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:

High-efficiency transgene integration by homology-directed repair in human primary cells using DNA-PKcs inhibition

Abstract

Therapeutic applications of nuclease-based genome editing would benefit from improved methods for transgene integration via homology-directed repair (HDR). To improve HDR efficiency, we screened six small-molecule inhibitors of DNA-dependent protein kinase catalytic subunit (DNA-PKcs), a key protein in the alternative repair pathway of non-homologous end joining (NHEJ), which generates genomic insertions/deletions (INDELs). From this screen, we identified AZD7648 as the most potent compound. The use of AZD7648 significantly increased HDR (up to 50-fold) and concomitantly decreased INDELs across different genomic loci in various therapeutically relevant primary human cell types. In all cases, the ratio of HDR to INDELs markedly increased, and, in certain situations, INDEL-free high-frequency (>50%) targeted integration was achieved. This approach has the potential to improve the therapeutic efficacy of cell-based therapies and broaden the use of targeted integration as a research tool.

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: AZD7648 is the most potent DNA-PKcs inhibitor for improving gene targeting in PSCs.
Fig. 2: AZD7648 improves gene targeting in HSPCs.
Fig. 3: AZD7648 treatment improves gene targeting with seemingly inactive and low-activity gRNAs.
Fig. 4: AZD7648 improves gene targeting with lower amounts of RNP and AAV6.
Fig. 5: AZD7648 improves gene targeting in primary human T cells and HBECs.

Similar content being viewed by others

Data availability

NGS data for off-target analysis have been deposited in the National Center for Biotechnology Informationʼs Sequence Read Archive database (accession number PRJNA982854)87 and can be accessed at http://www.ncbi.nlm.nih.gov/bioproject/982854. Sequences of the gRNA, PCR primers, ddPCR primers and probes have been included in Supplementary Table 1. Raw data for ICE analysis and screenshots of Sanger sequencing chromatograms are provided in the Supplementary Information. Source data are provided with this paper.

References

  1. Carroll, D. Genome engineering with targetable nucleases. Annu. Rev. Biochem. 83, 409–439 (2014).

    Article  CAS  PubMed  Google Scholar 

  2. Gaj, T., Gersbach, C. A. & Barbas, C. F. III. ZFN, TALEN, and CRISPR/Cas-based methods for genome engineering. Trends Biotechnol. 31, 397–405 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Horvath, P. & Barrangou, R. CRISPR/Cas, the immune system of bacteria and archaea. Science 327, 167–170 (2010).

    Article  CAS  PubMed  Google Scholar 

  4. Doudna, J. A. & Charpentier, E. Genome editing. The new frontier of genome engineering with CRISPR–Cas9. Science 346, 1258096 (2014).

    Article  PubMed  Google Scholar 

  5. Hsu, P. D., Lander, E. S. & Zhang, F. Development and applications of CRISPR–Cas9 for genome engineering. Cell 157, 1262–1278 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Nambiar, T. S., Baudrier, L., Billon, P. & Ciccia, A. CRISPR-based genome editing through the lens of DNA repair. Mol. Cell 82, 348–388 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Betermier, 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  Google Scholar 

  10. Kanaar, R., Hoeijmakers, J. H. & van Gent, D. C. Molecular mechanisms of DNA double strand break repair. Trends Cell Biol. 8, 483–489 (1998).

    Article  CAS  PubMed  Google Scholar 

  11. Chang, H. H. Y., Pannunzio, N. R., Adachi, N. & Lieber, M. R. Non-homologous DNA end joining and alternative pathways to double-strand break repair. Nat. Rev. Mol. Cell Biol. 18, 495–506 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Sfeir, A. & Symington, L. S. Microhomology-mediated end joining: a back-up survival mechanism or dedicated pathway? Trends Biochem. Sci. 40, 701–714 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Seol, J. H., Shim, E. Y. & Lee, S. E. Microhomology-mediated end joining: good, bad and ugly. Mutat. Res. 809, 81–87 (2018).

    Article  CAS  PubMed  Google Scholar 

  14. Doudna, J. A. The promise and challenge of therapeutic genome editing. Nature 578, 229–236 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Porteus, M. H. A new class of medicines through DNA editing. N. Engl. J. Med. 380, 947–959 (2019).

    Article  CAS  PubMed  Google Scholar 

  16. Fu, Y. W. et al. Dynamics and competition of CRISPR–Cas9 ribonucleoproteins and AAV donor-mediated NHEJ, MMEJ and HDR editing. Nucleic Acids Res. 49, 969–985 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Tatiossian, K. J. et al. Rational selection of CRISPR–Cas9 guide RNAs for homology-directed genome editing. Mol. Ther. 29, 1057–1069 (2021).

