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:

A compact Cascade–Cas3 system for targeted genome engineering

Abstract

CRISPR–Cas technologies have enabled programmable gene editing in eukaryotes and prokaryotes. However, the leading Cas9 and Cas12a enzymes are limited in their ability to make large deletions. Here, we used the processive nuclease Cas3, together with a minimal Type I-C Cascade-based system for targeted genome engineering in bacteria. DNA cleavage guided by a single CRISPR RNA generated large deletions (7–424 kilobases) in Pseudomonas aeruginosa with near-100% efficiency, while Cas9 yielded small deletions and point mutations. Cas3 generated bidirectional deletions originating from the programmed site, which was exploited to reduce the P. aeruginosa genome by 837 kb (13.5%). Large deletion boundaries were efficiently specified by a homology-directed repair template during editing with Cascade–Cas3, but not Cas9. A transferable ‘all-in-one’ vector was functional in Escherichia coli, Pseudomonas syringae and Klebsiella pneumoniae, and endogenous CRISPR–Cas use was enhanced with an ‘anti-anti-CRISPR’ strategy. P. aeruginosa Type I-C Cascade–Cas3 (PaeCas3c) facilitates rapid strain manipulation with applications in synthetic biology, genome minimization and the removal of large genomic regions.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Type I-C CRISPR-mediated self-targeting leads to genomic deletions.
Fig. 2: Optimization and characterization of Cascade–Cas3-directed genomic editing.
Fig. 3: Iterative generation of multiple genomic deletions in P. aeruginosa.
Fig. 4: Cascade–Cas3-mediated heterologous editing in various bacteria.
Fig. 5: Cascade–Cas3-mediated gene editing in native settings.

Similar content being viewed by others

Data availability

Plasmids pJW31 and pCas3cRh are available through Addgene (numbers 136423 and 133773, respectively). Raw WGS data associated with Figs. 1d, 3b, 4a,c,f and 5a) have been uploaded to GenBank (accession numbers CP047061CP047079) and are also available, along with bacterial strains, upon request from the corresponding author. P. aeruginosa strains available for laboratories with BSL-2 clearance. Source data are provided with this paper.

References

  1. Makarova, K. S. et al. An updated evolutionary classification of CRISPR–Cas systems. Nat. Rev. Microbiol. 13, 722–736 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  3. 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 

  4. Barrangou, R. & Doudna, J. A. Applications of CRISPR technologies in research and beyond. Nat. Biotechnol. 34, 933–941 (2016).

    CAS  PubMed  Google Scholar 

  5. 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 

  6. Hidalgo-Cantabrana, C. & Barrangou, R. Characterization and applications of Type I CRISPR-Cas systems. Biochem. Soc. Trans. 48, 15–23 (2020).

    CAS  PubMed  Google Scholar 

  7. Sinkunas, T. et al. Cas3 is a single‐stranded DNA nuclease and ATP‐dependent helicase in the CRISPR/Cas immune system. EMBO J. 30, 1335–1342 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Sinkunas, T. et al. In vitro reconstitution of Cascade-mediated CRISPR immunity in Streptococcus thermophilus. EMBO J. 32, 385–394 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Mulepati, S. & Bailey, S. In vitro reconstitution of an Escherichia coli RNA-guided immune system reveals unidirectional, ATP-dependent degradation of DNA target. J. Biol. Chem. 288, 22184–22192 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Redding, S. et al. Surveillance and processing of foreign DNA by the Escherichia coli CRISPR-Cas system. Cell 163, 854–865 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Vercoe, R. B. et al. Cytotoxic chromosomal targeting by CRISPR/Cas systems can reshape bacterial genomes and expel or remodel pathogenicity islands. PLoS Genet. 9, e1003454 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Gomaa, A. A. et al. Programmable removal of bacterial strains by use of genome-targeting CRISPR-Cas systems. mBio 5, e00928–00913 (2014).

    PubMed  PubMed Central  Google Scholar 

  13. Kiro, R., Shitrit, D. & Qimron, U. Efficient engineering of a bacteriophage genome using the type I-E CRISPR-Cas system. RNA Biol. 11, 42–44 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Li, Y. et al. Harnessing Type I and Type III CRISPR-Cas systems for genome editing. Nucleic Acids Res. 44, e34–e34 (2016).

