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.

  • Article
  • Published:

Catalytic mechanism and endo-to-exo selectivity reversion of an octalin-forming natural Diels–Alderase

Abstract

We have previously reported the identification of CghA, a proposed Diels–Alderase responsible for the formation of the bicyclic octalin core of the fungal secondary metabolite Sch210972. Here we show the crystal structure of the CghA–product complex at a resolution of 2.0 Å. Our result provides the second structural determination of eukaryotic Diels–Alderases and adds yet another fold to the family of proteins reported to catalyse [4 + 2] cycloaddition reactions. Site-directed mutagenesis-coupled kinetic characterization and computational analyses allowed us to identify key catalytic residues and propose a possible catalytic mechanism. Most interestingly, we were able to rationally engineer CghA such that the mutant was able to catalyse preferentially the formation of the energetically disfavoured exo adduct. This work expands our knowledge and understanding of the emerging and potentially widespread class of natural enzymes capable of catalysing stereoselective Diels–Alder reactions and paves the way towards developing enzymes potentially useful in various bio/synthetic applications.

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

Fig. 1: Octalin-containing fungal secondary metabolites and proposed involvement of Diels–Alder reactions in their biosynthesis.
Fig. 2: Crystal structure of CghA in complex with 1.
Fig. 3: In vitro assays of the CghA mutants.
Fig. 4: Computational analyses of the transition states and energetics of the octalin-forming intramolecular Diels–Alder reactions.

Similar content being viewed by others

Data availability

The crystal structure data are available on the PDB (accession code: 6KAW for the apo CghA structure, 6KBC for the CghA–1 complex structure). The coordinates of calculated structures are available online at https://doi.org/10.19061/iochem-bd-6-67. The GenBank accession numbers of the amino acid sequences of the enzymes referenced in this study are provided in this paper. The source data underlying Supplementary Fig. 3a are provided as a Source Data file. Source data are provided with this paper. All other data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  1. Kim, H. J., Ruszczycky, M. W. & Liu, H. W. Current developments and challenges in the search for a naturally selected Diels–Alderase. Curr. Opin. Chem. Biol. 16, 124–131 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Klas, K., Tsukamoto, S., Sherman, D. H. & Williams, R. M. Natural Diels–Alderases: elusive and irresistable. J. Org. Chem. 80, 11672–11685 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  Google Scholar 

  4. Sato, M. et al. Involvement of lipocalin-like CghA in decalin-forming stereoselective intramolecular [4 + 2] cycloaddition. ChemBioChem 16, 2294–2298 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Yang, S. W. et al. Chemokine receptor CCR-5 inhibitors produced by Chaetomium globosum. J. Nat. Prod. 69, 1025–1028 (2006).

    Article  CAS  PubMed  Google Scholar 

  6. Kakule, T. B. et al. Native promoter strategy for high-yielding synthesis and engineering of fungal secondary metabolites. ACS Synth. Biol. 4, 625–633 (2015).

    Article  CAS  PubMed  Google Scholar 

  7. Kato, N. et al. A new enzyme involved in the control of the stereochemistry in the decalin formation during equisetin biosynthesis. Biochem. Biophys. Res. Commun. 460, 210–215 (2015).

    Article  CAS  PubMed  Google Scholar 

  8. 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 

  9. Kato, N. et al. Control of the stereochemical course of [4 + 2] cycloaddition during trans-decalin formation by Fsa2-family enzymes. Angew. Chem. Int. Ed. Engl. 57, 9754–9758 (2018).

    Article  CAS  PubMed  Google Scholar 

  10. Tan, D. et al. Genome-mined Diels–Alderase catalyses formation of the cis-octahydrodecalins of varicidin A and B. J. Am. Chem. Soc. 141, 769–773 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Zhang, Z. et al. Enzyme-catalysed inverse-electron demand Diels–Alder reaction in the biosynthesis of antifungal ilicicolin H. J. Am. Chem. Soc. 141, 5659–5663 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Hantke, V., Skellam, E. J. & Cox, R. J. Evidence for enzyme catalysed intramolecular [4 + 2] Diels–Alder cyclization during the biosynthesis of pyrichalasin H. Chem. Commun. 56, 2925–2928 (2020).

