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:

SAM-dependent enzyme-catalysed pericyclic reactions in natural product biosynthesis

Abstract

Pericyclic reactions—which proceed in a concerted fashion through a cyclic transition state—are among the most powerful synthetic transformations used to make multiple regioselective and stereoselective carbon–carbon bonds1. They have been widely applied to the synthesis of biologically active complex natural products containing contiguous stereogenic carbon centres2,3,4,5,6. Despite the prominence of pericyclic reactions in total synthesis, only three naturally existing enzymatic examples (the intramolecular Diels–Alder reaction7, and the Cope8 and the Claisen rearrangements9) have been characterized. Here we report a versatile S-adenosyl-l-methionine (SAM)-dependent enzyme, LepI, that can catalyse stereoselective dehydration followed by three pericyclic transformations: intramolecular Diels–Alder and hetero-Diels–Alder reactions via a single ambimodal transition state, and a retro-Claisen rearrangement. Together, these transformations lead to the formation of the dihydropyran core of the fungal natural product, leporin10. Combined in vitro enzymatic characterization and computational studies provide insight into how LepI regulates these bifurcating biosynthetic reaction pathways by using SAM as the cofactor. These pathways converge to the desired biosynthetic end product via the (SAM-dependent) retro-Claisen rearrangement catalysed by LepI. We expect that more pericyclic biosynthetic enzymatic transformations remain to be discovered in naturally occurring enzyme ‘toolboxes’11. The new role of the versatile cofactor SAM is likely to be found in other examples of enzyme catalysis.

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

Access options

Buy this article

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

Figure 1: Enzyme-catalysed pericyclic reactions and the proposed naturally occurring inverse electron demand hetero-Diels-Alder (HDA) reactions.
Figure 2: HPLC analysis showing the reactions catalysed by LepI.
Figure 3: LepI-catalysed reactions are SAM-dependent.
Figure 4: Energetics and transition states of the ambimodal IMDA/HDA, and retro-Claisen pericyclic reactions leading to formation of 2 from 4.

Similar content being viewed by others

References

  1. Hoffmann, R. & Woodward, R. B. The conservation of orbital symmetry. Acc. Chem. Res. 1, 17–22 (1968)

    Article  CAS  Google Scholar 

  2. Takao, K., Munakata, R. & Tadano, K. Recent advances in natural product synthesis by using intramolecular Diels-Alder reactions. Chem. Rev. 105, 4779–4807 (2005)

    Article  CAS  PubMed  Google Scholar 

  3. 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 

  4. Ilardi, E. A., Stivala, C. E. & Zakarian, A. [3,3]-Sigmatropic rearrangements: recent applications in the total synthesis of natural products. Chem. Soc. Rev. 38, 3133–3148 (2009)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Ardkhean, R. et al. Cascade polycyclizations in natural product synthesis. Chem. Soc. Rev. 45, 1557–1569 (2016)

    Article  CAS  PubMed  Google Scholar 

  6. Nicolaou, K. C., Vourloumis, D., Winssinger, N. & Baran, P. S. The art and science of total synthesis at the dawn of the twenty-first century. Angew. Chem. Int. Ed. 39, 44–122 (2000)

    Article  CAS  Google Scholar 

  7. Kim, H. J., Ruszczycky, M. W., Choi, S. H., Liu, Y. N. & Liu, H. W. Enzyme-catalysed [4+2] cycloaddition is a key step in the biosynthesis of spinosyn A. Nature 473, 109–112 (2011)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  8. Li, S. et al. Hapalindole/ambiguine biogenesis is mediated by a Cope rearrangement, C–C bond-forming cascade. J. Am. Chem. Soc. 137, 15366–15369 (2015)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Andrews, P. R., Smith, G. D. & Young, I. G. Transition-state stabilization and enzymic catalysis. Kinetic and molecular orbital studies of the rearrangement of chorismate to prephenate. Biochemistry 12, 3492–3498 (1973)

    Article  CAS  PubMed  Google Scholar 

  10. Cary, J. W. et al. An Aspergillus flavus secondary metabolic gene cluster containing a hybrid PKS-NRPS is necessary for synthesis of the 2-pyridones, leporins. Fungal Genet. Biol. 81, 88–97 (2015)

