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:

Evolution-guided discovery of antibiotics that inhibit peptidoglycan remodelling

Abstract

Addressing the ongoing antibiotic crisis requires the discovery of compounds with novel mechanisms of action that are capable of treating drug-resistant infections1. Many antibiotics are sourced from specialized metabolites produced by bacteria, particularly those of the Actinomycetes family2. Although actinomycete extracts have traditionally been screened using activity-based platforms, this approach has become unfavourable owing to the frequent rediscovery of known compounds. Genome sequencing of actinomycetes reveals an untapped reservoir of biosynthetic gene clusters, but prioritization is required to predict which gene clusters may yield promising new chemical matter2. Here we make use of the phylogeny of biosynthetic genes along with the lack of known resistance determinants to predict divergent members of the glycopeptide family of antibiotics that are likely to possess new biological activities. Using these predictions, we uncovered two members of a new functional class of glycopeptide antibiotics—the known glycopeptide antibiotic complestatin and a newly discovered compound we call corbomycin—that have a novel mode of action. We show that by binding to peptidoglycan, complestatin and corbomycin block the action of autolysins—essential peptidoglycan hydrolases that are required for remodelling of the cell wall during growth. Corbomycin and complestatin have low levels of resistance development and are effective in reducing bacterial burden in a mouse model of skin MRSA infection.

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: Phylogeny-guided discovery of complestatin and corbomycin.
Fig. 2: Corbomycin and complestatin affect peptidoglycan metabolism.
Fig. 3: Corbomycin and complestatin inhibit autolysins.
Fig. 4: Corbomycin and complestatin bind to peptidoglycan.

Similar content being viewed by others

Data availability

Phylogenetic trees, including those of 84 GPA genes and gene segments from 71 BGCs that were analysed in this study, are available at https://github.com/waglecn/GPA_evolution. Streptomyces sp. WAC01529 and Streptomyces sp. WAC01325 genome sequences are available in GenBank with accession numbers NZ_CP029617.1 and NZ_QHKK00000000.1, respectively. Source Data for the animal experiments shown in Extended Data Fig. 8 are included online.

References

  1. Laxminarayan, R. et al. Antibiotic resistance—the need for global solutions. Lancet Infect. Dis. 13, 1057–1098 (2013).

    Article  PubMed  Google Scholar 

  2. Wright, G. D. Something old, something new: revisiting natural products in antibiotic drug discovery. Can. J. Microbiol. 60, 147–154 (2014).

    Article  CAS  PubMed  MathSciNet  Google Scholar 

  3. Kang, H. S. & Brady, S. F. Arixanthomycins A–C: phylogeny-guided discovery of biologically active eDNA-derived pentangular polyphenols. ACS Chem. Biol. 9, 1267–1272 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Peek, J. et al. Rifamycin congeners kanglemycins are active against rifampicin-resistant bacteria via a distinct mechanism. Nat. Commun. 9, 4147 (2018).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  5. Hover, B. M. et al. Culture-independent discovery of the malacidins as calcium-dependent antibiotics with activity against multidrug-resistant Gram-positive pathogens. Nat. Microbiol. 3, 415–422 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Thaker, M. N. et al. Identifying producers of antibacterial compounds by screening for antibiotic resistance. Nat. Biotechnol. 31, 922–927 (2013).

    Article  CAS  PubMed  Google Scholar 

  7. Yan, Y. et al. Resistance-gene-directed discovery of a natural-product herbicide with a new mode of action. Nature 559, 415–418 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  8. Tang, X. et al. Identification of thiotetronic acid antibiotic biosynthetic pathways by target-directed genome mining. ACS Chem. Biol. 10, 2841–2849 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Waglechner, N., McArthur, A. G. & Wright, G. D. Phylogenetic reconciliation reveals the natural history of glycopeptide antibiotic biosynthesis and resistance. Nat. Microbiol. 4, 1862–1871 (2019).

    Article  CAS  PubMed  Google Scholar 

  10. Chiu, H.-T. et al. Molecular cloning and sequence analysis of the complestatin biosynthetic gene cluster. Proc. Natl Acad. Sci. USA 98, 8548–8553 (2001).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  11. Nicolaou, K. C., Boddy, C. N. C., Bräse, S. & Winssinger, N. Chemistry, biology, and medicine of the glycopeptide antibiotics. Angew. Chem. Int. Ed. 38, 2096–2152 (1999).

