Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Letter
  • Published:

Building C(sp3)-rich complexity by combining cycloaddition and C–C cross-coupling reactions

Abstract

Prized for their ability to rapidly generate chemical complexity by building new ring systems and stereocentres1, cycloaddition reactions have featured in numerous total syntheses2 and are a key component in the education of chemistry students3. Similarly, carbon–carbon (C–C) cross-coupling methods are integral to synthesis because of their programmability, modularity and reliability4. Within the area of drug discovery, an overreliance on cross-coupling has led to a disproportionate representation of flat architectures that are rich in carbon atoms with orbitals hybridized in an sp2 manner5. Despite the ability of cycloadditions to introduce multiple carbon sp3 centres in a single step, they are less used6. This is probably because of their lack of modularity, stemming from the idiosyncratic steric and electronic rules for each specific type of cycloaddition. Here we demonstrate a strategy for combining the optimal features of these two chemical transformations into one simple sequence, to enable the modular, enantioselective, scalable and programmable preparation of useful building blocks, natural products and lead scaffolds for drug discovery.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Combining the logic of cycloaddition and C–C cross-coupling.
Fig. 2: Substrate scope of combining cycloaddition and C–C cross-coupling.
Fig. 3: Applications of combining cycloaddition and C–C cross-coupling.

Similar content being viewed by others

References

  1. Fleming, I. Pericyclic Reactions (Oxford Univ. Press, Oxford, 2015).

    Google Scholar 

  2. Nicolaou, K. C., Snyder, S. A., Montagnon, T. & Vassilikogiannakis, G. The Diels–Alder reaction in total synthesis. Angew. Chem. Int. Ed. 41, 1668–1698 (2002).

    Article  CAS  Google Scholar 

  3. Corey, E. J. & Cheng, X. M. The Logic of Chemical Synthesis (Wiley, New York, 1989).

  4. de Meijere, A., Bräse, S. & Oestreich, M. Metal Catalyzed Cross-Coupling Reactions and More (Wiley-VCH, New York, 2014).

  5. Lovering, F., Bikker, J. & Humblet, C. Escape from Flatland: increasing saturation as an approach to improving clinical success. J. Med. Chem. 52, 6752–6756 (2009).

    Article  PubMed  CAS  Google Scholar 

  6. Brown, D. G. & Boström, J. Analysis of past and present synthetic methodologies on medicinal chemistry: where have all the new reactions gone? J. Med. Chem. 59, 4443–4458 (2016).

    Article  PubMed  CAS  Google Scholar 

  7. Schneider, N., Lowe, D. M., Sayle, R. A., Tarselli, M. A. & Landrum, G. A. Big data from pharmaceutical patents: a computational analysis of medicinal chemists’ bread and butter. J. Med. Chem. 59, 4385–4402 (2016).

    Article  PubMed  CAS  Google Scholar 

  8. Fleming, I. Frontier Orbitals and Organic Chemical Reactions (Wiley, New York, 1991).

  9. Olivo, H. F. & Hemenway, M. S. Recent syntheses of epibatidine. A review. Org. Prep. Proced. Int. 34, 1–25 (2002).

    Article  CAS  Google Scholar 

  10. Carini, D. J. et al. Nonpeptide angiotensin II receptor antagonists: the discovery of a series of N-(biphenylylmethyl)imidazoles as potent, orally active antihypertensives. J. Med. Chem. 34, 2525–2547 (1991).

    Article  PubMed  CAS  Google Scholar 

  11. Dolitzky, B.-Z., Nisnevich, G., Ruchman, I. & Kaftanov, J. Processes for preparing losartan and losartan potassium. Canadian patent CA2482857A1 (2003).

  12. Larsen, R. D. et al. Efficient synthesis of losartan, a nonpeptide angiotensin II receptor antagonist. J. Org. Chem. 59, 6391–6394 (1994).

