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.

In vivo genome editing via CRISPR/Cas9 mediated homology-independent targeted integration

Abstract

Targeted genome editing via engineered nucleases is an exciting area of biomedical research and holds potential for clinical applications. Despite rapid advances in the field, in vivo targeted transgene integration is still infeasible because current tools are inefficient1, especially for non-dividing cells, which compose most adult tissues. This poses a barrier for uncovering fundamental biological principles and developing treatments for a broad range of genetic disorders2. Based on clustered regularly interspaced short palindromic repeat/Cas9 (CRISPR/Cas9)3,4 technology, here we devise a homology-independent targeted integration (HITI) strategy, which allows for robust DNA knock-in in both dividing and non-dividing cells in vitro and, more importantly, in vivo (for example, in neurons of postnatal mammals). As a proof of concept of its therapeutic potential, we demonstrate the efficacy of HITI in improving visual function using a rat model of the retinal degeneration condition retinitis pigmentosa. The HITI method presented here establishes new avenues for basic research and targeted gene therapies.

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

Figure 1: HITI-mediated in vitro genome editing.
Figure 2: HITI-mediated in vivo genome editing in neurons.
Figure 3: HITI-mediated gene correction of a rat model of retinitis pigmentosa.
Figure 4: AAV-mediated systemic HITI in vivo.

References

  1. Naldini, L. Gene therapy returns to centre stage. Nature 526, 351–360 (2015)

    Article  ADS  CAS  PubMed  Google Scholar 

  2. Cox, D. B. T., Platt, R. J. & Zhang, F. Therapeutic genome editing: prospects and challenges. Nat. Med. 21, 121–131 (2015)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Mali, P. et al. RNA-guided human genome engineering via Cas9. Science 339, 823–826 (2013)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  4. Cong, L. et al. Multiplex genome engineering using CRISPR/Cas systems. Science 339, 819–823 (2013)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  5. Lombardo, A. et al. Site-specific integration and tailoring of cassette design for sustainable gene transfer. Nat. Methods 8, 861–869 (2011)

    Article  CAS  PubMed  Google Scholar 

  6. Genovese, P. et al. Targeted genome editing in human repopulating haematopoietic stem cells. Nature 510, 235–240 (2014)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  7. Luchetti, A., Malgieri, A. & Sangiuolo, F. Small fragment homologous replacement (SFHR): sequence-specific modification of genomic DNA in eukaryotic cells by small DNA fragments. Methods Mol. Biol . 1114, 85–101 (2014)

    Article  CAS  PubMed  Google Scholar 

  8. Orthwein, A. et al. A mechanism for the suppression of homologous recombination in G1 cells. Nature 528, 422–426 (2015)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  9. Lieber, M. R. The mechanism of double-strand DNA break repair by the nonhomologous DNA end-joining pathway. Annu. Rev. Biochem. 79, 181–211 (2010)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Bétermier, M., Bertrand, P. & Lopez, B. S. Is non-homologous end-joining really an inherently error-prone process? PLoS Genet. 10, e1004086 (2014)

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  11. Maresca, M., Lin, V. G., Guo, N. & Yang, Y. Obligate ligation-gated recombination (ObLiGaRe): custom-designed nuclease-mediated targeted integration through nonhomologous end joining. Genome Res. 23, 539–546 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Auer, T. O., Duroure, K., De Cian, A., Concordet, J.-P. & Del Bene, F. Highly efficient CRISPR/Cas9-mediated knock-in in zebrafish by homology-independent DNA repair. Genome Res. 24, 142–153 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Nakade, S. et al. Microhomology-mediated end-joining-dependent integration of donor DNA in cells and animals using TALENs and CRISPR/Cas9. Nat. Commun. 5, 5560 (2014)

    Article  ADS  CAS  PubMed  Google Scholar 

  14. Chen, Z.-Y. et al. Linear DNAs concatemerize in vivo and result in sustained transgene expression in mouse liver. Mol. Ther. 3, 403–410 (2001)

    Article  CAS  PubMed  Google Scholar 

  15. Chen, Z.-Y., He, C.-Y., Ehrhardt, A. & Kay, M. A. Minicircle DNA vectors devoid of bacterial DNA result in persistent and high-level transgene expression in vivo. Mol. Ther. 8, 495–500 (2003)