    Article  CAS  PubMed  Google Scholar 

  18. Hustedt, N. & Durocher, D. The control of DNA repair by the cell cycle. Nat. Cell Biol. 19, 1–9 (2016).

    Article  PubMed  Google Scholar 

  19. Xue, C. & Greene, E. C. DNA repair pathway choices in CRISPR–Cas9-mediated genome editing. Trends Genet. 37, 639–656 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Ray, U. & Raghavan, S. C. Modulation of DNA double-strand break repair as a strategy to improve precise genome editing. Oncogene 39, 6393–6405 (2020).

    Article  CAS  PubMed  Google Scholar 

  21. Sun, W. et al. Strategies for enhancing the homology-directed repair efficiency of CRISPR–Cas systems. CRISPR J. 5, 7–18 (2022).

    Article  CAS  PubMed  Google Scholar 

  22. Bischoff, N., Wimberger, S., Maresca, M. & Brakebusch, C. Improving precise CRISPR genome editing by small molecules: is there a magic potion? Cells 9, 1318 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Fattah, F. et al. Ku regulates the non-homologous end joining pathway choice of DNA double-strand break repair in human somatic cells. PLoS Genet. 6, e1000855 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  24. Frit, P., Ropars, V., Modesti, M., Charbonnier, J. B. & Calsou, P. Plugged into the Ku-DNA hub: the NHEJ network. Prog. Biophys. Mol. Biol. 147, 62–76 (2019).

    Article  CAS  PubMed  Google Scholar 

  25. Zhao, B. et al. The essential elements for the noncovalent association of two DNA ends during NHEJ synapsis. Nat. Commun. 10, 3588 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Chu, V. T. et al. Increasing the efficiency of homology-directed repair for CRISPR–Cas9-induced precise gene editing in mammalian cells. Nat. Biotechnol. 33, 543–548 (2015).

    Article  CAS  PubMed  Google Scholar 

  27. Robert, F., Barbeau, M., Ethier, S., Dostie, J. & Pelletier, J. Pharmacological inhibition of DNA-PK stimulates Cas9-mediated genome editing. Genome Med. 7, 93 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Maruyama, T. et al. Increasing the efficiency of precise genome editing with CRISPR–Cas9 by inhibition of nonhomologous end joining. Nat. Biotechnol. 33, 538–542 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Pinder, J., Salsman, J. & Dellaire, G. Nuclear domain ‘knock-in’ screen for the evaluation and identification of small molecule enhancers of CRISPR-based genome editing. Nucleic Acids Res. 43, 9379–9392 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Canny, M. D. et al. Inhibition of 53BP1 favors homology-dependent DNA repair and increases CRISPR–Cas9 genome-editing efficiency. Nat. Biotechnol. 36, 95–102 (2018).

    Article  CAS  PubMed  Google Scholar 

  31. Riesenberg, S. et al. Simultaneous precise editing of multiple genes in human cells. Nucleic Acids Res. 47, e116 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Fok, J. H. L. et al. AZD7648 is a potent and selective DNA-PK inhibitor that enhances radiation, chemotherapy and olaparib activity. Nat. Commun. 10, 5065 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  33. Martin, R. M. et al. Highly efficient and marker-free genome editing of human pluripotent stem cells by CRISPR–Cas9 RNP and AAV6 donor-mediated homologous recombination. Cell Stem Cell 24, 821–828 (2019).

    Article  CAS  PubMed  Google Scholar 

  34. Goldberg, F. W. et al. The discovery of 7-methyl-2-[(7-methyl[1,2,4]triazolo[1,5-apyridin-6-yl)amino]-9-(tetrahydro-2H-pyran-4-yl)-7,9-dihydro-8H-purin-8-one (AZD7648), a potent and selective DNA-dependent protein kinase (DNA-PK) inhibitor. J. Med. Chem. 63, 3461–3471 (2020).

    Article  CAS  PubMed  Google Scholar 

  35. Zenke, F. T. et al. Pharmacologic inhibitor of DNA-PK, M3814, potentiates radiotherapy and regresses human tumors in mouse models. Mol. Cancer Ther. 19, 1091–1101 (2020).

    Article  CAS  PubMed  Google Scholar 

  36. Khan, A. J. et al. VX-984 is a selective inhibitor of non-homologous end joining, with possible preferential activity in transformed cells. Oncotarget 9, 25833–25841 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Hardcastle, I. R. et al. Discovery of potent chromen-4-one inhibitors of the DNA-dependent protein kinase (DNA-PK) using a small-molecule library approach. J. Med. Chem. 48, 7829–7846 (2005).

    Article  CAS  PubMed  Google Scholar 

  38. Berger, M. et al. BAY-8400: a novel potent and selective DNA-PK inhibitor which shows synergistic efficacy in combination with targeted alpha therapies. J. Med. Chem. 64, 12723–12737 (2021).

