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:

Bacterial retrons encode phage-defending tripartite toxin–antitoxin systems

Abstract

Retrons are prokaryotic genetic retroelements encoding a reverse transcriptase that produces multi-copy single-stranded DNA1 (msDNA). Despite decades of research on the biosynthesis of msDNA2, the function and physiological roles of retrons have remained unknown. Here we show that Retron-Sen2 of Salmonella enterica serovar Typhimurium encodes an accessory toxin protein, STM14_4640, which we renamed as RcaT. RcaT is neutralized by the reverse transcriptase–msDNA antitoxin complex, and becomes active upon perturbation of msDNA biosynthesis. The reverse transcriptase is required for binding to RcaT, and the msDNA is required for the antitoxin activity. The highly prevalent RcaT-containing retron family constitutes a new type of tripartite DNA-containing toxin–antitoxin system. To understand the physiological roles of such toxin–antitoxin systems, we developed toxin activation–inhibition conjugation (TAC-TIC), a high-throughput reverse genetics approach that identifies the molecular triggers and blockers of toxin–antitoxin systems. By applying TAC-TIC to Retron-Sen2, we identified multiple trigger and blocker proteins of phage origin. We demonstrate that phage-related triggers directly modify the msDNA, thereby activating RcaT and inhibiting bacterial growth. By contrast, prophage proteins circumvent retrons by directly blocking RcaT. Consistently, retron toxin–antitoxin systems act as abortive infection anti-phage defence systems, in line with recent reports3,4. Thus, RcaT retrons are tripartite DNA-regulated toxin–antitoxin systems, which use the reverse transcriptase–msDNA complex both as an antitoxin and as a sensor of phage protein 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: Retron-Sen2 encodes and inhibits the toxin protein RcaT.
Fig. 2: RcaT proteins interact and co-evolve with their cognate retron reverse transcriptases.
Fig. 3: Toxin activation–inhibition conjugation (TAC-TIC) is a genetic screen for discovery of TA triggers and blockers.
Fig. 4: Triggers activate retron toxins by directly affecting the msDNA.
Fig. 5: Retron TAs protect against phages.

Similar content being viewed by others

Data availability

Raw reads from whole-genome sequencing are deposited at the European Nucleotide Archive under accession PRJEB38324. Proteomics data from RcaT and RT immunoprecipitations and phage infections are in Supplementary Tables 1 and 7, and the raw data are deposited at ProteomeXchange under accession PXD022376. All TAC-TIC data are in Supplementary Table 5. Source images and replicates are provided at figshare (https://doi.org/10.6084/m9.figshare.17376488).

Code availability

TAC-TIC data were analysed with R (3.6.1), with the full code available at https://git.embl.de/kritikos/tic-tac.

References

  1. Yee, T., Furuichi, T., Inouye, S. & Inouye, M. Multicopy single-stranded DNA isolated from a Gram-negative bacterium, Myxococcus xanthus. Cell 38, 203–209 (1984).

    Article  CAS  PubMed  Google Scholar 

  2. Simon, A. J., Ellington, A. D. & Finkelstein, I. J. Retrons and their applications in genome engineering. Nucleic Acids Res. 47, 11007–11019 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Gao, L. et al. Diverse enzymatic activities mediate antiviral immunity in prokaryotes. Science 369, 1077–1084 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  4. Millman, A. et al. Bacterial retrons function in anti-phage defense. Cell 183, 1551–1561.e12 (2020).

    Article  CAS  PubMed  Google Scholar 

  5. Lampson, B. C., Inouye, M. & Inouye, S. Reverse transcriptase with concomitant ribonuclease H activity in the cell-free synthesis of branched RNA-linked msDNA of Myxococcus xanthus. Cell 56, 701–707 (1989).

    Article  CAS  PubMed  Google Scholar 

  6. Lampson, B. C. et al. Reverse transcriptase in a clinical strain of Escherichia coli: production of branched RNA-linked msDNA. Science 243, 1033–1038 (1989).

    Article  ADS  CAS  PubMed  Google Scholar 

  7. Mestre, M. R., González-Delgado, A., Gutiérrez-Rus, L. I., Martínez-Abarca, F. & Toro, N. Systematic prediction of genes functionally associated with bacterial retrons and classification of the encoded tripartite systems. Nucleic Acids Res. 48, 12632–12647 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Dhundale, A., Lampson, B., Furuichi, T., Inouye, M. & Inouye, S. Structure of msDNA from Myxococcus xanthus: evidence for a long, self-annealing RNA precursor for the covalently linked, branched RNA. Cell 51, 1105–1112 (1987).

    Article  CAS  PubMed  Google Scholar 

  9. Hsu, M. Y., Inouye, S. & Inouye, M. Structural requirements of the RNA precursor for the biosynthesis of the branched RNA-linked multicopy single-stranded DNA of Myxococcus xanthus. J. Biol. Chem. 264, 6214–6219 (1989).

    Article  CAS  PubMed  Google Scholar 

  10. Lim, D. Structure and biosynthesis of unbranched multicopy single‐stranded DNA by reverse transcriptase in a clinical Escherichia coli isolate. Mol. Microbiol. 6, 3531–3542 (1992).

