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.

  • Letter
  • Published:

A jumbo phage that forms a nucleus-like structure evades CRISPR–Cas DNA targeting but is vulnerable to type III RNA-based immunity

Abstract

CRISPR–Cas systems provide bacteria with adaptive immunity against bacteriophages1. However, DNA modification2,3, the production of anti-CRISPR proteins4,5 and potentially other strategies enable phages to evade CRISPR–Cas. Here, we discovered a Serratia jumbo phage that evades type I CRISPR–Cas systems, but is sensitive to type III immunity. Jumbo phage infection resulted in a nucleus-like structure enclosed by a proteinaceous phage shell—a phenomenon only reported recently for distantly related Pseudomonas phages6,7. All three native CRISPR–Cas complexes in Serratia—type I-E, I-F and III-A—were spatially excluded from the phage nucleus and phage DNA was not targeted. However, the type III-A system still arrested jumbo phage infection by targeting phage RNA in the cytoplasm in a process requiring Cas7, Cas10 and an accessory nuclease. Type III, but not type I, systems frequently targeted nucleus-forming jumbo phages that were identified in global viral sequence datasets. The ability to recognize jumbo phage RNA and elicit immunity probably contributes to the presence of both RNA- and DNA-targeting CRISPR–Cas systems in many bacteria1,8. Together, our results support the model that jumbo phage nucleus-like compartments serve as a barrier to DNA-targeting, but not RNA-targeting, defences, and that this phenomenon is widespread among jumbo phages.

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: Jumbo phage PCH45 evades type I, but not type III, CRISPR–Cas immunity.
Fig. 2: The jumbo phage assembles a DNA-containing protein shell during infection that excludes CRISPR–Cas complexes.
Fig. 3: Jumbo phage infection is inhibited by the type III CRISPR–Cas system.
Fig. 4: Jumbo phages are targeted by type III CRISPR–Cas systems in different bacterial classes.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author on request. The genome sequence of bacteriophage PCH45 is available in GenBank under accession number MN334766. Data and R scripts for the bioinformatics analyses are available at https://github.com/JacksonLab/Jumbophages.

References

  1. Hille, F. et al. The biology of CRISPR-Cas: backward and forward. Cell 172, 1239–1259 (2018).

    Article  CAS  PubMed  Google Scholar 

  2. Bryson, A. L. et al. Covalent modification of bacteriophage T4 DNA inhibits CRISPR-Cas9. mBio 6, e00648 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Vlot, M. et al. Bacteriophage DNA glucosylation impairs target DNA binding by type I and II but not by type V CRISPR–Cas effector complexes. Nucleic Acids Res. 46, 873–885 (2018).

    CAS  PubMed  Google Scholar 

  4. Pawluk, A., Davidson, A. R. & Maxwell, K. L. Anti-CRISPR: discovery, mechanism and function. Nat. Rev. Microbiol. 16, 12–17 (2018).

    CAS  PubMed  Google Scholar 

  5. Bondy-Denomy, J., Pawluk, A., Maxwell, K. L. & Davidson, A. R. Bacteriophage genes that inactivate the CRISPR/Cas bacterial immune system. Nature 493, 429–432 (2013).

    CAS  PubMed  Google Scholar 

  6. Chaikeeratisak, V. et al. Assembly of a nucleus-like structure during viral replication in bacteria. Science 355, 194–197 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Chaikeeratisak, V. et al. The phage nucleus and tubulin spindle are conserved among large Pseudomonas phages. Cell Rep. 20, 1563–1571 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Makarova, K. S., Wolf, Y. I. & Koonin, E. V. Classification and nomenclature of CRISPR-Cas systems: where from here? CRISPR J. 1, 325–336 (2018).

    PubMed  PubMed Central  Google Scholar 

  9. Barrangou, R. et al. CRISPR provides acquired resistance against viruses in prokaryotes. Science 315, 1709–1712 (2007).

    CAS  PubMed  Google Scholar 

  10. Jackson, S. A. et al. CRISPR-Cas: adapting to change. Science 356, eaal5056 (2017).

    PubMed  Google Scholar 

  11. Brouns, S. J. J. et al. Small CRISPR RNAs guide antiviral defense in prokaryotes. Science 321, 960–964 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Hale, C. R. et al. RNA-guided RNA cleavage by a CRISPR RNA-Cas protein complex. Cell 139, 945–956 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Garneau, J. E. et al. The CRISPR/Cas bacterial immune system cleaves bacteriophage and plasmid DNA. Nature 468, 67–71 (2010).

    CAS  PubMed  Google Scholar 

  14. Patterson, A. G. et al. Quorum sensing controls adaptive immunity through the regulation of multiple CRISPR-Cas systems. Mol. Cell 64, 1102–1108 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Yuan, Y. & Gao, M. Jumbo bacteriophages: an overview. Front. Microbiol. 8, 403 (2017).

