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:

Inhibition of translation termination by the antimicrobial peptide Drosocin

Abstract

The proline-rich antimicrobial peptide (PrAMP) Drosocin (Dro) from fruit flies shows sequence similarity to other PrAMPs that bind to the ribosome and inhibit protein synthesis by varying mechanisms. The target and mechanism of action of Dro, however, remain unknown. Here we show that Dro arrests ribosomes at stop codons, probably sequestering class 1 release factors associated with the ribosome. This mode of action is comparable to that of apidaecin (Api) from honeybees, making Dro the second member of the type II PrAMP class. Nonetheless, analysis of a comprehensive library of endogenously expressed Dro mutants shows that the interactions of Dro and Api with the target are markedly distinct. While only a few C-terminal amino acids of Api are critical for binding, the interaction of Dro with the ribosome relies on multiple amino acid residues distributed throughout the PrAMP. Single-residue substitutions can substantially enhance the on-target activity of Dro.

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: The PrAMP Dro stalls ribosomes at stop codons during in vitro translation.
Fig. 2: Dro induces stop codon readthrough.
Fig. 3: Endogenous expression of Dro is toxic for bacteria due to the interactions of the PrAMP with the ribosome and RFs.
Fig. 4: The growth inhibition effect of Dro mutants expressed in cells.
Fig. 5: Synthetic Dro variants with single amino acid substitutions preserve their ability to interfere with translation termination.

Similar content being viewed by others

Data availability

The uncropped gels, the raw data used for calculating the DS and statistics data for Fig. 2b and Extended Data Fig. 1a are shown in the Source Data file associated with the manuscript. As raw sequencing data for the mutant Dro gene libraries do not represent genomic, RNA-seq or Ribo-seq results, they were not deposited to the public databases but are available from the corresponding authors upon request. Source data are provided with this paper.

References

  1. Lazzaro, B. P., Zasloff, M. & Rolff, J. Antimicrobial peptides: application informed by evolution. Science 368, eaau5480 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Cardoso, M. H. et al. Non-lytic antibacterial peptides that translocate through bacterial membranes to act on intracellular targets. Int. J. Mol. Sci. 20, 4877 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Otvos, L. Jr. The short proline-rich antibacterial peptide family. Cell. Mol. Life Sci. 59, 1138–1150 (2002).

    CAS  PubMed  Google Scholar 

  4. Graf, M. & Wilson, D. N. Intracellular antimicrobial peptides targeting the protein synthesis machinery. Adv. Exp. Med. Biol. 1117, 73–89 (2019).

    CAS  PubMed  Google Scholar 

  5. Scocchi, M., Tossi, A. & Gennaro, R. Proline-rich antimicrobial peptides: converging to a non-lytic mechanism of action. Cell. Mol. Life Sci. 68, 2317–2330 (2011).

    CAS  PubMed  Google Scholar 

  6. Benincasa, M. et al. The proline-rich peptide Bac7(1-35) reduces mortality from Salmonella typhimurium in a mouse model of infection. BMC Microbiol. 10, 178 (2010).

    PubMed  PubMed Central  Google Scholar 

  7. Baliga, C. et al. Charting the sequence–activity landscape of peptide inhibitors of translation termination. Proc. Natl Acad. Sci. USA 118, e2026465118 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Bulet, P. et al. A novel inducible antibacterial peptide of Drosophila carries an O-glycosylated substitution. J. Biol. Chem. 268, 14893–14897 (1993).

    CAS  PubMed  Google Scholar 

  9. Cociancich, S. et al. Novel inducible antibacterial peptides from a hemipteran insect, the sap-sucking bug Pyrrhocoris apterus. Biochem. J. 300, 567–575 (1994).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Mattiuzzo, M. et al. Role of the Escherichia coli SbmA in the antimicrobial activity of proline-rich peptides. Mol. Microbiol. 66, 151–163 (2007).

