Organic Synthesis
Latest Publications


TOTAL DOCUMENTS

104
(FIVE YEARS 0)

H-INDEX

0
(FIVE YEARS 0)

Published By Oxford University Press

9780190646165, 9780197559734

Author(s):  
Douglass F. Taber

Paclitaxel (Taxol®) 3 is widely used in the clinical treatment of a variety of cancers. Takaaki Sato and Noritaka Chida of Keio University envisioned (Org. Lett. 2015, 17, 2570, 2574) establishing the central eight-membered ring of 3 by the SmI2-mediated cyclization of 1 to 2. The starting point for the synthesis was the enantiomerically-pure enone 5, pre­pared from the carbohydrate precursor 4. Conjugate addition to 5 proceeded anti to the benzyloxy substituent to give, after trapping with formaldehyde and protection, the ketone 6. Reduction and protection followed by hydroboration led to 7, that was, after protection and deprotection, oxidized to 8. The second ring of 3 was added in the form of the alkenyl lithium derivative 9, prepared from the trisylhydrazone of the corresponding ketone. Hydroxyl-directed epoxidation of 10 proceeded with high facial selectivity, leading, after reduction and protection, to the cyclic carbonate 11. Allylic oxidation converted the alkene into the enone, while at the same time oxidizing the benzyl protecting group to the ben­zoate, to give 12. Reduction of the ketone 12 led to a mixture of diastereomers. In practice, only one of the diastereomers of 1 cyclized cleanly to 2, as illustrated, so the undesired diastereomer from the NaBH4 reduction was oxidized back to the enone for recycling. For convenience, only one of the diastereomers of 2 was carried forward. To establish the tetrasubstituted alkene of 3, the alkene of 2 was converted to the cis diol and on to the bis xanthate 13. Warming to 50°C led to the desired tet­rasubstituted alkene, sparing the oxygenation that is eventually required for 3. For convenience, to intercept 16, the intermediate in the Takahashi total synthesis, both xanthates were eliminated to give 14. Hydrogenation removed the disubsti­tuted alkene, and also deprotected the benzyl ether. Oxidation followed by Peterson alkene formation led to 15, that was carried on to the Takahashi intermediate 16 using the now-standard protocol for oxetane construction. It is a measure of the strength of the science of organic synthesis that Masahisa Nakada of Waseda University also reported (Chem. Eur. J. 2015, 21, 355) an elegant synthesis of 3 (not illustrated).


Author(s):  
Douglass F. Taber

Disorazole C1 3, isolated from fermentation of the myxobacterium Sorangium cellu­losum, shows antifungal and anticancer activity. Amir H. Hoveyda of Boston College applied (J. Am. Chem. Soc. 2014, 136, 16136) recent advances in alkene metathesis from his group to enable the efficient assembly of 2 and so of 3. The ester 1 was assembled from the alcohol 11 and the acid 18. The preparation of 11 began with the enantioselective addition of 5 to 4 to give 6 and then 7, as described by Kalesse (Angew. Chem. Int. Ed. 2010, 49, 1619). Leighton allylation led to 8, that was then coupled with 9 to give 10 with high Z selectivity. Iodination of 10 followed by deprotection then completed the assembly of 11. The starting material for the acid 18 was the allylic alcohol 13. As reported by Cramer (Angew. Chem. Int. Ed. 2008, 47, 6483), exposure of the racemic alcohol 12 to vinyl acetate in the presence of Amano lipase PS converted one enantiomer to the acetate, leaving 13. Methylation of the secondary alcohol followed by acid-mediated removal of the t-butyl ester led to the acid 14, that was converted to the correspond­ing acyl fluoride and coupled with serine Me ester 15 to give 16. After cyclization to the oxazole 17, cross metathesis with five equivalents of 4-bromo-1-butene gave the homoallylic bromide, that was readily eliminated with DBU to give, after saponifica­tion, the acid 18. The cross metathesis of the coupled ester 1, a polyene, with 9 proceeded with remarkable selectivity to give 2, again as the Z geometric isomer. On exposure to the Heck catalyst Pd [(o-tolyl)3P]2, 2 dimerized efficiently. The deprotection was not straightforward, but conditions (H2SiF6, CH3OH, 4°C, 72 h) were found that deliv­ered 3 in 68% yield.


