<<

Asymmetric catalysis in complex target synthesis

Mark S. Taylor and Eric N. Jacobsen* Department of Chemistry and Chemical Biology, Harvard University, Cambridge, MA 02138

Edited by Jack Halpern, University of Chicago, Chicago, IL, and approved January 30, 2004 (received for review November 26, 2003)

This article describes three distinct strategies by which stereochemically complex molecules are synthesized and the ways asymmetric ca- talysis can impact on all three. The development of general methods to prepare synthetically useful building blocks leads to an expanded ‘‘’’ of potential starting materials for asymmetric synthesis. The possibility of discovering new reactions to access new types of building blocks is particularly attractive and serves to help define the frontiers of the field. Asymmetric catalysis can also be applied to di- astereoselective synthesis such that the of the catalyst, and not that of the substrate, determines the relative configura- tion of the product. Finally, in reactions where multiple are generated simultaneously or in tandem, catalyst and substrate control can operate in a complementary manner to achieve one of many possible stereochemical outcomes selectively.

he synthesis of both natural and unnatural organic compounds in optically active form is a cen- T tral challenge in chemistry, es- pecially in relation to the study of bio- logically active compounds. Natural products and synthetic pharmaceutical agents display a rich diversity of molecu- lar structures that range from very sim- ple to astonishingly intricate. A feature common to many such compounds is the

presence of multiple stereocenters, giv- Scheme 1. Masamune–Sharpless synthesis of the L-hexoses. Asterisks denote stereocenters under ing rise to the possibility of large num- absolute control. bers of stereoisomers (2n, where n is the number of stereocenters). The control of the relationship between stereo- substrate control. Other methods for rap- With the discovery of the first general centers is therefore an essential element idly achieving stereochemical complexity asymmetric synthetic reactions in the to the synthesis of many complex targets under catalyst control are possible, includ- 1980s, the way that stereochemically com- of interest. ing the desymmetrization of meso com- plex targets could be accessed was ex- There are several ways to generate ste- pounds and stereospecific reactions of panded in two crucial ways. Asymmetric reochemical complexity in a selective geometrically defined precursors. synthesis began to provide access to a manner (Fig. 1). Chiral building block Before the advent of efficient asymmet- greatly expanded pool of readily available assembly, for which peptide synthesis rep- ric synthetic methods, application of the chiral starting materials (2), and synthetic resents the prototype, relies on the avail- chiral building block approach was con- chemists could now devise strategies to ability of the appropriate optically pure strained by the very limited pool of opti- prepare target structures based on chiral starting materials along with the reactions cally active compounds provided by na- starting materials unavailable from nature. to elaborate and couple them. An alterna- ture in abundance (the ‘‘chiral pool’’) (1). In addition, ‘‘powerful’’ chiral reagents tive is diastereoselective synthesis, wherein Because of this constraint, strategies in and catalysts were identified that can in- new stereocenters are introduced into target-oriented synthesis were largely pre- duce highly diastereoselective reactions compounds bearing preexisting ones. The through external stereocontrol, overcom- selectivity may arise either from the prop- mised on the fact that optically active ing or complementing the inherent sub- erties inherent to the substrate (substrate starting materials were particularly pre- strate bias (3). control) or from an external agent (re- cious, and the application of substrate- Masamune and Sharpless provided a agent or catalyst control). A third strategy controlled diastereoselective processes was involves the use of transformations that emphasized. Although this now-classic seminal and brilliant illustration of the simultaneously introduce multiple stereo- approach has proven enormously power- potential of catalyst-induced double ste- centers in a single operation. For example, ful and will surely always play a critical reodifferentiation in natural product syn- the coupling of two prochiral substrates role in complex molecule construction, it thesis in their stereocontrolled synthesis of may result in four (or more) stereochemi- is also inherently limited because it re- the L-hexoses. The then-recently devel- cal outcomes; enantio- and diastereoselec- quires a specific design concept for every oped titanium tartrate (Sharpless) epoxi- tive variants of such processes usually in- new target (including of the dation catalyst was used to generate the corporate elements of catalyst and same target). key chiral starting material and was subse- quently applied in the diastereoselective elaboration of a late-stage intermediate

