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 NmrA-like enzyme-catalysed redox-mediated Diels–Alder cycloaddition with anti-selectivity

Abstract

The Diels–Alder cycloaddition is one of the most powerful approaches in organic synthesis and is often used in the synthesis of important pharmaceuticals. Yet, strictly controlling the stereoselectivity of the Diels–Alder reactions is challenging, and great efforts are needed to construct complex molecules with desired chirality via organocatalysis or transition-metal strategies. Nature has evolved different types of enzymes to exquisitely control cyclization stereochemistry; however, most of the reported Diels–Alderases have been shown to only facilitate the energetically favourable diastereoselective cycloadditions. Here we report the discovery and characterization of CtdP, a member of a new class of bifunctional oxidoreductase/Diels–Alderase, which was previously annotated as an NmrA-like transcriptional regulator. We demonstrate that CtdP catalyses the inherently disfavoured cycloaddition to form the bicyclo[2.2.2]diazaoctane scaffold with a strict α-anti-selectivity. Guided by computational studies, we reveal a NADP+/NADPH-dependent redox mechanism for the CtdP-catalysed inverse electron demand Diels–Alder cycloaddition, which serves as the first example of a bifunctional Diels–Alderase that utilizes this mechanism.

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: Fungal bicyclo[2.2.2]diazaoctane alkaloid biosynthesis and functional characterization of CtdP.
Fig. 2: Crystal structure analysis of CtdP complex.
Fig. 3: CtdP active site view and mutagenesis study.
Fig. 4: Proposed mechanism of CtdP-catalysed α-anti-selective IMDA reaction.

Similar content being viewed by others

Data availability

Data supporting the findings of this work are available within the Article, Extended Data, Source Data and the Supplementary Information files. Data supporting the current study are also available from the corresponding author upon request. The genome sequences of P. citrinum ATCC 9849 are accessible from the GeneBank database (BCKA00000000.1). Coordinates and associated structure factors of CtdP have been deposited in the Protein Data Bank (PDB) database (PDB code: 7UF8). The crystallographic data of small molecules have been deposited at the CCDC (2127333 for 5 and 2127332 for 10). Energies and molecular coordinates of calculated structures are provided in the Supplementary Information file. Source data are provided with this paper.

References

  1. 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 

  2. Houk, K. N., Liu, F., Yang, Z. & Seeman, J. I. Evolution of the Diels–Alder reaction mechanism since the 1930s: Woodward, Houk with Woodward, and the influence of computational chemistry on understanding cycloadditions. Angew. Chem. Int. Ed. 60, 12660–12681 (2021).

    Article  CAS  Google Scholar 

  3. 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 

  4. Gao, L. et al. FAD-dependent enzyme-catalysed intermolecular [4+2] cycloaddition in natural product biosynthesis. Nat. Chem. 12, 620–628 (2020).

    Article  CAS  PubMed  Google Scholar 

  5. Caputi, L. et al. Missing enzymes in the biosynthesis of the anticancer drug vinblastine in Madagascar periwinkle. Science 360, 1235–1239 (2018).

    Article  CAS  PubMed  Google Scholar 

  6. Tian, Z. et al. An enzymatic [4+2] cyclization cascade creates the pentacyclic core of pyrroindomycins. Nat. Chem. Biol. 11, 259–265 (2015).

    Article  CAS  PubMed  Google Scholar 

  7. Tan, D. et al. Genome-mined Diels–Alderase catalyzes 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 

  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. Zhang, Z. et al. Enzyme-catalyzed 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 

  10. Chen, Q. et al. Enzymatic intermolecular hetero-Diels–Alder reaction in the biosynthesis of tropolonic sesquiterpenes. J. Am. Chem. Soc. 141, 14052–14056 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Jamieson, C. S., Ohashi, M., Liu, F., Tang, Y. & Houk, K. N. The expanding world of biosynthetic pericyclases: cooperation of experiment and theory for discovery. Nat. Prod. Rep. 36, 698–713 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Qian-Cutrone, J. et al. Stephacidin A and B: two structurally novel, selective inhibitors of the testosterone-dependent prostate LNCaP cells. J. Am. Chem. Soc. 124, 14556–14557 (2002).