    Article  CAS  PubMed  Google Scholar 

  11. Lima, T. M. O. & Lim, D. A novel retron that produces RNA-less msDNA in Escherichia coli using reverse transcriptase. Plasmid 38, 25–33 (1997).

    Article  CAS  PubMed  Google Scholar 

  12. Jung, H., Liang, J., Jung, Y. & Lim, D. Characterization of cell death in Escherichia coli mediated by XseA, a large subunit of exonuclease VII. J. Microbiol. 53, 820–828 (2015).

    Article  CAS  PubMed  Google Scholar 

  13. Schubert, M. G. et al. High-throughput functional variant screens via in vivo production of single-stranded DNA. Proc. Natl Acad. Sci. USA 118, e2018181118 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Farzadfard, F., Gharaei, N., Citorik, R. J. & Lu, T. K. Efficient retroelement-mediated DNA writing in bacteria. Cell Syst. 12, 860–872.e5 (2021).

    Article  CAS  PubMed  Google Scholar 

  15. Porwollik, S. et al. Defined single-gene and multi-gene deletion mutant collections in Salmonella enterica sv Typhimurium. PLoS ONE 9, e99820 (2014).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  16. Pfalz, B. Comparing bacterial gene networks based on high-throughput phenomics. Preprint at heiDOK https://doi.org/10.11588/heidok.00022920 (2017).

  17. Elfenbein, J. R. et al. Multicopy single-stranded DNA directs intestinal colonization of enteric pathogens. PLoS Genet. 11, e1005472 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  18. Jumper, J. et al. Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  19. Zhao, G. et al. Structure of the N-glycosidase MilB in complex with hydroxymethyl CMP reveals its Arg23 specifically recognizes the substrate and controls its entry. Nucleic Acids Res. 42, 8115–8124 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Lampson, B. C., Viswanathan, M., Inouye, M. & Inouye, S. Reverse transcriptase from Escherichia coli exists as a complex with msDNA and is able to synthesize double-stranded DNA. J. Biol. Chem. 265, 8490–8496 (1990).

    Article  CAS  PubMed  Google Scholar 

  21. Jeong, D. W., Kim, K. & Lim, D. Evidence for the complex formation between reverse transcriptase and multicopy single-stranded DNA in retron EC83. Mol. Cells 7, 347–351 (1997).

    CAS  PubMed  Google Scholar 

  22. Almeida, A. et al. A unified catalog of 204,938 reference genomes from the human gut microbiome. Nat. Biotechnol. 39, 105–114 (2021).

    Article  CAS  PubMed  Google Scholar 

  23. Bleibtreu, A. et al. The rpoS gene is predominantly inactivated during laboratory storage and undergoes source-sink evolution in Escherichia coli species. J. Bacteriol. 196, 4276–4284 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  24. Inouye, S., Hsu, M.-Y., Xu, A. & Inouye, M. Highly specific recognition of primer RNA structures for 2′-OH priming reaction by bacterial reverse transcriptases. J. Biol. Chem. 274, 31236–31244 (1999).

    Article  CAS  PubMed  Google Scholar 

  25. Fraikin, N., Goormaghtigh, F. & Van Melderen, L. Type II toxin–antitoxin systems: evolution and revolutions. J. Bacteriol. 202, e00763–19 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Saka, K. A complete set of Escherichia coli open reading frames in mobile plasmids facilitating genetic studies. DNA Res. 12, 63–68 (2005).

    Article  CAS  PubMed  Google Scholar 

  27. Otsuka, Y. et al. GenoBase: comprehensive resource database of Escherichia coli K-12. Nucleic Acids Res. 43, D606–D617 (2015).

    Article  CAS  PubMed  Google Scholar 

  28. Bolden, A., Ward, C., Siedlecki, J. A. & Weissbach, A. DNA methylation. J. Biol. Chem. 259, 12437–12443 (1984).

    Article  CAS  PubMed  Google Scholar 

  29. Lampson, B. C., Inouye, M. & Inouye, S. Retrons, msDNA, and the bacterial genome. Cytogenet. Genome Res. 110, 491–499 (2005).

    Article  CAS  PubMed  Google Scholar 

  30. Geier, G. E. & Modrich, P. Recognition sequence of the dam methylase of Escherichia coli K-12 and mode of cleavage of Dpn I endonuclease. J. Biol. Chem. 254, 1408–1413 (1979).

    Article  CAS  PubMed  Google Scholar 

  31. Szyf, M., Avraham-Haetzni, K. & Reifman, A. DNA methylation pattern is determined by the intracellular level of the methylase. Proc. Natl Acad. Sci. USA 81, 3278–3282 (1984).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  32. Joseph, J. W. & Kolodner, R. Exonuclease VIII of Escherichia coli. II. Mechanism of action. J. Biol. Chem. 258, 10418–10424 (1983).

    Article  CAS  PubMed  Google Scholar 

  33. Nichols, R. J. et al. Phenotypic landscape of a bacterial cell. Cell 144, 143–156 (2011).

    Article  CAS  PubMed  Google Scholar 

  34. Price, M. N. et al. Mutant phenotypes for thousands of bacterial genes of unknown function. Nature 557, 503–509 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  35. Typas, A. et al. Regulation of peptidoglycan synthesis by outer-membrane proteins. Cell 143, 1097–1109 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Harms, A., Brodersen, D. E., Mitarai, N. & Gerdes, K. Toxins, targets, and triggers: an overview of toxin–antitoxin biology. Mol. Cell 70, 768–784 (2018).

