phosphine. . . . . . . . . . . . Cis,endo—2,6-bis(amidinothio)methyl- bicyclol3.3.0loctane Dibromide The Addition of Thiourea to Dibromide g3. . . . . . . . . . . . . . . . . . . . Cis,endo-2,6-bis(mercaptomethyl)bi- cyclo[3.3.0]octane 88' Hydrolysis of Bis-thiouronium Bromlde Q2. . . . . . . . The Reaction of Dithiol 66 with Di- bromide 62. . . . . . . . . . . . . . . . Cerous Chloride Mediated Sodium Boro- hydride Reduction of Bicyclo[3.3.0]octa- 3,7-diene-2,6-dione zg. Cis,endo-2,6- dihydroxybicyclo[3.3. Jocta-3,7-diene éZ' Temperature Effects in the Sodium Borohydride/Cerric Chloride Reduction of Dienedione 1g. . . . . . . . . . . . . Iodomethyltri-n-butyltin. . . . . . . . . Cis,endo-2,6-bis[(tri-n-butylstannyl)- methoxyIbicyclol3.3.0]octa-3,7-diene SQ . 2-0xa-tricyclo[5.2.l.o4'lojnona-5,8- diene i6. Reaction of 38 with Two Equiva ents of neButylllthium . . . . . . Low Temperature Quenching of Dianion i Derived from gig-stannane and n- Butyllithium. Cis,endo-2,6- 1methoxy- bicyclo[3.3.0]octa-3,7-diene 2g . . . . Page 94 96 96 97 99 100 101 102 103 104 104 105 107 Chapter Page 2-Oxa-tricyclc[5.2.l.04’lojnona- 5,8-diene, Z . Reaction of 8 with One Equlvalent of n-ButyI- lithium . . . . . . . . . . . . . . . . . . . . 108 Cis,endo-Z- hydroxy-6- [(tri- n-butyl- stannyl)methoxy]bicyc10[3. 3. 0]octa- 3, 7- diene 87. . . . . . . . . . . . . . . . 109 n-Butyllithium Exchange with Cis, endo-2- -hydroxy-6- (tri- n-butylstanny1)- methoxybicyclo[3. 3. 0]octa—3, 7- diene Z. Cis,endo-Z- hydroxy- 6- -hydroxymethy1bi- cyclol3. 3. 0]octa-3, 7- diene g, and 2- Endo-hydroxymethyl- 6- -buty1b1cyclo[3. 3. 0]- octa-3, 7- diene 82. . . . . . . . . . . . . . 110 n-Butyllithium Cleavage of Tricyclic Ether é. Endc-Z- (hydroxymethyl)- -6- -n- butylblcycloIB. 3. 0]octa-3, 7- diene 82. . . . . 112 Sodium Borodeuteride/Cerous Chloride Reduction of Bicyclo[3. 3. 0]octa-3, 7- diene- 2, 6- dione . Cis, endo- 2, 6- -dihydroxy- bicycloEB. .O]octa-3, 7- diene-2 d2 12. . . 113 Alkylation of[2, 6- dJDiol 17 with One Equi- valent of Iodomethy tri-n-butyltin. . . . . . . 114 2-Oxa-tricyclo[5.2.1.04’10]deca-5,8- diene-1,6-Q2. . . . . . . . . . . . . . . . . . 115 Cis, endo-Z- -hydroxy-6-(hydroxy- methyl)bicyclo[L 3. 0]octa-3, 7- diene- g§ and endo-2- (hydroxymethyl)- 6- :n-baty bicycIo I3. 3. O]octa-3, 7- diene- 812181 1.............15 Cis, endo- 2, 6- -bis(hydroxymethy1)- bicycloIL 3. OISEta-B, 7-diene 6. Nucleophilic Ring Opening of ri- cyclic Ether 16 with Methanol Di- anion . . . . . . . . . . . . . . . . . . . . . 116 Cis, endo- 2, 6- -bis(acetoxymethy1)- bicyclol3. 3. 0|octa-3, 7- diene 112. . . . . . 119 SUMMARY . . . . . . . . . . . . . . . . . . . . . . 121 vi Chapter Page APPENDIX. . . . . . . . . . . . . . . . . . . . . . 124 REFERENCES. . . . . . . . . . . . . . . . . . . . . 153 vii LIST OF TABLES Table Page 1 Predicted Energies for Various Dodecahedrane Inclusion Compounds . . . . 4 2 1H NMR Chemical Shifts and Multi- plicites in Diol fig . . . . . . . . . . . 61 3 Coupling Constants in p—Nitro- benzoate 2Q (Hz). . . . . . . . . . . . . 62 4 Temperature Dependence of [1,2] 25. [2,3] Rearrangements of 3% and gé. . . . . . . . . . . . . . . . . . . . 70 13 5 Comparison of C Chemical Shifts of Bridgehead Carbons . . . . . . . . . . 86 viii Figure LIST OF FIGURES The tetrahedron, hexahedron, octa- hedron, dodecahedron and icosahedron. Platonic solids or derivatives which have been synthesized: tetra- E-butyltetrahedrane, tetralithio- tetrahedrane, cubane and dodeca- hedrane . . . . . . . . . . . . . Proposed routes to hexaene 1% . . Summary of sulfide to olefin re- actions . . . . . . . . . . . . . Proposed route to disulfide fig. . 1H NMR spectrum (250 MHz) of tri- cyclic ether Zé (top) and its 1,6-_d_2 analog, 1% (bottom). . . . 1H NMR (250 MHz) of diol gg (top) and its 2,6—d2 analog, 2% (bottom). . . Examples of anionic [2,3] sigma- tropic rearrangements . . . . . . 1H NMR spectrum (250 MHz) of butylated product g2 (top) and ix Page . 1 18 49 65 69 analog, 191 (bottom) . . . . . 82 Figure Page 10 Summary of the sulfide contrac- tion route to hexaene 1%. . . . . . . . . 122 Al Proton NMR spectrum (CDCl 60 MHz) 5,11]_ 3! of i(2-oxa-3-oxotricycloE7.2.l.0 undeca-6,9-dien-4-yl 2x . . . . . . . . . 126 A2 Proton NMR spectrum (CDC13, 250 MHz) of 2,6-bis-methylenebicyclo- [3.3.0]octane 6% . . . . . . . . . . . . . 128 A3 Proton NMR spectrum (CDCl 250 MHz) 3! of cis,endo-2,6—bis(hydroxymethyl)- bicyclo[3.3.0]octane 63 . . . . . . . . . 129 A4 Proton NMR spectrum (CDCl 250 3, MHz) of gi§,e§gg-2,6-bi§(bromo- methyl)bicyclo[3.3.0]octane 62. . . . . . 132 A5 Proton NMR spectrum (CDC13, 250 MHz) of gi§,§2dgf2,6-bi§(mercaptomethyl)- bicyclo[3.3.0]octane 66 . . . . . . . . . 134 A6 Proton NMR spectrum (CDCl 250 MHz) 3! of the major product (93. 30%) in the coupling reaction of dibromide 62 and dithiol 66. . . . . . . . . . . . . . . . 136 A7 Proton NMR spectrum (CDCl 60 MHz) 3! of cis,endo-2,6-bis[(tri-n-butyl- stannyl)methoxy]bicyclo[3.3.0]octa—3,7- diene ég. . . . . . . . . . . . . . . . . 138 Figure A8 A9 A10 A11 A12 A13 A14 Page Proton NMR spectrum (CDC13, 60 MHz) of 2-oxa-tricyclo[5.2.1.04’10 Jnona- 5,8-diene Zé. . . . . . . . . . . . . . . 140 Proton NMR spectrum (CDC13, 250 MHz) 4'lojnona- of 2-oxa-tricyclo[5.2.l.0 5,8-diene] Zé o o o o o o o o o o o o o o 142 Proton NMR spectrum (CDCl 60 MHz) 3, of Z-ggggfhydroxymethyl-6-butylbi- cyclo[3.3.0]octa-3,7-diene 82 . . . . . . 144 Proton NMR spectrum (CDC13, 60 MHz) of gi§,gndg-2-hydroxy-6-(tri-n—butyl- stanny1)methoxybicyclo[3.3.0]octa- 3,7-diene 8;. . . . . . . . . . . . . . . 146 Proton NMR spectrum (CDC13, 250 MHz) of gi§,gngg-2-hydroxy-6-hydroxy- methylbicyclo[3.3.0]octa—3,7- diene 88. . . . . . . . . . . . . . . . . 148 Proton NMR spectrum (CDC13, 250 MHz) of gi§,gndg-2,6-bi§(hydroxymethyl)- bicyclo[3.3.0]octa-3,7—diene £6 . . . . . 150 Proton NMR spectrum (CDCl 250 MHz) 3' of cis,endo-2,6-bis(acetoxymethyl)- bicyclo[3.3.0]octa-3,7-diene 11%. . . . . 152 xi INTRODUCTION The Platonic solids comprise a set of five regular polyhedra: the tetrahedron, hexahedron, octahedron, dodeca- hedron and icosahedron shown in Figure 1. Of these, the structural framework of three can be translated to a molecu- lar framework of carbon-carbon bonds without demanding a radical alteration in the directionality of sp3 hybridized bonding, or introducing pentavalent carbons. Figure 1. The tetrahedron, hexahedron, octahedron, dodeca- hedron and icosahedron. The synethsis of organic equivalents of the tetrahedron (tetrahedrane), hexahedron (cubane), and dodecahedron (do- decahedrane) has been realized; the first as the tetra-E7 l 2 butyl (Maier, 1978) and tetralithio (Schleyer 1978) derivatives, the latter two as the parent hydrocarbonsB’4 (Eaton, 1966 and Paquette, 1982, respectively) (Figure 2). U -—r— _,___ LI ---— LI Figure 2. Platonic solids or derivatives which have been synthesized: tetra-Efbutyltetrahedrane, tetra- lithiotetrahedrane, cubane and dodecahedrane. The very recent achievement of dodecahedrane perhaps reflects the degree of difficulty encountered in its 5 synthesis. As a member of the Ih point group, it has 120 symmetry operations; each of twenty methine units is identi- cal to the remainder. The high level of symmetry in this 4'6 and molecule results in some interesting properties predictions: it shows no sign of melting up to tempera- tures of 450°C (for comparison, eicosane-C20H42- has a melt- 1 13 ing point of 36-38°C); the H and C NMR spectra each consists of one singlet; and calculations indicate very little strain energy.5'7 One of the most fascinating features of this hydro- carbon ball is the significant cavity "inside" the carbon framework. Assuming each carbon atom is a point, the in- ternal cavity along one of the 3-fold axes measures 4.32 A. As literal examples of cage compounds, various encapsulated species have been considered theoretically.8 Molecular orbital calculations predict the energies of these "en- 'capsulanes" range from a net destabilization of 48 kcal/mol for a trapped hydrogen molecule, to a net stabilization of 519 kcal/mol for a trapped berylium atom (Table 1). These calculations indicate considerable charge transfer between the trapped species and dodecahedrane sigma bonds in the case of Li+, Be and Na+. Such trapped species would have interesting physical properties and potentially valuable uses. The inclusion compound would simulate an element or iron in the vapor phase. It would represent an organic soluble metal cation. The benefits of a caged radioactive cation in medical re- search range from chemotherapy to organ imaging. Table 1. Predicted Energies for Various Dodecahedrane Inclusion Compounds.a X AE kcal/molQ e‘ +149 H+ -130 H° - 30 H" - 52 He + 22 Li+ -162, -1833 Be -519 Na+ +420.3 H2 + 48 aTaken from Reference 8. b«X at dodecahedrane midpoint 0 relative to X at infinite separation. CFrom CNDO and INDO closed shell calculations, respectively. It is the goal of the work described herein to explore a novel synthetic route to dodecahedrane and metal en- capculated dodecahedranes. If dodecahedrane 1 resisted synthesis for so long, it was not from lack of effort.5 The main avenues of syn- thetic design are readily classified as either convergent or linear sequences. The synthesis of triquinacene 29 in 1964 led to efforts to effect its dimerization to dodecahedrane. Attempts along /\ @ $4 ”"- $537 4 these lines have failed however; for example, photolysis of 2 effects only polymerization,10 while heating 2 at 400°C for two hours under 40,000 atmospheres of pressure resulted in recovery of starting material.11 Connecting two triquinacenes greatly enhances the pos- sibility of isomerization to dodecahedrane by removing translational and rotational degrees of freedom from the monomer. However, the cyclization of bivalvanes like 3 3H have not proved successful.5’12 Alternatively, the synthesis of another triquinacene dimer, 6, was attempted in the hope that homolytic scis- 'sion of one of the cyclobutane bonds would result in a 13 diradical which would then cyclize to dodecahedrane. Unfortunately, the target molecule 6 was never realized. A doubly linked triquinacene dimer 6 has recently been 14 synthesized. Efforts are now underway to transform 6 to symmetrically substituted dodecahedranes 1. Linear approaches to dodecahedrane have been pursued by the Eaton and Paquette groups. Eaton began with bicyclo[3.3.0]octanedione g and through successive cyclopentane annulations gained entry to the "peristylane" series, 2.15 Attempts to cap <01? @7 Q 2 r@ COOMe COO Me LOOMe CKKMHe Us well-functionalized peristylanes are in progress. The only successful synthesis of dodecahedrane to date uses the domino Diels-Alder reaction between 9,10-dihydro- fulvalene 1Q and dimethyl acetylene dicarboxylate to gen- 'erate pentacyclic adduct 11, which is then elegantly trans- formed into C16-hexaquinacene 12.16 Variations on the A, Cl6-hexaquinacene approach eventually led to the synthesis 4 of secododecahedrane 13. Dehydrogenation led to the long-sought-after dodecahedrane. The penultimate molecule in our approach to dodeca- hedrane is hexaene 14. There are two distinct conforma- tions 14 can adopt. One is a compact conformation with 'the two bicyclooctadienes directly facing each other (14a), 10 whereas the extended conformation (14%) has one bicyclo- octadiene twisted 90° with respect to the other. Both conformations have D2 symmetry, however 14a is geometrically similar to dodecahedrane, disregarding distortions due to differences in C-C single bond and double bond lengths and angles. Hexaene 14a is in fact a hexakissecododecahedra- hexaene. By a series of electron reorganizations, 14a can isomerize to dodecahedrane 1. While this type of bond ~L reorganization might have some of the quality of Equation 10 in Reference 5, there are some intriguing points well worth considering. The dodecahedrane precursors in this synthesis would all have an element of symmetry, which may result in a certain economy of steps in attaining the product. The hexaene 14 itself could coordinate a small metal ion, thus trapping it inside the dodecahedrane cavity upon cycli- zation. 11 The final cyclization is a 12n electron reorganization, thus a concerted [n23 + 1T2S + n23 + 1T2s + Tr2S + 1T28] hexa- merization requires photolysis, in accord with orbital sym- metry rules.17 On the other hand, triplet sensitization may promote cyclization gig a diyl intermediate. The hexamerization of six nonconjugated olefins could conceivably be accomplished in the presence of a transition metal catalyst. Studies on the copper(I) photo-assisted di- merization of norbornene lg suggest that a 1:2 c0pper-olefin complex is the direct precursor of the dimer, 16.18 (Scheme_l) f \ ems Scheme 1 12 It has been demonstrated that unstrained olefins such as cyclopentene and cyclohexene undergo c0pper(I) promoted di- 19 merizations. Recently, several natural products have been synthesized in which the key step was copper(I) catalyzed photobicyclization.20 Thus the 1,6-diene 17 afforded 18 ’b’b ’b upon photolysis, which was then transformed to a-pana- sinsene 19 and B-panasinsene 20. II ” .- $2 22 A copper-olefin complex in hexaene 14 could stabilize conformation a which upon irradiation could cyclize to dodecahedrane. Molecular models indicate the two trans double bonds in 14% are close enough to dimerize to a cyclo- butane 21. Interestingly, an olefin methathesis reaction of these two double bonds would only serve to alter the con- formation of hexaene 14! This is due to the equivalence 13 of the four cyclobutane carbons; the metathesis, however, is not degenerate. Finally, cationic or free radical induced cyclization of 14% could lead to dodecahedrane l, or a seggfderivative, 22. mm 14 The most serious drawback to the proposed cycliza- tion of 14 to dodecahedrane concerns the dependence of cyclization on only one of several conformational isomers. Molecular models indicate the olefin protons of the trans double bond in 14a are proximate to the ring olefin; simple twisting of the trans double bond in this conformation does not relieve this interaction. In addition, the starred carbon-carbon bond in 14a exists in a sort of s—gis arrangement with respect to the trans olefin, as opposed to g-Egans in 14b. The conformation 14% is simply a combination of 14a and 14b. Conformation 14a, however, 15 is ideally suited for a metal-olefin "internal" complex, and may prove to be the more reactive conformation. It is also possible that rotation about carbon-carbon bonds 7: in the activated complex 14 would serve to establish the conformation required for dodecahedrane. RESULTS AND DISCUSSION 1. £10 Lactone (or Diester) Dimerization Approach Dissection of hexaene 14 shows it to be two bicyclo- octadienes linked by a trans olefin. The immediate goal of this project was then synthesis of these two halves, and dimerization to hexaene 14. Each half must be of the same chirality to result in 14; two halves of opposite chirality would lead to the isomeric hexaene 23, a megg form which cannot be converted to dodecahedrane. To achieve maximum efficiency then, a resolution should be performed somewhere in the synthesis prior to the dimerization step. The initial approach to hexaene 14 centered on the c0pper(II) oxidative dimerization of diester 24 di- 21 lithium enolate, or of lactone 26 lithium enolate 16 17 [‘15 1‘45 659 " m R (Figure 3). In the first instance, the basic carbon skele- ton of 66 is generated in one step. In the second case, each of the carbon bridges in 16 would be formed sequen- tially. The results in this area were discouraging, however. Both diester 26 and lactone 26 were difficult to obtain (66: 9 steps from commercially available material, 6.3% overall yield; 26: 9 steps, 2.5% overall yield);22 more- over, 26 required a lengthy chromatography to effect separation from numerous by-products accompanying its formation.23 Addition of diester 26 to two equivalents of LDA, fol- lowed by addition of cupric triflate gave a complex mixture of products. Analysis by GLC/MS showed the two highest molecular weight components (besides recovered diester 26) had parent ions at m/e 276, strongly suggesting an intra- molecular reaction of the bis-enolate resulting in diesters 18 cog: a 2 sq IDA A on TH?» '78"? Lia no; no): : Cu(II) I a: w : 31130.: H’.) g. '5. \- 2) 1>t>(om:),,\~£ A: V - a: 0 uo’. H“ 1) IDA A a ______ -9 o 2) Cu(II) . 