Organic Synthesis
Latest Publications


TOTAL DOCUMENTS

104
(FIVE YEARS 0)

H-INDEX

0
(FIVE YEARS 0)

Published By Oxford University Press

9780199965724, 9780197563328

Author(s):  
Douglass F. Taber

The pentacyclic Apocynaceae alkaloid aspidophylline A 3 reverses drug resistance in resistant KB cells. In developing a strategy for the assembly of 3, Neil K. Garg of UCLA envisioned (J. Am. Chem. Soc. 2011, 133, 8877) the intramolecular Pd-catalyzed cyclization of 1 to 2. The starting material for the cyclohexenone derivative 1 was the known tricyclic anhydride 7. This was readily available in gram quantities by oxidation of the commercial pyridone 4. The double decarboxylation to 8 was delicate but could be effected by iterative small-batch microwave heating. Protection of 8 followed by fragmentation and alkylation than delivered 1. The intramolecular Heck cyclization of 1 indeed proceeded smoothly, giving the bicyclic diene 2. Deprotection of the ketone revealed a doubly activated enone, which could be selectively reduced under modifi ed dissolving metal conditions to give the keto ester 12. Alkylation of the lithium enolate with allyl iodide then gave 13, predominantly as the diastereomer illustrated. Reduction followed by selective Johnson-Lemieux oxidative cleavage of the terminal alkene then completed the construction of the diol 14. The vision for the final assembly of the alkaloid was to effect interrupted Fischer indolization of an alkylated cyclohexanone such as 15. To this end, several bicyclic ketones were explored, but none was successful. Finally, attention was turned to the more rigid tricyclic lactone 15. Happily, exposure of 15 to phenylhydrazine in the presence of trifluoroacetic acid led to an intermediate that was not isolated, but directly combined with methanolic K2CO3 to open the lactone, allowing closure of the tetrahydrofuran ring, to give 16. Simple arene sulfonamides can be advantageous in synthesis, as they do not appear as rotameric mixtures in NMR, and are often crystalline. Nevertheless, they have not commonly been used because of the perceived difficulty of deprotection. Sonication of 16 with Mg powder in methanol containing solid NH4Cl led to smooth desulfonylation. Formylation then completed the synthesis of aspidophylline A 3.


Author(s):  
Douglass F. Taber

(+)-Complestatin 3 shows promising activity against HIV infectivity. Dale L. Boger of Scripps/La Jolla described (J. Am. Chem. Soc. 2010, 132, 7776) an elegant multicomponent assembly of 3, the key step of which was the atropisomer-selective intramolecular Larock cyclization of 1 to 2. The preparation of 1 began with the protected phenethylamine 5, prepared by Sharpless asymmetric aminohydroxylation of the styrene 4. Conversion of 5 to the areneboronic acid followed by coupling with 6 delivered 7. Acylation led to 8, with the stage set for nitro-assisted addition-elimination, to form the first bis-aryl ether of 3. The product was a mixture of atropisomers, subsequently symmetrized to 9 by removal of the nitro group. Acylation of 9 led to 1. The role of the silyl group on the alkyne of 1 was to direct the regioselectivity of the intramolecular Larock indole synthesis. Again, two atropisomers were possible from the cyclization. Earlier model studies had suggested some preference for one over the other. As it turned out, in this case the desired atropisomer was the only one observed. It is particularly striking that the coupling was efficient even in the presence of the readily reduced and unprotected chlorophenols. The modular nature of this route to (+)-complestatin 3 will make it possible to prepare a variety of analogues. As long as only the substituents on the periphery are changed, the atropisomer selectivity in the Larock cyclization should be maintained.