    CAS  PubMed  Google Scholar 

  11. Krizsan, A., Knappe, D. & Hoffmann, R. Influence of the yjiL-mdtM gene cluster on the antibacterial activity of proline-rich antimicrobial peptides overcoming Escherichia coli resistance induced by the missing SbmA transporter system. Antimicrob. Agents Chemother. 59, 5992–5998 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Otvos, L. Jr. et al. Interaction between heat shock proteins and antimicrobial peptides. Biochemistry 39, 14150–14159 (2000).

    CAS  PubMed  Google Scholar 

  13. Scocchi, M. et al. The proline-rich antibacterial peptide Bac7 binds to and inhibits in vitro the molecular chaperone DnaK. Int. J. Pept. Res. Ther. 15, 147–155 (2009).

    CAS  Google Scholar 

  14. Czihal, P. et al. Api88 is a novel antibacterial designer peptide to treat systemic infections with multidrug-resistant Gram-negative pathogens. ACS Chem. Biol. 7, 1281–1291 (2012).

    CAS  PubMed  Google Scholar 

  15. Krizsan, A. et al. Insect-derived proline-rich antimicrobial peptides kill bacteria by inhibiting bacterial protein translation at the 70S ribosome. Angew. Chem. Int. Ed. Engl. 53, 12236–12239 (2014).

    CAS  PubMed  Google Scholar 

  16. Krizsan, A., Prahl, C., Goldbach, T., Knappe, D. & Hoffmann, R. Short proline-rich antimicrobial peptides inhibit either the bacterial 70S ribosome or the assembly of its large 50S subunit. Chem. Bio. Chem. 16, 2304–2308 (2015).

    CAS  PubMed  Google Scholar 

  17. Seefeldt, A. C. et al. The proline-rich antimicrobial peptide Onc112 inhibits translation by blocking and destabilizing the initiation complex. Nat. Struct. Mol. Biol. 22, 470–475 (2015).

    CAS  PubMed  Google Scholar 

  18. Roy, R. N., Lomakin, I. B., Gagnon, M. G. & Steitz, T. A. The mechanism of inhibition of protein synthesis by the proline-rich peptide oncocin. Nat. Struct. Mol. Biol. 22, 466–469 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Gagnon, M. G. et al. Structures of proline-rich peptides bound to the ribosome reveal a common mechanism of protein synthesis inhibition. Nucleic Acids Res. 44, 2439–2450 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Seefeldt, A. C. et al. Structure of the mammalian antimicrobial peptide Bac7(1-16) bound within the exit tunnel of a bacterial ribosome. Nucleic Acids Res. 44, 2429–2438 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Florin, T. et al. An antimicrobial peptide that inhibits translation by trapping release factors on the ribosome. Nat. Struct. Mol. Biol. 24, 752–757 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Weaver, J., Mohammad, F., Buskirk, A. R. & Storz, G. Identifying small proteins by ribosome profiling with stalled initiation complexes. mBio 10, e02819–18 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Peng, S. et al. Mechanism of actions of Oncocin, a proline-rich antimicrobial peptide, in early elongation revealed by single-molecule FRET. Protein Cell 9, 890–895 (2018).

    CAS  PubMed  Google Scholar 

  24. Berthold, N. et al. Novel apidaecin 1b analogs with superior serum stabilities for treatment of infections by Gram-negative pathogens. Antimicrob. Agents Chemother. 57, 402–409 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Graf, M. et al. Visualization of translation termination intermediates trapped by the Apidaecin 137 peptide during RF3-mediated recycling of RF1. Nat. Commun. 9, 3053 (2018).

    PubMed  PubMed Central  Google Scholar 

  26. Mangano, K. et al. Genome-wide effects of the antimicrobial peptide apidaecin on translation termination in bacteria. eLife 9, e62655 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Bulet, P., Urge, L., Ohresser, S., Hetru, C. & Otvos, L. Jr. Enlarged scale chemical synthesis and range of activity of drosocin, an O-glycosylated antibacterial peptide of Drosophila. Eur. J. Biochem. 238, 64–69 (1996).

    CAS  PubMed  Google Scholar 

  28. Bikker, F. J. et al. Evaluation of the antibacterial spectrum of drosocin analogues. Chem. Biol. Drug Des. 68, 148–153 (2006).