    Article  CAS  Google Scholar 

  13. Cornella, J. et al. Practical Ni-catalyzed aryl–alkyl cross-coupling of secondary redox-active esters. J. Am. Chem. Soc. 138, 2174–2177 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  14. Wang, J. et al. Nickel-catalyzed cross-coupling of redox-active esters with boronic acids. Angew. Chem. Int. Ed. 55, 9676–9679 (2016).

    Article  CAS  Google Scholar 

  15. Edwards, J. T. et al. Decarboxylative alkenylation. Nature 545, 213–218 (2017).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  16. Smith, J. M. et al. Decarboxylative alkynylation. Angew. Chem. Int. Ed. 56, 11906–11910 (2017).

    Article  CAS  Google Scholar 

  17. Qin, T. et al. A general alkyl-alkyl cross-coupling enabled by redox-active esters and alkylzinc reagents. Science 352, 801–805 (2016).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  18. Chen, Y., Tian, S.-K. & Deng, L. A highly enantioselective catalytic desymmetrization of cyclic anhydrides with modified cinchona alkaloids. J. Am. Chem. Soc. 122, 9542–9543 (2000).

    Article  CAS  Google Scholar 

  19. Padwa, A. & Dent, W. Use of N-[(trimethylsilyl)methyl]amino ethers as capped azomethine ylide equivalents. J. Org. Chem. 52, 235–244 (1987).

    Article  CAS  Google Scholar 

  20. Vitaku, E., Smith, D. T. & Njardarson, J. T. Analysis of the structural diversity, substitution patterns, and frequency of nitrogen heterocycles among U.S. FDA approved pharmaceuticals. J. Med. Chem. 57, 10257–10274 (2014).

    Article  PubMed  CAS  Google Scholar 

  21. Trost, B. M. & Chan, D. M. T. Palladium-mediated cycloaddition approach to cyclopentanoids. Introduction and initial studies. J. Am. Chem. Soc. 105, 2315–2325 (1983).

    Article  CAS  Google Scholar 

  22. Poplata, S., Tröster, A., Zou, Y.-Q. & Bach, T. Recent advances in the synthesis of cyclobutanes by olefin [2 + 2] photocycloaddition reactions. Chem. Rev. 116, 9748–9815 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Fan, Y.-Y., Gao, X.-H. & Yue, J.-M. Attractive natural products with strained cyclopropane and/or cyclobutane ring systems. Sci. China Chem. 59, 1126–1141 (2016).

    Article  CAS  Google Scholar 

  24. Bartoli, G., Bencivenni, G. & Dalpozzo, R. Asymmetric cyclopropanation reactions. Synthesis 46, 979–1029 (2014).

    Article  CAS  Google Scholar 

  25. Committee for Medicinal Products for Human Use (CHMP) assessment report EMA/CHMP/583011/2010 (2010).

  26. Anugu, R. R., Mainkar, P. S., Sridhar, B. & Chandrasekhar, S. The Ireland-Claisen rearrangement strategy towards the synthesis of the schizophrenia drug, (+)-asenapine. Org. Biomol. Chem. 14, 1332–1337 (2016).

    Article  PubMed  CAS  Google Scholar 

  27. Trivedi, M., Budihardjo, I., Loureiro, K., Reid, T. R. & Ma, J. D. Epothilones: a novel class of microtubule-stabilizing drugs for the treatment of cancer. Future Oncol. 4, 483–500 (2008).

    Article  PubMed  CAS  Google Scholar 

  28. Nicolaou, K. C. et al. Chemical synthesis and biological evaluation of cis- and trans-12,13-cyclopropyl and 12,13-cyclobutyl epothilones and related pyridine side chain analogues. J. Am. Chem. Soc. 123, 9313–9323 (2001).

    Article  PubMed  CAS  Google Scholar 

  29. Myers, M. R. et al. Potent quinoxaline-based inhibitors of PDGF receptor tyrosine kinase activity. Part 1: SAR exploration and effective bioisosteric replacement of a phenyl substituent. Bioorg. Med. Chem. Lett. 13, 3091–3095 (2003).

