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:

An enzymatic dual-oxa Diels–Alder reaction constructs the oxygen-bridged tricyclic acetal unit of (–)-anthrabenzoxocinone

Abstract

The hetero-Diels–Alder (HDA) reaction is a key method for synthesizing six-membered heterocyclic rings in natural products and bioactive compounds. Despite its importance in synthetic chemistry, naturally occurring enzymatic HDA reactions are rare and limited to a single heteroatom. Here we report Abx(−)F, a bifunctional vicinal oxygen chelate (VOC)-like protein that catalyses dehydration and dual-oxa Diels–Alder reactions to stereoselectively form the oxygen-bridged tricyclic acetal of (–)-anthrabenzoxocinone ((−)-ABX). Isotope assays and density functional theory calculations reveal a dehydration-coordinated, concerted HDA mechanism. The crystal structure of Abx(−)F and NMR complex structures of Abx(−)F with its substrate analogue and (−)-ABX define the reaction’s structural basis. Mutational analysis identifies Asp17 as a general base that mediates dehydration, forming an o-quinone methide intermediate for stereoselective dual-oxa HDA. This work establishes the molecular and structural basis of a polyheteroatomic Diels–Alderase, paving the way for designing polyheteroatomic DA enzymatic tools.

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: An overall schematic view of the oxa DA reaction and natural products containing oxygen-bridged tricyclic acetal.
Fig. 2: (−)-ABX biosynthetic gene cluster and key biosynthetic steps.
Fig. 3: Structure of the methylated derivatives of 2 and functional characterization of Abx(−)E and Abx(−)F.
Fig. 4: The calculated reaction energy profile shows the plausible reactions of 2 towards (±)-ABX.
Fig. 5: Structure and site-directed mutagenesis characterization of Abx(−)F.
Fig. 6: Proposed catalytic mechanism of Abx(−)F.

Similar content being viewed by others

Data availability

The crystal structure of Abx(−)F is available in the PDB under accession no. of 9JT3 (revised structure with 2.00 Å resolution and 30.1 Wilson B-factor) and 8STB (original structure with 2.22 Å resolution and 40.45 Wilson B-factor). The NMR structures of Abx(−)F and Abx(−)F in complex with (−)-ABX are available in the PDB under accession nos. 8EO9 and 8EPY, respectively. The complete chemical-shift assignments for Abx(−)F and Abx(−)F in the presence of (−)-ABX are available from BioMagResBank under BMRB ID 31044 and 31046, respectively. Crystallographic data for compound 5 have been deposited at the Cambridge Crystallographic Data Centre, under deposition no. CCDC 2308930. Source data are provided with this paper.

References

  1. Eschenbrenner-Lux, V., Kumar, K. & Waldmann, H. The asymmetric hetero-Diels–Alder reaction in the syntheses of biologically relevant compounds. Angew. Chem. Int. Ed. 53, 11146–11157 (2014).

    Article  CAS  Google Scholar 

  2. Ishihara, K. & Sakakura, A. in Comprehensive Organic Synthesis II, 2nd edn, Vol. 5 (eds Molander, G. & Knochel, A. P.) 409–465 (Elsevier, 2014).

  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 (2022).

    Article  Google Scholar 

  4. Caplan, S. M. & Floreancig, P. E. Total synthesis of divergolides E and H. Angew. Chem. Int. Ed. 57, 15866–15870 (2018).

    Article  CAS  Google Scholar 

  5. Takao, K.-I. et al. Total synthesis of (+)-cytosporolide A via a biomimetic hetero-Diels–Alder reaction. J. Am. Chem. Soc. 137, 15971–15977 (2015).