This paper was submitted directly (Track II) to the PNAS office. Abbreviation: HDA, hetero-Diels–Alder. *To whom correspondence should be addressed. E-mail: Fig. 1. Strategies for the construction of stereochemically complex targets. Asterisks denote the [email protected]. presence of defined stereocenters. © 2004 by The National Academy of Sciences of the USA

5368–5373 ͉ PNAS ͉ April 13, 2004 ͉ vol. 101 ͉ no. 15 www.pnas.org͞cgi͞doi͞10.1073͞pnas.0307893101 Downloaded by guest on September 29, 2021 PCA ETR PERSPECTIVE FEATURE SPECIAL

Fig. 2. Selected applications of the HKR reaction in natural product synthesis.

(Scheme 1) (4). Ultimately, all four con- metric catalytic reactions. The fruit of this many cases, this has forced synthetic tiguous stereocenters were controlled in- effort has been the recent discovery of a chemists to develop lengthy, inefficient dependently, providing access to each of variety of remarkably general catalytic syntheses to make use of available the eight hexose diastereomers by essen- oxidation, reduction, and C–C bond-form- compounds. tially the same route. ing reactions (5). The impact of these Asymmetric catalysis can provide a The strategy underlying the synthesis of methods on the manner in which complex powerful solution to this problem by ren- the hexoses is so compelling that one targets can be synthesized is just begin- dering accessible chiral materials not might ask why asymmetric catalysis has ning to be felt (6) and is discussed here in readily provided by nature. In principle, not become the approach by which all several representative examples. small molecule synthetic catalysts can be complex target synthesis is undertaken. applied to the synthesis of building blocks One answer is that certain targets are best An Expanded Chiral Pool as either and with nearly com- accessed from nature’s chiral pool or are The chiral pool approach to synthesis of plete generality. For example, whereas well suited to substrate-controlled diaste- optically active targets can be extremely nature’s chiral pool is not a useful source reoselective synthesis strategies, and asym- attractive if nature happens to provide of optically active terminal epoxides, these metric catalytic methods do not offer any an abundant supply of a starting mate- very useful compounds are now readily advantage. However, another reason is rial appropriate for the synthetic target. accessible in one step from the corre- that the hexose synthesis relied on a reac- Accordingly, considerable effort and sponding racemates by the (salen)Co- tion that was far ahead of its time, and creativity has been directed toward us- catalyzed hydrolytic (9). that there are still too few reactions as ing inexpensive chiral pool elements in As a result, it is now possible to evaluate effective and as general as the Sharpless target-oriented synthesis. In the best complex targets strategically with the epoxidation. As a result, it has not been cases, some of the most practical and knowledge that enantiopure terminal ep- possible to devise attractive synthetic brilliant syntheses ever devised have oxides represent viable starting points. routes to target structures based on avail- used this approach. The Stork syntheses Various groups have incorporated this ␣ able methods. of prostaglandin A2 and F2 are classic recently developed transformation into The latter consideration has motivated examples (7, 8). However, nature pro- synthetic routes to natural products of several research groups throughout the vides only a limited number of chiral varying complexity (Fig. 2) (10–13). world to overcome this limitation, by ex- starting materials, and usually only one In many cases, retrosynthetic analysis of panding the availability of powerful asym- enantiomer is produced naturally. In a complex target leads to chiral starting

Scheme 2. Synthesis of the central fragment of FR901464.

Taylor and Jacobsen PNAS ͉ April 13, 2004 ͉ vol. 101 ͉ no. 15 ͉ 5369 Downloaded by guest on September 29, 2021 Scheme 3. Catalyst-controlled synthesis of iridoid natural products.

