Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2023 Apr 22.
Published in final edited form as: J Am Chem Soc. 2021 Dec 27;144(1):118–122. doi: 10.1021/jacs.1c12283

A Concise Enantioselective Approach to Quassinoids

William P Thomas 1, Sergey V Pronin 2
PMCID: PMC10122274  NIHMSID: NIHMS1890695  PMID: 34958202

Abstract

A synthetic approach to quassinoids is described. The route to the tetracyclic core relies on an efficient and selective annulation between two unsaturated carbonyl components that is initiated by catalytic hydrogen atom transfer from an iron hydride to an alkene. Application of this strategy allows for enantioselective synthesis of quassin, which is prepared in 14 steps from commercially available starting material.


Quassinoids comprise a large group of terpenoid natural products isolated from plants of the Simaroubaceae family, which have been used in folk medicine to treat a variety of diseases.1 These secondary metabolites exhibit a diverse set of biological activities that became a subject of frequent scientific inquiry in the latter half of the 20th century. Notable examples include potent cytotoxicity toward different cancer cell lines and antimalarial activity against drug-resistant strains of P. falciparum.2,3 These effects have been commonly attributed to inhibition of protein synthesis in both mammalian and protozoan cells.4,5 However, recent studies suggest a far more complex mechanism of action that accounts for the observed antineoplastic properties, including down-regulation of c-MYC oncoproteins and inhibition of NF-κB.6,7 These findings spurred several subsequent investigations to further evaluate the therapeutic potential of quassinoids, driving a renewed interest in these natural products.8,9

Biosynthetic assembly of the shared tricarbocyclic motif found in the structure of quassinoids involves oxidative degradation of a triterpenoid precursor.10 The resulting common core can be found in a variety of functionalization patterns in different congeners and often contains a fused δ-valerolactone moiety (e.g., 1, 2, and 3, Figure 1).11 The combination of complex structural features and promising biological activities has prompted numerous investigations into the chemical synthesis of quassinoids over the past five decades.1217 These studies accumulated a wealth of knowledge in reactivity of complex polycyclic motifs, and the developed multistep sequences underscored the synthetic challenge associated with the target compounds, which continues to drive innovative research in the field of organic chemistry.18 Here, we demonstrate a new approach to quassinoids that provides rapid asymmetric access to the functionalized polycyclic core of these natural products. Application of our strategy has allowed for the synthesis of (+)-quassin (1) in 14 steps from commercially available material.

Figure 1.

Figure 1.

Representative quassinoids and our approach to the common polycyclic motif.

In devising our route, we aimed to develop a concise approach to the tricarbocyclic motif of quassinoids. At the same time, we recognized that redox manipulations following assembly of the carbon framework had historically accounted for the bulk of transformations in prior campaigns, therefore judicious selection of the functionalization patterns would be requisite for a succinct synthesis. With these considerations in mind, we envisioned that annulation between two unsaturated carbonyl components would allow for convergent union of cyclic fragments found in the perhydrophenanthrene motif (Figure 1).19 The desired transformation was expected to take advantage of hydrogen atom transfer (HAT) from an iron hydride to the 1,1-disubstituted alkene followed by Giese addition of the resulting alkyl radical to the unsaturated ketone and ultimate cyclization onto the pendant aldehyde.20,21 The functionalities formed at C7 and C14 would serve as handles for assembly of the fused lactone moiety, which is found in the majority of congeners. Although direct installation of the desired trans fusion at C8 and C9 during the annulation could be difficult to achieve, previous observations of facile proton exchange at C9 provided an outlet for possible solutions to the anticipated stereochemical challenge.22 Notably, the HAT-initiated annulation was expected to offer good tolerance of functional groups in the reaction partners, promising a considerable degree of flexibility during optimization of this crucial transformation.

After extensive experimentation with various annulation partners, we identified an optimal combination of unsaturated carbonyl components for construction of the desired tricarbocyclic motif. Synthesis of aldehyde 4 began with alkylation of a lithium enolate of unsaturated ketone 5, which could be prepared by IBX-mediated oxidation of (+)-3-methylcyclohexanone (Scheme 1).2325 Direct treatment of the reaction mixture with hydrogen peroxide and DBU delivered epoxide 6. Wittig olefination of the ketone produced exocyclic alkene 7, which underwent reduction of the nitrile and allylic epoxide upon exposure to DIBAL-H. Protection of the resulting secondary alcohol secured access to the desired annulation partner 4 in five steps from commercially available material.