    Article  CAS  PubMed  Google Scholar 

  13. Madariaga-Mazón, A., Hernández-Abreu, O., Estrada-Soto, S. & Mata, R. Insights on the vasorelaxant mode of action of malbrancheamide. J. Pharm. Pharmacol 67, 551–558 (2015).

    Article  PubMed  Google Scholar 

  14. Zinser, E. W. et al. Anthelmintic paraherquamides are cholinergic antagonists in gastrointestinal nematodes and mammals. J. Vet. Pharmacol. Ther. 25, 241–250 (2002).

    Article  CAS  PubMed  Google Scholar 

  15. Paterson, R. R. M., Simmonds, M. S. J. & Blaney, W. M. Mycopesticidal effects of characterized extracts of Penicillium isolates and purified secondary metabolites (including mycotoxins) on Drosophila melanogaster and Spodoptora littoralis. J. Invertebr. Pathol. 50, 124–133 (1987).

    Article  Google Scholar 

  16. Klas, K. R. et al. Structural and stereochemical diversity in prenylated indole alkaloids containing the bicyclo[2.2.2]diazaoctane ring system from marine and terrestrial fungi. Nat. Prod. Rep. 35, 532–558 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Fraley, A. E. & Sherman, D. H. Enzyme evolution in fungal indole alkaloid biosynthesis. FEBS J. 287, 1381–1402 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Domingo, L. R., Zaragozá, R. J. & Williams, R. M. Studies on the biosynthesis of paraherquamide A and VM99955. A theoretical study of intramolecular Diels−Alder cycloaddition. J. Org. Chem. 68, 2895–2902 (2003).

  19. Tsuda, M. et al. Citrinadin A, a novel pentacyclic alkaloid from marine-derived fungus Penicillium citrinum. Org. Lett. 6, 3087–3089 (2004).

    Article  CAS  PubMed  Google Scholar 

  20. Mercado-Marin, E. V. et al. Total synthesis and isolation of citrinalin and cyclopiamine congeners. Nature 509, 318–324 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Zhang, P., Li, X.-M., Liu, H., Li, X. & Wang, B.-G. Two new alkaloids from Penicillium oxalicum EN-201, an endophytic fungus derived from the marine mangrove plant Rhizophora stylosa. Phytochem. Lett. 13, 160–164 (2015).

    Article  CAS  Google Scholar 

  22. Liu, Z. et al. Structural basis of the stereoselective formation of the spirooxindole ring in the biosynthesis of citrinadins. Nat. Commun. 12, 4158 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Goswami, R. S. Targeted gene replacement in fungi using a split-marker approach. Methods Mol. Biol. 835, 255–269 (2012).

    Article  CAS  PubMed  Google Scholar 

  24. Chen, M., Liu, C.-T. & Tang, Y. Discovery and biocatalytic application of a PLP-dependent amino acid γ-substitution enzyme that catalyzes C–C bond formation. J. Am. Chem. Soc. 142, 10506–10515 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Andrianopoulos, A., Kourambas, S., Sharp, J. A., Davis, M. A. & Hynes, M. J. Characterization of the Aspergillus nidulans nmrA gene involved in nitrogen metabolite repression. J. Bacteriol. 180, 1973–1977 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Stammers, D. K. et al. The structure of the negative transcriptional regulator NmrA reveals a structural superfamily which includes the short-chain dehydrogenase/reductases. EMBO J. 20, 6619–6626 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Zang, W. & Zheng, X. Structure and functions of cellular redox sensor HSCARG/NMRAL1, a linkage among redox status, innate immunity, DNA damage response, and cancer. Free Radic. Biol. Med. 160, 768–774 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Zheng, X. et al. Restructuring of the dinucleotide-binding fold in an NADP(H) sensor protein. Proc. Natl Acad. Sci. USA 104, 8809–8814 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Trottmann, F. et al. Pathogenic bacteria remodel central metabolic enzyme to build a cyclopropanol warhead. Nat. Chem. 14, 884–890 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Wu, C. J., Li, C. W., Gao, H., Huang, X. J. & Cui, C. B. Penicimutamides D–E: two new prenylated indole alkaloids from a mutant of the marine-derived Penicillium purpurogenum G59. RSC Adv. 7, 24718–24722 (2017).