    Article  CAS  PubMed  Google Scholar 

  16. Wu, J., Corbett, A. H. & Berland, K. M. The intracellular mobility of nuclear import receptors and NLS cargoes. Biophys. J. 96, 3840–3849 (2009)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  17. Kameda, Y., Kameya, T. & Frankfurter, A. Immunohistochemical localization of a neuron-specific β-tubulin isotype in the developing chicken ultimobranchial glands. Brain Res. 628, 121–127 (1993)

    Article  CAS  PubMed  Google Scholar 

  18. Tabata, H. & Nakajima, K. Efficient in utero gene transfer system to the developing mouse brain using electroporation: visualization of neuronal migration in the developing cortex. Neuroscience 103, 865–872 (2001)

    Article  CAS  PubMed  Google Scholar 

  19. Mizutani, K. & Saito, T. Progenitors resume generating neurons after temporary inhibition of neurogenesis by Notch activation in the mammalian cerebral cortex. Development 132, 1295–1304 (2005)

    Article  CAS  PubMed  Google Scholar 

  20. Madisen, L. et al. A robust and high-throughput Cre reporting and characterization system for the whole mouse brain. Nat. Neurosci. 13, 133–140 (2010)

    Article  CAS  PubMed  Google Scholar 

  21. Samulski, R. J. & Muzyczka, N. AAV-mediated gene therapy for research and therapeutic purposes. Annu. Rev. Virol . 1, 427–451 (2014)

    Article  PubMed  CAS  Google Scholar 

  22. Bainbridge, J. W. B. et al. Effect of gene therapy on visual function in Leber’s congenital amaurosis. N. Engl. J. Med. 358, 2231–2239 (2008)

    Article  CAS  PubMed  Google Scholar 

  23. Maguire, A. M. et al. Safety and efficacy of gene transfer for Leber’s congenital amaurosis. N. Engl. J. Med. 358, 2240–2248 (2008)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Nathwani, A. C. et al. Long-term safety and efficacy of factor IX gene therapy in hemophilia B. N. Engl. J. Med . 371, 1994–2004 (2014)

    Google Scholar 

  25. Zincarelli, C., Soltys, S., Rengo, G. & Rabinowitz, J. E. Analysis of AAV serotypes 1–9 mediated gene expression and tropism in mice after systemic injection. Mol. Ther. 16, 1073–1080 (2008)

    Article  CAS  PubMed  Google Scholar 

  26. Luo, J. et al. Human retinal progenitor cell transplantation preserves vision. J. Biol. Chem. 289, 6362–6371 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. D’Cruz, P. M. et al. Mutation of the receptor tyrosine kinase gene Mertk in the retinal dystrophic RCS rat. Hum. Mol. Genet . 9, 645–651 (2000)

    Article  PubMed  Google Scholar 

  28. Boyden, E. S., Zhang, F., Bamberg, E., Nagel, G. & Deisseroth, K. Millisecond-timescale, genetically targeted optical control of neural activity. Nat. Neurosci. 8, 1263–1268 (2005)

    Article  CAS  PubMed  Google Scholar 

  29. Izpisua Belmonte, J. C. et al. Brains, genes, and primates. Neuron 86, 617–631 (2015)

    Article  PubMed  CAS  Google Scholar 

  30. Wu, J. et al. An alternative pluripotent state confers interspecies chimaeric competency. Nature 521, 316–321 (2015)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  31. Shitamukai, A., Konno, D. & Matsuzaki, F. Oblique radial glial divisions in the developing mouse neocortex induce self-renewing progenitors outside the germinal zone that resemble primate outer subventricular zone progenitors. J. Neurosci. 31, 3683–3695 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Liu, G.-H. G. et al. Targeted gene correction of laminopathy-associated LMNA mutations in patient-specific iPSCs. Cell Stem Cell 8, 688–694 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Liu, G.-H. G. et al. Recapitulation of premature ageing with iPSCs from Hutchinson-Gilford progeria syndrome. Nature 472, 221–225 (2011)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  34. Kay, M. A., He, C.-Y. & Chen, Z.-Y. A robust system for production of minicircle DNA vectors. Nat. Biotechnol. 28, 1287–1289 (2010)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Sanjana, N. E. et al. A transcription activator-like effector toolbox for genome engineering. Nat. Protocols 7, 171–192 (2012)