    Article  CAS  PubMed  Google Scholar 

  37. LeRoux, M. & Laub, M. T. Toxin–antitoxin systems as phage defense elements. Annu. Rev. Microbiol. 76, 21–43 (2022).

    Article  Google Scholar 

  38. Sberro, H. et al. Discovery of functional toxin/antitoxin systems in bacteria by shotgun cloning. Mol. Cell 50, 136–148 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Christensen-Dalsgaard, M., Jørgensen, M. G. & Gerdes, K. Three new RelE-homologous mRNA interferases of Escherichia coli differentially induced by environmental stresses. Mol. Microbiol. 75, 333–348 (2010).

    Article  CAS  PubMed  Google Scholar 

  40. Germain, E., Roghanian, M., Gerdes, K. & Maisonneuve, E. Stochastic induction of persister cells by HipA through (p)ppGpp-mediated activation of mRNA endonucleases. Proc. Natl Acad. Sci. USA 112, 5171–5176 (2015).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  41. Negri, A., Jąkalski, M., Szczuka, A., Pryszcz, L. P. & Mruk, I. Transcriptome analyses of cells carrying the type II Csp231I restriction-modification system reveal cross-talk between two unrelated transcription factors: C protein and the Rac prophage repressor. Nucleic Acids Res. 47, 9542–9556 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Hattman, S. DNA methylation of T-even bacteriophages and of their nonglucosylated mutants: Its role in P1-directed restriction. Virology 42, 359–367 (1970).

    Article  CAS  PubMed  Google Scholar 

  43. Handa, N. & Kobayashi, I. Type III restriction is alleviated by bacteriophage (RecE) homologous recombination function but enhanced by bacterial (RecBCD) function. J. Bacteriol. 187, 7362–7373 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Kritikos, G. et al. A tool named Iris for versatile high-throughput phenotyping in microorganisms. Nat. Microbiol. 2, 17014 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Datsenko, K. A. & Wanner, B. L. One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proc. Natl Acad. Sci. USA 97, 6640–6645 (2000).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  46. 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).

    Article  PubMed  PubMed Central  Google Scholar 

  47. Cherepanov, P. P. & Wackernagel, W. Gene disruption in Escherichia coli: TcR and KmR cassettes with the option of Flp-catalyzed excision of the antibiotic-resistance determinant. Gene 158, 9–14 (1995).

    Article  CAS  PubMed  Google Scholar 

  48. Ferrières, L. et al. Silent mischief: bacteriophage Mu insertions contaminate products of Escherichia coli random mutagenesis performed using suicidal transposon delivery plasmids mobilized by broad-host-range RP4 conjugative machinery. J. Bacteriol. 192, 6418–6427 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Chung, C. T., Niemela, S. L. & Miller, R. H. One-step preparation of competent Escherichia coli: transformation and storage of bacterial cells in the same solution. Proc. Natl Acad. Sci. USA 86, 2172–2175 (1989).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  50. Chassy, B. Transformation of bacteria by electroporation. Trends Biotechnol. 6, 303–309 (1988).

    Article  CAS  Google Scholar 

  51. Green, M. R. & Sambrook, J. Preparation of plasmid DNA by alkaline lysis with sodium dodecyl sulfate: minipreps. Cold Spring Harb. Protoc. 2016, pdb.prot093344 (2016).

    Article  Google Scholar 

  52. Green, M. R. & Sambrook, J. Isolation of DNA fragments from polyacrylamide gels by the crush and soak method. Cold Spring Harb. Protoc. 2019, pdb.prot100479 (2019).

    Article  Google Scholar 

  53. Summer, H., Grämer, R. & Dröge, P. Denaturing urea polyacrylamide gel electrophoresis (urea PAGE). J. Vis. Exp. 32, 1485 (2009).

    Google Scholar 

  54. Bassam, B. J. & Gresshoff, P. M. Silver staining DNA in polyacrylamide gels. Nat. Protoc. 2, 2649–2654 (2007).

    Article  CAS  PubMed  Google Scholar 

  55. Deatherage, D. E. & Barrick, J. E. in Engineering and Analyzing Multicellular Systems: Methods and Protocols (eds Sun, L. & Shou, W.) 165–188 (Springer, 2014).

  56. Hughes, C. S. et al. Single-pot, solid-phase-enhanced sample preparation for proteomics experiments. Nat. Protoc. 14, 68–85 (2019).

    Article  CAS  PubMed  Google Scholar 

  57. Mateus, A. et al. Thermal proteome profiling in bacteria: probing protein state in vivo. Mol. Syst. Biol. 14, e8242 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Ritchie, M. E. et al. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 43, e47 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Studier, F. W. Protein production by auto-induction in high-density shaking cultures. Protein Expr. Purif. 41, 207–234 (2005).

    Article  CAS  PubMed  Google Scholar 

  60. Katoh, K. MAFFT: a novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res. 30, 3059–3066 (2002).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  61. Price, M. N., Dehal, P. S. & Arkin, A. P. FastTree 2—approximately maximum-likelihood trees for large alignments. PLoS ONE 5, e9490 (2010).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  62. Benjamini, Y. & Hochberg, Y. Controlling the false discovery rate: a practical and powerful approach to multiple testing. J. R. Stat. Soc. B 57, 289–300 (1995).

