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:

Commensal production of a broad-spectrum and short-lived antimicrobial peptide polyene eliminates nasal Staphylococcus aureus

Abstract

Antagonistic bacterial interactions often rely on antimicrobial bacteriocins, which attack only a narrow range of target bacteria. However, antimicrobials with broader activity may be advantageous. Here we identify an antimicrobial called epifadin, which is produced by nasal Staphylococcus epidermidis IVK83. It has an unprecedented architecture consisting of a non-ribosomally synthesized peptide, a polyketide component and a terminal modified amino acid moiety. Epifadin combines a wide antimicrobial target spectrum with a short life span of only a few hours. It is highly unstable under in vivo-like conditions, potentially as a means to limit collateral damage of bacterial mutualists. However, Staphylococcus aureus is eliminated by epifadin-producing S. epidermidis during co-cultivation in vitro and in vivo, indicating that epifadin-producing commensals could help prevent nasal S. aureus carriage. These insights into a microbiome-derived, previously unknown antimicrobial compound class suggest that limiting the half-life of an antimicrobial may help to balance its beneficial and detrimental activities.

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: Plasmid and epifadin BGC composition, biosynthetic modules and domain organization of the epifadin NRPS–PKS enzymes of IVK83.
Fig. 2: High instability of epifadin under in vivo-like conditions.
Fig. 3: Detection, isolation and absorption spectrum of epifadin.
Fig. 4: Molecular mass and structure of epifadin.
Fig. 5: Broad antimicrobial activity of epifadin-producing S. epidermidis IVK83.
Fig. 6: Epifadin-producing S. epidermidis IVK83 restricts S. aureus growth in vitro and in vivo in cotton rats.

Similar content being viewed by others

Data availability

All data supporting the findings of this study are available within the paper, its extended data or supplementary information. Whole-genome sequencing data obtained for S. epidermidis IVK83 were deposited in the NCBI Sequence Read Archive (genome available under accession number CP088002, plasmid pIVK83 under CP088003). Sequence of strain S. epidermidis B155 (Liverpool, UK) was deposited as BioSample SAMEA12384066 (BioProject PRJEB50307). Representative microscopy images are included in the extended data figures and the supplementary videos, which were deposited at Figshare (https://doi.org/10.6084/m9.figshare.24125589). NMR data were deposited at nmrXive and are available under the project identifier NMRXIV:P18 (https://doi.org/10.57992/nmrxiv.p18; https://nmrxiv.org/P18). Source data are provided with this paper.

References

  1. Bomar, L., Brugger, S. D. & Lemon, K. P. Bacterial microbiota of the nasal passages across the span of human life. Curr. Opin. Microbiol. 41, 8–14 (2018).

    PubMed  Google Scholar 

  2. Paller, A. S. et al. The microbiome in patients with atopic dermatitis. J. Allergy Clin. Immunol. 143, 26–35 (2019).

    PubMed  Google Scholar 

  3. Totté, J. E. et al. Prevalence and odds of Staphylococcus aureus carriage in atopic dermatitis: a systematic review and meta-analysis. Br. J. Dermatol. 175, 687–695 (2016).

    PubMed  Google Scholar 

  4. Lee, Y. B., Byun, E. J. & Kim, H. S. Potential role of the microbiome in acne: a comprehensive review. J. Clin. Med. 8, 987 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Lee, A. S. et al. Methicillin-resistant Staphylococcus aureus. Nat. Rev. Dis. Prim. 4, 18033 (2018).

    PubMed  Google Scholar 

  6. Heilbronner, S., Krismer, B., Brotz-Oesterhelt, H. & Peschel, A. The microbiome-shaping roles of bacteriocins. Nat. Rev. Microbiol. https://doi.org/10.1038/s41579-021-00569-w (2021).