    Article  CAS  Google Scholar 

  12. Breazzano, S. P. & Boger, D. L. Synthesis and stereochemical determination of complestatin A and B (neuroprotectin A and B). J. Am. Chem. Soc. 133, 18495–18502 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Nazari, B. et al. Nonomuraea sp. ATCC 55076 harbours the largest actinomycete chromosome to date and the kistamicin biosynthetic gene cluster. MedChemComm 8, 780–788 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Naruse, N. et al. New antiviral antibiotics, kistamicins A and B. I. Taxonomy, production, isolation, physico-chemical properties and biological activities. J. Antibiot. 46, 1804–1811 (1993).

    Article  CAS  Google Scholar 

  15. Kwon, Y. J., Kim, H. J. & Kim, W. G. Complestatin exerts antibacterial activity by the inhibition of fatty acid synthesis. Biol. Pharm. Bull. 38, 715–721 (2015).

    Article  CAS  PubMed  Google Scholar 

  16. Park, O. K., Choi, H. Y., Kim, G. W. & Kim, W. G. Generation of new complestatin analogues by heterologous expression of the complestatin biosynthetic gene cluster from Streptomyces chartreusis AN1542. Chem BioChem 17, 1725–1731 (2016).

    CAS  Google Scholar 

  17. Czarny, T. L., Perri, A. L., French, S. & Brown, E. D. Discovery of novel cell wall-active compounds using PywaC, a sensitive reporter of cell wall stress, in the model Gram-positive bacterium Bacillus subtilis. Antimicrob. Agents Chemother. 58, 3261–3269 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. D’Elia, M. A., Millar, K. E., Beveridge, T. J. & Brown, E. D. Wall teichoic acid polymers are dispensable for cell viability in Bacillus subtilis. J. Bacteriol. 188, 8313–8316 (2006).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  19. Kuru, E., Tekkam, S., Hall, E., Brun, Y. V. & Van Nieuwenhze, M. S. Synthesis of fluorescent d-amino acids and their use for probing peptidoglycan synthesis and bacterial growth in situ. Nat. Protoc. 10, 33–52 (2015).

    Article  CAS  PubMed  Google Scholar 

  20. Domínguez-Cuevas, P., Porcelli, I., Daniel, R. A. & Errington, J. Differentiated roles for MreB-actin isologues and autolytic enzymes in Bacillus subtilis morphogenesis. Mol. Microbiol. 89, 1084–1098 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  21. Blackman, S. A., Smith, T. J. & Foster, S. J. The role of autolysins during vegetative growth of Bacillus subtilis 168. Microbiology 144, 73–82 (1998).

    Article  CAS  PubMed  Google Scholar 

  22. Diethmaier, C. et al. A novel factor controlling bistability in Bacillus subtilis: the YmdB protein affects flagellin expression and biofilm formation. J. Bacteriol. 193, 5997–6007 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Meisner, J. et al. FtsEX is required for CwlO peptidoglycan hydrolase activity during cell wall elongation in Bacillus subtilis. Mol. Microbiol. 89, 1069–1083 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Yepes, A. et al. The biofilm formation defect of a Bacillus subtilis flotillin-defective mutant involves the protease FtsH. Mol. Microbiol. 86, 457–471 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Monahan, L. G., Hajduk, I. V., Blaber, S. P., Charles, I. G. & Harry, E. J. Coordinating bacterial cell division with nutrient availability: a role for glycolysis. MBio 5, e00935-14 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Kudrin, P. et al. Subinhibitory concentrations of bacteriostatic antibiotics induce relA-dependent and relA-independent tolerance to β-lactams. Antimicrob. Agents Chemother. 61, e02173-16 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Jolliffe, L. K., Doyle, R. J. & Streips, U. N. The energized membrane and cellular autolysis in Bacillus subtilis. Cell 25, 753–763 (1981).

    Article  CAS  PubMed  Google Scholar 

  28. Calamita, H. G., Ehringer, W. D., Koch, A. L. & Doyle, R. J. Evidence that the cell wall of Bacillus subtilis is protonated during respiration. Proc. Natl Acad. Sci. USA 98, 15260–15263 (2001).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  29. Yamaguchi, H., Furuhata, K., Fukushima, T., Yamamoto, H. & Sekiguchi, J. Characterization of a new Bacillus subtilis peptidoglycan hydrolase gene, yvcE (named cwlO), and the enzymatic properties of its encoded protein. J. Biosci. Bioeng. 98, 174–181 (2004).