    MathSciNet  MATH  Google Scholar 

  63. Arndt, D. et al. PHASTER: a better, faster version of the PHAST phage search tool. Nucleic Acids Res. 44, W16–W21 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Khlebnikov, A. & Keasling, J. D. Effect of lacY expression on homogeneity of induction from the Ptac and Ptrc promoters by natural and synthetic inducers. Biotechnol. Prog. 18, 672–674 (2002).

    Article  CAS  PubMed  Google Scholar 

  65. Kropinski, A. M., Mazzocco, A., Waddell, T. E., Lingohr, E. & Johnson, R. P. Enumeration of bacteriophages by double agar overlay plaque assay. Methods Mol. Biol. 501, 69–76 (2009).

    Article  CAS  PubMed  Google Scholar 

  66. Collins, S. R., Schuldiner, M., Krogan, N. J. & Weissman, J. S. A strategy for extracting and analyzing large-scale quantitative epistatic interaction data. Genome Biol. 7, R63 (2006).

    Article  PubMed  PubMed Central  Google Scholar 

  67. Zuker, M. Mfold web server for nucleic acid folding and hybridization prediction. Nucleic Acids Res. 31, 3406–3415 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Jahn, M., Vorpahl, C., Hübschmann, T., Harms, H. & Müller, S. Copy number variability of expression plasmids determined by cell sorting and droplet digital PCR. Microb. Cell Fact. 15, 211 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  69. Walch, P. D. K. et al. Global mapping of Salmonella enterica–host protein–protein interactions during infection. Cell Host Microbe 29, 1316–1332.e12 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Datta, S., Costantino, N. & Court, D. L. A set of recombineering plasmids for Gram-negative bacteria. Gene 379, 109–115 (2006).

    Article  CAS  PubMed  Google Scholar 

  71. Uehara, T., Parzych, K. R., Dinh, T. & Bernhardt, T. G. Daughter cell separation is controlled by cytokinetic ring-activated cell wall hydrolysis. EMBO J. 29, 1412–1422 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Guzman, L. M., Belin, D., Carson, M. J. & Beckwith, J. Tight regulation, modulation, and high-levels expression by vectors containing the arabinose PBAD promoter. J. Bacteriol. 177, 4121–4130 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank the EMBL Genomics Core for WGS; the EMBL PEP Core for help with RT-Sen2 purification; the EMBL Proteomics Core for proteomics; the Typas laboratory for discussions, especially M. Wartel for helping in establishing high-throughput conjugation; E. van Nimwegen for E. coli SC402; and V. Mutalik and B. Adler for Ffm and Br60 phages. This work was supported by EMBL and the Sofja Kovaleskaja Award of the Alexander von Humboldt Foundation. K.M. and A.M. were supported by a fellowship from the EMBL Interdisciplinary Postdoc (EI3POD) programme (MSCA COFUND; 664726). M.G. was funded by the DFG under Germany's Excellence Strategy–EXC 2155–390874280. J.R.E. was supported by the NIH (K08AI108794). H.L.A.-P. is supported by NIFA (NIFA 2016-11004 and 2017-08881) and DARPA.

Author information

Authors and Affiliations

Authors

Contributions

A.B., G.Z., M.M.S., J.R.E., H.L.A.-P. and A.T. supervised the study. J.B. and A.T. conceived the study with inputs from J.R.E. and H.L.A.-P. J.B., K.M., A.M. and A.T. designed the experiments, and J.B., K.M., A.M., J.S., M.Z., S.G.-S., A.S. and C.K. performed the experiments. N.K. and M.G. constructed the RcaT and reverse transcriptase phylogenetic trees. V.M. predicted the RcaT-Sen2 structure. The STm chemical genetics screen data was from B.P. and M.G. M.G. analysed the STm genome sequencing data and A.M. and F.S. analysed the proteomics data. G.K. and J.B. analysed the TAC-TIC screen. J.B., A.M., N.K., K.M. and G.K. designed figures, with inputs from A.T. J.B. and A.T. wrote the different manuscript versions with input from all authors.

Corresponding authors

Correspondence to Johanna R. Elfenbein, Helene L. Andrews-Polymenis or Athanasios Typas.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature thanks Masayori Inouye and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer review 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 figures and tables

Extended Data Fig. 1 Retron phenotypes are suppressed by inactivating the accessory retron gene, which encodes a cytoplasmic toxin RcaT.