    Article  PubMed  CAS  Google Scholar 

  30. Hatakeyama, T. et al. Iron-catalyzed Suzuki−Miyaura coupling of alkyl halides. J. Am. Chem. Soc. 132, 10674–10676 (2010).

    Article  PubMed  CAS  Google Scholar 

  31. Corey, E. J., Shibata, T. & Lee, T. W. Asymmetric Diels−Alder reactions catalyzed by a triflic acid activated chiral oxazaborolidine. J. Am. Chem. Soc. 124, 3808–3809 (2002).

    Article  PubMed  CAS  Google Scholar 

  32. He, Y. et al. The EED protein–protein interaction inhibitor A-395 inactivates the PRC2 complex. Nat. Chem. Biol. 13, 389–395 (2017); erratum 13, 922 (2017).

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgements

Financial support for this work was provided by Leo Pharma and the US National Institutes of Health (NIH)/National Institute of General Medical Sciences (NIGMS; grant GM-118176). Shenzhen Haiwei M&E Co. Ltd supported a fellowship to T.–G.C.; the Uehara Memorial Foundation supported a research fellowship to S.A.; the Basque Government supported a fellowship to I.B.; Nankai University supported Y.L. and C.B.; the University of Science and Technology of China supported J.T.; and the Swiss National Science Foundation supported an Early Postdoc Mobility Fellowship to D.K. We thank L. Buzzetti for the synthesis of intermediates; D.-H. Huang and L. Pasternack for assistance with nuclear magnetic resonance spectroscopy; and A.L. Rheingold, M. Gembicky and C.E. Moore for X-ray crystallographic analysis.

Author information

Authors and Affiliations

Authors

Contributions

T.–G.C., T.Q. and P.S.B. conceived the work. T.–G.C., L.M.B., Y.L., J.T., D.K., I.B., S.A., C.B., J.S.C., M.S., H.F., F.G.F., H.-W.C., L.H., T.Q. and P.S.B. designed the experiments and analysed the data. T.–G.C., L.M.B., Y.L., J.T., D.K., I.B., S.A. and C.B. performed the experiments. M.S., H.F., F.G.F., H.-W.C. and L.H. performed the experiments described in Fig. 3f. P.S.B. wrote the manuscript. T.–G.C., L.M.B., Y.L., J.T., D.K., I.B., S.A., C.B., J.S.C. and T.Q. assisted in writing and editing the manuscript.

Corresponding author

Correspondence to Phil S. Baran.

Ethics declarations

Competing interests

M.S., H.F., F.G.F., H.-W.C. and L.H. are employees of Eisai Inc. This work was part-funded by Leo Pharma.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Complete substrate scope of [4+2] cycloadditions/cross-couplings.

See Supplementary Information for synthetic details. R1, R2 = (Het)Aryl, alkyl, alkenyl, alkynyl. X-ray structure data are available for compounds 11, 19, 23, 25, 28, 44, 45 and 49.

Extended Data Fig. 2 Complete substrate scope of [3+2], [2+2] and [2+1] sections.

See Supplementary Information for synthetic details. R1, R2 = (Het)Aryl, alkyl, alkenyl, alkynyl. X-ray structure data are available for compounds C2 and 65.

Supplementary information

Supplementary Information

This file contains Supplementary Text and Data – see contents pages for details.

Supplementary Data

This file contains NMR Spectra data.

Supplementary Data

This zipped file contains the crystallographic data files for the compounds used.

Supplementary Data

This zipped file contains the checkcif files for the compounds used.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, T., Barton, L.M., Lin, Y. et al. Building C(sp3)-rich complexity by combining cycloaddition and C–C cross-coupling reactions. Nature 560, 350–354 (2018). https://doi.org/10.1038/s41586-018-0391-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-018-0391-9

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing