\ I . giiiij $32331. . . v. 57.. 1 . .. . Ll 592.13.?! 1 >. i . L « .2. if: .y as: LIB! .m. . r! I... I ZhflnMW’L351353lPS Z: 7....) u u... u .yM..5)55 5h...‘.llurl .h?dfin‘i§§f¢fi9>fl’.1hfilfll¥ufiHALUQhMHPngthgx.Allis): . . .5. n‘i\.:hvfi\wn$ufiw§vi§haoihaL§.§? .:..L|\..th;flo>\.rnfn. h. ‘\h>?.)\i£ r u. h . ~ a \Lt‘us :U..).vs.s.=)afi (JVH’U.§3.9K§I5!.§§\£\R NIH-AL. .. . u) a . ..§VM.\2. . v I; - ON“ I diva...” 5|». 7:. I . r . . «at: “mehvuzktfi it... ‘H’Hvsxav..lnuh,u%.:nuhw ihfi.\...1lflflg. a “(yang-Is; 53% ’2: "7": 3"! n v I u v. 1.): . . u:.i’.lz\ - i luv. .0. 7:..1!..5IVY9VNK~MM. ..... .. y . 7.“.257 ahuéfii l .i v: ((5 gig Fifi-32:; w‘vnzr’.slyficwfivns: a pa... 7 u! that. .51. .. . . . . . I. E. Ehrgga :vguix .5 l... Pitt‘s. (221.1931... 2. 711.132... aaaaaa 3 .zxfiianz a: 2.. 1 .. 2 «i .5. . - u 5‘ 1973. lbs:\\)9 .. gsé )1\$b..rr~t. rXrafivlS.§.t.Lr/t+hn.fivsflsturu9i.hht\$¢\fl { {IV ‘£.V§}.l¢z . 51.7.5. 7......9. .1) :fi. H: v... Ix.‘1:t~¥v.{13vruvr . .1745)?! 1. «3.55 L.) 7.....\ . 1.. IA {slings-(5W3 I159E.»!a..-\3 .‘Piilry\t22£§ 11.43- 3;:L1571rt .51; flag 2:. I7.)..1...i.£...nt:a.ch .x :1 z . at}!§>hV1..1&? L2 32%.“? kl.ii§i-§§i1§ . 5 fi\...li’.£h5§. :vi». 7. :stEE.‘ Irvugzh’ 1.1.2.2325 Z4133? ... .. 3... 3... 1).. 4?. I: .v A... M35: 3. £16312: i3 32$}intttsnlri. {:19 ir‘.r:n7:!xtc.t. {79.35 in): £.;ta1l)t.iu . 2 ¢ \ tut-v.13... i; git-IA". ‘ Sikhs... Aiigz‘iiri «5‘4 If S ’58.??‘5‘?lf:tl— :3: a £5.33: . . $125.91.; : 3:31.. . . ., .3: :....?:ihi£:z . u 53......»152: $98.11;}: .3::. .02; 3'1. 5".5llflvv 1v 1 1.91.1! 1.3T: t,‘vi?tfi 3 5.5.3324 t I 25110. £52. 9-. 31.333.55.59 l:::§lli 5 tithé‘itx 021! :29) L4. .5...) .. dip-n: 25.1.1.5. i xiv-1&5 .. Jug ~2E..g e: I: G533».§.§15n.ifl~.:uiiii.§sg‘flfi xggfgé.“ , Q! r , ’5 I ‘ g. ! z;:;,§ ‘ t s ) it: ~ a viazvr 37 3%. L5... 31.2.: zlxvt-ztl : .... Nv 71m: 9 .. . ti 3 séiusxbf} .l u §.1£§ is}... .1; .. Hialfiuwnmuvfflfiifi wan: 8%. kg?! .1315: {gauné EVLfio$PPn3§§FS§§ xiii”. fikitn .Sfifxuu... haiifa _ . . . t .4 l... 28“ . g5 . z. .. .. 5.. ts $5.53.??? . g‘: int! 1.19.1.5 Lilith”?! 7‘1““ x . ~. a . )5tibsh!.fll\ ’9‘ i}, \uvas‘gla h {“Enl‘ ~ 9 . v. .. .. . .g . . i... 3. .. . . as: v. I . 5.1.3.... iglzvzluammurz. . fl! .. . I}!!! n 3..» ..... L3:$.Svl.; .5... 9!. ngiafizuwhgflaalywhh 1.... $114.... 1 {1:83 1 . ~ . $1“: ‘9 z . a . ~ .. I . fizéaas .1.) n 009:9..‘ullk‘39ti «H»I~.lv;VJ-u.,.‘1!2!vSign .HaizfisnrwlnulIWIN’ I t . 7 x. ~ . .Iflfisgtvllvi.§w :ll'flHaiu-Lf 568.41.!1. )5 I I. 2c.li.§a'.wlA-.Esh.'flh‘tti 00-h" ...«w,hVQh—I~Au.fld.flu4i I; i 5!. . fin. . i532 . 1......ivnIJmatvltlJavIJ.....~.n..-.H.3N.tt‘$t y‘tv'vfi 2......tl1fn'nth‘2anflnZy7ldfiu .. “new“... lily IAN“! 1...”...55 .. z .. . T u: Edvixgnv‘n Eviarfvairfiflci s. tintil.7‘.l. . 523.019.. tgfia 16 “Ian Eyizlnvflfltliultut . 1.1.1.1.... :5... .3 1.... 1.1.. . 1.. 3L 33%, gay. 8 221...... ifivzydirgs; 6E§.t.zflz.h . lihiiiihvflflihnnflfiis. “$92-1. . 12.1%. i: xiii..1tn.i!¢i..s , .. ..m.....i!.nu.tsv.sun§. «hum...» -gsffivmwabihn L... . . § .1831}... . .. .. Bum... S:..In.3§ii .351 . g .11. .3434... .. (3.3.... LR: . .i Izfignflu .. gguzizu 2.2.1.3. . :2 3.... u... 1.. 153.51....ng . I: . 01.98! jib-’31 .2... .b’it‘lfl.‘ ‘5: avv‘nggilliiitll flinch“: g \ti‘§§§.égfi ,. - ...:.. . z .1... (twangfiq. . ..... 1c. szfiqflhfiflfib $1513....a4i $1.32 SEER... t . «Nana. . . .1...‘ . :v: lttflJ-I‘: : §HQ.? A 151ng a 34.17 3 l: . . a .. . R3503... 1.... $35th . .. . l w. . p ‘i! 17% Q- i u th’fi‘WF-fiv arguggaabu .mtaiadwtt. 1%.... v-w : ‘ ‘ 77 . .1nu§.i+.§dnuiyfisvlvl§i 58.51.. Wiinsg .. :7 . . .. Elana 1...? E... 5.9.... . . r .1 ii 1.9... .QWWWELI I'lfluntle. 1&1. Lu. :5 ; f. .. 6 2 3 {€111th . v1. 13.}.11‘5!" 57.! . . ‘3‘; if.“ 131312;. 03? . Nahuunxvasiwlcvi iiiilx .f~%%§3 . .siliqivis . . v . 7R I . a . . . . \1 \w . . . . 1%vii‘ é I114}! Jill-71.1% if). ‘1‘}..1511 . 12.1.: IL." .G‘i‘i} .. i1. . . .\ . . . . I I... In». ..T‘ 4:9): Estfr? .410 gle A 214%.? fig. {’7}!!- : a .. . . n . . . .. . Eirbwzx. i; . . i . It§{.flf L... ........... n Bl§5tsjr~5i§lu v T c; . ‘1}! $5.. 3 II... 111%! iiiwf‘iiuiii . lhhsxrhu‘n E; ii"! (.1 . .itfil‘ls._..os n... 13111.... ‘ .l 7111: Lil???) a...) v' 5%? .I. 3 .l q :1 v.75} . a! $11. iii‘fi‘i . lllt.‘ IvoL 10% I giqui‘i‘tu. c. 11%! 1%" . i173§§ . 3.5.31! 19% 1... 1.19 kg Iii-Hz. r73.\\1‘..1§114§51i )i‘lxls‘ .‘anu. 12.3.”: 3351(“3: gtfivvi$iilliw . Isl. It?” . {I 1 2. 1311-13.“ .\ ‘3 II. 13.... )2... 7‘ gi‘!)1vls%fifl‘\\ll.l 11‘): 13!»! g. . . . #333. i H.) . .11}? 71“: .2... x d. ‘1‘ \ ‘11.. (1‘7... .2. $1....X.‘11111\\Y\)..1$}1 .. 1115\13111‘i\s‘\‘q I \ It‘ll. Egtx . . ,‘\.1 ti...) , I .. .. . .J'. ... ‘ 1 § . . .1.v1\i ,..v.1| 1...... .41. .5 ~ . .7:.x.:. . . . It .y . . . \z TH E315 This is to certify that the dissertation entitled FACTORS AFFECTING DIASTEREOSELECTIVE REDUCTIONS OF 9 , 10-DIHYDRO-9 , l 0- (1 1-KETO- 12-ALKYLETHANO)-ANTHRACENES presented by BRUCE R. OSTERBY has been accepted towards fulfillment of the requirements for Ph. D. CHEMISTRY degree in MR Major prol'éssor Date I”?! I 3)! 3k MSU is an Afflmmlive Anion/Equal Opportunity Institution 0-12771 RETURNING MATERIALS: IV1ESI_J Place in book drop to LIBRARJES remove this checkout from as. your record. FINES Will V be charged if book is returned after the date stamped below. FACTORS AFFECTING DIASTEREOSELECTIVE REDUCTIONS 0F 9 10-DIHYDRO-9 10- 11-KETO-12-ALKYLETHANO -ANTHRACENES by Bruce R. Osterby A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1982 ABSTRACT FACTORS AFFECTING DIASTEREOSELECTIVE REDUCTIONS 0F 9.10-DIHYDEO-9L10-(ll-EETO-IZ-ALKYLETHANOz-ANTHRACENES by Bruce R. Osterby A cornerstone in theories of diastereoselective reductions of ketones by complex metal hydrides is that the reagent attacks the least hindered side. Varech and Jacquesl’z, however, reported that this axiom does not hold true for a series of bicyclic ketones, but was in fact directly opposite to it. When 9,10 - dihydro - 9,10 - (11 - keto - 12 - alkylethano) - anthracenes (alkyl = methyl, ethyl, and iggépropyl) were reduced by lithium aluminum hydride, attack of the hindered (alkyl) side increased as the steric bulk of the alkyl group increased. The percentage of attack of the alkyl side was 35, 51 and 80% for methyl, ethyl, and iggfpropyl, respectively. When the bulky group was trbutyl or phenyl, attack of the hindered side was 23 and 0%, respectively.. These "surprising" results were attributed to an anistrOpic inductive effect.3 I have examined the product ratios as a function of temperature and calculated the parameters AAHI and AASI. The study was done for lithium aluminum hydride reductions in diethyl ether and tetrahydrofuran and for sodium borohydride reductions in isopropyl alcohol. The AAH‘ component indicates that attack of the more hindered side is due to enthalpy factors. This has been attributed to product development control. The AAS:F component shows a strong preference for attack of the least hindered side (to give the gig alcohol). This preference is more pronounced for the trbutyl group than for the other alkyl groups and is more pronounced in tetrahydrofuran than diethyl ether. These results are most likely due to solvent entropy effects. In the phenyl substituted case the hydride attacked from the least hindered side in all cases studied. This has been interpreted as the result of a strong coulombic repulsion between the negatively charged nucleophile and the electron density of the phenyl n system. 1. D. Varech and J. Jacques, Tetrahedron Lett., g_, 4443(1973). 2. M. J. Brienne, D. Varech and J. Jacques, Tetrahedron Lett., fig, 1233(1974). 3. M. C. Cherest, H. Felkin, P. Tacheau, I. Jacques and D. Varech, Chem. Comm., 372(1977). ACKNOWLEDGMENTS I would like to express my deepest gratitude to Dr. Gerasimos I. Karabatsos, under whose guidance this research was performed. I also would like to acknowledge the faculty, staff and graduate students of Michigan State University's chemistry department for their help and friendship. I would like to give a special thanks to my mother and father for their unfailing support and love throughout my education. Finally, I wish to express my appreciation and gratitude to my wife, Meg, for her unselfish support and immense help in making this all possible. Her effort and assistance will always be remembered. ii TABLE 93 CONTENTS page LI ST OF TABLES O O O O O O O O O O O O O O O 0 v LIST OF FIGMS O O C O C O O O O O O O O O O 0 v i INTRODUCTION I. The Scope of the Problem . . . . . . . . . . . 1 II. The Stereochemical and Energy Relationships for Carbonyl Addition Reactions . . . . . . . . . . 2 III. Models for Diastereoselective Additions to and Reductions of Ketones (Factors Affecting Asymmetric Induction) . . . . . . . . . . . . 6 A. Steric and Conformational Concepts . . . . . . 7 l. Acyclic Ketones and Aldehydes . . . . . . . 7 2. Cyclic Ketones . . . . . . . . . . . . 11 B. Molecular Orbital Models . . . . . . . . . . 16 IV. Hydride Reductions of 3 - Alkyl (and Phenyl) Bicyclo[2.2.2]octan - ;,- ones and 9,10 - Dihydro - 9,10 - (11 - Keto - 12 - Alkyl (and Phenyl) ethano) - Anthracenes . . . . . . . . . . . . . . . 19 RESULTS AND DISCUSSION I. RESULTS . . . . . . . . . . . . . . . . . 23 II. DISCUSSION . . . . . . . . . . . . . . . . 33 A. The Composition of Lithium Aluminum Hydride in Diethyl Ether and Tetrahyrdofuran . . . . . . . 33 B. The Mechanism of Lithium Aluminum Hydride Reduction of Ketones and the Role of Alkoxyaluminohydrides . . 34 C. The Mechanism of Sodium Borohydride Reduction of Ketones in Alcohols . . . . . . . . . . . 37 D. The Activation Parameters . . . . . . . . . 38 iii 1. The Role of Enthalpy . . . . . . . 2. The Role of Entropy . . . . . . . . 