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:

Conformational remodeling enhances activity of lanthipeptide zinc-metallopeptidases

Abstract

Lanthipeptides are an important group of natural products with diverse biological functions, and their biosynthesis requires the removal of N-terminal leader peptides (LPs) by designated proteases. LanPM1 enzymes, a subgroup of M1 zinc-metallopeptidases, have been recently identified as bifunctional proteases with both endo- and aminopeptidase activities to remove LPs of class III and class IV lanthipeptides. Herein, we report the biochemical and structural characterization of EryP as the LanPM1 enzyme from the biosynthesis of class III lanthipeptide erythreapeptin. We determined X-ray crystal structures of EryP in three conformational states, the open, intermediate and closed states, and identified a unique interdomain Ca2+ binding site as a regulatory element that modulates its domain dynamics and proteolytic activity. Inspired by this regulatory Ca2+ binding, we developed a strategy to engineer LanPM1 enzymes for enhanced catalytic activities by strengthening interdomain associations and driving the conformational equilibrium toward their closed forms.

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: LanPM1 enzymes in lanthipeptide biosynthesis.
Fig. 2: Bifunctional protease EryP processes both EryALP and AplAcyc peptides.
Fig. 3: Overall structure and the metal ion binding sites of EryP.
Fig. 4: The Ca2+ ion binding site regulates the conformational equilibrium and catalytic activity of EryP.
Fig. 5: The E802R mutation enhanced the interdomain interaction and the catalytic activity of EryP.
Fig. 6: AplPR98E-A368E-A779R display improved efficiency in removing the LP of AplAcyc peptide in comparison with AplP.

Similar content being viewed by others

Data availability

All crystal structures have been deposited in the PDB under accession number 7V9N for EryPclosed, 7V9P for EryPintermediate, 7V9Q for EryPopen and 7V9O for EryPE802R. All other data are contained in the published article or are available on request.

References

  1. Rawlings, N. D., Barrett, A. J. & Bateman, A. MEROPS: the peptidase database. Nucleic Acids Res. 38, D227–D233 (2010).

    Article  CAS  PubMed  Google Scholar 

  2. Chandu, D., Kumar, A. & Nandi, D. PepN, the major Suc-LLVY-AMC-hydrolyzing enzyme in Escherichia coli, displays functional similarity with downstream processing enzymes in Archaea and Eukarya. Implications in cytosolic protein degradation. J. Biol. Chem. 278, 5548–5556 (2003).

    Article  CAS  PubMed  Google Scholar 

  3. Chandu, D. & Nandi, D. PepN is the major aminopeptidase in Escherichia coli: insights on substrate specificity and role during sodium-salicylate-induced stress. Microbiology 149, 3437–3447 (2003).

    Article  CAS  PubMed  Google Scholar 

  4. Tamura, N., Lottspeich, F., Baumeister, W. & Tamura, T. The role of tricorn protease and its aminopeptidase-interacting factors in cellular protein degradation. Cell 95, 637–648 (1998).

    Article  CAS  PubMed  Google Scholar 

  5. Hermans, S. J. et al. Crystal structure of human insulin-regulated aminopeptidase with specificity for cyclic peptides. Protein Sci. 24, 190–199 (2015).

    Article  CAS  PubMed  Google Scholar 

  6. Zervoudi, E. et al. Rationally designed inhibitor targeting antigen-trimming aminopeptidases enhances antigen presentation and cytotoxic T-cell responses. Proc. Natl Acad. Sci. USA 110, 19890–19895 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Nguyen, T. T. et al. Structural basis for antigenic peptide precursor processing by the endoplasmic reticulum aminopeptidase ERAP1. Nat. Struct. Mol. Biol. 18, 604–613 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. 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 

  9. 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 

  10. Jabes, D. et al. Efficacy of the new lantibiotic NAI-107 in experimental infections induced by multidrug-resistant Gram-positive pathogens. Antimicrob. Agents Chemother. 55, 1671–1676 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Mathur, H., O’Connor, P. M., Hill, C., Cotter, P. D. & Ross, R. P. Analysis of anti-Clostridium difficile activity of thuricin CD, vancomycin, metronidazole, ramoplanin, and actagardine, both singly and in paired combinations. Antimicrob. Agents Chemother. 57, 2882–2886 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Goldstein, B. P., Wei, J., Greenberg, K. & Novick, R. Activity of nisin against Streptococcus pneumoniae, in vitro, and in a mouse infection model. J. Antimicrob. Chemother. 42, 277–278 (1998).

