ACS Publications. Most Trusted. Most Cited. Most Read
Stereocontrolled Synthesis of the Portimine Skeleton via Organocatalyst-Mediated Asymmetric Stannylation and Stereoretentive C(sp3)–C(sp2) Stille Coupling
My Activity
  • Open Access
Letter

Stereocontrolled Synthesis of the Portimine Skeleton via Organocatalyst-Mediated Asymmetric Stannylation and Stereoretentive C(sp3)–C(sp2) Stille Coupling
Click to copy article linkArticle link copied!

Open PDFSupporting Information (1)

Organic Letters

Cite this: Org. Lett. 2025, 27, 4, 942–947
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.orglett.4c04245
Published January 22, 2025

Copyright © 2025 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

Our efforts toward the synthesis of the marine natural product portimine are described. The key to the synthesis of the skeleton is a stereoretentive copper-catalyzed C(sp3)–C(sp2) Stille-type cross-coupling that enables the convergent assembly of functionalized fragments. The core skeleton of portimine was constructed via ring-closing metathesis and transannular acetal formation.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2025 The Authors. Published by American Chemical Society

Of the many known marine natural products, cyclic imine toxins from dinoflagellates have in particular attracted much interest due to their extremely diverse biological activities and unique chemical structures. (1) However, the strong neurotoxicity of these marine natural products complicates their use as lead compounds in the development of pharmaceuticals. Portimine (1) is a cyclic imine natural product isolated originally from a marine dinoflagellate Vulcanodinium rugosum in New Zealand, (2) and is of particular interest to our laboratory. Its relative and absolute configuration was determined by Tanaka and co-workers (Figure 1). (3) Portimine (1) exhibits potent anticancer, (2) antifouling, (4) and anti-HIV-1 activities. (5) Furthermore, portimine (1) was recently shown to selectively induce apoptosis in human cancer cell lines (6) yet shows low acute toxicity in mice. Given these potent biological activities and low toxicity, portimine (1) has significant potential as a pharmaceutical lead compound and/or reagent. Portimine (1) has a 5-membered cyclic imine structure, which is very rare in cyclic imine natural products. Other unique molecular features of portimine (1) include a cyclohexene ring containing a quaternary carbon center (C-3), a highly oxidized 14-membered macrocyclic core skeleton, and a medium-sized cyclic acetal. The remarkable biological activities and unique chemical structure of portimine (1) make it an attractive synthetic target. Synthetic studies of portimine (1) have been reported by the Fujiwara, (7) Brimble, (8) Harran, (9) and Zakarian groups, (10) and recently Baran and co-workers reported the first elegant total synthesis of portimine (1). (11)

Figure 1

Figure 1. Structure and our retrosynthetic analysis of portimine (1).

Our group has been working on the total synthesis of marine natural products using powerful synthetic methodologies (i.e., transition-metal-catalyzed carbon–carbon bond-forming coupling chemistry). (12) In the course of this research, we began investigating the total synthesis of portimine (1). We describe herein our synthetic efforts toward the total synthesis of portimine (1) based on our original retrosynthesis approach.

This approach toward portimine (1) is illustrated in Figure 1. Our synthetic target was compound 2 containing a macrocyclic core skeleton which an intermediate toward (1) by cyclic imine formation and oxidation at C-13, C-14, and C-21 in a late stage of synthesis. We envisioned that macrocyclic compound 2 could be synthesized from compound 3 by ring-closing metathesis (RCM) (13) between C-13 and C-14, followed by transannular acetal formation. Bond disconnection at the α-alkoxyketone moiety (between C-4 and C-5) in compound 3 gives chiral α-alkoxyalkylstannane 4 and thioester 5, which would then be convergently coupled by copper-catalyzed C(sp3)-C(sp2) Stille-type cross-coupling. (14) The advantage of this key last step is that the coupling of chiral α-alkoxyalkylstannane would proceed with high stereochemical retention (specifically at the C-5 position for stannane 4), but there are no examples in the literature of the application of this coupling reaction to substrates with complex structures such as 4 and 5. In addition, the steric bulkiness of thioester 5 is likely to hinder efficient coupling. It was assumed that chiral α-alkoxyalkylstannane 4 could be easily synthesized from aldehyde 6 via organocatalyst-mediated asymmetric stannylation. (15) Our aim was to synthesize thioester 5 from readily available starting materials 7 and 8, controlling the configuration at C-3 and C-16 using the asymmetric Diels–Alder reaction developed by Evans. (16)

The synthesis of chiral α-alkoxyalkylstannane is illustrated in Scheme 1. The Evans syn-aldol reaction (17) of aldehyde 9 (18) and N-acyloxazolidinone 10 (19) proceeded smoothly to give alcohol 11 in 87% yield with excellent diastereoselectivity (dr >20:1). Reduction of 11 using LiBH4 and subsequent tosylation gave alcohol 12 in 90% yield over two steps. After protection of the secondary alcohol of 12 as a diisopropylsilyl (TIPS) ether, the cyano group was introduced by treatment with NaCN, and then reduced with diisobutylaluminum hydride (DIBALH) at −78 °C. Methylenation of the resulting aldehyde under Wittig olefination conditions provided compound 13 in 72% yield over four steps. Treatment of compound 13 with NaOH at 60 °C smoothly and selectively removed the tert-butyldimethylsilyl (TBS) group, and the corresponding primary alcohol was obtained in 99% yield. Parikh-Doering oxidation (SO3·pyridine, Et3N, CH2Cl2, DMSO) (20) of this alcohol provided aldehyde 14 in 94% yield. Various methods for the asymmetric synthesis of α-alkoxyalkylstannanes have been reported to date. (21) We chose a facile organocatalyst-mediated asymmetric stannylation requiring only mild reaction conditions. (15) Thus, in accordance with Falck’s report, treatment of aldehyde 14 with (R)- diphenylprolinol in the presence of n-butyltin hydride and diethylzinc led to asymmetric stannylation with excellent diastereoselectivity (dr 97:3). However, the obtained alcohol was unstable and thus was immediately converted to thiocarbamate 15 using a two-step protection sequence. The configuration of the newly created stereogenic center by organocatalyst-mediated asymmetric stannylation was established by a modified Mosher analysis approach (see Supporting Information). (22)