    Article  CAS  PubMed  Google Scholar 

  36. Kutner, R. H., Zhang, X.-Y. & Reiser, J. Production, concentration and titration of pseudotyped HIV-1-based lentiviral vectors. Nat. Protocols 4, 495–505 (2009)

    Article  CAS  PubMed  Google Scholar 

  37. Liu, G.-H. G. et al. Modelling Fanconi anemia pathogenesis and therapeutics using integration-free patient-derived iPSCs. Nat. Commun. 5, 4330 (2014)

    Article  ADS  CAS  PubMed  Google Scholar 

  38. Takahashi, M., Nomura, T. & Osumi, N. Transferring genes into cultured mammalian embryos by electroporation. Dev. Growth Differ . 50, 485–497 (2008)

    Article  CAS  PubMed  Google Scholar 

  39. Mukai, H., Kawakami, S. & Hashida, M. Renal press-mediated transfection method for plasmid DNA and siRNA to the kidney. Biochem. Biophys. Res. Commun. 372, 383–387 (2008)

    Article  CAS  PubMed  Google Scholar 

  40. Lampe, S. E. G., Kaspar, B. K. & Foust, K. D. Intravenous injections in neonatal mice. J. Vis. Exp. e52037 (2014)

  41. Pasut, A., Oleynik, P. & Rudnicki, M. A. Isolation of muscle stem cells by fluorescence activated cell sorting cytometry. Methods Mol. Biol . 798, 53–64 (2012)

    Article  CAS  PubMed  Google Scholar 

  42. Pinto, A. R., Chandran, A., Rosenthal, N. A. & Godwin, J. W. Isolation and analysis of single cells from the mouse heart. J. Immunol. Methods 393, 74–80 (2013)

    Article  CAS  PubMed  Google Scholar 

  43. Hsu, P. D. et al. DNA targeting specificity of RNA-guided Cas9 nucleases. Nat. Biotechnol. 31, 827–832 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics 25, 1754–1760 (2009)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Sauvé, Y., Lu, B. & Lund, R. D. The relationship between full field electroretinogram and perimetry-like visual thresholds in RCS rats during photoreceptor degeneration and rescue by cell transplants. Vision Res. 44, 9–18 (2004)

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We are grateful to M. Kay, Z. Y. Chen, G. Lemke and P. G. Burrola for sharing experimental materials; J. Naughton, L. Lisowski and J. Marlett for AAV production; C. Fine, J. Olvera, E. O’Connor and K. E. Marquez for cell sorting; D. Okamura and M. Jacobs for mouse surgery and histology processing; D. Skowronska-Krawczyk for rat experiments; N. V. Gohad, T. Whitfield, I. M. Verma, J. Ogawa, T. Hara, U. Manor and J. Santini for imaging; L. Greg, Y. S. Kida and F. Osakada for valuable discussions; D. O’Keefe for proofreading the manuscript and M. Schwarz for administrative help. Core Facilities were utilized at the Salk Institute (support from: NIH-NCI CCSG: P30 014195, NINDS R24NS092943, and NEI P30 EY019005) and UCSD Neuroscience core grant P30 NS047101. R.H.B. was supported by a CONACYT fellowship of Mexico. J.Z. and T.J. were supported by 973 Program (2013CB967504, 2015CB964600) and 863 Program (2014AA021604). T.H. was partially supported by a Nomis Foundation Fellowship. E.J.K. is a Biogen-IDEC Fellow of the Life Science Research Foundation. M.Y. was partially supported by the Salk Women & Science Special Award. X.F. was supported by NSFC (No. 81601872). G.H.L. and J.Q. were supported by the National Basic Research Program of China (973 Program; 2015CB964800, 2014CB910503, 2013CB967504), National Natural Science Foundation of China (81625009, 81371342, 81271266), the National High Technology Research and Development Program of China (2015AA020307, 2014AA021604), and Program of Beijing Municipal Science and Technology Commission (Z151100003915072). F.M. was supported by RIKEN funding for Development and Regeneration. Ku.Z. was supported by NIH grant R01HL123755. P.J.M. and J.C.I.B. were supported by the King Abdullah University of Science and Technology (KAUST) Office of Sponsored Research (OSR) under award no. OSR-2015-CRG4-2631. Work in the laboratory of J.C.I.B. was supported by The Leona M. and Harry B. Helmsley Charitable Trust (2012-PG-MED002), the G. Harold and Leila Y. Mathers Charitable Foundation, NIH (R01HL123755), The McKnight Foundation, The Moxie Foundation, Fundacion Dr. Pedro Guillen and Universidad Católica San Antonio de Murcia (UCAM).