    Article  CAS  PubMed  Google Scholar 

  13. Mohr, K. I. et al. Pinensins: the first antifungal lantibiotics. Angew. Chem. Int. Ed. 54, 11254–11258 (2015).

    Article  CAS  Google Scholar 

  14. Ferir, G. et al. The lantibiotic peptide labyrinthopeptin A1 demonstrates broad anti-HIV and anti-HSV activity with potential for microbicidal applications. PloS ONE 8, e64010 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Smith, T. E. et al. Accessing chemical diversity from the uncultivated symbionts of small marine animals. Nat. Chem. Biol. 14, 179–185 (2018).

  16. Iorio, M. et al. A glycosylated, labionin-containing lanthipeptide with marked antinociceptive activity. ACS Chem. Biol. 9, 398–404 (2014).

    Article  CAS  PubMed  Google Scholar 

  17. Meindl, K. et al. Labyrinthopeptins: a new class of carbacyclic lantibiotics. Angew. Chem. Int. Ed. 49, 1151–1154 (2010).

    Article  CAS  Google Scholar 

  18. Repka, L. M., Chekan, J. R., Nair, S. K. & van der Donk, W. A. Mechanistic understanding of lanthipeptide biosynthetic enzymes. Chem. Rev. 117, 5457–5520 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Kozakai, R. et al. Acyltransferase that catalyses the condensation of polyketide and peptide moieties of goadvionin hybrid lipopeptides. Nat. Chem. 12, 869–877 (2020).

    Article  CAS  PubMed  Google Scholar 

  20. Wiebach, V. et al. The anti-staphylococcal lipolanthines are ribosomally synthesized lipopeptides. Nat. Chem. Biol. 14, 652–654 (2018).

    Article  CAS  PubMed  Google Scholar 

  21. Hegemann, J. D. & Süssmuth, R. D. Matters of class: coming of age of class III and IV lanthipeptides. RSC Chem. Biol. 1, 110–127 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Chen, S. et al. Zn-dependent bifunctional proteases are responsible for leader peptide processing of class III lanthipeptides. Proc. Natl Acad. Sci. USA 116, 2533–2538 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Ren, H., Shi, C., Bothwell, I. R., van der Donk, W. A. & Zhao, H. Discovery and characterization of a class IV lanthipeptide with a non-overlapping ring pattern. ACS Chem. Biol. 15, 1642–1649 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Wiebach, V. et al. An amphipathic alpha-helix guides maturation of the ribosomally-synthesized lipolanthines. Angew. Chem. Int. Ed. 59, 16777–16785 (2020).

    Article  CAS  Google Scholar 

  25. Lazdunski, C., Busuttil, J. & Lazdunski, A. Purification and properties of a periplasmic aminoendopeptidase from Escherichia coli. Eur. J. Biochem. 60, 363–369 (1975).

    Article  CAS  PubMed  Google Scholar 

  26. Mccaman, M. T. & Villarejo, M. R. Structural and catalytic properties of peptidase-N from Escherichia-coli K-12. Arch. Biochem. Biophys. 213, 384–394 (1982).

    Article  CAS  PubMed  Google Scholar 

  27. Voller, G. H. et al. Characterization of new class III lantibiotics–erythreapeptin, avermipeptin and griseopeptin from Saccharopolyspora erythraea, Streptomyces avermitilis and Streptomyces griseus demonstrates stepwise N-terminal leader processing. Chem. Bio. Chem. 13, 1174–1183 (2012).

    Article  PubMed  Google Scholar 

  28. Cerda-Costa, N. & Gomis-Ruth, F. X. Architecture and function of metallopeptidase catalytic domains. Protein Sci. 23, 123–144 (2014).