    Article  CAS  PubMed  Google Scholar 

  39. Morrison, R., Al-Rawi, J. M., Jennings, I. G., Thompson, P. E. & Angove, M. J. Synthesis, structure elucidation, DNA-PK and PI3K and anti-cancer activity of 8- and 6-aryl-substituted-1-3-benzoxazines. Eur. J. Med. Chem. 110, 326–339 (2016).

    Article  CAS  PubMed  Google Scholar 

  40. Liu, R. et al. Homozygous defect in HIV-1 coreceptor accounts for resistance of some multiply-exposed individuals to HIV-1 infection. Cell 86, 367–377 (1996).

    Article  CAS  PubMed  Google Scholar 

  41. Papapetrou, E. P. & Schambach, A. Gene insertion into genomic safe harbors for human gene therapy. Mol. Ther. 24, 678–684 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Gomez-Ospina, N. et al. Human genome-edited hematopoietic stem cells phenotypically correct mucopolysaccharidosis type I. Nat. Commun. 10, 4045 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Scharenberg, S. G. et al. Engineering monocyte/macrophage-specific glucocerebrosidase expression in human hematopoietic stem cells using genome editing. Nat. Commun. 11, 3327 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Brown, N., Song, L., Kollu, N. R. & Hirsch, M. L. Adeno-associated virus vectors and stem cells: friends or foes? Hum. Gene Ther. 28, 450–463 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Tsuchiya, A., Kanno, T. & Nishizaki, T. PI3 kinase directly phosphorylates Akt1/2 at Ser 473/474 in the insulin signal transduction pathway. J. Endocrinol. 220, 49–59 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Chen, J., Ghorai, M. K., Kenney, G. & Stubbe, J. Mechanistic studies on bleomycin-mediated DNA damage: multiple binding modes can result in double-stranded DNA cleavage. Nucleic Acids Res. 36, 3781–3790 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Dever, D. P. et al. CRISPR/Cas9 β-globin gene targeting in human haematopoietic stem cells. Nature 539, 384–389 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Lattanzi, A. et al. Development of β-globin gene correction in human hematopoietic stem cells as a potential durable treatment for sickle cell disease. Sci. Transl. Med. 13, eabf2444 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Vaidyanathan, S. et al. High-efficiency, selection-free gene repair in airway stem cells from cystic fibrosis patients rescues CFTR function in differentiated epithelia. Cell Stem Cell 26, 161–171 (2020).

    Article  CAS  PubMed  Google Scholar 

  50. Cromer, M. K. et al. Gene replacement of α-globin with β-globin restores hemoglobin balance in β-thalassemia-derived hematopoietic stem and progenitor cells. Nat. Med. 27, 677–687 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Pavel-Dinu, M. et al. Gene correction for SCID-X1 in long-term hematopoietic stem cells. Nat. Commun. 10, 1634 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  52. Romero, Z. et al. Editing the sickle cell disease mutation in human hematopoietic stem cells: comparison of endonucleases and homologous donor templates. Mol. Ther. 27, 1389–1406 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Pavani, G. et al. Ex vivo editing of human hematopoietic stem cells for erythroid expression of therapeutic proteins. Nat. Commun. 11, 3778 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Dudek, A. M. & Porteus, M. H. Answered and unanswered questions in early-stage viral vector transduction biology and innate primary cell toxicity for ex-vivo gene editing. Front. Immunol. 12, 660302 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Rose, J. C. et al. Suppression of unwanted CRISPR–Cas9 editing by co-administration of catalytically inactivating truncated guide RNAs. Nat. Commun. 11, 2697 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Coelho, M. A. et al. CRISPR GUARD protects off-target sites from Cas9 nuclease activity using short guide RNAs. Nat. Commun. 11, 4132 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. DeWitt, M. A. et al. Selection-free genome editing of the sickle mutation in human adult hematopoietic stem/progenitor cells. Sci. Transl. Med. 8, 360ra134 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Cavazzana, M., Antoniani, C. & Miccio, A. Gene therapy for β-hemoglobinopathies. Mol. Ther. 25, 1142–1154 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Gilman, J. G. et al. Distal CCAAT box deletion in the A γ globin gene of two black adolescents with elevated fetal A γ globin. Nucleic Acids Res. 16, 10635–10642 (1988).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Chen, F. et al. High-frequency genome editing using ssDNA oligonucleotides with zinc-finger nucleases. Nat. Methods 8, 753–755 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Traxler, E. A. et al. A genome-editing strategy to treat β-hemoglobinopathies that recapitulates a mutation associated with a benign genetic condition. Nat. Med. 22, 987–990 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Liu, Y. et al. Activated STING in a vascular and pulmonary syndrome. N. Engl. J. Med. 371, 507–518 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Piras, F. et al. Lentiviral vectors escape innate sensing but trigger p53 in human hematopoietic stem and progenitor cells. EMBO Mol. Med. 9, 1198–1211 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Ferrari, S. et al. Efficient gene editing of human long-term hematopoietic stem cells validated by clonal tracking. Nat. Biotechnol. 38, 1298–1308 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Wiebking, V. et al. Genome editing of donor-derived T-cells to generate allogenic chimeric antigen receptor-modified T cells: optimizing αβ T cell-depleted haploidentical hematopoietic stem cell transplantation. Haematologica 106, 847–858 (2021).