    Article  CAS  PubMed  Google Scholar 

  6. Kim, H. J. et al. Enzyme-catalysed [4 + 2] cycloaddition is a key step in the biosynthesis of spinosyn A. Nature 473, 109–112 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Cogan, D. P. et al. Structural insights into enzymatic [4 + 2] aza-cycloaddition in thiopeptide antibiotic biosynthesis. Proc. Natl Acad. Sci. USA 114, 12928–12933 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Dan, Q. et al. Fungal indole alkaloid biogenesis through evolution of a bifunctional reductase/Diels–Alderase. Nat. Chem. 11, 972–980 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Liu, Z. et al. An NmrA-like enzyme-catalysed redox-mediated Diels–Alder cycloaddition with anti-selectivity. Nat. Chem. 15, 526–534 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Ohashi, M. et al. SAM-dependent enzyme-catalysed pericyclic reactions in natural product biosynthesis. Nature 549, 502–506 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Little, R. F., Samborskyy, M. & Leadlay, P. F. The biosynthetic pathway to tetromadurin (SF2487/A80577), a polyether tetronate antibiotic. PLoS One 15, e0239054 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Ohashi, M. et al. An enzymatic Alder-ene reaction. Nature 586, 64–69 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Liu, J. et al. Tandem intermolecular [4 + 2] cycloadditions are catalysed by glycosylated enzymes for natural product biosynthesis. Nat. Chem. 15, 1083–1090 (2023).

    Article  CAS  PubMed  Google Scholar 

  15. Schotte, C., Li, L., Wibberg, D., Kalinowski, J. & Cox, R. J. Synthetic biology driven biosynthesis of unnatural tropolone sesquiterpenoids. Angew. Chem. Int. Ed. 59, 23870–23878 (2020).

    Article  CAS  Google Scholar 

  16. Basler, S. et al. Efficient Lewis acid catalysis of an abiological reaction in a de novo protein scaffold. Nat. Chem. 13, 231–235 (2021).

    Article  CAS  PubMed  Google Scholar 

  17. Blond, G., Gulea, M. & Mamane, V. Recent contributions to hetero Diels–Alder reactions. Curr. Org. Chem. 20, 2161–2210 (2016).

    Article  CAS  Google Scholar 

  18. Nicolaou, K. C. et al. Total synthesis and structural elucidation of azaspiracid-1. Final assignment and total synthesis of the correct structure of azaspiracid-1. J. Am. Chem. Soc. 128, 2859–2872 (2006).

    Article  CAS  PubMed  Google Scholar 

  19. Singh, S. B. et al. Integrastatins: structure and HIV-1 integrase inhibitory activities of two novel racemic tetracyclic aromatic heterocycles produced by two fungal species. Tetrahedron Lett. 43, 2351–2354 (2002).

    Article  CAS  Google Scholar 

  20. Sakuno, E., Yabe, K. & Nakajima, H. Involvement of two cytosolic enzymes and a novel intermediate, 5′-oxoaverantin, in the pathway from 5′-hydroxyaverantin to averufin in aflatoxin biosynthesis. Appl. Environ. Microbiol. 69, 6418–6426 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Ding, L. et al. Divergolides|A-D from a mangrove endophyte reveal an unparalleled plasticity in ansa-macrolide biosynthesis. Angew. Chem. Int. Ed. 50, 1630–1634 (2011).

    Article  CAS  Google Scholar 

  22. Tan, H., Chen, X., Chen, H., Liu, H. & Qiu, S. Proline-catalyzed knoevenagel condensation/[4 + 2] cycloaddition cascade reaction: application to formal synthesis of averufin. Eur. J. Org. Chem. 2015, 4956–4963 (2015).

    Article  CAS  Google Scholar 

  23. More, A. A. & Ramana, C. V. Total synthesis of integrastatin B enabled by a benzofuran oxidative dearomatization cascade. Org. Lett. 18, 1458–1461 (2016).

    Article  CAS  PubMed  Google Scholar 

  24. Herath, K. B. et al. Anthrabenzoxocinones from Streptomyces sp. as liver X receptor ligands and antibacterial agents. J. Nat. Prod. 68, 1437–1440 (2005).

    Article  CAS  PubMed  Google Scholar 

  25. Huang, J. K. & Shia, K. S. Development of a cross‐conjugated vinylogous [4 + 2] anionic annulation and application to the total synthesis of natural antibiotic (±)‐ABX. Angew. Chem. Int. Ed. 59, 6540–6545 (2020).