Scheme 1.

Scheme 1.

Synthesis of γ,δ-Unsaturated Aldehyde 4

Synthesis of unsaturated ketone 8 relied on regio- and diastereoselective epoxidation of tricyclic diene 9, which was prepared from 2,6-dimethylbenzoquinone in an enantioselective manner by taking advantage of oxazaborolidine catalysis (Scheme 2).26 Thermal retro-Diels–Alder reaction accomplished efficient extrusion of cyclopentadiene and delivered annulation partner 8 in three steps from commercially available material.27 Notably, this epoxyquinone derivative demonstrated superior performance in the HAT-initiated annulation reactions en route to the shared polycyclic motif of quassinoids when compared to diene 9, which was previously employed in our synthesis of forskolin.19

Scheme 2.

Scheme 2.

Synthesis of Epoxyquinone 8

Annulation of epoxyquinone 8 and aldehyde 4 resulted in highly selective formation of polycyclic intermediate 10 (Scheme 3).28,29 To our surprise, the epoxide fragment exerted a substantial degree of stereocontrol during the construction of the new bond between C9 and C10.

Scheme 3.

Scheme 3.

Synthesis of (+)-Quassin (1)

This stereochemical relay is noteworthy and may prove valuable in the synthesis of other terpenoid scaffolds. In the annulation toward quassin, the resulting secondary alcohol was selectively produced with an undesired configuration at C7, but the observed conformational preferences of product 10 suggested a potential solution. Crystallographic studies with a TBS analog in the solid state revealed that the cyclohexanol motif adopts a twist boat conformation, which was also suggested for the solutions in deuterated chloroform by 1H NMR analysis. We therefore speculated that the pseudoaxial orientation of the secondary hydroxy group could drive epimerization of the stereocenter at C7 upon a reversible retro-aldol reaction. Indeed, treatment of intermediate 10 with a strong base followed by addition of diethylphosphonoacetic acid and a coupling agent produced ester 11 with corrected configuration at C7. We empirically found that the application of toluene as a solvent in this transformation allowed us to selectively epimerize C7 without affecting the configuration of the stereocenter at C9, which was inverted under a variety of other conditions. This is noteworthy because epimerization of C9 was invariably accompanied by equilibration to the undesired configuration at C7 and therefore was necessary to avoid at this point in the sequence.

Initial attempts to affect the Horner–Wadsworth–Emmons reaction of phosphonate 11 were unsuccessful, likely due to unfavorable steric interactions associated with approach of the ester fragment from the concave face of the polycyclic system. We speculated that inversion of the C9 stereocenter would be beneficial at this stage, as it would likely lead to a less sterically encumbered transition state in the requisite addition to the ketone. We ultimately found that treatment with cesium fluoride in dimethyl sulfoxide induced epimerization of C9 and intramolecular olefination, delivering the desired tetracyclic core of the quassinoids. The resulting product mixtures contained varying amounts of deprotected secondary alcohol at C2, which we converged to the corresponding cyclohexanone 12 by oxidation with IBX in the presence of tosic acid hydrate.30

With access to the complete polycyclic scaffold of quassinoids, we proceeded to explore its application in the synthesis of a representative congener, quassin (1). Conversion of the unsaturated epoxide to the corresponding ketone was accomplished upon isomerization of intermediate 12 with catalytic amounts of a palladium phosphine complex.31,32 Direct addition of palladium on carbon and hydrogen to the reaction mixture allowed for chemoselective hydrogenation of the alkene to deliver lactone 13. Site-selective oxidation of related intermediates containing the saturated lactone moiety presented a significant challenge in previous studies and required protecting group manipulations to mitigate competing reactivity patterns.33 After significant experimentation with precursors and conditions, we found that the desired functionalization of our system could be accomplished by allylic hydroxylation of silyl enol ether 14, which in turn was selectively prepared by soft enolization of cyclohexanone 13.34 Subsequent oxidation of alcohol 15 delivered penultimate intermediate 16 containing the requisite polycarbonyl motif of the target natural product.35 Ultimately, treatment of silyl enol ether 16 with a source of fluoride ions in the presence of methyl iodide resulted in direct production of (+)-quassin, completing the synthesis in 14 steps from a commercially available starting material.