Author(s):  
Douglass F. Taber

The periwinkle-derived alkaloids vinblastine 2a and vincristine 2b are still mainstays of cancer chemotherapy. The more complex half of these dimeric alkaloids, vindoline 1, presents a formidable challenge for total synthesis. Building on his previous work (Organic Lett. 2005, 7, 4539), Dale L. Boger of Scripps/La Jolla devised (J. Am. Chem. Soc. 2010, 132, 3685) a strikingly simple solution to this problem based on sequential cycloaddition. The starting point for the synthesis was the ester 3, derived from D-asparagine. This was extended to 4, condensation of which with 5 gave the enol ether 6. On heating, 7 cyclized to 8, which lost N2 to give the zwitterion 9. Addition of the intermediate 9 to the indole then gave 10. In one reaction, the entire ring system of vindoline, appropriately oxygenated, was assembled, with the original stereogenic center from D-asparagine directing the relative and absolute configuration of the final product. To complete the synthesis, the pendant carbon on 11 had to be incorporated into the pentacyclic skeleton. After adjusting the relative configuration of the secondary alcohol, the N was rendered nucleophilic by reduction of the amide to the amine. Oxidation delivered 14, which on activation as the tosylate smoothly rearranged to the ketone 15. Reduction and regioselective dehydration then completed the synthesis of vindoline 1.


Author(s):  
Douglass F. Taber

Varinder K. Aggarwal of the University of Bristol described (Angew. Chem. Int. Ed. 2010, 49, 6673) the conversion of the Sharpless-derived epoxide 1 into the cyclopropane 2. Christopher D. Bray of Queen Mary University of London established (Chem. Commun. 2010, 46, 5867) that the related conversion of 3 to 5 proceeded with high diastereocontrol. Javier Read de Alaniz of the University of California, Santa Barbara, extended (Angew. Chem. Int. Ed. 2010, 49, 9484) the Piancatelli rearrangement of a furyl carbinol 6 to allow inclusion of an amine 7, to give 8. Issa Yavari of Tarbiat Modares University described (Synlett 2010, 2293) the dimerization of 9 with an amine to give 10. Jeremy E. Wulff of the University of Victoria condensed (J. Org. Chem. 2010, 75, 6312) the dienone 11 with the commercial butadiene sulfone 12 to give the highly substituted cyclopentane 13. Robert M. Williams of Colorado State University showed (Tetrahedron Lett. 2010, 51, 6557) that the condensation of 14 with formaldehyde delivered the cyclopentanone 15 with high diastereocontrol. D. Srinivasa Reddy of Advinus Therapeutics devised (Tetrahedron Lett. 2010, 51, 5291) conditions for the tandem conjugate addition/intramolecular alkylation conversion of 16 to 17. Marie E. Krafft of Florida State University reported (Synlett 2010, 2583) a related intramolecular alkylation protocol. Takao Ikariya of the Tokyo Institute of Technology effected (J. Am. Chem. Soc. 2010, 132, 11414) the enantioselective Ru-mediated hydrogenation of bicyclic imides such as 18. This transformation worked equally well for three-, four-, five-, six-, and seven-membered rings. Stefan France of the Georgia Institute of Technology developed (Org. Lett. 2010, 12, 5684) a catalytic protocol for the homo-Nazarov rearrangement of the doubly activated cyclopropane 20 to the cyclohexanone 21. Richard P. Hsung of the University of Wisconsin effected (Org. Lett. 2010, 12, 5768) the highly diastereoselective rearrangement of the triene 22 to the cyclohexadiene 23. Strategies for polycyclic construction are also important. Sylvain Canesi of the Université de Québec devised (Org. Lett. 2010, 12, 4368) the oxidative cyclization of 24 to 25.