    Article  CAS  Google Scholar 

  26. Jiang, D. et al. Total synthesis of three families of natural antibiotics: anthrabenzoxocinones, fasamycins/naphthacemycins and benastatins. CCS Chem. 2, 800–812 (2020).

    Article  CAS  Google Scholar 

  27. Mei, X. et al. Expanding the bioactive chemical space of anthrabenzoxocinones through engineering the highly promiscuous biosynthetic modification steps. ACS Chem. Biol. 13, 200–206 (2017).

    Article  PubMed  Google Scholar 

  28. Jiang, K. et al. An unusual aromatase/cyclase programs the formation of the phenyldimethylanthrone framework in anthrabenzoxocinones and fasamycin. Proc. Natl Acad. Sci. USA 121, e2321722121 (2024).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Chen, H. et al. Isolation of an anthrabenzoxocinone 1.264-C from Streptomyces sp. FXJ1. 264 and absolute configuration determination of the anthrabenzoxocinones. Tetrahedron Asymmetry 25, 113–116 (2014).

    Article  Google Scholar 

  30. Lam, Y. K. et al. L-755,805, a new polyketide endothelin binding inhibitor from an actinomycete. Tetrahedron Lett. 36, 2013–2016 (1995).

    Article  CAS  Google Scholar 

  31. Wang, W. et al. An engineered strong promoter for Streptomycetes. Appl. Environ. Microbiol. 79, 4484–4492 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Takahashi, S. et al. Reveromycin A biosynthesis uses RevG and RevJ for stereospecific spiroacetal formation. Nat. Chem. Biol. 7, 461–468 (2011).

    Article  CAS  PubMed  Google Scholar 

  33. Sun, P. et al. Spiroketal formation and modification in avermectin biosynthesis involves a dual activity of AveC. J. Am. Chem. Soc. 135, 1540–1548 (2013).

    Article  CAS  PubMed  Google Scholar 

  34. Bilyk, O. et al. Enzyme-catalyzed spiroacetal formation in polyketide antibiotic biosynthesis. J. Am. Chem. Soc. 144, 14555–14563 (2022).

    Article  CAS  PubMed  Google Scholar 

  35. Frank, B. et al. Spiroketal polyketide formation in sorangium: identification and analysis of the biosynthetic gene cluster for the highly cytotoxic spirangienes. Chem. Biol. 14, 221–233 (2007).

    Article  CAS  PubMed  Google Scholar 

  36. Hotta, K. et al. Enzymatic catalysis of anti-Baldwin ring closure in polyether biosynthesis. Nature 483, 355–358 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Jaggavarapu, S. R. et al. Facile access to novel chromeno-2,6,9-trioxabicyclo[3.3.1] nonadienes via tandem nucleophilic substitution and [4 + 2] hetero Diels–Alder reaction: experimental and theoretical study. Tetrahedron 69, 2142–2149 (2013).

    Article  CAS  Google Scholar 

  38. Fan, J. et al. Peniphenone and penilactone formation in Penicillium crustosum via 1,4-Michael additions of ortho-quinone methide from hydroxyclavatol to γ-butyrolactones from crustosic acid. J. Am. Chem. Soc. 141, 4225–4229 (2019).

    Article  CAS  PubMed  Google Scholar 

  39. Willis, N. J. & Bray, C. D. Ortho-quinone methides in natural product synthesis. Chem. Eur. J. 18, 9160–9173 (2012).

    Article  CAS  PubMed  Google Scholar 

  40. Tantillo, D. J., Chen, J. & Houk, K. N. Theozymes and compuzymes: theoretical models for biological catalysis. Curr. Op. Chem. Biol. 2, 743–750 (1998).

    Article  CAS  Google Scholar 

  41. Buchko, G. W. et al. Structural and biophysical characterization of the Mycobacterium tuberculosis protein Rv0577, a protein associated with neutral red staining of virulent tuberculosis strains and homologue of the Streptomyces coelicolor protein KbpA. Biochemistry 56, 4015–4027 (2017).