    Article  CAS  PubMed  Google Scholar 

  66. Mohiuddin, I. S. & Kang, M. H. DNA-PK as an emerging therapeutic target in cancer. Front. Oncol. 9, 635 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  67. Hu, S. et al. Small molecule DNA-PK inhibitors as potential cancer therapy: a patent review (2010–present). Expert Opin. Ther. Pat. 31, 435–452 (2021).

    Article  CAS  PubMed  Google Scholar 

  68. Yue, X., Bai, C., Xie, D., Ma, T. & Zhou, P. K. DNA-PKcs: a multi-faceted player in DNA damage response. Front. Genet. 11, 607428 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Vakulskas, C. A. et al. A high-fidelity Cas9 mutant delivered as a ribonucleoprotein complex enables efficient gene editing in human hematopoietic stem and progenitor cells. Nat. Med. 24, 1216–1224 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Cromer, M. K. et al. Comparative analysis of CRISPR off-target discovery tools following ex vivo editing of CD34+ hematopoietic stem and progenitor cells. Mol. Ther. 31, 1074–1087 (2023).

    Article  CAS  PubMed  Google Scholar 

  71. Shi, Y., Inoue, H., Wu, J. C. & Yamanaka, S. Induced pluripotent stem cell technology: a decade of progress. Nat. Rev. Drug Discov. 16, 115–130 (2017).

    Article  CAS  PubMed  Google Scholar 

  72. Yamanaka, S. Pluripotent stem cell-based cell therapy—promise and challenges. Cell Stem Cell 27, 523–531 (2020).

    Article  CAS  PubMed  Google Scholar 

  73. Hockemeyer, D. & Jaenisch, R. Induced pluripotent stem cells meet genome editing. Cell Stem Cell 18, 573–586 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Sweeney, C. L. et al. Correction of X-CGD patient HSPCs by targeted CYBB cDNA insertion using CRISPR/Cas9 with 53BP1 inhibition for enhanced homology-directed repair. Gene Ther. 28, 373–390 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. June, C. H., O’Connor, R. S., Kawalekar, O. U., Ghassemi, S. & Milone, M. C. CAR T cell immunotherapy for human cancer. Science 359, 1361–1365 (2018).

    Article  CAS  PubMed  Google Scholar 

  77. Sather, B. D. et al. Efficient modification of CCR5 in primary human hematopoietic cells using a megaTAL nuclease and AAV donor template. Sci. Transl. Med. 7, 307ra156 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  78. Eyquem, J. et al. Targeting a CAR to the TRAC locus with CRISPR/Cas9 enhances tumour rejection. Nature 543, 113–117 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Berical, A., Lee, R. E., Randell, S. H. & Hawkins, F. Challenges facing airway epithelial cell-based therapy for cystic fibrosis. Front. Pharmacol. 10, 74 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Vaidyanathan, S. et al. Targeted replacement of full-length CFTR in human airway stem cells by CRISPR–Cas9 for pan-mutation correction in the endogenous locus. Mol. Ther. 30, 223–237 (2022).

    Article  CAS  PubMed  Google Scholar 

  81. Wilson, L. O. W., O’Brien, A. R. & Bauer, D. C. The current state and future of CRISPR–Cas9 gRNA design tools. Front. Pharmacol. 9, 749 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  82. Hendel, A. et al. Chemically modified guide RNAs enhance CRISPR–Cas genome editing in human primary cells. Nat. Biotechnol. 33, 985–989 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Aurnhammer, C. et al. Universal real-time PCR for the detection and quantification of adeno-associated virus serotype 2-derived inverted terminal repeat sequences. Hum. Gene Ther. Methods 23, 18–28 (2012).

    Article  CAS  PubMed  Google Scholar 

  84. Khan, T. A. et al. Neuronal defects in a human cellular model of 22q11.2 deletion syndrome. Nat. Med. 26, 1888–1898 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Hung, K. L. et al. Engineering protein-secreting plasma cells by homology-directed repair in primary human B cells. Mol. Ther. 26, 456–467 (2018).

    Article  CAS  PubMed  Google Scholar 

  86. Rogers, G. L. et al. Optimization of AAV6 transduction enhances site-specific genome editing of primary human lymphocytes. Mol. Ther. Methods Clin. Dev. 23, 198–209 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Selvaraj, S. et al. High-efficiency transgene integration by homology-directed repair in human primary cells using DNA-PKcs inhibition. Sequence Read Archive http://www.ncbi.nlm.nih.gov/bioproject/982854 (2023).