Author(s):  
Douglass F. Taber

Adriaan J. Minnaard and Ben L. Feringa of the University of Groningen devised (J. Am. Chem. Soc. 2010, 132, 14349) what promises to be a general strategy for the construction of enantiomerically pure cyclopropanes, based on conjugate addition to acceptors such as 1 . X. Peter Zhang of the University of South Florida developed (J. Am. Chem. Soc. 2010, 132, 12796) a Co catalyst for the enantioselective cyclopropanation of α-olefins such as 3. Seiji Iwasa of Toyohashi University of Technology designed (Angew. Chem. Int. Ed. 2010, 49, 8439) a resin-bound Ru catalyst that could be used repeatedly for the enantioselective cyclization of the ester 6. Rai-Shung Lin of National Tsing-Hua University showed (Angew. Chem. Int. Ed. 2010, 49, 9891) that a gold catalyst could expand the alkyne 8 to the cyclobutene 9. Takao Ikariya of the Tokyo Institute of Technology reported (J. Am. Chem. Soc. 2010, 132, 16637) a detailed study of the enantioselective conjugate addition of malonate 11 to cyclopentenone 10. Vladimir A. D’yakonov of the Russian Academy of Sciences, Ufa, showed (Tetrahedron Lett. 2010, 51, 5886) that a cyclic alkyne 13 could be annulated to the cyclopentenone 14. Shunichi Hashimoto of Hokkaido University also designed (Angew. Chem. Int. Ed. 2010, 49, 6979) a resin-bound Rh catalyst that could also be used repeatedly for the enantioselective cyclization of the ester 15. Tushar Kanti Chakraborty of the Central Drug Research Institute used (Tetrahedron Lett. 2010, 51, 4425) Ti(III) to mediate the diastereoselective cyclization of 17 to 18. Alexandre Alexakis of the University of Geneva extended (Synlett 2010, 1694) enantioselective conjugate addition of isopropenyl to the more difficult enone 19. Joseph P. A. Harrity of the University of Sheffield showed (Org. Lett. 2010, 12, 4832) that Pd could catalyze the rearrangement of 21 to 22. Strategies for the controlled construction of polycyclic ring systems are also important. Günter Helmchen of the Universität Heidelberg showed (J. Org. Chem. 2010, 75, 7917) that 23 was efficiently cyclized to the diene with Pt catalyst. The reaction could be carried out in the presence of the dienophile 24 to give 25 directly.


Author(s):  
Douglass F. Taber

Alessandro Palmieri of the University of Camerino developed (Synlett 2010, 2468) the condensation of a nitro acrylate 1 with a 1,3-dicarbonyl partner 2 to give the furan 3. Chaozhong Li of the Shanghai Institute of Organic Chemistry showed (Tetrahedron Lett. 2010, 51, 3678) that an alkenyl halide 4 could be cyclized to the furan 5. Ayhan S. Demir of Middle East Technical University established (Chem. Commun. 2010, 46, 8032) that a Au catalyst could catalyze the addition of an amine 7 to a cyanoester 6 to give the pyrrole 8 . Bruce A. Arndtsen of McGill University effected (Org. Lett. 2010, 12, 4916) the net three-component coupling of an imine 9, an acid chloride 10, and an alkyne 11 to deliver the pyrrole 12. Bernard Delpech of CNRS Gif-sur-Yvette prepared (Org. Lett. 2010, 12, 4760) the pyridine 15 by combining the diene 13 with the incipient carbocation 14. Max Malacria, Vincent Gandon, and Corinne Aubert of UPMC Paris optimized (Synlett 2010, 2314) the internal Co-mediated cyclization of a nitrile alkyne 5 to the tetrasubstituted pyridine 17. Yoshiaki Nakao of Kyoto University and Tamejiro Hiyama, now at Chuo University, effected (J. Am. Chem. Soc. 2010, 132, 13666) selective substitution of a preformed pyridine 18 at the C-4 position by coupling with an alkene 19. We showed (J. Org. Chem. 2010, 75, 5737) that the anion from deprotonation of a pyridine 21 could be added in a conjugate sense to 22 to give 23. Other particularly useful strategies for further substitution of preformed pyridines have been described by Olafs Daugulis of the University of Houston (Org. Lett. 2010, 12, 4277), by Phil S. Baran of Scripps/La Jolla (J. Am. Chem. Soc. 2010, 132, 13194), and by Robert G. Bergmann of the University of California, Berkeley, and Jonathan A. Ellman of Yale University (J. Org. Chem. 2010, 75, 7863). K. C. Majumdar of the University of Kalyani developed (Tetrahedron Lett. 2010, 51, 3807) the oxidative Pd-catalyzed cylization of 24 to the indole 25. Nan Zheng of the University of Arkansas showed (Org. Lett. 2010, 12, 3736) that Fe could be used to catalyze the rearrangement of the azirine 26 to the indole 27.