    PubMed  PubMed Central  Google Scholar 

  16. Mesyanzhinov, V. V. et al. The genome of bacteriophage φKZ of Pseudomonas aeruginosa. J. Mol. Biol. 317, 1–19 (2002).

    CAS  PubMed  Google Scholar 

  17. Ceyssens, P. J. et al. Development of giant bacteriophage φKZ is independent of the host transcription apparatus. J. Virol. 88, 10501–10510 (2014).

    PubMed  PubMed Central  Google Scholar 

  18. Jackson, S. A., Birkholz, N., Malone, L. M. & Fineran, P. C. Imprecise spacer acquisition generates CRISPR-Cas immune diversity through primed adaptation. Cell Host Microbe 25, 250–260 (2019).

    CAS  PubMed  Google Scholar 

  19. Kraemer, J. A. et al. A phage tubulin assembles dynamic filaments by an atypical mechanism to center viral DNA within the host cell. Cell 149, 1488–1499 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Erb, M. L. et al. A bacteriophage tubulin harnesses dynamic instability to center DNA in infected cells. eLife 3, e03197 (2014).

    PubMed Central  Google Scholar 

  21. Samai, P. et al. Co-transcriptional DNA and RNA cleavage during type III CRISPR-Cas immunity. Cell 161, 1164–1174 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Goldberg, G. W., Jiang, W., Bikard, D. & Marraffini, L. A. Conditional tolerance of temperate phages via transcription-dependent CRISPR–Cas targeting. Nature 514, 633–637 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Jia, N. et al. Type III-A CRISPR-Cas Csm complexes: assembly, periodic RNA cleavage, DNase activity regulation, and autoimmunity. Mol. Cell 73, 264–277 (2019).

    CAS  PubMed  Google Scholar 

  24. Liu, T. Y., Iavarone, A. T. & Doudna, J. A. RNA and DNA targeting by a reconstituted Thermus thermophilus type III-A CRISPR-Cas system. PLoS ONE 12, e0170552 (2017).

    PubMed  PubMed Central  Google Scholar 

  25. Niewoehner, O. et al. Type III CRISPR–Cas systems produce cyclic oligoadenylate second messengers. Nature 548, 543–548 (2017).

    CAS  PubMed  Google Scholar 

  26. Kazlauskiene, M., Kostiuk, G., Venclovas, Č., Tamulaitis, G. & Siksnys, V. A cyclic oligonucleotide signaling pathway in type III CRISPR-Cas systems. Science 357, 605–609 (2017).

    CAS  PubMed  Google Scholar 

  27. Rouillon, C., Athukoralage, J. S., Graham, S., Gruschow, S. & White, M. F. Control of cyclic oligoadenylate synthesis in a type III CRISPR system. eLife 7, e36734 (2018).

    PubMed  PubMed Central  Google Scholar 

  28. Varble, A. & Marraffini, L. A. Three new Cs for CRISPR: collateral, communicate, cooperate. Trends Genet. 35, 446–456 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Lau, R. K. et al. Structure and mechanism of a cyclic trinucleotide-activated bacterial endonuclease mediating bacteriophage immunity. Preprint at https://doi.org/10.1101/694703 (2019).

  30. Silas, S. et al. Direct CRISPR spacer acquisition from RNA by a natural reverse transcriptase-Cas1 fusion protein. Science 351, aad4234 (2016).

    PubMed  PubMed Central  Google Scholar 

  31. Hynes, A. P., Villion, M. & Moineau, S. Adaptation in bacterial CRISPR-Cas immunity can be driven by defective phages. Nat. Commun. 5, 4399 (2014).

    CAS  PubMed  Google Scholar 

  32. Mendoza, S. D. et al. A nucleus-like compartment shields bacteriophage DNA from CRISPR-Cas and restriction nucleases. Preprint at https://doi.org/10.1101/370791 (2018).

  33. Watson, B. N. J., Staals, R. H. J. & Fineran, P. C. CRISPR-Cas-mediated phage resistance enhances horizontal gene transfer by transduction. mBio 9, e02406-17 (2018).

    PubMed  PubMed Central  Google Scholar 

  34. Frampton, R. A. et al. Identification of bacteriophages for biocontrol of the kiwifruit canker phytopathogen Pseudomonas syringae pv. actinidiae. Appl. Environ. Microbiol. 80, 2216–2228 (2014).

    PubMed  PubMed Central  Google Scholar 

  35. Bankevich, A. et al. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J. Comput. Biol. 19, 455–477 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Brettin, T. et al. RASTtk: a modular and extensible implementation of the RAST algorithm for building custom annotation pipelines and annotating batches of genomes. Sci. Rep. 5, 8365 (2015).