In summary, we disclose a new synthetic approach to quassinoids that relies on a HAT-initiated annulation of two unsaturated carbonyl components for rapid access to the shared polycyclic motif. The judicious choice of annulation partners ensures efficient formation of the key structural bonds and provides pragmatic functional handles for subsequent elaborations. Application of this strategy to the synthesis of (+)-quassin allows for asymmetric assembly of the target natural product in 14 steps for commercially available material, which significantly reduces unproductive functional group manipulations and is notably shorter than the previous efforts in the field.36 We expect our annulation-based approach to quassinoids to find application in the synthesis of other congeners, and corresponding studies are currently underway.

Supplementary Material

Supporting Information

ACKNOWLEDGMENTS

Financial support from the National Institutes of Health (R01GM121678), the University of California, Irvine, and Amgen is gratefully acknowledged. We thank Dr. Joseph Ziller for X-ray crystallographic analysis and Dr. Felix Grun for mass spectrometric analysis. We also thank Professors Larry Overman, Chris Vanderwal, and Scott Rychnovsky for providing routine access to their instrumentation and helpful discussions.

Footnotes

The authors declare no competing financial interest.

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.1c12283.

Experimental procedures, characterization data for all new compounds, and CIF file for the TBS analog of intermediate 10 (PDF)

Accession Codes

CCDC 2123424 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

Complete contact information is available at: https://pubs.acs.org/10.1021/jacs.1c12283

Contributor Information

William P. Thomas, Department of Chemistry, University of California, Irvine, California 92697-2025, United States

Sergey V. Pronin, Department of Chemistry, University of California, Irvine, California 92697-2025, United States