  7. Kloos, W. E. & Musselwhite, M. S. Distribution and persistence of Staphylococcus and Micrococcus species and other aerobic bacteria on human skin. Appl. Microbiol. 30, 381–385 (1975).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Coates, R., Moran, J. & Horsburgh, M. J. Staphylococci: colonizers and pathogens of human skin. Future Microbiol. 9, 75–91 (2014).

    CAS  PubMed  Google Scholar 

  9. Du, X. et al. Staphylococcus epidermidis clones express Staphylococcus aureus-type wall teichoic acid to shift from a commensal to pathogen lifestyle. Nat. Microbiol. 6, 757–768 (2021).

    CAS  PubMed  Google Scholar 

  10. Cogen, A. L. et al. Staphylococcus epidermidis antimicrobial delta-toxin (phenol-soluble modulin-gamma) cooperates with host antimicrobial peptides to kill group A Streptococcus. PLoS ONE 5, e8557 (2010).

    PubMed  PubMed Central  Google Scholar 

  11. Cogen, A. L. et al. Selective antimicrobial action is provided by phenol-soluble modulins derived from Staphylococcus epidermidis, a normal resident of the skin. J. Invest. Dermatol. 130, 192–200 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Otto, M. Phenol-soluble modulins. Int. J. Med. Microbiol. 304, 164–169 (2014).

    CAS  PubMed  Google Scholar 

  13. Janek, D., Zipperer, A., Kulik, A., Krismer, B. & Peschel, A. High frequency and diversity of antimicrobial activities produced by nasal Staphylococcus strains against bacterial competitors. PLoS Pathog. 12, e1005812 (2016).

    PubMed  PubMed Central  Google Scholar 

  14. O’Sullivan, J. N., Rea, M. C., O’Connor, P. M., Hill, C. & Ross, R. P. Human skin microbiota is a rich source of bacteriocin-producing staphylococci that kill human pathogens. FEMS Microbiol. Ecol. https://doi.org/10.1093/femsec/fiy241 (2018).

  15. Götz, F., Perconti, S., Popella, P., Werner, R. & Schlag, M. Epidermin and gallidermin: staphylococcal lantibiotics. Int. J. Med. Microbiol. 304, 63–71 (2014).

    PubMed  Google Scholar 

  16. Ekkelenkamp, M. B. et al. Isolation and structural characterization of epilancin 15X, a novel lantibiotic from a clinical strain of Staphylococcus epidermidis. FEBS Lett. 579, 1917–1922 (2005).

    CAS  PubMed  Google Scholar 

  17. Molloy, E. M., Cotter, P. D., Hill, C., Mitchell, D. A. & Ross, R. P. Streptolysin S-like virulence factors: the continuing sagA. Nat. Rev. Microbiol. 9, 670–681 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Van Tyne, D., Martin, M. J. & Gilmore, M. S. Structure, function, and biology of the Enterococcus faecalis cytolysin. Toxins 5, 895–911 (2013).

    PubMed  PubMed Central  Google Scholar 

  19. Cascales, E. et al. Colicin biology. Microbiol. Mol. Biol. Rev. 71, 158–229 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Baquero, F., Lanza, V. F., Baquero, M.-R., del Campo, R. & Bravo-Vázquez, D. A. Microcins in Enterobacteriaceae: peptide antimicrobials in the eco-active intestinal chemosphere. Front. Microbiol. https://doi.org/10.3389/fmicb.2019.02261 (2019).

  21. Klein, T. A., Ahmad, S. & Whitney, J. C. Contact-dependent interbacterial antagonism mediated by protein secretion machines. Trends Microbiol. 28, 387–400 (2020).

    CAS  PubMed  Google Scholar 

  22. Cao, Z., Casabona, M. G., Kneuper, H., Chalmers, J. D. & Palmer, T. The type VII secretion system of Staphylococcus aureus secretes a nuclease toxin that targets competitor bacteria. Nat. Microbiol. 2, 16183 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Acosta, E. M. et al. Bacterial DNA on the skin surface overrepresents the viable skin microbiome. eLife https://doi.org/10.7554/eLife.87192.1 (2021).