Author information

Authors and Affiliations

Authors

Contributions

K.S., Y.T., R.H.B., J.W. and J.C.I.B. conceived the project and designed experiments. K.S., Y.T. and E.A. constructed plasmids. K.S., R.H.B., M.Y. and M.L. generated minicircle DNA vectors. K.S., Y.T., R.H.B., M.L., E.A., A.G. and R.D.S. performed work on HEK293 cells. K.S., Y.T., R.H.B. and E.A. performed bisulfite sequencing. K.S., Y.T. and R.H.B. measured intracellular localization of dCas9. K.S. and Y.T. performed the Surveyor assay. R.H.B. performed work on primary neurons. K.S., Y.T., R.H.B., E.A. and A.G. performed work on human ES-cell-derived pan neurons. Y.T. and F.M. performed work on in utero electroporation. J.Z., T.J., X.F., M.J. and Ka.Z. performed work on RCS rats. E.J.K. and E.M.C. performed work on adult mouse brain. F.H., T.A., M.K. and T.H. performed in vivo mouse electroporation. F.H., M.Y. and T.A. performed AAV IV and IM injection in neonatal or adult mice. Z.L., S.G., S.C. and Ku.Z. performed deep sequencing and analysed data. K.S. and E.A. performed single-cell genotyping. J.W., J.Q., C.R.E, W.T.B., J.L., E.N.D., P.G., J.M.C., G.H.L., P.M. and J.C.I.B. supervised the project or related experiments. K.S., Y.T., J.W. and J.C.I.B. wrote the manuscript with input from all the authors.

Corresponding author

Correspondence to Juan Carlos Izpisua Belmonte.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Additional information

Reviewer Information Nature thanks M. Porteus and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Extended data figures and tables

Extended Data Figure 1 Optimization of donor vectors for HITI in HEK293 cells.

a, Schematic of NHEJ-mediated targeted genome editing and different HITI donor vectors with Streptococcus pyogenes Cas9. Blue pentagon, Cas9/gRNA target sequence. Black line within blue pentagon, Cas9 cleavage site. GOI, gene of interest; DSBs, double strand breaks; Indels, insertions and deletions. b, Surveyor nuclease assay performed with Cas9, Scramble-gRNA, and different donor plasmids (IRESmCherry-MC or IRESmCherry-MC-scramble) in the GFP-correction HEK293 line. The two lower bands are cleaved DNA products by Surveyor nuclease, indicating that Cas9/Scramble-gRNA cut scramble target sequence on the donor, but not genomic sequence. c, Representative immunofluorescence images of the targeted gene integration in the GFP-correction HEK293 line by HDR and HITI. Scale bar, 100 μm. d, Time course studies of the percentages of mCherry+ cells with different HITI targeting vectors. e, The percentage of mCherry+ cells on day 79 versus day 7. f, The CpG methylation status of the mCherry gene at early (day 10) and late passages (day 79) with different HITI targeting vectors. Two black half-arrows indicate primers for bisulphite sequencing. g, The effect of the NHEJ inhibitor (NU7026; 30 μM) on knock-in efficiencies by HDR and HITI. Results were obtained from three replicate wells and presented as mean ± s.d. The input data points were shown as black dots. N.S., not significant. **P < 0.01, unpaired Student’s t-test. For gel source image, see Supplementary Fig. 1. For source data, see Supplementary Table 5.