    Article  CAS  PubMed  Google Scholar 

  42. He, P. & Moran, G. R. Structural and mechanistic comparisons of the metal-binding members of the vicinal oxygen chelate (VOC) superfamily. J. Inorg. Biochem. 105, 1259–1272 (2011).

    Article  CAS  PubMed  Google Scholar 

  43. Wang, Y. S. et al. Molecular basis for the final oxidative rearrangement steps in chartreusin biosynthesis. J. Am. Chem. Soc. 140, 10909–10914 (2018).

    Article  CAS  PubMed  Google Scholar 

  44. van Zundert, G. et al. The HADDOCK2.2 web server: user-friendly integrative modeling of biomolecular complexes. J. Mol. Biol. 428, 720–725 (2016).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Vreven, T. et al. Combining quantum mechanics methods with molecular mechanics methods in ONIOM. J. Chem. Theory Comput. 2, 815–826 (2006).

    Article  CAS  PubMed  Google Scholar 

  47. Zhang, Z., Qiao, T., Watanabe, K. & Tang, Y. Concise biosynthesis of phenylfuropyridones in fungi. Angew. Chem. Int. Ed. 59, 19889–19893 (2020).

    Article  CAS  Google Scholar 

  48. Doyon, T. J. et al. Chemoenzymatic o-quinone methide formation. J. Am. Chem. Soc. 141, 20269–20277 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Purdy, T. N., Moore, B. S. & Lukowski, A. L. Harnessing ortho-quinone methides in natural product biosynthesis and biocatalysis. J. Nat. Prod. 85, 688–701 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Hertweck, C. The biosynthetic logic of polyketide diversity. Angew. Chem. Int. Ed. 48, 4688–4716 (2009).

    Article  CAS  Google Scholar 

  51. Xie, S. & Zhang, L. Type II polyketide synthases: a bioinformatics‐driven approach. ChemBioChem 24, e202200775 (2023).

    Article  CAS  PubMed  Google Scholar 

  52. Kieser, T., Bibb, M. J., Buttner, M. J., Chater, K. F. & Hopwood, D. A. Practical Streptomyces Genetics (John Innes Foundation, 2000).

  53. Fu, J. et al. Full-length RecE enhances linear–linear homologous recombination and facilitates direct cloning for bioprospecting. Nat. Biotechnol. 30, 440–446 (2012).

    Article  CAS  PubMed  Google Scholar 

  54. Gildea, R. J. et al. Xia2.multiplex: a multi-crystal data-analysis pipeline. Acta Crystallogr. D Struct. Biol. 78, 752–769 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Winn, M. D. et al. Overview of the CCP4 suite and current developments. Acta Crystallogr. D Biol. Crystallogr. 67, 235–242 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Liebschner, D. et al. Macromolecular structure determination using X-rays, neutrons and electrons: recent developments in Phenix. Acta Crystallogr. D Struct. Biol. 75, 861–877 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Emsley, P. et al. Features and development of Coot. Acta Crystallogr. D Biol. Crystallogr. 66, 486–501 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Mobli, M., Maciejewski, M. W., Gryk, M. R. & Hoch, J. C. An automated tool for maximum entropy reconstruction of biomolecular NMR spectra. Nat. Methods 4, 467–468 (2007).

    Article  CAS  PubMed  Google Scholar 

  59. Vranken, W. F. et al. The CCPN data model for NMR spectroscopy: development of a software pipeline. Proteins 59, 687–696 (2005).

    Article  CAS  PubMed  Google Scholar 

  60. Shen, Y. & Bax, A. Protein backbone and sidechain torsion angles predicted from NMR chemical shifts using artificial neural networks. J. Biomol. NMR 56, 227–241 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Güntert, P. & Buchner, L. Combined automated NOE assignment and structure calculation with CYANA. J. Biomol. NMR 62, 453–471 (2015).

    Article  PubMed  Google Scholar 

  62. Malde, A. K. et al. An automated force field topology builder (ATB) and repository: version 1.0. J. Chem. Theory Comput. 7, 4026–4037 (2011).