3. Conclusions . . . . . . . . . . . EXPERIMENTAL I. General Methods . . . . . . . . . . . . A. Stock solutions of complex metal hydrides . B. Temperature Control . . . . . . . C. General Procedure for Reductions . . . II. Syntheses . . . . . . . . . . . . . . III. Product Ratio Analysis . . . . IV. Error Analysis . . . . . . . . . . . . APPENDIX A MAGNESIUM-25 NMR STUDY OF ORGANOMAGNESIUM COMPOUNDS AND MAGNESIUM HALIDES IN DIETHYL ETHER AND TETRAHYDROFURAN REFERENCES . . . . . . . . . . . . . . . iv . 38 45 46 47 47 47 48 49 . 62 66 . 68 . 75 LIST 0_F TABLES Percentages of alcohols from equations 2 and 3 (%trans = 100% - %cis). Reactions were run at 20°C in diethyl ether. 0 O C O O O O O O I O O O O 0 Product ratios from lithium aluminum hydride reductions in diethyl ether. 0 O O O O O O O O O 0 Product ratios from lithium aluminum hydride reductions in tetrahydrofuran. . . . . . . . . . . Product ratios from sodium borohydride reductions in isoprOpyl alcohol. . . . . . . . . . . . Results from thermodynamic equilibration via the MPVO reaction. Lithium tri - t_- butoxyaluminohydride reductions. . . . . . . . . . . . . . Activation parameters from Figure 11. Lithium aluminum hydride in diethyl ether. . . . . . . . . Activation parameters from Figure 12. Lithium aluminum hydride in tetrahydrofuran. . . . . . . . . Activation parameters from Figure 13. Sodium borohydride in isopropyl alcohol. . . . . . . stg nmr Data for some Grignard Reagents and Magnesium Halides in Diethyl Ether and Tetrahydrofuran (THF). page . 25 . 27 . 28 10. 11. 12. 13. 14. 15. LIST OF FIGURES The three stereotopic relationships. . . . . . . Enantioselective (left) and diastereoselective (right) reaction profiles. . . . . . . . . . . . . The open chain, cyclic and dipolar models (8, M, L are small, medium, and large, respectively). . . . . The two transition states proposed by Karabatsos. Felkin's model for 1,2-asymmetric induction. The preferred transition state is on the left (S, M, L are small, medium, and large, respectively). . . . Felkin's model for cyclohexanone systems. . . . . Wertz and Allinger's model based on fewer gauche vicinal H/H interactions. . . . . . . . . . Felkin's torsional strain adapted to cyclopentanone systems. I O I O O I O O O O O O O 0 O The case of camphor and norcamphor. . . . . . . Ahn's antiperplanar effect. . . . . . . . . . Lithium aluminum hydride in diethyl ether. . . . . Lithium aluminum hydride in tetrahydrofuran. . . . Sodium borohydride in isopropyl alcohol. . . . . 4 - t - butylcylcohexanone and 2,2 - dimethyl - 4 - t - butylcyclohexanone. . . . . . . . . . . . The bicyclo[2.2.2] system. . . . . . . . . . vi page . 10 . 15 . 19 . 31 . 41 INTRODUCTION I. The Scope _f the Problem When 4-trbutylcyclohexanone is reduced by lithium aluminum hydride a pair of diastereomeric alcohols is formed in the ratio of 9:1 (Eq 1).1 0 L M IH H" H H W 5204’ m 90% ' l0°Io V The reaction is stereoselective, which should not be surprising. What is surprising is that the reason(s) for the observed selectivity is not known for certain. Several theories have been postulated to explain why axial attack is preferred, but no one theory has been overwhelmingly satisfying. This reaction exemplifies, .perhaps more than any other, the problem of stereochemistry in complex metal hydride reductions of cyclic ketones. The factors which control the stereochemistry of ketone (both cyclic and acyclic) reductions have been speculated about for thirty years. With almost an unlimited assortment of ketones, no single generalization is applicable to all systems. Studies of acyclic ketones and aldehydes have focused primarily on those with an asymmetric center adjacent to the carbonyl group (1,2-asymmetric induction). Studies of cyclic ketones have been divided into groups depending on the conformation(s) of either rigid or flexible systems. Aside from an intrinsic reason for studying factors which control the stereochemistry of ketone reductions, there is a practical reason. A large number of multistep total syntheses require at some point one or more asymmetric centers. Very often the methodology used in solving the stereochemical problem has been empirical. Understanding the factors which control stereoselectivity have greatly increased the efficiency of such syntheses. Such knowledge has also been useful for configurational assignments and mechanistic studies. There are a vast number of carbonyl addition reactions. One must be careful in extrapolating data from one type of reaction to another. As with complex metal hydride reductions and organo-metallic additions, most of these reactions are kinetically controlled. Therefore, the stereoselectivity is directly proportional to the difference in transition state energies, AAGI. Since transition state structures can only be surmised, determining the factors which control stereoselectivity is, by nature, a very subtle problem. II. The Stereochemical and Energy Relationships jg; Carbonyl Addition Reactions The terms enantiomeric and diastereomeric are commonly used to describe the relationship between pairs of stereoisomers. In a similar manner the relationship between portions of a molecule may be compared. Recognizing the stereochemical relationships between parts within a molecule is extremely valuable in solving stereochemical problems. Consider the two faces of a carbonyl double bond. The stereoisomeric environments of the two faces may be equivalent, enantiotOpic or diastereotopic. The classification depends upon the symmetry e1ement(s) associated with the molecule.“ If the molecule has a C3 axis along the C-0 bond, then, by definition, the faces are equivalent. They are indistinguishable under all circumstances and addition reactions would give no stereoisomers. If the faces can only be interchanged by an S axis (mirror plane containing the sp2 1 carbonyl bonds), then the faces are enantiotopic. They are indistinguishable under achiral conditions and distinguishable under chiral conditions. Addition of an achiral reagent would give a racemic mixture of enantiomers. If the reagent is chiral, then only by chance would the mixture be racemic. Finally, if the faces are not interchangeable by any symmetry operation, then they are diastereotOpic. They are distinguishable under all circumstances and addition reactions would give a pair of diastereomers in, most likely, unequal amounts. With the stereotOpic relationships for the two faces of a carbonyl defined, consider now the two reaction paths. With achiral reagents the reaction paths are related in the same manner as the products (there may be certain exceptions with diastereotOpic faces because of a preliminary reaction or coordination on one of the sides). For ketones with diastereotopic faces the transition state energies are most likely different. In a kinetically controlled reaction this difference is directly pr0portional to the product ratio. Such processes are termed diastereoselective. There is no energy difference between the transition states for ketones with enantiotopic faces using achiral reagents. A chiral reagent, however, turns an enantiotomeric relationship into a diastereomeric one. Thus ketones with enantiotopic faces go through diastereomeric transition states with chiral reagents. The term enantioselective applies to this type of reaction. The hydride reductions of the three ketones in Figure 1 illustrate these points. The energy relationships for enantioselective and diastereoselective reactions are shown in Figure 2. Although the terms enantioselective and diastereoselective have been used recently to describe reactions similar to those just discussed, the terms asymmetric synthesis or asymmetric induction are more prevalent in the literature. Any reaction where the epimers of a new asymmetric center are not formed equally is classified as an asymmetric synthesis. An enantioselective reaction is an asymmetric synthesis. A diastereoselective reaction is usually but not always an asymmetric synthesis. The example of 4 - £_- butylcyclohexanone illustrates this point. Nevertheless, most ketones with diastereotopic faces have an asymmetric center incorporated in the molecule and upon addition reactions form epimers of a new asymmetric center (provided an identical ligand is not added to the carbonyl). In studying the factors which affect diastereoselective asymmetric syntheses it is next to impossible to use isomerically pure enantiomers. This is no problem, however, since a racemic mixture gives the same information. For example, consider an Optically active ketone of S configuration. Upon reaction it gives a ratio EQUIVALENT FACES o H. OH 1C: ’2] / \ /"\ H3C CH3 H3C CH3 ENANTIOTOPIC FACES liq, H 0 ti OH I ks H / k, I . / \ .. . / \ A/ \ H3C HM szCH3 H3C CH2CH3 HBL I CHZCH3 ‘ k5 =i000 -3. 0- '400 I I I 0.0 1.0 2b .330 430 52 (l/T) x 10‘3 I Figure 11. Lithium aluminum hydride in diethyl ether. 31 3-0 "l 2.0- <><> —-<> @— -i E //E -13 / In (us/trans) I . :o [> [> O -c (CH3)3 i [:J ‘13ft3 ‘200‘ O A ‘3.C}— H OH trans ‘4‘0 ' F “n 1 I 00 L0 2.0 3.0 4.0 (l/T) x 10*3 I Figure 12. Lithium aluminum hydride in tetrahydrofuran. 32 3.0 l R g ‘ O “C (CHJIJ E] °C+t3 2.0" C) ‘Cifi207fl3 . Z: ~Cii(Cit3)2 ‘________________ 1.0- -0 -0 -<> 00- “CD—Q “‘6 \o‘ ‘23 if, A t \A ‘Efi - ‘\\\\\\£> 8 is d ‘\\\\\\\\\\\‘z: I I“ <3 I NaBH4 iFPrOH “4.0 r r I I i I ' l f T 0.0 [.0 2.0 3.0 4.0 5.0 “/77 x 173 Figure 13. Sodium borohydride in isopropyl alcohol. 