Source data

Extended Data Figure 2 Sequencing analyses of the IRESmCherry knock-in clones via HITI in HEK293 cells.

a, Analysis of the direction of targeted insertion for HITI with IRESmCherry-MC donor. Detection of reverse integrated IRESmCherry-MC from mCherry single-cell colonies via PCR. Only one (no. 30, highlighted in red) out of 48 mCherry clones (non-transfected, non-edited or reverse integrated) was integrated in reverse direction with a 10-bp deletion at junction site revealed by PCR and sequencing analysis. b, Sequences of the 5′ and 3′ junction sites of mCherry+ clones after IRESmCherry knock-in by HITI in the GFP-correction HEK293 line with IRESmCherry-1c donor. The blue pentagon and sequence highlighted in yellow indicate the Cas9/gRNA target sequence. The black line within the blue pentagon indicates the Cas9 cleavage site. The PAM sequence is underlined. c, Sequences of the 5′ and 3′ junction sites of mCherry+ clones after IRESmCherry knock-in by HITI in the GFP-correction HEK293 line with IRESmCherry-2c donor. d, Sequences of the 5′ and 3′ junction sites of mCherry+ clones after IRESmCherry knock-in by HITI in the GFP-correction HEK293 line with IRESmCherry-MC donor. e, Sequences of the 5′ junction site of mCherry+ clones after IRESmCherry knock-in by PITCh or HITI in the GFP-correction HEK293 line with IRESmCherry-MH donor. f, Sequences of the 5′ junction site of mCherry+ clones after IRESmCherry knock-in by HDR or HITI in the GFP-correction HEK293 line with IRESmCherry-HDR-2c donor. g, The fidelity of 5′ and 3′ junction sites of mCherry+ clones after IRESmCherry knock-in by HITI in the GFP-correction HEK293 line with IRESmCherry-1c, IRESmCherry-2c and IRESmCherry-MC donor. n, number of analysed clones. h, IRESmCherry-MC donor was transfected into the GFP-correction HEK293 line which has five copies of the target sequence in the genome. The ratio of IRESmCherry knock-in and mutation at all target loci of mCherry+ clones and the ratio of unmodified, mutated and reverse integrated target loci of mCherry clones were examined by sequencing. n, number of analysed clones. For gel source images, see Supplementary Fig. 1. For source data, see Supplementary Table 6.

Source data

Extended Data Figure 3 Optimization of nuclear transport of Cas9.

a, Schematic representation of a series of dCas9–Flag constructs with different nuclear localization signals. b, Representative immunofluorescence images of the transfected HEK293 cells stained with Flag antibody to visualize dCas9 localization. DNA was counterstained with DAPI. Scale bar, 50 μm. c, The nuclear/cytoplasm ratio of dCas9 with different NLS signals. n, number of analysed cells. Results were presented as mean ± s.e.m. The input data points were shown as black dots. **P < 0.01, unpaired Student’s t-test. d, e, Characterization of Cas9 nuclease activity in human ES cells. Agarose gel image (d) and quantification (e) show Surveyor nuclease assay performed with the gRNA targeting KCNQ1 gene and Cas9 with different NLSs, that is, Cas9 –NLS (Cas9 (no NLS)), Cas9 +NLS (1NLS-Cas9-1NLS) and Cas9 +BPNLS (1BPNLS-Cas9-1BPNLS). The two lower bands are cleaved DNA products by Surveyor nuclease. NHEJ (%) indicates the percentage of Cas9/gRNA-mediated gene modification. Results were obtained from three technical replicates and presented as mean ± s.e.m. The input data points were shown as black dots. *P < 0.05, paired Student’s t-test. For gel source image, see Supplementary Fig. 1. For source data, see Supplementary Table 7.

