The Lee Synthesis of (−)-Crinipellin A

Author(s):  
Douglass F. Taber

The crinipellins are the only tetraquinane natural products. The enone crinipellins, including crinipellin A 3, have anticancer activity. Hee-Yoon Lee of the Korea Advanced Institute of Science and Technology (KAIST) envisioned (J. Am. Chem. Soc. 2014, 136, 10274) the assembly of 2 and thus 3 by the intramolecular dipolar cycloaddition of the diazoalkane derived from the tosylhydrazone 1. The initial cyclopentene was prepared from commercial 4 following the Williams procedure. Conjugate addition of the Grignard reagent 5 in the presence of TMS-Cl led to the silyl enol ether 6. Regeneration of the enolate followed by allylation gave 7. The preparation of the racemic ketone was completed by ozonolysis followed by selec­tive reduction and protection. Addition of hydride in an absolute sense led to separa­ble 1:1 mixture of diastereomers. Reoxidation of one of the diastereomers delivered enantiomerically enriched 8. A few steps later, after coupling with 10, the sidechain stereocenter was set by Sharpless asymmetric epoxidation. Oxidation of 11 gave the aldehyde, that was converted to the alkyne 12 by the Ohira protocol. Addition of the Grignard reagent 13 gave the allene 14 as an inconse­quential 1:1 mixture of diastereomers. Deprotection then led to the tosylhydrazone 1. The transformation of 1 to 2 proceeded by initial formation of the diazo alkane 15. Intramolecular dipolar cycloaddition gave 16, that lost N2 to give the trimethylene–methane diradical 17. The insertion into the distal alkene proceeded with remarkable stereocontrol, to give 2 as a single diastereomer—in 87% yield from 1. Direct α-hydroxylation of the ketone derived from 2 gave the wrong diastereo­mer, and hydride addition to 18 reduced the wrong ketone. As an alternative, the enantiomerically-pure sulfoximine anion was added to the more reactive ketone, and the product was reduced and protected to give 19. Allylic oxidation converted the alkene to the enone, and heating to reflux in toluene reversed the sulfoximine addi­tion, leading to 20. Epoxidation of 20 followed by α-methylenation delivered the enone 21, that proved to be particularly sensitive. Eventually, success was found with TASF. With a similarly sensitive substrate, Douglass F. Taber of the University of Delaware observed (J. Am. Chem. Soc. 1998, 120, 13285) that TBAF in THF buffered with solid NH4Cl worked well.

Author(s):  
Douglass Taber

Alkene metathesis has been used to prepare more and more challenging natural products. The first and second generation Grubbs catalysts 1 and 2 and the Hoveyda catalyst 3 are the most widely used. Daesung Lee of the University of Illinois at Chicago designed (Organic Lett. 2008, 10, 257) a clever chain-walking cross metathesis, combining 4 and 5 to make 6. The diyne 3 was carried on (3R, 9R, 10R )-Panaxytriol 7. Tatsushi Imahori and Hiroki Takahata of Tohoku Pharmaceutical University found (Tetrahedron Lett. 2008, 49, 265) that of the several derivatives investigated, the unprotected alcohol 8 cyclized most efficiently. Selective cleavage of the monosubstituted alkene followed by hydroboration delivered the alkaloid Isofagomine 10. Brian M. Stoltz of Caltech established (J. Am. Chem. Soc. 2008 , 130 , 810) the absolute configuration of the halogenated chamigrene Elatol 14 using the enantioselective enolate allylation that he had previously devised. A key feature of this synthesis was the stereocontrolled preparation of the cis bromohydrin. Marc L. Snapper of Boston College opened (J. Org. Chem. 2008, 73, 3754) the strained cyclobutene 15 with ethylene to give the diene 16. Remarkably, cross metathesis with 17 delivered 18 with high regioselectivity, setting the stage for the preparation of the 5-F2t - Isoprostane 19. Derrick L. J. Clive of the University of Alberta assembled (J. Org. Chem. 2008, 73, 3078) Ottelione B 26 from the enantiomerically-pure aldehyde 20. Conjugate addition of the Grignard reagent 21 derived from chloroprene gave the kinetic product 22, that was equilibrated to the more stable 23. Addition of vinyl Grignard followed by selective ring-closing metathesis then led to 26.