    Article  CAS  PubMed  Google Scholar 

  63. Stroet, M. et al. Automated topology builder version 3.0: prediction of solvation free enthalpies in water and hexane. J. Chem. Theory Comput. 14, 5834–5845 (2018).

    Article  CAS  PubMed  Google Scholar 

  64. Brunger, A. T. et al. Crystallography & NMR system: a new software suite for macromolecular structure determination. Acta Crystallogr. D Biol. Crystallogr. D54, 905–921 (1998).

    Article  CAS  Google Scholar 

  65. Williams, C. J. et al. MolProbity: more and better reference data for improved all-atom structure validation. Protein Sci. 27, 293–315 (2018).

    Article  CAS  PubMed  Google Scholar 

  66. Bannwarth, C. et al. Extended tight-binding quantum chemistry methods. Wiley Interdiscip. Rev. Comput. Mol. Sci. 11, e1493 (2021).

    Article  CAS  Google Scholar 

  67. Bannwarth, C. et al. GFN2-xTB-an accurate and broadly parametrized self-consistent tight-binding quantum chemical method with multipole electrostatics and density-dependent dispersion contributions. J. Chem. Theory Comput. 15, 1652–1671 (2019).

    Article  CAS  PubMed  Google Scholar 

  68. Kollman, P. A. et al. Calculating structures and free energies of complex molecules: combining molecular mechanics and continuum models. Acc. Chem. Res. 33, 889–897 (2000).

    Article  CAS  PubMed  Google Scholar 

  69. Dolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. & Puschmann, H. OLEX2: a complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 42, 339–341 (2009).

    Article  CAS  Google Scholar 

  70. Sheldrick, G. M. SHELXT–Integrated space-group and crystal-structure determination. Acta Crystallogr. A 71, 3–8 (2015).

    Article  Google Scholar 

  71. Tirado-Rives, J. & Jorgensen, W. L. Performance of B3LYP density functional methods for a large set of organic molecules. J. Chem. Theory Comput. 4, 297–306 (2008).

    Article  CAS  PubMed  Google Scholar 

  72. Weigend, F., Häser, M., Patzelt, H. & Ahlrichs, R. RI-MP2: optimized auxiliary basis sets and demonstration of efficiency. Chem. Phys. Lett. 294, 143–152 (1998).

    Article  CAS  Google Scholar 

  73. Frisch, M. J. Gaussian 92, Revision E. 3 (Gaussian, 1992).

  74. Grimme, S., Antony, J., Ehrlich, S. & Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 132, 154104 (2010).

    Article  PubMed  Google Scholar 

  75. 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 

  76. Pracht, P., Bohle, F. & Grimme, S. Automated exploration of the low-energy chemical space with fast quantum chemical methods. Phys. Chem. Chem. Phys. 22, 7169–7192 (2020).

    Article  CAS  PubMed  Google Scholar 

  77. Grimme, S. Exploration of chemical compound, conformer and reaction space with meta-dynamics simulations based on tight-binding quantum chemical calculations. J. Chem. Theory Comput. 15, 2847–2862 (2019).

    Article  CAS  PubMed  Google Scholar 

  78. Maeda, S., Harabuchi, Y., Ono, Y., Taketsugu, T. & Morokuma, K. Intrinsic reaction coordinate: calculation, bifurcation and automated search. Int. J. Quantum Chem. 115, 258–269 (2015).