33 II. DISCUSSION The composition of a complex metal hydride and the mechanism of its reduction reactions is fundamental to any discussion concerning the stereoselectivity of such reactions. For that reason, a brief account of these tapics is given. A. Thy Composition pf Lithium Aluminum Hydride ip Diethyl gfippg ppd Tetrahyrdofuran Most of the information concerning the composition of lithium aluminum hydride in diethyl ether and tetrahydrofuran has come from conductance and molecular association (ebullioscopic) studies. The equivalent conductance of lithium aluminum hydride in diethyl ether is very small and concentration independent. For the same species in tetrahydrofuran, however, the equivalent conductance is substantially larger and increases with increasing concentration.‘0 This indicates that lithium aluminum hydride in diethyl ether and tetrahydrofuran exists as contact ions and solvent-separated ions respectively. Molecular association studies of lithium aluminum hydride in diethyl ether and tetrahydrofuran were done by ebullioscopic molecular weight studies. In tetrahydrofuran, lithium aluminum hydride has an association value (defined as the ratio of experimentally determined weight to formula weight) of 1.0 in dilute solution (0.05 m). This value increases with increasing concentration to a limiting value of 1.8 (0.5 m and greater). The association value for lithium aluminum hydride in diethyl ether is somewhat higher, starting at 1.7 in dilute solution (0.05 m) and increasing to 2.3 at the highest concentration (0.45 m). 34 Ashby4° has interpreted the results for tetrahydrofuran solutions as being due to solvent-separated ion pairs in dilute solution, while triple ions are formed at higher concentrations. Lithium aluminum hydride in diethyl ether appears to exist as contact ion pairs in equilibrium with higher aggregates. The better solvating ability of tetrahydrofuran and its higher dielectric constant (7.6 versus 4.34) is consistent with these interpretations. B. The Mechanism pf Lithium Aluminum Hydride Reduction pf The kinetics of ketone reductions by lithium aluminum hydride has been followed spectrOphotometrically by observing the disappearance of the n 9 n. band in the ultraviolet. In tetrahydrofuran the reduction is first order in both ketone and lithium aluminum hydride.‘1 In diethyl ether the reduction is first order in ketone, but about 1/2 order in lithium aluminum hydrids.‘° In diethyl ether this indicates participation of both monomeric and dimeric species. The role of the metal ion appears to be important but not essential for reaction. For example, Ashby has reported that lithium aluminum hydride is about ten times more reactive than sodium 1 borohydride.‘ He has also used tri - p - octyl -‘p - prepylammonium aluminum hydride as an effective reducing agent. Chelating agents, such as N, N, N', N", N", N. - hexamethyltriethylenetetraamine (for lithium ion) and 18 - crown - 6 (for sodium ion), do not inhibit the reduction. However, it appears that the metal ion is involved in the transition state, especially since different metals can show marked changes in the stereochemistry. The fact that mesityl phenyl ketone 35 is about 40 times more reactive in diethyl other than in tetrahydrofuran can also be attributed to the metal ion. That is, since tetrahydrofuran can solvate cations to a greater extent than diethyl ether, the ketone can compete more favorably for the metal ion in diethyl ether. The enthalpies of activation for the reduction of mesityl phenyl ketone by lithium aluminum hydride and sodium aluminum hydride are 10.5 and 18.1 kcal/mole respectively.‘1 The corresponding entropies of activation are -26.2 e.u. for lithium aluminum hydride and -5.4 e.u. for sodium aluminum hydride. This is consistent with cation association of the carbonyl oxygen in the transition state. The lithium cation can polarize the carbonyl to a greater extent, thus lowering the enthalpy of activation relative to the sodium cation. The more negative entropy of activation for lithium is consistent with the ability of that cation to order the ketone and a number of solvent molecules to a greater extent than the sodium cation. Ashby has also suggested that the lithium cation may be associated with one of the non - reacting hydrogens of lithium aluminum hydride, whereas the sodium cation would not. Kinetic isotOpe effects have been calculated for the reduction of mesityl phenyl ketone in diethyl ether and tetrahydrofuran. In both ‘1’41 Other than demonstrating that the cases the kH/kD was about 1.3. hydride is transferred in the transition state, it is difficult to interpret these results because of secondary isotope effects and the uncertainty of the Al-H(D)-C bond angle. Ashby has prOposed the following mechanism based largely on these 36 observations (Eq 7).‘1 S \lfl AIH: + C=0 =———‘ ./ S S a \ .. /CK\ AIH. + 8 ——~ 3 o\ \C/ / “i l s) [3,3 8 H H \C=0/§\S \ ./ \ / , , - —‘ S + LI Al (7) /: : H / \O/ \ H-A'I/ S I H /\ C-—H H H /\ The role of alkoxyaluminohydrides, has been investigated by Smith‘3 and Brown.‘3'4‘ It has been established by Smith that the transfers of the second, third, and fourth hydrogens from aluminum hydride are slower than the first transfer. However, these rate differences are not great enough to disregard the role of the alkoxyaluminohydrides, and may be significant when the ketone/lithium aluminum hydride ratio is close to one. Smith's studies were done on tertiary butoxy derivatives. Brown has determined that secondary alkoxyaluminohydrides (products from ketone reductions) disproportionate rapidly to regenerate aluminum hydride plus the tetraalkoxyaluminate. If the alkoxy group is not secondary, the disproportionation is much slower. 37 C. (Th2 Mechanism pf Sodium Borohydride Reduction pf Ketones ip Alcohols The composition of sodium borohydride in hydrolytic solvents is not known. It is generally assumed that it is strongly solvated and it is known to react with some alcoholic solvents to form alkoxy intermediates. Brown has reported that it is stable in isoprOpyl alcohol at temperatures as high as 60°C.‘5 The rate of reduction has been determined to be first order in “"’ Wigfield“, has ketone and first order in sodium borohydride. also determined the rate to be 3/2 order in alcohol. This has been explained by assuming that an alkoxide (from ionization of the solvent) and a solvent molecule are involved in the transition state. Wigfield reported that the products of reduction are the free alcohol, derived from the ketone, and the alkoxy borohydride, in which the alkoxy group is derived from the solvent. He has proposed the following mechanism (Eq 8). RO ------ B ----- H ----- c--—--=o ----- H ----- 0R (.9) The sodium ion is not included in the transition state. Studies“"’, including the removal of sodium cation by cryptatesso, have shown that this is so. Unlike lithium aluminum hydride and its alkoxy derivatives, the 38 transfer of the second, third, or fourth hydrogen of sodium 46,51 borohydride is faster than the initial reduction step. Secondary and tertiary alkoxy derivatives of sodium borohydride are fairly stable to disproportionation.’z"3 D. 132 Activation Parameters The activation parameters calculated define the relationships - AH? enthalpy favors attack from the least hindered side to give the cis a: - a: and (Ase-ii Astrans)° (Aniis rans) When R is a methyl group, isomer. When R is an ethyl or isopropyl group, there is a crossover and enthalpy favors attack of the more hindered side. Attack of the more hindered side is favored more for isOpropyl than for ethyl. In the case of pfbutyl, there is a slight preference for attack from the more hindered side in diethyl ether, essentially no preference in tetrahydrofuran, and a slight preference for attack from the least hindered side in isopropyl alcohol. In all cases studied, entrapy favors attack from the least hindered side to give the pig isomer. When R is a prbutyl group, the attack is favored strongly. No activation parameters could be calculated for the phenyl substituted case. The cis isomer is essentially quantitatively favored in all three systems. 1. Ihg.gplg pf Enthalpy It is interesting to note that the trend observed by Varech and Jacques" and later by Felkin and co-workers37 is indeed due to enthalpy factors. Felkin and co-workers proposed an anisotropic inductive effect to rationalize the stereoselectivity. An effect such .—__. §__._.._ __- 39 as this would be apparent in the enthalpies of activation. Intuitively, one would expect the two faces of any of the ketones studied to have slightly different electron densities. The importance of this in the transition state would be difficult, at best, to predict. It should be noted in this regard, that calculations of the electron density on the diastereotopic faces of carbonyls have been done." These calculations predict just the opposite to an anisotrOpic inductive effect when comparing a methyl to a hydrogen. That is, the face pppi to the methyl group has less electron density than the face pppi to the hydrogen. It would also require that the transition state be very "reactant like". In other words, the orbitals of the n system must still be largely intact. It is noteworthy to point out that the results in this study do not agree with Ahn's antiperiplanar effect. An antiperiplanar effect would also show up in the enthalpies of activation. The geometries of the transition states in these ketones would most likely not be identical to those proposed by Ahn, but would appear to be similar. It would also seem unlikely that any small distortions in the bicyclic skeleton could account for the stereochemistry. There are no eclipsing interactions of substituents on the ring system, except for the carbonyl oxygen and the bridgehead hydrogen. It would appear that such an interaction would not interfere, since in the ground state carbonyls prefer an eclipsing interaction.