Scheme 1

Scheme 1. Synthesis of α-Alkoxyalkylstannane 15

The synthesis of the thioester is shown in Scheme 2. Asymmetric Diels–Alder reaction of α-methylene γ-lactam 7 (23) and diene 16 (23) proceeded smoothly under Evans reaction conditions (16) in the presence of [Cu((S,S)-tert-Bu-box)](SbF6)2 to give spirocyclic cyclohexene 17 in 89% yield with excellent diastereo and enantioselectivity (dr >20:1 and er 98:2). (24) Evans conditions have been applied to the synthesis of δ-lactam (25) and ε-lactam, (26) but to our knowledge, the present report is the first example of its application for γ-lactam synthesis. Removal of the Cbz group using Et2AlCl and PhSMe (27) gave the corresponding N–H lactam in quantitative yield, and subsequent introduction of a 2-nitrobenzenesulfonyl (Ns) group by using n-BuLi and NsCl gave Ns lactam 18 in 90% yield. Thus, the Cbz group was converted to the Ns group. Ring-opening of the lactam was carried out using LiOH, and carboxylic acid 19 was obtained in 86% yield. Various methods to introduce thioesters were unsuccessful (i.e., when the carboxyl group in 19 was activated, lactam 18 was easily formed by the nucleophilic attack of Ns amide). This problem was addressed by masking the carboxyl group in situ by treating 19 with TMS-imidazole. Next, the Ns amide was protected with tert-butoxycarbonyl (Boc) under standard reaction conditions (Boc2O, Et3N, DMAP) to obtain carboxylic acid 20 in 80% yield in a one-pot operation. Finally, 20 was converted into the corresponding acid chloride, then treatment with 4-nitrothiophenol and pyridine completed the synthesis of thioester 21.

Scheme 2

Scheme 2. Synthesis of Thioester 21

With the coupling substrates 15 and 21 in hand, we examined the key copper-catalyzed C(sp3)-C(sp2) Stille-type cross-coupling reaction (Scheme 3). (14) Investigation of the reaction conditions pointed to the use of copper(I) thiophene-2-carboxylate (CuTC) (14b) as a catalyst and heating in toluene. This removed the Boc group and thus we reintroduced the Boc group after the coupling reaction to afford the coupling product 22 in 61% overall yield with excellent diastereoselectivity (dr >20:1). However, the coupling reaction proceeded very slowly, taking 5 days to reach 48% yield. Further extension of reaction time or addition of catalysts resulted in lower yields. This was attributed to the byproduct Bu3SnS-p-NO2Ph produced during the reaction process, (14b) which was coordinated to the copper catalyst and inhibited cross-coupling. Therefore, the reaction was carried out in a short time, and the resulting Bu3SnS-p-NO2Ph was removed. Then the same reaction was repeated three times and the product was pooled to obtain the total yield. We confirmed the configuration of the stereogenic center at the C-5 position in compound 22 as follows. Exposure of 22 to m-CPBA resulted in the selective removal of thiocarbamate (28) to give formyl ester 23. Subsequent treatment of 23 with K2CO3 in MeOH provided alcohol 24 in 78% yield over two steps. Alcohol 24 was converted to (R)- and (S)-MTPA esters and their configurations were confirmed by a modified Mosher analysis, (22) as shown in Scheme 3 (see Supporting Information for details). The key coupling reaction proceeded with extremely high stereochemical retention, and the chirality of α-alkoxyalkylstannane 15 was completely transferred to the coupling product 22. Similar couplings have occasionally been applied to the synthesis of natural products. (29) However, to our knowledge, the present study is the first to apply stereoretentive copper-catalyzed C(sp3)-C(sp2) Stille-type cross-coupling between diversely functionalized compounds such as 15 and 21.

Scheme 3

Scheme 3. Key Coupling and Stereochemical Conformation of Alcohol 24 through Modified Mosher Analysis

Having achieved efficient coupling, the next challenge was the formation of a 14-membered macrocycle by RCM (Scheme 4). The tert-butyldiphenylsilyl (TBDPS) group was selectively removed using a mixture of TBAF/AcOH to obtain primary alcohol 26 in 91% yield. Swern oxidation of alcohol 26 ((COCl)2, Et3N, DMSO) (30) proceeded smoothly, but the corresponding aldehyde was very unstable and decomposed during workup and purification. Therefore, the subsequent Grignard reaction was carried out immediately in a one-pot operation after the Swern oxidation step to obtain propargyl alcohol 27 in 98% overall yield with excellent diastereoselectivity (dr >20:1). The configuration of the newly generated stereogenic center at the C-15 position by Grignard reaction was established by a modified Mosher analysis (see Supporting Information). (22) We also attempted to introduce vinyl group but the reaction was surprisingly very sluggish, in contrast to the introduction of ethynyl group, and thus our attempts to introduce functional groups other than ethynyl were unsuccessful. Removal of the Ns group with PhSH, K2CO3, in DMF provided the corresponding amide in 91% yield. The terminal alkyne (at C-14) was partially hydrogenated by using Pd/CaCO3, 3,6-dithia-1,8-octanediol (A) (31) under a H2 atmosphere to afford the corresponding allylic alcohol in quantitative yield. Furthermore, PDC oxidation of allylic alcohol (at C-15) gave enone 28 in 88% yield. Having obtained substrate 28, we attempted the RCM reaction. RCM reaction using the second Hoveyda-Grubbs catalyst (HG-II, 21 mol %) (13) in toluene (1.0 mM) provided 14-membered macrocyclic compound 29 in excellent yield (82%, E/Z = 5.8/1).