Author(s):  
Douglass Taber

The tetracyclic Lycopodium alkaloid fawcettimine 3 and its derivatives are of interest as inhibitors of acetylcholine esterase. F. Dean Toste of the University of California, Berkeley recently reported (Angew. Chem. Int. Ed. 2007, 46, 7671) the first enantioselective synthesis of 3. The key to the synthesis was the rapid assembly of the enantiomerically-enriched hydrindane 2. The preparation of 2 began with the enantioselective Robinson annulation of the β-keto ester 4 with crotonaldehyde 5, mediated by the organocatalyst 6. In this protocol, originally developed by Karl Anker Jørgensen, the single stereogenic center was established by conjugate addition, presumably to the chiral iminium salt generated by the condensation of 5 with 6. Subsequent aldol (or more likely Mannich) cyclization followed by elimination gave 7. Hydrolysis and decarboxylation by heating with p-TsOH converted 7 to 1. This procedure was robust enough to allow preparation of a ten gram batch of 1. This Jørgensen annulation is the current method of choice for the enantioselective preparation of 2,5-dialkyl cyclohexenones. Conjugate addition of the propargyl anion equivalent 8 to 1 proceeded with the expected > 95:5 axial diastereoselectivity, to give the silyl enol ether 9. Exposure of the derived iodide 10 to catalytic [Ph3 PAu]Cl and AgBF4 induced smooth cyclization to the cis hydrindane 2. Before constructing the nine-membered ring amine of fawcettimine 3, it was first necessary to protect the ketone as the ketal. Pd-mediated coupling of the alkenyl iodide with the organoborane derived from 11 then proceeded smoothly, as did the subsequent hydroboration of the terminal alkene. Neither the mesylate nor the tosylate derived from 12 could be induced to cyclize. In contrast, intramolecular displacement of the iodide proceeded well, to give 13. Hydroboration followed by oxidation then gave 15, which on deprotection cyclized to (+)-fawcettimine 3. Several aspects of this synthesis are attractive. While the stereochemical outcome of the hydroboration of 14 could not necessarily be predicted with confidence, in fact it did not matter, as the stereogenic center adjacent to the ketone could be epimerized under the trifluoroacetic acid deprotection conditions, and only the desired diastereomer would be able to add in an intramolecular fashion to the cyclohexanone.


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

Thorsten Bach of the Technische Universität München used (Chem. Commun. 2014, 50, 3353) the chiral medium-mediated photochemical 2+2 cycloaddition that he devel­oped to prepare 3 by combining 1 with 2. Oxidative cleavage led to (−)-pinolinone 4. Robert A. Batey of the University of Toronto rearranged (Angew. Chem. Int. Ed. 2013, 52, 10862) furfural 5 in the presence of 6 to give the enone 7. Acylation fol­lowed by intramolecular conjugate addition delivered agelastatin A 8. Hee-Yoon Lee of KAIST prepared (Org. Lett. 2014, 16, 2466) the tosylhydrazone Na salt 9 from citronellal. Thermolysis led, via a dialkyl diazo intermediate, to the tricy­clic 10. Direct comparison of synthetic material with the natural product panaginsene 11 enabled the assignment of the relative configuration of the pendant methyl group. Hanfeng Ding of Zhejiang University eliminated (Org. Lett. 2014, 16, 3376) HBr from 12 to give, after rearrangement, the cycloheptadiene salvileucalin D 13. Irradiation converted 13 to the cyclobutene salvileucalin C 14. In a recent chapter of his continuing work on the morphine alkaloids, Tomas Hudlicky of Brock University described (Adv. Synth. Catal. 2014, 356, 333) the intra­molecular [3+2] cycloaddition of the nitrone derived from 15 to give 16. This was readily carried on to ent-codeine 17. Xingang Xie and Xuegong She of Lanzhou University used (Org. Lett. 2014, 16, 1996) Shi epoxidation and Itsuno–Corey reduction to prepare 18 in enantiomerically-pure form. Cationic cyclization converted 18 to 19, that was oxidized to (−)-walsucochin B 20.