Download references

Acknowledgements

We thank the Binns Program for Cord Blood Research for providing cord blood CD34+ HSPCs. We thank the FACS Core Facility at the Stanford Institute of Stem Cell Biology and Regenerative Medicine for access to the FACS machines. We would like to thank the Taube and Koret Foundation for funding support. We thank F. Suchy and the laboratory of H. Nakauchi at Stanford University for help with ddPCR primer/probe design.

Author information

Authors and Affiliations

Authors

Contributions

S.S., W.N.F., S. Viel, S. Vaidyanathan, A.M.D., M.G., S.J.R., F.K.E., A.R.O., L.X., M.P.-D. and S.E.L. designed and performed experiments and analyzed data. M.K.C. provided reagents for experiments. R.S. performed experiments and analyzed data. N.G.-O. and M.H.P. supervised the project. S.S. and M.H.P wrote the initial and final drafts of the manuscript with input from other authors.

Corresponding author

Correspondence to Matthew H. Porteus.

Ethics declarations

Competing interests

M.H.P. is a member of the scientific advisory board of Allogene Therapeutics and a member of the board of directors of Graphite Bio. S.S. and M.H.P. are inventors on intellectual property related to this work.

Peer review

Peer review information

Nature Biotechnology thanks the anonymous reviewers 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

Extended Data Fig. 1 Comparison of different DNA-PKcs inhibitors for gene targeting with AZD7648.

a. Allelic gene targeting efficiency post gene editing at the CCR5 locus for introduction of two stop codons with different concentrations of DNA-PKcs inhibitors (AZD7648, KU57788, LTURM34 and BAY8400) in comparison with the untreated cells (UNT) as measured by ICE analysis at 4 days post-gene editing (n = 3). b. Allelic gene targeting efficiency at the CCR5 locus for knock-in of UBC-GFP-bGHpA sequence with two different concentrations (0.5 and 0.25 μM) of various DNA-PKcs inhibitors (AZD7648, M3814, VX984 and BAY8400) as measured by ddPCR analysis at 3 days post-gene editing (n = 3). c. MTT cell viability assay on gene targeted PSCs (b) at 24 h, 48 h and 72 h post-gene editing represented as percent cell viability normalized to the mock cells (n = 3). All data in a, b and c are shown as mean ± SEM. Data in a, b and c were compared with one-way ANOVA and Tukey’s multiple comparisons test, *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001 and ns denotes not significant.

Source data

Extended Data Fig. 2 Biochemical validation of DNA-PKcs inhibition and CCR5 gene targeting with AZD7648 in PSC.

a. Western blot analysis for phospho DNA-PKcs (pSer2056), DNA-PKcs and phospho AKT (pSer473) in PSCs treated with bleomycin ± AZD for 2 h or gene targeted with and without AZD treatment at 2 h post gene editing. ACTB is used as a loading control. Control denotes sample not treated with bleomycin. Mock denotes control sample nucleofected without RNP. UNT denotes sample either treated with bleomycin or RNP-AAV6 gene editing and AZD denotes AZD7648 treated sample with either bleomycin or RNP-AAV6 gene editing treatment as indicated. b. Allelic gene targeting efficiency of PSCs gene targeted at the CCR5 locus for knock-in of UBC-GFP-bGHpA sequence (a) with and without AZD treatment at 72 h post gene editing as measured by ddPCR (n = 1). c. Percentage of gene targeted cells at the CCR5 locus for knock-in of UBC-GFP-bGHpA sequence with two different concentrations of AZD7648 (0.5 and 0.25 μM) in comparison with the untreated cells (UNT) as measured by flow cytometry for GFP at 5-days post-gene editing (n = 3). Mock and RNP only cells were used as negative controls. d. Allelic distribution of WT, INDEL, HDR frequencies in CCR5 gene edited PSCs (c) (n = 3). HDR frequency was measured by ddPCR analysis, WT and INDEL frequencies were measured using ICE analysis. Mean HDR to INDEL ratio is represented above the bars. All data in c and d are shown as mean ± SEM.