2 O £8 31 1) IDA .4, <--—---—- 1‘ 2) Cu(II) no; __ 3) Ho': 11" V a) Pb(0Ac)u CH3C(or:t)3 Figure 3. Proposed routes to hexaene 16. 19 61. This type of product is not unreasonable, given the :==§flfi s on 1 63(11) A / 0,5: IJO é} V proximity of the enolate anions and the entropy factor operating in favor of intramolecular reaction. Addition of lactone 26 to a solution of LDA, followed by addition of cupric triflate yielded the dimer, 61, in addition to recovered starting material. 1 H NMR, IR, and MS spectra all supported the assignment of structure 21 to this compound. The isolated yield of crystalline 21 is only 23- 15%, however. In reactions using cupric bromide or cupric chloride, the d-bromolactone 6% or a-chloro- lactone 66 was observed to the extent of 36% (isolated yield) and 14% (1H NMR yield), respectively. The use of cupric triflate24 for the enolate dimerization effectively CG 1%" o x ggx c1 0 suppressed this side reaction. Br 20 2. Sulfide Contraction/Elimination Approach Hexaene 66 can also be approached from bis-sulfide 66. A number of rearrangements may serve to transform the sul- 5 fide linkage to an olefin (see Figure 4).2 Fortuitously : ----- -> — w \e 84. $33, in this molecule, the rearrangement-elimination sequence is forced to proceed to a trans olefin, as molecular models indicate that a trans double bond is the only olefin geometry 16 can adopt. A gig olefin at this juncture would be impossibly sterically crowded. The approach to 16 described herein has been used suc- cessfully in the synthesis of various cyclophanes. The cyclodimerization is achieved by reaction of dithiolate 66 with dibromide 66 under conditions of high dilutions. Ring contraction of 61 gig a Stevens rearrangement, fol- lowed by Hofmann elimination of dimethylsulfide yields the cyclophane 69.26 This sequence has several advantages over a hypothetical 21 1) BuLi 2) CH3 I (Ref. 327) .l 1) r(on 30)zcn+ BF“- 2) t- BuOK (Ref. 27) (Ref. 25) 5"“ SPh 5°"! 5 Ph 1) [OJ 1) (CH30)20H+ BFu- 2) 11 2) NaH (Ref. 28) (Ref. 25) C§§uzf> 14m .3' mm f' 2) BuLis arZ' r' -soz r" (Rot. 29) Figure 4. Summary of sulfide to olefin reactions. 22 (If) 2) g-nuox sea, 1) lo} 0+ n?“- 2) run SCH, R 38 NM one step dimerization (2.3;, a Wittig reaction between ap- propriate components). The two subunits are brought to- gether gradually with three atoms separating the bulk of each molecule from the other, thus minimizing steric ef- fects in the crucial dimerization step. The starting materials are usually obtained from the same source, since the dibromide is easily converted to the dithiol. Finally, sulfur extrusion may be accomplished in a number of ways (Figure 4). Recently, while this work was in progress, dithia- 30 hexahydro[3.3]paracyclophane, 66, and the two diastereomers 23 of dithia[3.3](2,6)triquinacenophane, 61 and 66,14 were synthesized by this route. In addition, there has recently appeared a publication which details an improved synthesis of sulfide macrocycles, based on the reactions of cesium dithiolates with dibro- mides.31 The improved yields over literature values in some reactions were truely spectacular, as in the synthesis of 1,7—dithiacyclododecane, 66, prepared in 80% yield as compared to 0.8% previously reported.32 CSCO3 BrCH2(CH2)nCH2Br + HSCH2(CH2)mCHZSH ————) WSW (cm). (cm)... L.) n- 4% The obvious route to bis-sulfide 66 is by reaction of :3 ll m E II m 10 ll m E II dithiol 66 with dibromide 66. The dithiol can be derived 24 from dibromide 66, while it was anticipated that diol 66 would serve as a convenient precursor to the dibromide. As all of these compounds were unknown, attention was focused on the synthesis of diol 66. Most notably, 66 is a homologated version of diol 61, preserving the cis, endo H0) H5» BK} \\ ‘\ OH 8' SH 3, 4,2 :4, configuration at C-2 and C-6. The utilization of diol 61 to introduce gig-endo alkyl substituents via a [3,3] oxy— Cope rearrangement (Claisen rearrangement) had proved suc- cessful in other work in this laboratory (Scheme 2).23’33 Consideration of various methods to degrade products from these Claisen rearrangements led to the unsavory prospect of multistep degradative sequences34 in which some of the intermediates (E;E;r d-silyloxyaldehyde 66) or products (229;! dialdehyde 66) might be unstable. The most promising route to 66 was homologation of diol 61 via a bi§[2,3] sigmatropic rearrangement (Equa- tion 1). In this way, a side chain with the correct number of carbons (one) would be introduced, thus 25 cog: .0" / mama), A H+ ’ [3.31 HO 6?. 5:0,c/ 8% cuz—cuzoczas 115mm); 1 OH \L p ‘f ‘ ' *1 «w.» A ' I [3.3] ’ o HOJ BtzA101°PPh3 [3.31 /CH0 8' ”\ OH Scheme 2 26 Me,SiO g CHO EHO Mask) i die \CHo 51 alleviating the need for degradation. The Wittig [2.3] sigmatropic rearrangement of allyl ethers appeared ideal Y=sx [231 be CO x “0!“ ‘—Y for the purpose at hand. While the classical reaction in- volves treating an allyl benzyl ether with g-butyllithium 35 "W (Equation 2), Still has reported a [2,3] rearrangement :n and Ph 2) 320 | (2) ‘OH 27 in a-alkoxyorganolithium reagents derived from allylic alcohols. Generation of the carbanion exploits the ready exchange of tin for lithium in organostannane 61.36 \ 1):!!! A \ 2) masncazz I’ 0" 66 OVSn Bu, Bum, -78 C TH? (Ni ‘V 1) -78 C / ‘1___ 2) 11+ \ + BuuSn 3% con, Li [2,3] Rearrangement of the a-alkoxy anion leads to homo- allylic alcohol 66 in good yield. This methodology was used in a highly stereoselective synthesis of the C13 CecrOpia juvenile hormone, 66.37 Although this rearrangement had been demonstrated to work well in some cyclic systems, octalols 66 and 61 were reported to give a mixture of [1,2] and [2,3]rearranged 28 \ 1) m __\ at: 2) m35ncazx ’ ‘OH (OH \ occ Bum/MgA -20 cfi Bufihl lhhsn 79’ HO) EN) \ \ \ E 6 55 wm . . 36 products in low yield. 93‘ = o-OH B-OH 29 The application of this rearrangement in cyclopentenol systems appears unexplored, however. Indeed, this aspect of anionic [2,3] sigmatropic rearrangements is not cited in the literature at all. Thus an Opportunity presented itself for evaluation of endocyclic anionic [2,3] sigma- tropic rearrangements in a cyclopentene ring system (Scheme 3). The remainder of this thesis pursues this idea, as well as investigating reactions in model compounds which would transform diol 66 to sulfide dimer 16. All of the following‘ 0". e0\\ _ _ _ _ _ 9 3035,, SnBu, Ho’ of .972 I I I /°" $ 7; FOR,“ -32 33:2]- ., _ g 2)H+ H02 ”CH5 Scheme 3 30 reactions were performed on racemic material. 2.1. Model Studies To demonstrate the feasibility of the synthetic approach to hexaene 1 and to work out reaction conditions for some of the required conversions, the synthesis of diene 66 was undertaken. The immediate goal was sulfide dimer 66 which would serve as the precursor to diene 66. The starting material in this synthesis is bicyclooctanedione 61, which is available in hundred gram quantities in three steps from 38 The route devised to convert dione dimethylglutarate. 61 to sulfide dimer 66 is straightforward, and is outlined in Figure 5. Methylenation of the dione was anticipated to be the most troublesome step, given the well-known propensity of cyclopentanones towards enolization in the presence of base, leading to self-condensation reactions. The methoxy- methylenation of cyclOpentanone gave 14.5% of the desired 31 IMjF-GQZ --------> 0 cu, 3% .62 I : 1’ 32% . 2) H202. hon I W pm Samoa 1) I801 2)N¢I <— ------- (I) Icu, HOCH, 85. 8% I I I -33 I I I l w pmsu : —————— S + g; 9 <<::§E§;7 HSCH, 66 'V'v Figure 5. Proposed route to disulfide 66. 32 vinyl ether, along with a 33% conversion to 2-cyclopentyli- dene cyclopentanone when the ylide was prepared in the o I ocu, o O W“ A O + C55 I 14.5% 33% 39 traditional way with gfbutyllithium in THF. Several improvements or modifications of the original procedure have been devised to deal specifically with easily enolizable and/or hindered ketones.40 It appears that the choice of a base and a solvent for generation of the ylide component greatly affects the success of the re- action. For example, sodium i-amylate deprotonation of methyltriphenylphosphonium bromide in xylene gives the phosphorane, which reacts with cyclopentanone to yield 81% methylenecyclopentane.38a The presence of teamyl- alcohol in the reaction is thought to render the enoliza- tion process reversible. Several intermediates in the synthesis of triquinane natural products such as mode- hephene 6141 and isocomene 6642 were successfully prepared in this way. A solution of sodium ifamylate in toluene was prepared, and used to generate methylene triphenylphosphorane from methyl triphenylphosphonium bromide. Addition of the 33 m. 44-) . . diketone in toluene and refluxing overnight, followed by the usual workup led to diene 61, readily identified by the four proton AB quartet at 64.84 in the 250 MHz 1H NMR, and a sharp absorption at 1650 cm-1 in the IR. This particular preparation gave a low isolated yield of the diene (g3. 20%) and was contaminated by ifamyl alcohol. It was found that potassium i—butoxide in benzene gave map-CH2 lhfl.ru£uu ’ . 81-8 0 9f 34 improved results, although deprotonation of the phosphon- ium salt proceeded at a slower rate owing to the insolu- bility of the base in this solvent. The isolated yield of diene (bp 170°C) improved considerably when it was dis- covered that it codistilled with the aromatic solvents under reduced pressure, but not at atmospheric pressure. A substantial excess of ylide (93. 1.7 equivalents per carbonyl) gave the best results. With these modifications pure diene 61 could be prepared on a ten gram scale in 81-89% yield. It was anticipated that hydroboration/oxidation would transform diene 61 to the cis, endo diol 66.43 Treatment of 61 with two moles BH -THF complex followed by alkaline 3 hydrogen peroxide gave diol 66 in quantitative yield. Analysis by GLC showed the crude product was a mixture of stereoisomers. The major isomer had the longer reten- tion time on a QF-l column, and predominated to the extent .CHzO CH OH 1) mtg-m *2] H202. N10)! 7 008 HOCH, I ”I; N HOC 8% of 6:1 over the minor isomer when the reaction was run at -22°C. Lowering the reaction temperature to -45°C increased this ratio to 9.5:1, or 90% one isomer. Thus, the major isomer was assigned the cis, endo stereochemistry while 35 the minor product was presumed to be the trans isomer. Acylation of the 6:1 diol isomer mixture followed by chromotography gave a product which displayed a small acetate methyl at 62.06 in the 60 MHz 1 H NMR, and a larger one at 62.03 in approximately an 11:1 ratio, or a 6:1 ratio of cis:trans. Considering that the ultimate coupling reaction required gig, endo material and that both the dibromide and dithiol were to be derived from diol éé' even 10% of the trans isomer would limit the cyclized sulfide dimer to a maximum of 80% yield. Therefore it would be advantageous to increase the exgzendg selectivity in the hydroboration step. The 9-BBN°THF complex has been used as a highly hindered hydroborating reagent,44 but removal of the oxidation product of 9-BBN, 1,5-cyclooctane- diol, would require a chromatography. Finally, the use of disiamylborane4S (Sia BH) resulted in virtually exclusive 2 exo attack on diene gg (>98% selectivity by GLC). The amyl alcohol oxidation by-product (bp 111°C) could be conveniently removed under high vacuum at room tempera- ture. When this was done, the diol was pure enough to 9H,OH 1) sugar: m :; 2) 11202. man 1001 ~ 3% HOCH, '93, "'I, "a, 36 use directly in the next step, but it could also be kugelrohr distilled to provide an analytical sample. Treatment of this diol with mesyl chloride in pyridine gave a 95% yield of the dimesylate g3, which slowly solidified on standing. Exposure of this material to sodium iodide in refluxing acetone provided diiodide Qé in 63-70% yield after chromatography. 9H,0H ‘ ngMs . I301 4; m pyridine I 95$ . HOCH2 "5°C": 6 63 M» N11 19"" acetone = 63-70! "I. a .s=~ Alternatively, a one step conversion of diol g; to the dibromide fig could be accomplished with triphenylphoSphine/ 46 carbon tetrabromide. When a large scale (22' 10 9) reaction was conducted in ether, the amount of precipitated 37 triphenylphosphine oxide was sufficient to interfere with the mechanical stirring. Removal of triphenylphosphine oxide gave a yellow oil which contained large quantities of bromoform and carbon tetrabromide. The former could be removed under high vacuum at ambient temperature (bp 146-150°C). The latter contaminant was clearly visible in the 13 C NMR of the crude product at -29.52 ppm, initially appearing as a fold-over at 12.52 ppm. Removal could only be effected by chromatography, at which point pure di- bromide 57 was obtained in 55% yield. Changing the solvent. SCH 20H SCH 28' ”If/031'“ A man or c1130): 7 now, 55‘7““ area, 0% to acetonitrile resulted in a slightly improved yield, 74% on a small scale, 61% on a larger (2 g) scale. In any event, either the diiodide or dibromide could be transformed to dithiol fig any number of ways. Direct treatment with sodium sulfide or sodium hydrogen sulfide was not attempted, in anticipation of unwanted intramolecu- lar displacements yielding sulfide ég. Alternatively, treatment of either the dibromide or diiodide with sodium 38 thiolacetate followed by lithium aluminum hydride reduction, or with thiourea followed by hydrolysis,47 would effec- tively circumvent sulfide formation. The latter classical method was used due to the availability of thiourea in the lab. Heating a 0.30 M solution of dibromide $9 in 95% ethanol with thiourea in a closed tube at 100°C led to the crystallization of bis-thiouronium bromide 11 from the reaction mixture. Chilling the solution and filtration of the solid gave a 95% yield of the salt. Alkaline Bre 3H2 9H,Br $2st— 3 .