a, ΔrrtT and ΔxseA STm are cold-sensitive. 1536 colony-arrays of the STm gene-deletion library15 were pinned onto LB plates, and strains were grown at room temperature. Colony sizes44 were used to calculate a fitness S-score for each strain66. S-scores were calculated from n = 8 biological. Dashed vertical line denotes the mean S-score of all strains (n = 3781); negative or positive S-scores indicate sensitive or resistant mutants. b, Perturbing msDNA biogenesis leads to cold-sensitivity. STm wild-type and retron-deletion strains were serially diluted and spotted on LB plates as in Fig. 1b. Plates were incubated at indicated temperatures (n = 4 biological). c, Retron mutants grow less in anaerobic conditions. Growth curves of STm wild-type and retron-deletion strains were obtained by measuring OD578 in microtiter plates, under anaerobic conditions at 37 °C; n = 11 biological, symbols denote the mean and error bars the standard deviation (not shown if smaller than symbols). d, Retron mutants are not affected in aerobic conditions. Experiment and plotting as in panel c, but strains were grown aerobically (n = 11 biological). e, RNAse H and Exo VII are involved in msDNA biosynthesis. msDNA was extracted from STm wild-type and retron-deletion strains carrying plasmid p-retron-ΔrcaT. Extracted msDNA was electrophoresed in TBE-Polyacrylamide gels (n = 3 biological). f, Deleting rcaT reverts the cold-sensitivity of retron mutants. STm strains were grown and spotted as in Fig. 1d (n = 2 biological). g, Deleting rcaT reverts the anaerobic sensitivity of retron mutants. Growth curves of STm strains were obtained and plotted as in panel c (n = 11 biological). h, RcaT is a soluble protein. Untagged and rcaT-3xFlag STm WT were grown in LB at 37 °C, lysed, and samples were separated into soluble and membrane fractions through ultra-centrifugation steps. Different fractions were analysed by SDS-PAGE and immunoblotting. LpoA and RecA were used as controls for the membrane and soluble fraction, respectively (n = 2 biological). ij, Suppressors grow like wild type in cold temperatures. Suppressors isolated from cold-sensitive STm mutants (ΔrrtT, ΔxseA, Δmsrmsd; panel i) or the catalytic rcaT-E107Q mutants (panel j) were grown, serially diluted, and spotted on LB plates as in Fig. 1b. Identified suppressor mutations are indicated (n = 2 biological). kl, Loss-of-function point mutations do not affect RcaT levels. RcaT protein levels were quantified by mass spectrometry in mutant strains (indicated by dashed lines in panels ij). Y-axis is the ratio of RcaT protein levels (log2 fold-change) in retron deletion strains compared to WT (panel k) or the same ratio in RcaT point mutants compared to the background strain, in which suppressor was isolated in (panel l). The grey dotted line represents no change to RcaT protein levels. Black lines denote the mean (n = 2–4 biological).

Extended Data Fig. 2 Retron-Sen2 functions as a toxin-antitoxin system in E. coli and the antitoxin does not shut down RcaT expression.

a, RcaT can be inhibited by msrmsd-rrtT in trans. E. coli carrying binary combinations of plasmids p-rcaT, p-retron-ΔrcaT, and empty vectors, were grown in LB with appropriate antibiotics, serially diluted and spotted as in Fig. 1f (n = 3 biological). b, RcaT inhibition requires both msrmsd and rrtT in E. coli. E. coli with plasmids carrying retron-components were grown and spotted as in Fig. 1f. Only the intact retron could restore growth. RcaT expression is sufficient to inhibit growth (n = 3 biological). c, RcaT inhibition requires RNAse H and Exo VII in E. coli. E. coli strains (WT, ΔxseA, ΔxseB, ΔrnhA) carrying plasmid p-retron-ΔrcaT or p-retron were grown, serially diluted and spotted as in panel b (n = 2 biological). d, RcaT levels are only slightly affected by antitoxin deletions. rcaT-3xFlag STm strains (WT and retron-deletions) and the STm untagged strain (native) were either grown in LB at 37 °C, or shifted to 20 °C for 5 h. Protein samples from strains were analysed by SDS-PAGE and immunoblotting. LpoA levels (α-LpoA antibody) were used as loading control. Bars and error bars indicate mean and standard deviation, respectively (n = 5 biological). e, Quantification of RcaT-3xFlag signal from 37 °C immunoblots in panel d using ImageJ (pixel-density). Bars and error bars depict the mean and standard deviation, respectively (n = 5 biological).

Extended Data Fig. 3 Functional Flag-tagged RcaT-RT co-immunoprecipitate independently of temperature.

ab, Volcano plots of affinity purifications of 3xFlag-tagged RrtT (panel a) and RcaT (panel b) at 37 °C and 20 °C in wild-type and different mutant backgrounds. Same experiment as in Fig. 2a, b, presenting additionally the AP data from 37 °C. The x-axis represents the average log2 ratio of identified proteins (by more than 2 independent peptides) in rrtT-3xFlag and rcaT-3xFlag AP samples compared to an untagged STm WT control strain (n = 2 biological), and the y-axis represents p values of these log2 ratios (two-tailed limma). c, Flag-tagged RcaT retains its function. Untagged and rcaT-3xFlag tagged STm strains (WT and retron-deletions) were grown for 5-6 h at 37 °C in LB, serially diluted, spotted on LB plates, and incubated at indicated temperatures. * denotes ΔSTM14_4645::cat, used to co-transduce the scarless Flag-tagged rcaT (co-transduction verified by PCR) (n = 2 biological). d, Flag-tagged RrtT retains its function. Untagged and rrtT-3xFlag tagged STm strains (WT and retron-deletions) were grown for 5-6 h at 37 °C in LB, serially diluted, spotted on LB plates, and plates were incubated either at indicated temperatures (n = 2 biological). e, Flag-tagging rrtT does not alter retron protein expression, whereas flag-tagging rcaT decreases levels of both retron proteins. Proteins in input (whole proteome) samples used for samples shown in Fig. 2a, b were quantified by MS. Protein levels (log2-fold change) of RcaT and RT were compared between Flag-tagged and untagged STm WT strains (x-axis). Lines denote the mean and come from two biological replicates. RT was downregulated in rcaT-3xFlag tagged strains, whereas RcaT was not even detected (n.d.), presumably due to lower levels (chromosomal RcaT is at the level of detection by MS, so small changes can bring it below detection). Note that despite lower levels of both retron components, rcaT-3xFlag retains its function (panel c) and is expressed (Extended Data Fig. 2d).