    Article  CAS  Google Scholar 

  79. Tao, P. & Schlegel, H. B. A toolkit to assist ONIOM calculations. J. Comput. Chem. 31, 2363–2369 (2010).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported in part by the NSFC (32425033 to X.Q., 32300044 to X.Y., 31970041 to Y.-L.Z.), the National Key R&D Program of China (2018YFA0900400 to X.Q., 2020YFA0907700 to Y.-L.Z.), the Program of Shanghai Academic/Technology Research Leader (22XD1421300 to X.Q.), the Shanghai Municipal Science and Technology Major Project (to X.Q.), the Natural Science Foundation of Shanghai (23ZR1432800 to X.Q.), the University of Queensland (UQ Postdoctoral Research Fellowship to X.J., UQ Principal Research Fellowship to M.M.), and the Australian Research Council (ARC Laureate Fellowship FL180100109 to B.K.). We acknowledge the facilities, and the scientific and technical assistance of UQ-ROCX and the Australian Microscopy & Microanalysis Research Facilities at the Centre for Microscopy and Microanalysis, the University of Queensland; as well as the Macromolecular Crystallography (MX) beamlines at the Australian Synchrotron, Melbourne, Victoria, Australia and the Shanghai Synchrotron Radiation Facility BL10U2 beamline. We also thank S. Singh from Merck for kindly gifting the (–)-ABX producing strain S. sp. MA6657, W. Xie for critical discussions, D. Ericsson at Australian Synchrotron for his assistance with screening a potential anomalous signal from metal ions when collecting the diffraction data of Abx(-)F crystals, the Queensland NMR network and the Instrumental Analysis Center of Shanghai Jiao Tong University for access to the 600-, 700- and 900-MHz NMR spectrometer equipped with cryoprobes. We thank X. Bo at the National University of Singapore for his help and discussions on phenix refinement.

Author information

Authors and Affiliations

Authors

Contributions

X.Y., X.J., Y.-L.Z., M.M., Z.D. and X.Q. conceived this project. X.Y. performed fermentation, compound isolation, characterization, compound syntheses and in vitro experiments. H.Z. and M.Y. performed in vivo experiments. Z. Luo, M.-J.Z., X.J., X.-D.K. and B.K. solved the crystal structure of Abx(−)F. X.J. prepared labelled protein samples, analysed NMR spectra, performed NMR assignments and determined solution structures. M.M. acquired and processed NMR spectra using non-uniform sampling schemes. S.J. and Y.-L.Z. conducted the computational studies. J.O. performed water refinements of NMR structures. J.O. and M.M. provided advice and help in solving structures by NMR. K.J. and Z. Lin provided advice and help for in vitro experiments. X.Y., X.J., Y.-L.Z. and X.Q. prepared the paper. All authors contributed to paper editing.

Corresponding authors

Correspondence to Yi-Lei Zhao, Mehdi Mobli or Xudong Qu.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Chemistry thanks Garry Buchko, Marcio Dias and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 Cofactor preference of Abx(−)E and characteristics of products of Abx(−)E.

a, Cofactor preference of Abx(−)E. (i) standard (+)-ABX; (ii) standard (−)-ABX; (iii) 100 μM 1 in 80 μM Abx(−)E; (iv) 100 μM 1 in 80 μM Abx(−)E and 1 mM NADPH. (v) 100 μM 1 in 80 μM Abx(−)E and 1 mM NADH; (vi) 100 μM 1 in 80 μM Abx(−)E, 5 μM GDH, 20 mM glucose. There was no addition of NADPH to the reaction system, and the activity detected implies that the endogenous cofactor (*NADP+) was present in Abx(−)E. b, Time-course conversion of 100 μM 1 in the presence of 80 μM Abx(−)E which was analyzed by normal UPLC (λ = 365 nm). The 8 multiplier mean the heights of the tiny peaks corresponding to compound 5 are magnified eight times for a better visibility. c, UV spectra and mass of compound 2, 5 and (±)-ABX.

Extended Data Fig. 2 Kinetic analysis of Abx(−)E and Abx(−)F, and time-course analysis of 2 in the presence of Abx(−)E and Abx(−)F.

a, Michaelis-Menten kinetic analysis of Abx(−)E with 1. Data were presented as mean ± s.d. from triplicate independent experiments (n = 3). b, Michaelis-Menten kinetic analysis of Abx(−)F with 2. Data were presented as mean ± s.d. from triplicate independent experiments (n = 3). c, Time-course analysis of 100 μM 2 in the presence of Abx(−)E which was analyzed by Chiral UPLC (λ = 365 nm). d, Time-course of 100 μM 2 in the presence of Abx(−)F which was analyzed by Chiral UPLC (λ = 365 nm).