Scheme 4. Catalyst-controlled syntheses of manzacidin A and C.

materials that are not readily accessible by thetic chemists no longer must work only FR901464 synthesis suggest the feasibility available methods. In these situations, nat- with what nature provides directly. In sev- of attaining it. ural products can provide the inspiration eral cases, there is now general access to for the discovery of new asymmetric cata- entire classes of synthetically useful chiral Catalyst-Controlled Stereochemical lytic methods. For example, consideration building blocks. Ultimately, it should be Elaboration of Chiral Intermediates of synthetic approaches to the structurally possible to apply existing asymmetric cata- The classic approach to generating interesting antitumor agent FR901464 re- lytic reactions to afford a practical route stereochemical complexity in target- vealed an intriguing asymmetric catalytic to any building block of interest or to de- oriented synthesis has involved reliance route to the central fragment (Scheme 2). vise new processes when the need arises. on the conformational properties of the However, asymmetric cycloadditions in- Although the latter, rather ambitious goal substrate to direct the diastereoselective volving weakly nucleophilic diene partners remains far from met, examples such as generation of new stereocenters (sub- such as 2 and simple aldehydes were un- the HDA reaction developed for the strate control). To access the specific known, and as such this target provided the impetus to develop new methodology. In this context, it was discovered that novel (Schiff base)Cr(III) complexes such as 1 are remarkably reactive chiral Lewis acid catalysts, promoting cycloaddition between diene 2 and ynal 3 to establish the complete carbon framework of the central fragment while setting three of the four stereocenters simultaneously (Scheme 3) (14). Catalyst 1 was also applied successfully to the enantioselective synthesis of the right-hand fragment of FR901464. The total synthesis was completed in a conver- gent manner by using high yielding Pd- cross coupling and amide bond-forming reactions (15, 16). Subsequently, the het- ero-Diels–Alder (HDA) chemistry devel- oped specifically for this synthesis proved applicable to the synthesis of a number of other, structurally distinct natural products (17–21). Thus, advances in asymmetric catalysis have lifted many of the restrictions on the chiral pool approach to the synthesis of stereochemically complex targets, as syn- Scheme 5. Asymmetric catalytic syntheses of quinine and quinidine.