Extended Data Fig. 4 RT interacts with msDNA.

a, Purification of protein RT-Sen2. An E. coli BL21 (DE3) CodonPlus-RIL strain carrying plasmid p-msrmsd-rrtT-6xHis was used to purify protein RT-Sen2-6xHis (C-terminal fusion), by nickel-column immobilized metal-affinity chromatography (n = 2 purifications). b, His-tagging rrtT does not affect its antitoxin activity. E. coli BL21-AI strains carrying binary combinations of plasmids p-msrmsd-rrtT-6xHis, p-rcaT, and empty vectors, were grown for 5-6 h at 37 °C in kanamycin-LB, serially diluted, spotted on LB-kanamycin plates with or without arabinose, and plates were incubated at 20 °C (n = 6 biological). c, E. coli BL21 does not normally produce DNA of similar size to msDNA-Sen2. msDNA was extracted from STm or E. coli BL21 carrying plasmid p-msrmsd-rrtT or an empty plasmid. Extracted msDNAs were electrophoresed in TBE-polyacrylamide gels (n = 4 biological). d, Isolation of msDNA from purified RT-Sen2 from E. coli (same prep as in panel a). Total DNA was extracted from 500 μg of purified RT-Sen2-6xHis protein and electrophoresed on a TBE-polyacrylamide gel (n = 2 purifications).

Extended Data Fig. 5 Different retrons inhibit growth in different environments.

a, STm and E. coli DE-COMM-4976 (Eco10), SC402 (Eco3), and REL606 (Eco1) (WT, ΔrnhA, ΔrcaT, or double mutant strains) were grown for 5-6 h at 37 °C in LB, serially diluted, and spotted on LB or LB pH 5.5 plates. Plates were incubated at indicated temperatures (n = 2 biological). b, Same strains as in panel a were grown anaerobically at 37 °C in LB for 24h in microtiter plates while measuring OD578. Here plotted the yield of mutant strains compared to the yield of the corresponding WT strain. n = 4 (2 biological/2 technical). c, Only RcaT-Sen2 and -Eco9 act as toxins when overexpressed alone. E. coli BW25113 strains carrying retron- or rcaT-encoding plasmids (Sen2, Eco9, Eco10, Eco3, and Eco1), were grown for 5-6 h at 37 °C in LB-spectinomycin, serially diluted, spotted on LB-spectinomycin plates with or without arabinose, and plates were incubated at 37 °C (n = 2 biological).

Extended Data Fig. 6 Retron-Eco9 is similar to Retron-Sen2.

a, Retron-Eco9 has a similar operon structure to Retron-Sen2. Retron-Eco9 contains msr, msd, rcaT and rrtT regions, 85%, 58%, 43%, and 49% identical to the corresponding Retron-Sen2 regions (first two nucleotide, last two protein level); rrt = retron reverse transcriptase. b, msDNA-Eco9 is similar to msDNA-Sen2. Models of msDNAs-Sen2 and -Eco967. c, msDNA-Eco9 biosynthesis requires ExoVII and RNase H. msDNA was extracted from E. coli BW25113 strains (wild-type, ΔxseA, ΔxseB, ΔrnhA) carrying plasmid PBAD-msrmsd-RT-Eco9 and electrophoresed in a TBE-Polyacrylamide gel (n = 2 biological). d, RcaT-Eco9 inhibition requires RNAse H and Exo VII in E. coli. E. coli BW25113 strains (wild-type, ΔxseA, ΔxseB, ΔrnhA) carrying plasmid Ptac-msrmsd-Eco9, along with either PBAD-RT-Eco9 (−), or PBAD-RT-RcaT-Eco9 (+), were grown for 5-6 h at 37 °C in LB with appropriate antibiotics, serially diluted, spotted on antibiotics-LB plates, with/without arabinose, and plates were incubated overnight at 37 °C (n = 2 biological). e, Non-cognate RT-msrmsd pairs produce msDNA. msDNA were extracted from E. coli BW25113 co-expressing the RT of Retrons-Sen2 (Se) or -Eco9 (Ec) (PBAD-RT; Se or Ec – arabinose induction), and Ec/Se msrmsd (Ptac-msrmsd or an empty vector (-)). Extracted msDNA were electrophoresed in a TBE-polyacrylamide gel (n = 2 biological). f, Ec RT cannot act as antitoxin at basal msrmsd (Se) levels, but can when latter is induced (Fig. 2e), in line with lower capacity for Ec RT to reverse transcribe the Se msrmsd (panel e). Experiment conducted as in Fig. 2e (n = 2 biological).