    Article  CAS  Google Scholar 

  13. Little, R. et al. Unexpected enzyme-catalysed [4 + 2] cycloaddition and rearrangement in polyether antibiotic biosynthesis. Nat. Catal. 2, 1045–1054 (2019).

    Article  CAS  Google Scholar 

  14. 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 

  15. Chiu, H. J. et al. Structure of the first representative of Pfam family PF09410 (DUF2006) reveals a structural signature of the calycin superfamily that suggests a role in lipid metabolism. Acta Crystallogr. F 66, 1153–1159 (2010).

    Article  CAS  Google Scholar 

  16. Zheng, Q. et al. Enzyme-dependent [4 + 2] cycloaddition depends on lid-like interaction of the N-terminal sequence with the catalytic core in PyrI4. Cell Chem. Biol. 23, 352–360 (2016).

    Article  CAS  PubMed  Google Scholar 

  17. Byrne, M. J. et al. The catalytic mechanism of a natural Diels–Alderase revealed in molecular detail. J. Am. Chem. Soc. 138, 6095–6098 (2016).

    Article  CAS  PubMed  Google Scholar 

  18. Fage, C. D. et al. The structure of SpnF, a standalone enzyme that catalyses [4 + 2] cycloaddition. Nat. Chem. Biol. 11, 256–258 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Drulyte, I. et al. Crystal structure of the putative cyclase IdmH from the indanomycin nonribosomal peptide synthase/polyketide synthase. IUCrJ 6, 1120–1133 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Lakshmi, B., Mishra, M., Srinivasan, N. & Archunan, G. Structure-based phylogenetic analysis of the lipocalin superfamily. PLoS One 10, e0135507 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  21. Zheng, Y. & Thiel, W. Computational Insights into an enzyme-catalysed [4 + 2] cycloaddition. J. Org. Chem. 82, 13563–13571 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Hugot, M. et al. A structural basis for the activity of retro-Diels–Alder catalytic antibodies: evidence for a catalytic aromatic residue. Proc. Natl Acad. Sci. USA 99, 9674–9678 (2002).

    Article  CAS  PubMed  Google Scholar 

  23. Ishiuchi, K. et al. Combinatorial generation of complexity by redox enzymes in the chaetoglobosin A biosynthesis. J. Am. Chem. Soc. 135, 7371–7377 (2013).

    Article  CAS  PubMed  Google Scholar 

  24. Qiao, K., Chooi, Y. H. & Tang, Y. Identification and engineering of the cytochalasin gene cluster from Aspergillus clavatus NRRL 1. Metab. Eng. 13, 723–732 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Heine, A. et al. An antibody exo Diels–Alderase inhibitor complex at 1.95 angstrom resolution. Science 279, 1934–1940 (1998).

    Article  CAS  PubMed  Google Scholar 

  26. Xu, J. et al. Evolution of shape complementarity and catalytic efficiency from a primordial antibody template. Science 286, 2345–2348 (1999).

    Article  CAS  PubMed  Google Scholar 

  27. Cannizzaro, C. E., Ashley, J. A., Janda, K. D. & Houk, K. N. Experimental determination of the absolute enantioselectivity of an antibody-catalysed Diels–Alder reaction and theoretical explorations of the origins of stereoselectivity. J. Am. Chem. Soc. 125, 2489–2506 (2003).

    Article  CAS  PubMed  Google Scholar 

  28. Kasahara, K. et al. Solanapyrone synthase, a possible Diels–Alderase and iterative type I polyketide synthase encoded in a biosynthetic gene cluster from Alternaria solani. ChemBioChem 11, 1245–1252 (2010).

    Article  CAS  PubMed  Google Scholar 

  29. Gottlieb, H. E., Kotlyar, V. & Nudelman, A. NMR chemical shifts of common laboratory solvents as trace impurities. J. Org. Chem. 62, 7512–7515 (1997).

    Article  CAS  PubMed  Google Scholar 

  30. Kabsch, W. XDS. Acta Crystallogr. D 66, 125–132 (2010).

    Article  CAS  Google Scholar 

  31. Schrödinger. Release 2017-2: MacroModel v.11.2.014 (Schrödinger, LLC, 2017).

  32. Frisch, M. J. et al. Gaussian 16 Rev. A.03 (Gaussian Inc., 2016).

  33. Krishnan, R., Binkley, J. S., Seeger, R. & Pople, J. A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 72, 650–654 (1980).