    PubMed  PubMed Central  Google Scholar 

  37. Lowe, T. M. & Chan, P. P. tRNAscan-SE On-line: integrating search and context for analysis of transfer RNA genes. Nucleic Acids Res. 44, W54–W57 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Carver, T., Thomson, N., Bleasby, A., Berriman, M. & Parkhill, J. DNAPlotter: circular and linear interactive genome visualization. Bioinformatics 25, 119–120 (2009).

    CAS  PubMed  Google Scholar 

  39. Bao, Y., Chetvernin, V. & Tatusova, T. Improvements to pairwise sequence comparison (PASC): a genome-based web tool for virus classification. Arch. Virol. 159, 3293–3304 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Sullivan, M. J., Petty, N. K. & Beatson, S. A. Easyfig: a genome comparison visualizer. Bioinformatics 27, 1009–1010 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Meier-Kolthoff, J. P. & Göker, M. VICTOR: genome-based phylogeny and classification of prokaryotic viruses. Bioinformatics 33, 3396–3404 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Pawluk, A. et al. Inactivation of CRISPR-Cas systems by anti-CRISPR proteins in diverse bacterial species. Nat. Microbiol. 1, 16085 (2016).

    CAS  PubMed  Google Scholar 

  43. Stamatakis, A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30, 1312–1313 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Hampton, H. G. et al. CRISPR-Cas gene-editing reveals RsmA and RsmC act through FlhDC to repress the SdhE flavinylation factor and control motility and prodigiosin production in Serratia. Microbiology 162, 1047–1058 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Kaniga, K., Delor, I. & Cornelis, G. R. A wide-host-range suicide vector for improving reverse genetics in Gram-negative bacteria: inactivation of the blaA gene of Yersinia enterocolitica. Gene 109, 137–141 (1991).

    CAS  PubMed  Google Scholar 

  46. Watson, B. N. J. et al. Type I-F CRISPR-Cas resistance against virulent phage infection triggers abortive infection and provides population-level immunity. Preprint at https://doi.org/10.1101/679308 (2019).

  47. Silva-Rocha, R. et al. The Standard European Vector Architecture (SEVA): a coherent platform for the analysis and deployment of complex prokaryotic phenotypes. Nucleic Acids Res. 41, D666–D675 (2013).

    CAS  PubMed  Google Scholar 

  48. Biswas, A., Staals, R. H., Morales, S. E., Fineran, P. C. & Brown, C. M. CRISPRDetect: a flexible algorithm to define CRISPR arrays. BMC Genomics 17, 356 (2016).

    PubMed  PubMed Central  Google Scholar 

  49. Nicholson, T. J. et al. Bioinformatic evidence of widespread priming in type I and II CRISPR-Cas systems. RNA Biol. 16, 566–576 (2019).

    PubMed  Google Scholar 

  50. Paez-Espino, D. et al. IMG/VR v.2.0: an integrated data management and analysis system for cultivated and environmental viral genomes. Nucleic Acids Res. 47, D678–D686 (2019).

    CAS  PubMed  Google Scholar 

  51. Rizk, G. & Lavenier, D. GASSST: global alignment short sequence search tool. Bioinformatics 26, 2534–2540 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Johnson, L. S., Eddy, S. R. & Portugaly, E. Hidden Markov model speed heuristic and iterative HMM search procedure. BMC Bioinformatics 11, 431 (2010).

    PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Finn, R. D., Clements, J. & Eddy, S. R. HMMER web server: interactive sequence similarity searching. Nucleic Acids Res. 39, W29–W37 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by the Marsden Fund from the Royal Society of New Zealand and a University of Otago Research Grant. L.M.M. was supported by a University of Otago Doctoral Scholarship. We thank staff of the Otago Micro and Nano Imaging facility for assistance with electron and confocal microscopy and the Otago Genomics Facility for genome sequencing. We thank members of the Fineran laboratory for helpful discussions, S. Shehreen, T. Nicholson and X. Morgan for bioinformatics advice. We also acknowledge the use of the New Zealand eScience Infrastructure (NeSI) high-performance computing facilities in this research. NeSI’s facilities are provided by and funded jointly by NeSI’s collaborator institutions and through the Ministry of Business, Innovation and Employment’s Research Infrastructure programme.

Author information

Authors and Affiliations

Authors

Contributions

L.M.M., S.L.W., C.W. and L.F.G. performed the experiments. L.M.M., S.L.W., S.A.J. and C.W. generated the strains and plasmids. L.M.M., S.L.W. and L.F.G. performed the microscopy. L.M.M. and S.A.J. performed the bioinformatics analysis with input from P.P.G. and P.C.F. L.M.M. and P.C.F. conceived the project with input from all authors. P.C.F. supervised the project. L.M.M. and P.C.F. wrote the manuscript. All authors edited the manuscript.