« m. “2" N“2 2 non :\’ f 5‘ NH,\ 5 . 9 : BrCH, c—s CH, 69 / 7 NH, Bra «A1. hydrolysis was achieved by refluxing the salt in aqueous sodium hydroxide, followed by acidification of the 39 hydrolysate. In this respect, it is desirable not to include unreacted thiourea in the hydrolysis, due to the distinctive smell of hydrogen sulfide liberated during acidification. The crude dithiol was isolated as a clear, viscous oil after flash chromatography, possessed of the characteristic thiol smell. Protection from light and "'00: n I: "in I: 21% 1) Nam-I, reflux% 2)H+ f HSCH, oxygen was essential to inhibit the formation of disulfides; even storage under nitrogen in the refrigerator was not sufficient to completely preclude this process. The preceeding steps are summarized in Scheme 3. Initial studies on the reaction of dibromide £9 with dithiol QQ followed the general procedures used in the synthesis of dithiacyclophanes, i;g;, slow addition of the dihalide and dithiol to a dilute solution of base over a period of several days in order to achieve high dilution conditions.14'26’30 The high yield synthesis of some sulfur containing macrocycles via cesium thiolates appeared a promising route to the desired sulfide dimer Q0. 40 H2 mamnanr _t_-BuOK 7 Dawn“ 0 CH 8 2 61 9‘ 5 93 'VD 1) 813235 2) 11202. than 1001 \L CH,Br 5: H10" 4 map/car“ 033cm 451-7“ HOCH arcu, ‘ 8% .5. ‘1) nzncrmz 2) N301!) 11+ 89! w 9H,SH g. . HSCHI Scheme 3 41 Thus dibromide £2 and dithiol gé in DMF were added syn- chronously to a well-stirred suspension of cesium carbonate in DMF at 95°C. The addition took four days, after which time analysis of an aliquot indicated complete consumption of QQ and g2. Filtration of the excess base and removal of DMF led to the isolation of a yellow-white precipitate. Column chromatography led to the isolation of three distinct components in 1%, 3% and 30% yields, respectively, as well as small amounts of polymeric material. The smallest fraction eluted first from the column, and is tentatively identified on the basis of its mass spectrum as the intra- molecularly oxidized product, disulfide 2% (Scheme 4). The second component to elute from the column, isolated in 3% yield, was not readily identified by its 1H NMR spectrum. The mass spectrum indicated this product was a trimer which incorporated four sulfur atoms. This product might be the diastereomeric cyclic trimers with a disulfide, instead of a sulfide, linkage at one juncture. The last and largest fraction also had no distinguishing features in its 1H NMR, and additionally, was not sufficiently volatile to be analyzed by mass spectroscopy. The molecular weight of this component will have to be assessed by other means, which may include freezing point depression measurements, or osmometry. 42 9,13, EH25}! 5 § CD + CO BIC“; a "SCH! 8% 032003 DEF 95‘0 Jr . 1» Scheme 4 43 2.2. The Fate of One Bis-Wittig Rearrangement In extending the reactions performed on the model com- pounds to the unsaturated system, the synthesis of diol 4Q had to be addressed. As outlined previously, a bistittig rearrangement in a derivative of diol 41 was envisioned to proceed to diol 45. Diol 47 is available in multigram OH on { 3. HO/ 4'3 :6 quantities from the diisobutylaluminum hydride reduction of 22,23 dienedione 13, and is obtained in 88% yield as a 4:1 mixture of cis, endo:trans stereoisomers. The cis isomer can be fractionally recrystallized from acetone and is ob- tained pure in 57% overall yield. Recent endeavors in this lab have improved on the synthesis of diol 47 by using a cerous chloride mediated sodium borohydride reduction of dione 73.48 This combination of reagents results ex- clusively in carbonyl reduction, and diol 47 is easily prepared in quantitative yield. The reaction is fast, clean, and relatively easy to work up; moreover, the use of an alkyl aluminum reagent is avoided. In a brief study, 44 the ratio of gig, endo : trans isomers was studied as a function of reaction temperature. Thus, at 0°C, this ratio was 6:1; at -10°C, 7.6:1; and at -15°C, 8:1 (GLC analysis, two consecutive runs). (Interestingly, this temperature dependence parallels that observed in the BH3°THF hydroboration of saturated dione 61.) o P” Km 00 1 6: a co CBC ' O C: 032 (88‘) ° 3 Ho -15°C 31?. 4‘72 9H + co H0) Treatment of diol 41 with two equivalents of potassium hydride in THF, followed by iodomethyltri-n-butyltin49 gave the bigftnbgfbutylstannylmethyl ether 58 in 95% yield, after workup and chromatography. Although the alkylation is not complete until after 36 h at room temperature, it has been noted that l8-crown-6 will enhance the rate of alkylation of hindered secondary alcohols.36 45 Addition of bis-stannane 58 to two equivalents of n- butyllithium in THF at -78°C produced dianion i, which could be protonated at low temperature to give the dimethyl ether 24. Upon warming the solution of dianion i to -30°C for 36 h, the reaction changed from clear and colorless to a deep reddish-orange. Warming of this solution to room temperature, followed by workup and chromatography to remove tetra-nebutyltin, gave a clear colorless oil homo- geneous by TLC, which exhibited a complex 60 MHz 1H NMR spectrum. Both the absence of exchangeable protons in the NMR, and the absence of an -OH stretch in the IR indicated the product was not an alcohol, although a peak at 64.95 in its 1H NMR was a characteristic chemical shift for the allylic proton a to oxygen in diol 41 and bigfstannane 58. The appearance of a parent ion at m/e 134 confirmed the suspicion that the product was an ether of structure 74 or 76 (see Scheme 5). Moreover, only two out of ten protons 1H NMR, making exhibited overlapping signals in the 250 MHz single frequency homonuclear decoupling feasible. The distinguishing features in the NMR include one proton at 64.92 (dd, J=8 Hz, J=2 Hz), protons at 63.93 and 63.68 comprising part of an AMX system, and one proton at 63.49 (q, J=8 Hz) (see Figure A9). The signal at 64.92 was easily identified as H2, and the AMX splitting almost surely arose from the diastereotOpic protons on C-9 in 15 or C-lO in 16. The quartet at 63.49 could be ascribed 46 (M1 £p~—\ 1) KH m SnBu; :% l3 s 2) BUBSnCHZI "3" L: " 96% 0 $1 2% n-BuLi THF, -78'c W o ‘ gOCH,“ O 1* {D C ‘ o 70% . ZR LICH,O . 1 or __ H+, -78°C 0/’ v 23 SW3". CH4; 2% Scheme 5 47 to only one proton out of both structures, that of C-10 in 15. Irradiation at the frequency corresponding to 63.49 collapsed the doublet of doublets at 64.92 to a doublet (J=2 Hz), but had no effect on any of the olefinic pro- tons. Conversely, irradiation of the doublet of doublets at 64.92 transformed the quartet into a triplet (J=8 Hz), and removed a small (2 Hz) coupling constant from the pro- ton at 65.68. Therefore, the proton which appears as a quartet at 63.49 is not allylic, and structure 15 was assigned to the product. Later experiments proved that 75 was the isomer ob- tained. 2,6-g2 Biol 11 was synthesized from sodium boro- deuteride/cerous chloride reduction of dienedione 73. Alkylation of 17 with tri-n—butyliodomethyltin resulted in 18, which when treated with nebutyllithium produced a single product 79 whose 1H NMR clearly indicated the ab- sence of protons at 64.92 and 65.75 (see Scheme 6 and Figure 6). As mentioned before, dianion i could be trapped at low temperature. gig-stannane 58 was added to nebutyllithium in THF at -78°C. After 15 minutes aqueous acid was added, and the solution warmed to room temperature. Workup and analysis showed only the presence of dimethyl ether 71, >98% by NMR. Thus dianion i was being formed at -78°C, but some other reaction intervened before rearrangement could occur. 48 .Asouuonv ww .moncm mmuo.H mun cam Agony w» umcum oflaosowuu mo lam: ommc asuuommm mzz ma .m wusmwm m musqflm ‘0 C In D 49 3) g V 50 ' Dyna nun“ .3 -+ 00013: 61:20 7 0 011301) HO“; D «7,3 «7,2 1)KH 2) Bu’SnCHzI I) {‘II'IIIF> z;r Edd D ‘ W 1O Scheme 6 The failure of dianion i to undergo rearrangement at the low temperatures characteristic of [2,3] sigmatropic pro- cesses (vide infra) indicates that at -30°C, the dianion reacts exclusively by a pathway lower in energy than even the normally facile [2,3] rearrangement. During the 51 course of the reaction, it was noted that the solvent gradually assumed a deep orange-red color, suggesting THF was being slowly degraded. It therefore seemed probable that dianion i was deprotonating THF to give anion ii, which quickly cyclized to tricyclic ether 15. gocmu (pen, THF ~ % 5 i ' H ‘ 2 t L“: J) (3 CFQLI _i_ . E «moon3 As a test of this mechanism, bigestannane 58 in ether was treated with one equivalent of nebutyllithium at -78°C, and allowed to warm to 0°C over the course of several hours. An aliquot analyzed by TLC showed no less than six com- ponents! Of these, five were identified by comparison with authentic isolated samples (see Scheme 7). The sixth 52 O 8 5'18“. 3) 1g Bum. -78.C-0'C% Buusn 0 “i8 2) H+ I3u£5n L: a 3g ocu Qj £ §j cap - cm; 74 75 '39 «A: «A: . BuBSnCHZOH 6% 1 1 m. as 3% i ) 9q 7; 2) 0°C. 6 h: 22°C, 8 h 3) m 0 Zé + BuQSn + BHSSnCHZOH 3% Scheme 7 53 component, which possessed the third highest R in the f group, was most probably stannane 88. Allowing the reaction to stir at 0°C for six hours and at room temperature over- night, produced a greatly simplified TLC analysis, showing only three components: tetra-nfbutyltin, hydroxymethyltin 8k, and tricyclic ether 18 (Scheme 7). The various inter- mediates had evidently funneled off to tricyclic ether 88 by scrambling of tri-gfbutyltin in bis-stannane 88 and dianion i. After chromatography to remove the tin com- pounds, Zé was isolated in 72% yield, 69% from diol 88. This method represents a convenient preparation of 18, and supports the prOposed mechanism for its formation. In another experiment designed to suppress protonation of dianion i by the solvent, big-stannane 88 was added to 2.3 equivalents of nebutyllithium in ether at -78°C, al- lowed to warm to 0°C, and finally stirred overnight at room temperature. Workup of the reaction led to two pro- ducts (besides tetra-g—butyltin), tricyclic ether 88 and another compound which had a much longer retention time on GLC and TLC; both products were isolated by column chromatography. The new compound, obtained in approximately 30% yield, showed an -OH stretch in the IR, a doublet at 63.7 in the 60 MHz 1 H NMR, and what appeared to be a con— densed AB quartet for the olefinic protons at 65.6, as well as some higher field multiplets (60.6-l.6) attributed to contaminating tetra-n-butyltin. It appeared that simply 54 changing the solvent had produced the desired rearrange- ment of i to diol 88! This new compound exhibited some worrisome characteris- tics, however. These included the "wrong" migratory ap— ptitude on TLC for a diol in a given solvent system (too large) and the observation that the new compound could be easily analyzed by GLC under conditions where the saturated analogue, diol 88, failed completely to elute. A GLC/MS proved the intuitive conclusion of these observations, £;E;I that the new compound was not diol 88. A parent ion at m/e. 192 did not represent diol 88 (MW=166) but rather, tri- cyclic ether 7 (MW=134) plus butane (MW=58), or something ’1: H of structure 82. ’b’b °\... 38 Again, quenching experiments performed at -78°C on bigf stannane 88 added to nebutyllithium in ether gave almost exclusively dimethyl ether 11 and some 18 (96:4, 18:18, from GLC), indicating almost complete formation of di- anion 1. Interestingly, the amount of 81 isolated in the previous reaction was almost identical to the amount of ex- cess nebutyllithium employed. Since the -CH20H group was positively present in the molecule, it appeared that 55 dianion 1 closed to tricyclic ether 18 as before, but also that excess nfbutyllithium cleaved 18 in some way. To examine this possibility, tricyclic ether 18 in ether was subjected to nebutyllithium at 0°C. Following the reaction by TLC and GLC showed the formation of a compound with the identical Rf and retention time as 81. It appeared I: C) \.«I“ that some of the nebutyllithium was destroyed by the sol- vent, as more than a stoichiometric amount was required to complete the conversion (approximately three equivalents). Workup of the reaction gave an oily compound identical to 81 isolated earlier. The stereo- and regiochemical features of 81 were not readily apparent from inspection of its 1 H NMR (Figure A10). Overlapping multiplets for the bridgehead protons suggests the 3,7-diene and not the 3,6-diene is formed, as the latter isomer would be expected to have quite different bridgehead proton chemical shifts. While organolithium.reagents generally do not attack 56 unconjugated olefins,50 there is precedent for cleavage of allylic ethers. Broaddus demonstrated in 1965 that allylic ethers 88 refluxed in hexane with nebutyllithium undergo a displacement reaction.51 Complexation of the alkyllithium l)BuLi/hexane reflux ROCH CH=CH > ROH + BuCH CH=CH 2 2 2 2 2)H20 8% 60% is considered the activating factor in these additions. The proposed mechanism involves a six membered cyclic transi- tion state resulting in SNZ' displacement, but as only allyl ethers were used in this study, a direct displace- ment cannot be ruled out. It was noted that lower reaction temperatures (BO-40°C) resulted in significant isomeriza- tion to the alkyl gis-propenyl ether, which was stable to the reaction conditions. Other examples of alkyllithium displacements include the addition of isopropyllithium to allylic ether 88 and homoallylic ether 88.52 Allylic rearrangement was not demonstrated in 81 however. Moreover, it is well known that alkyllithium reagents add to the strained double bond 53 in norbornene. In 88, chelation of the ether oxygen with lithium is essential for the addition, and directs the 54 regiochemistry of attack. The relative configuration of 57 i-PrLi 4¢x\//ocn, «36 A i-Pl’Li OCHJ A 54% 85 M» 1, g. 80% (L the addend in 88 was assumed, however, and not rigorously proven. R” A, ”’° % . ocu3 Li'!