    PubMed  Google Scholar 

  15. Pyne, M. E., Bruder, M. R., Moo-Young, M., Chung, D. A. & Chou, C. P. Harnessing heterologous and endogenous CRISPR-Cas machineries for efficient markerless genome editing in Clostridium. Sci. Rep. 6, 25666 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Hidalgo-Cantabrana, C., Goh, Y. J., Pan, M., Sanozky-Dawes, R. & Barrangou, R. Genome editing using the endogenous type I CRISPR-Cas system in Lactobacillus crispatus. Proc. Natl Acad. Sci. USA 116, 15774–15783 (2019).

    CAS  PubMed  Google Scholar 

  17. 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 

  18. Cheng, F. et al. Harnessing the native type I-B CRISPR-Cas for genome editing in a polyploid archaeon. J. Genet. Genomics Yi Chuan Xue Bao 44, 541–548 (2017).

    PubMed  Google Scholar 

  19. Cañez, C., Selle, K., Goh, Y. J. & Barrangou, R. Outcomes and characterization of chromosomal self-targeting by native CRISPR-Cas systems in Streptococcus thermophilus. FEMS Microbiol. Lett. 366, fnz105 (2019).

    PubMed  Google Scholar 

  20. Xu, Z. et al. Native CRISPR-Cas-mediated genome editing enables dissecting and sensitizing clinical multidrug-resistant P. aeruginosa. Cell Rep. 29, 1707–1717.e3 (2019).

    CAS  PubMed  Google Scholar 

  21. Zheng, Y. et al. Characterization and repurposing of the endogenous Type I-F CRISPR-Cas system of Zymomonas mobilis for genome engineering. Nucleic Acids Res. 47, 11461–11475 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Edgar, R. & Qimron, U. The Escherichia coli CRISPR system protects from λ Lysogenization, Lysogens, and prophage induction. J. Bacteriol. 192, 6291–6294 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Dolan, A. E. et al. Introducing a spectrum of long-range genomic deletions in human embryonic stem cells using type I CRISPR-Cas. Mol. Cell 74, 936–950.e5 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Morisaka, H. et al. CRISPR–Cas3 induces broad and unidirectional genome editing in human cells. Nat. Commun. 10, 5302 (2019).

  25. Cameron, P. et al. Harnessing type I CRISPR–Cas systems for genome engineering in human cells. Nat. Biotechnol. 37, 1471–1477 (2019).

    CAS  PubMed  Google Scholar 

  26. Pickar-Oliver, A. et al. Targeted transcriptional modulation with type I CRISPR–Cas systems in human cells. Nat. Biotechnol. 37, 1493–1501 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Chen, Y. et al. Repurposing type I–F CRISPR–Cas system as a transcriptional activation tool in human cells. Nat. Commun. 11, 3136 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Young, J. K. et al. The repurposing of type I-E CRISPR-Cascade for gene activation in plants. Commun. Biol. 2, 383 (2019).

  29. Nam, K. H. et al. Cas5d protein processes Pre-crRNA and assembles into a cascade-like interference complex in subtype I-C/Dvulg CRISPR-Cas system. Structure 20, 1574–1584 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Hochstrasser, M. L., Taylor, D. W., Kornfeld, J. E., Nogales, E. & Doudna, J. A. DNA targeting by a minimal CRISPR RNA-guided cascade. Mol. Cell 63, 840–851 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Marino, N. D. et al. Discovery of widespread type I and type V CRISPR-Cas inhibitors. Science 362, 240–242 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Turner, K. H., Wessel, A. K., Palmer, G. C., Murray, J. L. & Whiteley, M. Essential genome of Pseudomonas aeruginosa in cystic fibrosis sputum. Proc. Natl Acad. Sci. USA 112, 4110–4115 (2015).

    CAS  PubMed  Google Scholar 

  33. Selle, K., Klaenhammer, T. R. & Barrangou, R. CRISPR-based screening of genomic island excision events in bacteria. Proc. Natl Acad. Sci. USA 112, 8076–8081 (2015).

    CAS  PubMed  Google Scholar 

  34. Chayot, R., Montagne, B., Mazel, D. & Ricchetti, M. An end-joining repair mechanism in Escherichia coli. Proc. Natl Acad. Sci. USA 107, 2141–2146 (2010).