    Article  CAS  PubMed  Google Scholar 

  29. Thompson, M. W., Archer, E. D., Romer, C. E. & Seipelt, R. L. A conserved tyrosine residue of Saccharomyces cerevisiae leukotriene A4 hydrolase stabilizes the transition state of the peptidase activity. Peptides 27, 1701–1709 (2006).

    Article  CAS  PubMed  Google Scholar 

  30. Giastas, P. et al. Mechanism for antigenic peptide selection by endoplasmic reticulum aminopeptidase 1. Proc. Natl Acad. Sci. USA 116, 26709–26716 (2019).

    Article  CAS  PubMed Central  Google Scholar 

  31. Chen, L., Lin, Y. L., Peng, G. & Li, F. Structural basis for multifunctional roles of mammalian aminopeptidase N. Proc. Natl Acad. Sci. USA 109, 17966–17971 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Kochan, G. et al. Crystal structures of the endoplasmic reticulum aminopeptidase-1 (ERAP1) reveal the molecular basis for N-terminal peptide trimming. Proc. Natl Acad. Sci. USA 108, 7745–7750 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Kyrieleis, O. J., Goettig, P., Kiefersauer, R., Huber, R. & Brandstetter, H. Crystal structures of the tricorn interacting factor F3 from Thermoplasma acidophilum, a zinc aminopeptidase in three different conformations. J. Mol. Biol. 349, 787–800 (2005).

    Article  CAS  PubMed  Google Scholar 

  34. Addlagatta, A., Gay, L. & Matthews, B. W. Structure of aminopeptidase N from Escherichia coli suggests a compartmentalized, gated active site. Proc. Natl Acad. Sci. USA 103, 13339–13344 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Maben, Z., Arya, R., Georgiadis, D., Stratikos, E. & Stern, L. J. Conformational dynamics linked to domain closure and substrate binding explain the ERAP1 allosteric regulation mechanism. Nat. Commun. 12, 5302 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Tian, W., Chen, C., Lei, X., Zhao, J. & Liang, J. CASTp 3.0: computed atlas of surface topography of proteins. Nucleic Acids Res. 46, W363–W367 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Gangola, P. & Rosen, B. P. Maintenance of intracellular calcium in Escherichia coli. J. Biol. Chem. 262, 12570–12574 (1987).

    Article  CAS  PubMed  Google Scholar 

  38. Grubbs, R. D. Intracellular magnesium and magnesium buffering. Biometals 15, 251–259 (2002).

    Article  CAS  PubMed  Google Scholar 

  39. Price, I. R., Gaballa, A., Ding, F., Helmann, J. D. & Ke, A. Mn(2+)-sensing mechanisms of yybP-ykoY orphan riboswitches. Mol. Cell 57, 1110–1123 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Cadel, S., Darmon, C., Pernier, J., Herve, G. & Foulon, T. The M1 family of vertebrate aminopeptidases: role of evolutionarily conserved tyrosines in the enzymatic mechanism of aminopeptidase B. Biochimie 109, 67–77 (2015).

    Article  CAS  PubMed  Google Scholar 

  41. Tholander, F. et al. Structure-based dissection of the active site chemistry of leukotriene A4 hydrolase: implications for M1 aminopeptidases and inhibitor design. Chem. Biol. 15, 920–929 (2008).

    Article  CAS  PubMed  Google Scholar 

  42. Sui, L. & Guo, H. C. ERAP1 binds peptide C-termini of different sequences and/or lengths by a common recognition mechanism. Immunobiology 226, 152112 (2021).

    Article  CAS  PubMed  Google Scholar 

  43. Liddle, J. et al. Targeting the regulatory site of ER aminopeptidase 1 leads to the discovery of a natural product modulator of antigen presentation. J. Med. Chem. 63, 3348–3358 (2020).