‘ OCH. OCH: «‘39 RLi I 7‘ NR That nebutyllithium adds easily to tricyclic ether 18 perhaps reflects the ideal geometry for a six membered transition state when the lithium chelates with the ether oxygen. The same type of process may also be responsible 58 ® 1‘; \‘ b/R ‘Li—CH,C.H, - 31—33” for the facile intramolecular displacement of alkoxide in ii. 2.3. Another Wittig Rearrangement An alternate route to diol 88 would involve two sequen- tial rearrangements. It seemed clear that suppression of cyclization of i to 18 could be achieved by converting one of the side chains to a poor leaving group. This neces- sarily destroys the symmetry of the molecule, and moreover renders what was a one-step reaction a multi-step sequence. But it would be interesting to see if a [2,3] rearrangement actually could take place in this system. The best protective group with a poor leaving ability was conveniently already part of diol 81; elimination of L120 appeared unlikely (however, see below). Hydroxystan- nane 81 was prepared in 26% yield from the reaction of diol 81 with potassium hydride and one equivalent of iodomethyl- tri-nebutyltin (Scheme 8). Separation from bigestannane 88 (31%) and unreacted diol 81 (23%) was achieved by column chromatography. Addition of hydroxystannane 81 to 59 OH 3 1)KH ‘4—3- Egaliu' 2) 13“3‘°"“’“21 ' + 3&3 + «473 HO ,6, 31¢ '63 20am 8t2°- ~78°c .OCH,Li f Lid 1) O'C- 22°C 2) 11+ 9".on 3 °’ CC) HO HO cHon 88 'Vb 38 Scheme 8 6O nebutyllithium in ether at -78°C, followed by gradual warm- ing to 0°C, and finally to room temperature, produced a heavy white precipitate. Acidification and work up of the reaction led to a multi-component crude product. Analysis by TLC showed two spots with the "correct" Rf for a diol, perhaps indicating regioisomers resulting from [1,2] and [2,3] rearrangement. The presence of other products was mysterious, especially as GLC analysis of the crude reac- tion mixture showed these were not trace impurities, but major by-products (4 components, 2 of which were not com- pletely resolved, in approximately 2:1:l:3.5 relative ratio). The various products were isolated by column chromatography. The major product, formed to the extent of 40%, was a diol of either structure Qg or g2. TLC, GLC 1H NMR (250 MHz) indicated the major product was a 1 and single isomer. Several features of the H NMR argued in favor of structure gg (see Table 2 and Figure A12); most notably, the protons at 63.42 and 63.56 assigned to the bridgehead carbons both appeared as triplets with a large (93. 8 Hz) coupling constant. Structure $2 on the other hand would be expected to exhibit a quartet for one of the bridgehead protons, as in tricyclic ether 2%. The trans isomer of fig would be expected to have a smaller J5,6 (approximately 4 Hz, see below); therefore H would appear 5 as a doublet of doublets (J=8 Hz, J=4 Hz), and this is not observed. 61 q gaps-I ,~, . H3 .8 1"» Table 2. 1H NMR Chemical Shifts and Multiplicites in Diol 88. W Chemical Coupling Constants Proton Shift (6) Multiplicity (Hz) 1 3.56 t of q Jt=8, Jq=2 2 4.81 d J1'2-8 3 5.94 ddd J3, =6, J3'5=l, J2, =0.5 4 5.74 d of t J3'4=6, Jt=l 5 3.42 t of m Jt=8 6 3.00 br m ------- 7 5.64 d of t J7,8=6' Jt=2 8 5.80 d of t J7,8=6' Jt=2 9A 3.68 dd JAB=11’ JA,6=5 9B 3.62 dd JAB=11, JB,6=7 62 The coupling constants in the penitrobenzoate 29 were determined with the use of a europium shift reagent (Table 3).55 In this case, J1 2 is 4 Hz, although interestingly, I OPNB 39 Table 3. Coupling Constants in p-Nitrobenzoate 20 (Hz). Jl,8 2.0 J3,4 5.6 J7,8 5.4 J2'3 2.3 J6a,7 2.3 J1,2 4.0 J68,7 2.3 J1,5 7.2 J4'5 2.0 J6a,7 = J68,7 = J2,3 = 2.3 Hz. The one diagnostic 1H NMR spectral feature in the determination of configuration at C-2 is the coupling constant corresponding to J1,2 in 29. Finally, it should be noted that one of the upfield triplets observed in the 1H NMR of the rearranged product is not due to the -C§CH20H proton. This proton is coupled unequally to the methylene protons, as observed in the 63 AMX system of the latter. Initial assignment of chemical shifts to individual protons was accomplished with decoupling experiments. These experiments showed 1) the protons at 64.81 and 63.56 are coupled with J=8 Hz, 2) protons at 63.00 and 63.42 are also coupled with a large (23' 8 Hz) coupling constant, 3) irradiation of the signal at either 64.81 or 3.42 affects olefinic protons at 65.94 and 65.74, and 4) irradiation of the signal at 63.00 collapses the AMX system at 63.68 and 3.62 as well as affecting olefinic protons at 65.64 and 65.80. The observation that irradiation of signals at 63.42,. 3.00 and 4.81 all affect two olefinic protons suggests the double bonds are at C-3 and C-7 (the signal at 63.56 could not be irradiated, because it was obscured by adjacent pro- tons). In contrast, irradiation of one bridgehead proton in alternate structure 89 would affect all the olefinic protons, whereas irradiation of the other would have no effect. As a final proof of structure, 2,6-d2 diol 77 was mono- alkylated with iodomethyltri-n—butyltin yielding 91, which was then treated with n—butyllithium as before to effect rearrangement (Scheme 9). Workup and chromatography led to the same mixture of products observed earlier; the component with the identical Rf as 88 on TLC was isolated, l and analyzed. The H NMR of this compound clearly showed the absence of signals at 64.81 and 63.00 (Figure 7), thus 64 .AEouaonv ww .moncm NW! m.m mun can Agony ww H036 «0 Ana: ommc mzz m H .5 musmflm 65 66 Dye c H, Li 2¢m_mnd 31:20 -78°C V? \ o \“\‘ [JO Scheme 9 validating the earlier structural assignment. The major product in this reaction therefore results from a [1,2] rearrangement of carbanion iii. Moreover, the proton at 63.42 in 88 (t of m, Jt=8 Hz), was now in 92 a broad doublet, J=8 Hz, thereby supporting the assigned configura- tion. The by-products in the rearrangement of 81 to diol 88 were isolated, and identified on the basis of spectral properties to be compounds 82, 93, and 47, in 12%, 7%, 67 and 14% isolated yields, respectively. Thus the other so -00.. product with a "diol Rf" on TLC was not an isomer of diol 88, but rather the starting material one step earlier. The origin of these products will be discussed later; it is significant to note, however, that 88 represents the only rearranged diol. Anion iii therefore rearranges exclusively by the [1,2] pathway. The [2,3] rearrangement is depicted as a concerted 56,57 process (path A in the scheme below) whereas the [1,2] b (Ref. 56) Q. 0 68 rearrangement is formulated as a nonconcerted radical 56,58 In dissociation/recombination process (path B). systems where either a [1,2] or a [2,3] rearrangement may occur, studies have shown that: l) the [2,3] rearrangement proceeds rapidly and smoothly at low temperatures (-78°C to -20°C) and 2) that the [1,2] rearrangement is a competing, 56 A few examples of anionic [2,3] higher energy pathway. sigmatrOpic rearrangements are outlined in Figure 8. In all of these examples, rearrangement takes place at -20°C or below. In a study of the benzyl ethers of substituted allylic alcohols, Baldwin found that the reaction temperature af- fected the relative amounts of [2,3] 35. [1,2] rearranged products (Table 4).56 (OH ems/\OML (“Haw c." ADM 95 Mb In both cases, the yield of [2,3] rearranged product in- creased with a decrease in temperature. In the rearrangement of 28 at 0°C, the [1,2] process occurred to the extent of LDA/THF c Lo -LiCN Qg-WC ——_')_§ o<"‘N—-)\_2J__) __)\_73j 90% (Ref. 59) C’H" \ 2 eq LDA/THF j \ '78 C 7" 0 fl euo MMcoH j com co,e Lao (Ref. 60) OH! C.H, BuLi A m aka c'"’ D THF -80°c7 D ———) _ CH, >___)\cn, / “cu, CH, CH, (Ref. 57) + 20% [1,2] (Ref. 61) Figure 8. Examples of anionic [2,3] sigmatrOpic rearrange- ments. 70 Table 4. Temperature Dependence of [1,2] vs. [2,3] Rear- rangements of 94 and 95.3 Reactant Temperature (°C) Product Distribution (GLC %) 96 97 mm mm 94 25 95 5 mm 94 -10 98 2 mm 95 25 23 77 mm 95 -20 17 83 mm aTaken from Reference 56. 14%, while at -80°C, this value fell to 0 indicating com- plete dominance of the [2,3] rearrangement.57 3111.1 D O 4, MO“ " p“. M Ph I) gag [1.2] product [2.3] product That a [2,3] rearrangement of carbanion iii was not observed at all then was surprising, given that in most cases, [2,3] rearrangement predominated over [1,2] rear- rangement. A concerted pathway in iii_might be precluded by the geometrical requirements of a [2,3] transition 71 state. In this bicyclo[3.3.0]octadienyl system, the "bridged" transition state geometry required resembles an anti-Bredt norbornene. The activation energy needed to iii aggain this conformation could very well exceed that for a dissociation/recombination mechanism. The approximate dis- tance between the bond forming terminii in iii is 2.8 A, as measured on molecular models, quite a bit larger than the normal carbon-carbon single bond length of 1.54 A. To date, there are no literature reports of an anionic [2,3] rearrangement occurring in a cyclopentene ring, although sulfenate 99 rearranges to the sulfoxide,62 albeit slug- gishly when compared with acyclic systems. The different H u 72 bond lengths in this example, as well as the use of the lone pair of sulfur or oxygen, might account for the difference in reactivity. Indeed, the geometrical factor may not be all that important in [2,3] rearrangements of sulfur ylides, as there are several reports where the double bond is re- placed with an acetylene.63 cc": éc/cofls c/CJ'I, It is also interesting to note that the selenium dioxide oxidation of allylic alcohols appears to proceed by a radi- cal dissociation/recombination mechanism in five- and six- membered rings, but through a concerted [2,3] process in 64 Endocyclic acylic and large ring systems (Scheme 10). olefin oxidations in small rings often give poor yields of allylic.alcohols, with concurrent formation of numerous allylic ether and peroxide by-products. These can be accounted for if the allyl radical intermediate traps the various nucleophiles in solution. In addition to the geometrical restraint in iii, the system is so arranged that a [2,3] transition state results in two negative charges being brought into close proximity 73 0* \\ \ ”one“ [2 3] \ Sc + ———9 HOSe ' 9 o” > g /\n H HOSeO ROH \ HOSeOR + HCD O’ 0’ PK) 4* H . "O’Sefl ___> m ———) products A, a: fi, A! Scheme 10 to each other. Whether the incipient charge repulsion is sufficient to drive the reaction exclusively through a [1,2] pathway is unknown. Elimination of this factor may of course be accomplished by attempting the rearrangement in a suitably substituted cyc10pentenol. D o \| BuLi \ 9 Q SnBu3 ’ ' The other products in the rearrangement of carbanion iii to 88 are 88, 88 and 81. The hydroxymethyl ether 88 is 74 \wm II Om I: C) HO) simply the protonation product of anion iii. The formation of compounds 88 and 88 were totally unexpected, however. The presence of diol 81 among the products indicated either a cleavage reaction, A, or a disproportionation, B, was occurring. In view of the ease of destannylation effected A {OW cm: 1 SnBu, m % co 4. csxusnm3 Li- LK) ER 33 ‘OLi i B SL‘W Sn Bu, ”0 8N7 (a .+ E 0.. ,7 $2 SnBu, 8035" 75 by nfbutyllithium, it is diffucult to rationalize the dis- placement depicted in A. In both A and B, the presence of 81 is invoked to account for the formation of diol 88. At the elevated temperatures ‘23- 0°C) required to effect rearrangement, the initially formed carbanion iii must participate in the equilibrium shown below. QCH,Li 1 09 + U0 iii .0 f 7‘] +- BuLi There is ample precedent for carbanion-tetraalkyltin equilib- ria. For example, the reaction of vinyl tributyltin with phenyllithium led to 73% tributylphenyltin and 12% tri- butylvinyltin, while the reaction of vinyllithium with tri- butylphenyltin gave 5% tributylvinyltin and 75% tributyl- phenyltin (Equation 3).65 76 EtZO Bu SnCH = CH2 + PhLi ‘___——_-£ Bu 3 2 SnPh + CH = CHLi 3 2 (3) The equilibrium of stannyl vinyllithium 888 with dilithio- ethylene 888 has also been proposed to account for anomalous Bu,5n H Li H )c =c< Bu Li (:3 >c=c< + Bu,sii «1,99 49% 66 In reactions of the dilithioethylene thus generated. this last example, it is not known whether the production of dilithioethylene is hindered by thermodynamic factors, or by kinetic considerations. In the event of an equilibrium which produced 88, dis- proportionation to 88 and 81 could follow along the lines of equation B. The fate of the other product, bis-stannane 88, would have to account for the butylated product 88 also isolated. The disproportionation of two molecules of 88 to 88 and 88 would liberate two molecules of g-butyllithium. It is already known that big-stannane 88 reacts with one equivalent of E—butyllithium to give tricyclic ether 88; thus it is conceivable that the second‘equivalent was reacting with the tricyclic ether initially formed to give 77 butylated product 88, as previously demonstrated (Scheme 11). QCH,5nBu, Qj . ms. m sm’ 3“” 9 m L! ODIN" \Em v83 75 'VM 8% \n‘“ Scheme 11 Another mechanism would serve to account for the forma- tion of butylated product 88, in which a [2,3] rearrangement occurs followed (or preceded) by an SN2' displacement of lithium oxide with gfbutyllithium (Scheme 12). It is known that allyl alcohol adds alkyllithiums to yield 2- substituted l-propanols 888, however the reaction of 78 .OCHzLi BuL1 , 4.120 LIO ii; [zgn _ Lid CHzoLi 89 'Vb [zan BuLi no} 82 'Vb C‘Hg c4...” 102 Li 103 m 0 OH 104 Wmm SCheme 12 79 cyclopentenol and (under forcing conditions) cyclohexenol with n—butyllithium gives exclusively olefin formation 67 Allylic rearrangement from displacement of lithium oxide. accompanies displacement, as demonstrated in the reaction of 2-cyc10pentenol-l-d lgé (Scheme 12). The difference in regioselectivity for these additions is not well understood, although it may result in part from the greater stability of a primary gs. a secondary carbanion in the case of allyl alcohol. This scenario is less likely, however, given that no re- arrangement products corresponding to diol 82 were isolated, and that 3 would have to rearrange exclusively gig a [2,3] process, as Opposed to the [1,2] observed in iii. This type of displacement reaction introduces the pos- sibility that tricyclic ether Zé is formed by elimination of Lizo, followed by addition of n-butyllithium. Clearly, there is a surfeit of possible mechanisms to rationalize the formation of butylated product §%' Only the disprOportionation scheme, however, accounts for the formation of both 8% and diol 41. 80 In the rearrangement of 2,6-d2 hydroxystannane 9%, products lgé, lgz, lgg, and £1 were isolated. The position Dd“) D ACHJDH «'BUU \\ U, a W .5 Jr of the deuteriums in lgz (Figure 9) unfortunately does not distinguish between the proposed mechanisms for its forma- tion, as all involve allylic rearrangement. The isolation of $81 does, however, confirm the assigned 2,6- regiochem- istry of the substituents. 2.4. The Synthesis of Cis, endo-2,6-bis(hydroxymethyl)- bicyclo[3.3.0]octa-3,7-diene ég In any case it seemed worthwhile to explore the reaction of other nucleophiles with tricyclic ether 1Q, in an effort to utilize this ring Opening reaction to generate diol QQ. Since Seebach had used methanol dianion in some carbonyl alkylations with success, this appeared to be a good 81 ml .AEouuonV www .ooamcm pum.