Scheme 4

Scheme 4. Construction of 14-Membered Macrocycles by RCM

The next challenge was to construct the portimine core skeleton by transannular acetal formation. We began our investigations using 29-(E) (Scheme 5, upper column). Luche reduction (NaBH4, CeCl3·7H2O, EtOH, – 78 to −60 °C) (32) of 29-(E) proceeded regioselectively at the C-15 ketone to give the desired allylic alcohol in 73% yield, along with its diastereomer in 18% yield. Removal of the TIPS group using HF·pyridine in THF/pyridine (1:1) gave diol 30 in 92% yield. We investigated many reaction conditions for transannular acetal formation using diol 30 but were unable to obtain the desired acetal 31, possibly because distortion of acetal 31 by E-olefin made the cyclization process unfavorable. We therefore examined transannular acetal formation starting from 29-(Z) (Scheme 5, lower column).

Scheme 5

Scheme 5. Challenges of Transannular Acetal Formation and Construction of the Macrocyclic Acetal Skeleton of Portimine

Luche reduction (NaBH4, CeCl3·7H2O, EtOH, −78 to −60 °C) (32) of 29-(Z) proceeded with excellent regio and diastereoselectivity to give the allylic alcohol in 90% yield with excellent diastereoselectivity (dr >20:1). The configuration of the newly generated stereogenic center at C-15 from ketone 29-(Z) by Luche reduction was established by a modified Mosher analysis (see Supporting Information). (22) Subsequent deprotection of the TIPS group gave diol 32 in 90% yield. To our delight, transannular acetal formation of 32 proceeded smoothly by the action of TsOH·H2O in CH2Cl2 to afford acetal 33 in 91% yield. Selective removal of the thiocarbamate group (28) by exposure of 33 to m-CPBA provided formyl ester 34. Finally, alcohol 35, with the same core skeleton as portimine, was synthesized by removal of the formyl group in 63% yield over two steps. Thus, our next challenge will be oxidation at C-13 and C-14.

The highlights of our synthesis are organocatalyst-mediated asymmetric stannylation to synthesize chiral α-alkoxyalkylstannane 15, and subsequent stereoretentive copper-catalyzed C(sp3)-C(sp2) Stille-type cross-coupling to achieve convergent fragment assembly of 15 and 21. It is worth noting that this coupling proceeded with high stereochemical retention. Furthermore, RCM and transannular acetal formation allowed us to construct the core skeleton of portimine. Further studies are currently underway toward the total synthesis of the natural product portimine (1).

Data Availability

Click to copy section linkSection link copied!

The data underlying this study are available in the published article and its Supporting Information.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.orglett.4c04245.

  • Experimental procedures, characterization data, and NMR spectra of all newly synthesized compounds. (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
  • Author
    • Daisuke Sato - Graduate School of Life Sciences, Tohoku University, 2-1-1 Katahira, 980-8577 Aoba-ku, Sendai, Japan
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

This work was financially supported by JSPS KAKENHI Grant Number JP23K14314. We thank Dr. Shota Nagasawa (Tohoku University) for FAB and ESI-TOF mass measurements.

References

Click to copy section linkSection link copied!

This article references 32 other publications.