Source data

Extended Data Figure 4 In-depth analyses of GFP knock-in in mouse primary neurons.

a, b, Characterization of cultured primary neurons. Representative immunofluorescence images (a) and quantification (b) show that all of the neurons are EdU negative in this culture condition. All EdU+ cells were GFAP-positive glia, fractin-positive apoptotic cells, or fragmented cells. Results were obtained from 15 technical replicates and presented as mean. c, The percentage of knock-in cells (GFP+) per transfected cells (mCherry+) with HDR donor (Tubb3-HDR), PITCh donor (Tubb3-MH), 1-cut donor (Tubb3-1c), 2-cut donor (Tubb3-2c), 2-cut donor without polyA (Tubb3-2cd), or minicircle donor (Tubb3-MC). n, number of technical replicates. Results were presented as mean ± s.d. The input data points were shown as black dots. d, Left panel, schematic showing inserted DNA sequences, with or without bacteria backbone, with different HITI donors. Blue pentagon and yellow highlight, Cas9/gRNA target sequence. Black line within blue pentagon, Cas9 cleavage site. Underlined sequence, PAM sequence. pA, polyA. Right panel, subcellular distribution of TUBB3-GFP. n, number of analysed cells. e, Representative fluorescence images of the primary neurons transfected with BPNLS-Cas9, gRNA, and different donor plasmids (Tubb3-MC-scramble, Tubb3-HDR, Tubb3-MH, Tubb3-1c, Tubb3-2c, Tubb3-2cd or Tubb3-MC). Different intracellular localization patterns of TUBB3-GFP were observed for different donors. Scale bar, 100 μm. f, Left panel, schematic of knock-in by Tubb3-2c and Tubb3-MC donor at the 5′ and 3′ ends of Tubb3 coding region. Black half-arrows indicate PCR primers for detecting integrated sequences. Right panel shows the PCR result. gi, Sequences of the 5′ and 3′ junction sites after GFP knock-in by HITI in mouse primary neurons with Tubb3-2c (g), Tubb3-2cd (h) and Tubb3-MC (i) donor plasmids. For gel source images, see Supplementary Fig. 1. For source data, see Supplementary Table 8.

Source data

Extended Data Figure 5 HITI-mediated GFP knock-in in neurons in vitro and in vivo.

a, The percentage of knock-in (GFP+) cells in mouse primary neurons per transfected cells (mCherry+) with empty vector (−Cas9), Cas9 (+NLS) and Cas9 (+BPNLS). n, number of technical replicates. Results were presented as mean ± s.e.m. **P < 0.01, unpaired Student’s t-test. b, Representative immunofluorescence images of human ES cell-derived pan neurons transfected with BPNLS-Cas9, gRNA, and different donor plasmids (hTUBB3-1c or hTUBB3-2c). Scale bar, 100 μm. c, PCR analysis of integrated GFP gene at TUBB3 locus in human ES cell-derived pan neurons. d, e, Sequences of the 5′ junction sites after GFP knock-in by HITI in human ES cell-derived pan neurons with TUBB3-1c donor (d) and TUBB3-2c donor (e). f, Upper panel, schematic of GFP knock-in at the 3′ end of the Tubb3 coding region via Tubb3-MC donor in the neonatal mouse brain. Black half-arrows indicate PCR primers for detecting integrated sequences. Lower panel, genomic PCR results showing transgene integration at both 5′ and 3′ ends. g, Schematic of in vivo targeted GFP knock-in by HITI in the neonatal mouse brain. CAG-floxSTOP-Cas9, inducible BPNLS-Cas9 expression plasmid. ERT2-Cre-ERT2, tamoxifen (TAM) inducible Cre expression plasmid. Donor plasmids: Tubb3-HDR or Tubb3-MC. Tamoxifen was injected at P10 and P11. Mice were analysed at P21. h, Representative fluorescence images of GFP knock-in at the Tubb3 locus in neonatal mouse brain by inducible Cas9 expression with HDR donor (Tubb3-HDR) or minicircle HITI donor (Tubb3-MC). Scale bar, 100 μm. i, Relative knock-in efficiencies of HDR and HITI donors with or without tamoxifen treatment. n, number of pups obtained from two pregnant mice. Results were presented as mean ± s.d. The input data points were shown as black dots. **P < 0.01. N.S., not significant. Unpaired student’s t-test. For gel source images, see Supplementary Fig. 1. For source data, see Supplementary Table 9.