    Article  CAS  PubMed  Google Scholar 

  11. Walsh, C. T. A chemocentric view of the natural product inventory. Nat. Chem. Biol. 11, 620–624 (2015)

    Article  CAS  PubMed  Google Scholar 

  12. Lin, C.-I., McCarty, R. M. & Liu, H.-W. The enzymology of organic transformations: a survey of name reactions in biological systems. Angew. Chem. Int. Ed. 56, 3446–3489 (2017)

    Article  CAS  Google Scholar 

  13. Minami, A. & Oikawa, H. Recent advances of Diels-Alderases involved in natural product biosynthesis. J. Antibiot. 69, 500–506 (2016)

    Article  CAS  Google Scholar 

  14. Tang, M.-C., Zou, Y., Watanabe, K., Walsh, C. T. & Tang, Y. Oxidative cyclization in natural product biosynthesis. Chem. Rev. 117, 5226–5333 (2017)

    Article  CAS  PubMed  Google Scholar 

  15. Stocking, E. M . & Williams, R. M. Chemistry and biology of biosynthetic Diels-Alder reactions. Angew. Chem. Int. Ed. 42, 3078–3115 (2003)

    Article  CAS  Google Scholar 

  16. Oikawa, H. & Tokiwano, T. Enzymatic catalysis of the Diels-Alder reaction in the biosynthesis of natural products. Nat. Prod. Rep. 21, 321–352 (2004)

    Article  CAS  PubMed  Google Scholar 

  17. Desimoni, G. & Tacconi, G. Heterodiene syntheses with α,β-unsaturated carbonyl compounds. Chem. Rev. 75, 651–692 (1975)

    Article  CAS  Google Scholar 

  18. Jessen, H. J. & Gademann, K. 4-Hydroxy-2-pyridone alkaloids: structures and synthetic approaches. Nat. Prod. Rep. 27, 1168–1185 (2010)

    Article  CAS  PubMed  Google Scholar 

  19. Van De Water, R. W. & Pettus, T. R. R. o-Quinone methides: intermediates underdeveloped and underutilized in organic synthesis. Tetrahedron 58, 5367–5405 (2002)

    Article  CAS  Google Scholar 

  20. Snider, B. B. & Lu, Q. Total synthesis of (±)-leporin A. J. Org. Chem. 61, 2839–2844 (1996)

    Article  CAS  PubMed  Google Scholar 

  21. Li, L. et al. Biochemical characterization of a eukaryotic decalin-forming Diels–Alderase. J. Am. Chem. Soc. 138, 15837–15840 (2016)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Halo, L. M. et al. Late stage oxidations during the biosynthesis of the 2-pyridone tenellin in the entomopathogenic fungus Beauveria bassiana. J. Am. Chem. Soc. 130, 17988–17996 (2008)

    Article  CAS  PubMed  Google Scholar 

  23. Jansson, A. et al. Aclacinomycin 10-hydroxylase is a novel substrate-assisted hydroxylase requiring S-adenosyl-L-methionine as cofactor. J. Biol. Chem. 280, 3636–3644 (2005)

    Article  CAS  PubMed  Google Scholar 

  24. Jiang, C. et al. Formation of the Δ18,19 double bond and bis(spiroacetal) in salinomycin is atypically catalyzed by SlnM, a methyltransferase-like enzyme. Angew. Chem. Int. Ed. 54, 9097–9100 (2015)

    Article  CAS  Google Scholar 

  25. Iwig, D. F. & Booker, S. J. Insight into the polar reactivity of the onium chalcogen analogues of S-adenosyl-L-methionine. Biochemistry 43, 13496–13509 (2004)

    Article  CAS  PubMed  Google Scholar 

  26. Coward, J. K. & Slisz, E. P. Analogs of S-adenosylhomocysteine as potential inhibitors of biological transmethylation. Specificity of the S-adenosylhomocysteine binding site. J. Med. Chem. 16, 460–463 (1973)

    Article  CAS  PubMed  Google Scholar 

  27. Bauer, N. J., Kreuzman, A. J., Dotzlaf, J. E. & Yeh, W. K. Purification, characterization, and kinetic mechanism of S-adenosyl-L-methionine:macrocin O-methyltransferase from Streptomyces fradiae. J. Biol. Chem. 263, 15619–15625 (1988)

