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:

Molecular mechanism underlying substrate recognition of the peptide macrocyclase PsnB

Abstract

Graspetides, also known as ω-ester-containing peptides (OEPs), are a family of ribosomally synthesized and post-translationally modified peptides (RiPPs) bearing side chain-to-side chain macrolactone or macrolactam linkages. Here, we present the molecular details of precursor peptide recognition by the macrocyclase enzyme PsnB in the biosynthesis of plesiocin, a group 2 graspetide. Biochemical analysis revealed that, in contrast to other RiPPs, the core region of the plesiocin precursor peptide noticeably enhanced the enzyme–precursor interaction via the conserved glutamate residues. We obtained four crystal structures of symmetric or asymmetric PsnB dimers, including those with a bound core peptide and a nucleotide, and suggest that the highly conserved Arg213 at the enzyme active site specifically recognizes a ring-forming acidic residue before phosphorylation. Collectively, this study provides insights into the mechanism underlying substrate recognition in graspetide biosynthesis and lays a foundation for engineering new variants.

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: Introduction of plesiocin and its minimal precursor peptide.
Fig. 2: Roles of LPs and CPs in enzyme behavior.
Fig. 3: ADP and AMPPNP enhance the affinity of the CP to PsnB.
Fig. 4: Crystal structures of PsnB reveal asymmetric dimers with bound nucleotide and CP.
Fig. 5: Molecular interactions between PsnB and its precursor peptide substrate.
Fig. 6: Conformational changes in the PsnB dimer during substrate binding.

Similar content being viewed by others

Data availability

Coordinates and structure factors for the reported crystal structures in this work were deposited in the RCSB PDB under accession numbers 7DRM (MP- and ADP-bound PsnB), 7DRN (MP- and AMPPNP-bound PsnB), 7DRP (MP(pE37)- and ADP-bound PsnB) and 7DRO (MP-bound PsnB). Source data are provided with this paper.

References

  1. Newman, D. J. & Cragg, G. M. Natural products as sources of new drugs from 1981 to 2014. J. Nat. Prod. 79, 629–661 (2016).

    Article  CAS  PubMed  Google Scholar 

  2. Arnison, P. G. et al. Ribosomally synthesized and post-translationally modified peptide natural products: overview and recommendations for a universal nomenclature. Nat. Prod. Rep. 30, 108–160 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Cimermancic, P. et al. Insights into secondary metabolism from a global analysis of prokaryotic biosynthetic gene clusters. Cell 158, 412–421 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Montalban-Lopez, M. et al. New developments in RiPP discovery, enzymology and engineering. Nat. Prod. Rep. 38, 130–239 (2021).

    Article  CAS  PubMed  Google Scholar 

  5. Ortega, M. A. et al. Structure and mechanism of the tRNA-dependent lantibiotic dehydratase NisB. Nature 517, 509–512 (2015).

    Article  CAS  PubMed  Google Scholar 

  6. Koehnke, J. et al. Structural analysis of leader peptide binding enables leader-free cyanobactin processing. Nat. Chem. Biol. 11, 558–563 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Regni, C. A. et al. How the MccB bacterial ancestor of ubiquitin E1 initiates biosynthesis of the microcin C7 antibiotic. EMBO J. 28, 1953–1964 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Evans, R. L. III, Latham, J. A., Xia, Y., Klinman, J. P. & Wilmot, C. M. Nuclear magnetic resonance structure and binding studies of PqqD, a chaperone required in the biosynthesis of the bacterial dehydrogenase cofactor pyrroloquinoline quinone. Biochemistry 56, 2735–2746 (2017).

    Article  CAS  PubMed  Google Scholar 

  9. Sumida, T., Dubiley, S., Wilcox, B., Severinov, K. & Tagami, S. Structural basis of leader peptide recognition in lasso peptide biosynthesis pathway. ACS Chem. Biol. 14, 1619–1627 (2019).

