Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 Aug 1.
Published in final edited form as: Tetrahedron Lett. 2017 Jul 25;58(35):3478–3481. doi: 10.1016/j.tetlet.2017.07.080

Synthesis of cananodine by intramolecular epoxide opening

Patrick Shelton 1, Toby J Ligon 1, Jennifer M Dell (née Meyer) 1, Loagan Yarbrough 1, James R Vyvyan 1,*
PMCID: PMC5722248  NIHMSID: NIHMS897154  PMID: 29230072

Abstract

Cananodine is a guaipyridine alkaloid with activity against liver cancer. Cananodine was synthesized using a remarkable intramolecular opening of a trisubstituted epoxide as the key step in construction of the seven-membered carbocycle of the target. The epoxide opening strategy allows all four stereoisomers of cananodine to be prepared.

Graphical abstract

graphic file with name nihms897154u1.jpg


Hepatocellular carcinoma (HCC) is one of the most common cancers worldwide, with an estimated 750,000 new cases each year leading to nearly as many deaths.1 Although HCC is particularly prevalent in Asia, the incidence of HCC in the U.S. has increased dramatically over the last two decades, to an estimated 27,000 new cases per year,2 probably as a result of the increase in chronic hepatitis C virus (HCV) infections in this country, which has resulted in the age-related incidence of HCC shifting to younger people.1,3 The prognosis for patients diagnosed with HCC is grave. The primary treatment and only proven curative therapy is liver resection or transplant, and although the 5-year survival rate for these surgical patients is 25–50%, there is a high rate of recurrence.4,5 Worse, only 10–15% of HCC patients are surgical candidates due to decreased liver function or metastasis of the tumor.5,6 A variety of anticancer drugs have been tried as systemic HCC treatments, but exceedingly few have proven effective. Currently the only approved drug for HCC is sorafenib (Nexavar®), a multi-kinase inhibitor containing a pyridine head group and diaryl urea tail that was introduced in 2007.7 Sorafenib improved overall survival time in patients with advanced HCC from 7.9 to 10.7 months.7a Clearly there is room for progress.8

graphic file with name nihms897154u2.jpg

Cananodine (1) is a guaipyridine alkaloid isolated in small quantities from the fruits of Cananga odorata (10 mg from 3.5 kg fruit) (Figure 1).9 C. odorata is a member of the Annonaceae family and commonly known as “ylang-ylang,” since steam distillation of the flowers provides fragrant ylang-ylang oil. Moreover, C. odorata is known to be rich in alkaloids,10 and has been used in traditional folk medicine in Southeast Asia.11,12 Cananodine was also recently isolated from Cyperus scariosus.13 Most importantly, cananodine (1) has activity against both the Hep G2 and Hep 2,2,15 hepatocarcinoma cell lines, with IC50 values of 0.22 and 3.8 μg/mL, respectively.9 This compares favorably to the in vitro activity of sorafenib against Hep G2 of 2.0 μg/mL.14 Among natural sesquiterpene pyridine alkaloids, the guaipyridine skeleton is rare.15 Recently discovered guaipyridine alkaloids include the rupestines, isolated from Artemsia rupestris (Figure 1).16 No biological activity for the rupestines has been reported yet.

Figure 1.

Figure 1

Guaipyridine alkaloids.

Despite the anticancer activity and unusual structure of cananodine, only one synthetic investigation has been reported to date, a total synthesis reported by Craig and Henry, which raised a question about the optical rotation of the natural product.17 The only other synthetic approaches to guaipyridines in the literature are best described as regio- and/or stereo-random, and all of them began with an intact seven-membered carbocycle.18 Herein we report the synthesis of optically active cananodine, its enantiomer and their diastereomers.

In our retrosynthesis of cananodine (1), we desired a strategy that would allow for the preparation of either enantiomer of the natural product. Thus, we planned to form the seven-membered ring through intramolecular attack of a picolyl anion on an optically active epoxide 2 (Scheme 1).19 The ‘Z’ group in 2 was intended to be electron withdrawing to make the α-protons at C-2 of the pyridine ring more acidic than the protons on the C-6 methyl group, thereby allowing for selective deprotonation to promote the intramolecular reaction. The epoxide 2 would be prepared from pyridyl diene 3, which in turn is formed by Suzuki-Miyaura coupling of pyridyl halide/triflate 4 and alkenyl boronate 5.20

Scheme 1.