  24. Brüggemann, H. et al. Staphylococcus saccharolyticus isolated from blood cultures and prosthetic joint infections exhibits excessive genome decay. Front. Microbiol. 10, 478 (2019).

    PubMed  PubMed Central  Google Scholar 

  25. Zipperer, A. et al. Human commensals producing a novel antibiotic impair pathogen colonization. Nature 535, 511–516 (2016).

    CAS  PubMed  Google Scholar 

  26. Schnell, N. et al. Prepeptide sequence of epidermin, a ribosomally synthesized antibiotic with four sulphide-rings. Nature 333, 276–278 (1988).

    CAS  PubMed  Google Scholar 

  27. Rogers, L. A. & Whittier, E. O. Limiting factors in the lactic fermentation. J. Bacteriol. 16, 211–229 (1928).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Blin, K. et al. antiSMASH 5.0: updates to the secondary metabolite genome mining pipeline. Nucleic Acids Res. 47, W81–W87 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Gao, X. et al. Cyclization of fungal nonribosomal peptides by a terminal condensation-like domain. Nat. Chem. Biol. 8, 823–830 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Law, B. J. C. et al. A vitamin K-dependent carboxylase orthologue is involved in antibiotic biosynthesis. Nat. Catal. 1, 977–984 (2018).

    CAS  Google Scholar 

  31. Fujii, I. Functional analysis of fungal polyketide biosynthesis genes. J. Antibiot. 63, 207–218 (2010).

    CAS  Google Scholar 

  32. Ng, B. G., Han, J. W., Lee, D. W., Choi, G. J. & Kim, B. S. The chejuenolide biosynthetic gene cluster harboring an iterative trans-AT PKS system in Hahella chejuensis strain MB-1084. J. Antibiot. 71, 495–505 (2018).

    CAS  Google Scholar 

  33. Cavassin, F. B., Baú-Carneiro, J. L., Vilas-Boas, R. R. & Queiroz-Telles, F. Sixty years of amphotericin B: an overview of the main antifungal agent used to treat invasive fungal infections. Infect. Dis. Ther. 10, 115–147 (2021).

    PubMed  PubMed Central  Google Scholar 

  34. Rapun-Araiz, B. et al. Systematic reconstruction of the complete two-component sensorial network in Staphylococcus aureus. mSystems 5, e00511-20 (2020).

    PubMed  PubMed Central  Google Scholar 

  35. Krismer, B. et al. Nutrient limitation governs Staphylococcus aureus metabolism and niche adaptation in the human nose. PLoS Pathog. 10, e1003862 (2014).

    PubMed  PubMed Central  Google Scholar 

  36. Baur, S. et al. A nasal epithelial receptor for Staphylococcus aureus WTA governs adhesion to epithelial cells and modulates nasal colonization. PLoS Pathog. 10, e1004089 (2014).

    PubMed  PubMed Central  Google Scholar 

  37. Donia, M. S. et al. A systematic analysis of biosynthetic gene clusters in the human microbiome reveals a common family of antibiotics. Cell 158, 1402–1414 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Sugimoto, Y. et al. A metagenomic strategy for harnessing the chemical repertoire of the human microbiome. Science 366, eaax9176 (2019).

    CAS  PubMed  Google Scholar 

  39. Donia, M. S. & Fischbach, M. A. Human microbiota. Small molecules from the human microbiota. Science 349, 1254766 (2015).

    PubMed  PubMed Central  Google Scholar 

  40. Myrtle, J. D., Beekman, A. M. & Barrow, R. A. Ravynic acid, an antibiotic polyeneyne tetramic acid from Penicillium sp. elucidated through synthesis. Org. Biomol. Chem. 14, 8253–8260 (2016).