Source data

Source data

Extended Data Fig. 3 Pluripotency, trilineage differentiation and single cell cloning analysis of CCR5 gene targeted PSCs.

a. Percentage of cells positive for various markers of pluripotency (SSEA4, OCT3/4, SOX2 and NANOG) in the CCR5 gene edited PSCs (n = 3). SSEA4 expression was measured by flow cytometry analysis. OCT3/4, SOX2 and NANOG expression was assessed by quantification of immunofluorescence staining images for corresponding markers in fixed PSCs and normalized to the total cell count measured through DAPI staining (n = 3). All data are shown as mean ± SEM. b. CCR5 gene edited PSCs were differentiated into the three germ layers. Mean frequency of differentiated cells as assessed by flow cytometry for the expression of corresponding markers for ectoderm (PAX6 and NES), mesoderm (CD56 and T) and endoderm (CXCR4 and SOX17) (n = 2). c. PSCs were gene targeted at the CCR5 locus with and without different concentrations of AZD7648 (0.5, 0.25 and 0.1 µM) for single cell cloning analysis. Allelic gene targeting efficiency was measured using ddPCR analysis (n = 1). d. Gene targeted PSCs (c) were subjected to single cell cloning and the frequency of clones with mono-, bi-allelic and no gene targeting was measured using ddPCR and PCR analysis. For each condition, 9-10 clones were picked and analyzed (n = 1). e. Mean allelic gene targeting efficiency in PSC at the CCR5 locus for the knock-in of UBC-GFP-bGHpA sequence with pre-treatment only (preAZD-UNT), post-treatment (AZD) only and pre+post treatment (pre+post AZD) with AZD7648 (0.5 µM) (n = 2). For pre-treatment, cells were treated with AZD7648 for 24 hours before gene targeting and for post-treatment, cells were treated with AZD7648 for 24 hours post gene editing.

Source data

Extended Data Fig. 4 AZD7648 treatment improves gene targeting at HBB, CFTR and HBA1 loci in PSC.

a. Schematic for gene targeting at the HBB locus to edit SCD (E6V) mutation in exon 1 using RNP/AAV6 gene editing (upper panel). Allelic distribution of WT, INDEL and HDR frequencies following gene editing at the HBB locus with and without AZD7648 treatment as measured by ICE analysis (lower panel) (n = 3). Mean HDR to INDEL ratio is represented above the bars. b. Schematic for gene targeting at the CFTR locus to edit CF disease mutation (ΔF508) in exon 11 using RNP/AAV6 gene editing (upper panel). Allelic distribution of WT, INDEL and HDR frequencies following gene editing with and without AZD7648 treatment at the CFTR locus as measured by ICE analysis (lower panel) (n = 3). Mean HDR to INDEL ratio is represented above the bars. c. Schematic for gene targeting at the HBB locus for knock-in of UBC-GFP-bGHpA sequence using RNP/AAV6 gene editing (upper panel). Allelic distribution of WT, INDEL and HDR frequencies following gene editing at the HBB locus with and without AZD7648 and M3814 (2 and 0.5 μM) treatment (lower panel) (n = 3). HDR efficiency was measured through ddPCR analysis. WT and INDEL frequencies were measured by ICE analysis. Mean HDR to INDEL ratio is represented above the bars. d. Schematic for gene replacement at the HBA1 locus for knock-in of transgene-2A-YFP sequence using RNP/AAV6 gene editing (upper panel). Allelic distribution of WT, INDEL and HDR frequencies following gene editing at the HBA1 locus with and without AZD7648 and M3814 (2 and 0.5 μM) treatment (lower panel) (n = 3). HDR efficiency was measured through ddPCR analysis. WT and INDEL frequencies were measured by ICE analysis. Mean HDR to INDEL ratio is represented above the bars. All data in a, b, c and d are shown as mean ± SEM.

Source data

Extended Data Fig. 5 Gene targeting in HSPCs at CCR5 and HBA1 loci with AZD7648 treatment.

a. Comparison of different DNA-PKcs inhibitors for gene targeting at the CCR5 locus for knock-in of UBC-GFP-bGHpA sequence at two different concentrations (2 and 0.5 µM) in HSPCs. Allelic gene targeting efficiency was measured by ddPCR analysis and cell viability was assessed by measuring live cell count and normalizing the number to the Mock sample (n = 3). All data are shown as mean ± SEM. b. Allelic distribution of WT, INDEL and HDR frequencies following gene editing at the CCR5 locus for knock-in of UBC-GFP-bGHpA sequence with different concentrations of AZD7648 as indicated (n = 1). HDR frequency was measured using ddPCR, WT and INDEL frequencies were measured using ICE analysis. c. CFU assay was performed on Mock, RNP, AAV6 and RNP + AAV6 treated HSPCs following gene editing at the CCR5 locus with or without AZD7648 (0.5 μM) treatment. Plot shows the distribution of the mean percentages of CFU-GEMM (multi-potent granulocyte, erythroid, macrophage, megakaryocyte progenitor cells), CFU-GM (colony forming unit-granulocytes and monocytes) and BFU-E (erythroid burst forming units) colonies (n = 2). d. Schematic for gene replacement at the HBA1 locus to replace HBA1 with HBB sequence using RNP/AAV6 gene editing. e. Allelic distribution of WT, INDEL and HDR frequencies following gene editing at the HBA1 locus with or without AZD7648 treatment (0.5 μM) (n = 3). HDR efficiency was measured through ddPCR analysis. WT and INDEL frequencies were measured by ICE analysis. Mean HDR to INDEL ratio is represented above the bars. All data are shown as mean ± SEM. f. Mean percentage of viable cell count of HSPCs at 3 days post-gene targeting (GT) at CCR5 (n = 4), HBB (n = 2) and HBA1 (n = 2) loci in untreated (UNT) and AZD7648 (AZD, 0.5 μM) treated cells. Mock, AAV only and RNP only treated cells were included as controls. Viable cell counts were measured at 72 h post-gene editing and plotted as percentage relative to the mock cell count. For CCR5 gene targeting, cell viability is shown as mean ± SEM.