‘4 There are some features in this system (bicyclo[2.2.2]). however, which set it apart from other ketones. First of all, the incoming hydride must eclipse another bond in the transition state. This would 4O certainly raise the enthalpy of activation compared to acyclic or simple monocyclic ketones. Another feature is that the oxygen of the carbonyl must, in the course of the reaction eclipse another bond. Only if the carbonyl remained trigonal in the transition state, which seems unlikely, could eclipsing be ignored. Again, this eclipsing would add to the enthalpy of activation, especially since the oxygen would need a metal counterion and/or a high degree of salvation. Thus, the overall enthalpies of activation would, to a first approximation, most likely be greater than in more flexible systems. Even without eclipsing considerations, the system itself is fairly hindered. It is reasonable to assume that the transition states would be further along the reaction coordinate and thus resemble products more so than for "normal" ketones. A comparison of the eclipsing interaction of a hydride with a methyl group versus a hydride with another hydrogen is in order. Competitive rate studies of the two molecules in Figure 14 have been carried out by Eliel and Senda.17 W1 W111 l.2£§ l.0 Figure 14. 4 -‘£ - butylcylcohexanone and 2,2 - dimethyl - 4 - p - butylcyclohexanone. They determined that equatorial attack of 4-pfbutylcyclohexanone 41 by lithium aluminum hydride is only 1.25 times faster than equatorial attack of 2,2-dimethyl-4-prbutylcyclohexanone. If tetrahydrofuran is used as the solvent instead of diethyl ether, this value is even smaller (1.15). It is apparent that the axial methyl group does not affect equatorial attack nearly as much as one might expect based on steric hindrance. In fact, when lithium pgiéprbutoxyaluminohydride is used to reduce these two compounds in tetrahydrofuran, equatorial attack is still favored for 4-prbutylcyclohexanone by only a factor of 1.5. Returning to the bicyclo[2.2.2] system, an analagous type of situation is occurring. The torsional strain due to an eclipsing hydrOgen is of the same order of magnitude as the steric hindrance due to a methyl group. Furthermore, in the series methyl, ethyl, isopropyl, the situation should not change significantly since there will always be a pathway which would in a sense ignore the "extra" methyl groups. This pathway is depicted in Figure 15. Figure 15. The bicyclo[2.2.2] system. On the side of the carbonyl opposite to the attack, the oxygen begins to eclipse either a hydrogen or an alkyl group. There is no 42 question that it would prefer to eclipse a hydrogen. The thermodynamic equilibration ratios bear this out (Table 5). Furthermore, these values indicate that the "extra" methyl group in the series is noticed. In other words, the reason that complex metal hydrides attack the more hindered side is the unfavorable eclipsing interaction of the oxygen with the alkyl group when the hydride attacks from the least hindered side. This implies that the interaction of the oxygen with the alkyl group is more significant than the interaction of the hydride with the alkyl group. In fact this is not unreasonable since complex metal hydrides invariably act as "small reagents" when compared to other nucleOphiles. Equations 9 through 14 illustrate this point.55 Q) 00) {l l) 02) ’ 0H /0 W HCECMQBF CECH (l3) EIZO 0H CHJMgBr Ma’s (14) 57‘20 43 If the oxygen-alkyl group interaction is greater than the metal hydride-alkyl group interaction then the transition state resembles the products to a certain extent. There is evidence to suggest that rehybridization of the carbonyl to sp3 is significant. For example, Hammett studies have shown that reaction of sodium borohydride with "57 and acetophenones” gives large positive substituted fluorenones‘ p values of 2.65 and 3.06 respectively. The corresponding value for substituted benzophenones with lithium aluminum hydride is 1.95.‘2 Lamaty and Geneste have argued that the high values for the sodium borohydride reductions indicate that such reactions pass through product-like transition states." Wigfield has reported an interesting difference between the sodium borohydride reductions of unhindered and hindered cyclohexanones.‘7 Specific isokinetic plots for axial and equatorial attack were constructed. (The enthalpy of activation for axial attack was plotted against the entropy of activation for axial attack. The same was done for equatorial attack.) These plots showed a clear cut difference in isokinetic behavior between hindered and unhindered cyclohexanones. It has been argued by Leffler that such a distinct difference corresponds to a qualitatively different ‘transition state.‘0 Isokinetic relationships have been criticized‘l, but Wigfield's plots seem to indicate a different transition state for hindered cyclohexanones versus unhindered ones. Wigfield has also noted a similar phenomenon in a kinetic isotope effect study.‘z Reduction of a number of ketones with lithium tri - p - butoxyaluminohydride(deuteride) gave isotOpe effects ranging from 44 1.0 to 1.5. The corresponding isotope effects for two highly hindered ketones, 3,3,5,5 - tetramethylcylcohexanone and 4 - acetyl - 3,3,5,5 - tetramethylcyclohexanone are inverse, 0.79 and 0.70, respectively. An inverse isotOpe effect can be explained by invoking a late transition state since a weaker Al-H(D) bond is being replaced by a stronger C-H(D) bond. The isotope effects for sodium borohydride reductions 59,63,64 are inverse (kn/kn = 0.59 - 0.77). The inverse isotope effects are difficult to interpret since the primary effect is masked by three secondary effects. Both an early and a late transition state have ”“ Lithium aluminum hydride has been reported to have been proposed.‘ an isotope effect which is normal and small (kH/kD = 1.3).‘3 Again, this result is difficult to interpret. The trend in stereoselectivity for the bicyclo[2.2.2] system can be explained by invoking a late transition state. Product stabilities are reflected in the transition state. The prbutyl case is consistent with this because the hindered side is effectively blocked and steric approach control is dominant. However, even in this case a significant prOportion of attack is from the hindered side, thus product stability is still important. The phenyl substituted system appears out of line since it favors the .21; product almost exclusively. In examining the following two systems (Eqs 15,16), however, it is apparent that a n system can hinder attack.3"‘5 45 C) LIAIH4 05) EQO @700»; 307° H OH O NaBH HO & (I 6) / i'—Pr0H / . Iqu Presumably this is due to a coulombic repulsion between the high electron density of the n system with the negatively charged nucleophile. 2. TE; 3215 pf Entropy Entropy favors the pig product in all reductions. This effect is more pronounced for the prbutyl group than for the other alkyl groups. The difference in results for diethyl ether and tetrahydrofuran as solvent is significant. These results are most likely due to solvent entropy effects. That is, in the pig case the carbonyl oxygen starts eclipsing the alkyl group; the disordering of solvent molecules around the oxygen metal bond is greater than in the pgppg case. This explains why the pfbutyl group has such a larger AASI: It cannot rotate out of the way as ethyl and isopropyl can. This also explains the difference between reductions in tetrahydrofuran and diethyl ether. In tetrahydrofuran, the cation is solvated to a greater degree and thus the effect is magnified. In isopropyl alcohol, the cation is 46 not involved in the transition state, thus salvation is apparently not as important. This interpretation is consistent with a late transition state. It is difficult to assess the role of rotational entropy for the alkyl groups. A late transition state would hinder rotation for both routes of attack. 3. Conclusions The results of this study show that the observed trend in stereoselectivity in the bicyclo[2.2.2]-octanone system is due to enthalpy factors. The enthalpy factors which control this selectivity can be attributed to product stability. This implies that the transition state resembles products to a certain extent. It does not imply that this is true in all cases. This appears to be a special case. The results of this study also show the importance of considering salvation in evaluating entropy effects. EXPERIMENTAL I. General Methods A. fippgk solutions pf complex ppppl hydrides Diethyl ether and tetrahydrofuran were dried over lithium aluminum hydride, distilled and stored over molecular sieves (4A). Is0pr0pyl alcohol was dried over calcium sulfate and distilled. Stock solutions of lithium aluminum hydride were prepared as follows. Lithium aluminum hydride (Aldrich, 95%) and the solvent were stirred together for 1h under an argon atmosphere. The crude slurry was then siphoned through a U-shaped glass tube, which contained a glass wool plug, into a Schlenk filter tube packed with CeliteR. After filtering, the clear solution was stored under argon. Lithium aluminum deuteride and lithium ppitprbutoxyaluminohydride solutions were made by stirring them, under an argon atmosphere, in solvent. Sodium borohydride solutions were made by stirring the hydride in isoprOpyl alcohol until all material was dissolved (several hours, a drying tube was used to keep moisture out). The solutions made in this manner were between 0.05 and 0.10 molar. B. Temperature Control Reductions done at -77°C were controlled by using a dry ice-acetone bath. A dry ice-carbon tetrachloride bath and an ice water bath were used for reductions at -22°C and 2°C, respectively. Temperatures at reflux were controlled using an ordinary oil bath. 47 48 C. General Procedure