    CAS  PubMed  Google Scholar 

  41. Uranga, C., Nelson, K. E., Edlund, A. & Baker, J. L. Tetramic acids mutanocyclin and reutericyclin A, produced by Streptococcus mutans strain B04Sm5 modulate the ecology of an in vitro oral biofilm. Front. Oral Health 2, 796140 (2021).

    PubMed  Google Scholar 

  42. Ganzle, M. G. & Vogel, R. F. Studies on the mode of action of reutericyclin. Appl. Environ. Microbiol. 69, 1305–1307 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Li, Z. R. et al. Mutanofactin promotes adhesion and biofilm formation of cariogenic Streptococcus mutans. Nat. Chem. Biol. 17, 576–584 (2021).

    CAS  PubMed  Google Scholar 

  44. Moldenhauer, J., Chen, X.-H., Borriss, R. & Piel, J. Biosynthesis of the antibiotic Bacillaene, the product of a giant polyketide synthase complex of the trans-AT family. Angew. Chem. Int. Ed. 46, 8195–8197 (2007).

    CAS  Google Scholar 

  45. Pacheco, A. R. & Segrè, D. A multidimensional perspective on microbial interactions. FEMS Microbiol. Lett. 366, fnz125 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Byrd, A. L., Belkaid, Y. & Segre, J. A. The human skin microbiome. Nat. Rev. Microbiol. 16, 143–155 (2018).

    CAS  PubMed  Google Scholar 

  47. Garcia-Bayona, L. & Comstock, L. E. Bacterial antagonism in host-associated microbial communities. Science 361, eaat2456 (2018).

    PubMed  Google Scholar 

  48. Moghadam, Z. M., Henneke, P. & Kolter, J. From flies to men: ROS and the NADPH oxidase in phagocytes. Front. Cell Dev. Biol. 9, 628991 (2021).

    PubMed  PubMed Central  Google Scholar 

  49. Khayatt, B. I., van Noort, V. & Siezen, R. J. The genome of the plant-associated lactic acid bacterium Lactococcus lactis KF147 harbors a hybrid NRPS–PKS system conserved in strains of the dental cariogenic Streptococcus mutans. Curr. Microbiol. 77, 136–145 (2020).

    CAS  PubMed  Google Scholar 

  50. Wu, C. et al. Genomic island TnSmu2 of Streptococcus mutans harbors a nonribosomal peptide synthetase-polyketide synthase gene cluster responsible for the biosynthesis of pigments involved in oxygen and H2O2 tolerance. Appl. Environ. Microbiol. 76, 5815–5826 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Neubauer, H., Pantel, I. & Götz, F. Molecular characterization of the nitrite-reducing system of Staphylococcus carnosus. J. Bacteriol. 181, 1481–1488 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Gutierrez, J. A. et al. Insertional mutagenesis and recovery of interrupted genes of Streptococcus mutans by using transposon Tn917: preliminary characterization of mutants displaying acid sensitivity and nutritional requirements. J. Bacteriol. 178, 4166–4175 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Youngman, P. J., Perkins, J. B. & Losick, R. Genetic transposition and insertional mutagenesis in Bacillus subtilis with Streptococcus faecalis transposon Tn917. Proc. Natl Acad. Sci. USA 80, 2305–2309 (1983).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Zerbino, D. R. & Birney, E. Velvet: algorithms for de novo short read assembly using de Bruijn graphs. Genome Res. 18, 821–829 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Darling, A. E., Mau, B. & Perna, N. T. progressiveMauve: multiple genome alignment with gene gain, loss and rearrangement. PLoS ONE 5, e11147 (2010).

    PubMed  PubMed Central  Google Scholar 

  56. Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Biopharma Bioinformatics. FastQC https://www.bioinformatics.babraham.ac.uk/projects/fastqc/ (2018).

  58. Ewels, P., Magnusson, M., Lundin, S. & Kaller, M. MultiQC: summarize analysis results for multiple tools and samples in a single report. Bioinformatics 32, 3047–3048 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Wick, R. R., Judd, L. M., Gorrie, C. L. & Holt, K. E. Unicycler: resolving bacterial genome assemblies from short and long sequencing reads. PLoS Comput. Biol. 13, e1005595 (2017).