5370 ͉ www.pnas.org͞cgi͞doi͞10.1073͞pnas.0307893101 Taylor and Jacobsen Downloaded by guest on September 29, 2021 and elaboration in the synthesis of diaste- reomeric natural products is also illus- trated in the synthesis of the cinchona al- kaloids quinine and quinidine (Scheme 5) (33). Aluminum-catalyzed asymmetric Michael addition methodology provided ready access to enantio-enriched adduct 8, which was converted in several steps to the highly functionalized alkene 9. Asym- metric dihydroxylation using the pseudo- enantiomeric catalyst systems ADmix-␣ and ADmix-␤ afforded diastereomeric diols 10 and 11 with high selectivity, and these were converted stereospecifically to quinidine and quinine. The formal synthesis of fostriecin re- ported by Shibasaki and coworkers fea- tures the use of four catalytic asymmetric processes, of which three are C–C bond- forming reactions, to generate each of the Scheme 6. Asymmetric catalytic synthesis of fostriecin. four stereocenters present in this promis- ing anticancer agent (Scheme 6) (34). The configuration of the C8 tertiary alcohol, stereochemical relationship in complex the iridoid family of natural products (30). one of the challenging stereochemical molecules, this has often required ex- The requisite aldehyde 5 was prepared in features of the natural product, was traordinary levels of creativity in syn- two steps from citronellal (4), a chiral established at an early stage by titanium- thetic design, and this is the basis for building block that is produced commer- catalyzed enantioselective ketone cyana- much of the ‘‘art’’ of organic synthesis. cially in either enantiomeric form by a tion. The resulting chiral building block Indeed, the history of modern organic highly efficient catalytic asymmetric was elaborated by catalyst-controlled di- synthesis can be tied closely to the ques- isomerization reaction (31). Selective astereoselective transformations, including tion of how stereocontrol has been cycloaddition with either enantiomer of 5 a direct aldol fragment coupling catalyzed achieved in complex target synthesis was achieved by using chiral chromium by a chiral lanthanum complex and a late- (22, 23). Early successes relied princi- catalyst 6. The diastereomeric products stage enantioselective ketone reduction pally on cyclic stereocontrol on confor- were thus generated selectively, and under transfer hydrogenation conditions mationally restricted substrates. Wood- readily elaborated to the natural products using Noyori’s ruthenium catalyst. ward’s (24–26) work in total synthesis is boschnialactone and iridomyrmecin. The synthesis of the antifungal agent appreciated widely in this regard, with The catalyst-controlled hydrogenation ambruticin further demonstrates how his synthesis of reserpine being a dra- of homoallylic alcohol 7 was a key trans- modern catalyst- and reagent-controlled matic, albeit representative example. formation in the synthesis of the bro- asymmetric reactions can serve as a guid- The development of more sophisticated mopyrrole alkaloids manzacidin A and C ing strategy in complex target synthesis principles in conformational analysis (Scheme 4) (32). The requisite building (Scheme 7) (35). Two new applications of opened the door to successes in acylic block was prepared by a highly enantiose- HDA methodology served to establish the stereocontrol, as demonstrated in Kishi’s lective catalytic ene reaction. Stereoselec- left- and right-hand pyran units, and the (27–29) pioneering work in the synthesis tive hydrogenation of 7 could not be first application of asymmetric hydro- of monensin. achieved by using substrate control, but formylation to target-oriented synthesis The alternative possibility of controlling either diastereoisomer could be accessed allowed effective construction of the C15 relative stereochemistry predictably by efficiently by choice of the appropriate under complete catalyst using an external controlling element (cat- chiral rhodium catalyst. control. Asymmetric cyclopropanation was alyst control) holds significant appeal. The The combined application of asymmet- applied in a doubly diastereoselective con- synthetic plan is no longer contingent on ric catalysis to building block preparation text to introduce the central, trisubstituted the properties of the substrate, but rather on the ability of the catalyst to exert ste- reoselectivity on a complex substrate. The value of catalyst-controlled transforma- tions is particularly evident when selective access to multiple diastereomers is re- quired in the course of a synthetic effort. Catalyst control introduces the possibility of generating stereoisomeric products by the same route, changing only the enan- tiomer of catalyst or of substrate to access diastereomeric products selectively. For example, the inverse electron-demand HDA reaction of chiral aldehyde 5 with ethyl vinyl ether was applied toward the preparation of the fused cyclopenta[c]- pyran bicyclic ring system characteristic of Scheme 7. Retrosynthetic analysis applied to the asymmetric catalytic synthesis of ambruticin.