    Article  CAS  PubMed  Google Scholar 

  10. Chekan, J. R., Ongpipattanakul, C. & Nair, S. K. Steric complementarity directs sequence promiscuous leader binding in RiPP biosynthesis. Proc. Natl Acad. Sci. USA 116, 24049–24055 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Ghilarov, D. et al. Architecture of microcin B17 synthetase: an octameric protein complex converting a ribosomally synthesized peptide into a DNA gyrase poison. Mol. Cell 73, 749–762 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Grove, T. L. et al. Structural insights into thioether bond formation in the biosynthesis of sactipeptides. J. Am. Chem. Soc. 139, 11734–11744 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Davis, K. M. et al. Structures of the peptide-modifying radical SAM enzyme SuiB elucidate the basis of substrate recognition. Proc. Natl Acad. Sci. USA 114, 10420–10425 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Li, K., Condurso, H. L., Li, G., Ding, Y. & Bruner, S. D. Structural basis for precursor protein-directed ribosomal peptide macrocyclization. Nat. Chem. Biol. 12, 973–979 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  15. Chekan, J. R., Estrada, P., Covello, P. S. & Nair, S. K. Characterization of the macrocyclase involved in the biosynthesis of RiPP cyclic peptides in plants. Proc. Natl Acad. Sci. USA 114, 6551–6556 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Ludewig, H. et al. Characterization of the fast and promiscuous macrocyclase from plant PCY1 enables the use of simple substrates. ACS Chem. Biol. 13, 801–811 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Song, H. et al. A molecular mechanism for the enzymatic methylation of nitrogen atoms within peptide bonds. Sci. Adv. 4, eaat2720 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Ongpipattanakul, C. & Nair, S. K. Molecular basis for autocatalytic backbone N-methylation in RiPP natural product biosynthesis. ACS Chem. Biol. 13, 2989–2999 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Bothwell, I. R. et al. Characterization of glutamyl-tRNA-dependent dehydratases using nonreactive substrate mimics. Proc. Natl Acad. Sci. USA 116, 17245–17250 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Dong, S. H., Liu, A., Mahanta, N., Mitchell, D. A. & Nair, S. K. Mechanistic basis for ribosomal peptide backbone modifications. ACS Cent. Sci. 5, 842–851 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Thibodeaux, G. N., McClerren, A. L., Ma, Y., Gancayco, M. R. & van der Donk, W. A. Synergistic binding of the leader and core peptides by the lantibiotic synthetase HalM2. ACS Chem. Biol. 10, 970–977 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. 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 

  23. Ishitsuka, M. O., Kusumi, T., Kakisawa, H., Kaya, K. & Watanabe, M. M. Microviridin—a novel tricyclic depsipeptide from the toxic cyanobacterium Microcystis viridis. JACS 112, 8180–8182 (1990).

    Article  CAS  Google Scholar 

  24. Ziemert, N., Ishida, K., Liaimer, A., Hertweck, C. & Dittmann, E. Ribosomal synthesis of tricyclic depsipeptides in bloom-forming cyanobacteria. Angew. Chem. Int. Ed. Engl. 47, 7756–7759 (2008).

    Article  CAS  PubMed  Google Scholar 

  25. Philmus, B., Christiansen, G., Yoshida, W. Y. & Hemscheidt, T. K. Post-translational modification in microviridin biosynthesis. Chembiochem 9, 3066–3073 (2008).

    Article  CAS  PubMed  Google Scholar 

  26. Lee, H., Park, Y. & Kim, S. Enzymatic cross-linking of side chains generates a modified peptide with four hairpin-like bicyclic repeats. Biochemistry 56, 4927–4930 (2017).

    Article  CAS  PubMed  Google Scholar 

  27. Roh, H., Han, Y., Lee, H. & Kim, S. A topologically distinct modified peptide with multiple bicyclic core motifs expands the diversity of microviridin-like peptides. Chembiochem 20, 1051–1059 (2019).

    Article  CAS  PubMed  Google Scholar 

  28. Lee, H., Choi, M., Park, J. U., Roh, H. & Kim, S. Genome mining reveals high topological diversity of ω-ester-containing peptides and divergent evolution of ATP-grasp macrocyclases. J. Am. Chem. Soc. 142, 3013–3023 (2020).

    Article  CAS  PubMed  Google Scholar 

  29. Unno, K. & Kodani, S. Heterologous expression of cryptic biosynthetic gene cluster from Streptomyces prunicolor yields novel bicyclic peptide prunipeptin. Microbiol. Res. 244, 126669 (2021).

    Article  CAS  PubMed  Google Scholar 

  30. Ahmed, M. N. et al. Phylogenomic analysis of the microviridin biosynthetic pathway coupled with targeted chemo-enzymatic synthesis yields potent protease inhibitors. ACS Chem. Biol. 12, 1538–1546 (2017).