    Article  CAS  Google Scholar 

  31. Makino, Y., Negoro, S., Urabe, I. & Okada, H. Stability-increasing mutants of glucose dehydrogenase from Bacillus megaterium IWG3. J. Biol. Chem. 264, 6381–6385 (1989).

    Article  CAS  PubMed  Google Scholar 

  32. Zhang, G. et al. Mechanistic insights into polycycle formation by reductive cyclization in ikarugamycin biosynthesis. Angew. Chem. Int. Ed. 53, 4840–4844 (2014).

    Article  CAS  Google Scholar 

  33. Wen, W.-H. et al. Reductive inactivation of the hemiaminal pharmacophore for resistance against tetrahydroisoquinoline antibiotics. Nat. Commun. 12, 7085 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Zhang, Z., Chen, L., Liu, L., Su, X. & Rabinowitz, J. D. Chemical basis for deuterium labeling of fat and NADPH. J. Am. Chem. Soc. 139, 14368–14371 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Laina-Martín, V., Fernández-Salas, J. A. & Alemán, J. Organocatalytic strategies for the development of the enantioselective inverse-electron-demand hetero-Diels–Alder reaction. Chem. Eur. J. 27, 12509–12520 (2021).

    Article  PubMed  Google Scholar 

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

    Article  Google Scholar 

  37. Sheldrick, G. M. Crystal structure refinement with SHELXL. Acta Crystallogr. C 71, 3–8 (2015).

    Article  Google Scholar 

  38. 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 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Terwilliger, T. C. et al. Decision-making in structure solution using Bayesian estimates of map quality: the PHENIX AutoSol wizard. Acta Crystallogr. D 65, 582–601 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. Features and development of Coot. Acta Crystallogr. D 66, 486–501 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Adams, P. D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D 66, 213–221 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Chen, V. B. et al. MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr. D 66, 12–21 (2010).

    Article  CAS  PubMed  Google Scholar 

  44. Schrödinger Release 2017-2: MacroModel, version 11.2.014 (Schrödinger, 2017).

  45. Frisch, M. J. et al. Gaussian 16 Revision C.01 (Gaussian, 2016).

  46. Zhao, Y. & Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 120, 215–241 (2008).

    Article  CAS  Google Scholar 

  47. Miertuš, S., Scrocco, E. & Tomasi, J. Electrostatic interaction of a solute with a continuum. A direct utilizaion of AB initio molecular potentials for the prevision of solvent effects. Chem. Phys. 55, 117–129 (1981).

    Article  Google Scholar 

  48. Miertus̃, S. & Tomasi, J. Approximate evaluations of the electrostatic free energy and internal energy changes in solution processes. Chem. Phys. 65, 239–245 (1982).

    Article  Google Scholar 

  49. Pascual-ahuir, J. L., Silla, E. & Tuñon, I. GEPOL: an improved description of molecular surfaces. III. A new algorithm for the computation of a solvent-excluding surface. J. Comput. Chem. 15, 1127–1138 (1994).

    Article  CAS  Google Scholar 

  50. Ribeiro, R. F., Marenich, A. V., Cramer, C. J. & Truhlar, D. G. Use of solution-phase vibrational frequencies in continuum models for the free energy of solvation. J. Phys. Chem. 115, 14556–14562 (2011).

    Article  CAS  Google Scholar 

  51. Luchini, G., Alegre-Requena, J., Funes-Ardoiz, I. & Paton, R. GoodVibes: automated thermochemistry for heterogeneous computational chemistry data [version 1; peer review: 2 approved with reservations]. F1000Research https://doi.org/10.12688/f1000research.22758.1 (2020).

  52. Legault, C. Y. CYLview20 (Université de Sherbrooke, 2020).

  53. Salomon-Ferrer, R., Götz, A. W., Poole, D., Le Grand, S. & Walker, R. C. Routine microsecond molecular dynamics simulations with AMBER on GPUs. 2. Explicit solvent particle mesh Ewald. J. Chem. Theory Comput. 9, 3878–3888 (2013).