    Article  CAS  PubMed  Google Scholar 

  30. Fukushima, T., Yao, Y., Kitajima, T., Yamamoto, H. & Sekiguchi, J. Characterization of new l,d-endopeptidase gene product CwlK (previous YcdD) that hydrolyzes peptidoglycan in Bacillus subtilis. Mol. Genet. Genomics 278, 371–383 (2007).

    Article  CAS  PubMed  Google Scholar 

  31. Davies, A. Peptidoglycan Architecture and Dynamics in Bacillus subtilis. PhD thesis, Univ. Sheffield (2014).

  32. Pletzer, D., Mansour, S. C., Wuerth, K., Rahanjam, N. & Hancock, R. E. W. New mouse model for chronic infections by Gram-negative bacteria enabling the study of anti-infective efficacy and host–microbe interactions. MBio 8, e00140-17 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Chiang, N. et al. Infection regulates pro-resolving mediators that lower antibiotic requirements. Nature 484, 524–528 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  34. Kugelberg, E. et al. Establishment of a superficial skin infection model in mice by using Staphylococcus aureus and Streptococcus pyogenes. Antimicrob. Agents Chemother. 49, 3435–3441 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Guo, J., Ran, H., Zeng, J., Liu, D. & Xin, Z. Tafuketide, a phylogeny-guided discovery of a new polyketide from Talaromyces funiculosus Salicorn 58. Appl. Microbiol. Biotechnol. 100, 5323–5338 (2016).

    Article  CAS  PubMed  Google Scholar 

  36. Atilano, M. L. et al. Bacterial autolysins trim cell surface peptidoglycan to prevent detection by the Drosophila innate immune system. eLife 3, e02277 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Humann, J. & Lenz, L. L. Bacterial peptidoglycan degrading enzymes and their impact on host muropeptide detection. J. Innate Immun. 1, 88–97 (2009).

    Article  CAS  PubMed  Google Scholar 

  38. Skalweit, M. J. & Li, M. Bulgecin A as a β-lactam enhancer for carbapenem-resistant Pseudomonas aeruginosa and carbapenem-resistant Acinetobacter baumannii clinical isolates containing various resistance mechanisms. Drug Des. Devel. Ther. 10, 3013–3020 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. King, A. M. et al. Aspergillomarasmine A overcomes metallo-β-lactamase antibiotic resistance. Nature 510, 503–506 (2014).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  40. Unemo, M. et al. The novel 2016 WHO Neisseria gonorrhoeae reference strains for global quality assurance of laboratory investigations: phenotypic, genetic and reference genome characterization. J. Antimicrob. Chemother. 71, 3096–3108 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Koo, B.-M. et al. Construction and analysis of two genome-scale deletion libraries for Bacillus subtilis. Cell Syst. 4, 291–305 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Blin, K. et al. antiSMASH 4.0-improvements in chemistry prediction and gene cluster boundary identification. Nucleic Acids Res. 45, W36–W41 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Haslinger, K., Peschke, M., Brieke, C., Maximowitsch, E. & Cryle, M. J. X-domain of peptide synthetases recruits oxygenases crucial for glycopeptide biosynthesis. Nature 521, 105–109 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  44. Wade, J. J. & Graver, M. A. A fully defined, clear and protein-free liquid medium permitting dense growth of Neisseria gonorrhoeae from very low inocula. FEMS Microbiol. Lett. 273, 35–37 (2007).

    Article  CAS  PubMed  Google Scholar 

  45. Gonsior, M. et al. Biosynthesis of the peptide antibiotic feglymycin by a linear nonribosomal peptide synthetase mechanism. ChemBioChem 16, 2610–2614 (2015).

    Article  CAS  PubMed  Google Scholar 

  46. Bhavsar, A. P., Zhao, X. & Brown, E. D. Development and characterization of a xylose-dependent system for expression of cloned genes in Bacillus subtilis: conditional complementation of a teichoic acid mutant. Appl. Environ. Microbiol. 67, 403–410 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Liu, T.-Y., Chu, S.-H. & Shaw, G.-C. Deletion of the cell wall peptidoglycan hydrolase gene cwlO or lytE severely impairs transformation efficiency in Bacillus subtilis. J. Gen. Appl. Microbiol. 64, 139–144 (2018).