jg; Reductions Method A: A solution of the complex metal hydride (8-12 mL, 0.5-1.0 mmol) was put into a 25 mL round-bottomed flask equipped with a magnetic stir bar, gas inlet valve and septum. The solution was allowed to come to thermal equilibrium under an argon atmosphere. In a separate 25 mL round-bottomed flask equipped with a gas inlet valve and septum were combined the ketone (20-30 mg, 0.07-0.12 mmol) and 5 mL of solvent. If the reduction was run below room temperature the ketone solution was also cooled to the reaction temperature. The ketone solution was then slowly added (10-15 min) to the complex metal hydride solution by using a syringe. If the reduction was run at reflux, a reflux condensor was connected to the flask. The following reaction times were sufficient for complete reduction: refluxing diethyl ether, tetrahydrofuran and isopropyl alcohol (10 min); 22°C diethyl ether, tetrahydrofuran and isopropyl alcohol (15 min); -22°C diethyl ether and tetrahydrofuran (0.5h), isopropyl alcohol (1h); -77°C diethyl ether and tetrahydrofuran (2h), isopropyl alcohol (6h) and very slow for the p-butyl system. The solution was hydrolyzed with 20 mL of water and extracted with three 30 mL portions of ether. The ether extracts were combined and dried over anhydrous magnesium sulfate, filtered and the solvent removed under reduced pressure. The crude product mixture (>90% yield) was used for NMR spectra. Method B: Using the identical experimental set—up as Method A, the metal hydride solution was added to the ketone solution (inverse addition). Care was taken to initially add the metal hydride very slowly. Other than that, reaction times were the same. 49 Method A was used throughout the study. Method B was used periodically to see if inverse addition would affect product ratios. No such affect was detected. II. Syptheses 2-Nitroethanol (1) 2-Nitroethanol was prepared according to the procedure of Noland.“ Paraformaldehyde (75.0 g, 2.5 mol) and nitromethane (1.70 kg, 28.0 mol) were combined in a 3 L, three-necked round-bottomed flask equipped with a mechanical stirrer, a dropping funnel and a thermometer. The mixture was vigorously stirred and a solution of potassium hydroxide in methanol (3 ‘3) was added dropwise until an apparent pH of 8 (pH paper). The paraformaldehyde completely dissolved and the solution temperature increased to 40°C. Stirring was continued for 1h after which concentrated sulfuric acid (18 M) was added drapwise until an apparent pH of 4 (pH paper). Stirring was continued for an additional hour. Precipitates, formed during the acidification, were removed by filtration and the nitromethane removed under reduced pressure. The crude 2-nitroethanol was vacuum distilled with an equal volume of diphenyl ether (heat dispersing agent), bp 59-650C (0.20 mm). The two phase distillate was separated and the crude 2-nitroethanol (114.0 g, 50%) was used without further purification; ‘H NMR (CDCl,) 5 4.60-4.25 (2a, br m), 4.20—3.80 (2a, br m), 3.50 (1H, br s). 50 l-Nitro-Z- ro anol (2) l-Nitro-Z-propanol was prepared according to the procedure of Hard and Nilson." Nitromethane (183.0 g, 3.0 mol) and 155 mL of water were combined in a 2 L, three-necked round-bottomed flask equipped with a mechanical stirrer, drapping funnel and a reflux condensor. To this was added enough sodium carbonate to give an apparent pH of 10 (pH paper). A solution of acetaldehyde (132.0 g, 3.0 mol) and 120 mL of water was added dropwise to the solution over a 90 min period, during which time the reaction temperature increased to 60°C. Sodium carbonate was added periodically to maintain a pH of about 10. The solution was stirred for 2h and then let stand for an additional 2-3h. The contents of the flask were poured into a separatory funnel and extracted with three 200 mL portions of ether. The combined ethereal solution was dried over anhydrous sodium sulfate, filtered, and the solvent removed under reduced pressure to afford the crude 1-nitro-2-propanol. Vacuum distillation gave pure l~nitro~2~propanol (214.2 g, 2.04 mol, 68%), bp 76—77°c (1.5 mm). [1it" bp 86—89°c (8 mm)]; ‘H NMR (C001,) 6 4.60—4.25 (3H. br m). 3.20 (1H, br s), 1.21 (3H. d). 1-Nitr012-butppplqigl 1-Nitro-2-butanol was prepared in the same manner as described for the preparation of 1-nitro-2-propanol.‘7 The crude nitro alcohol was purified by vacuum distillation, 64% yield, bp 75-76°c (1.1 mm). [lit‘7 98-99°c (9 mm)]: ‘H NMR (CDCl,) 8 4.47-3.93 (3a, br m). 3.02 (1H, br s). 1.80-1.32 (2H, m), 1.03 (3H, t). 51 3-Methyl-l-nitro-2-butanol (4) Using the same reaction conditions as previously described for 1-nitro-2-pr0panol (2) and 1-nitro-2-butanol (i). 2-methylpr0panal (iso-butyraldehyde) was condensed with nitromethane. The crude nitro alcohol was purified by vacuum distillation, 56% yield, bp 79-80°C (1.0 mm); 1H NMR (cnc1,) 8 4.48-4.00 (3a, br m). 3.05 (1H, br s). 1.85-1.36 (1H. m). 1.23 (3H. d). 1.12 (3H. d). Nitroethylene (5) Nitroethylene was prepared according to the procedure of Buckley and Scaife.fl 2-Nitroethanol (100.0 g, 1.10 mol) and phthalic anhydride (180.0 g, 1.22 mol) were combined in a vacuum distillation apparatus with a short fractionating column and a stir bar. Under reduced pressure (80 mm) the mixture was heated to approximately 170°C (oil bath temp 170-175°C). A two phase distillate of water and nitroethylene was collected. The nitroethylene was separated from the water, dried over anyhdrous magnesium sulfate, filtered and redistilled (34.5 g, 43%). bp 42-44°c (88 mm), [1it" hp 38—39°c (80 mm)]; 1H NMR (CDCl3) 6 7.55-6.68 (m). The nitroethylene was used immediately. 1-Nitropropepp‘jgl l-Nitro-Z-propanol was dehydrated as described for the dehydration of 2-nitroethanol." The two phase distillate of crude product and water was collected at an oil bath temperature of 180°C (80-100 mm). The layers were separated and the crude product was 52 dried over anhydrous magnesium sulfate, filtered, and vacuum 1 distilled, 65% yield, hp 50-51°c (80 mm). [lit" hp 54°C (88 mm)]; H NMR (CDC13) 6 7.60-6.67 (2H, m). 1.88 (3H, d). l-Nitropropene prepared in this manner is predominantly the ppgpg isomer based upon comparison with identical spectra published by Baskov and co-workers." The pig isomer is present in about 5% by proton NMR integration."0 pgpps-l-Nitro-l-bppgpg‘11) ppppg-l-Nitro-l-butene was prepared in the same manner as nitroethylene and l-nitrOpropene. Heating to an oil bath temperature of 180°C (80-100 mm) resulted in only partial dehydration with significant amounts of 1-nitro-2-butanol being collected. After removing the water from the two phase distillate, the mixture of product and starting material was resubjected to the dehydration conditions. The crude nitro alkene was purified by vacuum distillation, 47% yield, hp 58-59°c (75 mm); 1H NMR (CDCl,) 8 7.50-6.65 (2H, m). 2.58-2.02 (2H. m). 1.10 (3H, t). l-Nitro-l-butene prepared in this manner is assigned the ppppp configuration since the olefinic pattern in the proton NMR is similar to those published by 69 Baskov and co-workers. trans-3-Methyl-l-pitro-l-butene (8) The dehydratrion of 3-methyl-1-nitro-2-butanol was carried out as described for the dehydration of l-nitro-Z-butanol. The nitro alkene was purified by vacuum distillation, bp 54-55°C (55 mm); 1H NMR 53 (CDCl,) 6 7.28 (1H, d of d, J= 16 and 7 Hz), 6.98 (1H, d, J= 16 Hz), 2.64 (1H, octet). 1.00 (6H, d). trans-B-Nitrostyrene (9) pgppgrB-Nitrostyrene was prepared by the method of Worrall."1 Nitromethane (61.0 g, 1.0 mol), benzaldehyde (106.0 g, 1.0 mol) and 100 mL of methanol were combined in a 2 L, three-necked round-bottomed flask equipped with a mechanical stirrer, dropping funnel and thermometer. After cooling (0°C) and with stirring, a solution of sodium hydroxide (42.0 g, 1.05 mol) and ice water containing crushed ice (100 mL total volume) was slowly added at such a rate that the temperature was kept at 10-15°C. During the addition a white precipitate formed and stirring became difficult (20 mL of methanol was added to aid stirring). After 15 min of standing, 600 mL of ice water containing crushed ice was added. The cooled solution (0-5°C) was then immediately added, with stirring, to 500 mL of hydrochloric acid (7.2 M). A yellow crystalline mass was formed as the solution came in contact with the acid. The solid mass was suction filtered and washed several times with water. The crude nitrostyrene was purified by two recrystallizations from ethanol (110.3 g, 0.74 mol, 74%), mp 57-58°c, (lit'1 mp 57-58°C): ‘H NMR (CDCl,) 5 7.29 (5H, m). 7.21 (1H, d, J= 14 Hz). 6.93 (1H, d, J= 14 Hz). 9,10-Dihydro-9,10-(trpps-ll-pitro-lZ-methylethpno)-anthracene (10) 9,10-Dihydro-9,lO-(trans-11-nitro-12-methylethano)-anthracene was prepared according to the method of Noland and co--workers."z 54 Anthracene (30.0 g, 0.168 mol) and 50 mL of prdichlorobenzene were combined in a 200 mL round-bottomed flask equipped with a magnetic stir bar and addition funnel. The flask was cooled (0°C), evacuated and purged with argon. 1-Nitropropene (3.7 g, 0.042 mol) and 15 mL of pfdichlorobenzene were quickly added to the addition funnel under an argon atmosphere. The solution was heated to reflux and with stirring, the nitro alkene was added dropwise over a 5 min period. The solution was refluxed for 80 min and after cooling the solvent was removed under reduced pressure. The residue was dissolved in boiling benzene and upon concentrating and cooling, anthracene was removed. The crude residue was chromatographed on alumina (300 g) packed wet with petroleum ether (60-70). Elation with 10% ether-petroleum ether (60-70) removed anthracene. Elution with methanol gave the crude product (6.95 g. 0.0262 mol, 62%), mp 115—119°c. Two recrystallizations from ethanol gave pure adduct (6.50 g, 0.0245 mol, 58%), mp 122-123°c, (lit" mp 122.5-123.5°C); 1H NMR (CDCl,) 8 7.28-6.88 (8H, m). 