Source data

Extended Data Fig. 3 Characteristics of the enantiomers (S)-5 and (R)-5.

a, Analysis (λ = 400 nm) of the enantiomers (S)-5 and (R)-5 using HPLC equipped with a chiral column (OD-H, 5 μm, 4.6 × 250 mm). (i) The mixture of (S)-5 and (R)-5 isolated from the fermentation extract of the ∆abx()F mutant strain; (ii) The mixture of (S)-5 and (R)-5 from the crystal. b, The measured CD and calculated ECD spectra of (S)-5 and (R)-5. c, Structures of (S)-5 and (R)-5.

Extended Data Fig. 4 Proposed dehydrative cyclization model of Abx(−)F.

a, Proposed reaction diagram of dehydrative cyclization. b, MS spectra of 18O labeled or unlabeled substrate and products in Abx(−)E and Abx(−)F assay. (i) MS spectra of 18O-1 in boiled Abx(−)E and boiled Abx(−)F assay; (ii) MS spectra of 1 in boiled Abx(−)E and boiled Abx(−)F assay; (iii) MS spectra of products 18O-2, 18O-5, 18O-(+)-ABX and 18O-(−)-ABX in Abx(−)E and boiled Abx(−)F assay; (iv) MS spectra of products 2, (+)-ABX and (−)-ABX in Abx(−)E and boiled Abx(−)F assay; (v) MS spectra of product 18O-(−)-ABX in Abx(−)E and Abx(−)F assay; (vi) MS spectra of product (−)-ABX in Abx(−)E and Abx(−)F assay.

Extended Data Fig. 5 Proposed SN1 and SN2 model of Abx(−)F and the DFT calculation.

a, Proposed reaction diagram of SN1 and SN2. b, The high strain in the 8-membered ring prevents the C24 hydroxyl group from approaching the C8 atom. (i) The ground state of the 8-membered ring intermediate upon hemiketalization, in which the C24 hydroxyl group points outward; (ii) The local minimal with the shortest C8–O(C24) distance, the conformation of which is higher by 8.3 kcal mol−1 at the B3LYP(D3)/def2tzvp (SMD) level of theory. c, The energy will further increase to as high as 83.1 kcal mol−1 when the distance decreases to 1.8 Å. Thus, an SN2 mechanism is unlikely to take place. Moreover, there is no specific reason to form high-energy carbon cation species through an SN1 mechanism as well.

Extended Data Fig. 6 Proposed annulation of phenol and the dual-oxa DA reaction model of Abx(−)F.

a, Proposed reaction diagram of annulation of phenol and the dual-oxa DA reaction. Intermediate 3 either undergo a disrotatory oxa-6π electrocyclization or a 1,4-Michael addition to form compound 5. b, MS spectra of 2 and (−)-ABX in Abx(−)F assay which buffer configured with D2O and H2O. (i) MS spectra of 2 in boiled Abx(−)F and D2O buffer; (ii) MS spectra of (−)-ABX in Abx(−)F and D2O buffer; (iii) MS spectra of 2 in boiled Abx(−)F and H2O buffer; (iv) MS spectra of (−)-ABX in Abx(−)F and H2O buffer.

Extended Data Fig. 7 The quantum chemical calculations support and explain the innate stereospecific dual-oxa Diels-Alder cyclization of Abx(−)F.

a, Transition states for the formation of (Z, Ra)-3 in both the non-enzymatic (TS1 and TS2) and enzymatic pathways (TS0). In TS0, the assumed base is displayed as faint sticks behind the solid TS0. b, Comparison of energy profiles between the non-enzymatic and enzymatic pathway. The enzymatic pathway was calculated using the theozyme model. c, Four plausible orientations of the oxa Diels-Alder reactions between the o-QM and the ketone group. d, Transition states and e, energy profiles for the oxa-DA reaction of both E- and Z- o-QM, along with enthalpies and Gibbs free energies (in parentheses).

Extended Data Fig. 8 The calculated reaction energy profile shows the plausible reactions of 2 toward 5.