Scheme 1

Retrosynthetic analysis of cananodine.

Our first investigation involved use of a sulfone as the aforementioned directing group. Thus, cleavage of the pivaloate group in 620 gave primary alcohol 7 which was converted to sulfone 8 in a two-step process (Scheme 2). Cleavage of the acetonide followed by conversion of the resulting diol to the trisubstituted epoxide via the mesylate yielded 9. Treatment of sulfone epoxide 9 with various bases (t-BuOK, LDA, n-BuLi) produced the orange-red color of the substituted picolyl anion,19 but no cyclized product 10 was obtained. Even with prolonged reaction times, and in some cases elevated temperature, most of the starting epoxide 9 was recovered in each case. Attempts to activate the epoxide after deprotonation (LiBr, SnCl2,21 BF3•OEt2) were unsuccessful in producing cyclized products also, again returning mostly unreacted 9. The highly delocalized anion that is produced upon deprotonation of the activated methylene in 9 is apparently not nucleophilic enough to open the relatively hindered epoxide.

Scheme 2.

Scheme 2

Initial epoxide opening investigation.

To see if reducing steric demands of the epoxide would facilitate cyclization, we next examined a monosubstituted epoxide with a single picolyl position to eliminate the need for a group to direct deprotonation. Chemoselective asymmetric dihydroxylation of the monosubstituted olefin of 1120 produced diol 12 in good yield (Scheme 3). A portion of the diol was converted to the corresponding acetonide for analysis by chiral GC, which showed an enantiomeric excess of 76%. Sharpless’ one pot procedure22 was used to convert the diol to epoxide 13 in a relatively low yield, but provided sufficient material to study the cyclization.

Scheme 3.

Scheme 3

Cyclization of a monosubstituted epoxide.

Baldwin’s rules for ring closure23 do not address medium ring cases. Although the tendency for epoxides to undergo nucleophilic attack at the less substituted position under basic conditions is well-established for acyclic cases and in the formation of small rings, we predicted that the transannular strain in the incipient 8-membered ring would disfavor the 8-endo cyclization and make the 7-exo mode more likely.24 Treatment of 13 with LDA, however, produced only the 8-membered product 14 along with a significant amount of recovered starting material. No 16 was detected in the 1H NMR spectrum of the crude reaction mixture. The oxygenated methine of 14 (δ 3.70 for 1H and δ 72.2 for 13C) was the key to assigning the 8-membered ring structure to the product, along with the splitting pattern of the benzylic protons at 3.06 and 2.95 ppm (each a ddd). The use of Et2AlCl Lewis acid to coordinate to the epoxide and in an attempt to alter the regioselectivity of the cyclization was unsuccessful, instead producing the chloride 15 after conversion of the initially produced chlorohydrin to the acetate ester to aid in purification and characterization. Use of BF3•OEt2 resulted in products arising from rearrangement of the epoxide to the corresponding aldehyde. In no case was the seven-membered product 16 observed by 1H NMR analysis of the crude products.

Despite the fact that our earlier investigation of the intermolecular reaction of a picolyl anion with a trisubstituted epoxide gave relatively low-yields,19 we proceeded to prepare a trisubstituted epoxide cyclization precursor expecting that the intramolecular reaction would be more successful than the intermolecular case. (Scheme 4). Pd-catalyzed cross coupling of pyridyl iodide 17 with dienylboronate 5b20 in the presence of Ag2O25 was found to be the most efficient method to obtain 18 in high yield. Asymmetric dihydroxylation of the more electron rich olefin of 18 gave diol 19 in moderate yield.26 Conversion of the diol to the corresponding epoxide via the intermediate mesylate27 provided epoxide 20, which was determined to have an enantiomeric excess of 96% by chiral GC analysis. Treatment of 20 with butyllithium resulted in deprotonation at the picolyl positions and 7-exo opening of the epoxide to produce 21 in modest yield. Although this reaction is intramolecular, deprotonation of the picolyl positions is non-selective and deprotonation at C6 is unproductive in terms of opening the epoxide, making the maximum yield 50% assuming irreversible deprotonation.