Author(s):  
Douglass F. Taber

Following the Szymoniak protocol, Morwenna S. M. Pearson-Long and Philippe Bertus of the Université du Maine added (Synthesis 2015, 47, 992) the Grignard rea­gent 2 to the nitrile 1 to give the cyclopropyl amine 3. Chen-Guo Feng of the Shanghai Institute of Organic Chemistry prepared (Chem. Commun. 2015, 51, 8773) the cyclobutane 6 by enantioselective conjugate addition of 5 to the unsaturated ester 4. Martin Kotora of Charles University showed (Eur. J. Org. Chem. 2015, 2868) that the zirconacycle from the eneyne 7 reacted with the aldehyde 8 to give, after iodina­tion, the alcohol 9. Xiaoming Feng of Sichuan University used (Angew. Chem. Int. Ed. 2015, 54, 1608) a scandium catalyst to effect the intramolecular Roskamp cyclization of 10 to 11. Celia Dominguez of CHDI observed (Org. Lett. 2015, 17, 1401) that the double alkylation of the ester 12 with the dibromide 13 proceeded with high diaste­reoselectivity, to give 14. Hirokazu Tsukamoto of Tohoku University cyclized (Chem. Commun. 2015, 51, 8027) 15 to 16 in high ee. Daniel J. Weix of the University of Rochester found (J. Am. Chem. Soc. 2015, 137, 3237) that under the influence of an enantiomerically-pure Ti catalyst, the organon­ickel species derived from 18 opened the prochiral epoxide 17 to give 19 in high ee. John F. Bower of the University of Bristol optimized (J. Am. Chem. Soc. 2015, 137, 463) conditions for the highly diastereoselective Rh-mediated cyclocarbonylation of 20 to 21. Margaret A. Brimble of the University of Auckland initiated (J. Org. Chem. 2015, 80, 2231) the construction of the cyclohexenone 24 by the diastereoselective addition of 23 to the unsaturated ester 22. Olivier Baslé and Marc Maduit of ENSC Rennes devised (Chem. Eur. J. 2015, 21, 993) conditions for the preparation of 26 by enantioselective conjugate addition to the cyclohexenone 25. Yoshito Kishi of Harvard University demonstrated (Tetrahedron Lett. 2015, 56, 3220) that the carbenoid generated from the epoxide 27 cyclized to 28 with high dia­stereoselectivity. Wenjun Tang, also of the Shanghai Institute of Organic Chemistry, developed (Angew. Chem. Int. Ed. 2015, 54, 3033) a Pd catalyst for the diastereoselec­tive (because it is enantioselective) cyclization of 29 to 30.


Author(s):  
Douglass Taber

Ryoichi Kuwano of Kyushu University showed (J. Am. Chem. Soc. 2008, 130, 808) that diastereomerically and enantiomerically pure pyrollidines such as 2 could be prepared by hydrogenation of the corresponding pyrrole. Victor S. Martín of Universidad de la Laguna found (Organic Lett. 2008, 10, 2349) that the stereochemical outcome of the pyrrolidine-forming Nicholas cyclization could be directed by the protecting group on the N. Jianbo Wang of Peking University established (J. Org. Chem. 2008, 73, 1971) a convenient route to diazo esters such as 6. N-H insertion led to the pyrrolidine, which Zhen-Jiang Xu of the Shanghai Institute of Organic Chemistry and Chi-Ming Che of the University of Hong Kong showed (Organic Lett. 2008, 10, 1529) could be reduced with high diastereoselectivity to the hydroxy ester 7. Alternatively, Professor Wang found that photochemical Wolff rearrangement of 6 delivered the pyrrolidone 8 . Martin J. Slater and Shiping Xie of GlaxoSmithKline optimized (J. Org. Chem. 2008, 73, 3094) the hydroquinine catalyzed enantioselective 3+2 cycloaddition of 9 and 10, leading to the pyrrolidine 11 with high diastereocontrol. Shu Kobayashi of the University of Tokyo developed (Adv. Synth. Cat. 2008, 350, 647) a practical protocol for the aza Diels-Alder construction of enantiomerically-pure piperidines such as 14 . Biao Yu of the Shanghai Institute of Organic Chemistry cyclized (Tetrahedron Lett. 2008, 49, 672) the product from the proline-catalyzed enantioselective aldol of 15 and 16, leading to the substituted piperidine 17 . Michael Shipman of the University of Warwick described (Tetrahedron Lett. 2008, 49, 250) the cyclization of the aziridine derived from 18, that proceeded to give 19 as a single diastereomer, apparently via kinetic side-chain protonation. Takeo Kawabata of Kyoto University found (J. Am. Chem. Soc. 2008, 130, 4153) that intramolecular alkylation to form four, five and six-membered rings from amino esters such as 21 proceeded with remarkable enantioretention. Géraldine Masson and Jieping Zhu of CNRS, Gif-sur-Yvette, condensed (Organic Lett. 2008, 10, 1509) cinnamaldehyde 23 with cyanide and an ω-alkenyl amine to give the intramolecular aza-Diels-Alder substrate 24. Hongbin Zhai of the Shanghai Institute of Organic Chemistry acylated (J. Org. Chem. 2008, 73, 3589) 26 with 27, leading to the ring-closing metathesis precursor 28.