    PubMed  PubMed Central  Google Scholar 

  60. Prjibelski, A., Antipov, D., Meleshko, D., Lapidus, A. & Korobeynikov, A. Using SPAdes de novo assembler. Curr. Protoc. Bioinforma. 70, e102 (2020).

    CAS  Google Scholar 

  61. Seemann, T. Prokka: rapid prokaryotic genome annotation. Bioinformatics 30, 2068–2069 (2014).

    CAS  PubMed  Google Scholar 

  62. Geiger, T. et al. The stringent response of Staphylococcus aureus and its impact on survival after phagocytosis through the induction of intracellular PSMs expression. PLoS Pathog. 8, e1003016 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Brückner, R. Gene replacement in Staphylococcus carnosus and Staphylococcus xylosus. FEMS Microbiol. Lett. 151, 1–8 (1997).

    PubMed  Google Scholar 

  64. Bruckner, R. A series of shuttle vectors for Bacillus subtilis and Escherichia coli. Gene 122, 187–192 (1992).

    CAS  PubMed  Google Scholar 

  65. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    CAS  PubMed  Google Scholar 

  66. Saising, J. et al. Rhodomyrtone (Rom) is a membrane-active compound. Biochim. Biophys. Acta Biomembr. 1860, 1114–1124 (2018).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank D. Belikova, V. Augsburger and J. Straetner for excellent technical support, M. Hamburger (Pharmaceutical Biology, University of Basel, Switzerland) for providing authentic sample of militarinone C, and A. Tooming-Klunderud (Center for Ecological and Evolutionary Synthesis, Department of Biosciences, University of Oslo, Norway) for PacBio sequencing of strain IVK83. The sequencing company MicrobesNG (Birmingham, UK) is supported by the Biotechnology and Biological Sciences Research Council; grant number BB/L024209/1). The authors’ work is financed by grants from Deutsche Forschungsgemeinschaft (DFG) TRR261 (A.P., H.B.-O. and S.G.; project ID 398967434), GRK1708 (S.G., H.B.-O. and A.P.) and Cluster of Excellence EXC2124 Controlling Microbes to Fight Infection (CMFI, S.G., B.K., H.B.-O. and A.P.; project ID 390838134), TRR156 (A.P.; project ID 246807620), and ZUK 63 (N.A.S.); from the German Center of Infection Research (DZIF) to B.K., H.B.-O. and A.P.; from the Novo Nordisk Foundation (T.W., project ID NNF20CC0035580); from the German Ministry of Research and Education (BMBF) Culture Challenge to A.P.; and from the European Innovative Medicines Initiate IMI (COMBACTE) to A.P. We acknowledge support by the High Performance and Cloud Computing Group at the Zentrum für Datenverarbeitung of the University of Tübingen, the state of Baden-Württemberg through bwHPC and the German Research Foundation (DFG) through grant no. INST 37/935-1 FUGG.

Author information

Authors and Affiliations

Authors

Contributions

B.O.T.S. performed and analysed most of the bacteriological, molecular and compound isolation experiments with help by D.J., S.K. and B.K., who originally isolated strain IVK83; animal experiments were performed by B.O.T.S. and B.K.; T.D., N.A.S. and S.G. elucidated the structure of epifadin with support from J.M.B.-B.; T.D. performed total syntheses of all tetrapeptides, their purification and chemical analyses; A.B. analysed epifadin toxicity and membrane potential effects; J.B. performed all microscopic experiments; A.M.A.E. performed the bioinformatic search for epifadin-like BGCs; M. Li., M.J.H. and S.K. identified and provided epifadin-producing S. epidermidis strains.; N.A. and M.J.H. performed the experimental evolution and analysed the epifadin-resistant mutants; S.J.J. and M. Lämmerhofer confirmed the absolute configuration of the tetrapeptide via chiral HPLC; T.W. analysed the potential epifadin biosynthesis enzymes; H.B.-O., S.G., B.K. and A.P. supervised the experiments and wrote the manuscript.