Author(s):  
Douglass F. Taber

Jianbo Wang of Peking University described (Angew. Chem. Int. Ed. 2010, 49, 2028) the Au-promoted bromination of a benzene derivative such as 1 with N-bromosuccinimide. In a one-pot procedure, addition of a Cu catalyst followed by microwave heating delivered the aminated product 2. Jian-Ping Zou of Suzhou University and Wei Zhang of the University of Massachusetts, Boston, observed (Tetrahedron Lett. 2010, 51, 2639) that the phosphonylation of an arene 3 proceeded with substantial ortho selectivity. Yonghong Gu of the University of Science and Technology, Hefei, showed (Tetrahedron Lett. 2010, 51, 192) that an arylpropanoic acid 6 could be ortho hydroxylated with PIFA to give 7. Louis Fensterbank, Max Malacria, and Emmanuel Lacôte of UMPC Paris found (Angew. Chem. Int. Ed. 2010, 49, 2178) that a benzoic acid could be ortho aminated by way of the cyano amide 8. Daniel J. Weix of the University of Rochester developed (J. Am. Chem. Soc. 2010, 132, 920) a protocol for coupling an aryl iodide 10 with an alkyl iodide 11 to give 12. Professor Wang devised (Angew. Chem. Int. Ed. 2010, 49, 1139) a mechanistically intriguing alkyl carbonylation of an iodobenzene 10. This is presumably proceeding by way of the intermediate diazo alkane. Usually, benzonitriles are prepared by cyanation of the halo aromatic. Hideo Togo of Chiba University established (Synlett 2010, 1067) a protocol for the direct electrophilic cyanation of an electron-rich aromatic 15. Thomas E. Cole of San Diego State University observed (Tetrahedron Lett. 2010, 51, 3033) that an alkyl dimethyl borane, readily prepared by hydroboration of the alkene with BCl3 and Et3 SiH, reacted with benzoquinone 17 to give 18. Presumably this transformation could also be applied to substituted benzoquinones. When a highly substituted benzene derivative is needed, it is sometimes more economical to construct the aromatic ring. Joseph P. A. Harrity of the University of Sheffield and Gerhard Hilt of Philipps-Universität Marburg showed (J. Org. Chem. 2010, 75, 3893) that the Co-catalyzed Diels-Alder cyloaddition of alkynyl borinate 21 with a diene 20 proceeded with high regiocontrol, to give, after oxidation, the aryl borinate 22.


Author(s):  
Douglass F. Taber

Richard J. K. Taylor of the University of York employed (Tetrahedron Lett. 2011, 52, 2024) the Jørgensen protocol to add 2 to 1, to give the enantiomerically enriched cyclohexenone 3. Condensation of 3 with aqueous ammonia led directly to (-)-mearsine 4. Wei-Dong Z. Li of Nankai University found (Org. Lett. 2011, 13, 3538) that the intermediate from Dibal reduction of the lactone 5 underwent Nazarov cyclization, giving the α-hydroxy cyclopentenone 6. After acetylation, deprotection gave an amine that cyclized with high diastereocontrol, leading to (±)-cephalotaxine 7. Tony K. M. Shing of the Chinese University of Hong Kong cyclized (Org. Lett. 2011, 13, 2916) the aldehyde 8 by exposure to 9. The product 10 was carried on to (-)-cocaine 11, as well as several hydroxylated cocaine derivatives. Susumi Hatakeyama of Nagasaki University found (Tetrahedron Lett. 2011, 52, 923) that exposure of the simple prochiral aldehyde 12 to catalytic proline transformed it, after reduction, into the cyclized diol 13 in high ee. The diol 13 was readily carried on to quinine 14. M.-Lluïsa Bennasar of the University of Barcelona devised (Org. Lett. 2011, 13, 2042) Pd-catalyzed conditions for the cyclization of 15 that selectively delivered the unstable kinetic product 18. Selective hydrogenation of the more reactive bridgehead alkene then led to cleavamine 17. The alkene 16 is also prochiral, so it is possible that a catalyst could be found that would deliver 17 in high ee. The synthesis of the heptacyclic alkaloid strychnine 23 would, in the past, have been a major undertaking. Christopher D. Vanderwal of the University of California, Irvine, prepared (Chem. Sci. 2011, 2, 649) 23 in just six linear steps. The dienyl aldehyde 18 was available in two steps from tryptophyl bromide. Exposure to t -BuOK cyclized 18 to 19. N-deallylation followed by alkylation with 20 provided 21, setting the stage for a truly spectacular Brook rearrangement/conjugate addition, to give the Wieland-Gumlich aldehyde 22. The known condensation with malonic acid completed the preparation of 23.