    CAS  PubMed  Google Scholar 

  35. Lindeberg, M., Cunnac, S. & Collmer, A. Pseudomonas syringae type III effector repertoires: last words in endless arguments. Trends Microbiol. 20, 199–208 (2012).

    CAS  PubMed  Google Scholar 

  36. Kvitko, B. H. et al. Deletions in the repertoire of Pseudomonas syringae pv. tomato DC3000 type III secretion effector genes reveal functional overlap among effectors. PLoS Pathog. 5, e1000388 (2009).

    PubMed  PubMed Central  Google Scholar 

  37. Caliando, B. J. & Voigt, C. A. Targeted DNA degradation using a CRISPR device stably carried in the host genome. Nat. Commun. 6, 6989 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Bachman, M. A. et al. Genome-wide identification of Klebsiella pneumoniae fitness genes during lung infection. mBio 6, e00775 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Cady, K. C., Bondy-Denomy, J., Heussler, G. E., Davidson, A. R. & O’Toole, G. A. The CRISPR/Cas adaptive immune system of Pseudomonas aeruginosa mediates resistance to naturally occurring and engineered phages. J. Bacteriol. 194, 5728–5738 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. 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 

  41. Rauch, B. J. et al. Inhibition of CRISPR-Cas9 with bacteriophage proteins. Cell 168, 150–158.e10 (2017).

    CAS  PubMed  Google Scholar 

  42. Stanley, S. Y. et al. Anti-CRISPR-associated proteins are crucial repressors of Anti-CRISPR transcription. Cell 178, 1452–1464.e13 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Ha, A. D. & Denver, D. R. Comparative genomic analysis of 130 bacteriophages infecting bacteria in the genus Pseudomonas. Front. Microbiol. 9, 1456 (2018).

    PubMed  PubMed Central  Google Scholar 

  44. Pósfai, G. et al. Emergent properties of reduced-genome Escherichia coli. Science 312, 1044–1046 (2006).

    PubMed  Google Scholar 

  45. Csörgő, B., Nyerges, Á., Pósfai, G. & Fehér, T. System-level genome editing in microbes. Curr. Opin. Microbiol. 33, 113–122 (2016).

    PubMed  Google Scholar 

  46. Képès, F. et al. The layout of a bacterial genome. FEBS Lett. 586, 2043–2048 (2012).

    PubMed  Google Scholar 

  47. Cui, L. & Bikard, D. Consequences of Cas9 cleavage in the chromosome of Escherichia coli.Nucleic Acids Res. 44, 4243–4251 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Bowater, R. & Doherty, A. J. Making ends meet: repairing breaks in bacterial DNA by non-homologous end-joining. PLoS Genet. 2, e8 (2006).

  49. Tuladhar, R. et al. CRISPR–Cas9-based mutagenesis frequently provokes on-target mRNA misregulation. Nat. Commun. 10, 4056 (2019).

  50. Smits, A. H. et al. Biological plasticity rescues target activity in CRISPR knock outs. Nat. Methods 16, 1087–1093 (2019).

    CAS  PubMed  Google Scholar 

  51. Choi, K.-H. et al. A Tn7-based broad-range bacterial cloning and expression system. Nat. Methods 2, 443–448 (2005).

    CAS  PubMed  Google Scholar 

  52. Stover, C. K. et al. Complete genome sequence of Pseudomonas aeruginosa PAO1, an opportunistic pathogen. Nature 406, 959–964 (2000).

    CAS  PubMed  Google Scholar 

  53. Choi, K.-H. & Schweizer, H. P. mini-Tn7 insertion in bacteria with single attTn7 sites: example Pseudomonas aeruginosa. Nat. Protoc. 1, 153–161 (2006).

    CAS  PubMed  Google Scholar 

  54. Buell, C. R. et al. The complete genome sequence of the Arabidopsis and tomato pathogen Pseudomonas syringae pv. tomato DC3000. Proc. Natl Acad. Sci. USA 100, 10181–10186 (2003).