Corresponding authors

Correspondence to Stephanie Grond or Bernhard Krismer.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Microbiology thanks Rolf Müller, James O’Gara and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

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 Comparison of MS/MS spectra of the synthetic and natural peptide amide fragments of epifadin.

a, MS/MS spectrum of the natural peptide amide after decomposition of epifadin. b, MS/MS spectrum of the synthetic peptide amide 2. c, Fragmentation pattern of the synthetic and natural peptide amides 2. Fragmentation pattern for the peptide amide 2 is shown in black. F, phenylalanine; D, aspartate; N, asparagine; CO, carbon monoxide; NH3, ammonia.

Extended Data Fig. 2 Proton NMR spectrum of the synthetic peptide amide FfDn-NH2 (DMSO-d6, 600MHz, 303K).

The integrals of the proton signals are depicted as black curves. The scale shows the chemical shift δ in parts per million (ppm).

Extended Data Fig. 3 1H-1H-ROESY NMR spectrum of the purified epifadin in DMSO-d6 (700 MHz, 303 K).

The red circles highlight coupling between the NH-proton and the protons of the methyl group of the alanine residue (9.14 ppm/1.95 ppm) and to the proton of the adjacent methine group (9.14 ppm/6.77 ppm). Also, the coupling of the protons of the methyl group from the alanine residue to the methine group is shown (6.77 ppm/1.95 ppm).

Extended Data Fig. 4 1H NMR spectrum (DMSO-d6, 700 MHz, 303 K) of purified epifadin and its decomposition analyzed by HPLC-MS.

a, DMSO-d6 signal at 2.50 ppm as reference. The integrals of the proton signals are depicted as black curves. The scale shows the chemical shift δ in parts per million (ppm). b,c, The epifadin-enriched material was dissolved in a mixture of acetonitrile and water (1:1) with 0.05% trifluoroacetic acid, resulting in a concentration of 0.2 mg/mL and analyzed by HPLC-ESI-TOF-high resolution MS. The extracted ion chromatograms (EICs) of epifadin (C51H61N7O12 [M+H]+, m/z 964.4451 ± 0.005) are depicted in red (retention time 15.2 min and 15.6 min) and the base peak chromatograms (BPCs) in gray. EICs of the peptide amide (C26H32N6O7 [M+H]+, m/z 541.2405 ± 0.005) are depicted in blue (retention time 7.4 min) accumulating by strong decomposition of epifadin in the mentioned solvent after storage at −20 °C. b, analyzed after purification. c, Analyzed after 14 days of storage at −20 °C.

Extended Data Fig. 5 Deduced fragmentation pattern for the peptide amide and the PKS/NRPS moiety.

From a six-membered transition state a rearrangement results in a neutral loss of the peptide amide moiety. The newly formed allene (m/z 424.2131) decomposes into further fragments.

Extended Data Fig. 6 MS/MS spectra of epifadin showing fragmentation products from ionization in MS.

a, The mass of 964 Da corresponds to the intact proton adduct (m/z 964.4) of epifadin. 524 Da (m/z 524.2) corresponds to the proton adduct of the tetrapeptide EfiA product, and the mass of 441 Da (m/z 441.2) is assigned to the proton adduct of the EfiBCDE product (expansion shows also minor signals of peptide fragments). [M+H]+, monoisotopic positively charged ion; F, phenylalanine; D, aspartate; N, asparagine; CO, carbon monoxide. b, The fragmentation pattern for the peptide moiety in epifadin is shown. Numbering of amino acids and carbon atoms of PKS chain in red.