Author(s):  
Douglass F. Taber

The amphidinolides, having zero, one, or (as exemplified by amphidinolide F 3) two tetrahydrofuran rings, have shown interesting antineoplastic activity. It is a tribute to his development of robust Mo catalysts for alkyne metathesis that Alois Fürstner of the Max-Planck-Institut für Kohlenforschung Mülheim could with confidence design (Angew. Chem. Int. Ed. 2013, 52, 9534) a route to 3 that relied on the ring-closing metathesis of 1 to 2 very late in the synthesis. Three components were prepared for the assembly of 1. Julia had already reported (J. Organomet. Chem. 1989, 379, 201) the preparation of the E bromodiene 5 from the sulfone 4. The alcohol 7 was available by the opening of the enantiomerically-pure epoxide 6 with propynyl lithium, followed by oxidation following the Pagenkopf pro­tocol. Amino alcohol-directed addition of the organozinc derived from 5 to the alde­hyde from oxidation of 7 completed the assembly of 8. Addition of the enantiomer 10 of the Marshall butynyl reagent to 9 followed by protection, oxidation to 11, and addition of, conveniently, the other Marshall enan­tiomer 12 led to the protected diol 13. Silylcupration–methylation of the free alkyne set the stage for selective desilylation and methylation of the other alkyne. Iodination then completed the trisubstituted alkene of 14. Methylation of the crystalline lactone 15, readily prepared from D-glutamic acid, led to a mixture of diastereomers. Deprotonation of that product followed by an aque­ous quench delivered 16. Reduction followed by reaction with the phosphorane 17 gave the unsaturated ester, that cyclized with TBAF to the crystalline 18. The last ste­reogenic center of 22 was established by proline-mediated aldol condensation of the aldehyde 19 with the ketone 20. To assemble the three fragments, the ketone of 21 was converted to the enol triflate and thence to the alkenyl stannane. Saponification gave the free acid 22, that was acti­vated, then esterified with the alcohol 18. Coupling of the stannane with the iodide 14 followed by removal of the TES group led to the desired diyne 1. It is noteworthy that the Mo metathesis catalyst is stable enough to tolerate the free alcohol of 1 in the cyclization to 2.


Author(s):  
Douglass F. Taber

The early promise for the biological activity of the derivatives of ingenol 3 has been borne out by the clinical efficacy of the derived angelate, recently approved by the US Food and Drug Administration for the treatment of actinic keratosis. Phil S. Baran of Scripps La Jolla envisioned (Science 2013, 341, 878) a route to 3 based on a rearrange­ment of 2, available by the Pauson–Khand cyclization of the allenyl alkyne 1. One of the partners for the preparation of 1 was available following the Sugai (Synlett 1997, 1297) procedure, by the Claisen rearrangement of triethyl orthopro­pionate 5 with the propargyl alcohol 4 to give 6. Reduction delivered a racemic mix­ture of alcohols. On exposure of the mixture to vinyl acetate and Pseudomonas cepacia lipase, the undesired enantiomer was selectively acetylated to 7, leaving residual 8 of high ee. IBX was found by the Scripps group to be effective at oxidizing 8 without racemization. The other component of 1 was prepared from the inexpensive (+)-3-carene 10. Chlorination followed by ozonolysis delivered 11, that was reduced to the enolate, then alkylated with methyl iodide. Exposure to LiHMDS gave a new enolate, that was added to the aldehyde 9 to give 12. Addition of ethynyl magnesium bromide to the now more open face of 12 proceeded with high diastereoselectivity. Selective silylation of the secondary alcohol followed by silylation of the tertiary alcohol set the stage for the Pauson–Khand cyclization. Following the Brummond protocol, 1 was cyclized to 2. Methyl magnesium bro­mide was added, again to the more open face of the ketone, to give a new tertiary alco­hol. Exposure to stoichiometric OsO4 converted the more available alkene to the cis diol, that was protected as its cyclic carbonate 13. A central challenge in the total synthesis of the ingenanes is the construction of the “inside–outside” skeleton. This was achieved by the pinacol rearrangement of 13 with BF3•OEt2, to give 14. All that remained to complete the synthesis was selective oxidation. Allylic oxi­dation with stoichiometric SeO2 installed the secondary alcohol, that was acety­lated to give 15. The other secondary alcohol was then freed, and dehydrated with the Martin sulfurane, to give 16. A last allylic oxidation completed the synthesis of ingenol 3.