    CAS  PubMed  Google Scholar 

  28. Patel, A. et al. Dynamically complex [6+4] and [4+2] cycloadditions in the biosynthesis of spinosyn A. J. Am. Chem. Soc. 138, 3631–3634 (2016)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Ess, D. H. et al. Bifurcations on potential energy surfaces of organic reactions. Angew. Chem. Int. Ed. 47, 7592–7601 (2008)

    Article  CAS  Google Scholar 

  30. Hong, Y. J. & Tantillo, D. J. Biosynthetic consequences of multiple sequential post-transition-state bifurcations. Nat. Chem. 6, 104–111 (2014)

    Article  CAS  PubMed  Google Scholar 

  31. Xu, W., Cai, X., Jung, M. E. & Tang, Y. Analysis of intact and dissected fungal polyketide synthase-nonribosomal peptide synthetase in vitro and in Saccharomyces cerevisiae. J. Am. Chem. Soc. 132, 13604–13607 (2010)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Frisch, M. J. et al. Gaussian 09 Rev. C.01 (Gaussian Inc., 2010)

Download references

Acknowledgements

This work was supported by the NIH (1DP1GM106413 and 1R35GM118056), the NSF (CHE-1361104 to K.N.H.), and the JSPS Program for Advancing Strategic International Networks to Accelerate the Circulation of Talented Researchers (G2604 to K.W.).

Author information

Authors and Affiliations

Authors

Contributions

M.O., F.L., Y.H., K.N.H. and Y.T. developed the hypothesis and designed the study. M.O. performed all in vivo and in vitro experiments, as well as compound isolation and characterization. M.O., Y.H. and M.C. performed protein purification. M.S. and M.-C.T. performed compound characterization. F.L., Z.Y. and K.N.H. performed the computational experiments. All authors analysed and discussed the results. M.O., F.L., K.W., K.N.H. and Y.T. prepared the manuscript.

Corresponding authors

Correspondence to Kenji Watanabe, K. N. Houk or Yi Tang.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

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 Figure 1 LC–MS analysis of the in vitro reaction of 3 with LepF.

The extracted ion chromatograms (EIC) under positive ionization are shown. The mass/charge (m/z) ratio of alcohols 4 and 4′ is 354 under positive ionization. Because the enzymatic activity of LepF is low and 4 is very unstable, we were not able to obtain enough 4 to use it as the substrate for in vitro reaction of LepI. Thus, we obtained 4 by reducing ketone 3 with NaBH4. Since this reduction proceeds non-stereoselectively, 4 and diastereomer 4′ were formed. After the isolation of 4 and 4′ by HPLC, the fractions containing 4 and 4′ were not concentrated and were immediately used as the substrate. The stereochemistry of the secondary alcohol in 4 and 4′ was not determined.

Extended Data Figure 2 HPLC analysis of the chemical reduction of 3 with NaBH4.

The reaction mixture containing 1 mM 3 and 10 mM NaBH4 with EtOH (50 μl) was incubated at 0 °C for 1 min, then the reaction was quenched with water. After centrifugation, the supernatant was analysed by HPLC. The reduction of 3 gave the alcohol 4 and diastereomer 4′. The spontaneous dehydration of both alcohols resulted in the formation of HDA and IMDA products via the E/Z mixture of quinone methide 5. The isolated 4 and 4′ also readily dehydrated and converted to a mixture of the desired HDA (2) and the undesired HDA (9) and IMDA (68) products, showing the instability of these compounds. The structures (right) show the relative stereochemistry.

Extended Data Figure 3 Reaction analysis of 6–9 under heating.

6 (dissolved in 5% DMSO with H2O) was heated at 95 °C for 1 h. 7–9 (dissolved in 5% DMSO with H2O) were heated at 95 °C for 10 h. 6 was completely converted to 2 via [3,3]-sigmatropic retro-Claisen rearrangement. This reaction is irreversible under these conditions. It should be noted that the conversion of 6 to 2 via cycloreversion can be ruled out, since 6 was completely converted to 2 without any other IMDA/HDA side products. No reactions occurred in the case of 7. 8 and 9 can be interconverted via Claisen rearrangement. In this case, retro-Claisen rearrangement (8 to 9) is preferable to forward Claisen rearrangement (9 to 8). The structures show the relative stereochemistry.