    CAS  PubMed  Google Scholar 

  55. Blattner, F. R. et al. The complete genome sequence of Escherichia coli K-12. Science 277, 1453–1462 (1997).

    CAS  PubMed  Google Scholar 

  56. Broberg, C. A., Wu, W., Cavalcoli, J. D., Miller, V. L. & Bachman, M. A. Complete genome sequence of Klebsiella pneumoniae strain ATCC 43816 KPPR1, a Rifampin-resistant mutant commonly used in animal, genetic, and molecular biology studies. Genome Announc. 2, e00924–14 (2014).

    PubMed  PubMed Central  Google Scholar 

  57. Qiu, D., Damron, F. H., Mima, T., Schweizer, H. P. & Yu, H. D. PBAD-based shuttle vectors for functional analysis of toxic and highly regulated genes in Pseudomonas and Burkholderia spp. and other bacteria. Appl. Environ. Microbiol. 74, 7422–7426 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Gibson, D. G. et al. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat. Methods 6, 343–345 (2009).

    CAS  PubMed  Google Scholar 

  59. Meisner, J. & Goldberg, J. B. The Escherichia coli rhaSR-PrhaBAD inducible promoter system allows tightly controlled gene expression over a wide range in Pseudomonas aeruginosa. Appl. Environ. Microbiol. 82, 6715–6727 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Borges, A. L. et al. Bacteriophage cooperation suppresses CRISPR-Cas3 and Cas9 immunity. Cell 174, 917–925.e10 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Nyerges, Á. et al. Directed evolution of multiple genomic loci allows the prediction of antibiotic resistance. Proc. Natl Acad. Sci. USA 115, E5726–E5735 (2018).

    CAS  PubMed  Google Scholar 

  62. Huynh, T. V., Dahlbeck, D. & Staskawicz, B. J. Bacterial blight of soybean: regulation of a pathogen gene determining host cultivar specificity. Science 245, 1374–1377 (1989).

    CAS  PubMed  Google Scholar 

  63. Kropinski, A. M. Sequence of the genome of the temperate, serotype-converting, Pseudomonas aeruginosa bacteriophage D3. J. Bacteriol. 182, 6066–6074 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Budzik, J. M., Rosche, W. A., Rietsch, A. & O’Toole, G. A. Isolation and characterization of a generalized transducing phage for Pseudomonas aeruginosa strains PAO1 and PA14. J. Bacteriol. 186, 3270–3273 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Alikhan, N.-F., Petty, N. K., Ben Zakour, N. L. & Beatson, S. A. BLAST Ring Image Generator (BRIG): simple prokaryote genome comparisons. BMC Genomics 12, 402 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

B.C. is supported by the Eötvös National Scholarship of Hungary and a Marie Skłodowska-Curie Actions Individual Global Fellowship (‘GenDels’, no. 844093) of the Horizon 2020 Research Program of the European Commission. L.M.L. is supported by the HHMI Gilliam Fellowship for Advanced Study and the UCSF Discovery Fellowship. Research on plant immunity in the Lewis laboratory is supported by the USDA grant nos. ARS 2030-21000-046-00D and 2030-21000-050-00D (J.D.L.), and the NSF Directorate for Biological Sciences grant no.IOS-1557661 (J.D.L.). I.J.C.-L. is supported by a Grace Kase fellowship from UC Berkeley and the NSF Graduate Research Fellowship Program. A.V.R. is supported by funding from the Pew Charitable Trusts. E.D.C. is funded by the Chan Zuckerberg Biohub. CRISPR–Cas3 projects in the Bondy-Denomy Laboratory are supported by the UCSF Program for Breakthrough Biomedical Research funded in part by the Sandler Foundation, the Innovative Genomics Institute and an NIH Director’s Early Independence Award DP5-OD021344. We thank J.B. Goldberg (Emory University) for providing the plasmid pJM230, and A. Borges (UCSF) for providing pAB01 to clone Type I-F crRNAs. We thank the Bondy-Denomy laboratory for productive conversations pertaining to this project.

Author information

Authors and Affiliations

Authors

Contributions

B.C. and L.M.L. participated in designing and performing experiments, analyzing data, acquiring funding for the project and writing the manuscript. I.J.C.-L. performed in vitro and in planta P. syringae experiments. A.V.-R. performed and analyzed Type I-F CRISPR–Cas editing experiments. J.D.B. assisted in designing and constructing the all-in-one pCas3cRh vector. C.M. constructed the PAO1IIA strain and Cas9 gRNA expression vector. E.D.C. performed WGS. J.D.L. designed experiments with P. syringae. J.B.-D. conceived the study, designed experiments, analyzed data, acquired funding for the project and wrote the manuscript.