    Article  CAS  PubMed  Google Scholar 

  48. Ling, L. L. et al. A new antibiotic kills pathogens without detectable resistance. Nature 517, 455–459 (2015).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  49. Deatherage, D. E. & Barrick, J. E. Identification of mutations in laboratory-evolved microbes from next-generation sequencing data using breseq. Methods Mol. Biol. 1151, 165–188 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Dajkovic, A. et al. Hydrolysis of peptidoglycan is modulated by amidation of meso-diaminopimelic acid and Mg2+ in Bacillus subtilis. Mol. Microbiol. 104, 972–988 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Formstone, A. & Errington, J. A magnesium-dependent mreB null mutant: implications for the role of mreB in Bacillus subtilis. Mol. Microbiol. 55, 1646–1657 (2005).

    Article  CAS  PubMed  Google Scholar 

  52. Schaub, R. E. & Dillard, J. P. Digestion of peptidoglycan and analysis of soluble fragments. Bio Protoc. 7, 2438 (2017).

    Article  Google Scholar 

  53. Tiyanont, K. et al. Imaging peptidoglycan biosynthesis in Bacillus subtilis with fluorescent antibiotics. Proc. Natl Acad. Sci. USA 103, 11033–11038 (2006).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  54. Yan, A. et al. Transformation of the anticancer drug doxorubicin in the human gut microbiome. ACS Infect. Dis. 4, 68–76 (2018).

    Article  CAS  PubMed  Google Scholar 

  55. Verbist, L. The antimicrobial activity of fusidic acid. J. Antimicrob. Chemother. 25 (Suppl. B), 1–5 (1990).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank S. French for help with the imaging of vancomycin–BODIPY stained cells, and V. Rao and V. Yarlagadda for MIC measurement of N. gonorrhoeae. This research was funded by a Canadian Institutes of Health Research grant (FRN-148463), the Ontario Research Fund, and by a Canada Research Chair to G.D.W.; the work was also supported by National Institute of Health grants R35GM122556 to Y.V.B. and 5R01GM113172 to M.S.V. and Y.V.B., and a Canada 150 Research Chair in Bacterial Cell Biology to Y.V.B. E.J.C. was supported by a CIHR Vanier Canada Graduate Scholarship. N.W. was supported by a CIHR Canada Graduate Scholarship Doctoral Award.

Author information

Authors and Affiliations

Authors

Contributions

E.J.C., N.W. and G.D.W. conceived the study and designed experiments. N.W. performed phylogenetic analysis, structural predictions and resistant mutant genome analysis. W.W. performed compound purification and structural elucidation. E.J.C., A.A.F.-C. and B.K.C. designed animal studies and A.A.F.-C. performed animal experiments. E.J.C. and D.S. performed peptidoglycan and enzyme purification. E.J.C. and K.K. performed the synthesis of BODIPY derivatives. Y.-P.H. and Y.V.B. designed FDAA studies, M.S.V. provided FDAAs and Y.-P.H. performed experiments. E.J.C. performed all other experiments. E.J.C. and G.D.W. prepared the manuscript.

Corresponding author

Correspondence to Gerard D. Wright.

Ethics declarations

Competing interests

E.J.C., N.W., W.W. and G.D.W. are inventors on a provisional patent application that covers the use of complestatin and corbomycin as antimicrobial therapies.

Additional information

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

Extended data figures and tables

Extended Data Fig. 1 crb and com biosynthetic gene clusters.

a, Labelling scheme for the position of the condensation domains with respect to the centrally encoded 4-hydroxyphenylglycine. b, Full maximum-likelihood phylogenetic tree of condensation domains with labels and colours according to the scheme in a. Figure 1a is an enlargement of the C-2 region. The tree is shown as a rectangular cladogram with a midpoint root. Labels indicate well-supported clades (SH-like support values > 0.89) of condensation domains for the true GPAs, complestatin (comp)-type BGCs and corbomycin (corb)-type BGCs. c, d, Schematic and gene annotations of com, the BGC of Streptomyces sp. WAC01325 corresponding to complestatin (c) and crb, the BGC of Streptomyces sp. WAC01529 corresponding to corbomycin (d). Gene names in c are based on the previously published complestatin BGC in Streptomyces lavendulae10, the gene synteny of which matches exactly except for comE, an MbtH-like protein that appears fused to the C terminus of comD in the sequence of Streptomyces sp. WAC01325. This may be due to sequencing error or incorrect calling of the open reading frame, which we performed using Prodigal as part of the antiSMASH workflow. e, Proposed biosynthetic pathway for corbomycin. A full description is provided in the Supplementary Discussion.