REFERENCES

  • (1).For relevant reviews, see:; (a) Guo Z; Vangapandu S; Sindelar RW; Walker LA; Sindelar RD Biologically active quassinoids and their chemistry: potential leads for drug design. Curr. Med. Chem 2005, 12, 173. [DOI] [PubMed] [Google Scholar]; (b) Fiaschetti G; Grotzer MA; Shalaby T; Castelletti D; Arcaro A Quassinoids: from traditional drugs to new cancer therapeutics. Curr. Med. Chem 2011, 18, 316. [DOI] [PubMed] [Google Scholar]; (c) Houël E; Stien D; Bourdy G; Deharo E Quassinoids: anticancer and antimalarial activities. In Natural Products; Springer, 2013; pp 3775–3802. [Google Scholar]
  • (2).For notable examples, see:; (a) Kupchan SM; Britton RW; Lacadie JA; Ziegler MF; Sigel CW The isolation and structural elucidation of bruceantin and bruceantinol, new potent antileukemic quassinoids from Brucea antidysenterica. J. Org. Chem 1975, 40, 648. [DOI] [PubMed] [Google Scholar]; (b) Wiseman CL; Yap HY; Bedikian AY; Bodey GP; Blumenschein GR Phase II trial of bruceantin in metastatic breast carcinoma. Am. J. Clin. Oncol 1982, 5, 389. [DOI] [PubMed] [Google Scholar]; (c) Arseneau J; Wolter J; Kuperminc M; Ruckdeschel J A phase II study of bruceantin (NSC-165, 563) in advanced malignant melanoma. Invest. New Drugs 1983, 1, 239. [DOI] [PubMed] [Google Scholar]
  • (3).For examples, see:; (a) O’Neill MJ; Bray DH; Boardman P; Phillipson JD; Warhurst DC; Peters W; Suffness M Plants as sources of antimalarial drugs: in vitro antimalarial activities of some quassinoids. Antimicrob. Agents Chemother 1986, 30, 101. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Ang HH; Chan KL; Mak JW In vitro antimalarial activity of quassinoids from Eurycoma longifolia against Malaysian chloroquine-resistant Plasmodium falciparum isolates. Planta Med. 1995, 61, 177. [DOI] [PubMed] [Google Scholar]; (c) Ekong RM; Kirby GC; Patel G; Phillipson JD; Warhurst DC Comparison of the in vitro activities of quassinoids with activity against Plasmodium falciparum, anisomycin and some other inhibitors of eukaryotic protein synthesis. Biochem. Pharmacol 1990, 40, 297. [DOI] [PubMed] [Google Scholar]
  • (4).(a) Liao LL; Kupchan SM; Horwitz SB Mode of action of the antitumor compound bruceantin, and inhibitor of protein synthesis. Mol. Pharmacol 1976, 12, 167. [PubMed] [Google Scholar]; (b) Fresno M; Gonzales A; Vazquez D; Jimenez A Bruceantin, a novel inhibitor of peptide bond formation. Biochim. Biophys. Acta 1978, 518, 104. [DOI] [PubMed] [Google Scholar]
  • (5).Kirby GC; O’Neill MJ; Phillipson JD; Warhurst DC In vitro studies on the mode of action of quassinoids with activity against chloroquine-resistant Plasmodium falciparum. Biochem. Pharmacol 1989, 38, 4367. [DOI] [PubMed] [Google Scholar]
  • (6).Mata-Greenwood E; Cuendet M; Sher D; Gustin D; Stock W; Pezzuto JM Brusatol-mediated induction of leukemic cell differentiation and G1 arrest is associated with down-regulation of c-myc. Leukemia 2002, 16, 2275. [DOI] [PubMed] [Google Scholar]
  • (7).Cuendet M; Gills JJ; Pezzuto JM Brusatol-induced HL-60 cell differentiation involves NF-, B activation. Cancer Lett. 2004, 206, 43. [DOI] [PubMed] [Google Scholar]
  • (8).For examples, see:; (a) Cuendet M; Christov K; Lantvit DD; Deng Y; Hedayat S; Helson L; McChesney JD; Pezzuto JM Multiple myeloma regression mediated by bruceantin. Clin. Cancer Res 2004, 10, 1170. [DOI] [PubMed] [Google Scholar]; (b) Kim J-A; Lau EK; Pan L; Carcache De Blanco EJ NF-cB inhibitors from Brucea javanica exhibiting intracellular effects on reactive oxygen species. Anticancer Res. 2010, 30, 3295. [PMC free article] [PubMed] [Google Scholar]; (c) Issa ME; Berndt S; Carpentier G; Pezzuto JM; Cuendet M Bruceantin inhibits multiple myeloma cancer stem cell proliferation. Cancer Biol. Ther 2016, 17, 966. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (9).For relevant discussion, see:; (a) Cuendet M; Pezzuto JM Antitumor activity of bruceantin: an old drug with new promise. J. Nat. Prod 2004, 67, 269. [DOI] [PubMed] [Google Scholar]; (b) Z.-K. Duan; Z.-J. Zhang; S.-H. Dong; Y.-X. Wang; S.-J. Song; X.-X. Huang Quassinoids: phytochemistry and antitumor properties. Phytochemistry 2021, 187, 112769. [DOI] [PubMed] [Google Scholar]; For a recent review, see:.