Extended Data Fig. 7 TAC low-IPTG, reproducibility, hits heatmap and validation.

a-b, TAC results from low-IPTG induction. Experiments as described in Fig. 3b, c, but with low-IPTG plasmid-library induction for MOB (panel a; p1)26 and TransBac (panel b; p2)27. Prophage gene fold enrichment (FE) and p values (one-tailed Fisher’s exact test) are shown on top. cd, TAC reproducibility in control and experiment plates. Unprocessed colony opacity values of strains from control (ctrl) or experiment (exp) plates, derived from replicate plates 1 and 2, were plotted against each other. The overexpression library and IPTG concentration contained in each plate-pair are denoted above the plots. Results shown for MOB (panel c; p1)26 and TransBac (panel d; p2)27. Replication correlation (R2) is shown for each reproducibility plot. Green-colored dots correspond to the hits shown in panels a-b and in Fig. 3b, c. e, Non-parametric comparison of identified triggers across overexpression libraries. Trigger-genes identified from the MOB26 (p1) and the TransBac27 (p2) overexpression libraries were rank-ordered in percentiles based on their mean z-score difference value, as calculated from TAC screens conducted in low, or high-IPTG induction. Not available (NA) denotes genes for which measurements were flagged as problematic (see Methods) in the respective library/IPTG-concentration. Absent denotes genes which are missing from one library. Prophage and phage-related trigger-genes are in bold and underlined, respectively. f, Plasmids (p1/p2-trigger gene in green) were conjugated into E. coli BW25113 carrying p-empty, p-retron, p-retron-ΔrcaT, or p-rcaT. 384-colony-arrays of the transconjugants were pinned on LB plates containing appropriate antibiotics, arabinose, and IPTG concentrations (IPTG concentrations with strongest effect shown here). The y-axis represents the triggering-degree of each plasmid, measured as the colony opacity ratio of strains (p-retron-ΔrcaT + p1/p2-trigger) divided by the (p-retron + p1/p2-trigger). p1-control values were derived by measuring the same colony opacity ratio, for p1-blocker genes (n = 72 – 36 biological X 2 technical replicates), as a negative control. Ratios were calculated from n = 10 – 5 biological X 2 technical replicates for p1-trigger genes (except for p1-yfbO; n = 8 – 4 biological X 2 technical replicates), and from n = 24 – 12 biological X 2 technical replicates for p2-trigger genes. Representative colonies of strains carrying p1/p2-trigger plasmids are shown below the graphs. Horizontal lines denote the average fitness ratio, and error bars denote the standard deviation. The p1-control fitness data are shown in grey and p1-dam was plotted separately, to avoid compressing the scores of the rest of the trigger-genes. g, Plasmids (p1/p2-trigger gene in green, and p1-empty/p2-empty) from conjugation donor strains were conjugated with E. coli BW25113 carrying plasmids p-retron-ΔrcaT and p-retron. Transconjugants were grown for 5 h in LB with appropriate antibiotics, serially diluted, spotted on LB plates containing antibiotics, arabinose, and IPTG (no IPTG, low, or high), and incubated at 37 °C. Triggers specifically inhibit growth by activating RcaT, and triggering depends on induction levels.

Extended Data Fig. 8 TIC high-IPTG, reproducibility, hits heatmap, and validation.

ab, TIC results from high-IPTG induction. Experiments as described in Fig. 3d, e, but with low-IPTG plasmid-library induction for MOB (a; p1)26 and TransBac (b; p2)27. Prophage gene fold enrichment (FE) p values (one-tailed Fisher’s exact test) are shown on top. yfbO was selected as a blocker, despite not being over the cut-off, due to genetic linkage with yfbN. cd, TIC reproducibility in control and experiment plates – as in Extended Data Fig. 7c, d. Blue-colored dots correspond to the hits shown in panels ab and in Fig. 3d, e. e, Non-parametric comparison of identified blockers across overexpression libraries – as in Extended Data Fig. 7e. f, Plasmids (p1-blocker gene in blue and p1-empty in black) were conjugated into E. coli BW25113 carrying p-empty, p-retron, p-retron-ΔrcaT, or p-rcaT. 384-colony-arrays of the transconjugants were pinned on LB plates containing appropriate antibiotics, arabinose, and IPTG concentrations (IPTG concentrations with strongest effect shown here). The y-axis represents the blocking degree of each plasmid, measured as the colony opacity ratio of strains carrying the different blockers (p-rcaT + p1-blocker) divided by the average of a strain carrying trigger plasmids (p-rcaT + p1-trigger), as a negative control. Fitness ratios calculated from n = 8 – 4 biological X 2 technical replicates. The mean colony opacity for p1-trigger was derived from n = 110 – 55 biological X 2 technical replicates. Representative colonies of strains carrying p1-blocker plasmids shown below the graphs. Horizontal lines denote the average fitness ratio, and error bars denote the standard deviation. The p1-control fitness data are shown in grey (full toxicity of RcaT without any blocker). g, Plasmids (p1/p2-blocker gene in blue, and p1-empty/p2-empty) from conjugation donor strains were conjugated with E. coli BW25113 carrying plasmid p-rcaT or a p1-empty plasmid. Transconjugants were grown as in Extended Data Fig. 7g. Blockers specifically alleviate RcaT-mediated toxicity in a dose-dependent manner.

Extended Data Fig. 9 Dam-mediated retron-TA triggering.

a, Bacterial/phage dam homologues trigger Retron-Sen2. E. coli BW25113 carrying combinations of p-retron-ΔrcaT, p-retron, p1-damEc, p1-damSTm, p1-damP1 and p1-empty, were grown for 5 h in LB with appropriate antibiotics, serially diluted, spotted on LB plates with antibiotics, with/without arabinose, and IPTG (low), and incubated at 37 °C (n = 2 biological). b, Phage P1 dam triggers the chromosomal Retron-Eco9 in E. coli NILS-16. E. coli NILS-16 WT and ΔrcaT-Eco9 strains carrying plasmids p2-damP1 or p2-empty, were grown for 5 h in LB-hygromycin, serially diluted, spotted on LB-hygromycin plates, with or without IPTG, and incubated at 37 °C (n = 2 biological). c, Phage P1 Dam is expressed at comparable levels to the p2-damP1 plasmid. An E. coli BW25113 ΔlacY strain carrying an empty plasmid was infected with phage P1vir (MOI 2) for 0, 5, or 35 min (uninfected, early, and late infection). The same strain, but carrying a single-copy vector expressing DamP1 was grown in LB and the plasmid was induced with 60 μM of IPTG. The levels of the DamP1 protein were quantified by proteomics (TMT labeling), and DamP1 amounts were compared between phage-infection and plasmid-induction (X-axis). Phage infection or DamP1 induction does not alter the expression of endogenous Dam (Table S7), which has diverse enough sequence to be distinguished from DamP1 (Dam 39% identical for 266/754 residues of DamP1) (n = 2 biological). d, The DamP1 protein levels produced by P1vir are enough to trigger Retron-Eco9. The same strains as in panel b were grown in LB-hygromycin at 37 °C with IPTG (0 – 1000 μM), their optical density was measured at 8 h post-induction, and the fitness of the WT was compared to the ΔrcaT-Eco9 strain (X-axis). Vertical lines denote the mean fitness ratio, and horizontal lines denote standard deviation (n = 3 biological). e, Dam overexpression does not impact msDNA levels. msDNA was isolated from STm carrying combinations of plasmids p-retron-ΔrcaT, p1-dam, or p1-empty. Plasmids were co-induced with arabinose and IPTG (low). Extracted msDNA was electrophoresed in a TBE-polyacrylamide gel (n = 4 biological). f, RcaT in p-retronmut is functional. E. coli BW25113 ΔxseA strains carrying combinations of plasmids p-retronWT, p-retronmut, p1-dam, and p1-empty, were grown and spotted as described in panel a (n = 3 biological). g, p-retronmut produces same levels of msDNA as wild-type retron. msDNA was extracted from STm strains carrying p-retronWT, or p-retronmut. Extracted msDNA was electrophoresed in a TBE-polyacrylamide gel (n = 2 biological).

Extended Data Fig. 10 racC-recE are a blocker-trigger gene-pair.

a, RecE triggers the Retron-Eco9. E. coli BW25113 carrying combinations of plasmids p-retron-Eco9, p-retron-ΔrcaT-Eco9, p1-recE, and p1-empty, were grown for 5 h in LB with appropriate antibiotics, serially diluted, spotted on LB plates with antibiotics, arabinose, and IPTG (low), and incubated at 37 °C (n = 3 biological). b, RecE degrades mature and immature msDNA in vitro. msDNA was extracted from STm strains (WT, or ΔxseA) carrying plasmid p-retron-ΔrcaT. msDNA extracts were incubated with recombinant RecE (see Methods), and electrophoresed on a TBE-Polyacrylamide gel (n = 3 biological). c, RacC blocks RcaT-Eco9. E. coli BW25113 carrying combinations of plasmids p-rcaT-Eco9, p1-racC, and p1-empty, were grown and spotted as in panel a with different IPTG concentrations (n = 3 biological). d, RacC blocks cold-sensitivity of endogenous STm retron-antitoxin deletions. STm strains (WT, ΔrrtT, ΔxseA, ΔrnhA, Δmsrmsd) carrying plasmid p1-racC or the empty vector were grown and spotted as in panel a (n = 2 biological).

Extended Data Fig. 11 Phage defense assay data for Fig. 5a.

Growth curves of strains containing retrons in pSC101 vectors (3–5 copies/cell68,69,70,71,72) under the control of their native promoter or an empty vector (see Methods); n = 2 biological for each of the panels: a (37 °C) and b (25 °C).

Supplementary information

Reporting Summary

Peer Review File

Supplementary Table 1

Structural similarity search for RcaT domains.

Supplementary Table 2

RcaT and RT affinity purification proteomics data.

Supplementary Table 3

RcaT and RT phylogenetic trees source data.

Supplementary Table 4

Enrichment of prophage genes in TAC-TIC hit genes.

Supplementary Table 5

Raw and processed TAC-TIC fitness data.

Supplementary Table 6

Functional descriptions of trigger and blocker genes.

Supplementary Table 7

Phage P1vir proteomics.

Supplementary Table 8

Strains, plasmids and primers used in this study.

Rights and permissions

Springer Nature or its licensor 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

Bobonis, J., Mitosch, K., Mateus, A. et al. Bacterial retrons encode phage-defending tripartite toxin–antitoxin systems. Nature 609, 144–150 (2022). https://doi.org/10.1038/s41586-022-05091-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-022-05091-4

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 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