Source data

Extended Data Figure 6 HITI via in vivo DNA transfection.

a, Schematic for in vivo targeted GFP-NLS or luciferase gene knock-in by HITI. CAG-Cas9, gRNA (Ai14gRNA or Scramble-gRNA)-mCherry and minicircle donor (Ai14-GFPNLS-MC-scramble, Ai14-luc-MC-scramble, Ai14-GFPNLS-MC or Ai14-luc-MC) were locally delivered to mouse kidney or muscle via pressure-mediated transfection and/or electroporation at 8 postnatal weeks and analysed 2 weeks later. b, In vivo imaging of luciferase signals at day 2, day 8, and day 14 post-intramuscular injection of luciferase HITI constructs. Right leg (−Cas9) was injected with empty plasmid, Ai14gRNA-mCherry, and Ai14-luc-MC. Left leg (+Cas9) was injected with CAG-Cas9, Ai14gRNA-mCherry, and Ai14-luc-MC. Top, control wild-type (WT) mouse. Bottom, Ai14 mouse. c, Representative immunofluorescence images of GFP expression after intramuscular electroporation of GFP-NLS HITI constructs into Ai14 mouse quadriceps (left panel) and panniculus carnosus (right panel). Top panel, donor-cut only control (CAG-Cas9, Scramble-gRNA-mCherry, and Ai14-GFPNLS-MC-scramble). Middle panel, no Cas9 control (empty plasmid, Ai14gRNA-mCherry, and Ai14-GFPNLS-MC). Bottom panel, GFP-NLS HITI (CAG-Cas9, Ai14gRNA-mCherry, and Ai14-GFPNLS-MC). Insets, higher magnification images. Scale bar, 100 μm. d, In vivo imaging of luciferase signals at days 7 and 14 after pressure-mediated kidney transfection of luciferase HITI constructs. Left mouse (−Cas9) was delivered with empty plasmid, Ai14gRNA-mCherry and Ai14-luc-MC. Right mouse (+Cas9) was delivered with CAG-Cas9, Ai14gRNA-mCherry and Ai14-luc-MC. Top, control wild-type mouse. Bottom, Ai14 mouse. e, Ex vivo luciferase imaging of stomach and oesophagus (St+E), heart (H), liver (Li), spleen (Sp), lungs (Lu), right (R) and left (L) kidney (K), pancreas (Pa), small intestine (SI), caecum (Ce) and colon (Co). Arrow shows luciferase signal in the right kidney. Top, wild-type mouse. Bottom, Ai14 mouse. f, Representative immunofluorescence images of GFP expression after electroporation of GFP-NLS HITI constructs into Ai14 mouse kidney. Top panel, donor-cut only control (CAG-Cas9, Scramble-gRNA-mCherry and Ai14-GFPNLS-MC-scramble). Middle panel, no Cas9 control (empty plasmid, Ai14gRNA-mCherry and Ai14-GFPNLS-MC). Bottom panel, GFP-NLS HITI (CAG-Cas9, Ai14gRNA-mCherry and Ai14-GFPNLS-MC). Scale bar, 100 μm.

Extended Data Figure 7 AAV8-mediated HITI in cultured mouse primary neurons.

a, Representative immunofluorescence images of neurons infected with AAV-Cas9 and AAV-mTubb3. AAVs were packaged with serotype 8. Insets, higher magnification images. Scale bar, 50 μm. b, Representative immunofluorescence images of GFP knock-in at the Tubb3 locus in pan neurons after AAV8 infections. Top panel, AAV-mTubb3 only. Bottom panel, AAV-Cas9 and AAV-mTubb3. Scale bar, 100 μm. The absolute GFP knock-in efficiency was shown in the upper right corner of the picture. Results were obtained from three technical replicates and presented as mean ± s.d. c, Intracellular distribution of the GFP expression after AAV8 infection (AAV-Cas9 and AAV-mTubb3). n, number of analysed cells. d, Validation of correct gene knock-in by PCR. e, Sequences of the 5′ and 3′ junction sites after GFP knock-in by HITI. For gel source image, see Supplementary Fig. 1. For source data, see Supplementary Table 10.