  • (10).Polonsky J Quassinoid bitter principles. Fortschr Chem. Org. Naturst 1973, 30, 101. [DOI] [PubMed] [Google Scholar]
  • (11).For examples, see:; (a) Valenta Z; Gray AH; Orr DE; Papadopoulos S; Podešva C The stereochemistry of quassin. Tetrahedron 1962, 18, 1433. [Google Scholar]; (b) Polonsky J; Bourguignon-Zylber N Etude des constituants amers des fruits de Hannoa klaineana (Simarubacée): chaparrione et klainéanone. Bull. Soc. Chim. Fr 1965, 10, 2793. [PubMed] [Google Scholar]; (c) Stöcklin W; Stefanović M; Geissman TA; Casinovi CG Amarolide, isolation from Castela nicholsoni hook and revision of its structure. Tetrahedron Lett 1970, 11, 2399. [Google Scholar]
  • (12).(a) Vidari G; Ferriño S; Grieco PA Quassinoids: total synthesis of d-quassin. J. Am. Chem. Soc 1984, 106, 3539. [Google Scholar]; (b) Grieco PA; Lis R; Ferrino S; Law JY Quassinoids: total synthesis of dl-castelanolide. J. Org. Chem 1984, 49, 2342. [Google Scholar]; (c) Grieco PA; Nargund RP; Parker DT Total synthesis of the highly oxygenated quassinoid (±)-klaineanone. J. Am. Chem. Soc 1989, 111, 6287. [Google Scholar]; (d) Collins JL; Grieco PA; Gross RS Synthetic studies on quassinoids: total synthesis of (±)-shinjulactone C. J. Org. Chem 1990, 55, 5816. [Google Scholar]; (e) Gross RS; Grieco PA; Collins JL Synthetic studies on quassinoids: total synthesis of (±)-chaparrinone. J. Am. Chem. Soc 1990, 112, 9436. [Google Scholar]; (f) Gross RS; Grieco PA; Collins JL Synthesis of the highly oxygenated quassinoid shinjulactone D. J. Org. Chem 1991, 56, 7167. [Google Scholar]; (g) Fleck TJ; Grieco PA Synthetic studies on quassinoids: total synthesis of (±)-glaucarubolone and (±)-holacanthone. Tetrahedron Lett 1992, 33, 1813. [Google Scholar]; (h) Moher ED; Collins JL; Grieco PA Synthetic studies on quassinoids: total synthesis of simalikalactone D and assignment of the absolute configuration of the u-methylbutyrate ester side chain. J. Am. Chem. Soc 1992, 114, 2764. [Google Scholar]; (i) Moher ED; Grieco PA; Collins JL (R)-(+)- and (S)-(−)-5-Ethyl-5-methyl-1,3-dioxolane-2,4-dione reagents for the direct preparation of α-hydroxy- α-methylbutyrate esters: assignment of the absolute configuration of the α-acetoxy- α-methylbutyrate ester side chain of quassimarin via total synthesis. J. Org. Chem 1993, 58, 3789. [Google Scholar]; (j) VanderRoest JM; Grieco PA Total synthesis of (±)-bruceantin. J. Am. Chem. Soc 1993, 115, 5841. [Google Scholar]; (k) Grieco PA; Collins JL; Moher ED; Fleck TJ; Gross RS Synthetic studies on quassinoids: total synthesis of (−)-chaparrinone, (−)-glaucarubolide, and (+)-glaucarubinone. J. Am. Chem. Soc 1993, 115, 6078. [Google Scholar]; (l) Grieco PA; Piñeiro-Nuñez MM C19 quassinoids: total synthesis of dl-samaderin B. J. Am. Chem. Soc 1994, 116, 7606. [Google Scholar]; (m) Grieco PA; Cowen SD; Mohammadi F Synthetic studies on highly oxygenated quassinoids: total synthesis of (±)-141,15,-dihydroxyklaineanone. Tetrahedron Lett 1996, 37, 2699. [Google Scholar]; (n) Grieco PA; Collins JL; Huffman JC Synthetic studies on quassinoids: synthesis of (±)-shinjudilactone and (±)-13-epi-shinjudilactone. J. Org. Chem 1998, 63, 9576. [DOI] [PubMed] [Google Scholar]
  • (13).Stojanac N; Valenta Z Total synthesis of d,l-quassin. Can. J. Chem 1991, 69, 853. [Google Scholar]
  • (14).Hirota H; Yokoyama A; Miyaji K; Nakamura T; Takahashi T Synthetic studies on quassinoids: total synthesis of (±)-amarolide. Tetrahedron Lett 1987, 28, 435. [Google Scholar]