    Article  CAS  Google Scholar 

  34. Rassolov, V. A., Pople, J. A., Ratner, M. A. & Windus, T. L. 6-31G* basis set for atoms K through Zn. J. Chem. Phys. 109, 1223–1229 (1998).

    Article  CAS  Google Scholar 

  35. Francl, M. M. et al. Self-consistent molecular orbital methods. XXIII. A polarization-type basis set for second-row elements. J. Chem. Phys. 77, 3654–3665 (1982).

    Article  CAS  Google Scholar 

  36. Dill, J. D. & Pople, J. A. Self-consistent molecular orbital methods. XV. Extended Gaussian-type basis sets for lithium, beryllium, and boron. J. Chem. Phys. 62, 2921–2923 (1975).

    Article  CAS  Google Scholar 

  37. Hehre, W. J., Ditchfield, R. & Pople, J. A. Self-consistent molecular orbital methods. XII. Further extensions of gaussian-type basis sets for use in molecular orbital studies of organic molecules. J. Chem. Phys. 56, 2257–2261 (1972).

    Article  CAS  Google Scholar 

  38. Chai, J. D. & Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections. Phys. Chem. Chem. Phys. 10, 6615–6620 (2008).

    Article  CAS  PubMed  Google Scholar 

  39. Linder, M. & Brinck, T. On the method-dependence of transition state asynchronicity in Diels–Alder reactions. Phys. Chem. Chem. Phys. 15, 5108–5114 (2013).

    Article  CAS  PubMed  Google Scholar 

  40. Mardirossian, N. & Head-Gordon, M. Thirty years of density functional theory in computational chemistry: an overview and extensive assessment of 200 density functionals. Mol. Phys. 115, 2315–2372 (2017).

    Article  CAS  Google Scholar 

  41. Clark, T., Chandrasekhar, J., Spitznagel, G. W. & Schleyer, P. V. R. Efficient diffuse function-augmented basis sets for anion calculations. III. The 3-21 + G basis set for first-row elements, Li–F. J. Comput. Chem. 4, 294–301 (1983).

    Article  CAS  Google Scholar 

  42. Blaudeau, J.-P., McGrath, M. P., Curtiss, L. A. & Radom, L. Extension of Gaussian-2 (G2) theory to molecules containing third-row atoms K and Ca. J. Chem. Phys. 107, 5016–5021 (1997).

    Article  CAS  Google Scholar 

  43. Schäfer, A., Huber, C. & Ahlrichs, R. Fully optimized contracted Gaussian basis sets of triple zeta valence quality for atoms Li to Kr. J. Chem. Phys. 100, 5829–5835 (1994).

    Article  Google Scholar 

  44. Weigend, F. & Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: design and assessment of accuracy. Phys. Chem. Chem. Phys. 7, 3297–3305 (2005).

    Article  CAS  PubMed  Google Scholar 

  45. Grimme, S. Improved second-order Møller–Plesset perturbation theory by separate scaling of parallel- and antiparallel-spin pair correlation energies. J. Chem. Phys. 118, 9095–9102 (2003).

    Article  CAS  Google Scholar 

  46. Dunning, T. H. J. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 90, 1007–1023 (1989).

    Article  CAS  Google Scholar 

  47. Weigend, F., Köhn, A. & Hättig, C. Efficient use of the correlation consistent basis sets in resolution of the identity MP2 calculations. J. Chem. Phys. 116, 3175–3183 (2002).

    Article  CAS  Google Scholar 

  48. Marenich, A. V., Cramer, C. J. & Truhlar, D. G. Universal solvation model based on solute electron density and on a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B 113, 6378–6396 (2009).

    Article  CAS  PubMed  Google Scholar 

  49. Neese, F. The ORCA program system. WIREs Comput. Mol. Sci. 2, 73–78 (2012).

    Article  CAS  Google Scholar 

  50. Neese, F. Software update: the ORCA program system, version 4.0. WIREs Comput. Mol. Sci. 8, e1327 (2018).

    Article  Google Scholar 

  51. Paton, L., Rodríguez-Guerra, J., Chen, J. & Funes-Ardoiz, I. bobbypaton/GoodVibes: GoodVibes v3.0.0 (Zenodo, 2019).

  52. Grimme, S. Supramolecular binding thermodynamics by dispersion-corrected density functional theory. Chem. Eur. J. 18, 9955–9964 (2012).