Corresponding author

Correspondence to Joseph Bondy-Denomy.

Ethics declarations

Competing interests

J.B.-D. is a scientific advisory board member of SNIPR Biome and Excision Biotherapeutics and a scientific advisory board member and cofounder of Acrigen Biosciences. J.B.-D., L.M.L. and B.C. have filed a patent application relating to various aspects of Cas3-based genome editing.

Additional information

Peer review information Lei Tang was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Type I-C CRISPR targeting leads to genomic deletions.

a, Comparison of Type I-C CRISPR system from P. aeruginosa used in the study, to various other previously identified I-C systems from a range of different bacteria. Values show query coverage and percent identity (ID) percentages comparing the four genes of the P. aeruginosa system to each of the other four. * Denotes the reference Type I-C CRISPR system referred to in Ref. 1. b, PCR amplification of a 3 kb genomic fragment flanking the phzM gene targeted using two different crRNAs, phzM_1 and phzM_2. Colony PCRs were performed on 18 biological replicates of self-targeted strains for each crRNA. The PAO1ICparental strain is used as a positive control (wt). L indicates a 1 kb DNA ladder.

Source data

Extended Data Fig. 2 Excision of plasmid-encoded spacer sequences.

a, Phage targeting assays with survivors that had no discernable deletion of the crRNA-targeted genomic site. Strains were transformed with a D3 phage-targeting crRNA to assay for IC CRISPR-Cas3 activity. Three unique survivors were isolated from six self-targeting assays for a total of 18 survivors. Control is a non-targeting crRNA. b, Schematic of spacer excision events where the two direct repeats recombine, resulting the loss of the targeting spacer. c, PCR amplification of the crRNA sequence from plasmids isolated from 17 non-deletion self-targeted survivors (selected from 3 biological replicates of 12 analyzed colonies (see Fig. 2a). Pl indicates the original plasmid as the PCR template, Ni indicates a sample where the crRNA was not induced, L indicates a 1 kb DNA ladder. d, Sample chromatogram of a sequenced plasmid with the spacer flipped out. Only one 32 bp repeat sequence remains in the plasmid, the 34 bp spacer sequence and other 32 bp repeat are missing.

Source data

Extended Data Fig. 3 Phage-targeting assays to confirm CRISPR-Cas functionality.

a, Phage-targeting assay showing the activity of the modified repeat crRNA constructs. Ten-fold serial dilutions of DMS3 phage and D3 phage were spotted on lawns of PAO1IC expressing either empty vector (top), a crRNA targeting D3 with WT direct repeats (middle), or a crRNA targeting D3 with modified repeats (bottom). b, Phage targeting assay of five non-deletion self-targeting survivors expressing a D3 phage targeting crRNA. Unsuccessful targeting of phage indicates a non-functional CRISPR-Cas system in these strains. The parental PAO1IC strain with a functional CRISPR-Cas system was used as a control.

Source data

Extended Data Fig. 4 Genomic targeting of essential gene rplQ.

a, Growth curves of 36 PAO1IC biological replicates targeting the essential gene, rplQ, using the MR crRNA plasmid. b, Phage targeting assays with eight isolated rplQ-targeted survivors to assay for I-C CRISPR-Cas activity. Serial dilutions of DMS3 phage and D3 phage were spotted on lawns of PAO1IC expressing a crRNA targeting phage D3. The parent PAO1IC strain expressing a D3 targeting crRNA (top left) was used as a positive control, while PAO1IC expressing a non-targeting crRNA was used as a negative control.

Source data

Extended Data Fig. 5 Genomic targeting using a Type II-A CRISPR-Cas system.

Growth of self-targeting strains of PAO1IIA expressing a self-targeting gRNA targeting the genome at phzM (Ind.). An empty vector (E.V.) and a non-induced phzM targeting strain (N.I.) were used as controls. Mean OD values measured at 600 nm are shown for 8 biological replicates each, error bars indicate SD values.