  • (15).Sasaki M; Murae T A formal synthesis of bruceantin. Tetrahedron Lett. 1989, 30, 355. [Google Scholar]
  • (16).Kim M; Kawada K; Gross RS; Watt DS An enantioselective synthesis of (+)-picrasin B, (+)-)2-picrasin B, and (+)-quassin from the R-(−) enantiomer of the Wieland-Miescher ketone. J. Org. Chem 1990, 55, 504. [Google Scholar]
  • (17).(a) Shing TKM; Jiang Q; Mak TCW Total synthesis of (+)-quassin from (+)-carvone. J. Org. Chem 1998, 63, 2056. [Google Scholar]; (b) Shing TKM; Jiang Q Total synthesis of (+)-quassin. J. Org. Chem 2000, 65, 7059. [DOI] [PubMed] [Google Scholar]; (c) Shing TKM; Yeung YY Total synthesis of (−)-samaderine Y from (S)-(+)-carvone. Angew. Chem., Int. Ed 2005, 44, 7981. [DOI] [PubMed] [Google Scholar]; (d) Shing TKM; Yeung YY Synthetic studies towards pentacyclic quassinoids: total synthesis of unnatural (−)-14-epi-samaderin E and natural (−)-samaderine Y from (S)-(+)-carvone. Chem.—Eur. J 2006, 12, 8367. [DOI] [PubMed] [Google Scholar]
  • (18).For innovative approaches to the quassinoids in the 21st century, see:; (a) Barrero AF; Alvarez-Manzaneda EJ; Alvarez-Manzaneda R; Chahboun R; Meneses R; Cuerva JM; Aparicio M; Romera JL Approach to the synthesis of antitumor quassinoids from labdane diterpenes: an efficient synthesis of a picrasane-related intermediate. Org. Lett 2001, 3, 647. [DOI] [PubMed] [Google Scholar]; (b) Spino C; Hill B; Dubé P; Gingras S A diene-transmissive approach to the quassinoid skeleton. Can. J. Chem 2003, 81, 81. [Google Scholar]; (c) Perreault S; Spino C Meldrum’s acid-derived thione dienophile in a convergent and stereoselective synthesis of a tetracyclic quassinoid intermediate. Org. Lett 2006, 8, 4385. [DOI] [PubMed] [Google Scholar]; (d) Marcos IS; García N; Sexmero MJ; Hernandez FA; Escola MA; Basabe P; Díez D; Urones JG Synthetic studies towards picrasane quassinoids. Tetrahedron 2007, 63, 2335. [Google Scholar]; (e) Caron P-Y; Deslongchamps P Versatile strategy to access tricycles related to quassinoids and triterpenes. Org. Lett 2010, 12, 508. [DOI] [PubMed] [Google Scholar]; (f) Burns DJ; Mommer S; O’Brien P; Taylor RJK; Whitwood AC; Hachisu S Stereocontrolled synthesis of the AB rings of samaderine C. Org. Lett 2013, 15, 394. [DOI] [PubMed] [Google Scholar]; (g) Ravindar K; Caron P-Y; Deslongchamps P Anionic polycyclization entry to tricycles related to quassinoids and terpenoids: a stereocontrolled total synthesis of (+)-cassaine. J. Org. Chem 2014, 79, 7979. [DOI] [PubMed] [Google Scholar]; (h) Usui K; Suzuki T; Nakada M A highly stereoselective intramolecular Diels–Alder reaction forconstruction of the AB ring moiety of bruceantin. Tetrahedron Lett. 2015, 56, 1247. [Google Scholar]; (i) Oki Y; Nakada M Research on Au(I)-catalyzed ene-yne cycloisomerization for construction of quassinoid scaffold. Tetrahedron Lett 2018, 59, 926. [Google Scholar]; (j) Condakes ML; Rosen RZ; Harwood SJ; Maimone TJ A copper-catalyzed double coupling enables a 3-step synthesis of the quassinoid core architecture. Chem. Sci 2019, 10, 768. [Google Scholar]
  • (19).Thomas WP; Schatz DJ; George DT; Pronin SV A radical-polar crossover annulation to access terpenoid motifs. J. Am. Chem. Soc 2019, 141, 12246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (20).For an excellent review of relevant HAT-initiated reactions, see:; (e) Crossley SWM; Martinez RM; Obradors C; Shenvi RA Mn, Fe, and Co-catalyzed radical hydrofunctionalizations of olefins. Chem. Rev 2016, 116, 8912. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (21).For pioneering examples of HAT-initiated reductive coupling of alkenes and mechanistic studies, see:; (a) Lo JC; Yabe Y; Baran PS A practical and catalytic reductive olefin coupling. J. Am. Chem. Soc 2014, 136, 1304. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Lo JC; Gui J; Yabe Y; Pan C-M; Baran PS Functionalized olefin cross-coupling to construct carbon–carbon bonds. Nature 2014, 516, 343. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Lo JC; Kim D; Pan C-M; Edwards JT; Yabe Y; Gui J; Qin T; Gutiérrez S; Giacoboni J; Smith MW; Holland PL; Baran PS Fe-catalyzed C–C bond construction from olefins and radicals. J. Am. Chem. Soc 2017, 139, 2484. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Kim D; Rahaman SMW; Mercado BQ; Poli R; Holland P Roles of iron complexes in catalytic radical alkene cross-coupling: a computational and mechanistic study. J. Am. Chem. Soc 2019, 141, 7473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (22).(a) Valenta Z; Papadopoulos S; Podešva C Quassin and neoquassin. Tetrahedron 1961, 15, 100. [Google Scholar]; (b) Stojanac N; Sood A; Stojanac Ž; Valenta Z A synthetic approach to quassin. Introduction of functionality and stereochemistry by a Diels-Alder reaction. Can. J. Chem 1975, 53, 619. [Google Scholar]
  • (23).Nicolaou KC; Zhong Y-L; Baran PS A new method for the one-step synthesis of α,β,-unsaturated carbonyl systems from saturated alcohols and carbonyl compounds. J. Am. Chem. Soc 2000, 122, 7596. [Google Scholar]
  • (24).Ketone 5 can also be prepared in two steps from (−)-isopulegol. See the Supporting Information for details.
  • (25).The desired product 6 is formed with high diastereoselectivity, but the efficiency appears to suffer from unidentified side processes.
  • (26).Liu D; Canales E; Corey EJ Chiral oxazaborolidine–aluminum bromide complexes are unusually powerful and effective catalysts for enantioselective Diels-Alder reactions. J. Am. Chem. Soc 2007, 129, 1498. [DOI] [PubMed] [Google Scholar]
  • (27).Alder K; Flock FH; Beumling H Darstellung von p-Chinon-epoxyden. Chem. Ber 1960, 93, 1896. [Google Scholar]
  • (28).The annulation could be performed with 1.5 equiv of aldehyde 4 without a significant loss in yield (54%). Further reduction in the relative amount of the γ,δ,-unsaturated carbonyl component led to a precipitous loss in efficiency.
  • (29).For the discovery of isopropoxy(phenyl)silane as an exceptional reductant in the HAT-initiated reactions, see:; Obradors C; Martinez RM; Shenvi RA Ph(i-PrO)SiH2: an exceptional reductant for metal-catalyzed hydrogen atom transfers. J. Am. Chem. Soc 2016, 138, 4962. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (30).Tosic acid hydrate likely affects deprotection of the silyl ether, but it is also known to accelerate IBX-mediated oxidations. See ref 23 for relevant discussion.
  • (31).Suzuki M; Oda Y; Noyori R Palladium(0) catalyzed reaction of 1,3-diene epoxides. A useful method for the site-specific oxygenation of 1,3-dienes. J. Am. Chem. Soc 1979, 101, 1623. [Google Scholar]
  • (32).Interestingly, attempts to hydrogenate the unsaturated epoxide motif of 12 to the corresponding saturated alcohol analogous to a transformation reported by the Shing group (see ref 17) on a related intermediate only resulted in hydrogenation of the alkene and no reduction of the epoxide.
  • (33).For relevant discussion, see refs 12a, 16, and 17b.
  • (34).Magnus P; Mugrage B New trialkylsilyl enol ether chemistry. Regiospecific and stereospecific sequential electrophilic addition. J. Am. Chem. Soc 1990, 112, 462. [Google Scholar]
  • (35).Omura K; Sharma AK; Swern D Dimethyl sulfoxide-trifluoroacetic anhydride. New reagent for oxidation of alcohols to carbonyls. J. Org. Chem 1976, 41, 957. [Google Scholar]
  • (36).According to our analysis of the existing reports, quassin was previously prepared in 23 steps (from Wieland-Miescher ketone12a),33 steps (from Wieland-Miescher ketone16), 31 steps (from 2-methylpenta-1,3-diene13), and 28 steps (from carvone17a).

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information

RESOURCES