Extended Data Figure 4 Analysis of the substrate specificity of LepI.

a, In vitro reactions of other IMDA products 7–9 with 30 μM LepI for 12 h. (i) 8 in buffer, (ii) 8 with LepI, (iii) 7 in buffer, (iv) 7 with LepI, (v) 9 in buffer, (vi) 9 with LepI. The experimental details are described in Methods. b, Elucidation of inhibitory activity of 7 on LepI-catalysed retro-Claisen rearrangement of 6 to 2. The experimental details are described in Methods. The IC50 value is mean ± standard deviation (s.d.) of three independent experiments. The structures show the relative stereochemistry.

Extended Data Figure 5 Time-course analysis of the LepI-catalysed retro-Claisen rearrangement of 6 to 2.

The experimental details are described in Methods. The data show one representative experiment from at least three independent replicates.

Extended Data Figure 6 HPLC analysis showing that purified LepI retains SAM.

SAM was detected in the supernatant of denatured (by acetonitrile) LepI. When LepI was denatured by heating the sample at 95 °C for 10 min, a single peak corresponding to 5′-deoxy-5′-(methylthio)adenosine (MTA), a major degradation product of SAM25, was detected from the supernatant of boiled LepI. Since SAM to MTA conversion is nearly quantitative and an MTA standard curve can be readily constructed, we found that about 90% of LepI still retains SAM after purification. Shown are HPLC profiles of (i) LepI denatured by acetonitrile, (ii) LepI heated at 95 °C for 10 min, (iii) the authentic reference of SAM, (iv) SAM heated at 95 °C for 10 min, and (v) the authentic reference of MTA. The experimental details are described in Methods.

Extended Data Figure 7 HPLC analysis of SAM-dependent LepI-catalysed reactions.

a, Analysis of in vitro reaction of 240 μM 4 with 300 nM LepI at 30 °C for 5 min in the presence and absence of cofactors. The concentrations of SAH, SAM and sinefungin used in this experiment are 250 μM, 100 μM and 100 μM, respectively. The data show one representative experiment from at least three independent replicates. b, Analysis of the in vitro reaction of 140 μM 6 with 300 nM LepI at 30 °C for 4 min in the presence and absence of cofactors. The concentrations of SAH, SAM and sinefungin used in this experiment are 250 μM, 100 μM and 100 μM, respectively. The data show one representative experiment from at least three independent replicates.

Extended Data Figure 8 SAH is a competitive inhibitor of LepI retro-Claisen rearrangement.

a, Dose-dependent inhibition of retro-Claisen rearrangement by SAH. b, Dose-dependent recovery of retro-Claisen rearrangement by SAM in the presence of 250 μM SAH. The experimental details are described in Methods. Error bars, s.d. of three independent experiments.

Extended Data Figure 9 Time-course analysis of the production of 2 divided by the sum of the production of 2 and 6.

The substrate used in this study is alcohol 4. a, LepI-catalysed reaction with or without SAH (250 μM). b, Non-enzymatic reaction. The initial production ratio (IMDA (6) versus HDA products (2)) between LepI-catalysed (about 1:1 periselectivity) and non-catalysed reactions (about 94:6 periselectivity) are clearly different. These data support the suggestion that LepI catalyses the competitive IMDA/HDA reactions by changing the product distribution resulting from IMDA versus HDA reactions.

Extended Data Figure 10 Calculated free energies and bond distances.

Data are shown for the ambimodal transition state (TS-1) and the transition state for the retro-Claisen rearrangement (TS-2), uncatalysed and with various catalysts, calculated with B3LYP-D3/6-311+G(d,p)//6-31G(d), CPCM water. Positions of the bonds are shown in the structures above.

Supplementary information

Supplementary Information

This file contains Supplementary Methods 1-4, Supplementary Tables 1-9, Supplementary Figures 1-32, Supplementary Data and additional references. (PDF 17133 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ohashi, M., Liu, F., Hai, Y. et al. SAM-dependent enzyme-catalysed pericyclic reactions in natural product biosynthesis. Nature 549, 502–506 (2017). https://doi.org/10.1038/nature23882

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature23882

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