Taylor and Jacobsen PNAS ͉ April 13, 2004 ͉ vol. 101 ͉ no. 15 ͉ 5371 Downloaded by guest on September 29, 2021 The synthesis of the pyrrolidinoindoline alkaloid quadrigemine C illustrates cata- lyst-controlled desymmetrization of meso compounds in total synthesis in a spectac- ular way (37). Symmetrical dimeric sub- strate 11 underwent palladium-catalyzed enantioselective double Heck cyclization, generating a product bearing four carbon- substituted quaternary stereocenters and six of the eight stereocenters present in the natural product (Scheme 9). Com- pound 11 is one the most complex sub- strates ever subjected successfully to an enantioselective catalytic reaction, and the tolerance of the palladium catalyst to the numerous functional groups present in Scheme 8. Asymmetric catalytic synthesis of myo-inositol-1-phosphate. this advanced intermediate underlies this success. Multiple stereocenters may also be gen- erated by tandem or serial transforma- tions in which a number of bonds are cre- ated sequentially in a single operation. The synthesis of the oxasqualenoid gla- brescol highlights an innovative implemen- tation of an asymmetric catalytic reaction to generate a highly stereochemically com- plex structure rapidly from a relatively simple starting material (Scheme 10) (38). Eight of the 10 stereocenters present in the natural product were introduced in a single transformation. Reaction of tet- raene 12 with Shi’s catalyst led to oxida- Scheme 9. Asymmetric catalytic synthesis of quadrigemine. tion of all four double bonds and genera- tion of the indicated tetraepoxide in Ϸ80% diastereomeric purity. Taking into cyclopropane ring. The overall sequence products lie at the heart of many of the account that the starting tetraene was was exceptionally efficient (16 steps and most elegant total syntheses accom- enantiomerically pure, each of the four 12% yield in the longest linear sequence), plished to date. Asymmetric catalysis epoxidation events involved a catalyst- and each of the stereochemical challenges may be used to dramatic effect in this controlled diastereoselective process. was met in a highly selective manner, with context, with both the relative and abso- Seven chemically distinct orderings of the recently developed enantioselective C–C lute configuration of multiple stereo- bond-forming reactions applied to the four epoxidations are possible, and it is centers controlled simultaneously. likely that all sequences occur to some direct introduction of 8 of the 10 stereo- The desymmetrization of meso sub- centers. extent given the similar steric and elec- strates is a powerful approach to the In general, the implementation of a tronic properties of each double bond. catalyst-controlled diastereoselective syn- enantioselective synthesis of compounds Per-epoxidation of tetraene 12 was thus thesis strategy places the most stringent of bearing multiple stereocenters. This achieved with an estimated selectivity of demands on the catalytic reaction in ques- strategy forms the basis for a highly effi- Ͼ20:1 for each reaction. This is a dra- tion. When elaborate synthetic intermedi- cient approach to the cellular signaling matic manifestation of catalyst generality ates are used as substrates, the catalyst agent myo-insotol-1-phosphate, wherein given that the process most likely involved must be tolerant of potentially reactive a simple synthetic pentapeptide cata- diastereoselective reactions on eight dif- functional groups, and may need to over- lyzed the phosphorylation of the pro- ferent substrates. The resulting tetraep- come inherent substrate bias to set the tected inositol derivative 10, generating oxide was elaborated to glabrescol in a required stereochemical relationships. a product bearing six stereocenters in short sequence featuring acid-catalyzed Furthermore, catalyst performance must one step from an achiral starting mate- cascade cyclization to form four tetrahy- be highly predictable, because it may be rial (Scheme 8) (36). drofuran rings in a single step. difficult to devise meaningful model sys- tems, and the cost of failure at the late stages of a synthetic effort is high. None- theless, the range of reactions that has been applied successfully in catalyst- controlled diastereoselective synthesis is expanding rapidly, as illustrated in the preceding examples. Setting Multiple Stereocenters Simultaneously Transformations of simple starting ma- terials to stereochemically complex Scheme 10. Asymmetric catalytic synthesis of glabrescol.