¢ mufl paw Amouv ww poopoum concamvsn mo Aux: ommv Eonuommm mzz ma .m madman ma." . m wusmflm . v . . w 1 q q Id dr- ~r~ 82 \n 83 nucleophile to try. CH,O Li H ‘LacanQ; o L, Dilithiomethanol was generated by the addition of freshly prepared hydroxymethyltributyltin 8%49 to two equivalents n-butyllithium in ether at -78°C. Addition of tricyclic ether lg to this cold, ethereal solution of LiOCHZLi, followed by warming to 0°C resulted in the ap- pearance of a thick, white precipitate. Neutralization, work- up and chromatography gave a single product, which on the basis of its spectral properties was identified as the desired diol 4% ‘OH 09 *LiCHzou; .0 7 /‘ “N" \ H0 «39 In addition to 3%, a small amount of butylation product 8% was again observed; retitration of the n—butyllithium indicated the previously determined molarity was correct. Thus it appeared another tin exchange equilibrium was 84 occurring (Equation 4). If this were the case, the equilib- rium could be shifted to the left by increasing the concen- tration of llg in the reaction. A 2:1 ratio of llg:lgg 0°C LiOCHzLi + Bu4Sn F===3 LiOCHZSnBu3 + BuLi (4) 109 110 mmm mmm completely suppresses the butylation reaction and diol 3% can be isolated in 47% yield by column chromatography. An excess of dilithiomethanol lgg relative to tricyclic ether Z8 was used, and one of the products isolated after a long reaction time appears to be stannane ill. This contaminant can only be separated from the product diol by careful column chromatography. Pure diol éé slowly crystallizes on standing. The relative configuration of the added hydroxymethyl 1H NMR. One substi- substituent was determined from the tuent is necessarily endo from the tricyclic ether start- ing material, and an endo orientation of the incoming 85 nucleophile would result in a compound with C symmetry 2v whereas exo attack results in a compound with C1 symmetry. 13 The C NMR spectrum of diol 39 showed only five signals. In addition the diacetate llg was synthesized from 3%, l and its 250 MHz H NMR showed one sharp singlet for the 13C NMR spectrum of diacetate acetate methyl groups. A llg showed only seven signals. When the crude reaction mixture is worked up with acetic anhydride, a small amount (<5%) of another acetate was observed in the 1H NMR. Although this material might be the isomeric trans-diacetate %%%' the 13C NMR of this mixture is not consistent with OM: OAc / i Z ACO/ ACO ) egg ll3 Mv» this hypothesis, as the few small miscellaneous signals were not close in value to the more intense signals. Fin- ally, H in diacetate llg was observed to be a triplet, l with J=8 Hz. l3C chemical shifts for the bridge- A comparison of head carbons of various bicyclo[3.3.0]octadienes studied in this project is found in Table 5. It is evident that the 3,7-dienes have similar bridgehead carbon chemical cupu Co / H0 .39 :1? 5‘3 1° 0 '2) A00 39 age Table 5. Comparison of 13C Chemical Shifts of Bridgehead Carbons.a Chemical Shift Compound Carbon (ppm) 47 C-1 51.59 24 C-1 51.03 88 C-1 51.04 88 C-5 53.52 46 C-1 51.22 112 C-1 50.52 26 C-8 57.18 26 c-11 40.92 (40.52)b 75 C-7 57.01 75 C-10 50.75 (50.63) al3C spectra obtained at 62.8 MHz in CDC13. bIndefinite assignment. 87 shifts around 51 ppm; on the other hand, the 3,6-dienes exhibit a large difference in chemical shift between the two bridgehead carbons. The doubly allylic C-S appears at 57 ppm, similar to that found in triquinacene (57 ppm). Both structures 1,5 and fig have one olefinic carbon at 138 ppm which is significantly different than the 131-134 ppm range found in the 3,7-dienes. The evidence therefore strongly supports the assignment Of regiochemistry in diol £6. The selectivity in the ring Opening Of tricyclic ether 75 with dilithiomethanol is surprising in view of the sig- nificant sterric bias in 15 towards exo attack. gngg selec- tivity may be due to chelation Of lithium with the ether oxygen similar to that proposed for the nebutyllithium cleavage of 82. A: ___, _ “~Li—cu,0u °"‘ up” EXPERIMENTAL General. Melting points were measured in Open capil- laries with a Thomas-Hoover apparatus and are uncorrected. Proton nuclear magnetic resonance (NMR) spectra were ob- tained on Varian T-60 or Bruker WM-250 spectrometers at 60 MHz and 250 MHz, respectively. Proton homonuclear de- coupling experiments were performed on a Bruker WM-250 spectrometer. Carbon-13 NMR spectra (proton decoupled) were' determined on a Bruker WM-250 spectrometer at 62.8 MHz. Chemical shifts are reported in parts per million downfield from tetramethylsilane internal standard. Infrared spectra (IR) were measured on a Perkin-Elmer 137 spectrometer and a Perkin-Elmer 237 spectrometer, and were calibrated with the polystyrene 1601 cm—1 peak. Mass spectra (MS) were Ob- tained on a Finnigan 4000 instrument with an ionizing voltage Of 70 eV, or when noted, with ionized methane (CI). Gas liquid chromatography (GLC) was performed on an F & M model 700 chromatograph equipped with a thermal conductivity detector. Component ratios were determined by a comparison of peak areas (determined by triangulation) and are uncor— rected for detector response. Columns used for analysis were 6' x 8" aluminum columns packed with the following stationary phases on Chromosorb G (acid and base washed and 88 89 silanized, 60/80 mesh): column A, 5% Carbowax 20M; column B, 4% QF-l; column C, 3% SE-30. Thin layer chromatograms (TLC) were run on Machery-Nagel Polygram SIL G/UV254 pre- coated 0.25 mm silica gel plates, and were developed with either iodine vapor, or an anisaldehyde spray reagent.68 Flash column chromatography69 was accomplished with What- man LPS-2 silica gel (37-53 pm). Normal column chromatog- raphy was performed on EM Reagents Silica Gel 60, 70-230 mesh (EM cat. #7734) rated at activity 2-3. Unless otherwise noted, reagents and solvents were reagent grade materials and were used as received. Ether, THF and benzene were dried by distillation from sodium or potassium benzophenone ketyl. Diisopropylamine and DMF were dried by distillation from calcium hydride. g-Butyllith- ium was Obtained from Aldrich as a 1.6 M solution in hexane, and was titrated according to the method of Watson and Eastham.70 5,11 Bi(2-oxa-3-oxotricyclo[7.2.1.0 ]undeca-6,9-dien-4-yl) 21. Oxidative Dimerization Of Lactoneggé Lithium Enolate. Cupric triflate was prepared according to the procedure of Jenkins and Kochi.24 Diisopropylamine (0.15 mL, 1.07 mmol) in dry THF (5 mL) was cooled to -78°C and B—butyl- lithium (0.83 mL of a 1.23 M solution, 1.02 mmol) added slowly gig syringe. Lactone 2623 (166 mg, 1.02 mmol) in THF (5 mL) was slowly added to the solution of base and 90 the reaction stirred at -78°C for 30 min. Then cupric triflate (369 mg, 1.02 mmol) in dry DMF (5 mL) was added, and the reaction stirred for 8 h at -78°C. The reaction was quenched with 2 N HCl (3 mL), more water was added (5 mL), and the whole extracted with ethyl acetate (3 x 50 mL). The organic layer was washed with water (1 x 10 mL) and saturated NaCl solution (1 x 10 mL). Drying over MgSO4 and removal Of solvent left a brown oil. Crystallization Of lactone dimer 27 was effected by dissolving the Oil in iSOprOpanOl/chloroform and cooling tO 0°C overnight. Dimer 27 was obtained as small colorless cubes (41 mg, 25%), with the following properties: mp 223-224 C; IR 1 1 (nujol) 1727 cm’ ; H NMR (CDCl 60 MHz) 66.18 (2H, dd, 3, J=6 Hz, J=2 Hz), 5.76 (4H, m), 5.36 (4H, m), 3.73 (2H, m), 3.30 (4H, m), 2.93 (2H, m); MS m/e 323 (M+1, 7%), 322 (M+, £3. 2%), 276 (21%), 232 (23%), 161 (93%), 117 (100%), 79 (27%). In runs where cupric bromide or cupric chloride were used, the a-halolactone 32 or 33 was also obtained. Fol- lowing the same procedure as before but using cupric bromide, lactone 26 (51 mg, 0.31 mmol) was transformed to a mixture of products which could be separated by column chromatography on silica gel packed with ether. Recovered lactone 26 (8 mg, 16%) and a-bromolactone 32 (27 mg, 36%) were isolated. The latter product was identi- l fied on the basis Of its spectral prOperties: H NMR (CDCl 60 MHz) 66.23 (1H, dd, J=6 Hz, J=2 Hz), 5.86 (2H, 3! 91 condensed AB q, J=6 Hz), 5.58 (1H, d, J=8 Hz), 5.46 (dd, 1H, J=6 Hz, J=2 Hz), 4.33 (1H, d, J=4 Hz), 04.00-2.93 (3H, m); MS m/e 243 (M+l), 242 (M+), 241 (M+l), 240 (M+), 161, 117, 104. Similarly, the use of cupric chloride converted lactone 26 (125 mg, 0.77 mmol) to a mixture Of products containing a-chlorolactone 33 to the extent Of 14% (NMR yield, from integration of the doublet at 04.40 (J=4 Hz)). The dimer 27 was Obtained by crystallization Of the crude product from chloroform/THF (17 mg, 14%). 2,6-Bis-methylenebicyclo[3.3.0]Octane 6%. Methyl tri- phenylphosphonium bromide (Aldrich, 97 g, 0.27 mol) was dried overnight in a vacuum oven at 80°C, and then trans— ferred to a dry 1 L three necked flask equipped with a mechanical stirrer, reflux condenser, addition funnel, and provision for inert atmosphere. The salt was suspended in dry benzene (500 mL) and potassium E-butoxide (Aldrich, 27.3 g, 0.24 mol) added all at once. A yellow color formed as the slurry became very thick and difficult to stir. Continued stirring at room temperature for 1 h resulted in a bright yellow solution which was quite fluid, and easily stirred. Dione 61 prepared by the method of Haga- 38 (9.35 g, 0.067 mol) in dry benzene dorn and Farnum, (100 mL) was added dropwise, and the well-stirred reaction refluxed under nitrogen for 24 h. After cooling to room 92 temperature, the precipitated solid was gravity filtered through a large funnel and washed with an additional por- tion of benzene (100 mL). The benzene was removed at am- bient pressure through a 30 cm Vigreaux,and the Oily resi- due transferred tO a l L Erlenmeyer flask while hot. Pen- tane (300 mL) was added to the cooled residue, whereupon triphenylphosphine oxide separated as a light tan solid. The lumps were crushed and the flask cooled to 5°C for several hours. The precipitate was filtered through celite packed in a fritted funnel, and washed thoroughly with cold pentane. The solution was concentrated by distilla- tion Of pentane at ambient pressure to a volume Of 93. 150 mL, transferred to a separatory funnel, and washed with water (2 x 30 mL) (to remove unreacted ylide) and saturated NaCl solution (1 x 30 mL). The organic phase was dried over Na2504, and concentrated as before. Repeated crystal- lization Of triphenylphosphine oxide from pentane served to remove most of it from the product, while the last traces were removed by flash chromatography (75 g Whatman silica gel packed with pentane). Diene 62 (7.83 g, 95%) was isolated as a clear, colorless, Oil with a strong Olefin smell. Analysis by GLC showed one component with a retention time of 5.2 min (column C, 90°C, 11 mL/min). Diene 62 had the following properties: bp 55°C (mlO mm aspirator pressure); IR (neat) 3180 (w), 2950, 1675 (m, doublet), 1440, 875 (s) cm'l; 1H NMR (CDC13, 250 MHz) 93 64.83 (4H, AB q, Av=45 Hz, J=l7 Hz), 2.94 (2H, m), 2.30 13 (4H, m), 1.92 (2H, m), 1.63 (2H, m): C NMR (CDCl 62.8 3’ MHz) 157.79, 104.80, 49.10, 33.87, 32.58 ppm; ms m/e 135 (M+l,7%), 134 (M+, 59%), 120 (7%), 106 (31%), 91 (100%), 80 (17%). Cis, endo-2,6-bis(hydroxymethy1)bicyclo[3.3.0]octane 63. Hydroboration Of Diene 62 with Disiamylborane. 2- Methyl-Z-butene (Aldrich) was distilled from lithium alum- inum hydride before use (bp 37.5-38°C). A 500 mL three necked flask fitted with an addition funnel, serum cap, stir bar and provision for inert atmosphere was flame dried under vacuum. A nitrogen atmosphere was established, and BH3-THF solution (Ventron, 105 mL of a l M solution, 0.105 mol) was introduced by a syringe. This solution was cooled in a NaCl-ice bath, and 2-methyl—2-butene (24 mL, 0.22 mol) in dry THF (26 mL) added dropwise over a period Of 30 min. Stirring at ice bath temperature was continued for 2.5 h, whereupon diene 62 (6.19 g, 46 mmol) in THF (50 mL) was slowly added. The reaction was stirred at 0°C for l h, then allowed tO warm to room temperature overnight. The reaction was cooled to 0°C, and carefully quenched with 1:1 THF:H O (30 mL). This was followed by the addition of 2 aqueous NaOH (35 mL Of a 3 M solution) and the slow addi- tion of H202 (34 mL Of a 30% solution), keeping the tem- perature below 50°C with an ice bath. After the addition 94 was complete, the reaction was warmed to 55°C and stirred for l h. Upon cooling, the layers were separated, and the aqueous phase extracted with chloroform (1 x 100 mL). This was combined with the THF portion, and dried over MgSO4. The solution was filtered, and the solvent removed to give an Oil which was subjected to high vacuum for 1 day to remove the amyl alcohol by-product. DiOl 63 was Ob- tained as a somewhat cloudy Oil (8.19 g, 104%). Kugel- rohr distillation (110°C, 0.05 mm) gave an analytical sam- ple of the clear, colorless Oil, which showed two peaks, in a relative ratio Of 1:123 (99.1% one isomer) when analyzed by GLC (column B, 155°C, 60 mL/min). DiOl 63 had the fol- lowing properties: IR (neat) 3325 (br), 2950, 2875, 1660 (w), 1475, 1065 (s) cm‘l; 1H NMR (CDC13/D20, 250 MHz) 03.68 (2H, dd, part of an AMX system, J =11 Hz, J AM AX: 7 Hz), 3.61 (2H, dd, part Of an AMX system, JAM=11 Hz, JMX=7 Hz), 2.60 (2H, sextet, J=6 Hz), 2.19 (2H, d of q, =20 Hz, Jq=7 Hz), 1.72 (2H, septet, J=6 Hz), 1.53 (4H, 13 Jd m), 1.18 (2H, d of t, Jd=18 Hz, Jt=8 Hz); C NMR (CDCl 3! 62.8 MHz) 64.51, 45.99, 45.75, 29.17, 24.34 ppm; MS m/e 171 (n+1, 1%), 151 (2%), 134 (13%), 121 (95%), 106 (20%), 93 (100%), 79 (84%). Hydroboration of Diene 62 with BH3-THF Complex. The following procedure is representative of the general method used in the hydroboration Of diene 6% with BH3-THF. A 95 three necked round bottom flask equipped with a magnetic stir bar, addition funnel, serum cap and provision for inert atmosphere was flame dried under vacuum. A nitrogen atmosphere was established, and BH -THF (7.60 mL Of a l M 3 solution, 7.60 mmol) was introduced gig syringe. The solu- tion was cooled to either -23°C or -45°C, and diene 62 (462 mg, 3.45 mmol) in dry THF (15 mL) was slowly added dropwise to the stirred solution. The reaction was crystal— clear initially, but eventually the borane adduct separat- ed as a gelatinous slurry. Stirring was continued for 3 h at the indicated temperature, and then the reaction warmed to 0°C and carefully quenched with 1:1 ethanol:water (10 mL). Stirring was continued at room temperature for 2 h to insure complete hydrolysis and then NaOH (3.5 mL Of a 3 N solution) and H (0.75 mL Of a 30% solution) were 202 added. The reaction was heated to 50-55°C for 1 h and then cooled to room temperature. The mixture was extracted with ethyl acetate, the organic phase dried over MgSO4, and the solvent removed to yield a clear Oil (609 mg, 100%). The crude product was analyzed by GLC (column B, 180°C, 60 mL/min), and exhibited two barely resolved peaks at 6.4 min and 7.0 min. The relative areas Of these two com- ponents were determined to be 1:12, respectively, at ~45°C, and 1:6 at -23-0°C. The spectral properties of this material were identical to those Obtained for the product of disiamylborane hydroboration/oxidation Of diene 62, 96 with the exception of shoulders on the doublet at 03.58 1 in the H NMR (60 MHZ). Cis, endo-2,6-bis(hydroxymethyl)bicyclo[3.3.0]octane Dimethanesulfonate 64. Diol 6% (1.7 g, 0.01 mol) was dis- solved in dry pyridine (15 mL) in a 25 mL round bottom flask, and cooled to 0°C. Methanesulfonyl chloride (7.7 mL, 0.10 mol, previously distilled (bp 54.5-55°C/m10 mm)), was added to the stirred solution. The reaction was stirred for 2 h, and then poured onto ice. Extraction with chloro- form (2 x 20 mL) was followed by multiple extractions Of the combined organic phases with cold 5% NaOH (3 x 20 mL), cold 1 M HC1 (4 x 20 mL), saturated NaHCO3 (1 x 20 mL), and saturated NaCl solution (1 x 20 mL). The extract was dried over MgSO4 and the solvent removed to afford di- mesylate 64 (2.96 g, 91%) as a brown Oil which solidified on standing. Similar runs Of this reaction gave yields ranging from 90-96%. A 1H NMR of this material (CDC13, 60 MHz) was as follows: 64.16 (4H, d, J=7 Hz), 2.96 (s, 6H), 2.76—0.96 (12H, m). This material was not characterized or purified further, but was used directly in the next step. Cis,endo-2,6—bis(iodomethyl)bicyclo[3.3.0]octane, 65. TO a solution Of crude dimesylate 6% (2.96 g, 9.0 mmol) in acetone (40 mL) was added sodium iodide (4.08 g, 27 mmol). 