4.80 (1H, d. J= 2.6 Hz). 4.10 (1H. d of d, J= 4.6 Hz and 2.6 Hz). 3.95 (1H, d, J= 2.2 Hz). 2.64 (1H, m), 1.02 (3H. d. J= 7.0 Hz). 55 9,10-Dihydro-9l10-5ll-Keto-lngethylethanoz-Anthracene.Llll 9,10 - Dihydro - 9,10 - (nggg - 11 - nitro - 12 - methylethano) - anthracene (2.45 g, 9.24 mmol) in 250 mL of ether was combined with 225 g of activated basic silica gel"3 (chromographic grade). Ether was removed under reduced pressure and the dry silica gel was warmed to 60°C for 7-8h. After cooling, the silica gel was eluted with ether to give 1.86 g of crude 9,10 - dihydro - 9,10 - (11 - keto - 12 - methylethano) - anthracene. The solid was chromatographed on alumina (55 g) with 50% ether-petroleum ether (60-70) to give the pure ketone (1.44 g, 6.15 mmol, 67%). mp 2121-122°c, (11t” mp 122°C); 1H NMR (CDC13) 5 7.34-6.88 (8H. m). 4.78 (H—lO, s). 4.28 (H-9, d, J= 2.4 Hz). 2.29 (H-12, d of q, J= 7.3 and 2.4 Hz). 0.94 (3H, d, J= 7.3 Hz). The residue. which was left on the chromatography column was mostly the starting nitro adduct which was recycled. 9IlO-Dihydro-9.10-(trags-ll-Nitro-lg-Ethylethanoz-Anthracene (12) The Diels-Alders reaction of anthracene and tgggg-l-nitro-l-butene was carried out in the same manner as described for anthracene and 1-nitropropene. The crude adduct was chromatographed, 7% yield, but not purified further; 1H NMR (CDCl,) 8 7.42-7.00 (33, m). 4.92 (H-10, d), 4.38-4.17 (H411 and H49, 56 superimposed), 2.54 (H-12, m), 1.24 (2H, quintet). 0.94 (3H, t). H\,CH2CH3 HHCH‘QCHJ OZN H at. + -— A 999 H a ,2 Q .7 — 9,10-Dihzdro-9,19-(ll-Keto-l2-Eth lethano -Anthracene (13) The oxidation of the nitro adduct to the ketone was carried out on activated basic silica gel as described for the 11-nitro—12-methylethano adduct, 1;. 59% yield, dec. 9s—11o°c, mt” dec. 95-1100C); 1H NMR (CD013) 8 7.44-7.00 (8H. m). 4.77 (H-lO, s). 4.46 (H-9, d), 2.10 (H-12, octet), 1.62 (2H, m), 1.03 (3H, t). H C H2CH3 H CH(CH3)2 + _ . @@@ OzN > H a Q 8 I4 --' ”meg. ...,, - 57 9,10-Dih1dro-9,10-(traps-ll-Nitro-IZ-iso-Propylethano)-Anthracene (142 The reaction between anthracene and trans-3-methyl-l-nitro—l-butene was carried out as described for anthracene and 1-nitr0pr0pene. The crude adduct was chromatographed, 4% yield, but not purified further; 1H NMR (CDCl,) 8 7.40-6.95 (83. m), 4.82 (H910, d), 4.32-4.13 (H911 and H99, superimposed), 2.52 (H912, m), 1.75 (1H, m), 1.10-1.00 (6H, two d superimposed). 9,10-Dihydro-9,10-gll-Keto-lz-iso-Propylethanoz-Anthracene (15) The oxidation of the nitro adduct to the ketone was carried out on activated basic silica gel as described for the 11-nitro-12-methy1ethano adduct, 1;, 55% yield, mp 128-129°c, (lit” mp 129°C); 1H NMR (cnc1,) 5 7.35-6.92 (8H. m). 4.66 (H-lO, s), 4.48 (H99, d). 2.00 (H912, d of d). 1.40 (1H, m). 0.97 (3H, d). 0.76 (3H, d). 9I10-Dihydro-9,10—(trags-ll-Nitro-lZ-Phen lethano -Anthracene (l6) 9,10-Dihydro-9,10-(trans-ll-nitro-lZ-phenylethano)-anthracene was synthesized as described by Noland and co-workers.12 B-Nitrostyrene (6.0 g, 0.040 mol), anthracene (14.6 g, 0.088 mol) and 100 mL of xylenes were combined in a 250 mL round-bottomed flask equipped with a 58 magnetic stir bar and a reflux condensor. Under an argon atmosphere the solution was refluxed for 20h. After cooling, the solvent was removed under reduced pressure. The residue was extracted six times with boiling ether (50 mL each). The combined extracts were concentrated and the crude residue chromatographed on alumina (300 g) with 10% ether-petroleum ether (60-70) to remove traces of anthracene. Elution with methanol, followed by recrystallization from ethanol gave pure _1_6 (4.91 g, 0.015 mol, 38%), mp 150-151°c. (111:3’ mp ISO-151°C); ‘H NMR (0001,) 6 7.40-6.78 (11H, m), 6.60-6.35 (2H, m), 5.10 (a—10, d, J= 2.7 Hz). 4.90 (H911, d of d. J= 4.7 and 2.7 Hz). 4.30 (H99. d. J= 2.2 Hz), 3.91 (H-12, d of d, J= 4.7 and 2.2 Hz). H CsHs ~06 H>=:<: _‘ 9 lO-Dih dro-9 10- 11-ketoethano -Anthracene (l9) The conversion of the nitro adduct to the corresponding ketone has been described, 11. The crude ketone was chromatographed on 60 alumina (50 g/ 1.5 g ketone) with 50% ether-petroleum ether (60-70) to give pure 12. mp 150-151°C, (1it” mp 151°C; ‘3 NMR (CDC1,) 6 7.40-6.92 (8H, m). 4.70 (H910, s), 4.45 (H—9, t, J= 2.5 Hz). 2.27 (2H, d, J= 2.5 Hz). H H H H GEN F1 C> g_© “ fl_© 9,10-Dihydro-9,10-(ll-Trimeth lsilox ethen 1 -Anthracene (20) To a 25-mL round-bottomed flask, equipped with a rubber septum, gas inlet valve and a stir bar were added 4.0 mL of a hexane solution of grbutyllithium (1.56 g, 6.24 mmol) and 4 mL of dry hexane. After cooling (0°C), dry diisopropylamine (0.90 mL, 0.65 g, 6.38 mmol) was slowly injected by syringe into the reaction flask. The resulting solution was stirred for 15 min at 0°C, after which the solvent was removed under reduced pressure. To the resulting white powder was added 7 mL of dry tetrahydrofuran. Ketone 12 (1.37 g, 6.24 mmol) in 4.5 mL of tetrahydrofuran was slowly added to the lithium diisopropylamide solution by means of a syringe. The resulting deep purple solution was stirred at 0°C for 10 min after which chlorotrimethylsilane (0.77 mL, 0.71 g, 6.55 mmol) was slowly injected into the reaction mixture by syringe. The resulting solution was stirred at room temperature for 15 min, diluted with 50 mL of hexane and washed twice with 30 mL of.10% cold aqueous sodium biéarbonate. 61 The organic layer was dried with anhydrous magnesium sulfate, filtered and the solvent removed under reduced pressure to give crude 20; 1H NMR (CDC13) 5 7.74-7.42 (4H. m). 7.28-7.15 (4H, m). 6.01 (H911, d of d. I: 16 and 2 Hz). 5.08 (H-12, d, J= 16 Hz). 4.88 (H—9, d, J= 2 Hz). 0.08 (9H. s). H (mama-ob H a (9 £2 9,10-Dihydro-9,10-(11:5eto-12-t-bgtylethanoz-Anthracene (2;) The silyl enol ether, ‘20, was converted to 2;, using Chan's £9buty1 alkylation procedure."4 Freshly distilled titanium tetrachloride (2.27 g, 1.32 mL, 12.0 mmol), £9butylchloride (0.56 g, 6.0 mmol) and 15 mL of dry methylene chloride were combined, under argon, in a 25 mL round-bottomed flask equipped with a magnetic stir bar, gas inlet valve and dropping funnel. After cooling the solution to -40°C and with stirring, the crude silyl enol ether, 29, (1.75 g, 5.99 mmol) in 5 mL of methylene chloride was added dropwise to it. The mixture was stirred for 3h at -40°C and then hydrolyzed with 5 mL of saturated sodium carbonate solution. The resultant mixture was extracted with 50 mL of carbon tetrachloride and washed with three 25 mL portions of sodium carbonate solution. The organic layer was dried with anhydrous magnesium sulfate, filtered and evaporatated under reduced pressure. Two recrystallizations from ethanol gave pure 2; m-.. _.q_.. 62 (0.92 g, 3.48 mmol, 58%), mp 131—132°C, (1iu3’ mp 132°C); 1H NMR (CDC1,) 6 7.45-7.05 (83, m), 4.80 (n+10, s), 4.69 (n+9, d), 2.16 (H-12, d). H H C(CH3)3 (CH3)aSi-O L o 32 ::J a :::’ Eguilibration _j the the Qiasteregmgric Algghglg.(ggyglrgagtiggl The equilibration of the alcohols was carried out as described by Eliel.1’ The mixture of diastereomeric alcohols (0.10 g), acetone (0.08 g), aluminum isopropoxide (0.10 g) and iSOprOpyl alcohol (7 mL) were combined in a 25 mL round-bottomed flask equipped with a magnetic stir bar, reflux condensor and drying tube. The solution was refluxed for 7d, after which it was quenched with 10 mL of 35% cold aqueous sulfuric acid. The resulting solution was extracted with three 25 mL portions of ether. The combined ethereal extracts were washed with aqueous sodium bicarbonate, water, and brine and dried over anhydrous magnesium sulfate. After filtering the solution the ether was removed under reduced pressure. The residue was analyzed by NMR. III. Product Ratio Analysis The product ratios (cis/trans) were determined by proton NMR peak integration. In the case of ketone 11, the ensuing diastereomers, 2; and 2;, were assigned the following chemical shifts: 22 (CDCl,) 8 63 4.38 (H-lO. d. J= 2.7Hz). 4.12 (H-ll. d of d. I= 8.8 and 2.7Hz). 3.96 (H99, d, J= 1.5Hz). 2.21 (H912, m). 0.74 (’CHS, d); 2; 4.25 (H—lO, d. J= 2.8). 3.91 (H-9, d, J=1.5Hz). 3.53 (H-ll. br t. J= 3 Hz), 1.60 (H-12, m). 0.90 (-CH3, d). The coupling constant of 8.8 Hz for H—ll of 2; is consistent with cis stereochemistry. Likewise, the coupling H CH3 H CH3 H CH3 <3 H (Ni PK) H constant of 3 Hz for H-ll of 23 is in line with Egggg stereochemistry. Reduction of 1; with lithium aluminum deuteride resulted in the loss of H-ll peaks, while H-10 peaks became singlets. Using homo-decoupling techniques, irradiation of 8 4.38 (H910, 22) produced a doublet at 8 4.12 (H911, 22, I= 8.8 Hz) and irradiation of 8 4.25 (H-10, 2;) resulted in a doublet at 8 3.53 (H-ll, 23, I= 2.7Hz). When the lanthanide shift reagent. tris(6,6,7.7,8,8,8 - heptafluoro - 2,2 - dimethyl - 3,5 - octanedionato) europium, (Eu(fod),). (5-50 mol per cent), was added step-wise with increasing concentration, the doublet at 8 0.74 was chemically shifted downfield a factor of three times that of the doublet at 8 0.90. Relative peak areas were consistent with the chemical shift assignments. Varech and Jacques reported H-ll chemical shifts for 22 and 23 of 8 4.12 and 3.5, respectively (no other chemical shifts were reported by them).3’ The H-10 peak areas were used to determine product ratios. 64 H CH2CH3 H CHpCHa H CHpCH3 Diastereomers 21 and 2§_ from 22 were assigned the following chemical shifts: 21 (CDCl,) 8 4.38 (H910, d, J= 2.5Hz), 4.16 (H99 and H-ll, superimposed; H-11, d of d, I= 9.0 and 2.5Hz; H—9, d, J= 1.5Hz), 1.98 (H912. m); g; 4.28 (H-10. d. J= 2.6Hz). 4.08 (H99, d, J 1.6Hz). 