5372 ͉ www.pnas.org͞cgi͞doi͞10.1073͞pnas.0307893101 Taylor and Jacobsen Downloaded by guest on September 29, 2021 nism. The successful application of these reactions to natural product targets re- quired that those stereochemical ele- ments controlled by intrinsic reaction bias could be successfully mapped onto the relative stereochemistry present in the natural products. Much greater free- dom in choice of synthetic targets would be possible if introduction of all stereo- chemical elements were under catalyst control, such that any could be accessed by choice of cata- lyst(s). This presents a daunting chal- lenge for reaction development, particu- Scheme 11. Asymmetric catalytic synthesis of prostaglandin E methyl ester. 1 larly in cases wherein the intrinsic substrate or reaction bias is high. Devis- Coupling between prochiral partners selective and simultaneous generation of ing solutions to this problem will likely can lead to stereochemically complex multiple stereocenters in enantioselective require new concepts in catalyst design; products in a single reaction, and chiral catalytic Diels–Alder reactions. In the syn- independent activation of each reaction catalysts introduce the possibility of effect- thesis of FR901464 discussed above partner in cooperative multicatalyst sys- ing such transformations enantioselec- (Scheme 2), the asymmetric catalytic tems represents a particularly intriguing approach (40, 41). tively. This strategy was used to establish HDA reaction was used to set three of The total syntheses described here the complete carbon skeleton and three of the four stereocenters of the central ring. provide an indication of how stereo- the four stereocenters of prostaglandin E1, This result reflects three different stereo- in a copper-catalyzed three-component chemically complex targets have become chemical control elements: the chiral cata- more readily accessible with the discov- coupling of a diorganozinc reagent, cyclo- lyst (dictating which enantioface of the penten-3,5-dione monoacetal, and an al- ery and reduction to practice of truly aldehyde reacts selectively), the stereospe- dehyde (Scheme 11) (39). The stereo- general and powerful asymmetric cata- cific nature of the cycloaddition reaction chemical outcome of this reaction lytic methods. With continuing progress reflected an interplay of catalyst and sub- (translating the diene configuration to the in catalyst design and development, ulti- strate control: the chiral copper catalyst relative stereochemistry of the two meth- mately one might imagine that practical dictated the facial selectivity of organozinc yl-substituted carbons), and the endo di- routes to any complex target of interest conjugate addition to the unsaturated ke- astereoselectivity (dictating the relative may be realized by using selective cata- tone, and the configurations of the two approach of aldehyde to diene). lytic reactions to introduce all stereo- other stereocenters resulted from an in- Whereas the absolute stereochemistry chemical elements independently and trinsic substrate bias in the subsequent of each of these complexity-generating efficiently. Although this goal remains aldol addition. reactions is under catalyst control, the quite distant, it is increasingly clear what The combination of substrate and cata- relative stereochemistry is dictated by sorts of advances in asymmetric catalysis lyst control also lies at the heart of the the substrate and the reaction mecha- will be required to attain it.