    Article  CAS  PubMed  Google Scholar 

  31. Weiz, A. R. et al. Harnessing the evolvability of tricyclic microviridins to dissect protease–inhibitor interactions. Angew. Chem. Int. Ed. Engl. 53, 3735–3738 (2014).

    Article  CAS  PubMed  Google Scholar 

  32. Reyna-Gonzalez, E., Schmid, B., Petras, D., Sussmuth, R. D. & Dittmann, E. Leader peptide-free in vitro reconstitution of microviridin biosynthesis enables design of synthetic protease-targeted libraries. Angew. Chem. Int. Ed. Engl. 55, 9398–9401 (2016).

    Article  CAS  PubMed  Google Scholar 

  33. Lee, C., Lee, H., Park, J. U. & Kim, S. Introduction of bifunctionality into the multidomain architecture of the ω-ester-containing peptide plesiocin. Biochemistry 59, 285–289 (2020).

    Article  CAS  PubMed  Google Scholar 

  34. Zhang, Y. et al. A distributive peptide cyclase processes multiple microviridin core peptides within a single polypeptide substrate. Nat. Commun. 9, 1780 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Tanaka, T., Nishioka, T. & Oda, J. Nicked multifunctional loop of glutathione synthetase still protects the catalytic intermediate. Arch. Biochem. Biophys. 339, 151–156 (1997).

    Article  CAS  PubMed  Google Scholar 

  36. Shi, Y. & Walsh, C. T. Active site mapping of Escherichia coli d-Ala-d-Ala ligase by structure-based mutagenesis. Biochemistry 34, 2768–2776 (1995).

    Article  CAS  PubMed  Google Scholar 

  37. Lipmann, F. & Tuttle, L. C. A specific micromethod for the determination of acyl phosphates. J. Biol. Chem. 159, 21–28 (1945).

    Article  CAS  Google Scholar 

  38. Jencks, W. P. The mechanism of the reaction of hydroxylamine with activated acyl groups. Biochim. Biophys. Acta 27, 417–418 (1958).

    Article  CAS  PubMed  Google Scholar 

  39. Fawaz, M. V., Topper, M. E. & Firestine, S. M. The ATP-grasp enzymes. Bioorg. Chem. 39, 185–191 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Weiz, A. R. et al. Leader peptide and a membrane protein scaffold guide the biosynthesis of the tricyclic peptide microviridin. Chem. Biol. 18, 1413–1421 (2011).

    Article  CAS  PubMed  Google Scholar 

  41. Oman, T. J., Knerr, P. J., Bindman, N. A., Velasquez, J. E. & van der Donk, W. A. An engineered lantibiotic synthetase that does not require a leader peptide on its substrate. J. Am. Chem. Soc. 134, 6952–6955 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Dunbar, K. L. & Mitchell, D. A. Insights into the mechanism of peptide cyclodehydrations achieved through the chemoenzymatic generation of amide derivatives. J. Am. Chem. Soc. 135, 8692–8701 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Wang, H. & van der Donk, W. A. Biosynthesis of the class III lantipeptide catenulipeptin. ACS Chem. Biol. 7, 1529–1535 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Zhao, G. et al. Structure and function of Escherichia coli RimK, an ATP-grasp fold, l-glutamyl ligase enzyme. Proteins 81, 1847–1854 (2013).

    Article  CAS  PubMed  Google Scholar 

  45. Hara, T., Kato, H., Katsube, Y. & Oda, J. A pseudo-Michaelis quaternary complex in the reverse reaction of a ligase: structure of Escherichia coli B glutathione synthetase complexed with ADP, glutathione, and sulfate at 2.0 Å resolution. Biochemistry 35, 11967–11974 (1996).

    Article  CAS  PubMed  Google Scholar 

  46. Ouchi, T. et al. Lysine and arginine biosyntheses mediated by a common carrier protein in Sulfolobus. Nat. Chem. Biol. 9, 277–283 (2013).

    Article  CAS  PubMed  Google Scholar 

  47. Thoden, J. B., Firestine, S., Nixon, A., Benkovic, S. J. & Holden, H. M. Molecular structure of Escherichia coli PurT-encoded glycinamide ribonucleotide transformylase. Biochemistry 39, 8791–8802 (2000).

    Article  CAS  PubMed  Google Scholar 

  48. Batson, S. et al. Inhibition of d-Ala:d-Ala ligase through a phosphorylated form of the antibiotic d-cycloserine. Nat. Commun. 8, 1939 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  49. Fan, C., Moews, P. C., Walsh, C. T. & Knox, J. R. Vancomycin resistance: structure of d-alanine:d-alanine ligase at 2.3 Å resolution. Science 266, 439–443 (1994).