Extended Data Fig. 2 FabI and FabL are not the target of complestatin or corbomycin.

a, b, Dose–response curves of B. subtilis growth against complestatin, corbomycin and triclosan. a, B. subtilis has two enoyl-ACP reductase orthologues, FabI and FabL. fabI and fabL gene deletions (BKK library strains BKK11720 and BKK08650, respectively41) do not affect susceptibility to complestatin or corbomycin, but sensitize cells to triclosan—a strong fabI inhibitor—when fabL is deleted. b, Overexpression of fabI and fabL does not affect susceptibility towards complestatin or corbomycin but protects cells from triclosan when fabL is overexpressed. Genes were expressed from the pSWEET integrative plasmid under the control of a xylose-inducible promoter46. The average and standard deviation of biological triplicates is shown for a and b. c, Supplementing exogenous fatty acid was previously reported to modestly reduce susceptibility to complestatin15, so was repeated here using 2D checkerboards of fatty acids against complestatin, corbomycin or triclosan in S. aureus ATCC 29213. There is no robust reduction in susceptibility to complestatin or corbomycin in contrast to triclosan. Colour is scaled to OD600 with white representing no growth and dark blue representing growth. d, Bright-field microscopy of S. aureus ATCC 29213 grown in 0.6 × MIC triclosan (0.004 μg ml−1), complestatin (0.6 μg ml−1) or corbomycin (0.6 μg ml−1) was used to examine the effect of these antibiotics on phenotype. Whereas complestatin and corbomycin cause a distinct clumping phenotype, triclosan shows no gross morphological defect. All experiments were performed on two independent occasions with similar results.

Extended Data Fig. 3 Peptidoglycan biosynthesis is not inhibited by complestatin or corbomycin.

a, b, Incorporation of fluorescent HADA into peptidoglycan of B. subtilis treated with complestatin or corbomycin for 5 min before continued exposure with HADA for 1 h, or with ampicillin for 1 min before FDAA incubation for 5 min. For a, samples left to right, the number of individual cells measured was n = 243, 242, 299, 271. Whiskers show the minimum and maximum values in the dataset, the box shows the upper and lower quartiles and the line shows the median value. The results in b are quantified in Fig. 2d. c, Incorporation of HALA or HADA, as in a. Cells were treated with 0.6 × MIC and HALA incorporation is shown to represent transpeptidase activity. d, Enumeration of cfu after treatment of B. subtilis 168 for 6 h shows that complestatin and corbomycin are bacteriostatic, similar to the control antibiotic chloramphenicol. Mean and standard deviation cfu values are shown for four independent experimental replicates. e, Washing and subculturing antibiotic-treated cells as prepared in d similarly shows successful outgrowth, indicating a bacteriostatic antibiotic. Mean and standard deviation is shown for three biological replicates. All experiments were performed twice with similar results.

Extended Data Fig. 4 Complestatin and corbomycin have a novel mode of action.

a, Two-dimensional checkerboards in E. coli BW25113 (polymyxcin B nonapeptide; PMBN) or B. subtilis 168 (remainder). For the complestatin versus corbomycin checkerboard, the fractional inhibitory concentration index is 0.75. The colour is scaled to OD600, with white representing no growth and dark blue representing growth. b, Phenotype of B. subtilis 168 grown in 2 × MIC (8 μg ml−1) complestatin or corbomycin for 2 h, or treated with sub-MIC fosfomycin (100 μg ml−1), ampicillin (0.2 μg ml−1), vancomycin (1 μg ml−1) and triclosan (0.5 μg ml−1) until growth to mid-log phase. All experiments were performed on two independent occasions with similar results.

Extended Data Fig. 5 Mutants raised with complestatin or corbomycin.

a, Results of serially passaging S. aureus in sub-MIC concentrations of antibiotic over 25 days. Passaging was performed on two independent lines with similar results. b, MICs of strains COM14 and COR20 against various antibiotics, with respect to B. subtilis 168. The mbl mutation in strain COR14 renders it not viable in 40 μg ml−1 ZnCl2 used for bacitracin, so MIC determination was not possible. c, Susceptibility of B. subtilis strains to moenomycin was assessed by a Kirby–Bauer assay on Mueller–Hinton agar, which results in a zone of sub-lethal growth inhibition in the wild-type strain. d, Bright-field, fluorescence and transmission electron microscopy of resistant mutants COM20 and COR14 in comparison to wild-type and independently generated mutants in relevant genes. e, Fold increase in MIC as compared with wild-type B. subtilis 168 of various independently generated mutants. All knockouts were obtained from a non-essential gene knockout library (see Extended Data Table 2). ND, not determined. *The MIC of the mbl mutant was measured in the presence of 25 mM MgCl2. f, The effects of CwlO knockout and overexpression, either full length or the truncated version identified in COM20, on antibiotic susceptibility was determined. MICs were measured in the presence of 3% xylose to induce expression from Pxyl. Results are described in the Supplementary Discussion. All experiments were performed on at least two independent occasions with similar results.