    Article  CAS  PubMed  Google Scholar 

  44. Eijsink, V. G., Matthews, B. W. & Vriend, G. The role of calcium ions in the stability and instability of a thermolysin-like protease. Protein Sci. 20, 1346–1355 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Papakyriakou, A. & Stratikos, E. The role of conformational dynamics in antigen trimming by intracellular aminopeptidases. Front. Immunol. 8, 946 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Bode, W., Gomis-Ruth, F. X., Huber, R., Zwilling, R. & Stocker, W. Structure of astacin and implications for activation of astacins and zinc-ligation of collagenases. Nature 358, 164–167 (1992).

    Article  CAS  PubMed  Google Scholar 

  47. Ito, K. et al. Crystal structure of aminopeptidase N (proteobacteria alanyl aminopeptidase) from Escherichia coli and conformational change of methionine 260 involved in substrate recognition. J. Biol. Chem. 281, 33664–33676 (2006).

    Article  CAS  PubMed  Google Scholar 

  48. Pettersen, E. F. et al. UCSF Chimera–a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004).

    Article  CAS  PubMed  Google Scholar 

  49. 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 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  Google Scholar 

  52. 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 

  53. Pettersen, E. F. et al. UCSF ChimeraX: structure visualization for researchers, educators, and developers. Protein Sci. 30, 70–82 (2021).

    Article  CAS  PubMed  Google Scholar 

  54. Rostkowski, M., Olsson, M. H. M., Sondergaard, C. R. & Jensen, J. H. Graphical analysis of pH-dependent properties of proteins predicted using PROPKA. BMC Struct. Biol. 11, 6–12 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Li, P. & Merz, K. M. Jr. MCPB.py: a Python based metal center parameter builder. J. Chem. Inf. Model 56, 599–604 (2016).

    Article  CAS  PubMed  Google Scholar 

  56. Case, D. A. et al. AMBER 16 (University of California, San Francisco, 2016).

  57. Seminario, J. M. Calculation of intramolecular force fields from second-derivative tensors. Int. J. Quantum Chem. 60, 1271–1277 (1996).

    Article  Google Scholar 

  58. 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 

  59. Friesner, R. A. et al. Extra precision glide: docking and scoring incorporating a model of hydrophobic enclosure for protein–ligand complexes. J. Med. Chem. 49, 6177–6196 (2006).

    Article  CAS  PubMed  Google Scholar 

  60. Ward, J. H. Hierarchical grouping to optimize an objective function. J. Am. Stat. Assoc. 58, 236–244 (1963).

    Article  Google Scholar 

  61. Barone, V. & Cossi, M. Quantum calculation of molecular energies and energy gradients in solution by a conductor solvent model. J. Phys. Chem. A 102, 1995–2001 (1998).

    Article  CAS  Google Scholar 

  62. Takano, Y. & Houk, K. N. Benchmarking the conductor-like polarizable continuum model (CPCM) for aqueous solvation free energies of neutral and ionic organic molecules. J. Chem. Theory Comput. 1, 70–77 (2005).

    Article  PubMed  Google Scholar 

  63. Cossi, M., Rega, N., Scalmani, G. & Barone, V. Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation model. J. Comput. Chem. 24, 669–681 (2003).

    Article  CAS  PubMed  Google Scholar 

  64. Morris, G. M. et al. Automated docking using a Lamarckian genetic algorithm and an empirical binding free energy function. J. Comput. Chem. 19, 1639–1662 (1998).

    Article  CAS  Google Scholar 

  65. Jakalian, A., Bush, B. L., Jack, D. B. & Bayly, C. I. Fast, efficient generation of high-quality atomic charges. AM1-BCC model: I. method. J. Comput. Chem. 21, 132–146 (2000).

    Article  CAS  Google Scholar 

  66. Jakalian, A., Jack, D. B. & Bayly, C. I. Fast, efficient generation of high-quality atomic charges. AM1-BCC model: II. parameterization and validation. J. Comput. Chem. 23, 1623–1641 (2002).