    Article  CAS  PubMed  Google Scholar 

  50. Salwiczek, M., Nyakatura, E. K., Gerling, U. I., Ye, S. & Koksch, B. Fluorinated amino acids: compatibility with native protein structures and effects on protein–protein interactions. Chem. Soc. Rev. 41, 2135–2171 (2012).

    Article  CAS  PubMed  Google Scholar 

  51. Hemsley, A., Arnheim, N., Toney, M. D., Cortopassi, G. & Galas, D. J. A simple method for site-directed mutagenesis using the polymerase chain reaction. Nucleic Acids Res. 17, 6545–6551 (1989).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Bouwman, S., Orru, R. V. & Ruijter, E. Stereoselective synthesis of fluorinated aminoglycosyl phosphonates. Org. Biomol. Chem. 13, 1317–1321 (2015).

    Article  CAS  PubMed  Google Scholar 

  53. Sausen, C. W., Rogers, C. M. & Bochman, M. L. Thin-layer chromatography and real-time coupled assays to measure ATP hydrolysis. Methods Mol. Biol. 1999, 245–253 (2019).

    Article  CAS  PubMed  Google Scholar 

  54. Otwinowski, Z. & Minor, W. Processing of X-ray diffraction data collected in oscillation mode. Methods Enzymol. 276, 307–326 (1997).

    Article  CAS  PubMed  Google Scholar 

  55. McCoy, A. J. et al. Phaser crystallographic software. J. Appl. Crystallogr. 40, 658–674 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. 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 

  57. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D 60, 2126–2132 (2004).

    Article  CAS  PubMed  Google Scholar 

  58. Liebschner, D. et al. Polder maps: improving OMIT maps by excluding bulk solvent. Acta Crystallogr. D 73, 148–157 (2017).

    Article  CAS  Google Scholar 

  59. Ko, J. et al. The FALC-Loop web server for protein loop modeling. Nucleic Acids Res. 39, W210–W214 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Park, H., Lee, G. R., Heo, L. & Seok, C. Protein loop modeling using a new hybrid energy function and its application to modeling in inaccurate structural environments. PLoS ONE 9, e113811 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

We thank H. Lee, G. Eom, C. Lee, H. Roh, H. Cho and H. Park for helpful discussions and technical assistance, H. Chung, Y.J. Lee and T.Y. Im for assistance in synthesizing compounds, Y.T. Kim, W.J. Jeong and Y. Choi for advice in crystallization and crystallographic analysis and H. Woo for help in loop modeling. We also thank the staff of the 5C and 7A beamlines at the Pohang Light Source. This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2020R1F1A1054191 and 2021R1A2C1008730) to S.K.

Author information

Authors and Affiliations

Authors

Contributions

I.S. and S.K. conceived the project, designed experiments, analyzed data and wrote the paper. S.K. supervised the project. I.S. performed the majority of experiments. Y.K. performed a part of the mutant analysis. I.S. and J.Y. conducted crystallographic experiments under the supervision of W.J.S. and S.K. I.S. and S.Y.G. synthesized the phosphomimetic glutamate under the supervision of H.G.L. and S.K.

Corresponding author

Correspondence to Seokhee Kim.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Chemical Biology thanks Jesko Koehnke 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 Designed minimal precursor recapitulates the reactivity of PsnB.

a, Sequence logo of precursors of Group 2a graspetides including the leader peptide (red region) and one core motif (blue region). The LFIEDL region is highly conserved in Group 2a graspetides (yellow box). b,c, ATPase assays titrating with ATP (b) in solutions containing 0.4 μM PsnB, 200 μM minimal precursor (MP), 100 mM Tris pH 7.3, 50 mM KCl, and 10 mM MgCl2, or titrating with MgCl2 (c) in solutions containing 0.4 μM PsnB, 200 μM MP, 100 mM Tris pH 7.3, 50 mM KCl, and 5 mM ATP. Data are presented as dot plots with mean ± 1 SD (n = 3 independent experiments) and fitted to a hyperbolic equation. d, Minimal precursor was successfully modified by PsnB with a modification rate similar to that of wild-type PsnA2. 50 μM MP (top panel; the product contains two ester bonds) or wild-type PsnA2 (bottom panel; the product contains eight ester bonds) was co-incubated with 0.5 μM PsnB in buffer A (100 mM Tris pH 7.3, 50 mM KCl, 5 mM ATP, and 10 mM MgCl2) at 37 °C, and the reaction solutions at designated time points were analyzed by MALDI-MS.