Author(s):  
Douglass F. Taber

(+)-Pachastrissamine 3, also known as jaspine B, induces apoptosis in melanoma cells by a caspase-dependent pathway. Nobutaka Fujii and Hiroaki Ohno of Kyoyo University developed (J. Org. Chem. 2010, 75, 3831) a practical route to 3 based on the Pd-mediated cyclization of 1 to 2. Thomas G. Minehan of California State University, Northridge, optimized (Organic Lett. 2010, 12, 1580) the condensation of 5 with the bis-pivalate 4. This opened a general route to C-aryl glycosides, including aspalathin 6. (+)-Varitriol 11 is vinylogously related to 7. A key step in the synthesis of 11 reported (J. Org. Chem. 2010, 75, 2107) by Subhash Ghosh of the Indian Institute of Chemical Technology was the intermolecular Heck coupling of 8 with 9. Aspercyclide A 14 and its more stable methyl ether are promising lead compounds for the treatment of asthma. In the course of a synthesis of 14, Alan C. Spivey of Imperial College developed (Chem. Commun. 2010, 1824) the intramolecular Heck cyclization of 12 to 13. In a bold synthesis of etnangien 17, Dirk Menche of Ruprecht-Karls-Universität Heidelberg showed (J. Org. Chem. 2010, 75, 2429) that the intramolecular Heck coupling of 15 to 16 proceeded efficiently. The substituents on 15 may be favoring conformations that lead to cyclization.


Author(s):  
Douglass F. Taber

The challenge of controlling the relative and absolute configuration of highly substituted cyclic ether-containing natural products continues to stimulate the development of new synthetic methods. Masahiro Murakami of Kyoto University showed (J. Org. Chem. 2009, 74, 6050) that Rh-mediated addition of an aryl boronic acid to 1 proceeded with high syn diastereocontrol, giving 3. This set the stage for Au-mediated rearrangement, leading to 4. We found (J. Org. Chem. 2009, 74, 5516) that asymmetric epoxidation of 5 followed by exposure to AD-mix could be used to prepare each of the four diastereomers of 6. We carried 6 on the isofuran 7, using a stereodivergent strategy that allowed the preparation of each of the 32 enantiomerically pure diastereomers of the natural product. Following up on the synthesis of brevisamide 16 described (Organic Highlights, November 16, 2009) by Kazuo Tachibana of the University of Tokyo, three groups reported alternative total syntheses. James S. Panek of Boston University prepared (Organic Lett. 2009, 11, 4390) the cyclic ether of 16 by addition of the enantiomerically pure silane 9 to 8. Craig W. Lindsley of Vanderbilt University used (Organic Lett. 2009, 11, 3950) SmI2 to effect the cyclization of 11 to 12. Arun K. Ghosh of Purdue University employed (Organic Lett. 2009, 11, 4164) an enantiomerically pure Cr catalyst to direct the absolute configuration in the hetero Diels-Alder addition of 14 to 13. Rubottom oxidation of the enol ether so formed led to the α-hydroxy ketone 15. Yuji Mori of Meijo University described (Organic Lett. 2009, 11, 4382) the total synthesis of the Gambierdiscus toxicus ladder ether gambierol 19. A key strategy, used repeatedly through the sequence, was the exo cyclization of an epoxy sulfone, illustrated by the conversion of 17 to 18. The epoxy sulfones were prepared by alkylating the anions derived from preformed epoxy sulfones such as 20.


Sign in / Sign up

Export Citation Format

Share Document