    CAS  PubMed  Google Scholar 

  29. Gobbo, M. et al. Antimicrobial peptides: synthesis and antibacterial activity of linear and cyclic drosocin and apidaecin 1b analogues. J. Med. Chem. 45, 4494–4504 (2002).

    CAS  PubMed  Google Scholar 

  30. Vonkavaara, M. et al. Francisella is sensitive to insect antimicrobial peptides. J. Innate Immun. 5, 50–59 (2013).

    CAS  PubMed  Google Scholar 

  31. Hoffmann, R., Bulet, P., Urge, L. & Otvos, L. Jr. Range of activity and metabolic stability of synthetic antibacterial glycopeptides from insects. Biochim. Biophys. Acta 1426, 459–467 (1999).

    CAS  PubMed  Google Scholar 

  32. Hanson, M. A., Kondo, S. & Lemaitre, B. Drosophila immunity: the Drosocin gene encodes two host defence peptides with pathogen-specific roles. Proc. Biol. Sci. 289, 20220773 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Ludwig, T., Krizsan, A., Mohammed, G. K. & Hoffmann, R. Antimicrobial activity and 70S ribosome binding of apidaecin-derived Api805 with increased bacterial uptake rate. Antibiotics 11, 430 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Orelle, C. et al. Tools for characterizing bacterial protein synthesis inhibitors. Antimicrob. Agents Chemother. 57, 5994–6004 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Uno, M., Ito, K. & Nakamura, Y. Functional specificity of amino acid at position 246 in the tRNA mimicry domain of bacterial release factor 2. Biochimie 78, 935–943 (1996).

    CAS  PubMed  Google Scholar 

  36. Kragol, G. et al. The antibacterial peptide pyrrhocoricin inhibits the ATPase actions of DnaK and prevents chaperone-assisted protein folding. Biochemistry 40, 3016–3026 (2001).

    CAS  PubMed  Google Scholar 

  37. Monk, J. W. et al. Rapid and inexpensive evaluation of nonstandard amino acid incorporation in Escherichia coli. ACS Synth. Biol. 6, 45–54 (2017).

    CAS  PubMed  Google Scholar 

  38. Knappe, D., Cassone, M., Nollmann, F. I., Otvos, L. & Hoffmann, R. Hydroxyproline substitutions stabilize non-glycosylated drosocin against serum proteases without challenging its antibacterial activity. Protein Pept. Lett. 21, 321–329 (2014).

    CAS  PubMed  Google Scholar 

  39. Lele, D. S., Dwivedi, R., Kumari, S. & Kaur, K. J. Effect of distal sugar and interglycosidic linkage of disaccharides on the activity of proline rich antimicrobial glycopeptides. J. Pept. Sci. 21, 833–844 (2015).

    CAS  PubMed  Google Scholar 

  40. de Visser, P. C. et al. Biological evaluation of Tyr6 and Ser7 modified drosocin analogues. Bioorg. Med. Chem. Lett. 15, 2902–2905 (2005).

    PubMed  Google Scholar 

  41. Lele, D. S., Talat, S., Kumari, S., Srivastava, N. & Kaur, K. J. Understanding the importance of glycosylated threonine and stereospecific action of Drosocin, a proline rich antimicrobial peptide. Eur. J. Med. Chem. 92, 637–647 (2015).

    CAS  PubMed  Google Scholar 

  42. Ahn, M. et al. Substitution of the GalNAc-α-O-Thr11 residue in drosocin with O-linked glyco-peptoid residue: effect on antibacterial activity and conformational change. Bioorg. Med. Chem. Lett. 21, 6148–6153 (2011).

    CAS  PubMed  Google Scholar 

  43. Taguchi, S., Mita, K., Ichinohe, K. & Hashimoto, S. Targeted engineering of the antibacterial peptide apidaecin, based on an in vivo monitoring assay system. Appl. Environ. Microbiol. 75, 1460–1464 (2009).