Source data

Extended Data Fig. 2 Acyl-phosphate intermediates were trapped by hydroxylamine.

a, Scheme of the acyl-phosphate trapping during the macrocyclization reaction. Nucleophilic attack of hydroxylamine rather than the core threonine generates the hydroxylamine adducts of the precursor. b, MALDI analysis of the reaction solution with (left) or without (right) hydroxylamine. 0.5 M hydroxylamine was added to the reaction mixture containing 100 μM MP (1-a) and 6 μM PsnB in buffer A. The reaction mixtures were analyzed by MALDI after 4 hour incubation at 37 °C. Co-incubation of 0.5 M hydroxylamine generated NH2OH-added precursor peptides (1-b and 2-b), which are the result of nucleophilic attack of hydroxylamine to acyl-phosphate intermediates. c, 2-b was purified by HPLC and methanolysis was performed with purified 2-b as previously reported26. The methanolysis product (2-c) was detected by MALDI. d, MALDI-MS/MS analysis of three hydroxylamine adducts. The connectivities of ester bonds were determined by MS/MS analysis with NH2OH-added precursor peptides (1-b and 2-b) and methanolysis product (2-c).

Extended Data Fig. 3 Binding of the leader peptide activates PsnB.

a, LP (PsnA214-24) enhances the affinity of the CP (PsnA228-38) to PsnB. Fraction bound of Fl_CP (0.1 µM) to PsnB (50 µM) was determined by fluorescence anisotropy. b, LP enhances the ATPase activity of PsnB. Basal ATP consumption rate of PsnB was 0.14 min-1enz-1, whereas the addition of 200 μM LP increased the ATP consumption rate to 0.89 min-1enz-1. P value < 0.01 by two-sided Student’s t-test. c, Fluorescence anisotropy of Fl_LP (0.1 µM) to wild-type PsnB or leader-fused PsnB (LP_PsnB). Data are presented as dot plots with mean ± 1 SD (n = 3 independent experiments; a-c) and fitted to a hyperbolic equation (c).

Source data

Extended Data Fig. 4 Binding and modification of ring-containing precursors.

a,b, Affinity of ring-containing MPs (MP-1H2O, the single-ring intermediate; MP-2H2O, the double-ring product) to PsnB was determined by fluorescence anisotropy without nucleotide (a) or with 1 mM ADP (b). c, ATPase activity of PsnB was measured with different concentrations of MP or the single-ring intermediate (MP-1H2O). Data are presented as dot plots with mean ± 1 SD (n = 3 independent experiments; a-c) and fitted to a hyperbolic equation (a-c).

Source data

Extended Data Fig. 5 Four crystal structures of PsnB complexes.

Structure of four different states of PsnB dimers (E, enzyme, yellow or pale cyan cartoon; N, nucleotide, cyan sticks; L, leader, magenta sticks; C, core, green sticks). Polder OMIT map (gray mesh) of each peptide or nucleotide is contoured at 4.0 σ.

Extended Data Fig. 6 Polder OMIT maps that are calculated without the model for LP, CP, or nucleotide (contour level at 4.0 σ).

For the LP and CP, resolved residues are listed on Supplementary Table 2. Phospho-mimetic side-chain in CP is only resolved well in chain E of 7DRP.

Extended Data Fig. 7 No higher oligomers were observed from PsnB.

a, Gel-filtration chromatogram of PsnB without MgCl2 (black solid line), with MgCl2 (red solid line), or with both MgCl2 and ATP (blue solid line). A chromatogram of molecular weight standard (black dashed line) is also shown with known molecular weights. Addition of nucleotide did not induce the formation of the stable higher oligomer of PsnB dimer. be, FRET of a solution containing both donor- and acceptor-labeled PsnB was measured with different amounts of MgCl2 (b), AMPPNP (c), LP (d), or MP (e). Additional components in solutions are listed above the plots. Neither nucleotide nor precursor induced stable intermolecular interaction of PsnB. f, Loop modeling revealed that β13β14 loop is long and flexible enough for intramolecular interaction. Residues between Ile227 and Ile247, the β13β14 loop region, were modeled by FALC59,60 and overall complex structure was optimized by relaxation. In the modeled structure, the conformation of the β13β14 loop was flipped and the DFR motif moved toward the enzyme active site. Also, Arg235 had hydrogen bonding with nucleotide which is similar to the intermolecular interaction scheme shown in crystal structures (Fig. 4c).