Author(s):  
Douglass F. Taber

Djamaladdin G. Musaev and Huw M. L. Davies of Emory University effected (Chem. Sci. 2013, 4, 2844) enantioselective cyclopropanation of ethyl acrylate 2 with the α-diazo ester 1 to give 3 in high ee. Philippe Compain of the Université de Strasbourg used (J. Org. Chem. 2013, 78, 6751) SmI2 to cyclize 4 to the cyclobutanol 5. Jianrong (Steve) Zhou of Nanyang Technological University effected (Chem. Commun. 2013, 49, 11758) enantioselective Heck addition of 7 to the prochiral ester 6 to give the cyclopentene 8. Liu-Zhu Gong of USTC, Hefei added (Org. Lett. 2013, 15, 3958) the Rh enolate from the enantioselective ring expansion of the α-diazo ester 9 to the nitroalkene 10, to give 11 in high de. Stephen P. Fletcher of the University of Oxford set (Angew. Chem. Int. Ed. 2013, 52, 7995) the cyclic quaternary center of 14 by the enantioselective conjugate addition to 12 of the alkyl zirconocene derived from 13. Alexandre Alexakis of the University of Geneva reported (Chem. Eur. J. 2013, 19, 15226) high ee from the conjugate addition of alkenyl Al reagents (not illustrated) to 12. Paultheo von Zezschwitz of Philipps-Universität Marburg prepared (Adv. Synth. Catal. 2013, 355, 2651) the silyl enol ether 17 by trapping the intermediate from the conjugate addition of 16 to 15. Stefan Bräse of the Karlsruhe Institute of Technology effected (Eur. J. Org. Chem. 2013, 7110) conjugate addition to the prochiral dienone 18 to give the highly substi­tuted cyclohexenone 19. Ping Tian and Guo-Qiang Lin of the Shanghai Institute of Organic Chemistry cyclized (J. Am. Chem. Soc. 2013, 135, 11700) 20 to the kinetic, less stable epimer of the diketone 21. Rh-mediated intramolecular C–H insertion has been a powerful tool for the con­struction of cyclopentane derivatives. Douglass F. Taber of the University of Delaware found (J. Org. Chem. 2013, 78, 9772) that the Rh carbene derived from 22 was dis­criminating enough to target the more nucleophilic C–H bond, leading to the cyclohexanone 23. Kozo Shishido of the University of Tokushima observed (Org. Lett. 2013, 15, 3666) high diastereoselectivity in the intramolecular Heck cyclization of 24 to 25.