    Article  CAS  PubMed  Google Scholar 

  54. Case, D. A. et al. AMBER 2016 (University of California, San Francisco, 2016).

  55. Wang, J., Wolf, R. M., Caldwell, J. W., Kollman, P. A. & Case, D. A. Development and testing of a general amber force field. J. Comput. Chem. 25, 1157–1174 (2004).

    Article  CAS  PubMed  Google Scholar 

  56. Bayly, C. I., Cieplak, P., Cornell, W. & Kollman, P. A. A well-behaved electrostatic potential based method using charge restraints for deriving atomic charges: the RESP model. J. Chem. Phys. 97, 10269–10280 (1993).

    Article  CAS  Google Scholar 

  57. Besler, B. H., Merz, K. M. Jr. & Kollman, P. A. Atomic charges derived from semiempirical methods. J. Comput. Chem. 11, 431–439 (1990).

    Article  CAS  Google Scholar 

  58. Singh, U. C. & Kollman, P. A. An approach to computing electrostatic charges for molecules. J. Comput. Chem. 5, 129–145 (1984).

    Article  CAS  Google Scholar 

  59. Jorgensen, W. L., Chandrasekhar, J., Madura, J. D., Impey, R. W. & Klein, M. L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 79, 926–935 (1983).

    Article  CAS  Google Scholar 

  60. Maier, J. A. et al. ff14SB: improving the accuracy of protein side chain and backbone parameters from ff99SB. J. Chem. Theory Comput. 11, 3696–3713 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Darden, T., York, D. & Pedersen, L. Particle mesh Ewald: an Nlog(N) method for Ewald sums in large systems. J. Chem. Phys. 98, 10089–10092 (1993).

    Article  CAS  Google Scholar 

  62. Shaw, D. E. et al. Anton 2: raising the bar for performance and programmability in a special-purpose molecular dynamics supercomputer. In SC ’14: Proceedings of the International Conference for High Performance Computing, Networking, Storage and Analysis (ed. Dongarra, J.) 41–53 (IEEE Press, 2014).

  63. Li, H. et al. Asperversiamides, linearly fused prenylated indole alkaloids from the marine-derived fungus Aspergillus versicolor. J. Org. Chem. 83, 8483–8492 (2018).

    Article  CAS  PubMed  Google Scholar 

  64. Kagiyama, I. et al. Taichunamides: prenylated indole alkaloids from Aspergillus taichungensis (IBT 19404). Angew. Chem. Int. Ed. 55, 1128–1132 (2016).

    Article  CAS  Google Scholar 

  65. Fraley, A. E. et al. Molecular basis for spirocycle formation in the paraherquamide biosynthetic pathway. J. Am. Chem. Soc. 142, 2244–2252 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Fraley, A. E. et al. Flavin-dependent monooxygenases NotI and NotI′ mediate spiro-oxindole formation in biosynthesis of the notoamides. ChemBioChem 21, 2449–2454 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Sievers, F. et al. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol. Syst. Biol. 7, 539 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  68. Robert, X. & Gouet, P. Deciphering key features in protein structures with the new ENDscript server. Nucleic Acids Res. 42, W320–W324 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Gerlt, J. A. et al. Enzyme function initiative-enzyme similarity tool (EFI-EST): a web tool for generating protein sequence similarity networks. Biochim. Biophys. Acta 1854, 1019–1037 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Shannon, P. et al. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 13, 2498–2504 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank L. Kürti and Y.-D. Kwon for their assistance with the high-resolution mass spectrometry experiments, J. D. Hartgerink for sharing the instruments for electronic circular dichroism measurements, C. Ajo-Franklin and S. Li for sharing the anaerobic workstation, and J. Smith for providing generous resources for protein crystallization, APS beamline access, and software for structure determination; and finally the 23-IDB@GM/CA beamline staff. This work was supported by a National Institute of Health (NIH) grant (R35GM138207) and the Robert A. Welch Foundation (C-1952) to X.G., an NIH grant (R35GM11810) and the Hans W. Vahlteich Professorship to D.H.S., and an NIH grant (AI141481) to K.N.H. J.N.S. acknowledges the support of the National Institute of General Medical Sciences (NIGMS) of the NIH under an F32 individual postdoctoral fellowship (F32GM122218). S.R. acknowledges support from the Michigan Chemistry–Biology Interface Training Program, funded by the NIGMS, NIH T32 Training Grant (5T32GM132046). Computational resources for DFT computations were provided by the UCLA Institute for Digital Research and Education (IDRE) and by the San Diego Supercomputing Center (SDSC) through XSEDE (ACI-1548562). Microsecond molecular dynamics simulations were performed using Anton 2 computer time provided by the Pittsburgh Supercomputing Center (PSC) through grant R01GM116961 from the NIH. The Anton 2 machine at PSC was generously made available by D.E. Shaw Research.