  1. 1
    (a) Stivala, C. E.; Benoit, E.; Aráoz, R.; Servent, D.; Novikov, A.; Molgó, J.; Zakarian, A. Synthesis and biology of cyclic imine toxins, an emerging class of potent, globally distributed marine toxins. Nat. Prod. Rep. 2015, 32, 411435,  DOI: 10.1039/C4NP00089G
    (b) Molgó, J.; Marchot, P.; Aráoz, R.; Benoit, E.; Iorga, B. I.; Zakarian, A.; Taylor, P.; Bourne, Y.; Servent, D. Cyclic imine toxins from dinoflagellates: A growing family of potent antagonists of the nicotinic acetylcholine receptors. J. Neurochem. 2017, 142, 4151,  DOI: 10.1111/jnc.13995
  2. 2
    Selwood, A. I.; Wilkins, A. L.; Munday, R.; Shi, F.; Rhodes, L. L.; Holland, P. T. Portimine: a bioactive metabolite from the benthic dinoflagellate Vulcanodinium rugosum. Tetrahedron Lett. 2013, 54, 47054707,  DOI: 10.1016/j.tetlet.2013.06.098
  3. 3
    Hermawan, I.; Higa, M.; Hutabarat, P. U. B.; Fujiwara, T.; Akiyama, K.; Kanamoto, A.; Haruyama, T.; Kobayashi, N.; Higashi, M.; Suda, S.; Tanaka, J. Kabirimine, a New Cyclic Imine from an Okinawan Dinoflagellate. Mar. Drugs 2019, 17, 353361,  DOI: 10.3390/md17060353
  4. 4
    Brooke, D. G.; Cervin, G.; Champeau, O.; Harwood, D. T.; Pavia, H.; Selwood, A. I.; Svenson, J.; Tremblay, L. A.; Cahill, P. L. Antifouling activity of portimine, select semisynthetic analogues, and other microalga-derived spirocyclic imines. Biofouling 2018, 34, 950961,  DOI: 10.1080/08927014.2018.1514461
  5. 5
    Izumida, M.; Suga, K.; Ishibashi, F.; Kubo, Y. The Spirocyclic Imine from a Marine Benthic Dinoflagellate, Portimine, Is a Potent Anti-Human Immunodeficiency Virus Type 1 Therapeutic Lead Compound. Mar. Drugs 2019, 17, 495,  DOI: 10.3390/md17090495
  6. 6
    Cuddihy, S. L.; Drake, S.; Harwood, D. T.; Selwood, A. I.; McNabb, P. S.; Hampton, M. B. The marine cytotoxin portimine is a potent and selective inducer of apoptosis. Apoptosis 2016, 21, 14471452,  DOI: 10.1007/s10495-016-1302-x
  7. 7
    Saito, T.; Fujiwara, K.; Kondo, Y.; Akiba, U.; Suzuki, T. Synthesis of the cyclohexene segment of portimine. Tetrahedron Lett. 2019, 60, 386389,  DOI: 10.1016/j.tetlet.2018.12.063
  8. 8
    (a) Aitken, H. R. M.; Brimble, M. A.; Furkert, D. P. A catalytic asymmetric ene reaction for direct preparation of α-hydroxy 1, 4- diketones as intermediates in natural product synthesis. Synlett 2020, 31, 687690,  DOI: 10.1055/s-0037-1610748
    (b) Ding, X.; Aitken, H. R. M.; Pearl, E. S.; Furkert, D. P.; Brimble, M. A. Synthesis of the C4–C16 Polyketide Fragment of Portimines A and B. J. Org. Chem. 2021, 86, 1284012850,  DOI: 10.1021/acs.joc.1c01463
  9. 9
    Li, L.; El Khoury, A.; Clement, B. O.; Wu, C.; Harran, P. G. Asymmetric organocatalysis enables rapid assembly of portimine precursor chains. Org. Lett. 2022, 24, 26072612,  DOI: 10.1021/acs.orglett.2c00556
  10. 10
    Lee, H. J.; Gladfelder, J. J.; Zakarian, A. Remoto Stereocontrol in the Ireland-Claisen Rearrangement by δ-Alkyl Group. J. Org. Chem. 2023, 88, 75607563,  DOI: 10.1021/acs.joc.3c00535
  11. 11
    Tang, J.; Li, W.; Chiu, T. Y.; Martínes-Peña, F.; Luo, Z.; Chong, C. T.; Wei, Q.; Gazaniga, N.; West, T. J.; See, Y. Y.; Lairson, L. L.; Parker, C. G.; Baran, P. S. Synthesis of portimines reveals the basis of their anti-cancer activity. Nature 2023, 622, 507513,  DOI: 10.1038/s41586-023-06535-1
  12. 12
    (a) Iwasaki, K.; Sasaki, S.; Kasai, Y.; Kawashima, Y.; Sasaki, S.; Ito, T.; Yotsu-Yamashita, M.; Sasaki, M. Total Synthesis of Polycavernosides A and B, Two Lethal Toxins from Red Alga. J. Org. Chem. 2017, 82, 1320413219,  DOI: 10.1021/acs.joc.7b02293
    (b) Sasaki, M.; Iwasaki, K.; Arai, K.; Hamada, N.; Umehara, A. Convergent Synthesis of the HIJKLMN-Ring Fragment of Caribbean Ciguatoxin C-CTX-1 by a Late-Stage Reductive Olefin Coupling Approach. Bull. Chem. Soc. Jpn. 2022, 95, 819824,  DOI: 10.1246/bcsj.20220070
    (c) Suwa, T.; Sasaki, M.; Umehara, A. Total Synthesis of (−)-Irijimaside A Enabled by Ni/Zr-Mediated Reductive Ketone Coupling. Org. Lett. 2024, 26, 43774382,  DOI: 10.1021/acs.orglett.4c01367
    (d) Sasaki, M.; Ohba, M.; Murakami, A.; Umehara, A. Convergent and Scalable Second-Generation Synthesis of the Fully Functionalized HIJKLMN-Ring Segment of Caribbean Ciguatoxin C-CTX-1. J. Org. Chem. 2024, 89, 1863118639,  DOI: 10.1021/acs.joc.4c02723
  13. 13
    (a) Kingsbury, J. S.; Harrity, J. P. A.; Bonitatebus, P. J., Jr.; Hoveyda, A. H. A Recyclable Ru-Based Metathesis Catalyst. J. Am. Chem. Soc. 1999, 121, 791799,  DOI: 10.1021/ja983222u
    (b) Garber, S. B.; Kingsbury, J. S.; Gray, B. L.; Hoveyda, A. H. Efficient and Recyclable Monomeric and Dendritic Ru-Based Metathesis Catalysts. J. Am. Chem. Soc. 2000, 122, 81688179,  DOI: 10.1021/ja001179g
  14. 14
    (a) Falck, J. R.; Bhatt, R. K.; Ye, J. Tin-Copper Transmetalation: Cross-Coupling of alpha-Heteroatom-Substituted Alkyltributylstannanes with Organohalides. J. Am. Chem. Soc. 1995, 117, 59735982,  DOI: 10.1021/ja00127a010
    (b) Li, H.; He, A.; Falck, J. R.; Liebeskind, L. S. Stereocontrolled Synthesis of α-Amino-α′-alkoxy Ketones by a Copper-Catalyzed Cross-Coupling of Peptidic Thiol Esters and α-Alkoxyalkylstannanes. Org. Lett. 2011, 13, 36823685,  DOI: 10.1021/ol201330j
    (c) Cordovilla, C.; Bartolomé, C.; Martínez-Ilarduya, J. M.; Espinet, P. The Stille Reaction, 38 Years Later. ACS Catal. 2015, 5, 30403053,  DOI: 10.1021/acscatal.5b00448
  15. 15
    He, A.; Falck, J. R. Synthesis of Enantioenriched α-(Hydroxyalkyl)-tri-n-butylstannanes. Angew. Chem., Int. Ed. 2008, 47, 65866589,  DOI: 10.1002/anie.200802313
  16. 16
    (a) Evans, D. A.; Miller, S. J.; Lectka, T.; von Matt, P. Chiral bis(oxazoline)copper(II) complexes as Lewis acid catalysts for the enantioselective Diels-Alder reaction. J. Am. Chem. Soc. 1999, 121, 75597573,  DOI: 10.1021/ja991190k
    (b) Evans, D. A.; Barnes, D. M.; Johnson, J. S.; Lectka, T.; von Matt, P.; Miller, S. J.; Murry, J. A.; Norcross, R. D.; Shaughnessy, E. A.; Campos, K. R. Bis(oxazoline) and Bis(oxazolinyl)pyridine Copper Complexes as Enantioselective Diels–Alder Catalysts: Reaction Scope and Synthetic Applications. J. Am. Chem. Soc. 1999, 121, 75827594,  DOI: 10.1021/ja991191c
  17. 17
    (a) Evans, D. A.; Bartroli, J.; Shih, T. L. Enantioselective Aldol Condensations. 2. Erythro-Selective Chiral Aldol Condensations via Boron Enolates. J. Am. Chem. Soc. 1981, 103, 21272129,  DOI: 10.1021/ja00398a058
    (b) Evans, D. A.; Nelson, J. V.; Vogel, E.; Taber, T. R. Stereoselective aldol condensations via boron enolates. J. Am. Chem. Soc. 1981, 103, 30993111,  DOI: 10.1021/ja00401a031
  18. 18

    Aldehyde 9 was readily synthesized from commercially available 1,4-cyclohexanedione monoethyleneketal in five steps. For details, see the Supporting Information.