Author(s):  
Douglass F. Taber

M. Kevin Brown of Indiana University prepared (J. Am. Chem. Soc. 2015, 137, 3482) the cyclobutane 3 by the organocatalyzed addition of 2 to the alkene 1. Karl Anker Jørgensen of Aarhus University assembled (J. Am. Chem. Soc. 2015, 137, 1685) the complex cyclobutane 7 by the addition of 5 to the acceptor 4, followed by conden­sation with the phosphorane 6. Zhi Li of the National University of Singapore balanced (ACS Catal. 2015, 5, 51) three enzymes to effect enantioselective opening of the epoxide 8 followed by air oxidation to 9. Gang Zhao of the Shanghai Institute of Organic Chemistry and Zhong Li of the East China University of Science and Technology added (Org. Lett. 2015, 17, 688) 10 to 11 to give 12 in high ee. Akkattu T. Biju of the National Chemical Laboratory combined (Chem. Commun. 2015, 51, 9559) 13 with 14 to give the β-lactone 15. Paul Ha-Yeon Cheong of Oregon State University and Karl A. Scheidt of Northwestern University reported (Chem. Commun. 2015, 51, 2690) related results. Dieter Enders of RWTH Aachen University constructed (Chem. Eur. J. 2015, 21, 1004) the complex cyclopentane 20 by the controlled com­bination of 16, 17, and 18, followed by addition of the phosphorane 19. Derek R. Boyd and Paul J. Stevenson of Queen’s University Belfast showed (J. Org. Chem. 2015, 80, 3429) that the product from the microbial oxidation of 21 could be protected as the acetonide 22. Ignacio Carrera of the Universidad de la República described (Org. Lett. 2015, 17, 684) the related oxidation of benzyl azide (not illustrated). Manfred T. Reetz of the Max-Planck-Institut für Kohlenforschung and the Philipps-Universität Marburg found (Angew. Chem. Int. Ed. 2014, 53, 8659) that cytochrome P450 could oxidize the cyclohexane 23 to the cyclohexanol 24. F. Dean Toste of the University of California, Berkeley aminated (J. Am. Chem. Soc. 2015, 137, 3205) the ketone 25 with 26 to give 27. Benjamin List, also of the Max-Planck-Institut für Kohlenforschung, reported (Synlett 2015, 26, 1413) a parallel investigation. Philip Kraft of Givaudan Schweiz AG and Professor List added (Angew. Chem. Int. Ed. 2015, 54, 1960) 28 to 29 to give 30 in high ee.


Author(s):  
Douglass F. Taber

Kyungsoo Oh of Chung-Ang University cyclized (Org. Lett. 2015, 17, 450) the chloro enone 1 with NBS to the furan 2. Hongwei Zhou of Zhejiang University acylated (Adv. Synth. Catal. 2015, 357, 389) the imine 3, leading to the furan 4. H. Surya Prakash Rao of Pondicherry University found (Synlett 2014, 26, 1059) that under Blaise conditions, exposure of 5 to three equivalents of 6 led to the pyrrole 7. Yoshiaki Nishibayashi of the University of Tokyo and Yoshihiro Miyake, now at Nagoya University, prepared (Chem. Commun. 2014, 50, 8900) the pyrrole 10 by adding the silane 9 to the enone 8. Barry M. Trost of Stanford University developed (Org. Lett. 2015, 17, 1433) the phosphine-mediated cyclization of 11 to an intermediate that on brief exposure to a Pd catalyst was converted to the pyridine 12. Nagatoshi Nishiwaki of the Kochi University of Technology added (Chem. Lett. 2015, 44, 776) the dinitrolactam 14 to the enone 13 to give the pyridine 15. Metin Balci of the Middle East Technical University assembled (Org. Lett. 2015, 17, 964) the tricyclic pyridine 18 by adding propargyl amine 17 to the aldehyde 16. Chada Raji Reddy of the Indian Institute of Chemical Technology cyclized (Org. Lett. 2015, 17, 896) the azido enyne 19 to the pyridine 20 by simple exposure to I2. Björn C. G. Söderberg of West Virginia University used (J. Org. Chem. 2015, 80, 4783) a Pd catalyst to simultaneously reduce and cyclize 21 to the indole 22. Ranjan Jana of the Indian Institute of Chemical Biology effected (Org. Lett. 2015, 17, 672) sequential ortho C–H activation and cyclization, adding 23 to 24 to give the 2-substituted indole 25. In a complementary approach, Debabrata Maiti of the Indian Institute of Technology Bombay added (Chem. Eur. J. 2015, 21, 8723) 27 to 26 to give the 3-substituted indole 28. In a Type 8 construction, Nobutaka Fujii and Hiroaki Ohno of Kyoto University employed (Chem. Eur. J. 2015, 21, 1463) a gold catalyst to add 30 to 29, leading to 31.