Author(s):  
Douglass F. Taber

The compact and highly functionalized Gelsemium alkaloids, exemplified by gelsemine (OHL20060403) and gelsemoxonine 3, offer a substantial challenge. The cytotoxicity of closely related alkaloids adds to the interest in this class. Tohru Fukuyama of the University of Tokyo envisioned (J. Am. Chem. Soc. 2011, 133, 17634) that cyclopropane-accelerated Cope rearrangement of 1 could deliver 2, ready for further functionalization to 3. The starting material for the synthesis was the enantiomerically pure acetate 4, for which a practical synthetic route was developed. Conjugate addition of 5 then proceeded away from the acetoxy group to give, after intramolecular alkylation, the cyclopropane 6. Selective protection of the derived triol 7 led to a monopivalate that was oxidized to the keto aldehyde 8. Condensation with the oxindole 9 followed by silylation then completed the assembly of 1. The trisubstituted alkene of 1 was established as a single geometric isomer. It followed that in the product 2, the oxindole and the bridging ether had the appropriate relative stereochemical arrangement. The product silyl enol ether was deprotected with fluoride to liberate the ketone 2. With 2 in hand, the next challenge was the kinetic installation of the less stable secondary aminated stereogenic center. To this end, the aldehyde 10 was exposed to TMS-CN and DBU. Under the reaction conditions, the alkene of the intermediate β,γ-unsaturated silylated cyanohydrin was brought into conjugation. Kinetic quench with allyl alcohol gave 11 with a 4:1 preference for the desired endo diastereomer 11. Inversion of the carboxyl then led to the protected amine 12. The ketone 12 was formylated under modified Vilsmeier-Haack conditions, first with Bredereck’s reagent 13 and then with oxalyl chloride, leading to the chloro aldehyde 14. The chlorine was removed by selective Pd-catalyzed reduction, and the product aldehyde was exposed to ethyl magnesium bromide followed by IBX to give the ethyl ketone 15. Epoxidation of the α,β-unsaturated ketone proceeded across the expected exo face leading to 16. The deprotected amine then opened the epoxide to establish the aminated quaternary center and complete the synthesis of gelsemoxonine 3.


Author(s):  
Douglass F. Taber

The nodulisporic acids, isolated from the endophytic fungus Nodulisporium sp., show promising insecticidal activity. Amos B. Smith III of the University of Pennsylvania envisioned (J. Am. Chem. Soc. 2015, 137, 7095) the construction of the central indole of nodulisporic acid D 4 by the convergent coupling of the chloroaniline 1 with the enol triflate 2. The preparation of 2 began (Org. Process Res. Dev. 2007, 11, 19) with the mono­ketal 5 of the Wieland–Miescher ketone, available in enantiomerically-pure form by organocatalyzed Robinson annulation. Condensation with thiophenol and formal­dehyde gave 6, which, under dissolving metal conditions, was reduced to an enolate that was trapped as the silyl enol ether 7. Condensation again with formaldehyde gave 8, that was converted by reduction and protecting group exchange to the ketone 9. Pd-catalyzed formylation of the derived enol triflate led to 10. The Cu-meditated conjugate addition of vinyl magnesium bromide to the unsatu­rated aldehyde 10 was carefully optimized to maximize equatorial addition, away from the angular methyl group. Subsequent C-methylation of the aldehyde was achieved by generating the Li enolate and carrying out the alkylation in diglyme. With 11 in hand, the third carbocyclic ring was assembled by 1,2-addition of vinylmagnesium bromide to the aldehyde followed by ring-closing metathesis and oxidation to give 12. Hydrogenation followed by functional group interconversion then completed the assembly of the enol triflate 2. The stereogenic center of 1 was established by Enders alkylation of 13 with the iodide 14. The ketone 15 was best liberated by ozonolysis under non-epimerizing conditions. The critical Barluenga indole construction that formed 3 also required careful optimization in a model study, the key observation being the value of the Buchwald ligand RuPhos. The conditions developed were found, remarkably, to be compatible with the aldehyde functional group, so subsequent Horner–Wadsworth–Emmons condensation with 16 could be carried out directly, to complete the synthe­sis of (−)-nodulisporic acid D 4.