3.60 (H911, br m), 1.57 (H-12, br m). These assignments were found to be consistent in the same manner as tested for 22 and 22, except no lanthanide shift study was done. The chemical shifts for the methylene hydrogens were superimposed between 8 1.40 and 1.10. The two doublets for the methyl groups were superimposed between 8 1.09 and 1.01. The H—ll assignments were consistent with those reported.” The H-lO peak areas were used to determine product ratios. H CH(CH;)2 H CH(CH,)2 H CH-(CH,)2 o b ' H OH Ho H i ,. Q g .. Q 5.. Q Alcohols 2g and 21 were assigned the following chemical shifts: (‘CDC1,) 2_6 6 4.43 (H-lO, d, 2.0 Hz) 4.28 (H-9, superimposed with 11-9 of 21), 4.23 (H-11,J= 8.0 and 2.0 Hz), 1.80 (H-12, br m); 21 4.32 65 (H910, d, J= 2.0Hz), 4.28 (H99, superimposed with H-9 of 2g). 3.54 (H—ll, br t, I= 2.5 Hz). 1.63 (H912, br m). Homo-decoupling experiments and reaction with lithium aluminum deuteride were used to verify these assignments. Relative peak areas were consistent. The chemical shifts for the H-ll hydrogens are reported to be 4.23 and 3.54, for 2§_and 21, respectively.” The H-10 peak area of 2g and the H—ll peak of 21 were used to determine product ratios. H C(CH.). H C(CH3)3 H C(CH.), o H OH Ho H “)2, —° @_© @_© The gig and $5232 alcohols, 22 and 21, were give the following proton NMR assignments: (CDC13) 2§_8 4.31 (H911, br d, J= 9.0 Hz). 1.00 (-C(CH,)3, s); 22 3.91 (H911, br t, J= 3.5 Hz). 0.80 (-C(CH,),. s). The bridgehead hydrogens (H99,10) from both isomers were superimposed in the region 4.48 - 4.34. The H-12 hydrogen from 22 and _2 were also superimposed, 1.70 - 1.60. Varech and Jacques reported chemical shifts of 4.30 and 3.91 for the H-ll hydrogen of 2§_ and 22, respectively.” The relative peak areas at 1.00 and 0.80 were used to determine product ratios. For this reason the lanthanide shift reagent Eu(fod)3 was used to help verify the assignments. The step-wise addition of 5-50 mole per cent of Eu(fod)3 to the mixture produced downfield chemical shifts for the peak at 1.00 which were a factor of 3.5 times greater than the 0.80 peak. These assignments are 66 consistent with the relative peak areas for the H911 hydrogens. H C6H5 H CGHS H C6H5 The following chemical shift assignments were made for 2g and 21: (CDCl,) 19 8 4.62 (H910, d. J= 2.8Hz). 4.48 (H911, d of d. J= 10.0 and 2.8Hz). 4.43 (H-9, d, J= 1Hz), 3.44 (H-12, d of d, J= 10.0 and 1H2); 21 4.59 (H910, d, J= 2.6 Hz). 4.48 (H-9, d, J= 2.7 Hz). 4.20 (H911, br t, J= 3.8), 3.20 (H-12, br t, J= 3.8 Hz). These assignments were consistent with homo-decoupling experiments and reduction with lithium aluminum deuteride. As in the previous cases, the H-11 chemical shifts are consistent with those published. The spectra were taken on either a Bruker WH9180 or Bruker WP-250 FT NMR spectrometer. The product ratios were calculated by cutting and weighing peak areas. IV. 2552; Analysis The errors involved in the calculation of product ratios are due to an error in the temperature and errors in the method of intergrating peak areas. A temperature variation of t 20 contributes approximately one per cent to the error of a product ratio. The error in the integration of peak areas is dependent upon the signal to noise ratio of the spectra and human error in cutting and weighing peaks 67 (the method used here). Since the signal to noise ratio was quite high (solid baseline) and the results reproducible, this error was estimated to be no greater than i 2 per cent. At least two experiments were run at each temperature. The error in calculating the difference in enthalpies of activation was determined by using Equation 17.1"71 5 = 2RT'Ta/(T'-T) (u << 1) (17) The symbols in the expression are defined as follows: 8 is the error in AAH:F in cal/mole, R is the universal gas constant, T' and T are the temperatures at the extreme of the temperature range and a represents the maximum fractional error of the ratios. The maximum fractional error for all data was assumed to be 0.03. The maximum possible error (a) in AAS; is given by Equation 18 16,17 6 = fill/T + (T'—T)/2T'T] (u << 1) (18) APPENDIX —————_———_— The composition of Grignards has been the subject of considerable interest since their discovery in 1900. Through the years a number of structures have been suggested for the complexes. Today there is no doubt that a single composition can not represent Grignards. A number of studies, primarily by Ashby" , have shown that Grignards are best represented by a complex equilibrium (Eq 17), referred to as the extended Schlenk equilibrium. etc . =R3Mg :Mng-——‘R3Mg + ngz= 2RMgX =(RMgX) ;— etc . (17) From the organic chemisfs viewpoint, the various complexes and equilibria should be identified so that mechanisms and transition states can be more completely described. A better understanding of a Grignard's composition would be useful in explaining its stereoselectivity in addition reactions to ketones and aldehydes. Ideally, one would want to probe the composition of Grignards spectroscopically. A limited amount of knowledge has been obtained using infrared”, 1H nmr and 13C nmr'1 spectroscopy. There have been a few reports in the literature on z’Mg nmr spectroscopy'z"3, however, no such report concerns Grignards. Reported here are some z5Mg nmr results on Grignard systems. The chemical shifts (8) and line widths (Ayl/z) are reported in Table 9. No 1’Mg resonances were observed for Grignards in diethyl ether. There were also no 68 . 69 resonances observed for bis-(phenylethynyl) magnesium in tetrahydrofuran or for dimethyl magnesium or diethyl magnesium in either solvent. The nmr linewidth of a quadrupolar nuclide, such as magnesium, is generally determined by the interaction of the nuclide's quadrupolar moment with electric field gradients at the nucleus. Since Mg+3 is strongly hydrated in aqueous solution. electric field gradients at the nucleus are most likely very small because of the highly symmetrical arrangement of water molecules around the ion. Thus, for the following discussion it is assumed that the quadrupolar relaxation mechanism dominates the 3‘Mg relaxation thereby accounting for the observed linewidths. It is currently accepted that in diethyl other the Schlenk equilibrium favors RMgX and aggregated species." These species would be expected to have low electonic symmetry at the metal site and have some degree of covalent bonding to the magnesium. Therefore, large local electronic field gradients could be expected at the magnesium nucleus. This, plus the dynamic nature of such systems broadens the HMg lines beyond detection.. Magnesium bromide in diethyl ether would be expected to have higher electronic symmetry at the nucleus than a Grignard. A resonance for magnesium bromide in diethyl ether is observed (Table 9). Crystal structures of phenyl magnesium bromide“ and magnesium bromideu (diethyl ether solvated) ‘ are consistent with this interpretation based upon electronic symmetry. 70 In tetrahydrofuran, Grignard complexes are widely thought to exist primarily as RzMg plus aggregated forms.7' One would expect that the Rqu species (in whatever aggregated and solvated form) could be characterized by the same types of arguments proposed for RMgX in diethyl ether. No signals were observed for dimethyl magnesium, diethyl magnesium or bis-(phenylethynyl) magnesium in the temperature range 210 to 293°K. However, magnesium bromide was observed in tetrahydrofuran (Table 9) and its linewidth indicates high electronic symetry around the magnesium nucleus. Table 9. “Mg nmr Data for some Grignard Reagents and Magnesium Halides in Diethyl Ether and Tetrahydrofuran (THF). Compound T(°K) 5 (PM!)a ATi/z Magnesium Sulfate 1.0 g in Water 293 0 5 Magnesium Bromide 0.1 M in diethyl ether 293 47 131 Magnesium Bromide 0.1 _M in THF 293 9 20 Magnesium Bromide 0.1 M in THF ‘ 272 6 5 Magnesium Chloride 0.1 M in THF 293 9 230 Magnesium Chloride 0.1 M in THF 262 10 215 Phenyl Magnesium Bromide 0.1 M in THF 303 12 400 Phenyl Magnesium Bromide 0.1 M in THF 293 11 287 71 (Table 9 continued) Phenyl Magnesium Bromide 0.1 M in THF 273 7 33 Phenyl Magnesium Bromide 0.1 M in THF 250 6 19 Phenyl Magnesium Bromide 0.1 g in m 2101’ 5 43 Ethyl Magnesium Bromide 0.1 M in THF 301 9 224 Ethyl Magnesium Bromide 0.1.! in THF 280 7 33 Ethyl Magnesium Bromide 0.1 M in THF 260 5 14 Phenylethynyl Magnesium Bromide 0.1 M in THF 293 9 245 Phenylethynyl Magnesium Bromide 0.1 M in THF 263 6 23 Methyl Magnesium Bromide 0.1 M in THF 293 no resonance observed Dicyc10pentadienyl Magnesium in THF 293 -39 43 (dilute) a Positive values are deshielded. b A white precipitate formed in the nmr tube. The resonances observed for Grignards in tetrahydrofuran are most likely due to magnesium bromide which may be more properly thought of as solvated ions or ion pairs. This species is probably in equilibrium with other species represented in the Schlenk equilibrium. Crystallographic stuctures of magnesium bromide" and methyl magnesium bromideH (tetrahydrofuran solvated) are consistent with the theory that line widths are a function of electronic symmetry at the 72 magnesium nucleus. It is of interest to note the large chemical shift difference between magnesium bromide in diethyl ether versus tetrahydrofuran. This no doubt is due to the differing solvating ability of the two solvents and perhaps the proximity of the bromide ion. The crystal structure of magnesium bromide derived from diethyl ether has two solvent molecules per magnesium. In the case of the less sterically demanding tetrahydrofuran there are four solvent molecules per magnesium. _‘ Since the line width of a stg resonance is dependent upon the electronic symmetry around the nucleus, a symmetrical organometallic species was sought. 