Source data

Extended Data Fig. 6 Off-target analysis, gene targeting at CCR5 and HBB loci in HSPCs.

a. HSPCs gene targeted at HBB, CCR5 and HBA1 loci with and without AZD7648 treatment (0.5 μM) were assessed for off-target activity at the top off-target sites (OT1 for HBB, OT39 for CCR5 and OT1 for HBA1) through next generation sequencing (NGS). Mean frequency of reads with insertion and deletion INDELs is shown following SNP/INDEL detection analysis of the NGS data (n = 2). b. HSPCs were gene edited with Cas9 RNP at the CCR5 locus with and without AZD7648 treatment in the presence or absence of AAV6 donor template (for knock-in of two stop codons). Allelic distribution of mean frequencies of WT, insertions, deletions and HDR were determined by ICE analysis (n = 2). c. HSPCs were gene edited with Cas9 RNP at the HBB locus with and without AZD7648 treatment in the presence or absence of AAV6 donor template (for SCD mutation editing). Allelic distribution of frequencies of WT, insertions, deletions and HDR were determined by using ICE analysis (n = 3). All data are shown as mean ± SEM. d. Allelic distribution of mean WT, INDEL and HDR frequencies following gene editing at the HBB locus for editing SCD mutation using RNP and ssODN donor with and without AZD7648 treatment (0.5 μM) (n = 2). ssODN donor was tested at two different concentrations as indicated (2.5 and 5 μM).

Source data

Extended Data Fig. 7 Gene targeting at the CCR5 and STING1 loci in HSPCs using seemingly low activity gRNAs.

a. HSPCs were gene targeted at the CCR5 locus for knock-in of UBC-GFP-bGH-pA sequence using high activity (sg11) and low activity gRNAs (sg1 and 4). Allelic distribution of WT, INDEL and HDR frequencies following gene editing at the CCR5 locus with and without AZD7648 treatment (n = 3). HDR frequency was determined by ddPCR analysis. WT and INDEL frequencies were determined by ICE analysis. Mean HDR to INDEL ratio is represented above the bars. All data are shown as mean ± SEM. b. Schematic for gene targeting at the IL2RG locus for knock-in of codon-optimized cDNA and bGHpA in exon 1 using RNP/AAV6 gene editing with a high activity gRNA (sg1) and a low activity gRNA (sg6). c. Schematic for gene targeting at the IL2RG locus for knock-in of codon-optimized cDNA and bGHpA in exon 1 using RNP/AAV6 gene editing with a high activity gRNA (sg1) and two low activity gRNAs (sg5, 7). d. Allelic distribution of WT, INDEL and HDR frequencies in HSPCs gene targeted at the IL2RG locus with or without AZD7648 treatment (0.5 μM) using sg1 and 6 gRNAs (n = 3) (b). HDR efficiency was measured through ddPCR analysis. WT and INDEL frequencies were measured by ICE analysis. Mean HDR to INDEL ratio is represented above the bars. All data are shown as mean ± SEM. e. Allelic distribution of mean WT, INDEL and HDR frequencies in HSPCs gene targeted at the IL2RG locus with or without AZD7648 treatment (0.5 μM) using sg1, 5 and 7 (n = 2) (c). HDR frequency was measured through ddPCR analysis. WT and INDEL frequencies were measured by ICE analysis. Mean HDR to INDEL ratio is represented above the bars.

Source data

Extended Data Fig. 8 AZD7648 improves gene targeting with lower amounts of RNP/AAV6 in PSC and HSPC.