Scheme 4.

Scheme 4

Synthesis of cananodine and its epimer.

The last step in the synthesis was hydrogenation of the exo methylene group of 21. We were hopeful that the tertiary alcohol of 21 would provide a handle for hydroxyl-directed hydrogenation.28 Crabtree’s catalyst29 was ineffective, both at 1 atm or 50 psi of hydrogen. Wilkinson’s catalyst30 accomplished the transformation under elevated hydrogen pressure, but with poor diastereoselectivity, producing cananodine (1) and its epimer in 49% yield as a 1:1 mixture. Either the tertiary hydroxyl group is too hindered to effectively coordinate to the metal, or the dominant conformation of the seven-membered ring does not bring the group sufficiently close to the olefin. Molecular modeling supports the latter. Unfortunately, we were unable to separate the diastereomers with either normal or reverse phase chromatography. Resonances belonging to cananodine (1) were easily assigned in the 1H and 13C NMR spectra of the mixture, however, and matched those previously reported.17

For the sake of completeness, we also prepared the enantiomeric series of intermediates (Scheme 5). Diene 18 was converted to the epoxide 24 via diol 23 in good yield and high enantiomeric excess. Cyclization of epoxide 24 proceeded in a yield consistent with previous results to produce 25. Hydrogenation with Wilkinson’s catalyst as before led to the diastereomeric mixture of ent-cananodine (26) and its epimer 27.

Scheme 5.

Scheme 5

Synthesis of ent-cananodine and its epimer.

In summary, an intramolecular epoxide opening strategy successfully produces the bicyclic guaipyridine core. Both enantiomers of the epoxide substrate are accessible through asymmetric dihydroxylation in the preparation of the epoxide substrate. Following the remarkable epoxide cyclization, hydrogenation of the exo-methylene in the penultimate intermediate was not diastereoselective, however. Cananodine was prepared in six steps from 3-iodo-2,6-dimethylpyridine in 4% overall yield as a mixture with its diastereomer. Cananodine’s enantiomer was prepared in 3% overall yield as a diastereomeric mixture. Thus, all four stereoisomers of the target compound were prepared from a common intermediate. Current efforts are aimed at more efficient construction of the guaipyridine core and enantioselective hydrogenation of the exo-methylene in the late stages of the synthesis.

Supplementary Material

supplement

Highlights.

  • Guaipyridine alkaloid cananodine was synthesized in 6 steps

  • 7-membered carbocycle was constructed using a remarkable intramolecular opening of a trisubstituted epoxide

  • Opening of a similar monosubstituted epoxide produced an 8-membered carbocycle

  • Epoxide opening strategy resulted in the synthesis of all four diastereomers of the target

Acknowledgments

This work was supported by the National Cancer Institute, National Institutes of Health (R15 CA122084) and the Western Washington University Fund for the Enhancement of Graduate Research. We thank Prof. Donald Craig for 1H and 13C NMR spectra of cananodine.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Supporting Information

Experimental procedures, compound characterization data, and copies of NMR spectra (PDF).