Author(s):  
Douglass F. Taber

Xin-Yan Wu of East China University of Science and Technology and Jun Yang of the Shanghai Institute of Organic Chemistry added (Tetrahedron Lett. 2014, 55, 4071) the Grignard reagent 1 to propargyl alcohol 2 to give an intermediate that could be bory­lated, then coupled under Pd catalysis with an anhydride, leading to the furan 3. Fuwei Li of the Lanzhou Institute of Chemical Physics constructed (Org. Lett. 2014, 16, 5992) the furan 6 by oxidizing the keto ester 4 in the presence of the enamide 5. Yuanhong Liu of the Shanghai Institute of Organic Chemistry prepared (Angew. Chem. Int. Ed. 2014, 53, 11596) the pyrrole 9 by reducing the azadiene 7 with the Negishi reagent, then adding the nitrile 8. Yefeng Tang of Tsinghua University found (Tetrahedron Lett. 2014, 55, 6455) that the Rh carbene derived from 11 could be added to an enol silyl ether 10 to give the pyrrole 12. Pazhamalai Anbarasan of the Indian Institute of Technology Madras reported (J. Org. Chem. 2014, 79, 8428) related results. Zheng Huang of the Shanghai Institute of Organic Chemistry established (Angew. Chem. Int. Ed. 2014, 53, 1390) a connection between substituted piperidines and pyridines by dehydrogenating 13 to 15, with 14 as the acceptor. Joseph P. A. Harrity of the University of Sheffield conceived (Chem. Eur. J. 2014, 20, 12889) the cascade assembly of the pyridine 18 by cycloaddition of 16 with 17 followed by Pd-catalyzed coupling. Teck-Peng Loh of Nanyang Technological University converted (Org. Lett. 2014, 16, 3432) the keto ester 19 into the azirine, then eliminated it to form an aza­triene that cyclized to the pyridine 20. En route to a cholesteryl ester transfer protein inhibitor, Zhengxu S. Han of Boehringer Ingelheim combined (Org. Lett. 2014, 16, 4142) 21 with 22 to give an intermediate that could be oxidized to 23. Magnus Rueping of RWTH Aachen used (Angew. Chem. Int. Ed. 2014, 53, 13264) an Ir photoredox catalyst in conjunction with a Pd catalyst to cyclize the enamine 24 to the indole 25. Yingming Yao and Yingsheng Zhao of Soochow University effected (Angew. Chem. Int. Ed. 2014, 53, 9884) oxidative cyclization of 26 to 27.