Extended Data Fig. 6 Complestatin and corbomycin block autolysins in vitro and in vivo.

a, Effect of chloramphenicol on the lysis of B. subtilis. Cells were treated with ampicillin (100 μg ml−1), fosfomycin (50 μg ml−1) or sodium azide (75 mM) to induce lysis, in addition to chloramphenicol, MgSO4 or solvent control. Excess Mg2+ inhibits azide-induced lysis by a mechanism that is not well-understood but is thought to be partially through the modulation of autolysin activity50,51. b, Complestatin and corbomycin antagonism of S. aureus lysis induced by fosfomycin (50 μg ml−1) or ramoplanin (50 μg ml−1). Average values and standard deviation from biological triplicates are plotted for a and b. c, Activity of hen egg white lysozyme on peptidoglycan incubated with 400 μg ml−1 complestatin or corbomycin. The assay was performed in technical triplicate with similar results. The legend refers to both b and c. d, Vancomycin–BODIPY FL (van–BODIPY) staining shows nascent peptidoglycan containing d-Ala-d-Ala in cells treated with 10 × MIC antibiotic for 30 min. All experiments were performed twice with similar results.

Extended Data Fig. 7 Tools used to measure peptidoglycan binding.

a, Activity of CwlO on peptidoglycan that was preincubated with 100 μg ml−1 antibiotic, then washed and resuspended in fresh buffer. The assay was performed in technical triplicate with similar results each time. b, HPLC chromatograms of antibiotic left unbound after incubation with either B. subtilis or S. aureus peptidoglycan. Additional negative control drugs rifampicin (absorbing at 330 nm) and daptomycin (absorbing at 280 nm) were tested under conditions identical to those described in Fig. 4a. Incubation with peptidoglycan does not reduce peak height, indicating the absence of non-specific binding. c, Chemical structures of corbomycin–BODIPY FL (Corb–BODIPY) and complestatin–BODIPY FL (Comp–BODIPY). d, BODIPY derivatives of corbomycin and complestatin induce the same twisted cell chaining as the parent compounds, so retain the same mechanism of action. All experiments were performed on two independent occasions with comparable results.

Extended Data Fig. 8 Complestatin and corbomycin are effective in a mouse skin infection model.

A topical application of antibiotic in petroleum jelly (%w/w) was used to treat MRSA in a skin infection model using neutropenic mice. Mice were assessed 33 h after infection and compared to those receiving the vehicle control (petroleum jelly containing 10% DMSO). a, The cfu per gram of tissue was determined. Significance was tested by one-way ANOVA on ranks with Kruskal–Wallis test, P values versus vehicle as follows: 1% complestatin, 0.0062; 1% corbomycin, 0.0157; 0.25% fusidic acid, 0.0003. b, The change in weight (%) after infection compared to pre-infection was measured for each mouse. Significance was tested by a two-sided unpaired Mann–Whitney t-test, P-values versus vehicle as follows: 1% complestatin, 0.0016; 1% corbomycin, 0.016; 0.25% fusidic acid, 0.0028. For a, b, * denotes significance (P < 0.05), and whiskers show the minimum and maximum values of the dataset, the box shows upper and lower quartiles and the line shows the median value. c, Visual inspection of the wounds of representative mice in each treatment group. Each treatment condition was tested with replicate mice (n = 6 complestatin, corbomycin, fusidic acid; n = 9 vehicle) divided between two independent experiments.

Source data

Extended Data Table 1 MICs of complestatin and corbomycin
Extended Data Table 2 Strains used in this study

Supplementary information

Supplementary Information

This file contains supplementary discussion, Fig. 1-6 and table 1. Discussion is of the structural characterization of corbomycin, the biosynthetic pathway and gene cluster for corbomycin, and further analysis of resistant mutants raised to complestatin and corbomycin. Figures and table are for NMR spectra of corbomycin.

Reporting Summary

Supplementary Data

This file contains complestatin and corbomycin structures.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Culp, E.J., Waglechner, N., Wang, W. et al. Evolution-guided discovery of antibiotics that inhibit peptidoglycan remodelling. Nature 578, 582–587 (2020). https://doi.org/10.1038/s41586-020-1990-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-020-1990-9

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research