    Article  CAS  PubMed  Google Scholar 

  53. Li, Y.-P., Gomes, J., Mallikarjun Sharada, S., Bell, A. T. & Head-Gordon, M. Improved force-field parameters for QM/MM simulations of the energies of adsorption for molecules in zeolites and a free rotor correction to the rigid rotor harmonic oscillator model for adsorption enthalpies. J. Phys. Chem. C 119, 1840–1850 (2015).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank T. Kan at the University of Shizuoka for providing us with important advice on the chemical synthesis of the substrate analogues. This work was supported by the NIH (GM118056 to Y. Tang and GM124480 to K.N.H.). This work was also supported in part by the Japan Society for the Promotion of Science (JSPS) (K.W., 15KT0068, 26560450 and 19KK0150; M.S., 19K07136) and Innovative Areas from MEXT, Japan (K.W., 16H06449).

Author information

Authors and Affiliations

Authors

Contributions

M.S., H.H. and K.W. conceived and designed the study. S.K. and K.N. chemically synthesized substrate analogues. Y. Tsunematsu and K.W. designed and performed molecular cloning. M.Y. performed the heterologous expression and purification as well as in vitro characterization of the enzymes. N.M. prepared the crystal of the CghA–1 complex. M.Y. and Y. Tsunematsu prepared and analysed SeMet-substituted CghA. M.S. and K.N.H. elucidated the chemical structures. C.S.J. and K.N.H. performed the computational analysis. K.H. and H.H. performed the crystallographic studies and structural analysis. Y. Tang and K.W. analysed sequence and structure comparison data. All authors analysed and discussed the results. K.W. and K.N.H. prepared the manuscript.

Corresponding author

Correspondence to Kenji Watanabe.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Catalysis thanks Marc van der Kamp and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Extended data

Extended Data Fig. 1 Classification of DAases by the mechanism of product release from the NRPS.

a, Proposed two distinct mechanisms of substrate release from the NRPS. One type of the embedded terminal reductase domain of an NRPS catalyses reduction and Knoevenagel condensation of the thioester intermediate to give a pyrrolin-2-one product via a rote A. Another type catalyses a Dieckmann condensation of the intermediate to give a pyrrolidine-2,4-dione product via a rote B. b, Phylogenetic tree of the lipocalin-type DAase homologs. Phylogenetic analysis of the amino acid sequences of the representative homologs found in the protein database revealed that they could be divided clearly into two groups based on the types of tetramic acid moieties described above that the corresponding products bear. Pink represents enzymes predicted to catalyse Knoevenagel condensation to release the pyrrolin-2-one-type (type A) product. Purple represents enzymes predicted to catalyse Dieckmann condensation to release the pyrrolidine-2,4-dione-type (type B) product. The length of the black bar is scaled to represent the evolutionary distance of 0.1 amino acid substitution per site. The enzyme names are abbreviated for the ease of representation.

Extended Data Fig. 2 Electron density of the bound ligand in the CghA co-crystal structure.

The FoFc map contoured at 3.0σ represented as thin blue mesh is shown to indicate the electron density of Sch210972 bound within the active site pocket of CghA. Colouring scheme for the ligand is the same as in Fig. 2a.

Extended Data Fig. 3 Activity assays on the wild-type CghA and its mutants to examine the roles of the active-site residues.

The in vitro analyses were performed using the substrate analogue 8 for the formation of the endo 9 and exo 10 adducts. Detailed reaction conditions are described in Methods unless otherwise specified below. a, HPLC profiles of the reaction mixtures containing 8 as a substrate and (i) the S65N mutant; (ii) the N82A mutant; (iii) the H94A mutant; (iv) the K352A mutant; (v) the N364A mutant; (vi) the wild-type CghA; (vii) buffer, as a background reaction showing the conversion of 8 to 9 and 10 after 10 minutes of incubation; (viii) the authentic reference of 9; and (ix) the authentic reference of 10. The reactions were monitored at 288 nm. N-Boc-L-tryptophan methyl ester was used as an internal standard (IS) throughout the study. See Methods for details. Peaks denoted by asterisks are impurities from the chemical synthesis of 8. b, Hydrogen-bonding interaction between the bound ligand P-1 and the S65 and N82 side chain groups. Two water molecules also link the S65 and N82 side chain groups through hydrogen bonds.