Source data

Extended Data Figure 8 HITI via in vivo local injection of AAVs.

a, Schematic of AAV vectors for inserting GFP-NLS downstream of the CAG promoter at the Ai14 Rosa26 locus. The AAVs were packaged with serotype 8. Half-arrows indicate PCR primers to validate correct gene knock-in. b, Schematic of the experimental design for in vivo targeted GFP-NLS gene knock-in by HITI via intramuscular (IM) injection in Ai14 mice. AAV8s (AAV-Cas9 and AAV-Ai14-HITI) were injected into quadriceps at 8 postnatal weeks and analysed at 12 weeks. c, Immunofluorescence analysis of GFP expression after IM delivery of AAV8s. Top panel, no Cas9 control (AAV-Ai14-HITI only). Bottom panel, GFP-NLS HITI (AAV-Cas9 and AAV-Ai14-HITI). Dystrophin was used as a marker for muscle cytoskeletal protein. Scale bar, 100 μm. d, Schematic and sequencing analyses of 5′ and 3′ junctions of the integration sites for Mertk exon 2 inserted by HITI in the eyes of RCS rats.

Extended Data Figure 9 HITI-mediated GFP-NLS knock-in via systemic intravenous injection in neonatal Ai14 mice.

a, Representative immunofluorescence images of GFP expression in the brain, muscle, kidney, adrenal gland, spleen, lung and choroid plexus of the eye after intravenous injection of AAV-Cas9 and AAV-Ai14-HITI. AAVs were packaged with serotype 8. Insets, higher magnification images. Scale bar, 100 μm. b, Comparison of HDR- and HITI-mediated targeted gene knock-in via systemic intravenous injection in neonatal mice. Left panel shows a schematic of AAV vectors for knock-in GFP downstream of the CAG promoter at the Ai14 Rosa26 locus. AAV-Ai14-scramble was used as a donor cut-only control. The HDR donor (AAV-Ai14-HDR) has homology arms at both ends of the GFP-NLS-pA cassette. The AAVs were packaged with serotype 8 and co-infected with AAV-Cas9 via IV, same as Fig. 4a. c, Absolute knock-in efficiency measured by the percentage of GFP+ cells in the liver and heart by HDR or HITI. Results were obtained from five slides and presented as mean ± s.d. **P < 0.01, unpaired Student’s t-test. d, Sequencing analyses of the 5′ and 3′ junction sites of heart and liver cells after GFP-NLS knock-in by HITI via intravenous AAV injections. For source data, see Supplementary Table 11.

Source data

Extended Data Figure 10 HITI via in vivo systemic injection of AAVs in mice.

a, Schematic of AAVs for knock-in luciferase downstream of the CAG promoter at the Ai14 Rosa26 locus. AAVs (AAV-Cas9 and AAV-Ai14-luc) were systemically delivered via tail vein injection in 8-week-old Ai14 mice and analysed at 12 weeks. The AAVs were packaged with serotype 9. b, In vivo imaging of luciferase signals at days 14 and 28 post-tail vein injection of luciferase HITI constructs. c, Ex vivo luciferase imaging analysis of testis (Te), stomach and oesophagus (St+E), heart (H), liver (Li), spleen (Sp), lungs (Lu), right (R) and left (L) kidney (K), pancreas (Pa), brain (Br), pituitary (Pi), right (R) and left (L) eye, (Ey), Tongue (To), small intestine (SI), caecum (Ce) and colon (Co). d, Representative immunofluorescence images of GFP expression in the liver after tail vein injection of HITI GFP-NLS gene knock-in AAV9. The absolute efficiency of GFP knock-in was shown in the bottom right corner. Results were obtained from five slides and presented as mean ± s.d. Scale bar, 200 μm. e, A list of on- and off-target sites that were used to determine the indel frequency of HITI mediated genome modifications using genomic DNA isolated from the liver. The nucleotide letters shown in red are the individual mismatches in predicted off-target sites.

Supplementary information

Supplementary Information

This file contains the uncropped gels for Figures 3d, 4c and Extended Data Figures 1b, 2a, 3d, 4f, 5c, f, 7d (PDF 787 kb)

PowerPoint slides

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Suzuki, K., Tsunekawa, Y., Hernandez-Benitez, R. et al. In vivo genome editing via CRISPR/Cas9 mediated homology-independent targeted integration. Nature 540, 144–149 (2016). https://doi.org/10.1038/nature20565

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature20565

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