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:

Transposon-encoded nucleases use guide RNAs to promote their selfish spread

Abstract

Insertion sequences are compact and pervasive transposable elements found in bacteria, which encode only the genes necessary for their mobilization and maintenance1. IS200- and IS605-family transposons undergo ‘peel-and-paste’ transposition catalysed by a TnpA transposase2, but they also encode diverse, TnpB- and IscB-family proteins that are evolutionarily related to the CRISPR-associated effectors Cas12 and Cas9, respectively3,4. Recent studies have demonstrated that TnpB and IscB function as RNA-guided DNA endonucleases5,6, but the broader biological role of this activity has remained enigmatic. Here we show that TnpB and IscB are essential to prevent permanent transposon loss as a consequence of the TnpA transposition mechanism. We selected a family of related insertion sequences from Geobacillus stearothermophilus that encode several TnpB and IscB orthologues, and showed that a single TnpA transposase was broadly active for transposon mobilization. The donor joints formed upon religation of transposon-flanking sequences were efficiently targeted for cleavage by RNA-guided TnpB and IscB nucleases, and co-expression of TnpB and TnpA led to substantially greater transposon retention relative to conditions in which TnpA was expressed alone. Notably, TnpA and TnpB also stimulated recombination frequencies, surpassing rates observed with TnpB alone. Collectively, this study reveals that RNA-guided DNA cleavage arose as a primal biochemical activity to bias the selfish inheritance and spread of transposable elements, which was later co-opted during the evolution of CRISPR–Cas adaptive immunity for antiviral defence.

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: Pervasive distribution of IS200/IS605-like elements in G. stearothermophilus.
Fig. 2: TnpA catalyses DNA excision for multiple families of IS elements.
Fig. 3: TnpB and IscB target donor joint molecules produced by TnpA.
Fig. 4: Unbiased identification of TnpB/IscB TAM specificity by ChIP–seq and library assays.
Fig. 5: RNA-guided nucleases promote transposon survival via recombination at donor sites.

Similar content being viewed by others

Data availability

Next generation sequencing data are available in the National Center for Biotechnology Information (NCBI) Sequence Read Archive: SRX19058888SRX19058905, SRR23476356SRR23476358, and SRR24994123 (BioProject Accession: PRJNA925099 and PRJNA986543) and the Gene Expression Omnibus (GSE223127). The published genome used for ChIP–seq analyses was obtained from NCBI (GenBank: CP001509.3). The published genome used for bioinformatics analyses of the G. stearothermophilus genome was obtained from NCBI (GenBank: NZ_CP016552.1). Datasets generated and analysed in the current study are available from the corresponding author upon reasonable request.

Code availability

Custom scripts used for bioinformatics, TAM library analyses and ChIP–seq data analyses are available at GitHub (https://github.com/sternberglab/Meers_et_al_2023).

References

  1. Siguier, P., Gourbeyre, E., Varani, A., Ton-Hoang, B. & Chandler, M. Everyman’s guide to bacterial insertion sequences. Microbiol. Spectr. 3, MDNA3-0030-2014 (2015).

    Article  PubMed  Google Scholar 

  2. He, S. et al. The IS200/IS605 family and “peel and paste” single-strand transposition mechanism. Microbiol. Spectr. 3, MDNA3-0039-2014 (2015).

    Article  ADS  Google Scholar 

  3. Kapitonov, V. V., Makarova, K. S. & Koonin, E. V. ISC, a novel group of bacterial and archaeal DNA transposons that encode Cas9 homologs. J. Bacteriol. 198, 797–807 (2015).

    Article  PubMed  Google Scholar 

  4. Chylinski, K., Makarova, K. S., Charpentier, E. & Koonin, E. V. Classification and evolution of type II CRISPR–Cas systems. Nucleic Acids Res. 42, 6091–6105 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Altae-Tran, H. et al. The widespread IS200/IS605 transposon family encodes diverse programmable RNA-guided endonucleases. Science 374, 57–65 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  6. Karvelis, T. et al. Transposon-associated TnpB is a programmable RNA-guided DNA endonuclease. Nature 599, 692–696 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  7. Haudiquet, M., de Sousa, J. M., Touchon, M. & Rocha, E. P. C. Selfish, promiscuous and sometimes useful: how mobile genetic elements drive horizontal gene transfer in microbial populations. Phil. Trans. R. Soc. B 377, 20210234 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Feschotte, C. & Pritham, E. J. DNA transposons and the evolution of eukaryotic genomes. Annu. Rev. Genet. 41, 331–368 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Benler, S. & Koonin, E. V. Recruitment of mobile genetic elements for diverse cellular functions in prokaryotes. Front. Mol. Biosci. 9, 821197 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Zimmerly, S. & Semper, C. Evolution of group II introns. Mob. DNA 6, 7 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  11. Nakamura, T. M. & Cech, T. R. Reversing time: origin of telomerase. Cell 92, 587–590 (1998).

    Article  CAS  PubMed  Google Scholar 

  12. Liu, C., Zhang, Y., Liu, C. C. & Schatz, D. G. Structural insights into the evolution of the RAG recombinase. Nat. Rev. Immunol. 22, 353–370 (2022).

    Article  CAS  PubMed  Google Scholar 

  13. Koonin, E. V. & Makarova, K. S. Origins and evolution of CRISPR–Cas systems. Phil. Trans. R. Soc. B 374, 20180087 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Cosby, R. L., Chang, N. C. & Feschotte, C. Host–transposon interactions: conflict, cooperation, and cooption. Genes Dev. 33, 1098–1116 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Koonin, E. V., Makarova, K. S., Wolf, Y. I. & Krupovic, M. Evolutionary entanglement of mobile genetic elements and host defence systems: guns for hire. Nat. Rev. Genet. 21, 119–131 (2020).

    Article  CAS  PubMed  Google Scholar 

  16. Anzalone, A. V., Koblan, L. W. & Liu, D. R. Genome editing with CRISPR–Cas nucleases, base editors, transposases and prime editors. Nat. Biotechnol. 38, 824–844 (2020).

    Article  CAS  PubMed  Google Scholar 

  17. Krupovic, M., Makarova, K. S., Forterre, P., Prangishvili, D. & Koonin, E. V. Casposons: a new superfamily of self-synthesizing DNA transposons at the origin of prokaryotic CRISPR–Cas immunity. BMC Biol. 12, 36 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  18. Özcan, A. et al. Type IV CRISPR RNA processing and effector complex formation in Aromatoleum aromaticum. Nat. Microbiol. 4, 89–96 (2019).

    Article  PubMed  Google Scholar 

  19. Seed, K. D., Lazinski, D. W., Calderwood, S. B. & Camilli, A. A bacteriophage encodes its own CRISPR/Cas adaptive response to evade host innate immunity. Nature 494, 489–491 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  20. Peters, J. E., Makarova, K. S., Shmakov, S. & Koonin, E. V. Recruitment of CRISPR–Cas systems by Tn7-like transposons. Proc. Natl Acad. Sci. USA 114, e7358–e7366 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  21. Sternberg, S. H., Redding, S., Jinek, M., Greene, E. C. & Doudna, J. A. DNA interrogation by the CRISPR RNA-guided endonuclease Cas9. Nature 507, 62–67 (2014).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  22. Strecker, J. et al. RNA-guided DNA insertion with CRISPR-associated transposases. Science 365, 48–53 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  23. Chandler, M. et al. Breaking and joining single-stranded DNA: the HUH endonuclease superfamily. Nat. Rev. Microbiol. 11, 525–538 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Pasternak, C. et al. ISDra2 transposition in Deinococcus radiodurans is downregulated by TnpB. Mol. Microbiol. 88, 443–455 (2013).

    Article  CAS  PubMed  Google Scholar 

  25. Filée, J., Siguier, P. & Chandler, M. I am what I eat and I eat what I am: acquisition of bacterial genes by giant viruses. Trends Genet. 23, 10–15 (2007).

    Article  PubMed  Google Scholar 

  26. Bao, W. & Jurka, J. Homologues of bacterial TnpB_IS605 are widespread in diverse eukaryotic transposable elements. Mob DNA 4, 12 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Saito, M. et al. Fanzor is a eukaryotic programmable RNA-guided endonuclease. Nature 620, 660–668 (2023).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  28. Kersulyte, D., Mukhopadhyay, A. K., Shirai, M., Nakazawa, T. & Berg, D. E. Functional organization and insertion specificity of IS607, a chimeric element of Helicobacter pylori. J. Bacteriol. 182, 5300–5308 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Weinberg, Z., Perreault, J., Meyer, M. M. & Breaker, R. R. Exceptional structured noncoding RNAs revealed by bacterial metagenome analysis. Nature 462, 656–659 (2009).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  30. Gomes-Filho, J. V. et al. Sense overlapping transcripts in IS1341-type transposase genes are functional non-coding RNAs in archaea. RNA Biol. 12, 490–500 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  31. Ibrahim, A., Vêncio, R. Z. N., Lorenzetti, A. P. R. & Koide, T. Halobacterium salinarum and Haloferax volcanii comparative transcriptomics reveals conserved transcriptional processing sites. Genes 12, 1018 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Nety, S. P. et al. The transposon-encoded protein TnpB processes its own mRNA into ωRNA for guided nuclease activity. CRISPR J. 6, 232–242 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Harrington, L. B. et al. A thermostable Cas9 with increased lifetime in human plasma. Nat. Commun. 8, 1424 (2017).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  34. Barabas, O. et al. Mechanism of IS200/IS605 family DNA transposases: activation and transposon-directed target site selection. Cell 132, 208–220 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Hickman, A. B. et al. DNA recognition and the precleavage state during single-stranded DNA transposition in D. radiodurans. EMBO J. 29, 3840–3852 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Morero, N. R. et al. Targeting IS608 transposon integration to highly specific sequences by structure-based transposon engineering. Nucleic Acids Res. 46, 4152–4163 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Lavatine, L. et al. Single strand transposition at the host replication fork. Nucleic Acids Res. 44, 7866–7883 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Pasternak, C. et al. Irradiation-induced Deinococcus radiodurans genome fragmentation triggers transposition of a single resident insertion sequence. PLoS Genet. 6, e1000799 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Cox, M. M. Recombinational DNA repair of damaged replication forks in Escherichia coli: questions. Annu. Rev. Genet. 35, 53–82 (2001).

    Article  CAS  PubMed  Google Scholar 

  40. Leenay, R. T. & Beisel, C. L. Deciphering, communicating, and engineering the CRISPR PAM. J. Mol. Biol. 429, 177–191 (2017).

    Article  CAS  PubMed  Google Scholar 

  41. Wu, X. et al. Genome-wide binding of the CRISPR endonuclease Cas9 in mammalian cells. Nat. Biotechnol. 32, 670–676 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Swarts, D. C., van der Oost, J. & Jinek, M. Structural basis for guide RNA processing and seed-dependent DNA targeting by CRISPR–Cas12a. Mol. Cell 66, 221–233.e224 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Jiang, F., Zhou, K., Ma, L., Gressel, S. & Doudna, J. A. A Cas9–guide RNA complex preorganized for target DNA recognition. Science 348, 1477–1481 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  44. Belfort, M. & Bonocora, R. P. Homing endonucleases: from genetic anomalies to programmable genomic clippers. Methods Mol. Biol. 1123, 1–26 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Tourasse, N. J., Stabell, F. B. & Kolstø, A. B. Survey of chimeric IStron elements in bacterial genomes: multiple molecular symbioses between group I intron ribozymes and DNA transposons. Nucleic Acids Res. 42, 12333–12351 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Ton-Hoang, B. et al. Single-stranded DNA transposition is coupled to host replication. Cell 142, 398–408 (2010).

    Article  CAS  PubMed  Google Scholar 

  47. Hickman, A. B. & Dyda, F. DNA transposition at work. Chem. Rev. 116, 12758–12784 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Roberts, D., Hoopes, B. C., McClure, W. R. & Kleckner, N. IS10 transposition is regulated by DNA adenine methylation. Cell 43, 117–130 (1985).

    Article  CAS  PubMed  Google Scholar 

  49. Yin, J. C., Krebs, M. P. & Reznikoff, W. S. Effect of dam methylation on Tn5 transposition. J. Mol. Biol. 199, 35–45 (1988).

    Article  CAS  PubMed  Google Scholar 

  50. Nicolas, E. et al. The Tn3-family of replicative transposons. Microbiol. Spectr. 3, https://doi.org/10.1128/microbiolspec.mdna3-0060-2014 (2015).

  51. Li, W. & Godzik, A. Cd-hit: a fast program for clustering and comparing large sets of protein or nucleotide sequences. Bioinformatics 22, 1658–1659 (2006).

    Article  CAS  PubMed  Google Scholar 

  52. Katoh, K. & Standley, D. M. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol. Biol. Evol. 30, 772–780 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Capella-Gutiérrez, S., Silla-Martínez, J. M. & Gabaldón, T. trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25, 1972–1973 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  54. Minh, B. Q. et al. IQ-TREE 2: new models and efficient methods for phylogenetic inference in the genomic era. Mol. Biol. Evol. 37, 1530–1534 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Kalvari, I. et al. Rfam 14: expanded coverage of metagenomic, viral and microRNA families. Nucleic Acids Res. 49, D192–d200 (2021).

    Article  CAS  PubMed  Google Scholar 

  56. Nawrocki, E. P. & Eddy, S. R. Infernal 1.1: 100-fold faster RNA homology searches. Bioinformatics 29, 2933–2935 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Yao, Z., Weinberg, Z. & Ruzzo, W. L. CMfinder-a covariance model based RNA motif finding algorithm. Bioinformatics 22, 445–452 (2006).

    Article  CAS  PubMed  Google Scholar 

  58. Will, S., Joshi, T., Hofacker, I. L., Stadler, P. F. & Backofen, R. LocARNA-P: accurate boundary prediction and improved detection of structural RNAs. RNA 18, 900–914 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Edgar, R. C. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 32, 1792–1797 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Rivas, E. RNA structure prediction using positive and negative evolutionary information. PLoS Comput. Biol. 16, e1008387 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  61. Hyatt, D. et al. Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinf. 11, 119 (2010).

    Article  Google Scholar 

  62. 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 

  63. Kechin, A., Boyarskikh, U., Kel, A. & Filipenko, M. cutPrimers: a new tool for accurate cutting of primers from reads of targeted next generation sequencing. J Comput. Biol. 24, 1138–1143 (2017).

    Article  CAS  PubMed  Google Scholar 

  64. Smith, T., Heger, A. & Sudbery, I. UMI-tools: modeling sequencing errors in unique molecular identifiers to improve quantification accuracy. Genome Res. 27, 491–499 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Vasimuddin, M., Misra, S., Li, H. & Aluru, S. Efficient architecture-aware acceleration of BWA-MEM for multicore systems. In 2019 IEEE International Parallel and Distributed Processing Symposium (IPDPS) 314–324 (IEEE, 2019).

  66. Danecek, P. et al. Twelve years of SAMtools and BCFtools. Gigascience 10, giab008 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  67. Ramírez, F. et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 44, W160–W165 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  68. Robinson, J. T., Thorvaldsdottir, H., Turner, D. & Mesirov, J. P. igv.js: an embeddable JavaScript implementation of the Integrative Genomics Viewer (IGV). Bioinformatics 39, btac830 (2023).

    Article  CAS  PubMed  Google Scholar 

  69. Hoffmann, F. T. et al. Selective TnsC recruitment enhances the fidelity of RNA-guided transposition. Nature 609, 384–393 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  70. Zhang, Y. et al. Model-based analysis of ChIP–seq (MACS). Genome Biol. 9, R137 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  71. Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Bailey, T. L. et al. MEME SUITE: tools for motif discovery and searching. Nucleic Acids Res. 37, W202–W208 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Crooks, G. E., Hon, G., Chandonia, J. M. & Brenner, S. E. WebLogo: a sequence logo generator. Genome Res. 14, 1188–1190 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank D. R. Gelsinger and A. Bernheim for helpful bioinformatics discussions, L. E. Berchowitz for helpful RNA biology discussions, J. C. Cheong for technical support, L. F. Landweber for qPCR instrument access, and the JP Sulzberger Columbia Genome Center for NGS support. C.M. was supported by NIH Postdoctoral Fellowship F32 GM143924-01A1. M.W.G.W. was supported by a National Science Foundation Graduate Research Fellowship. This research was supported by NSF Faculty Early Career Development Program (CAREER) Award 2239685, and by a generous start-up package from the Columbia University Irving Medical Center Dean’s Office and the Vagelos Precision Medicine Fund (to S.H.S.).

Author information

Authors and Affiliations

Authors

Contributions

C.M. and S.H.S. conceived and designed the project. C.M. performed most experiments, with assistance from S.R.P. on plasmid interference assays, and F.T.H. for transposon excision assays and ChIP–seq experiments and analyses. C.M. and H.C.L. performed all bioinformatics analyses. M.W.G.W. assisted with TAM library design and NGS analysis. J.G. assisted with plasmid interference assays. S.T. performed G. stearothermophilus RNA-seq experiments. C.M. and S.H.S. discussed the data and wrote the manuscript, with input from all authors.

Corresponding author

Correspondence to Samuel H. Sternberg.

Ethics declarations

Competing interests

Columbia University has filed US Patent Application Number 63/379,082 related to this work, for which C.M. and S.H.S. are inventors. S.H.S. is a co-founder and scientific advisor to Dahlia Biosciences, a scientific advisor to CrisprBits and Prime Medicine, and an equity holder in Dahlia Biosciences and CrisprBits.

Peer review

Peer review information

Nature thanks the 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 figures and tables

Extended Data Fig. 1 Bioinformatic analyses of IscB and TnpB homologs.

a, Phylogenetic tree of IscB and IsrB protein homologs; IscB contains HNH and RuvC nuclease domains, whereas IsrB lacks the HNH nuclease. Genetic neighborhood analyses demonstrate that most homologs are encoded proximal to a predicted ωRNA (inner ring), whereas the vast majority do not reside near a predicted tyrosine-family TnpA transposase gene (outer ring). The GstIscB homolog encoded by ISGst6 used in this study is indicated. Bootstrap values are indicated for major nodes. b, Schematic of a non-autonomous IS element encoding IscB and its associated ωRNA; a structural covariation model is shown in the inset (top). The red rectangle and dotted black line indicate the transposon boundaries, and the guide portion of the ωRNA is shown in blue. LE and RE, transposon left end and right end. c, Orientation bias of the nearest upstream ORFs to the indicated protein-coding gene (iscB, tnpB, or IS630 transposase), demonstrating that IS elements encoding IscB are preferentially integrated (or retained) in an orientation matching that of the upstream gene. The y-axis indicates the frequency of ORFs containing the same orientation, at a distance from the gene start codon defined by the x-axis. 242 bp represents the average length of IscB-associated ωRNAs upstream of IscB ORF. The spike at ~0-bp for TnpB corresponds to IS elements that encode adjacent/overlapping tnpA and tnpB genes. IS630 transposase genes are included as a representative gene from unrelated transposable elements. d, Phylogenetic tree of TnpB homologs, with bootstrap values shown for major nodes. Genetic neighborhood analyses demonstrate that most homologs are encoded proximal to a predicted ωRNA (inner ring), whereas the vast majority do not reside near a predicted tyrosine- or serine-family TnpA transposase genes (outer rings). Interestingly, TnpB homologs are associated with two distinct transposase families in prokaryotes: tyrosine transposases (denoted TnpA (Y)) within IS200/605-family elements, and serine transposases (denote TnpA (S)) within IS607-family elements. GstTnpB homologs used in this study are highlighted, along with the predicted structures of their associated ωRNAs based on covariance modeling. The ωRNA encoded by ISGst2 did not show strong covariation in structure and was therefore omitted. e, Read coverage for RNA-seq data from G. stearothermophilus strain DSM 458, demonstrating expression of putative ωRNAs from each of the indicated ISGst families. ISGst5 is a PATE-like element that lacks any protein-coding ORFs. Other TnpB-associated ωRNAs are encoded within/downstream of the ORF, whereas IscB-associated ωRNAs are encoded upstream of the ORF.

Extended Data Fig. 2 Classification of IS605-family elements encoded by G. stearothermophilus strain DSM 458.

a, DNA multiple sequence alignment of transposon left ends (LE) for IS200/IS605-family elements from G. stearothermophilus. The weblogo (top) is built from 47 unique elements (Supplementary Figs. 2 & 3), and one representative sequence from each family is shown below, with the TAM highlighted in yellow and DNA guide sequences highlighted in red. Nucleotides highlighted in black exhibit covarying mutations, relative to ISGst2. TAM, transposon-adjacent motif; the dotted red line indicates the transposon boundary. b, DNA multiple sequence alignment of transposon right ends (RE) for IS200/IS605-family elements from G. stearothermophilus, shown as in a. TEM, transposon-encoded motif is shown in orange. c, Phylogenetic tree of ISGst elements based on the transposon left end. Each colored clade encodes an associated TnpB/IscB protein homolog and is flanked by the indicated TAM sequence. d, Phylogenetic tree of ISGst elements based on the transposon right end, shown as in b but with TEM sequence in lieu of TAM. e, Schematic of PATEs (palindrome-associated transposable elements) related to ISGst2 and ISGst5, which contain similar transposon ends but no protein-coding genes. The percent sequence identity between shaded regions (black) is shown, as are the genomic accession IDs and coordinates.

Extended Data Fig. 3 Specificity and efficiency of transposon DNA excision by TnpA.

a, Schematic of heterologous E. coli transposon excision assay. Plasmids encode TnpA and a mini-transposon (Mini-Tn) substrate, whose loss is monitored by PCR using the indicated primers. The expected sizes of PCR products generated from donor joints that are produced upon religation of flanking sequences are shown, for both ISGst2 and H. pylori IS608. b, TnpA homologs do not cross-react with distinct IS elements, as assessed by analytical PCR. Cell lysates were tested after overnight expression of TnpA in combination with a mini-Tn substrate from either G. stearothermophilus (G) or H. pylori (H), and PCR products were resolved by agarose gel electrophoresis. M refers to catalytically inactive mutants. Note that HpyTnpA is substantially more active for DNA excision than GstTnpA under the tested conditions. U, unexcised; E, excised. c, Schematic of qPCR assay to quantify excision frequencies, in which one of the two primers anneals directly to the donor joint formed upon mini-Tn excision and religation. d, Comparison of simulated excision frequencies, generated by mixing clonally excised and unexcised lysate in known ratios, versus experimentally determined integration efficiencies measured by qPCR. e, qPCR-based quantification of TnpA-mediated excision of an ISGst2 mini-Tn substrate in E. coli. Excised refers to a cloned excision product; TnpAM denotes a TnpA mutant (Y125A); ND, not detected above a 0.0001% threshold. Bars indicate mean ± standard deviation (n = 3). f, Schematic of mini-Tn for ISGst3 element, highlighting the subterminal palindromic transposon ends located on the top strand (top). Transposon-adjacent and transposon-encoded motifs (TAM and TEM) are shown in yellow and orange, respectively; DNA guides are shown in orange, and their putative base-pairing interactions are indicated; dotted lines indicate transposon boundaries and thus the sites of ssDNA cleavage and religation. Sanger sequencing of excision events confirms the identity of the expected donor joint product formed upon transposon loss (bottom). Sanger sequencing results are duplicated from Fig. 2d. g, Schematic and Sanger sequencing data as in f, but for a modified ISGst3 substrate containing TEM mutations. Experimentally detected products erroneously excise at an alternative, aberrant TEM-like sequence located outside of the native transposon boundary (orange), presumably because of the need to maintain cognate base-pairing between the DNA guide (orange and TEM (blue).