  19. 19
    N-Acyloxazolidinone 10 was readily synthesized by DMAPO/Boc2O-madiated direct N-acylation method of oxazolidinone. For details, see the Supporting Information. Also see:Umehara, A.; Shimizu, S.; Sasaki, M. DMAPO/Boc2O-Mediated One-Pot Direct N-Acylation of Less Nucleophilic N-Heterocycles with Carboxylic Acids. ChemCatChem. 2023, 15, e202201596  DOI: 10.1002/cctc.202201596
  20. 20
    Parikh, J. R.; Doering, W. v. E. Sulfur Trioxide in the Oxidation of Alcohols by Dimethyl Sulfoxide. J. Am. Chem. Soc. 1967, 89, 55055507,  DOI: 10.1021/ja00997a067
  21. 21

    For selected methods for synthesis of chiral α-alkoxyalkylstannanes, see:

    (a) Still, W. C.; Sreekumar, C. α-alkoxyorganolithium reagents. A new class of configurationally stable carbanions for organic synthesis. J. Am. Chem. Soc. 1980, 102, 12011202,  DOI: 10.1021/ja00523a066
    (b) Chan, P. C. M.; Chong, J. M. Asymmetric Reduction of Acylstannanes. Preparation of Enantiomerically Enriched α-Alkoxystannanes. J. Org. Chem. 1988, 53, 55845586,  DOI: 10.1021/jo00258a048
    (c) Christoph, G.; Hoppe, D. Asymmetric Synthesis of 2-Alkenyl-1-cyclopentanols via Tin–Lithium Exchange and Intramolecular Cycloalkylation. Org. Lett. 2002, 4, 21892192,  DOI: 10.1021/ol026068w
  22. 22
    Ohtani, I.; Kusumi, T.; Kashman, Y.; Kakisawa, H. High-Field FT NMR Application of Mosher’s Method. The Absolute Configurations of Marine Terpenoids. J. Am. Chem. Soc. 1991, 113, 40924096,  DOI: 10.1021/ja00011a006
  23. 23

    γ-Lactam 7 and diene 16 were readily synthesized using a known procedure. For details, see the Supporting Information.

  24. 24

    The absolute configuration of the C16 stereogenic center was determined by the phenylglycine methyl ester (PGME) method, and the relative stereochemistry of the C16/C3 stereogenic centers was determined by NOE correlations. See Supporting Information for details.

  25. 25
    (a) Tsujimoto, T.; Ishihara, J.; Horie, M.; Murai, A. Asymmetric construction of the azaspiro[5.5]undec-8-ene system towards gymnodimine synthesis. Synlett 2002, 2002, 399402,  DOI: 10.1055/s-2002-20450
    (b) Kong, K.; Moussa, Z.; Lee, C.; Romo, D. Total Synthesis of the Spirocyclic Imine Marine Toxin (−)-Gymnodimine and an Unnatural C4-Epimer. J. Am. Chem. Soc. 2011, 133, 1984419856,  DOI: 10.1021/ja207385y
  26. 26
    (a) Ishihara, J.; Horie, M.; Shimada, Y.; Tojo, S.; Murai, A. Asymmetric construction of the azaspiro[5.6]dodec-9-ene system in marine natural toxins. Synlett 2002, 2002, 403406,  DOI: 10.1055/s-2002-20451
    (b) Tsuchikawa, H.; Minamino, K.; Hayashi, S.; Murata, M. Efficient Access to the Functionalized Bicyclic Pharmacophore of Spirolide C by Using a Selective Diels–Alder Reaction. Asian J. Org. Chem. 2017, 6, 13221327,  DOI: 10.1002/ajoc.201700164
  27. 27
    Tsujimoto, T.; Murai, A. Efficient Detachment of N-Benzyl Carbamate Group. Synlett 2002, 2002, 1283,  DOI: 10.1055/s-2002-32986
  28. 28
    (a) Barma, D. K.; Bandyopadhyay, A.; Capdevila, J. H.; Falck, J. R. Dimethylthiocarbamate (DMTC): An Alcohol Protecting Group. Org. Lett. 2003, 5, 47554757,  DOI: 10.1021/ol0354573
    (b) Kojima, M.; Nakamura, Y.; Ishikawa, T.; Takeuchi, S. Fluorous dimethylthiocarbamate (FDMTC) protecting groups for alcohols. Tetrahedron Lett. 2006, 47, 63096314,  DOI: 10.1016/j.tetlet.2006.05.142
  29. 29
    (a) Ye, J.; Bhatt, R. K.; Falck, J. R. Stereospecific α-alkoxystannane couplings with acyl chlorides: Total synthesis of (+)-goniofufurone. Tetrahedron Lett. 1993, 34, 80078010,  DOI: 10.1016/S0040-4039(00)61436-3
    (b) Wang, R.; Falck, J. R. Studies towards asymmetric synthesis of 4(S)-11-dihydroxydocosahexaenoic acid (diHDHA) featuring cross-coupling of chiral stannane under mild conditions. Org. Biomol. Chem. 2015, 13, 16241628,  DOI: 10.1039/C4OB02324B
  30. 30
    Huang, S. L.; Omura, K.; Swern, D. Oxidation of Sterically Hindered Alcohols to Carbonyls with Dimethyl Sulfoxide-Trifluor-acetic Anhydride. J. Org. Chem. 1976, 41, 33293331,  DOI: 10.1021/jo00882a030
  31. 31
    Lindlar, H.; Dubuis, R. Palladium Catalyst for Partial Reduction of Acetylenes. Org. Synth. 1966, 46, 8992,  DOI: 10.1002/0471264180.os046.27
  32. 32
    (a) Luche, J.-L. Lanthanides in Organic Synthesis. 1. Selective 1,2 Reduction of Conjugated Ketones. J. Am. Chem. Soc. 1978, 100, 22262227,  DOI: 10.1021/ja00475a040
    (b) Gemal, A. L.; Luche, J.-L. Lanthanoids in Organic Synthesis. 6. The reduction of α-enones by Sodium Borohydride in the Presence of Lanthanoid Chlorides: Mechanistic Aspects. J. Am. Chem. Soc. 1981, 103, 54545459,  DOI: 10.1021/ja00408a029