Author(s):  
Douglass Taber

The Zoanthus alkaloids, exemplified by (-)-norzoanthamine 3a and zoanthamine 3b, show promising activity against osteoporosis. Susumu Kobayashi of the Tokyo University of Science assembled (Angew. Chem. Int. Ed. 2009, 48, 1400; Angew. Chem. Int. Ed. 2009, 48, 1404) the challenging tricyclic core of 3a employing the intramolecular Diels-Alder cyclization of 1 to 2. The cyclopentane of 1 served as useful scaffolding, even though it was cleaved en route to 3a. The cyclohexane ring of 1 has five of its six positions substituted, including three that are alkylated quaternary centers. The starting point for the preparation of 1 was the enantiomerically-pure Hajos-Parrish ketone 4, containing the first of the those quaternary centers. Conjugate addition of MeLi established the second quaternary center. The less stable endo alkyl branch of 1 was installed by conjugate addition to the more reactive α-methylene ketone of the cross-conjugated 5, followed by kinetic quench. Addition of vinyl cuprate across the open face of the enone 7 then established the final quaternary center, setting the stage for the intramolecular Diels-Alder reaction. The silyl enol ether from the cyclization of 1 was not stable, so it was directly oxidized to the enone 2. The keto phosphonate 16 for the last two rings of 3a was prepared from the previously-reported crystalline glutamic acid-derived mesylate 12. Reduction and homologation delivered the ester 14, that was condensed with the phosphonate anion 15 to give 16. The congested cyclopentanone 17, derived from 2, was most efficiently deprotonated with n -BuLi. Exposure of the resulting silyl enol ether to ozone led to the α-hydroxylated product 18. Unexpectedly but happily, oxidative cleavage of 18 delivered, after deprotection and reprotection, the more congested aldehyde 19. This cleavage may be proceeding by tautomerization of 18 to the regioisomeric keto alcohol. The aldehyde 19 was condensed with the keto phosphonate 16, to give, after hydrogenation, the keto lactone 20. A series of oxidation state adjustments then completed the synthesis of (-)-norzoanthamine 3a.


Author(s):  
Douglass Taber

Carsten Bolm of RWTH Aachen developed (Angew. Chem. Int. Ed. 2008, 47, 8920) an Ir catalyst that effected hydrogenation of trisubstituted enones such as 1 with high ee. Benjamin List of the Max-Planck-Institut Mülheim devised (J. Am. Chem. Soc. 2008, 130, 13862) an organocatalyst for the enantioselective reduction of nitro acrylates such as 3 with the Hantzsch ester 4. Gregory C. Fu of MIT optimized (J. Am. Chem. Soc. 2008, 130, 12645) a Ni catalyst for the enantioselective arylation of propargylic halides such as 6. Both enantiomers of 6 were converted to the single enantiomer of 8. Michael C. Willis of the University of Oxford established (J. Am. Chem. Soc. 2008, 130, 17232) that hydroacylation with a Rh catalyst was selective for one enantiomer of the allene 9, delivering 11 in high ee. Similarly, José Luis García Ruano of the Universidad Autónoma de Madrid found (Angew. Chem. Int. Ed. 2008, 47, 6836) that one enantiomer of racemic 13 reacted selectively with the enantiomerically- pure anion 12, to give 14 in high diastereomeric excess. Ei-chi Negishi of Purdue University described (Organic Lett. 2008, 10, 4311) the Zr-catalyzed asymmetric carboalumination (ZACA reaction) of the alkene 15 to give the useful chiron 16. David W. C. MacMillan of Princeton University developed (Science 2008, 322, 77) an intriguing visible light-powered oxidation-reduction approach to enantioselective aldehyde alkylation. The catalytic chiral secondary amine adds to the aldehyde to form an enamine, that then couples with the radical produced by reduction of the haloester. Two other alkylations were based on readily-available chiral auxiliaries. Philippe Karoyan of the Université Pierre et Marie Curie observed (Tetrahedron Lett . 2008, 49, 4704) that the acylated Oppolzer camphor sultam 20 condensed with the Mannich reagent 21 to give 22 as a single diastereomer. Andrew G. Myers of Harvard University developed the pseudoephedrine chiral auxiliary of 23 to direct the construction of ternary alkylated centers. He has now established (J. Am. Chem. Soc. 2008, 130, 13231) that further alkylation gave 24, having a quaternary alkylated center, in high diastereomeric excess.


Sign in / Sign up

Export Citation Format

Share Document