Author information

Authors and Affiliations

Authors

Contributions

Z.L. carried out compound purification, synthesis and structural determination experiments, and in vitro assays and mutagenesis. Z.L., F.Z., H.-W.S. and S.P. performed protein expression and purification. Z.L., F.Z., Q.N. and S.L. performed in vivo gene deletion experiments. Z.L. and S.L. did the CtdP refolding experiments. Z.L. and Q.N. performed (S)-[4-2H] NADPH assays. J.D.F. and G.N.P. assisted in the single-crystal X-ray diffraction analysis of small molecules. S.R. and S.A.N. carried out the protein crystallization and crystal structure determinations. Z.L., S.R., S.A.N., W.X. and M.D.M. analysed the protein crystal structure. J.N.S. performed the DFT computations and MD simulations. Z.L., S.R., S.A.N., J.N.S., K.N.H., D.H.S. and X.G. analysed all the data and prepared the manuscript. All authors revised the manuscript.

Corresponding authors

Correspondence to K. N. Houk, David H. Sherman or Xue Gao.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Chemistry thanks Andrew Gulick, Shina Kamerlin 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 Representations of different classes of Diels–Alderases and the proposed biosynthetic pathway of 21R-citrinadin A.

a, Chemical structures of representative natural products synthesized by different types of Diels–Alderases. b, Chemical structures of natural PIAs with α-anti bicyclo[2.2.2]diazaoctane rings63,64. c, The proposed biosynthetic pathway of 21R-citrinadin A (1)24. PT: prenyltransferase, DAase: Diels–Alderase.

Extended Data Fig. 2 Comparison analysis of the biosynthetic gene clusters of PIAs and in vivo characterization of NmrA-like proteins CtdO, CtdR, and CtdP.

a, Annotation of the biosynthetic gene clusters of PIAs. 21R-citrinadin A (ctd in P. citrinum ATCC 9849), citrinadin A (cnd in P. citrinum DSM1997)24, paraherquamides (phq in P. fellutanum ATCC20841)65, (−)-notoamide A (not in A. protuberus MF297-2)66, (+)-notoamide A (not’ in A. versicolor NRRL35600)66 and malbrancheamide (mal in Malbranchea aurantiaca RRC1813)8. NRPS: non-ribosomal peptide synthetase; FPMO: flavoprotein monooxygenase; P450: cytochrome P450; SDR: short-chain dehydrogenase/reductase. SDRs MalC and PhqE, and NmrA-like oxidoreductase CtdP are characterized as Diels–Alderases. b, LC–MS analysis (extracted ion chromatogram, EIC) of P. citrinum wild-type (WT) and mutant ΔctdN. Four ΔctdN mutants all retained the production of compound 1. c, LC–MS analysis (EIC) of P. citrinum WT and mutants ΔctdO, ΔctdR, ΔctdN, and ΔctdP. The ΔctdR mutant major produces R1, which is identified as an α-anti product by NMR (Supplementary Table 10) and ECD (Supplementary Fig. 6b) analysis, indicating that CtdR might work on the biosynthetic steps after Diels–Alder cycloaddition. The ΔctdO mutant showed reduced production of citrinadin compounds, which possibly serves as a regulator. Both ΔctdR and ΔctdO produce trace amounts of 5 and 6, indicating the deletions of these genes have some influence and result in accumulations of the shunt products. Only the ΔctdP mutant could abolish the production of 1 and 2, as well as have a remarkable accumulation of 5 and 6.