    CAS  PubMed  Google Scholar 

  44. Taguchi, S., Nakagawa, K., Maeno, M. & Momose, H. In vivo monitoring system for structure–function relationship analysis of the antibacterial peptide apidaecin. Appl. Environ. Microbiol. 60, 3566–3572 (1994).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Taguchi, S., Ozaki, A., Nakagawa, K. & Momose, H. Functional mapping of amino acid residues responsible for the antibacterial action of apidaecin. Appl. Environ. Microbiol. 62, 4652–4655 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Taboureau, O. et al. Design of novispirin antimicrobial peptides by quantitative structure–activity relationship. Chem. Biol. Drug Des. 68, 48–57 (2006).

    CAS  PubMed  Google Scholar 

  47. Muthunayake, N. S. et al. Expression and in vivo characterization of the antimicrobial peptide oncocin and variants binding to ribosomes. Biochemistry 59, 3380–3391 (2020).

    CAS  PubMed  Google Scholar 

  48. Kuru, E. et al. Release factor inhibiting antimicrobial peptides improve nonstandard amino acid incorporation in wild-type bacterial cells. ACS Chem. Biol. 15, 1852–1861 (2020).

    CAS  PubMed  Google Scholar 

  49. DeJong, M. P. et al. A platform for deep sequence–activity mapping and engineering antimicrobial peptides. ACS Synth. Biol. 10, 2689–2704 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Koller, T. O. et al. Structural basis for translation inhibition by the glycosylated drosocin peptide. Nat. Chem. Biol. https://doi.org/10.1038/s41589-023-01293-7 (2023).

  51. Bundy, B. C. & Swartz, J. R. Site-specific incorporation of p-propargyloxyphenylalanine in a cell-free environment for direct protein-protein click conjugation. Bioconjug. Chem. 21, 255–263 (2010).

    CAS  PubMed  Google Scholar 

  52. Orelle, C. et al. Identifying the targets of aminoacyl-tRNA synthetase inhibitors by primer extension inhibition. Nucleic Acids Res. 41, e144 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Baba, T. et al. Construction of Escherichia coli K-12 in-frame, single-gene knockout mutants: the Keio collection. Mol. Syst. Biol. 2, 2006.0008 (2006).

    PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank T. Koller and D. Wilson (University of Hamburg) for generously sharing their results and the structure of the ribosome–drosocin complex. This work was supported in part by NIH grant R01 AI162961 (to A.S.M., N.V.-L. and Y.S.P.).

Author information

Authors and Affiliations

Authors

Contributions

R.H., N.V.-L. and A.S.M. conceived the study. K.M. guided and supervised preparation of the endogenously expressed wt Dro and Dro mutant library. D.K. carried out toeprinting and microbiological experiments. A.B. and A.K. synthesized peptides and carried out in vitro translation and MIC testing experiments. I.O. cloned the Dro gene and carried out library screening experiments. K.M. and C.B. analyzed library screening results. C.B. consulted on multiple experiments. W.H. analyzed stop codon readthrough and Dro-resistant mutants. Y.S.P. analyzed structural data. K.M., C.B., A.K., R.H., N.V.-L. and A.S.M. analyzed data. K.M., N.V.-L. and A.S.M. wrote the manuscript.

Corresponding authors

Correspondence to Ralf Hoffmann, Nora Vázquez-Laslop or Alexander S. Mankin.

Ethics declarations

Competing interests

Up until 1 year before submission of this manuscript, R.H. served as an advisor for the company EnBiotix, Inc. on a project unrelated to this study. The remaining authors declare no competing interests.

Peer review

Peer review information

Nature Chemical Biology thanks Alex Tossi 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 Dro acts upon translation termination but causes weak translation arrest at UGA stop codons.

a, Inhibition of in vitro GFP translation by synthetic PrAMPs: class-I Oncocin112 (Onc112), class-II Api137, or non-glycosylated Dro (Dro). Bar graphs represent the normalized values of GFP fluorescence from reactions where RF1 was depleted (light grey) or supplemented (dark grey), setting the fluorescence value from reactions with or without RF1 in the absence of PrAMP as 100%. The error bars show standard deviation from the mean in three independent experiments. Significance levels indicated as NS, not significant; *, p-value <0.05; **, p-value < 0.01; ***, value < 0.001 (One-way ANOVA with Tukey’s Multiple Comparison test by GraphPad Prism). b, In vitro toeprinting analysis of the Api137 or Dro-mediated ribosome arrest at the UGA stop codon (red arrowhead) of the model yrbA ORF. The control reaction with no added PrAMPs is labeled as ‘none’. The control antibiotic retapamulin (Ret) stalls ribosomes at the start codon (green arrowhead). Sequencing reactions are labeled as C, U, A, G.