Author(s):  
Douglass F. Taber

Stephen G. DiMagno of the University of Nebraska developed (Chem. Eur. J. 2015, 21, 6394) a protocol for the clean monoiodination of 1 to 2. The bromomethylation (or chloromethylation, with HCl) of a benzene derivative is straightforward with formal­dehyde and HBr. Naofumi Tsukada of Shizuoka University designed (Organometallics 2015, 34, 1191) a Cu catalyst that mediated the coupling of an alkyne with the benzyl bromide so produced, effecting net propargylation of 3 with 4 to give 5. Triazenes such as 7, versatile intermediates for organic synthesis, are usually prepared by diazotization of the corresponding aniline. Kay Severin of the Ecole Polytechnique Fédérale de Lausanne established (Angew. Chem. Int. Ed. 2015, 54, 302) an alternative route from the aryl Grignard reagent 6. Ping Lu and Yanguang Wang of Zhejiang University showed (Chem. Commun. 2015, 51, 2840) that dimethylformamide could serve as the carbon source for the conversion of 8 to the nitrile 9. Junha Jeon of the University of Texas at Arlington effected (J. Org. Chem. 2015, 80, 4661; Chem. Commun. 2015, 51, 3778) the reductive ortho silylation of 10 to give 11. Vladimir Gevorgyan of the University of Illinois at Chicago found (Angew. Chem. Int. Ed. 2015, 54, 2255) that the phenol derivative 12 could be ortho carboxylated, leading to 13. Lutz Ackermann of the Georg-August-Universität Göttingen, starting (Chem. Eur. J. 2015, 21, 8812) with the designed amide 14, effected ortho metala­tion followed by coupling, to give the methylated product 15. Tetsuya Satoh and Masahiro Miura of Osaka University used (Org. Lett. 2015, 17, 704) the dithiane of 16 to direct ortho metalation. Coupling with acrylate followed by reductive desulfu­rization led to the ester 17. Jin-Quan Yu of Scripps/La Jolla designed (Angew. Chem. Int. Ed. 2015, 54, 888) the phenylacetamide 18 to direct selective meta metalation, leading to the unsat­urated aldehyde 19. In an extension of the Catellani protocol, Guangbin Dong of the University of Texas prepared (J. Am. Chem. Soc. 2015, 137, 5887) the biphenyl 21 by net meta metalation of the benzylamine 20. Several methods for the de novo assembly of benzene derivatives have recently been put forward. Rajeev S. Menon of the Indian Institute of Chemical Technology condensed (Org. Lett. 2015, 17, 1449) the unsaturated aldehyde 22 with the sulfonyl ester 23 to give 24.


Author(s):  
Douglass F. Taber

Jin Kun Cha of Wayne State University prepared (Org. Lett. 2014, 16, 6208) the allene 1 by SN2′ coupling of a cyclopropanol with a propargylic tosylate. Silver-mediated cyclization converted 1 into 2, that was reduced with diimide to the Dendrobates alka­loid indolizidine 223AB 3. Sanghee Kim of Seoul National University observed (Chem. Eur. J. 2014, 20, 17433) high diastereoselectivity in the Ireland–Claisen rearrangement of 4 to 5. The acid 5 was the key intermediate for the synthesis of the tunicate alkaloid lepadiformine 6. Tohru Fukuyama of Nagoya University also used (Eur. J. Org. Chem. 2014, 4823) an ester enolate Claisen rearrangement to set the relative and absolute configuration of 7. Pd-catalyzed cyclization then led to 8, that was carried on to the excitatory amino acid receptor agonist kainic acid 9. Gephyrotoxin 12 was so named because it incorporates structural elements from two different classes of the Dendrobates alkaloids. Martin D. Smith of the University of Oxford envisioned (Angew. Chem. Int. Ed. 2014, 53, 13826) the cascade cyclization of deprotected 10 to give, after reduction, the ketone 11. Zhen Yang of the Peking University Shenzhen Graduate School showed (Chem. Eur. J. 2014, 20, 12881) that the Rh carbene derived from 13 readily cyclized to an imine. The facial selectivity of the addition of the Grignard reagent 14 to that imine depended on the temperature of the reaction. At room temperature, 15 was formed. At low temperature, the other diastereomer predominated. Ring-closing metathesis was used for the elaboration of 15 to the Stemona alkaloid tuberostemospiroline 16. Kevin A. Reynolds of Portland State University prepared (J. Org. Chem. 2014, 79, 11674) 19 by condensation of the pyrrole 17 with the aldehyde 18. The biosyn­thetic enzyme, that they had overexpressed, oxidized 19 to the antimalarial alkaloid permarineosin A 20.


Sign in / Sign up

Export Citation Format

Share Document