Extended Data Fig. 4 Mating-out assay to monitor transposition of ISGst3.

a, Schematic of mating-out assay, in which transposition events into the F-plasmid are monitored via drug selection. E. coli donor cells carrying an F-plasmid were transformed with a plasmid encoding either TnpA and ISGst3-derived mini-Tn (pDonor1) or an autonomous ISGst3 element with tnpA and tnpB (pDonor2). After induction of TnpA, conjugation was used to transfer the F-plasmid into the recipient strain, and transposition events were quantified by selecting for recipient cells (RifR) containing spectinomycin (F+) and kanamycin (mini-Tn+) resistance. b, Transposition frequency of ISGst3 mini-Tn deriving from pDonor1 into the F-plasmid was measured with and without tnpA. Bars indicate mean ± s.d. (n = 6). c, Drug-selected cells from mating-out assays from pDonor1 contain TAM-proximal IS insertions, as evidenced by long-read Nanopore sequencing. A genetic map of the F-plasmid is shown, along with the location of distinct ISGst3-derived mini-Tn integration events. The insets show a zoom-in view of each integration site at the nucleotide level, with the TAM highlighted in yellow and the integration site denoted by an arrow. d, Transposition frequency of an autonomous ISGst3 deriving from pDonor2 into the F-plasmid was measured with active and catalytically inactive forms of TnpA and TnpB. M, TnpA Y125A mutant; d, TnpB D196A mutant; ND, not detected. Bars indicate mean ± s.d. (n = 3).

Extended Data Fig. 5 Optimization and testing of DNA cleavage parameters with TnpB/IscB nucleases.

a, Promoter screen to optimize conditions for E. coli-based interference assays using plasmid-encoded ωRNA and TnpB2. P1 indicates the promoter for ωRNA expression, whereas P2 indicates the promoter for TnpB expression. Transformants with a targeting (T) or non-targeting (NT) ωRNA-pTarget combination were serially diluted, plated on selective media, and cultured at 37 °C for 24 h. b, Results from plasmid interference assays with HpyTnpB (IS608) and DraTnpB (ISDra2) using ωRNAs that target native donor joint products, which revealed an absence of activity for HpyTnpB. Experiments were performed as in a. c, DNA cleavage by TnpB2 is highly sensitive to TAM mutations, as assessed by plasmid interference assays. Data are shown as in a, with the indicated TAM sequences; TTTAT denotes the WT TAM, and NT denotes a non-targeting control. d, DNA cleavage by IscB is highly sensitive to TAM mutations, as assessed by plasmid interference assays. Data are shown as in a, with the indicated TAM sequences; TTCAT denotes the WT TAM, and NT denotes a non-targeting control. e, TnpB2 is only active for targeted genomic DNA cleavage using select ωRNAs, as assessed by genome targeting assays. Transformants with a non-targeting (NT) or one of three lacZ-targeting guides (T1–T3) were serially diluted, plated on selective media, and cultured at 37 °C for 24 h. f, Schematic of E. coli-based genome targeting assay, in which RNA-guided DNA cleavage of lacZ by TnpB results in cell death. Guides were designed to target three sites within the lacZ coding sequence. g, TnpB1 is active for targeted genomic DNA cleavage at elevated temperatures, as assessed by genome targeting assays. Transformants with a non-targeting (NT) or one of three lacZ-targeting guides (T1–T3) were serially diluted, plated on selective media, and cultured at 37 °C or 45 °C for 24 h, as shown. h, Schematic of E. coli-based plasmid interference assay using a native ISGst3 element. The transposon was cloned from G. stearothermophilus gDNA onto the pEffector plasmid, with 40 bp of flanking sequence to preserve the natural ωRNA. Targeted cleavage of pTarget results in a loss of kanamycin resistance and cell lethality on selective LB-agar plates. i, TnpB2 encoded by a native ISGst3 transposon is active for targeted cleavage of its donor joint. Transformants with a targeting (T) or non-targeting (NT) ωRNA-pTarget combination were serially diluted, plated on selective media, and cultured at 37 °C for 24 h. dTnpB2 contains an inactivating D196A mutation, whereas ΔTnpB2 contains a stop codon at codon 32. j, Schematic of E. coli-based genome targeting assay as in e, but with pEffector containing a native ISGst3 transposon cloned into lacZ. RNA-guided DNA cleavage of genomic lacZ results in cell death. k, TnpB2 encoded by a native ISGst3 transposon is active for targeted genomic cleavage of a donor joint mimic. Transformants were serially diluted, plated on selective media, and cultured at 37 °C for 24 h. dTnpB2 contains an inactivating D196A mutation, whereas ΔTnpB2 contains a stop codon at codon 32.