97 The reaction was refluxed with stirring for 12 h, after which time an additional portion of sodium iodide (1 g) was added and refluxing continued for an additional 6 h. The reaction was cooled, filtered free of precipitated salt, and the solvent removed under reduced pressure. The brown residue was taken up in ether and washed with satu- rated sodium sulfite (2x) and saturated NaCl solution (1x). The organic phase was dried and the solvent removed to give crude diiodide 65 as a brown oil. Flash chromatography on silica gel packed with hexane yielded diiodide 65 (1.94 g, 55%) as a clear, colorless Oil, which could be kugel- rohr distilled (100°C (0.05 mm)) only when very pure (g_.__g_i after chromatography). Attempted distillation Of the crude product resulted in a diminished yield of 65, Obtained as a dark purple Oil. The spectral characteristics were as follows: IR (neat) 2950, 2899, 1458 (triplet), 1282, 1182 1 1 cm- ; H NMR (CDCl 60 MHz) 63.18 (2H, dd, part of an 3' AMX system, J=10 Hz, J=8 Hz), 3.03 (2H, dd, part Of an AMX system, J=10 Hz, J=8 Hz), 2.83-0.93 (12H, two m centered at 2.46 and 1.56); MS m/e 391 (M+1, 33. 0.02%), 390 (M+, 0.8%), 263 (7%), 135 (100%), 107 (9%). Cis,endo-2,6-bis(bromomethyl)bicyclo[3.3.0]octane 62. Reaction Of Diol 6% with Carbon Tetrabromide/Triphenylphos- phine. Diol 63 (2.01 g, 11.8 mmol) was dissolved in aceto— nitrile (200 mL) in a 500 mL three necked flask equipped 98 with either a good magnetic stir bar or a mechanical stirrer, thermometer and stOpper. Carbon tetrabromide (16.1 g, 48.6 mmol) was added, and the solution stirred until it was all in solution. Triphenylphosphine (12.8 g, 48.9 mmol) was then added all at once, causing the reaction to warm up to 40°C. The solution immediately became a dark red, and began to turn brown as triphenylphosphine oxide precipi- tated. After stirring at room temperature for 2 h, TLC analysis (silica gel, ether) showed only one component. One half of the solvent was removed under reduced pressure (beware Of bumping!) and the remaining solvent and pre- cipitate poured into an equal volume Of ether. The solution was filtered through celite packed in a fritted funnel, and the filtered precipitate washed thoroughly with cold ether. The solvent was removed under reduced pressure, and more triphenylphosphine oxide crystallized from cold 1:1 ether:petroleum ether. The Oil finally Obtained was placed under high vacuum overnight to remove bromoform (bp 150°C). The remaining product was flash chromatographed on silica gel (Whatman) packed with hexane to remove contaminating carbon tetrabromide. The dibromide (2.12 g, 61%) was isolated as a clear, colorless Oil with the following spectral prOperties: IR (neat) 2950, 2875, 1440, 1275, 1225 cm-l; 1H NMR (CDC13, 250 MHz) 63.44 (2H, dd, part of an AMX system, JAM=12 Hz, =8 Hz), 3.37 (2H, dd, part JAX Of an AMX system, J =12 Hz, =9 Hz), 2.68 (2H, sextet, AM JMX J=6 Hz), 2.43 (2H, d of q, Jd=20 Hz, Jq=7 Hz), 1.84 (2H, 99 =4 Hz), 1.65 (2H, m), 1.52 (2H, m), 13 q Of d, Jq=7 Hz, Jd 1.22 (2H, t of t, J=12 Hz, J=8 Hz); C NMR (CDCl 62.8 3! MHz) 47.04, 46.57, 35.05, 31.11, 24.05 ppm; MS m/e 297 (0.04%), 296 (0.01%), 295 (0.10%), 293 (0.05%), 215 (11%), 135 (100%). Cis,endo-2,6-bis[(amidinothio)methyl]bicyc10[3.3.0]- Octane Dibromide 71. The Addition Of Thiourea to Dibromide 69. NB: The success of this reaction depends on the purity Of the dibromide; substantial amounts of other bromocarbons (gig; carbon tetrabromide) will seriously interfere-with the calculated stoichiometry. Pure dibromide 62 (890 mg, 3.00 mmol) and thiourea (Aldrich, 685 mg, 9 mmol, 50% excess) were placed in a screw-capped reaction tube, and 95% ethanol added (10 mL, iigi, to make the solution 0.3 M in dibromide). The tube was tightly capped and heated on a steam bath for 18 h. During this time, the product crystallized from the reaction. The solution was allowed to cool to room temperature, and then chilled to -15°C for 4 h. The resultant solid was filtered, washed with cold ethanol, and air dried to give a white solid (1.46 g, 108%), mp l62-164°C. This material was not purified further, although it Obviously contained some of the excess thiourea. 100 Cis,endo-2,6-bis(mercaptomethyl)bicyclo[3.3.0joctane 66. Hydrolysis Of Bis-thiouronium Bromide 62. The thio- uronium salt from above (1.46 g m3 mmol) was placed in a 100 mL round bottom flask, and suspended, with stirring, in aqueous NaOH (40 mL of a 1.5 N solution). A serum cap was fitted to the flask and the suspension stirred as nitrogen was bubbled through the solution with a syringe needle (another needle served as a vent). After 5 min, a nitrogen purged condenser was fitted to the flask, and the whole system kept under nitrogen for the duration Of the reaction. The solution was refluxed for 12 h, where— upon the suspension dissolved. After cooling tO 0°C, the reaction was acidified to pH 1-2 with cold HC1 (4N) in the hood. The mixture was transferred to a separatory funnel, and extracted with ether (2 x 75 mL). The organic phase was washed with water (2 x 20 mL) and saturated NaCl aqueous solution (1 x 20 mL), and dried over NaZSO4. Removal of solvent (stench!) left a yellow Oil (710 mg). This was flash chromatographed on silica gel (Whatman) with 5% ether in hexane. Dithiol 66 was obtained as a clear, colorless (but not odorless) oil (585 mg, 96% from dibromide 66), and was best stored in the dark under nitrogen to discourage formation of disulfides. Dithiol 66 had the following spectral properties: IR (neat) 2924, 2857, 2654 (w), 1449 (triplet), 1250, 758 cm—1; 1H NMR (CDC1 250 MHz) 62.57 (2H, m), 2.55 (4H, t, J=8 Hz), 3' 101 2.12 (2H, m), 1.79 (2H, d of pentet, Jp=7 Hz, Jd=5 Hz), 1.64-1.32 (4H, m), 1.36 (2H, t, J=8 Hz, -sg), 1.13 (2H, 13 tt, J=11 HZ, J=9 HZ); C NMR (CDC1 62.8 MHZ) 47.93, 46.51, 3! 31.52, 26.40, 24.05 ppm; MS m/e 203 (n+1, 8%), 202 (M+, 62%), 168 (37%), 135 (40%), 121 (82%), 93 (100%), 79 (88%). The Reaction Of Dithiol 66 with Dibromide 66; A l L three necked flask fitted with an addition funnel, ther— mometer, magnetic stir bar and provision for inert atmos- phere was flame dried under vacuum, and a nitrogen atmos- phere established. Cesium carbonate (1.83 g, 5.64 mmol, Alpha) was suspended in dry DMF (320 mL), and the tempera- ture brought to 95°C. Dithiol 66 (570 mg, 2.82 mmol) mixed with dibromide 62 (834 mg, 2.82 mmol) in DMF (70 mL) was slowly added over a period Of 3 days. The reaction was cooled to room temperature, and excess cesium carbonate re- moved by filtration. DMF was distilled at 25°C (0.10 mm). The residue was flash chromatographed on silica gel (75 g) packed with 20% methylene chloride in hexane. Only one component eluted in this solvent system (10 mg, ml%). Increasing the polarity of the solvent with 40% methylene chloride eluted another component (30 mg, %3%). Elution with straight methylene chloride gave a third fraction (280 mg, 30%). Further elution of the column led to the isola- tion Of small amounts of polymer. 1H NMR spectra Obtained on these materials did not unambiguously identify them. 102 The smallest fraction is tentatively identified as disul- fide 1%: MS m/e (CI) 241 (2.5%), 229 (5%), 201 (86%), 200 (8%), 167 (25%), 135 (100%). The 3% fraction showed MS m/e 537 (2%), 536 (9%), 371 (3%), 167 (100%). The largest fraction was insufficiently volatile tO be analyzed by MS. Cerous Chloride Mediated Sodium Borohydride Reduction Of Bicyclo[3.3.0JOcta-3,7-diene-2,6-dione 1%. Cis,endo-2,6- dihydroxybicycloE3.3.0]octa-3,7-diene 61. A three necked 1 L round bottom flask was equipped with a mechanical stirrer, thermometer and drying tube. Dienedione 16 (2.0 g, 0.15 mol) (prepared by the method Of Hagadorn and Farnum)38 in absolute methanol (600 mL) was added, and stirring com— menced. Cerous chloride hexahydrate (Fluka, 116 g, 0.33 mol) was added, and the solution cooled to -20°C with a dry ice-acetone bath (below m-25°C, the cerous chloride begins tO precipitate from solution). Sodium borohydride (22.6 g, 0.60 mol, 100% excess) was added slowly portionwise, allow- ing the foam to subside between additions and keeping the temperature below -10°C by the judicious addition of dry ice tO the cold acetone bath. The addition was com- plete after 1 h, and the acetone bath was replaced with a NaCl-ice bath. Stirring was continued at -15°C for 2 h, at which time TLC analysis (20% ethyl acetate in ether) showed the reaction was complete. Warming tO 0°C, water 103 (300 mL) was added and the reaction stirred for 15 min. Methanol and water were removed under reduced pressure, the last traces on a vacuum pump with a large capacity cold trap. The solid was transferred to a 2 L erlenmeyer flask, crushed and extracted with boiling chloroform (1 x 1 L, 2 x 500 mL). After each extraction, the solution was cooled and filtered through an 11 cm Buchner funnel. The combined filtrates were passed through a Drierite cone and the chloroform distilled under reduced pressure to afford the isomeric dienediols 61 as a waxy, white solid (22 g, 106% due to residual chloroform). Recrystallization of the crude product from acetone at -15°C, followed by a second re- crystallization Of the mother liquors at -30°C yielded gig, ggggfdiol 61 as hard, white granules (10.9 g, 53%), mp 92-93°C. The Spectral properties Of this material were identical to those reported earlier. Temperature Effects in the Sodium Borohydride/Cerous Chloride Reduction of DienedioneiZQ. The reduction was carried out as previously described on a l-lO 9 scale at one Of the following temperatures or temperature range: 0°C to 22°C; -10°C; -15°C. The crude product from each reaction was analyzed by GLC (column A, 190°C, 60 mL/min), and the relative ratio of cis:trans isomers determined (retention times of 10.1 and 12.2 min, respectively). At -15°C, this ratio was 8:1, at -10°C, 7.6:1, and at 0°C-22°C, 6:1. 104 Iodomethyltri-g-butyltin. The title compound was pre- pared from methylene iodide and iodomethylzinc iodide fol- 71 9 lowing the procedures of Seyferth and Andrews, and Still.4 The reaction was run on 0.02 to 0.80 molar scales, with iso— lated yields ranging from 78-85% after distillation (bp 118-119°C/0.65 mm), (literature lOO-llO°C/0.01 mm). TLC (hexane Rf 0.51; IR (neat) 2865, 1456 cm'l; NMR (c0c13, 60 MHz) overlapping absorptions at 61.90 (2H, s, flanked by two isotope satellite peaks, J=18 Hz, ICHZSn), 1.55 (m), 0.93 (m). Cis,endo-2,6-bis[(tri-g—butylstannyl)methoxy]bicyclo— [3.3.0jocta-3,7-diene 66. Potassium hydride (6.42 g of a 35% mineral Oil dispersion, 0.056 mol) was weighed out in a dry 500 mL three-necked flask. A mechanical stirrer, addition funnel,nitrogen inlet adapter and rubber serum cap were fitted to the flask, and a nitrogen atmosphere established. The potassium hydride was washed free of min- eral Oil with dry ether (3 x 20 mL), and then suspended in dry THF (100 mL). Dienediol 67 (2.58 g, 0.019 mol) in dry THF (100 mL) was added dropwise to the stirred suspension at room temperature. When gas evolution ceased (gg. 2 h), iodomethyltri-g-butyltin (17.7 g, 0.041 mol) in dry THF (75 mL) was added dropwise, and the reaction stirred for 24 h at room temperature. TLC analysis showed the reaction to be complete (20% ethyl acetate/hexanes) Rf 66, 0.68; 105 Rf 61, 0.30. Methanol (10 mL) was then added slowly to destroy excess hydride. The dark brown solution was added to an equal volume of petroleum ether and extracted with water (2 x 100 mL), and saturated aqueous sodium chloride (1 x 25 mL). The organic phase was dried over sodium sul- fate and the solvent evaporated to leave a light brown liquid (19.3 g). The crude product was applied directly to the top of a 6 cm diameter column containing silica gel (250 g) packed with hexane. When residual nonpolar com- pounds were no longer detected in the eluate (TLC, 1% ethyl acetate/hexane, Rf 0.67) the solvent was changed to 20% ethyl acetate/hexane whereupon the title compound eluted, and was isolated as a very pale yellow liquid (12.8 g, 96%) with the following properties: TLC (1% ethyl acetate/hexane) Rf 0.17; IR (neat) 3040, 2857, 1613, 1 1453 cm" ; NMR (CDC1 60 MHz) 65.70 (4H, condensed AB q, 3! J=7 Hz), 4.26 (2H, m), 3.75 (4H, 5, two isotope satellite peaks, J=18 Hz), 3.33 (2H, m), 2.16-0.53 (54H, two mul- tiplets, including a broad distorted triplet at 0.88). 2-Oxa-tricyclo[5.2.l.0.4’lo]nona-5,8-diene 16. Reaction of 66 with Two Eguivalents Of p-Butyllithium. A 100 mL three necked flask fitted with a serum cap, nitrogen inlet valve, magnetic stir bar and addition funnel was flame dried under vacuum. A nitrogen atmosphere was es- tablished, and g—butyllithium (6.6 mL of a 1.1 M solution in 106 hexane, 7.3 mmol) was transferred to the flask and cooled to -78°C. Dry THF (14 mL) was added slowly to the cold g-butyllithium gig syringe, followed by big-stannylmethyl ether 66 (2.5 g, 3.3 mmol) in THF (25 mL). The reaction was stirred at -78°C for 0.5 h, and then allowed to warm to -30°C. The nitrogen inlet valve was closed, and the reaction stored in a freezer at -30°C for 36 h. At the end Of this time, the reaction solution had become quite a deep reddish-orange. A nitrogen line was reattached to the ap— paratus and the reaction mixture allowed tO stand at room temperature for several hours. Water (6 mL) was added and the whole solution transferred to a separatory funnel. The organic phase was washed with saturated sodium chloride (1 x 20 mL) and then dried over MgSO4. The dry organic extracts were filtered into a 250 mL round bottom flask containing a magnetic stir bar, and the solvents removed at atmOSpheric pressure through a 28 cm Vigreaux. The red- orange residue left from the distillation was applied to the top of a 4 cm diameter column containing silica gel (65 g) packed in 20% ether/pentane. Fractions averaging 20 mL were collected, and the chromatography followed by TLC. Tetra-g—butyltin eluted first, followed by tricyclic ether 76. The fractions containing 66 were combined, and the bulk Of the solvent removed at atmospheric pressure followed by brief evacuation (aspirator pressure) to remove the last traces Of solvent. In this way, Z6 was Obtained 107 as a clear, colorless, volatile Oil. If further purifica- tion is desired, 66 can be distilled through a short path distillation unit (bp 80°C/10 mm) into a cold receiver. TLC (10% ether/pentane) Rf 0.20; IR (neat) 3067, 2941, 1 1618, 1076 cm” ; NMR (CDC1 250 MHz) 65.98 (1H, dd, J=6 3' Hz, J=2 Hz), 5.75 (1H,dt, J =6 Hz, Jt=2 Hz), 5.68 (1H, dt, d Jd=6 Hz, Jt=2 Hz), 5.50 (1H, dt, J=6 Hz, J=2 Hz), 4.92 (1H, dd, J=8 HZ, J=2 HZ), 3.93 (1H, dd, J=10 HZ, J=7 HZ), 3.68 (1H, dd, J=10 HZ, J=4 HZ), 3.64 (1H, m), 3.49 (1H, 13C q, J=8 Hz, with some fine structure), 3.26 (1H, m); NMR (CDC1 62.8 MHz) 138.76, 134.41, 131.29, 129.59, 87.95, 3, 74.95, 57.01, 50.75, 50.63 ppm; MS m/e 135 (M+1, 0.3%), 134 (M+, 4%), 104 (100%), 78 (17%). Tricyclic ether Z6 should be stored under nitrogen, and preferably kept cold, as it appears tO be rather sensitive to oxygen and susceptible to polymerization over a period of time. Low Temperature Quenching Of Dianion i_Derived from Bis-stannane 66 and g-Butyllithium. Cis,endo-2,6-dimethoxy— bicyclo[3.3.0]0cta-3,7-diene Z6. Big-stannane 66 (267 mg, 0.36 mmol) in either dry ether or dry THF (2 mL) was slowly added to a solution Of g-butyllithium (0.92 mL of a 1.11 M solution) in either dry ether or THF at -78°C. The reaction was stirred for 15 min at -78°C, and then quenched with 0.5 M HC1 (4 mL). The organic phase was separated from 108 the aqueous phase and washed with water, and then dried over MgSO4. The 60 MHz 1H NMR showed only one product for both solvents. The crude product from the reaction done in ether was analyzed by GLC (column A, 140°C, 60 mL/min) and consisted of 96% dimethyl ether 74 and 4% tricyclic ether Zé- 1 3, Hz), 4.46 (2H, br m), 3.43 (8H, a sharp s superimposed on H NMR (CDC1 60 MHz) 5.80 (4H, condensed AB q, J=6 a m). 