1. Hanessian, S. (1983) in Total Synthesis of Natural 14. Dossetter, A. G., Jamison, T. F. & Jacobsen, E. N. 28. Fukuyama, T., Wang, C.-L. J. & Kishi, Y. (1979) Products: The ‘‘Chiron’’ Approach, ed. Baldwin, J. E. (1999) Angew. Chem. Int. Ed. Engl. 38, 2398–2400. J. Am. Chem. Soc. 101, 260–262. (Pergamon, New York). 15. Thompson, C. F., Jamison, T. F. & Jacobsen, E. N. 29. Fukuyama, T., Akasaka, K., Karanewsky, D. S., 2. Nugent, W. A., RajanBabu, T. V. & Burk, M. J. (2000) J. Am. Chem. Soc. 122, 10482–10483. Wang, C.-L. J., Schmid, G. & Kishi, Y. (1979) J. Am. (1993) Science 259, 479–483. 16. Thompson, C. F., Jamison, T. F. & Jacobsen, E. N. Chem. Soc. 101, 262–264. 3. Masamune, S., Choy, W., Petersen, J. S. & Sita, (2001) J. Am. Chem. Soc. 123, 9974–9983. 30. Chavez, D. E. & Jacobsen, E. N. (2003) Org. Lett. 5, L. R. (1985) Angew. Chem. Int. Ed. Engl. 24,1–30. 17. Chavez, D. E. & Jacobsen, E. N. (2001) Angew. 2563–2565. 4. Ko, S. Y., Lee, A. W. M., Masamune, S., Reed, Chem. Int. Ed. Engl. 40, 3667–3670. 31. Akutagawa, S. (1999) in Comprehensive Asymmetric L. A., III, Sharpless, K. B. & Walker, F. J. (1983) 18. Paterson, I., De Savi, C. & Tudge, M. (2001) Org. Catalysis, eds. Jacobsen, E. N., Pfaltz, A. & Science 220, 949–951. Lett. 3, 3149–3152. Yamamoto, H. (Springer, New York), Vol. 3, Chap. 5. Jacobsen, E. N., Pfaltz, A. & Yamamoto, H., eds. 19. Wender, P. A., Baryza, J. L., Bennett, C. E., Bi, C., 41.4. (1999) Comprehensive Asymmetric Catalysis Brenner, S. E., Clarke, M. O., Horan, J. C., Kan, C., 32. Wehn, P. M. & Du Bois, J. (2002) J. Am. Chem. Soc. (Springer, New York), Vols. 1–3. Lacoˆte, E., Lippa, B., et al. (2002) J. Am. Chem. Soc. 124, 12950–12951. 6. Hoveyda, A. H. (2000) in Stimulating Concepts in 124, 13648–13649. 33. Raheem, I. T., Goodman, S. N. & Jacobsen, E. N. Chemistry, eds. Stoddart, F. J., Shibasaki, M. & 20. Paterson, I. & Tudge, M. (2003) Angew. Chem. Int. (2004) J. Am. Chem. Soc. 126, 706–707. Vogtle, F. (VCH, Weinheim, Germany), pp. 145– Ed. Engl. 42, 343–347. 34. Fujii, K., Maki, K., Kanai, M. & Shibasaki, M. 160. 21. Paterson, I. & Luckhurst, C. A. (2003) Tetrahedron (2003) Org. Lett. 5, 733–736. 7. Stork, G. & Raucher, S. (1976) J. Am. Chem. Soc. Lett. 44, 3749–3754. 35. Liu, P. & Jacobsen, E. N. (2001) J. Am. Chem. Soc. 98, 1583–1584. The Logic of 123, 8. Stork, G., Takahashi, T., Kawamoto, I. & Suzuki, T. 22. Corey, E. J. & Cheng, X.-M. (1995) 10772–10773. (1978) J. Am. Chem. Soc. 100, 8272–8273. Chemical Synthesis (Wiley, New York). 36. Sculimbrene, B. R. & Miller, S. J. (2001) J. Am. 9. Schaus, S. E., Brandes, B. D., Larrow, J. F., Toku- 23. Nicolaou, K. C. & Sorensen, E. J. (1996) Classics in Chem. Soc. 123, 10125–10126. naga, M., Hansen, K. B., Gould, A. E., Furrow, Total Synthesis (VCH, New York). 37. Lebsack, A. D., Link, J. T., Overman, L. E. & M. E. & Jacobsen, E. N. (2002) J. Am. Chem. Soc. 24. Woodward, R. B., Bader, F. E., Bickel, H., Frey, Stearns, B. A. (2002) J. Am. Chem. Soc. 124, 9008– 124, 1307–1315. A. J. & Kierstead, R. W. (1956) J. Am. Chem. Soc. 9009. 10. Schaus, S. E., Brånalt, J. E. & Jacobsen, E. N. (1998) 78, 2023–2025. 38. Xioing, Z. & Corey, E. J. (2000) J. Am. Chem. Soc. J. Org. Chem. 63, 4876–4877. 25. Woodward, R. B., Bader, F. E., Bickel, H., Frey, 122, 9328–9329. 11. Colby, E. A., O’Brien, K. C. & Jamison, T. F. (2004) A. J. & Kierstead, R. W. (1956) J. Am. Chem. Soc. 39. Arnold, L. A., Naasz, R., Minnaard, A. J. & Feringa, J. Am. Chem. Soc. 126, 998–999. 78, 2657–2857. B. L. (2001) J. Am. Chem. Soc. 123, 5841–5842. 12. Nagasawa, K. (2000) J. Synth. Org. Chem. Jpn. 58, 26. Woodward, R. B., Bader, F. E., Bickel, H., Frey, A. J. 40. Shibasaki, M., Sasai, H. & Arai, T. (1997) Angew. 877–886. & Kierstead, R. W. (1958) Tetrahedron 2, 1–57. Chem. Int. Ed. Engl. 36, 1236–1256. 13. Panek, J. S. & Liu, P. (2000) J. Am. Chem. Soc. 122, 27. Schmid, G., Fukuyama, T., Akasaka, K. & Kishi, Y. 41. Jacobsen, E. N. (2000) Acc. Chem. Res. 33, 421– 11090–11097. (1979) J. Am. Chem. Soc. 101, 259–260. 431.

Taylor and Jacobsen PNAS ͉ April 13, 2004 ͉ vol. 101 ͉ no. 15 ͉ 5373 Downloaded by guest on September 29, 2021