Corresponding author

Correspondence to Peter C. Fineran.

Ethics declarations

Competing interests

The authors declare no competing interests.

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 The jumbo phage is resistant to type I CRISPR-Cas immunity.

a. tblastx alignment of PCH45 with phages PhiEaH1 and 2050HW. b. Target location of chromosomally expressed anti-PCH45 type I-E (S4-7) and type I-F (S8) spacers targeting the major capsid gene (gp033). Phage resistance measured by c. EOP or d. plate reader assays for Serratia strains with type I-E, (S4, PCF591; S5, PCF593; S6, PCF545; S7, PCF544) and type I-F (S8, PCF548) infected with PCH45. Phage resistance measured by e. EOP or f. plate reader assays for Serratia carrying a type I-E (S4, pPF1460) spacer in mini-CRISPR arrays, infected with PCH45. Phage resistance measured by g. EOP or h. plate reader assays for Serratia carrying a type I-F (S8, pPF1461) spacer in a plasmid mini-CRISPR array, infected with PCH45. In c, e and g MOI=0.001. In c-h data are presented as mean ± s.d. (n=3 biologically independent samples). i. Restriction length fragment polymorphism (RLFP) analysis of phage gDNA treated with restriction enzymes MfeI, EcoRI, DpnI*, KpnI, AluI and NruI. Undigested PCH45 gDNA was used as a negative control. In parenthesis the number of restriction sites found in the genome of PCH45. (*): Cleaves only when the recognition motif is methylated. This experiment was performed three times with similar results and a representative gel is shown.

Extended Data Fig. 2 The shell and tubulin proteins in Serratia jumbo phage PCH45 possess low sequence similarity to homologues encoded by other jumbo phages.

Phylogenetic tree of a. the shell protein and b. PhuZ protein encoded by jumbo phages. The maximum likelihood trees were built for phage encoded shell and tubulin-like proteins (n=9 and n=11, respectively) using RaxML with 100 bootstrap replicates. The scale bar represents the approximate number of changes per amino acid position. c. EOP assay for Serratia strains PCF761 (mCherry2-cas8e), PCF763 (mCherry2-cas8f) and PCF765 (mCherry2-cas10) carrying type I-E, I-F and III-A anti-JS26 spacers in plasmids (pPF1485, pPF1489 and pPF1473 respectively). In c data are presented as mean ± s.d. (n=3 biologically independent samples).

Extended Data Fig. 3 Type III RNA targeting provides protection against jumbo phage infection.

a. EOP assay for type III-A mutant strains: cas10H17A, N18A (HD domain), cas10D618A, D619A (Palm domain), cas7D34A, and the accessory nuclease knock out carrying an anti-PCH45 spacers (RNA polymerase beta subunit, S9; anti-terminase S10; and anti-helicase, S12) overexpressed in trans from a plasmid mini-CRISPR array. b. Conjugation efficiency assay (transconjugants/recipients) of plasmids pPF781 (untargeted control) and pPF1043 targeted by the type III-A CRISPR-Cas systems for Serratia strains. The type III-A mutants: PCF683 (cas7D34A), PCF690 (cas10 HD mutant), PCF691 (cas10 Palm mutant) PCF686 (Δ accessory nuclease), and c. the chromosomal complementation with wild-type copies of the genes in PCF684 (cas7), PCF688 (cas10) and PCF687 (accessory nuclease). All data are presented as mean ± s.d. (n=3 biologically independent samples).

Extended Data Fig. 4 Type III CRISPR arrays are enriched in jumbo phage-targeting spacers.

a. Workflow used to obtain spacer-phage hits. Scores for spacer-target matches for targeted (black) and shuffled (grey) databases for b. type III c. type I-E and d. type I-F CRISPR-Cas systems. Scores with a false positive rate (FPR) < 0.01 were used as a cut-off to determine the spacer-protospacer hits; the FPR is defined as (the number of hits above the scoring threshold to the shuffled target database)/(the number of hits above the scoring threshold to the target database). e. Number of unique spacers in type I-E, I-F or type III systems matching nucleus-forming phages.

Extended Data Fig. 5

Spacer list in native CRISPR arrays.

Extended Data Fig. 6

Spacers expressed from mini-array in plasmid.

Supplementary information

Supplementary Information

Supplementary Tables 3–5.

Reporting Summary

Supplementary Data

Supplementary File 1.

Supplementary Data

Supplementary File 2.

Supplementary Tables

Supplementary Tables 1 and 2.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Malone, L.M., Warring, S.L., Jackson, S.A. et al. A jumbo phage that forms a nucleus-like structure evades CRISPR–Cas DNA targeting but is vulnerable to type III RNA-based immunity. Nat Microbiol 5, 48–55 (2020). https://doi.org/10.1038/s41564-019-0612-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-019-0612-5

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