2-Oxa-tricyclo[5.2.l.04'101nona-5,8-diene, 16. Reaction Of 66 with One Equivalent of g—Butyllithium. A 500 mL three necked flask equipped with a magnetic stir bar, nitrogen inlet adapter and a rubber serum cap was flame dried under vacuum, and a nitrogen atmosphere established. gig-stan- nylmethyl ether 66 (11.5 g, 15.5 mmol) in dry ether (175 mL) was introduced gig cannula to the flask. The solution was cooled to -78°C, and g-butyllithium (16.3 mL of a 0.95 M solution in hexane, 15.5 mmol) added slowly gig syringe. After stirring for 0.5 h at -78°C, an aliquot from the re- action showed the presence Of at least six components (TLC 20% ethyl acetate/hexanes Rf = 0.85, 0.68, 0.51, 0.33, 0.28, 0.17). These correspond in order of migratory ap- petitude on TLC, to tetra-g—butyltin, gig-stannylmethyl ether 66, methoxy stannylmethyl ether 66, hydroxymethyl- tri-g—butyltin 66, tricyclic ether 66 and dimethyl ether 66. By allowing the reaction to warm slowly to 0°C, and 109 stirring at 0°C for 5 h, followed by warming to room tem- perature overnight, TLC analysis showed only tricyclic ether 66, hydroxymethyltri-g-butyltin, and tetra-g—butyltin. Apparently, the various bicyclooctadiene compounds fun- neled off to the tricyclic ether. Workup as before gave pure Z6 (1.44 g, 72%). Cis,endo-Z-hydroxy-6-[(tri-g—butylotannyl)methoxy]- bicyclo[3.3.0]Octa-3,7-diene, 66. This reaction was run under conditions identical to those used for the prepara- tion of gig-stannylmethyl ether 66, save for the fact that only one equivalent of iodomethyltri-g-butyltin was used. Thus diene diol 66 (664 mg, 4.81 mmol) in dry THF (5 mL) was added dropwise to a stirred suspension Of potassium hydride (700 mg of a 35% Oil dispersion, 5.77 mmol) in THF; when gas evolution ceased, iodomethyltri-g—butyltin (2.06 g, 4.81 mmol) in THF (5 mL) was added drOpwise. After stirring at room temperature for 2.5 h the reaction was worked up as before to give a viscous brown oil (2.16 g). The products were separated by flash chromatography on 150 g silica gel packed in 20% ether/hexanes. gigfstannyl- methyl ether 16 eluted first (TLC 20% hexane Rf 0.68), (1.11 g, 31%), followed by the title compound 66 (TLC 20% ethyl acetate/hexane Rf 0.30) (0.50 g, 24%), which exhibited 110 the following properties: IR (neat) 3344 (br), 3077, 2907, l 1634, 1471, 1070 cm’ ; NMR (CDC1 60 MHz) 65.96-5.53 (4H, 3, s at 5.83 overlapping an AB q at 5.53, J=6 Hz), 4.71 (1H, br m), 4.21 (1H, m), 3.75 (2H, t, J=7 Hz), 3.33 (1H, m), 1.85 (1H, s, OH), 1.73-0.67 (12H, m at 1.33, distorted t at 0.90, J=6 Hz). Further elution of the column with 40% ethyl acetate in hexane yielded recovered dienediol 47 (150 mg, 23%). Thus the total yield of chromatographed material represents a 78% mass balance. n-Butyllithium Exchange with Cis,endo-Z-hydroxy-6—(tri-i n-butylstannyl)methoxybicyclo[3.3.0]octa-3,7-diene 87. Cis,endo-2-hydroxy-6-hydroxymethy1bicyclo[3.3.0]octa-3,7- diene 88, and 2-Endo—hydroxymethyl-6-butylbicyclo[3.3.0]- octa-3,7-diene 8%. A 250 mL three necked flask equipped with a magnetic stir bar, nitrogen inlet and a rubber serum cap was flame dried under vacuum and a nitrogen atmosphere established. Dry ether (100 mL) was transferred to the flask via cannula, and cooled to -78°C. n—Butyllithium (14.3 mL of a 0.95 M solution in hexane, 13.6 mmol) was slowly syringed into the cold ether, followed by (stannyl)- methoxy alcohol 87 (2.93 g, 6.65 mmol) in ether (20 mL). Stirring was continued at -78°C for 0.5 h; the reaction mixture then gradually warmed up to 0°C whereupon stirring was discontinued. Within one hour of standing at 0°C, a heavy white precipitate appeared. The reaction was 111 allowed to stand at 0°C for 6 h, and then at room tempera- ture overnight (12 h). Water (10 mL) was added and the layers separated. The aqueous phase was back-extracted with chloroform (l x 50 mL). The organic extracts were com- bined and washed with saturated, aqueous sodium chloride, and then dried over sodium sulfate. Removal of the solvent left an orange-red oil (3.49 9). Gas chromatographic analysis of the crude material (column A, 190°C) showed the presence of at least four major products (besides tetra-n-butyltin), two of which could not be completely resolved, in a 2:1:l:3.5 relative ratio. The entire crude product was taken up in a minimum of ether and chromato— graphed on silica gel (100 g) packed with 1% methanol/ ether. Following elution of tetra-n-butyltin, the solvent was changed to 25% ethylacetate in ether containing 1% methanol. Butylated product 82 was isolated as an oil (150 mg, 12%), and had the following properties: TLC (20% ethylacetate/hexane) Rf 0.50; IR (neat) 3356 (br), 3030, l l 2890, 1453, 1370, 1068 cm“ ; H NMR (0001 250 MHz) 05.71 3: (1H, d J=6 Hz), 5.62 (2H, AB q, J=6 Hz), 5.53 (1H, d of t, J=6Hz, J=2 Hz), 3.72 (1H, dd (AMX), J=10 Hz, J=6 Hz), 3.68 (1H, dd (AMX), J=10 Hz, J=7 Hz), 3.45 (2H, m), 3.01 (1H, m), 2.74 (1H, m), 1.66-0.80 (10H, m centered at 01.35, and a broad t at 0.89); ms m/e 192 (M+, 15%), 161 (64%), 135 (6%), 117 (85%), 105 (66%), 91 (83%), 41 (100%). The next product to elute from the column was methoxy 112 alcohol 9% (70 mg, 7%), easily identified by the sharp singlet at 03.4 in its 1H NMR spectrum: TLC (20% ethyl 1 acetate/hexane) Rf 0.51, H NMR (CDC1 60 MHz) 05.76 (4H, 3’ m), 4.76 (1H, br m), 4.35 (1H, br m), 3.20-3.80 (5H, sharp singlet at 03.40 superimposed on a multiplet), 2.68 (1H, S, OH). Biol 88 next eluted from the column, and was obtained as a clear, colorless, viscous oil (270 mg, 40%). TLC (20% ethyl acetate/ether) Rf 0.27; IR (neat) 3280 (br), 3021, 2874, 1639, 1020 cm’l; 1H NMR (c001 13 3, 250 MHz) (See Table 2, page 61); C NMR (CDC1 62.8 MHz) 134.14, 3! 133.62, 132.68, 130.38, 77.98, 63.01, 53.52, 51.04, 49.22 ppm; MS m/e 153 (M+l, 0.7%), 152 (M+, 7%), 134 (12%), 121 (49%), 104 (100%), 91 (58%), 78 (59%). Finally, as the last product to elute from the column, diol 47 was isolated as an oil which solidified on stand- ing. Its 1H NMR spectrum left no doubt as to its identity. TLC (20% ethyl acetate/ether) R 0.12. f n—Butyllithium Cleavage of Tricyclic Ether 15. Endo- 2-(hydroxymethyl)-6-n—buty1bicyclo[3.3.0]octa-3,7-diene 8%. Tricyclic ether 75 (61 mg, 0.45 mmol) in dry ether (5 mL) was cooled to 0°C and n-butyllithium (1.0 mL of a 0.95 M solution) added via syringe. After the reaction was stirred for l h at 0°C, an aliquot was quenched and analyzed by GLC (column A, 135°C-l900C, 60 mL/min). A EE- 1:1 ratio of 113 15 to product was observed. The product had a retention time identical to that of butylated material 82 isolated earlier. The relative ratio of 75:82 did not change over the course of 2 h, so additional n—butyllithium was added (2 mL). The reaction was stirred at 0°C for 1 h and a1- 1owed to warm to room temperature overnight, after which time GLC analysis of an aliquot indicated the reaction was complete. Careful quenching with methanol to destroy excess alkylithium, extraction with water and saturated aqueous NaCl was followed by dryingtflmaorganic phase over MgSO4. Removal of solvent at reduced pressure left a yellow oil (70 mg, 80%) which was identical in all respects to 82 isolated earlier. Sodium Borodeuteride/Cerous Chloride Reduction of Bi- cyclo[3.3.0]octa-3,7-diene-2,6-dione 73. Cis,endo-2,6- dihydroxybicyclo[3.3.0]octa-3,7—diene-2,6-d_2 77. The reduction was carried out as described previously, using sodium borohydride. Dienedione 7% (1.75 g, 13 mmol) and cerric chloride hexahydrate (10.1 g, 28 mmol) in methanol— d (35 mL) was stirred at -15°C, and sodium borodeuteride (2.2 g, 52 mmol, Merck Sharp & Dohme) added portionwise. Workup of the reaction followed the procedure previously outlined. The title compound was obtained as a mixture of stereoisomers (1.90 g, 100%). Recrystallization of the crude material from acetone afforded labeled Cis,endo diol 114 1 77 (632 mg, 35%). H NMR (c0013, 60 MHz) 04.10 (4H, AB q, 40:14 Hz, J=6 Hz), 3.36 (2H, br s), 2.83 (2H, br s, 03)? MS m/e 141 (M+1, 0.5%), 140 (M+, 4%), 122 (99%), 108 (3%), 92 (100%). Alkylation of 2,6-d Diol 77 with One Equivalent of Iodo- 2 methyltri-n—butyltin. Diol 77 (604 mg, 4.31 mmol) was alkylated with iodomethyltri-n—butyltin according to the procedure described earlier for the protio analog, g7. Labeled bis-stannane 78 was isolated as a clear liquid (620 mg, 19%) which was identical to 88 with the following 1 1 exceptions: IR (neat) 2079 cm- ; H NMR (CDC1 60 MHz) 3, 5.68 (4H, AB q, J=6 Hz, 00:10 Hz), 3.75 (4H, s, with two isotope satellite bands, J=8 Hz), 3.35 (2H, s), 2.15-0.60 (54H, 2 m centered at 01.38 and 00.90). Labeled hydroxystannane 87 was obtained as a clear, viscous oil (660 mg, 35%) identical to 87 with the follow- ing exceptions: IR (neat) 2151 cm-1; 1H NMR (CDC13, 60 MHz) 05.93-5.53 (4H, s at 05.83 overlapping an AB q cen- tered at 05.40, J=6 Hz), 3.70 (2H, s, two isotope satel- lite peaks, J=8 Hz), 3.33 (2H, br s), 1.75-0.80 (10H, 2 m centered at 01.36 and 00.90); MS (CI) m/e 443 (0.2%), 442 (0.7%), 387 (2%), 291 (100%), 251 (38%). 115 2—Oxa-tricyclo[5.2.l.04’lOJdeca-S,8-diene-1,6-g2. Bis-stannane 78 (505 mg, 0.67 mmol) was destannylated with two equivalents of n—butyllithium in THF following the pro- cedure outlined for the protio analog 88. Labeled tri- cyclic ether 78 was obtained as an oil (45 mg, 50%) which 1 had the following spectral properties: H NMR (CDC1 3! 250 MHz) 05.96 (1H, dd, J=6 Hz, J=2 Hz), 5.67 (1H, dd, J=6 Hz, J=1 Hz), 5.49 (1H, br s), 3.93 (1H, dd, J=9 Hz, J=7 Hz), 3.67 (1H, dd, J=9 Hz, J=4 Hz), 3.63 (1H, m), 3.49 (1H, t, J=8 Hz), 3.24 (1H, m). gi§,endo-2-hydroxy-6-(hydroxymethyl)bicyclo[3.3.0]octa- 3,7-diene-2,6-d2 88_and Endo-Z-(hydroxymethyl)-6-n—buty1-bi cyclof3.3.0]octa-3,7-diene-4,8-g 107. Labeled hydroxy- 2—’L"u’\a stannane 87 (456 mg, 1.03 mmol) was destannylated with n- butyllithium following the procedure outlined for the protio compound, 87. Labeled diol 706 was isolated as a clear, mm mm 1 colorless oil. H NMR (CDC13/D20, 250 MHz) 05.96 (1H, dd, J=6 Hz, J=2 Hz), 5.82 (1H, dd, J=6 Hz, J=2 Hz), 5.76 (1H, dd, J=6 Hz, J=2 Hz), 5.65 (1H, dd, J=6 Hz, J=2 Hz), 3.73 (1H, d, J=12 Hz), 3.64 (1H, d, J=12 Hz), 3.58 (1H, br d, J=7 Hz), 3.42 (1H, br d, J=7 Hz). Labeled butylated product 787 was obtained as an oil. 1H NMR (CDC1 250 MHz) 05.62 (1H, br s), 5.54 (1H, br s), 3! 3.72 (2H, ABX; HA centered at 03.74, J=11 Hz, J=7 Hz; H centered at 03.68, J=11 Hz, J=7 Hz), 3.46 (2H, m), 3.01 B 116 (1H, m), 2.75 (1H, m). Labeled methyl ether 788 was identical to the parent compound 88, save for the absence of signals at 04.76 and 04.33 in its 60 MHz 1H NMR spectrum. Cis,endo—2,6-bis(hydroxymethyl)bicyclo[3.3.0]octa-3,7- diene 88. Nucleophilic Ring Opening of Tricyclic Ether 78 with Methanol Dianion. Tri-n-butyltin hydride was prepared from bis(tributyltin)oxide (Pfaltz and Bauer) and poly- methylhydrosiloxane (Aldrich) according to the procedure 72 of Hayashi, Iyoda and Shiihara. Hydroxymethyltri-n- butyltin 87 was prepared by the method described by Still.49 Thus, a 250 mL three necked round bottom flask was fitted with a nitrogen inlet, stir bar, ground glass stopper and a serum cap, and then flame dried under vacuum. A nitrogen atmosphere was established, and dry THF (75 mL) transferred to the flask. Diisopropylamine (10.3 mL, 73.5 mmol) was added, and the solution cooled to 0°C, whereupon n—butyl- lithium (73.7 mL of a 0.95 M solution in hexane) was slowly added. Stirring was continued at 0°C for 5 min, and then tri-n—butyltin hydride (18.4 mL, 70 mmol) was added slowly from a syringe. After complete addition, the reaction was stirred 15 min at 0°C, and then paraformaldehyde (2.1 g, 70 mmol; dried overnight in a vacuum oven at 40°C) was added all at once. The reaction was stirred at room temperature for 3.5 h, and then poured into petroleum 117 ether (450 mL). This solution was washed with water (2 x 300 mL), saturated NaCl solution (1 x 50 mL), and dried over NaZSO4. Removal of the solvent at reduced pressure yielded a colorless oil (19 g). The crude product was chromatographed on a 2 inch diameter column containing silica gel (150 g) packed with hexane. One large fraction was collected (450 mL) in which residual nonpolar tin com— pounds resided. The solvent was changed to 10% ethyl acetate in hexane, whereupon hydroxymethyltri-n—butyltin 87 eluted. Removal of the solvent left a colorless oil (13.8 g, 61% (lit. 67%)). This material had the same Rf on TLC as reported in the literature: (10% ethyl acetate/ hexane) Rf 0.31. The NMR Spectrum was consistent with the 1 reported structure of the adduct. H NMR (CDC1 60 MHz) 3, 03.63 (2H, 5, two isotope satellite peaks, J=6 Hz), 1.93- 0.46 (28H, two overlapping multiplets). This material partially decomposes within a day, and therefore it should be used directly in the next step. A 250 mL three necked round bottom flask was prepared as before, and hydroxymethyltri-n—butyltin 87 (13.8 g, 43.1 mmol) in dry ether (75 mL) was transferred to the flask with a cannula. This solution was cooled to -78°C under a nitrogen atmosphere, and n-butyllithium added slowly from a syringe (68.1 mL of a 0.95 M solution in hexane, 64.5 mmol) while the solution was vigorously stirred. The re- action was stirred at ~78°C for l h, then warmed to -40°C, 118 and stirred for 45 min at this temperature. After cooling the reaction to -78°C, tricyclic ether 78 (1.01 g, 7.51 mmol) in dry ether (10 mL) was slowly added. After com- plete addition, the reaction was allowed to warm to 0°C and stirred at this temperature for 2 h. During this time a very heavy precipitate formed (the dilithium alkoxide of 88 is insoluble in ether). At this point, the reaction is probably complete. In this run, however, the reaction was stirred overnight at room temperature, and then cooled to 0°C and cautiously quenched with methanol (exothermic!). A minimum of water was added to dissolve the precipitate (10 mL) and the layers separated. The aqueous phase was back extracted with chloroform (2 x 50 mL) and the organic extracts combined. Drying over NaZSO4 and removal of sol- vent left two immicible oils, one very viscous, and the other quite fluid. The free flowing oil was removed from the flask with a pasteur pipette, and chromotographed on a 6 cm diameter column containing silica gel (125 g) packed in 20% ethylacetate/ether containing 1% methanol. Dienediol 88 eluted as a viscous, slightly yellow oil (140 mg) . The crude viscous oil was chromatographed on a 2 cm diameter column containing silica gel (20 g) packed with 20% ethyl acetate/ether and 1% methanol. Elution of the column led to dienediol 88 (430 mg). The combined chromato- graphed yield, therefore was 770 mg, or 62%. After isolation of the diol, it was clear from 119 miscellaneous absorptions in the NMR in the 1 ppm region that a small amount of tin contaminant remained. Thus, the product was rechromatographed on silica gel (25 g) with 3% methanol in ether to afford pure dienediol 88 as a clear, colorless oil (470 mg) which began to crystallize on stand- ing. A less pure fraction 93. 