Extended Data Fig. 3 In vitro assays of 3 in anaerobic and aerobic conditions.

a, EIC traces of in vitro assays of 3 with CtdP, CtdN, and MalC in an anaerobic chamber, respectively. b, Enzymatic and spontaneous transforms of 3. c, Ultraviolet (UV) spectrum of 12. d, X-ray structure of 5. e, Spontaneous reactions of 3. f, In vitro assays of 3 with CtdO and CtdR in aerobic conditions, respectively. The symbol * represents the compound identified by MS and UV spectra (Supplementary Figs. 4, 5).

Extended Data Fig. 4 Sequence alignment of CtdP and homologous NmrA-like proteins.

CtdP, HSCARG from Homo sapiens (PDB: 2exx), and NmrA-like family domain-containing protein 1 from H. sapiens (PDB: 2wm3) are aligned using the Clustal Omega server67. Secondary structural elements of CtdP are marked above the alignment. Residues are coloured based on their conservation using the ESPript 3.0 server68. The η symbol refers to a 310-helice.

Extended Data Fig. 5 Superposition of the different chains in CtdP.

Superposition of the different chains and calculated r.m.s.d. values. Chains A-D are coloured forest green, cyan, wheat, and warm pink, respectively. NADP+ (yellow) and penicimutamide E (10, salmon) are shown as sticks.

Extended Data Fig. 6 C-terminal tail and electrostatic surface structures of CtdP.

a, Interactions of C-terminal tail (orange) in CtdP chain A. NADP+ (yellow) and penicimutamide E (10, salmon) shown as sticks. Distances in angstroms shown in red. b, Electrostatic surface potential of CtdP-NADP+-10 complex (chain A, calculated by Adaptive Pisson-Boltzmann Sovler plugin, https://www.poissonboltzmann.org/).

Extended Data Fig. 7 Density functional theory calculations for tautomerization and redox mechanisms.

a, The redox-mediated pathway of CtdP catalysis in (S)-[4-2H] NADPH assay. b, Study of the inherent stereoselectivity of the IMDA reaction via the redox-mediated pathway by comparing the four diastereomeric transition states in the absence of the CtdP enzyme. c, DFT computations comparing the redox properties of CtdP substrate 3 and MalC substrate 9. The free energy change for the reaction above (∆G = +4.5 kcal mol–1) corresponds to CtdP substrate 3, with a 6-membered methylpiperidine ring, being over 1000 times harder to oxidize than the corresponding substrate 9-red with a 5-membered pyrrolidine ring.

Extended Data Fig. 8 MD simulations of substrate 3-ox in CtdP active site.

a, Tracking the distance between the carbon atom that gets oxidized to form 3-ox and the hydride transferred to form NADPH during three 1.2-microsecond MD simulations. b, The hydrogen bond between substrate 3-ox and NADPH in MD simulations. The amide N-H of substrate 3-ox is the H-bond donor and the ribose 2′-OH of NADPH is the H-bond acceptor. The plot illustrates the N-O distance as a function of time over three 1.2-microsecond MD simulations. c, MD snapshot of CtdP active site in which substrate 3-ox has an α-anti conformation and the shortest DA bond lengths.

Extended Data Fig. 9 Sequence similarity network analysis of NmrA-like protein CtdP.

Sequence similarity network of CtdP generated by the EFI-EST69 server at an E-value of 5 and alignment score of 35, 2000 sequences from UniProt (2022-02) database were used, filtered by 50% representative nodes, visualized in Cytoscape v3.9.170.

Supplementary Information

Supplementary Information

Supplementary Tables 1–14, Figs. 1–17, and energies and molecular coordinates of calculated structures.

Reporting Summary

Supplementary Data 1

Crystallographic data of the protein and small molecules.

Supplementary Data 2

Cif file with crystallographic data for compound 5.

Supplementary Data 3

Cif file with crystallographic data for compound 10.

Source data

Source Data Fig. 1

Source data of Fig. 3c and Supplementary Fig. 7a,d.

Source Data Fig. 2

Uncropped and unprocessed scan of SDS–PAGE images.

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

Liu, Z., Rivera, S., Newmister, S.A. et al. An NmrA-like enzyme-catalysed redox-mediated Diels–Alder cycloaddition with anti-selectivity. Nat. Chem. 15, 526–534 (2023). https://doi.org/10.1038/s41557-022-01117-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41557-022-01117-6

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