Cited By

Click to copy section linkSection link copied!

This article has not yet been cited by other publications.

Organic Letters

Cite this: Org. Lett. 2025, 27, 4, 942–947
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.orglett.4c04245
Published January 22, 2025

Copyright © 2025 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

2695

Altmetric

-

Citations

-
Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Structure and our retrosynthetic analysis of portimine (1).

    Scheme 1

    Scheme 1. Synthesis of α-Alkoxyalkylstannane 15

    Scheme 2

    Scheme 2. Synthesis of Thioester 21

    Scheme 3

    Scheme 3. Key Coupling and Stereochemical Conformation of Alcohol 24 through Modified Mosher Analysis

    Scheme 4

    Scheme 4. Construction of 14-Membered Macrocycles by RCM

    Scheme 5

    Scheme 5. Challenges of Transannular Acetal Formation and Construction of the Macrocyclic Acetal Skeleton of Portimine
  • References


    This article references 32 other publications.

    1. 1
      (a) Stivala, C. E.; Benoit, E.; Aráoz, R.; Servent, D.; Novikov, A.; Molgó, J.; Zakarian, A. Synthesis and biology of cyclic imine toxins, an emerging class of potent, globally distributed marine toxins. Nat. Prod. Rep. 2015, 32, 411435,  DOI: 10.1039/C4NP00089G
      (b) Molgó, J.; Marchot, P.; Aráoz, R.; Benoit, E.; Iorga, B. I.; Zakarian, A.; Taylor, P.; Bourne, Y.; Servent, D. Cyclic imine toxins from dinoflagellates: A growing family of potent antagonists of the nicotinic acetylcholine receptors. J. Neurochem. 2017, 142, 4151,  DOI: 10.1111/jnc.13995
    2. 2
      Selwood, A. I.; Wilkins, A. L.; Munday, R.; Shi, F.; Rhodes, L. L.; Holland, P. T. Portimine: a bioactive metabolite from the benthic dinoflagellate Vulcanodinium rugosum. Tetrahedron Lett. 2013, 54, 47054707,  DOI: 10.1016/j.tetlet.2013.06.098
    3. 3
      Hermawan, I.; Higa, M.; Hutabarat, P. U. B.; Fujiwara, T.; Akiyama, K.; Kanamoto, A.; Haruyama, T.; Kobayashi, N.; Higashi, M.; Suda, S.; Tanaka, J. Kabirimine, a New Cyclic Imine from an Okinawan Dinoflagellate. Mar. Drugs 2019, 17, 353361,  DOI: 10.3390/md17060353
    4. 4
      Brooke, D. G.; Cervin, G.; Champeau, O.; Harwood, D. T.; Pavia, H.; Selwood, A. I.; Svenson, J.; Tremblay, L. A.; Cahill, P. L. Antifouling activity of portimine, select semisynthetic analogues, and other microalga-derived spirocyclic imines. Biofouling 2018, 34, 950961,  DOI: 10.1080/08927014.2018.1514461
    5. 5
      Izumida, M.; Suga, K.; Ishibashi, F.; Kubo, Y. The Spirocyclic Imine from a Marine Benthic Dinoflagellate, Portimine, Is a Potent Anti-Human Immunodeficiency Virus Type 1 Therapeutic Lead Compound. Mar. Drugs 2019, 17, 495,  DOI: 10.3390/md17090495
    6. 6
      Cuddihy, S. L.; Drake, S.; Harwood, D. T.; Selwood, A. I.; McNabb, P. S.; Hampton, M. B. The marine cytotoxin portimine is a potent and selective inducer of apoptosis. Apoptosis 2016, 21, 14471452,  DOI: 10.1007/s10495-016-1302-x
    7. 7
      Saito, T.; Fujiwara, K.; Kondo, Y.; Akiba, U.; Suzuki, T. Synthesis of the cyclohexene segment of portimine. Tetrahedron Lett. 2019, 60, 386389,  DOI: 10.1016/j.tetlet.2018.12.063
    8. 8
      (a) Aitken, H. R. M.; Brimble, M. A.; Furkert, D. P. A catalytic asymmetric ene reaction for direct preparation of α-hydroxy 1, 4- diketones as intermediates in natural product synthesis. Synlett 2020, 31, 687690,  DOI: 10.1055/s-0037-1610748
      (b) Ding, X.; Aitken, H. R. M.; Pearl, E. S.; Furkert, D. P.; Brimble, M. A. Synthesis of the C4–C16 Polyketide Fragment of Portimines A and B. J. Org. Chem. 2021, 86, 1284012850,  DOI: 10.1021/acs.joc.1c01463
    9. 9
      Li, L.; El Khoury, A.; Clement, B. O.; Wu, C.; Harran, P. G. Asymmetric organocatalysis enables rapid assembly of portimine precursor chains. Org. Lett. 2022, 24, 26072612,  DOI: 10.1021/acs.orglett.2c00556
    10. 10