a. Frequency of gene targeted cells measured by flow cytometry for GFP following gene editing at the CCR5 locus for knock-in of UBC-GFP-bGH-pA sequence with or without AZD7648 treatment (0.25 μM) using varying amounts of Cas9-RNP as indicated and fixed amount of AAV6 donor (MOI:2500) in PSCs (n = 3). RNP-1X denotes 250 μg/ml of Cas9 protein complexed with 100 μg/ml of gRNA. b. Frequency of gene targeted cells measured by flow cytometry for GFP following gene editing at the CCR5 locus for knock-in of UBC-GFP-bGH-pA sequence in PSCs with or without AZD7648 treatment (0.25 μM) using a fixed amount of RNP and varying amounts of AAV6 donor (n = 3). c. Frequency of gene targeted cells measured by flow cytometry for GFP following gene editing at the CCR5 locus for knock-in of UBC-GFP-bGH-pA sequence in HSPCs with or without AZD7648 treatment (0.5 μM) using a fixed amount of RNP and varying amounts of AAV6 donor (n = 3). d. Frequency of gene targeted alleles measured by ICE analysis following gene editing at the HBB locus for editing the SCD mutation in HSPCs with or without AZD7648 treatment (0.5 μM) using a fixed amount of RNP and varying amounts of AAV6 donor (n = 3). e. Fold change in live cell count at D2, 4 and 6 post gene editing relative to D0 of HSPCs gene targeted at the HBB locus (d) (n = 3). f. p21 (CDKN1A) expression levels in gene targeted HSPCs (d) at D2, 4 and 6 post gene editing represented as values relative to non-gene edited control cells (n = 3). p21 expression levels were measured by RT-ddPCR and normalized to the levels of the house keeping gene, TBP. g. Number of AAV genomes in gene targeted HSPCs (d) at D2, 4 and 6 post gene editing represented as viral genomes (vg) per cell (n = 3). AAV genomes were quantified using ddPCR amplifying the AAV ITR and the values were normalized to a reference locus, ZEB2. Mean values are represented above the bars. All data are shown as mean ± SEM.

Source data

Extended Data Fig. 9 AZD7648 improves gene targeting in T cells.

a. Comparison of different DNA-PKcs inhibitors for gene targeting at the CCR5 locus for knock-in of UBC-GFP-bGHpA sequence at two different concentrations (2 and 0.5 µM) in T cells. Allelic gene targeting efficiency was measured by ddPCR analysis and cell viability was assessed by measuring live cell counts and normalizing the number to the Mock sample (n = 3). b. Frequency of gene targeted cells as measured by flow cytometry for GFP expression following UBC-GFP-bGH-pA sequence knock-in at the CCR5 locus in T cells with varying concentrations of AZD7648 using RNP/AAV6 gene editing (n = 3). c. Percentage of viable cell count for the CCR5 gene targeted T cells (b) at 72 h post gene editing (n = 3) relative to the mock cells. d. Frequency of gene targeted cells as measured by flow cytometry for GFP following UBC-GFP-bGH-pA sequence knock-in at the CCR5 locus in T cells with or without AZD7648 (0.5 μM) treatment using a fixed amount of RNP and varying amounts of AAV6 donor (n = 3). e. Percentage of viable cell count for the CCR5 gene targeted T cells (d) at 72 h post gene editing (n = 3) relative to the mock cells. f. T cells were gene targeted at the TRAC locus for knock-in of CD19-CAR with and without AZD treatment (n = 3). Engineered CD19 CAR T cells were challenged with GFP+ Nalm6 leukemia target cells in co-culture at an effector to target ratio of 1:1 for 72 hours. Potency of the CAR T cell cytotoxicity activity was monitored by the residual percentage of GFP+ target cells by flow cytometry at 24 h and 48 h post challenge. All data are shown as mean ± SEM.

Source data

Extended Data Fig. 10 AZD7648 improves gene targeting in B cells and HBECs.

a. Allelic distribution of mean WT, INDEL and HDR frequencies following gene editing at the CCR5 locus for knock-in of UBC-GFP-bGH-pA sequence in B cells with varying concentrations of AZD7648 using RNP/AAV6 gene editing (n = 2). HDR frequency was measured by ddPCR. WT and INDEL frequencies were measured through ICE analysis. Mean HDR to INDEL ratio is represented above the bars. b. Schematic for gene targeting at exon 1 of the CFTR gene for knock-in of SFFV-Citrine-pA sequence using RNP/AAV6 gene editing. c. Frequency of gene targeted cells as measured by flow cytometry for Citrine expression following CFTR-exon 1 gene editing (b) with or without AZD7648 treatment (0.5 μM) in HBECs (n = 4). d. Fold change in the frequency of CFTR gene targeted HBECs (c) with AZD7648 (0.5 μM) treatment relative to the untreated cells (n = 4). All data in c and d are shown as mean ± SEM.

Source data

Supplementary information

Supplementary Information

Supplementary Figs. 1–14.

Reporting Summary

Supplementary Table 1

This table includes gRNA sequences, antibodies, ddPCR primer/probe sequences and PCR primer sequences.

Source data

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

Selvaraj, S., Feist, W.N., Viel, S. et al. High-efficiency transgene integration by homology-directed repair in human primary cells using DNA-PKcs inhibition. Nat Biotechnol (2023). https://doi.org/10.1038/s41587-023-01888-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1038/s41587-023-01888-4

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research