References

  • 1.(a) Forner A, Llovet JM, Bruix J. Lancet. 2012;379:1245–55. doi: 10.1016/S0140-6736(11)61347-0. [DOI] [PubMed] [Google Scholar]; (b) Dhanasekaran R, Limaye A, Cabrera R. Hepatic Med: Evidence Res. 2012;4:19–37. doi: 10.2147/HMER.S16316. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Ahmad J, Rabinovitz M. Etiology and Epidemiology of Hepatocellular Carcinoma. In: Carr BI, editor. Current Clinical Oncology: Hepatocellular Cancer: Diagnosis and Treatment. Humana Press; Totowa, NJ: 2005. pp. 1–21. [Google Scholar]
  • 2.Facciorusso A, Licinio R, Carr BI, Di Leo A, Barone M. Expert Rev Gastroenterol Hepatol. 2015;9:993–1003. doi: 10.1586/17474124.2015.1040763. [DOI] [PubMed] [Google Scholar]
  • 3.El-Serag HB. N Engl J Med. 2011;365:1118–1127. doi: 10.1056/NEJMra1001683. [DOI] [PubMed] [Google Scholar]
  • 4.Hung H. Curr Cancer Drug Targets. 2005;5:131–8. doi: 10.2174/1568009053202063. [DOI] [PubMed] [Google Scholar]
  • 5.Ruff P. Am J Cancer. 2004;3:119–131. [Google Scholar]
  • 6.DeFrances MC, Michalopoulos GK. In: Current Clinical Oncology: Hepatocellular Cancer: Diagnosis and Treatment. Carr BI, editor. Humana Press; Totowa, NJ: 2005. pp. 23–57. [Google Scholar]
  • 7.(a) Llovet JM, Ricci S, Mazzaferro V, Hilgard P, Gane E, Blanc J-F, Cosme de Oliveira A, Santoro A, Raoul J-L, Forner A, Schwartz M, Porta C, Zeuzem S, Bolondi L, Greten TF, Galle PR, Seitz J-F, Borbath I, Haussinger D, Giannaris T, Shan M, Moscovici M, Voliotis D, Bruix J. N Engl J Med. 2008;359:378–390. doi: 10.1056/NEJMoa0708857. [DOI] [PubMed] [Google Scholar]; (b) Keating GM, Santoro A. Drugs. 2009;69:223–40. doi: 10.2165/00003495-200969020-00006. [DOI] [PubMed] [Google Scholar]; (c) Siegel AB, Olsen SK, Magun A, Brown RS., Jr Hepatol. 2010;52:360–369. doi: 10.1002/hep.23633. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Bupathi M, Kaseb A, Meric-Bernstam F, Naing A. Mol Oncol. 2015;9:1501–9. doi: 10.1016/j.molonc.2015.06.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Hsieh TJ, Chang FR, Chia YC, Chen CY, Chiu HF, Wu YC. J Nat Prod. 2001;64:616–9. doi: 10.1021/np0005208. [DOI] [PubMed] [Google Scholar]
  • 10.(a) Leboeuf M, Streith J, Cavé A. Ann Pharm Françaises. 1975;33:43–47. [PubMed] [Google Scholar]; (b) Rao JUM, Giri GS, Hanumaiah T, Rao KVJ. J Nat Prod. 1986;49:346–7. [Google Scholar]; (c) Yang TH, Huang WY. Chin Pharm J. 1989;41:279–87. [Google Scholar]; (d) Hseih TJ, Chang FR, Wu YC. J Chin Chem Soc. 1999;46:607–11. [Google Scholar]
  • 11.(a) Whistler WA. Tongan Herbal Medicine. Isle Botanica; Honolulu, HI: 1992. pp. 86–87. [Google Scholar]; (b) Weiner MA. Secrets of Fijian Medicine. Quantum; 1983. p. 60. [Google Scholar]
  • 12.(a) Sehrawat A, Kumar V. In: Bioactive Foods and Extracts. Watson RR, Preedy VR, editors. CRC Press; 2011. pp. 555–82. [Google Scholar]; (b) Zhao C-Y, Liu X, Shen C, Wang Y, Qiao L. In: Mol Aspects Hepatocellular Carcinoma. Qiao L, editor. Bentham Science; 2012. pp. 167–173. [Google Scholar]
  • 13.Clery RA, Cason JRL, Zelenay V. J Nat Prod. 2016;64:4566–73. doi: 10.1021/acs.jafc.6b00680. [DOI] [PubMed] [Google Scholar]
  • 14.Liu L, Cao Y, Chen C, Zhang X, McNabola A, Wilkie D, Wilhelm S, Lynch M, Carter C. Cancer Res. 2006;66:11851–8. doi: 10.1158/0008-5472.CAN-06-1377. [DOI] [PubMed] [Google Scholar]