90% in 88 was also isolated as an oil (90 mg). The yield of twice chromatographed material was then 44%. The actual yield is much higher, how— ever, as chromatography of a diol generally only leads to 70% recovery. It would be advantageous to recrystallize the product directly from the crude product mixture. The spectral properties of diol 88 are as follows: IR (neat) 3360 (br), 3045, 2895, 1740, 1020 cm—1; 1H NMR (CDC1 250 MHz) 6.82 (2H, dt, J =6 Hz, Jt=2 Hz), 6.61 3! (2H, dt, J d d=6 Hz, Jt=2 Hz), 3.73 (2H, dd (AMX), J=11 Hz, J=7 Hz), 3.66 (2H, dd (AMX), J=ll Hz, J=8 Hz), 3.58 (2H, t of m, Jt=8 Hz), 3.04 (2H, m), 1.65 (1H, s, OH), 1.49 (1H: S! OH); 13 C NMR (CDC13, 62.8 MHZ) 131.67, 131.35, 63.72, 50.48, 51.22 ppm; MS (CI) m/e 167 (M+1, 9%), 166 (M, 0.7%), 149 (37%), 131 (100%), 121 (30%). Cis,endo-2,6-bis(acetoxymethy1)bicyclo[3.3.0]octa-3,7- diene 777. Diol 88 (3 mg, 0.018 mmol) in 5 mL chloroform was refluxed for l h with dry pyridine (0.10 mL) and acetic anhydride (0.10 mL). After cooling the solution, ice and more chloroform were added, and the reaction was extracted 120 with dilute HC1 and saturated sodium bicarbonate (2 x 20 mL). Removal of the solvent gave an oil which was chromato- graphed on silica gel with ether. One compound was iso- lated (4.5 mg, 100%) which was quite pure from TLC analysis. IR (neat) 3045, 2950, 2880, 1735, 1230 cm’l; 18 NMR (00c13, 250 MHz) 5.69 (2H, dt, Jd=6 Hz, Jt=2 Hz), 5.55 (2H, dt, Jd=6 Hz, Jt=2 Hz), 4.10 (2H, dd (AMX), J=11 Hz, J=7 Hz), 3.99 (2H, dd (AMX), J=ll Hz, J=9 Hz), 3.56 (2H, t of m, Jt=8 Hz), 3.13 (2H, m), 2.08 (6H, s); 13c NMR (CDC1 250 MHz) 3’ 170.99, 131.47, 130.82, 65.34, 50.52, 47.48, 20.99 ppm; MS (c1) m/e 250 (M+, 0.5%), 249 (3%), 205 (4%), 177 (2%), 163 (10%), 129 (34%), 117 (4%), 61 (100%). SUMMARY The synthesis of gig, Egggfz,6-§i§(hydroxymethyl)bi- cyclo[3.3.0]octa-3,7-diene, a key intermediate in our ap- proach to the synthesis of dodecahedrane, has been achieved. Some interesting chemistry has been learned en route to diol 88. This includes the observation of an extraordinarily facile SNi' attack of an organolithium species on an un- strained double bond in the generation of tricyclic ether 78. It can also be concluded from this work that a [1,2] Wittig rearrangement rather than an anionic [2,3] sigma- tropic rearrangement is the preferred, if not exclusive, reaction path in a methyl-metallated bicyclo[3.3.0]-3,7- diene-2,6-diol methyl ether. Finally, eXploitation of the geometrical constraints in tricyclic ether 78 led to a chelation—controlled stereoselective SN2' ring opening of 78 with methanol dianion. Concerning the use of diol 88 in the synthesis of do- decahedrane, model studies have served to convert the saturated analog of 88 to the dibromide and dithiol in ac- ceptable yields. The successful application of these re- actions to diol 88 thus seems assured. An element of un- certainty is introduced into the overall synthetic plan (summarized in Figure 10) by the uncharacterized macro- molecular sulfide products obtained in the model dibromide- 121 122 /""'o SH Summary of the sulfide contraction route to Figure 10. hexaene 78. 123 dithiolate dimerization. The appearance of a major product in this reaction bodes well for its eventual characteriza- tion as the desired gig-sulfide 88. Completion of the model studies thus requires the isolation and identifica- tion of gig-sulfide 88, and its transformation to diene 88. The application of these reactions to the "real system" and the synthesis of hexaene 78 must then be addressed. Finally, the exciting possibility of an electrocyclic ring closure of hexaene 78 to dodecahedrane, and the possibility of trapping a metal cation inside the cavity, needs to be investigated. APPENDIX 124 125 a~HH.m o.H.N.NQOHoonMuOXOIMImeINVanmfioAwmz ow .m .Nw Aamlvlcmflpum.mnmompcs Hooov snuuommm mzz cououm .H¢ musmflm 126 Hfl wusmflm on o v AIA v.3“.- . mumL-DJ J - - 127 Imo.m.mQOHU>UHchmH>£umEImHQIm.N mo Aux: omm .m .ww wcmuoo HUQUV Esuuommm mzz cououm .N< musmflm 128 o m< mucosa 0 _ v m 1 4 d 11 A! 1‘ 111‘ 4w 14‘ 1‘ 1‘ 1 ‘ J‘ 1‘ 14 J 4 1‘ .1 ‘ Id 4 f 4 11 < J 4 4 fiJJ4+11¢JJJA 4 4 % I we. O .20 .30 \ \ 129 -Aasnumasxouvsncman-m.~-oocm.mao mo Awmz 0mm .m .Ww mangoomo.m.mHoHo>oHn HUQUV Esnpommm mzz cououm .m< musoflm 130 AL l1 0 L Figure A3 131 IaaxcumEOEounvmHQIo.NIO©:m.mflU mo ANmE 0mm .m .ww mcmuooho.m.mloHo>ofln HUQUV Esuuowmm mzz cowoum .v¢ wusmflm 132 O P fil 14 4 4‘ d d 1 l - I4 1 l4 1 1“ c C - J v< muomflm N IJIll—‘l-‘lll‘l 404.114 D 133 IAH>£umEouQmoumEVmflbno.mI0©cm.wflo mo Ammz 0mm .m .ww wcmuoomo.m.mgoHomofln HUQUV Esuuommm mzz :Ououm .m4 wusmflm 134 :m.: ”u" 0 m4 2J1; é 135 .ww Heacufla acm ww wofleounflm mo cofluommu mcaamsoo mcp CH Awom .mmv acupoum HohmE wnu mo ANmE omm .mauoov Esuuowmm mzz cououm .mfl wusmflm 136 $4 madman 137 .ww mamaouh.mumuoomo.m.onHomoflnflsxoaumsAH>: LcmumawvsnnclfluuvQmflnlm.NIOUcw.mHo mo Aux: on .m HUQUV Eduuowmm mzz cououm .nd musmfim 138 n4 wusmflm . _— __ —-— : a. ‘ 5...; :r. :C :\ d 1 fl 4 J a d 4 a d J _ q d 1 q H a 1 d 1 l4 1 fl 4 1 d d d - . d "anew 139 IGCOCfi OHJV o.a.m.m_oHo>ofluunmxoum mo ANmz om. m .ww mcmflcum.m Hooov gunpowmm mzz cououm .m< musmflm 140 m< wusmflm I :_ :x. .: 2.. as 5:; 2.. so 3\ 4 d‘ d _ q d T d —r 1 1 i 4 — d d d d _ d d d 1 1 d 1 d — 1 11 d I F q - q . 1 4 a < 4 4 — 4 q 4 a a 4 d — . a q q - a a . J < a q — . d 141 Imcocw oa.v o.a.m.mgoHo>oHuunmxoum mo “um: omm .m .$ 658$ .m HUQUV Esuuommm mzz coeoum .md wusmflm 142 m< ousmflm )3 5 g; k c _. _l l mm 143 .ww wcmflpuh.mumuoomo.m.muoHo>o Iflnazusnuolawsuwewxou©>£|0pcwlm mo Aux: ow .m Huoov Esuuommm mzz conoum .oam musmflm 144 oac musmflm .1. Sr 145 .Nw wcmfipnh.mumuoofio.m.mgoHozoflnwxonumEAawccmwm l'l'lIl‘IIll IawusnlmlfluuvIol>xou©>£nm100cw.mflo mo ANmz on .m HUQUV Esuuommm mzz cououm .HH¢ musmflm 146 Haa mucous 4:. : a. .1. 2.. 2.. L. 2.: on so .1 1 1 1 J‘l 1 1 41 + T 1| 1 1 1 1 d 1 1 _ 4 1 a d — M 4 - _‘ 4 d M — d M 1 4 d M _ J4 _ J m u _ v . u l . _ . l _ _ -~_—_’.._- ..- ._.__.._--.__ -—--— .-- 0041’I'II ,.__——-...—- _-—--—~ _--— ——- - on“. -—.—_——... “o- “cu-mm o“..- . --—_~o~-—-¢-o-..- 14? .WW mcmwplh.MImuoomo.m.mgoHo>oflQ lullllI' I| .Aacumexxoup>cuoI>x0HU>£Im10pcm.mflo mo Aux: 0mm .m Huoov Esuuowmm mzz cououm .Nad wusmflm r..- w 1.44444444444 41!..- 148 @ mafi mnsmflm v m 444444444 d 41444 4 4 <4444‘4I4'14J14lfi4l14’414l4ljsl14l14l1 \ift 149 .ww wcwflp|>.mumuoomo.m.onH0%oH£ Ifiamcqumxoupacvmflblm.m10©cw.mflo wo ANS: omm .m Hooov snuuommm mzz cououm .mam musmflm 150 ma< ousmflm o P N n —y* P F > FL F FLr \PLL F LVL+LPL P LP PL \P P ? v @ Qm P}Flrk—LVPFlrlrxrkrfihrPLrbeF*L*Lrer—VP**LFLIPrxPx—rL _ m .01 i, “"\ hm u.111111J—111111111d1141u4111—111141111‘1 is \w'“ 10. 151 .wew mamaonb.mumuoomo.m.onHosofln ill!!!" Iflaxnuwemxouwomvmanlo.mnoocw.mflo mo Ammz 0mm .m HUQUV Eduuowom mzz cououm .QH< musmam VHxN Ohsmflh 152 0.1 44441q1444d4444d J a... é N H H \..... 040 REFERENCES 10. 11. 12. 13. 14. 15. REFERENCES Maier, G.; Pfriem, S.; Schaffer, U., Matusch, R., Angew. Chem. Int. Ed. Eng. 1978, 11, 520. Rauscher, G., Clark, T., Poppinger, D., Schleyer, P. v.R. f Angew. Chem. Int. Ed. Eng. 1278, 11, 276. n i Eaton, P. B., Cole, T. W. Jr., J. Am. Chem. Soc. 1964, : El 9620 . Ternansky, R. J., Balogh, D. W., Paquette, L. A. 3 J. Am. Chem. Soc. 1282, 04, 4503. E For a review on synthetic efforts directed towards dodecahedrane, see Eaton, P. E. Tetrahedron 1312, 35, 2189. Schulman, J. M., Venanzi, T., Disch, R. L. J. Am. Chem. Soc. 1275, 91, 5335. Gasteiger, J., Dammer, O. Tetrahedron 1978, 34, 2939. Schulman, J. M., Disch, R. L. J. Am. Chem. Soc. 1978, 100, 5677. Woodward, R. B., Fukunaga, T., Kelley, R. C. J. Am. Chem. Soc. 196%, 86, 3162. (a) Paquette, L. A., Kramer, J. D., Lavrik, P. B., Wyvratt, M. J. J. Org, Chem. 19 , 42, 503. (b) Bosse, D., deMeijere, A. Angew. Chem. 7%, 86, 706. Gladysz, J. A. Chem. Tech. 1272, 9, 372. (a) Paquette, L. A., Itoh, I., Farnham, W. B. J. Am. Chem. Soc. 1975, 97, 7280. (b) Paquette, L. A., Itoh, I., LipkowiEE, K. B. J. Org. Chem. 1916, 41, 3524. Oljan, R. Ph.D. Thesis, Harvard University, 1976. Roberts, W. P., Shoham, G. Tetrahedron Lett. 1981, 4895. Eaton, P. B., Mueller, R. H., Carlson, G. R., Cullison, D. A., Cooper, G. P., Chou, T.-C., Krebs, E.-P. J. Am. Chem. Soc. 1911, 99, 2751. 153 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 154 Paquette, L. A., Snow, R. A., Muthard, J. L., Cynkow- ski, T. J. Am. Chem. Soc. 1918, 100, 1600. Woodward, R. B., Hoffmann, R. "The Conservation of Orbital Symmetry", Verlag Chemie, GmbH, Weinham/Bergstr, 1970, p. 70. Salomon, R. G., Kochi, J. K. J. Am. Chem. Soc. 1911, gg, 1137. Salomon, R. G., Folting, K., Streib, W. B., Kochi, J. K. J. Am. Chem. Soc. 1911, _9, 1145. (a) Salomon, R. G., Ghosh, S., Zagorski, M. G., Reitz, M. J. Org. Chem. 1, 41, 829. (b) McMurry, J. E., Choy, W. Tetrahedron Lett. 1980, 2477. Rathke, M. W., Lindert, A. J. Am. Chem. Soc. , 2;, 4605. 12% Farnum, D. G., Hagedorn, A. A. III Tetrahedron Lett. 1915, 3987. Hagadorn, A. A. III Ph.D. Thesis, Michigan State Uni- veristy, 1974. Jenkins, C. L., Kochi, J. K. J. Am. Chem. Soc. 1911, 99, 843. Taken, in part, from Boeckelheide, V. Acc. Chem. Res. 12§QI lg, 65. Mitchell, R. H., Bolkelheide, V. J. Am. Chem. Soc. 1911, 99, 1547. Mitchell, R. H., Otsubo, T., Boekelheide, V. Tetra— hedron Lett. 1915, 219. (a) Otsubo, T., Boekelheide, V. Tetrahedron Lett. 1915, 3881, (b) Otsubo, T., Boekelheide, V. J. Org. Chem. 1311, 33, 1085. For a review on the Ramberg-Backlund rearrangement, see Paquette, L. A., Organic Reactions Vol. 25, John Wiley and Sons: 1977, pp. 1-71. Swepston, P. N., Lin, S.-T., Hawkins, A., Humphrey, S., Siegel, S., Cordes, A. W. J. Org. Chem. 1981, 49, 3754. . Buter, J., Kellogg, R. M. J. Org. Chem. 1981, 49, 4481. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 155 Ochrymowycz, L. A., Mak, C., Michna, J. D. J. Org. Chem. 1919, 9, 2079. Farnum, D. G., Mohaddes, M., Monego, T. A. unpublished results. For organo aluminum-mediated aliphatic Claisen rearrangements, see Takai, K., Mori, T. Oshima, K., Nozaki, H. Tetrahedron Lett. 1991, 3985. Hassner, A., Reuss, R. H., Pinnick, H. W. J. Org. Chem. 1919, 99, 3427. For example, see Baldwin, J. B., DeBernardis, J., Patrick, J. E. Tetrahedron Lett. 1919, 353. Still, W. C. Mitra, A. J. Am. Chem. Soc. 1978, 00, 1927. “””” 7“ Still, W. C., McDonald, J. H. III, Collum, D. B., Mitra, A. Tetrahedron Lett. 1919, 593. Hagadorn, A. A. III, Farnum, D. G. J. Org. Chem. 1911, 99, 3765. Wittig, G., B611, W., Krfick, K.-L. Chem. Ber. 1991, 99, 2514. (a) Conia, J.-M., Limasset, J.-C. Bull. Soc. Chim. fig. 1997, 1936; (b) Greenwald, R. , Chaykovsky, M., Corey, E. J. J. Org. Chem. 12199933 9_, 1128, (c) Peter- son, D. J. J. Org. Chem. 780; Chan, T. H., Chang, E., Vinokur, E. Tetrahedron Lett. 1910, 1137. Schostarez, H., Paquette, L. A. J. Am. Chem. Soc. 1991, 103, 722. Dauben, W. G., Walker, D. M. J. Org. Chem. 1991, 1103. Brown, H. C. "Organic Synthesis Via Boranes", John Wiley and Sons: New York, 1975. Scouten, C. G., Brown, H. C. J. Org. Chem. 1913, _§, 4092. Brown, H. C., Zweifel, G. J. Am. Chem. HSoc. 199, 1512; Zweifel, G., Nagase, K., Brown, s93. 129%, g3, 190. Hooz, J., Gilani, S. S. H. Can. J. Chem. 1999, 99, 86. 8Chem. Urquhart, G. G., Gates, J. W. Jr., Connor, R. Org. Syn., Coll. Vol. 9, 1999, 363. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 156 Farnum, D. G., Gur, Y. T., unpublished results. See also Luche, J. L. J. Am. Chem. Soc. 1919,100, 2226. Still, W. C. J. Am. Chem. Soc. 1919,100, 1481. Wakefield, B. J. "The Chemistry of Organolithium Com- pounds", Pergamon Press Ltd.: Oxford, 1974, pp. 89- 108. Broaddus, C. D. J. Org. Chem. 1999, 99, 4431. Veefkind, A. H., Bickelhaupt, F., Klump, G. W. Rec. Trav. Chim. Pays-Bas 1999, 99, 1058. -. land.” Mulvaney, J. B., Gardlund, Z. G. J. Org. Chem. 1999, 99, 917. Veefkind, A. H., Schaaf, J. v.D., Bickelhaupt, P., Klumppr G. W. Chem. Comm. 1911, 722. Moriarty, R. M., Yeh, C. —L. Tetrahedron Lett.1979 383. (a) Baldwin, J. B., DeBernardis, J., Patrick, J. E. Tetrahedron Lett. 1919, 353; (b) Rautenstrauch,v. Chem. Comm. 1919, Baldwin, J. B., Patrick, J. E. J. Am; Chem. Soc. 1911, 99, 3556. Lansbury, P. T., Pattison, V. A. J. Org. Chem. 1999, 91, 1933. Cazes, B., Julia, S. Tetrahedron Lett. 1919, 2077. Nakai, T., Mikami, K., Taya, S., Kimura, Y., Mimura, T. Tetrahedron Lett. 1991, 69. Baldwin, J. B., Urban, F. J. Chem. Comm. 1919, 165. Evans, D. A., Andrews, G. C. Acc. Chem. Res. 1919, Z, 147. (a) Baldwin, J. B., Hackler, R. E., Kelleyl, D. P. Chem. Comm. 1999, 1083; (b) Greico, P. A., Meyers, M., Finkelhor, J. . Ogg. Chem. 1919, 99, 119. Warpehoski, M. A., Chabaud, B., Sharpless, K. B. J. Org. Chem. 1999, 91, 2897. Seyferth, D., Weiner, M. A. J. Am. Chem. Soc. 84, 361. 128% 66. 67. 68. 69. 70. 71. 72. 157 Seyferth, D., Vick, S. C. J. Organomet. Chem. 1978, 144, 1. “ ”m (a) Crandall, J. K., Clark, A. C. J. Org. Chem. lgzg, 31, 4236; (b) Crandall, J. K., Clark, A. C. Tetra- hedron Lett. lgég, 707. Gordon, A. J., Ford, R. A. "The Chemists Companion", John Wiley & Sons: New York, 1972, p. 379. Still, W. C., Kahn, M., Mitra, A. J. Org. Chem. , 33, 2923. $2Z§ Watson, 8. C., Eastham, J. F. J. Organomet. Chem. $967, 9, 165. Seyferth, D., Andrews, S. B. J. Organomet. Chem. we 3_o. 151- Hayashi, K., Iyoda, J., Shiihara, I. J. Organomet. Chem. lgéz, lg, 81.