    Article  CAS  PubMed  Google Scholar 

  67. Wang, J. M., 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 

  68. Darden, T., York, D. & Pedersen, L. Particle mesh Ewald - an n.log(n) method for Ewald sums in large systems. J. Chem. Phys. 98, 10089–10092 (1993).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work is supported by the National Science Foundation of China (grant nos. 21922703, 91953112 and 21861142005 to H.W.; 81871615 and 81670008 to R.B.), the Natural Science Foundation of Jiangsu Province (grant no. BK20190004 and BK20202004 to H.W. and BK20200335 to W.W.), the Fundamental Research Funds for the Central Universities (grant no. 14380131 to H.W.), the Jiangsu Innovation & Entrepreneurship Talents Plan to H.W., National Key R&D Program of China (grant no. 2019YFA0905800 to H.W.), Ministry of Science and Technology of the People’s Republic of China (grant no. 2018ZX09201018–005 to R.B.) and National Mega-project for Innovative Drugs (grant no.2019ZX09721001–001–001 to R.B.). We thank National Center for Protein Sciences Shanghai (NCPSS) beamlines BL18U and BL19U allowance. We thank the staffs of NCPSS beamlines BL18U and BL19U and SSRF BL17U, Shanghai, People’s Republic of China, for assistance during data collection. We thank the High-Performance Computing Center of Nanjing University for the numerical calculations on its blade cluster system.

Author information

Authors and Affiliations

Authors

Contributions

H.W. and R.B. initiated and directed this study together with W.W. W.S., C.Z., Y.W., X.X. and J.Z. prepared enzymes and peptides and performed biochemical assays. C.Z. performed structural biology experiments. W.W. and Y.L. performed computational studies on protein dynamics. All authors participate in the data analysis and manuscript preparation.

Corresponding authors

Correspondence to Wanqing Wei, Rui Bao or Huan Wang.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Chemical Biology thanks Efstratios Stratikos 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 Structural comparison between EryAcyc and AplAcyc.

Structural comparison between EryAcyc and AplAcyc.

Extended Data Fig. 2 EryP cleaves AplAcyc peptide as an endopeptidase at multiple sites, as determined by MALDI-TOF analysis.

Assay conditions: 100 μM AplAcyc peptide and 1.0 μM EryP were incubated in 20 mM Tris buffer, pH 8.0, at 37 °C for indicated time.

Extended Data Fig. 3 EryP contains a zinc binding motif that is highly conserved in M1 Zn2+-dependent metallopeptidases.

(a) The overall crystal structure of EryP and the close-up view of the Zn binding residues. (b) Sequence alignment of EryP with M1 Zn2+-dependent metallopeptidases. Catalytic residues are labeled with stars and the conserved HEXXH(X)18E is highlighted. Accession numbers of related proteins: EryP(WP_009950696.1), AplP(AHB63590.1), ePepN(AAA24317.1), ERAP1(NP_001185470.1).

Extended Data Fig. 4 Omit maps for the Zn2+ and Ca2+ in EryPclosed.

The Fo-Fc density maps (contoured at 3.0 σ) for the Zn2+ (a) and Ca2+ (b) with the ion removed are shown as grey mesh.

Extended Data Fig. 5 Both endopeptidase and aminopeptidase activities of EryPE307Q and EryPY392F were abolished.

(a) EryPE307Q and EryPY392F were inactive toward EryALP peptide. Assay conditions: EryALP peptide (100 μM) and EryP (1.0 μM) were incubated in 20 mM Tris buffer, pH 8.0, at 37 °C for 24 h. (b) The aminopeptidase activity of EryPE307Q and EryPY392F toward amino acid pNA derivatives were almost abolished compared with EryP. * represents the sodium adducts of peptides in MS. Error bars indicate standard deviation of three independent replicates. Data represent the mean ± s.d. from three replicates.

Extended Data Fig. 6 Bestatin and Zn2+-chelating reagent o-phenanthroline significantly decreased both endo- and aminopeptidase activities of EryP.

Bestatin and Zn2+-chelating reagent o-phenanthroline significantly decreased both endo- and aminopeptidase activities of EryP.

Extended Data Fig. 7 The docking model of a dipeptide Leu-Glu into the active site of EryP in the closed state.