  • 15.Lião LM. Alkaloids. 2003;60:287–343. doi: 10.1016/s0099-9598(03)60005-2. [DOI] [PubMed] [Google Scholar]
  • 16.(a) Su A, Wu HK, He H, Slukhan U, Aisa HA. Helv Chim Acta. 2010;93:33–38. [Google Scholar]; (b) He F, Nugroho AE, Wong CP, Hirasawa Y, Shirota O, Morita H, Aisa HA. Chem Pharm Bull. 2012;60:213–218. doi: 10.1248/cpb.60.213. [DOI] [PubMed] [Google Scholar]
  • 17.Craig D, Henry GD. Eur J Org Chem. 2006:3558–3561. [Google Scholar]
  • 18.(a) Büchi G, Goldman IM, Mayo DW. J Am Chem Soc. 1966;88:3109–3133. [Google Scholar]; (b) Cren MC, Defaye G, Fetizon M. Bull Soc Chim Fr. 1970:3020–3022. [Google Scholar]; (c) Van der Gen A, Van der Linde LM, Witteveen JG. Recueil Trav Chim Pays-Bas. 1972;91:1433–40. [Google Scholar]; (d) Okatani T, Koyama J, Tagahara K, Suzuta Y. Heterocycles. 1987;26:595–7. [Google Scholar]; (e) Koyama J, Okatani T, Tagahara K, Suzuta Y, Irie H. Heterocycles. 1987;26:925–7. [Google Scholar]
  • 19.Vyvyan JR, Brown RC, Woods BP. J Org Chem. 2009;74:1374–1376. doi: 10.1021/jo802267n. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Vyvyan JR, Dell JA, Ligon TJ, Motanic KK, Wall HS. Synthesis. 2010:3637–3644. doi: 10.1055/s-0030-1258237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Vyvyan JR, Meyer JA, Meyer KD. J Org Chem. 2003;68:9144–9147. doi: 10.1021/jo035112y. [DOI] [PubMed] [Google Scholar]
  • 22.Kolb HC, Sharpless KB. Tetrahedron. 1992;48:10515–30. [Google Scholar]
  • 23.Baldwin JE. J Chem Soc, Chem Commun. 1976:734–6. [Google Scholar]
  • 24.We have found only three reports of an 8-endo epoxide cyclization:; (a) Arnone A, Bernardi R, Bravo P, Frigerio M, Ticozzi C. Gazz Chem Ital. 1989;119:87–94. [Google Scholar]; (b) Takabatake K, Nishi I, Shindo M, Shishido K. J Chem Soc, Perkin Trans 1. 2000:1807–8. [Google Scholar]; (c) Yang L, Deng G, Wang D-X, Huang Z-T, Zhu J-P, Wang M-X. Org Lett. 2007;9:1387–1390. doi: 10.1021/ol070292+. [DOI] [PubMed] [Google Scholar]
  • 25.Zou G, Reddy YK, Falck JR. Tetrahedron Lett. 2001;42:7213–15. [Google Scholar]
  • 26.(a) Xu D, Crispino GA, Sharpless KB. J Am Chem Soc. 1992;114:7550–1. [Google Scholar]; (b) Crispino GA, Sharpless KB. Synlett. 1993:47–8. [Google Scholar]; (c) Becker H, Soler MA, Sharpless KB. Tetrahedron. 1995;51:1345–76. [Google Scholar]
  • 27.(a) Moore CJ, Possner S, Hayes P, Paddon-Jones GC, Kitching W. J Org Chem. 1999;64:9742–4. [Google Scholar]; (b) Winne JM, Guang B, D’herde J, De Clerq PJ. Org Lett. 2006;8:4815–8. doi: 10.1021/ol061962z. [DOI] [PubMed] [Google Scholar]
  • 28.(a) Evans DA, Morrissey MM. J Am Chem Soc. 1984;106:3866–8. [Google Scholar]; (b) Evans DA, Morrissey MM. Tetrahedron Lett. 1984;25:4637–40. [Google Scholar]; (c) Evans DA, Morrissey MM, Dow RL. Tetrahedron Lett. 1985;26:6005–8. [Google Scholar]; (d) Hoveyda AH, Evans DA, Fu GC. Chem Rev. 1993;93:1307–70. [Google Scholar]
  • 29.Crabtree RH, Davis MW. Organometallics. 1983;2:681–2. [Google Scholar]
  • 30.Thompson HW, McPherson E. J Am Chem Soc. 1974;96:6232–3. [Google Scholar]

Associated Data

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

Supplementary Materials

supplement

RESOURCES