      Lee, H. J.; Gladfelder, J. J.; Zakarian, A. Remoto Stereocontrol in the Ireland-Claisen Rearrangement by δ-Alkyl Group. J. Org. Chem. 2023, 88, 75607563,  DOI: 10.1021/acs.joc.3c00535
    11. 11
      Tang, J.; Li, W.; Chiu, T. Y.; Martínes-Peña, F.; Luo, Z.; Chong, C. T.; Wei, Q.; Gazaniga, N.; West, T. J.; See, Y. Y.; Lairson, L. L.; Parker, C. G.; Baran, P. S. Synthesis of portimines reveals the basis of their anti-cancer activity. Nature 2023, 622, 507513,  DOI: 10.1038/s41586-023-06535-1
    12. 12
      (a) Iwasaki, K.; Sasaki, S.; Kasai, Y.; Kawashima, Y.; Sasaki, S.; Ito, T.; Yotsu-Yamashita, M.; Sasaki, M. Total Synthesis of Polycavernosides A and B, Two Lethal Toxins from Red Alga. J. Org. Chem. 2017, 82, 1320413219,  DOI: 10.1021/acs.joc.7b02293
      (b) Sasaki, M.; Iwasaki, K.; Arai, K.; Hamada, N.; Umehara, A. Convergent Synthesis of the HIJKLMN-Ring Fragment of Caribbean Ciguatoxin C-CTX-1 by a Late-Stage Reductive Olefin Coupling Approach. Bull. Chem. Soc. Jpn. 2022, 95, 819824,  DOI: 10.1246/bcsj.20220070
      (c) Suwa, T.; Sasaki, M.; Umehara, A. Total Synthesis of (−)-Irijimaside A Enabled by Ni/Zr-Mediated Reductive Ketone Coupling. Org. Lett. 2024, 26, 43774382,  DOI: 10.1021/acs.orglett.4c01367
      (d) Sasaki, M.; Ohba, M.; Murakami, A.; Umehara, A. Convergent and Scalable Second-Generation Synthesis of the Fully Functionalized HIJKLMN-Ring Segment of Caribbean Ciguatoxin C-CTX-1. J. Org. Chem. 2024, 89, 1863118639,  DOI: 10.1021/acs.joc.4c02723
    13. 13
      (a) Kingsbury, J. S.; Harrity, J. P. A.; Bonitatebus, P. J., Jr.; Hoveyda, A. H. A Recyclable Ru-Based Metathesis Catalyst. J. Am. Chem. Soc. 1999, 121, 791799,  DOI: 10.1021/ja983222u
      (b) Garber, S. B.; Kingsbury, J. S.; Gray, B. L.; Hoveyda, A. H. Efficient and Recyclable Monomeric and Dendritic Ru-Based Metathesis Catalysts. J. Am. Chem. Soc. 2000, 122, 81688179,  DOI: 10.1021/ja001179g
    14. 14
      (a) Falck, J. R.; Bhatt, R. K.; Ye, J. Tin-Copper Transmetalation: Cross-Coupling of alpha-Heteroatom-Substituted Alkyltributylstannanes with Organohalides. J. Am. Chem. Soc. 1995, 117, 59735982,  DOI: 10.1021/ja00127a010
      (b) Li, H.; He, A.; Falck, J. R.; Liebeskind, L. S. Stereocontrolled Synthesis of α-Amino-α′-alkoxy Ketones by a Copper-Catalyzed Cross-Coupling of Peptidic Thiol Esters and α-Alkoxyalkylstannanes. Org. Lett. 2011, 13, 36823685,  DOI: 10.1021/ol201330j
      (c) Cordovilla, C.; Bartolomé, C.; Martínez-Ilarduya, J. M.; Espinet, P. The Stille Reaction, 38 Years Later. ACS Catal. 2015, 5, 30403053,  DOI: 10.1021/acscatal.5b00448
    15. 15
      He, A.; Falck, J. R. Synthesis of Enantioenriched α-(Hydroxyalkyl)-tri-n-butylstannanes. Angew. Chem., Int. Ed. 2008, 47, 65866589,  DOI: 10.1002/anie.200802313
    16. 16
      (a) Evans, D. A.; Miller, S. J.; Lectka, T.; von Matt, P. Chiral bis(oxazoline)copper(II) complexes as Lewis acid catalysts for the enantioselective Diels-Alder reaction. J. Am. Chem. Soc. 1999, 121, 75597573,  DOI: 10.1021/ja991190k
      (b) Evans, D. A.; Barnes, D. M.; Johnson, J. S.; Lectka, T.; von Matt, P.; Miller, S. J.; Murry, J. A.; Norcross, R. D.; Shaughnessy, E. A.; Campos, K. R. Bis(oxazoline) and Bis(oxazolinyl)pyridine Copper Complexes as Enantioselective Diels–Alder Catalysts: Reaction Scope and Synthetic Applications. J. Am. Chem. Soc. 1999, 121, 75827594,  DOI: 10.1021/ja991191c
    17. 17
      (a) Evans, D. A.; Bartroli, J.; Shih, T. L. Enantioselective Aldol Condensations. 2. Erythro-Selective Chiral Aldol Condensations via Boron Enolates. J. Am. Chem. Soc. 1981, 103, 21272129,  DOI: 10.1021/ja00398a058
      (b) Evans, D. A.; Nelson, J. V.; Vogel, E.; Taber, T. R. Stereoselective aldol condensations via boron enolates. J. Am. Chem. Soc. 1981, 103, 30993111,  DOI: 10.1021/ja00401a031
    18. 18

      Aldehyde 9 was readily synthesized from commercially available 1,4-cyclohexanedione monoethyleneketal in five steps. For details, see the Supporting Information.