Extended Data Fig. 7 Epifadin is bactericidal for susceptible bacterial cells but does not inhibit mammalian cells.

a, Time-dependent elimination of S. aureus by epifadin. Incubation of S. aureus USA300 LAC with epifadin concentrations of 24 µg/mL and 12 µg/mL led to a fast decline of CFUs reaching the detection limit of 1 × 103 CFU/mL after 210 min. Data represent means with SEM of three independent experiments. b,c, Cell viability assay. HeLa cells incubated with epifadin do not show increased cell death compared to mock-treated cells (DMSO treatment set as 100%) even at high concentrations of 12 µg/mL. Cycloheximide (CHM) was included as a positive control. Only at concentration of 24 µg/mL, epifadin shows a significant effect on cell viability, still leaving 84% of HeLa cells intact. Data points represent the mean ± SD of three independent experiments. Significant differences between lowest compound concentrations and higher concentrations were analyzed by one-way ANOVA (*P < 0.05; **P < 0.01; ***P < 0.001; ****P < 0.0001). Exact p values for the CHM treated HeLa cells were 0.0192 (6.25 × 10−2 µg/ml), 0.0009 (0.125 µg/ml), 0.0001 (0.25 µg/ml).

Source data

Extended Data Fig. 8 Epifadin-producing S. epidermidis IVK83 restricts S. aureus growth in vitro.

ac, in vitro competition assays in TSB. a, S. aureus growth is inhibited by IVK83 wild type (grey or light blue bars, respectively) already after 24 h of incubation in TSB inoculated at ratios of ~50:50. b, in contrast, the mutant IVK83 ΔefiTP is overgrown by S. aureus over time when inoculated at a 50:50 ratio. c, Complemented strain overgrew S. aureus for 48 h, after 72 h, ratio of complemented strain and S. aureus were similar to starting conditions. Data points represent mean value ± SD of three independent experiments. Significant differences between the starting condition and the indicated time points were analyzed by one-way ANOVA (**P < 0.01; ***P < 0.001; ****P < 0.0001).

Source data

Extended Data Fig. 9 Structures of semi-synthetic derivatives of peptide amide and MS/MS spectra.

a) methyl ester of natural peptide amide and b) acetylated natural peptide amide.

Extended Data Fig. 10 Epifadin leads to depolarization of the bacterial membrane and rapid cell lysis of S. aureus.

(a) S. aureus USA300 JE2 cells were applied to an agarose pad, onto which 2 µL of extracts (50 mg/mL) of the epifadin producer IVK83 (left) or ΔefiTP (right) had been previously spotted. Image acquisition was started in the surrounding of the respective extract spot 15 min after S. aureus application. The agarose contained FM4-64 (red, 0.25 µg/mL, membrane dye) and Sytox Green (green, 0.25 µM, only visible upon membrane barrier malfunction). Representative micrographs are depicted, all adjusted in the Sytox green channel to the same settings for qualitative comparison. White scale bar, 5µm. (b) Time-resolved effects of extracts of IVK83 wild type and ΔefiTP on the membrane potential of S. aureus NCTC8325 as monitored by DiOC2(3) staining. CCCP (5 µM), positive control; DMSO, untreated control. Mean and s.d. of two biological with two technical replicates (n = 4). Black arrow, time point of compound addition.

Source data

Supplementary information

Supplementary Information

Supplementary Figs. 1–8, information and Tables 1–7.

Reporting Summary

Peer Review File

Supplementary Video 1

Agarose-embedded S. aureus cells lyse within 15 min after contact with epifadin-containing DMSO extract.

Supplementary Video 2

Agarose-embedded S. aureus cells after contact with an epifadin-free DMSO extract. Growth and cell division is seen within 4 h.

Source data

Source Data Fig. 2

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

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

Torres Salazar, B.O., Dema, T., Schilling, N.A. et al. Commensal production of a broad-spectrum and short-lived antimicrobial peptide polyene eliminates nasal Staphylococcus aureus. Nat Microbiol 9, 200–213 (2024). https://doi.org/10.1038/s41564-023-01544-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-023-01544-2

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