The binding pocket (a) and 2D diagram (b) of dipeptide Leu-Glu binding with EryPclosed are shown. Following the convention for naming peptidase sites, the site responsible for accommodating the peptide side chain N-terminal to the cleavage site is named S1, and the subsequent position are named S1′. In the dipeptide docking model, the highly conserved G272AME275N motif in EryP binds to the dipeptide substrate through multiple hydrogen bonds in the docking model. In particular, residue E275 directly binds to the amino group at the N-terminus of the dipeptide with additional binding from residues E132 and E329. This docking model suggests that both S1 and S1′ pockets are spacious to accommodate amino acid side chains of various sizes in peptide substrates, allowing the enzyme to sequentially cleave various residues from EryALP as an aminopeptidase.

Extended Data Fig. 8 Computational docking of a hexapeptide (ELDAPN) as an endopeptidase substrate to be cleaved between residues Asp and Ala in EryPclosed.

(a-b) and structures of ERAP1 (PDB ID: 2YD0) (c) and IRAP (PDB ID: 5MJ6) (d) bound with peptide inhibitors. In the hexapeptide docking model, the hexapeptide adopted a bent conformation with the scissile peptide bond between Asp and Ala of the hexapeptide binding to the catalytic zinc ion with the amide oxygen. The N-terminal segment of the hexapeptide was accommodated in the internal cavity extending from the catalytic site in domain II toward the interdomain Ca2+ binding site. The N-terminal amino group of the hexapeptide is anchored firmly by electrostatic interactions with E132. The carboxylate of the Asp of the hexapeptide forms hydrogen bonds with N328 and Y392 in the S1 pocket as well as the side chain of the Ala forms hydrophobic interactions with residues in the S1′ pocket from the G272AMEN motif. This docking model also reveals that there is space to accommodate additional amino acid extending at both the N- and C-terminus of the hexapeptide, a configuration that would allow the binding of longer peptides as endopeptidase substrates in EryPclosed.

Extended Data Fig. 9 Structural comparison of EryP with peptide-binding M1 aminopeptidase.

Cutaway views of the binding pocket of EryPclosed (a), docked EryPclosed-(10 mer peptide) complex (b), ERAP1-(10 mer peptide) complex (PDB ID: 6RQX) (c) and APN-substance P complex (PDB ID: 6RQX) (d) are shown, respectively. M1 aminopeptidases are shown in surface mode, the ligands are shown in cartoon mode.

Extended Data Fig. 10 Close views of EryP-(10-mer peptide) MD representative snapshot, ERAP1-(10-mer peptide) complex and Porcine APN-(substance P) complex.

(a) EryP-(10-mer peptide) MD representative snapshot, (b) ERAP1-(10-mer peptide) complex crystal structure and (c) Porcine APN-(substance P) complex crystal structure. Ca2+ ion is shown in red sphere and Zn2+ ion in yellow sphere. As shown in (a), the 10-mer peptide lied in a fit inside the cavity formed around the active site in the space between domains I/ II and domain IV of EryP and showed an extended conformation similar to ERAP1-(10 mer peptide) and porcine APN-(substance P) complexes (b and c). The N-terminus of the 10-mer peptide that bears the phosphinic moiety is bound on the catalytic zinc ion (yellow sphere) in a motif observed previously in crystal structures of homologous aminopeptidases with bound phosphinic groups. Van der Waals interactions and electrostatic interactions with the main chain and side chains of residues of EryP were found to stabilize the substrate (a). In addition, space for additional atoms is available at the N/C-terminus of the 10-mer peptide, a configuration that would allow the accommodation of longer internal peptide and N-terminal peptide substrates in this conformation of EryP.

Supplementary information

Supplementary Information

Supplementary Tables 1–8 and Figs. 1–15.

Reporting Summary

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhao, C., Sheng, W., Wang, Y. et al. Conformational remodeling enhances activity of lanthipeptide zinc-metallopeptidases. Nat Chem Biol 18, 724–732 (2022). https://doi.org/10.1038/s41589-022-01018-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-022-01018-2

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