Source data

Extended Data Fig. 8 Binding property of core-binding site mutants of leader-fused PsnB.

Fluorescence anisotropy of Fl_CP (0.1 µM) with leader-fused PsnB mutants. Mutation of core-binding residues reduced the affinity between CP and the enzyme. Data are presented as dot plots with mean ± 1 SD (n = 3 independent experiments) and fitted to a hyperbolic equation.

Source data

Extended Data Fig. 9 Molecular interaction, binding affinity, and inhibition effect of MP(pE37).

a, Detailed interaction scheme of MP(pE37) and PsnB. pGlu37 makes no interaction, whereas Glu38 interacts with Arg213 (yellow sticks) of PsnB. b, Binding affinity of Fl-MP(pE37) to PsnB was measured by fluorescence anisotropy. MP(pE37) binds less tightly to PsnB than the MP no matter whether ADP is present. Data are presented as dot plots with mean ± 1 SD (n = 3 independent experiments) and fitted to a hyperbolic equation. c, Inhibition assay was performed with MP(pE37) or LP. 10 μM MP and 0.4 μM PsnB was co-incubated with various concentrations of MP(pE37) (left) or LP (right) in a buffer containing Tris-HCl (100 mM; pH 7.3), MgCl2 (10 mM), KCl (50 mM) and ATP (5 mM) at 37 °C for 10 min.

Source data

Extended Data Fig. 10 Precursor and nucleotide binding induces conformational change of PsnB.

a, Surface models of the leader-binding domain without (left) or with (right) the bound Phe15 in the LP (magenta sticks). Hydrophobic pocket for Phe15 is shown as magenta dashed lines. b, The nucleotide (cyan sticks) binding shifts the β9β10 loop (from gray to yellow cartoon). Lys172 and Thr180 that interact with a nucleotide are shown as green sticks. c, Two α3α4 pairs of a PsnB dimer show extensive interactions to each other to form a rigid body. Two PsnB subunits are shown as yellow and pale cyan cartoons, and their interacting residues are presented as orange and cyan sticks, respectively. d, Surface model of the ENLC-E complex (yellow and pale cyan for two PsnB subunits). Binding of LP (magenta) and nucleotide (cyan spheres) induces the conformational change of PsnB dimer to generate a compact core (green) binding site. e, Superposition of four LP and CP pairs from PsnB-MP-ADP (7DRM, green and cyan ribbons) and PsnB-MP-AMPPNP (7DRN, magenta and yellow ribbons). All LPs and CPs are well overlapped. PsnB is shown as gray cartoon and half-transparent surface. Nucleotide and Arg213 are shown as gray sticks. Glu37 in CP is shown as sticks.

Supplementary information

Supplementary Information

Supplementary Figs. 1–6, Tables 1–5 and Note.

Reporting Summary

Supplementary Video 1

Supplementary Video 1

Source data

Source Data Fig. 1

Statistical source data for Fig. 1d,e.

Source Data Fig. 2

Statistical source data for Fig. 2a,b,d–f.

Source Data Fig. 3

Statistical source data for Fig. 3a–c.

Source Data Fig. 4

Statistical source data for Fig. 4b.

Source Data Fig. 5

Statistical source data for Fig. 5b,d.

Source Data Extended Data Fig. 1

Statistical source data for Extended Data Fig. 1b,c.

Source Data Extended Data Fig. 3

Statistical source data for Extended Data Fig. 3a–c.

Source Data Extended Data Fig. 4

Statistical source data for Extended Data Fig. 4a–c.

Source Data Extended Data Fig. 7

Statistical source data for Extended Data Fig. 7b–e.

Source Data Extended Data Fig. 8

Statistical source data for Extended Data Fig. 8.

Source Data Extended Data Fig. 9

Statistical source data for Extended Data Fig. 9b.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Song, I., Kim, Y., Yu, J. et al. Molecular mechanism underlying substrate recognition of the peptide macrocyclase PsnB. Nat Chem Biol 17, 1123–1131 (2021). https://doi.org/10.1038/s41589-021-00855-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-021-00855-x

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