The Gibbs free energies, given in kcal mol−1, were used to quantify the thermodynamics of these processes at the B3LYP(D3)/def2-TZVP(SMD) level of theory. The energy profile of the two potential dehydration pathways involved in the conversion of 2 to enantiomeric intermediates 3. Following this, besides undergoing oxo-Diels-Alder cyclization to yield the products (−)-ABX and (+)-ABX, the compound 3 can undergo disrotatory oxa-6π electrocyclization to form less stable compounds 5 after hydrogen abstraction from the acetyl methylene substituent. Alternatively, compound 3 form an enolate ion under basic condition and go 1,4-Michael addition forming the anionic form of compound 5 (blue dotted box).

Extended Data Fig. 9 The critical steps of forming (±)-ABX, (R/S)-5 and 6.

‘spon.’ represent spontaneous reaction whose reaction pathways are indicated by grey dotted arrows; reactions catalyzed by Abx(−)E and Abx(−)F are indicated by blank and red arrows; chiral axle is highlighted by orange.

Extended Data Fig. 10 Abx(−)F structure and molecular dynamics simulations.

a, Crystal structure of Abx(−)F. the dashed line represents the amino acids whose electron densities are missing. b, Superimposed structures of crystal Abx(−)F (chain A with purple, chain B with lightblue) and solution NMR Abx(−)F (pink). For Abx(−)F_chain A and solution NMR Abx(−)F, RMSD is 1.370 Å; For Abx(−)F_chain B and solution NMR Abx(−)F, RMSD is 1.398 Å; For Abx(−)F_chain A and Abx(−)F_chain B, the RMSD is 0.522 Å. The missing loop regions in the crystal Abx(−)F structure are highlight in green in NMR structure. c, Superimposed structures of crystal Abx(−)F (purple), MtRv0577 (green, PDB: 3OXH) and the BphC enzyme (light pink, PDB: 1DHY). The backbone RMSD values were 1.168 Å between Abx(−)F and MtRv0577 and 10.261 Å between Abx(−)F and BphC. In BphC enzyme, Fe ion is coordinated by residues H145, H209 and E260. However, no residues in MtRv0577 or Abx(−)F were appropriate for binding a metal ion. d, Ribbon representation of Abx(−)F without (−)-ABX (pink) and Abx(−)F (purple) with (−)-ABX (cyan), including the active residues (W15, D17, L19, Y48, S60, N67, Y78, I105 and W125, pink represent residues of Abx(−)F without (−)-ABX; red represent residues of Abx(−)F with (−)-ABX) displayed as sticks. Regions that undergo apparent conformational changes are highlighted as a non-transparent cartoon against a transparent background. The blue spheres identify the backbone amides with significant chemical-shift perturbations in the presence of (−)-ABX (Supplementary Fig. 19). The conformational changes are manifested by three labelled distances: 11.8 Å between the Cα atoms of Gly44, 8.5 Å between the Cα atoms of Asn67, and 4.6 Å between the Cα atoms of Ile105. e, Molecular dynamics simulations of hydrogen bond between D17 and 6-OH group of 1.

Supplementary information

Supplementary Information

Supplementary Methods 1.1–1.4, Tables 1–11, Figs. 1–20, HADDOCK NOE restraints, computational analysis and references.

Reporting Summary

Supplementary Data 1

Source data for graphs in Supplementary Figs. 14b and 18.

Supplementary Data 2

Crystallographic data of compound 5.

Source data

Source Data Fig. 5

Source data for graph in Fig. 5d.

Source Data Extended Data Fig./Table 2

Source data for graph in Extended Data Fig. 2a and 2b

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yan, X., Jia, X., Luo, Z. et al. An enzymatic dual-oxa Diels–Alder reaction constructs the oxygen-bridged tricyclic acetal unit of (–)-anthrabenzoxocinone. Nat. Chem. 17, 1058–1066 (2025). https://doi.org/10.1038/s41557-025-01804-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue date:

  • DOI: https://doi.org/10.1038/s41557-025-01804-0

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