Extended Data Fig. 6 Genomic deletions and junction sites.

a, Deletion efficiencies observed over six cycles of iterative self-targeting. Six genomic targets were targeted in six different orders. Six survivors were analyzed using site-specific PCR after each cycle, for a total of 36 analyzed colonies (6*6) after each cycle, error bars represent standard deviations. b, Deletion junctions at XNES6 target site of the 6 PAO1IC strains with 6 iterative targeting events each. Sequences of each specific microhomology for the junctions are shown for each strain above the bars representing the given genomes at both ends, deletion sizes are shown below dashed lines for each strain. c, PCR analysis using a representative set of primers amplifying various large deletion junctions (at XNES1, 6, 8, and 9 regions) of the whole-genome sequenced Δ62 strain. Δ62 served as a positive control template, while wtC represents untargeted PAO1IC cells scraped from a lawn of colonies from a single overnight culture grown on plates serving as templates, and wtG represents isolated genomic DNA from a different 1.5 ml overnight culture of untargeted PAO1IC used as templates. Bands appearing for the XNES9 deletion junction for the PAO1IC samples were aspecific and when sequenced, did not match any genomic region of the PAO1IC genome. L indicates a 1 kb DNA ladder.

Source data

Extended Data Fig. 7 Genomic targeting of P. aeruginosa PAO1 with all-in-one vector pCas3ch.

a, Map of the I-C CRISPR-Cas all-in-one plasmid pCas3cRh carrying I-C crRNA and genes cas3, cas5, cas8, and cas7 under the control of the rhamnose-inducible rhaSR-PrhaBAD system. b, Growth curve of PAO1 transformed with the pCas3cRh vector expressing a self-targeting crRNA targeting phzM (Ind.). An empty vector (E.V.) and a non-induced phzM targeting strain (N.I.) were used as controls. Mean OD values measured at 600 nm are shown for six biological replicates each. c, Deletion efficiencies for WT PAO1 using the all-in-one vector pCas3cRh carrying all necessary components of the I-C CRISPR-Cas system. Values are averages of three replicates where 12 individual colonies were analyzed using site-specific PCR. Error bars show standard deviations. d, Transformation efficiencies with self-targeting pCas3cRh vectors expressing crRNAs for phzM or XNES 2 compared to a non-targeting control (green bar) in PAO1. Values are means of 3 replicates each, error bars represent SD values.

Extended Data Fig. 8 Genomic targeting of Pseudomonas syringae and growth phenotypes of deletion strains.

a, Growth of P. syringae DC3000 strains expressing the I-C system and distinct crRNAs. Constructs VI, IV-IX, and VIII target P. syringae DC3000 non-essential chromosomal genes, non-targeting crRNA (NT), empty vector (EV). b, Percentage of survivors with targeted deletions in clusters of non-essential virulence effector genes in P. syringae pv. tomato DC3000. Values are averages of three biological replicates where 12 individual colonies were analyzed using site-specific PCR for each, error bars show standard deviations. c, In vitro growth of cluster VI deletion strains in King’s medium B (KB). ΔCEL is the previously published polymutant, while ΔCVI-1 and ΔCVI-2 are Cas3-generated mutants. Values shown are the means of 4 biological replicates each, error bars represent standard deviations. d, In vitro growth of cluster IV, cluster IX deletion strains in KB. ΔCEL is the previously published polymutant, while ΔCIVΔCIX-1 and ΔCIVΔCIX-2 are Cas3-generated mutants. Values shown are the means of 4 biological replicates each, error bars represent standard deviations. e, In vitro growth of cluster X deletion strains in KB. ΔCEL is the previously published polymutant, while ΔCX-1 and ΔCX-2 are Cas3-generated mutants. Values shown are the means of 4 biological replicates each, error bars represent standard deviations. f, In vitro growth of cluster VI deletion strains in apoplast mimicking minimal media (MM). ΔCEL is the previously published polymutant, while ΔCVI-1 and ΔCVI-2 are Cas3-generated mutants. Values shown are the means of 4 biological replicates each, error bars represent standard deviations. g, In vitro growth of cluster IV, cluster IX deletion strains in MM. ΔCEL is the previously published polymutant, while ΔCIVΔCIX-1 and ΔCIVΔCIX-2 are Cas3-generated mutants. Values shown are the means of 4 biological replicates each, error bars represent standard deviations. h, In vitro growth of cluster X deletion strains in MM. ΔCEL is the previously published polymutant, while ΔCX-1 and ΔCX-2 are Cas3-generated mutants. Values shown are the means of 4 biological replicates each, error bars represent standard deviations.