Extended Data Fig. 6 Off-target ChIP-seq DNA binding analyses.

a, ChIP-seq experiments reveal recruitment of dIscB to the target site (blue triangle) with a targeting ωRNA. Genome-wide representation of ChIP-seq data for dIscB reshown from Fig. 4b, with addition of a second replicate. Representative off-target sites for dIscB identified by MACS3 are highlighted (OT1-4) and analyzed in the right panel. The middle panel highlights analysis of off-target binding events using MEME-ChIP, as shown in Fig. 4b. Motifs shared by off-target peaks reveal conserved TAM sequences and little conservation of the adjacent seed sequence. The sequence of the 5′ end of the corresponding ωRNA is shown below each motif. n indicates the number of peaks contributing to the motif and their percentage of total peaks called by MACS3; E, E-value significance of the motif generated from the MEME-ChIP analysis. DNA sequences corresponding to the on-target and off-target sites are shown on the right, with TAM (yellow) and mismatches (orange) highlighted. b, ChIP-seq experiments reveal recruitment of dTnpB2 to the target site (blue triangle) with a targeting ωRNA. Data shown as in a. Similar to dIscB, dTnpB2 shows limited seed sequence requirements. c, ChIP-seq experiments reveal recruitment of dCas9 to the target site (blue triangle) with a targeting ωRNA. Data shown as in a. Analysis of off-target sites reveal a short (3–4 nt) seed sequence adjacent to the PAM motif. d, ChIP-seq experiments reveal recruitment of dCas12a to the target site (blue triangle) with a targeting ωRNA. Data shown as in a. Analysis of off-target sites reveals a short (5–7 nt) seed sequence adjacent to the PAM motif.

Extended Data Fig. 7 qPCR analysis of IS element loss upon TnpA and TnpB co-expression.

a, TnpB promotes transposon retention at the donor site, as assessed by blue-white colony screening for the experiment in Fig. 5a–c; images of representative LB-agar plates are shown. TnpAM, TnpA Y125A mutant; dTnpB, TnpB D196A mutant. b, Schematic of qPCR-based strategy for quantifying excision. Primers are designed to flank the donor joint generated upon excision and religation. Selective PCR conditions with a shortened extension time allows for reduced amplification of the starting locus containing the mini-Tn. rssA was used as a reference gene for ΔCq calculations. c, Comparison of simulated excision frequencies, generated by mixing clonally excised and unexcised lysates in known ratios, versus experimentally determined integration efficiencies measured by qPCR. d, qPCR-based quantification of transposon excision for experiments presented in Fig. 5a–c. TnpA was present in all conditions marked with orange bars, and Empty refers to an experiment lacking TnpB; Excised and Unexcised Marker labels shown in blue correspond to clonal control strains with either wild-type lacZ or mini-Tn-interrupted lacZ, respectively. The detection limit is based on qPCR experiments tested on simulated excision frequency samples shown in b. Bars indicate mean ± s.d. (n = 3–9). e, Images of representative LB-agar plates accompanying the experiments presented in Fig. 5d–f, highlighting the roles of TnpA and TnpB in transposon maintenance and spread via recombination. TnpAM, TnpA Y125A mutant; dTnpB, TnpB D196A mutant.

Supplementary information

Supplementary Figures

This file contains Supplementary Figs. 1–3.

Reporting Summary

Supplementary Tables

This file contains Supplementary Tables 1–5.

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

Meers, C., Le, H.C., Pesari, S.R. et al. Transposon-encoded nucleases use guide RNAs to promote their selfish spread. Nature 622, 863–871 (2023). https://doi.org/10.1038/s41586-023-06597-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-023-06597-1

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

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