0f the available crystallograpic strutures reported in the literature, the most symmetrical was bis-(phenylethynyl) magnesium.u This compound has an octahedral arrangement of ligands with four tetrahydrofuran molecules in a plane and £522; phenylethynyl ligands. As mentioned, no resonance was observed between 210 and 293‘s. The case of dicyc10pentadienyl magnesium is interesting because it represents the only case in which an organomagnesium species is definitively detected. Here the line width is relatively narrow which is consistent with a symmetrical "sandwich" compound. The fact that the cyclopentadienyl rings are a bonded rather than a bonded to the magnesium may also be significant. The upfield (shielding) chemical shift is most likely due to a combination of the anistropic magnetic properties. of the aromatic rings plus the high shielding provided by the cyc10pentadienyl ligands. 73 Experimental Magnesiumrzs nmr spectra were obtained at 11.0 MHz using a Bruker WH+180 FT NMR. Twenty mm spinning sample tubes were used throughout this study. Field/Frequency lock was maintained by locking on the 2H resonance of deuterium oxide or d‘-acetone (for low temperatures) which was contained in a co-axial 5 mm tube centered within the 20 mm tube. Line widths were measured as the full width at half-height. Diethyl ether and tetrahydrofuran were purified by refluxing and distilling over sodium/benzOphenone and stored under argon and over molecular sieves. Organic halides were distilled before use. Sublimed magnesium. was obtained from the Dow Chemical Company, Midland, Michigan. The following is the general procedure for the preparation of Grignards. Tb a dry 250 mL, three-necked round‘bottomed .flask equipped with a dropping funnel, a magnetic stir bar, a condensor and an air inlet valve was added sublimed magnesium (1.1 equiv/1.0 equiv organic halide). The dropping funnel was equipped with a septum. The flask was evacuated and purged with argon. Using a syringe, 5-10 mL of solvent was added to the flask followed by 0.5 mL of the organic halide. After a short induction period or with gentle heating the reaction started. The remaining solvent and organic halide were combined in the dropping funnel using a syringe. This solution was added, with stirring, at a rate to maintain the reaction. After the addition, the reaction mixture was refluxed for 30 minutes. The mixture was allowed to cool to room temperature and settle. under an argon atmosphere, the Grignard solution was siphoned through a _... ~ aa—e- 74 U‘shaped tube equipped with a frit directly into the nmr tube. The solutions were clear and colorless except for phenyl magnesium bromide (which was only slightly discolored). Magnesium halides were prepared from the corresponding mercuric salts using an excess of magnesium as described by Ashby and Arnott.u Dimethyl and diethyl magnesium were prepared from the corresponding dialkyl mercury as described by Ashby and Arnott.u The dialkyl mercury was prepared by the method of Gilman and Brown.’0 Phenylethynyl magnesium bromide was prepared by adding a slight excess ‘ Bis-(phenylethynyl) of phenylacetylene to ethyl magnesium bromide.‘ magnesium was prepared by adding a slight excess of phenylacetylene to dimethyl magnesium.u Dicyclopentadienyl magnesium was kindly provided by Dr. Carl H. Brubaker. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. REFERENCES E. L. Eliel, Rec. Chem. Progr., 3;, 129(1961). K. Mislow and M. Raban, Top, Stereochem. ;, 1(1967). E. Fischer, Ber., 21, 3231. H. B. Burgi, J. M. Lehn and G. Wipf. IM,A!; Chem. Soc,, 26, 1956(1974). G. 8. Hammond, I; Am, Chem. Soc,, 11, 334(1955). For activation parameters see References 41,42 and 47. For heats of reaction see R. E. Davis and J. Carter, Tetrahedron, 22, 495(1966). D. I. Cram and F. A. Abd Elhafez, 1, Am, Chem, Soc., 11, 5828(1952); D. J. Cram and F. D. Greene, ibid., 1g, 6005(1953); D. J. Cram and D. R. Wilson, ibid., §§, 1254(1963). I. W. Cornforth, R. H. Cornforth and K. K. Mathews, 1, Chem, Soc,, 112(1959). G. J. Karabatsos, 14.52; Chem. Soc., §2, 1367(1967). M. Cherest, H. Felkin and N. Prudent, Tetrahedron Lett., 2201(1968). J. Richer, I; Org. Chem,, 39, 324(1965). 1. R. Boone and E. C. Ashby, Top, Stereochem. 1;, 53(1979). I. D. Morrison and H. S. Mosher, Asymmetric Organic Reactions, Prentice-Hall, Englewoods Cliffs, N. J., 1971; paperback reprint, American Chemical Society, Washington D. C. 1976. D. H. R. Barton, M, Chem, Soc,, 1027(1953). W. G. Dauben, G. I. Fonken and D. S. Noyce, I; Am. Chem. Soc., lfi. 2579(1956). H. C. Brown and H. R. Deck. J'_. _A!_,_ M m. fl. 5620(1965). E. L. Eliel and Y. Senda, Tetrahedron, 36, 2411(1970). I. Klein and E. Dunkelblum, Tetrahedron M££;,, 6127(1968). D. H. Wertz and N. L. Allinger, Tetrahedron, 39, 1597(1974). E. C. Ashby and I. T. Laemmle, Chemical Reviews, 15, 521(1975). 75 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 76 n. c. Wigfield and F. v. Gowland. L M... 21102.... 5;. 1108(1977). n. c. Wigfield. an L 91.12;... a. 646(1977). Reference 12, p. 77. Reference 13, p. 119 and Reference 12, p. 91. Reference 13, p. 120. Reference 12, p. 86. Reference 12, pp. 79-81 and Reference 13, pp. 129-130. N. T. Ahn, O. Eisenstein, I. M. Lefour and M. R. Tran Huu Dan, 1; Am. Chem, Soc., 2;, 6146(1973). I. KLein, Tetrahedron Lg£;,, 4307(1973). J. Klein and D. Lichtenberg, MM Q£Ba.§!2!;o.§2; 2654(1970). E. C. Ashby and J. R. Boone. l; Q;g,.§hgg,,lfil, 2890(1976). Reference 20, p. 544. N. T. Ahn and O. Eisenstein, Mggz,‘£&‘§Mig&, ;, 61(1977). Reference 12, p. 75. D. Varech and J. Jacques, Tetrahedron ngg,, 1;, 4443(1973). M. J. Brienne, D. Varech and J. Jacques, Tetrahedron Lett., 16, 1233(1974). M. C. Cherest, H. Felkin, P. Tacheau, I. Jacques and D. Varech, Chem. Comm., 372(1977). Y. Izumi and A. Tai, Stereo-Differentiating Reactions, Konansha Scientific Books, Tokyo, Japan, 1977, pp. 178-184. D. Varech, M. J. Brienne and I. Jacques, 1; Chem. Res. 81302., 12. 319(1979). E. C. Ashby, F. R. Dobbs and H. P. Hopkins, Jr., 1; Am, Chem. Soc,, 21, 3158(1975). A E. C. Ashby and J. R. Boone, I. Am. Chem. Soc., 2;, 5524(1976). K. E. Viegers and S. G. Smith, 1, Mg, Chem, Soc., 22, 1480(1977). H. C. Brown and R. F. McFarlin, l;.é!a Chem. Soc., £9, 5372(1958). 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 77 H. C. Brown and C. F. Shoaf, J. Am, Chem. Soc., gg, 1079(1964). H. C. Brown, E. J. Mead and B. C. Subba Rao, 1,,MM, Chem. Soc., 11. 6209(1955). H. C. Brown, 0. H. Wheeler and X. Ichikawa, Tetrahedron, ;, 214(1957). 1). c. Wigfield and F. w. Gowland. J_. ml @291. 5;. 2396(1976). n. c. Wigfield and F. w. Gowland. .J_.. m. 9.112... 1.2.. 1108(1977). H. C. Brown and K. Ichikawa, IM.A!;‘£MQQM.§22&..§§. 4372(1961). I. L. Pierre and H. Handel, Tetrahedron LQEEL. 2317(1974). H. C. Brown. E. J. Mead and C. I. Shoaf, I; Am. Chem. Soc., 18. 3616(1956). H. Haubenstock and E. L. Eliel, 11.52; Chem, Soc,, £1, 2368(1962). B. Rickborn and M. T. Wuesthoff,‘1, Am. Chem. Soc,, 22, 6894(1970). G. J. Karabatsos and D. I. Fenoglio, Topics Mg Stereochemistgy, é. 167(1970). For Equations 9, 10, and 11 see Reference 13, p. 123. For Equation 12 see Reference 12, p. 89. For Equations 13 and 14 see Reference 20, p. 532. G. G. Smith and R. P. Bayer, Tetrahedron, Mg, 323(1962). J. A. Perry and K. D. Warren, ;, Chem. Soc,, 4049(1965). K. Bowden and M. Hardy, Tetrahedron, 22, 1169(1966). P . Geneste and G. Lamaty, Bull. Soc. Chim. 25;, 669(1968). J. E. Leffler, 1‘ Org, Chem., 29, 1202(1955). R. C. Peterson, 1, Org. Chem,, 22, 3133(1964). n. c. Wigfield, n. J. Phelps, 3. F. Pottie and a. Sander, q, Ln, Chem. Soc., 21, 898(1975). D. C. Wigfield and D. J. Phelps, Chem. Commun,, 1152(1970). D. C. Wigfield and D. I. Phelps, Can. M, Chem., gg, 388(1972). H. C. Brown and J. Muzzio, erAm. Chem. Soc., MM, 2811(1966). 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 78 W. E. Noland, Org. Syp., Coll. Vol. 5, 833(1973). C. D. Hurd and M. E. Nilson, l; Org. Chem.. 29, 927(1955). G. D. Buckley and C. W. Scaife, M, Chem, Soc.. 1471(1947). Y. V. Baskov, T. Urbanski, M. Witanowski and L. Stefaniak, Tetrahedron, 29, 1519(1964). The methyl doublet for the cis isomer is at 8 2.02. D. W. Worrall, Org. 823,, Coll. vol. 1, 413(1941). W. E. Noland, H. I. Freeman and M. S. Baker, 11.52; Chem. Soc.. 1M, 188(1956). E. Keinan and Y. Mazur, 1;,AQ& Chem. Soc,, 22, 3861(1977). T. H. Chan, I. Paterson and J. Pinsonnault, Tetrahedron Lett., fifi. 4183(1977). E. L. Eliel and S. H. Schroeter, l; Am. Chem. Soc.. 81. 5031(1965). K. B. Wiberg, Physical Organic Chemistry, John Wiley and Sons, Inc., New York, 1964. pp. 376-379. R. C. Peterson, J. Markgraf and S. D. Ross, 1; Am. Chem. Soc.. £1. 3819(1961). E. c. Ashby, Quart, Rev,, 2;, 259(1967). E. c. Ashby, Bull. Soc. Chim, Fr,. 2133(1972). R. M. Salinger and H. S. Mosher, leAm. Chem. Soc., 85, 118(1963). G. E. Parris and E. C. Ashby, J; Am. Chem. Soc,, 23, 1206(1971) and references therein. D. Leibfritz, B. O. Wagner and J. D. Roberts, Leibigs Ann, Chem. 763. 173(1972). L. Simeral and c. a. Maciel. L. 2222.. Elan“ £9. 552(1976). P. H. Heubel and A. I. Popov, 1; Solution 9222,, g, 283(1979). 6. n. Stucky and a. E. Bundle. .1. _Am. £11.22... in... 132. 1002(1963). H. Schibilla and M. T. LeBihan. Acta Egyg£&,{;§, 232(1967). M. C. Perucaud and M. T. LeBihan, Acta Cryst,, Sect. B, 31, 1502(1968). 87. 89. 90. 79 M. Vallino, L, Organometal. Chem., LO, 1(1969). M. C. Perucaud, M. J. Ducom and M. Vallino. _C_. M, Acad. Sci.. Paris, ser. C, 264, 571(1967). E. C. Ashby and R. C. Arnott, l; Organometal. Chem,, 141, 1(1968) . H. Gilman and R. E. Brown, L a, Chem. Soc,, 52, 3514(1930). IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII , llHlHfllllUlMflm[[tfllLUlllflllffllejfltW”