Extended Data Fig. 4 Background conversion of the substrate analogue 8.

Spontaneous transformation of the substrate analogue 8 into the endo 9 and exo 10 adducts occurs almost entirely in the HPLC solvent supplemented with formic acid. 8 (150 µM) was incubated for 0 (labelled “buffer 0 min” in the plot), 10 (buffer 10 min) and 20 min (buffer 20 min) at 25 °C without CghA in the reaction buffer (100 mM potassium phosphate, 100 mM NaCl, pH 8.4) in a total reaction volume of 25 µl. After incubation, the reaction was quenched with 25 µl of MeOH and then centrifuged for removal of debris. Subsequently, the sample was subjected to LC–MS analysis. For the control experiment “buffer 10 min + CghA 10 min”, 8 (150 µM) was incubated for 10 min at 25 °C without CghA in the reaction buffer to allow the spontaneous transformation of 8. Subsequently, 0.1 µM of the wild-type CghA was added to the reaction mixture and the mixture was incubated for another 10 min at 25 °C in a total reaction volume of 25 µl to allow CghA to react on the remaining 8. For other control experiments “CghA 10 min”, “CghA + 10 10 min” and “CghA + 9”, 150 µM of 8, 10 and 9 was added to the reaction mixture, respectively. These mixtures were incubated for 10 min at 25 °C with CghA (0.1 µM wild-type CghA) under the same conditions and analysed as described above.

Extended Data Fig. 5 Total turnover rates of the wild-type CghA and its mutants designed for stereospecificity conversion.

The combined apparent turnover rates kcat(app) for the formation of 9 and 10 from 8 by the CghA mutants that were designed to examine the role of octalin-packing residues in determining the diastereoselectivity of the adducts formed are presented. The centre value is the mean of triplicate measurements with each data point shown as an open circle and the error bar representing the standard deviation. See Methods and Supplementary Note 5 for details.

Extended Data Fig. 6 Activity assays on CghA and its mutants to examine endo-to-exo stereospecificity conversion.

The in vitro analyses were performed using the substrate analogue 8 as the substrate. Detailed reaction conditions are described in Methods unless otherwise specified below. The reactions were monitored at 288 nm. N-Boc-L-tryptophan methyl ester was used as an internal standard (IS) throughout the study. See Methods for details. Peaks denoted by asterisks are impurities from the chemical synthesis of 8. a, HPLC profiles of the reaction mixtures containing 8 as a substrate and (i) buffer only as a background reaction; (ii) the wild-type CghA; (iii) the A242S mutant; (iv) the M257V mutant; (v) the V391L mutant; (vi) the A242S/M257V mutant; (vii) the A242S/M257V/V391L mutant; (viii) the authentic reference of 9; and (ix) the authentic reference of 10. b, HPLC profiles of the reaction mixtures containing 8 as a substrate and (i) buffer, as a background reaction; (ii) the wild-type CghA; (iii) the A242S/M257V/V391L mutant; (iv) the A242N/M257V/V391L mutant; (v) the authentic reference of 9; and (vi) the authentic reference of 10.

Extended Data Fig. 7 The kinetic assay to examine product inhibition of CghA.

In vitro analyses of CghA were performed with 8 as the substrate in the presence of an increasing concentration of the reaction product (the endo adduct 9) to examine the possible product inhibition of CghA. The activity at each endo product concentration is given relative to the activity in the absence of the product. The centre value is the mean of triplicate measurements with the error bar representing the standard deviation. See Methods for details.

Supplementary information

Supplementary Information

Supplementary Methods, Supplementary Tables 1–8, Figs. 1–58, Notes 1–7 and references.

Reporting Summary.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sato, M., Kishimoto, S., Yokoyama, M. et al. Catalytic mechanism and endo-to-exo selectivity reversion of an octalin-forming natural Diels–Alderase. Nat Catal 4, 223–232 (2021). https://doi.org/10.1038/s41929-021-00577-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41929-021-00577-2

This article is cited by

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