    19. 19
      N-Acyloxazolidinone 10 was readily synthesized by DMAPO/Boc2O-madiated direct N-acylation method of oxazolidinone. For details, see the Supporting Information. Also see:Umehara, A.; Shimizu, S.; Sasaki, M. DMAPO/Boc2O-Mediated One-Pot Direct N-Acylation of Less Nucleophilic N-Heterocycles with Carboxylic Acids. ChemCatChem. 2023, 15, e202201596  DOI: 10.1002/cctc.202201596
    20. 20
      Parikh, J. R.; Doering, W. v. E. Sulfur Trioxide in the Oxidation of Alcohols by Dimethyl Sulfoxide. J. Am. Chem. Soc. 1967, 89, 55055507,  DOI: 10.1021/ja00997a067
    21. 21

      For selected methods for synthesis of chiral α-alkoxyalkylstannanes, see:

      (a) Still, W. C.; Sreekumar, C. α-alkoxyorganolithium reagents. A new class of configurationally stable carbanions for organic synthesis. J. Am. Chem. Soc. 1980, 102, 12011202,  DOI: 10.1021/ja00523a066
      (b) Chan, P. C. M.; Chong, J. M. Asymmetric Reduction of Acylstannanes. Preparation of Enantiomerically Enriched α-Alkoxystannanes. J. Org. Chem. 1988, 53, 55845586,  DOI: 10.1021/jo00258a048
      (c) Christoph, G.; Hoppe, D. Asymmetric Synthesis of 2-Alkenyl-1-cyclopentanols via Tin–Lithium Exchange and Intramolecular Cycloalkylation. Org. Lett. 2002, 4, 21892192,  DOI: 10.1021/ol026068w
    22. 22
      Ohtani, I.; Kusumi, T.; Kashman, Y.; Kakisawa, H. High-Field FT NMR Application of Mosher’s Method. The Absolute Configurations of Marine Terpenoids. J. Am. Chem. Soc. 1991, 113, 40924096,  DOI: 10.1021/ja00011a006
    23. 23

      γ-Lactam 7 and diene 16 were readily synthesized using a known procedure. For details, see the Supporting Information.

    24. 24

      The absolute configuration of the C16 stereogenic center was determined by the phenylglycine methyl ester (PGME) method, and the relative stereochemistry of the C16/C3 stereogenic centers was determined by NOE correlations. See Supporting Information for details.

    25. 25
      (a) Tsujimoto, T.; Ishihara, J.; Horie, M.; Murai, A. Asymmetric construction of the azaspiro[5.5]undec-8-ene system towards gymnodimine synthesis. Synlett 2002, 2002, 399402,  DOI: 10.1055/s-2002-20450
      (b) Kong, K.; Moussa, Z.; Lee, C.; Romo, D. Total Synthesis of the Spirocyclic Imine Marine Toxin (−)-Gymnodimine and an Unnatural C4-Epimer. J. Am. Chem. Soc. 2011, 133, 1984419856,  DOI: 10.1021/ja207385y
    26. 26
      (a) Ishihara, J.; Horie, M.; Shimada, Y.; Tojo, S.; Murai, A. Asymmetric construction of the azaspiro[5.6]dodec-9-ene system in marine natural toxins. Synlett 2002, 2002, 403406,  DOI: 10.1055/s-2002-20451
      (b) Tsuchikawa, H.; Minamino, K.; Hayashi, S.; Murata, M. Efficient Access to the Functionalized Bicyclic Pharmacophore of Spirolide C by Using a Selective Diels–Alder Reaction. Asian J. Org. Chem. 2017, 6, 13221327,  DOI: 10.1002/ajoc.201700164
    27. 27
      Tsujimoto, T.; Murai, A. Efficient Detachment of N-Benzyl Carbamate Group. Synlett 2002, 2002, 1283,  DOI: 10.1055/s-2002-32986
    28. 28
      (a) Barma, D. K.; Bandyopadhyay, A.; Capdevila, J. H.; Falck, J. R. Dimethylthiocarbamate (DMTC): An Alcohol Protecting Group. Org. Lett. 2003, 5, 47554757,  DOI: 10.1021/ol0354573
      (b) Kojima, M.; Nakamura, Y.; Ishikawa, T.; Takeuchi, S. Fluorous dimethylthiocarbamate (FDMTC) protecting groups for alcohols. Tetrahedron Lett. 2006, 47, 63096314,  DOI: 10.1016/j.tetlet.2006.05.142
    29. 29
      (a) Ye, J.; Bhatt, R. K.; Falck, J. R. Stereospecific α-alkoxystannane couplings with acyl chlorides: Total synthesis of (+)-goniofufurone. Tetrahedron Lett. 1993, 34, 80078010,  DOI: 10.1016/S0040-4039(00)61436-3
      (b) Wang, R.; Falck, J. R. Studies towards asymmetric synthesis of 4(S)-11-dihydroxydocosahexaenoic acid (diHDHA) featuring cross-coupling of chiral stannane under mild conditions. Org. Biomol. Chem. 2015, 13, 16241628,  DOI: 10.1039/C4OB02324B
    30. 30
      Huang, S. L.; Omura, K.; Swern, D. Oxidation of Sterically Hindered Alcohols to Carbonyls with Dimethyl Sulfoxide-Trifluor-acetic Anhydride. J. Org. Chem. 1976, 41, 33293331,  DOI: 10.1021/jo00882a030
    31. 31
      Lindlar, H.; Dubuis, R. Palladium Catalyst for Partial Reduction of Acetylenes. Org. Synth. 1966, 46, 8992,  DOI: 10.1002/0471264180.os046.27
    32. 32
      (a) Luche, J.-L. Lanthanides in Organic Synthesis. 1. Selective 1,2 Reduction of Conjugated Ketones. J. Am. Chem. Soc. 1978, 100, 22262227,  DOI: 10.1021/ja00475a040
      (b) Gemal, A. L.; Luche, J.-L. Lanthanoids in Organic Synthesis. 6. The reduction of α-enones by Sodium Borohydride in the Presence of Lanthanoid Chlorides: Mechanistic Aspects. J. Am. Chem. Soc. 1981, 103, 54545459,  DOI: 10.1021/ja00408a029
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.orglett.4c04245.

    • Experimental procedures, characterization data, and NMR spectra of all newly synthesized compounds. (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.