Source data

Extended Data Fig. 2 Endogenous expression of Dro in cells grown in rich medium does not prevent cell growth.

Growth of E. coli BL21 cells transformed with pDro[UAG] or pDro[UGA] plasmids on agar lysogeny broth (LB) rich medium supplemented with glucose or L-arabinose. The toxic effect of endogenous expression of Api (in cells transformed with pApi) is shown for comparison. Cells transformed with empty pBAD vector were used as a negative control.

Extended Data Fig. 3 Robustness of experimental data and comparisons of the effect of endogenously expressed Dro and Api variants on cell growth.

a, Correlation of the depletion scores of E. coli clones from the single amino acid Dro mutant libraries generated on the bases of pDro[UAG] or pDro[UGA] expression plasmids. Pearson correlation coefficient (r) is indicated. b, Depletion scores of library clones endogenously expressing single-amino acid Dro variants from the pDro[UGA] plasmids. Coloring is according to the toxic effects (blue - highly toxic, yellow - mildly toxic, salmon - not toxic). The analogous data for the pDro[UAG] library are shown in Fig. 4. c, Similarly calculated depletion scores for endogenously expressed Api mutants from the pApi plasmid using data from the reference7.

Extended Data Fig. 4 Dro variants with single amino acid substitutions retain the ability to arrest ribosomes at stop codons.

Toeprinting analysis of the ribosome arrest of the UAG stop codon of the model yrbA ORF mediated by synthetic non-glycosylated Dro variants. Samples with no added synthetic peptide are labeled as ‘none’. Arrest at stop codons caused by Api137 or at start codons by retapamulin (Ret) are shown as reference. Toeprint bands from ribosomes stalled at start or stop codons are marked by green and red arrowheads, respectively. Shown are representative gels of two independent experiments that produced converging results.

Extended Data Fig. 5 Functionally critical contacts of Dro with the ribosome as revealed by mutational analysis.

The central image depicts the placement of glycosylated Dro in the NPET of the E. coli ribosome50. Functionally critical Dro residues are indicated in salmon; residues that tolerate multiple substitutions are marked in green. a-e, Functionally critical contacts involving Dro residues: a, Arg18, b, Arg15 and Pro16, c, Arg9, d, Arg4, e, Lys2. Panel f shows the ribosomal contacts of the Pro5-Pro8 segment of Dro where most amino acids substitutions do not impair the PrAMP’s activity.

Supplementary information

Supplementary Information

Supplementary Tables 1 and 2 combined in a single pdf file.

Reporting Summary

Source data

Source Data Fig. 1

Uncropped gels for Fig. 1c.

Source Data Fig. 2

Raw data and statistical values for Fig. 2b.

Source Data Fig. 4

Read counts for UAG and UGA codons (Fig. 4).

Source Data Fig. 5

Raw data and statistical values for Fig. 5a.

Source Data Fig. 5

Uncropped gels for Fig. 5a.

Source Data Extended Data Fig. 1

Raw data and statistical values for ED Fig. 1a.

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

Mangano, K., Klepacki, D., Ohanmu, I. et al. Inhibition of translation termination by the antimicrobial peptide Drosocin. Nat Chem Biol 19, 1082–1090 (2023). https://doi.org/10.1038/s41589-023-01300-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-023-01300-x

This article is cited by

Search

Quick links

Nature Briefing Microbiology

Sign up for the Nature Briefing: Microbiology newsletter — what matters in microbiology research, free to your inbox weekly.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing: Microbiology