Source data

Extended Data Fig. 9 CRISPR-Cas3 editing in Klebsiella pneumoniae.

a, Growth curves of K. pneumoniae strains expressing distinct crRNAs targeting rfaH and sacX (2 each). Non-targeting crRNA expressing control is marked in blue. Values depicted are averages of 8 biological replicates each. b, Representative gel electrophoresis of PCR fragments amplified from 8 total surviving colonies each from the 4 crRNA targeting constructs (representing 1 biological replicate of 3 total). Primer pairs amplified regions flanking the targeted position at rfaH and sacX. Wild-type KPPR1 (wt) colonies were used as controls, L represents 1 kb DNA marker ladder. c, Percentage of survivors with targeted deletions at the targeted genomic positions. Values are averages of three biological replicates where 8 individual colonies were analyzed using site-specific PCR for each, error bars show standard deviations. d, Colony morphologies of deletion candidate strains of rfaH and sacX compared to wild-type K. pneumoniae KPPR1.

Source data

Extended Data Fig. 10 Genomic editing in native host of Type I-C CRISPR-Cas system and effect of I-C specific anti-CRISPR protein on the process.

a, Editing efficiencies for the Pseudomonas aeruginosa environmental isolate naturally expressing the Type I-C cas genes, transformed with a plasmid targeting phzM with WT repeats or modified repeats. Each data point represents the fraction of isolates with the deletion out of ten isolates assayed. b, Genotyping results for the Pseudomonas aeruginosa environmental isolate using the 0.17 kb HDR template. 10 biological replicates were assayed. Larger band corresponds to the WT sequence, smaller band corresponds to a genome reduced by 0.17 kb. c, Genotyping results of PAO1IC AcrC1 lysogens after self-targeting induction in the presence or absence of aca1 and a non-targeted control. Ten biological replicates per strain were assayed. gDNA was extracted from each replicate and PCR analysis for the phzM gene (targeted gene, top row of gels) or cas5 gene (non-targeted gene, bottom row) was conducted. Only cells that co-expressed aca1 with the crRNA showed loss of the phzM band, indicating genome editing. All replicates had a cas5 band, indicating successful gDNA extraction and target specificity for the phzM locus.

Source data

Supplementary information

Supplementary Information

Supplementary Fig. 1.

Reporting Summary

Supplementary Tables 1–4

Tables listing XNES regions of P. aeruginosa PAO1 genome, genomic coordinates of large deletions generated in the study, summary of HR-mediated deletion experiments using I-F CRISPR–Cas system, and table of all oligonucleotides used in the study.

Source data

Source Data Extended Data Fig. 1

Unprocessed gel electrophoresis picture of Extended Data Fig. 1b

Source Data Extended Data Fig. 2

Unprocessed pictures of phage plaquing assays for Extended Data Fig. 2a, unprocessed gel electrophoresis picture of Extended Data Fig. 2c

Source Data Extended Data Fig. 3

Unprocessed pictures of phage plaquing assays for Extended Data Figs. 3a,b

Source Data Extended Data Fig. 4

Unprocessed pictures of phage plaquing assays for Extended Data Fig. 4b

Source Data Extended Data Fig. 6

Unprocessed gel electrophoresis picture of Extended Data Fig. 6c

Source Data Extended Data Fig. 8

Unprocessed pictures of P. syringae cultures for Extended Data Fig. 8a

Source Data Extended Data Fig. 9

Unprocessed gel electrophoresis picture of Extended Data Fig. 9b and unprocessed gel electrophoresis picture of Extended Data Fig. 9b

Source Data Extended Data Fig. 10

Unprocessed gel electrophoresis pictures of Extended Data Fig. 10

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Csörgő, B., León, L.M., Chau-Ly, I.J. et al. A compact Cascade–Cas3 system for targeted genome engineering. Nat Methods 17, 1183–1190 (2020). https://doi.org/10.1038/s41592-020-00980-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41592-020-00980-w

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

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