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:

Rapidly inducible Cas9 and DSB-ddPCR to probe editing kinetics

Abstract

We developed a chemically inducible Cas9 (ciCas9) and a droplet digital PCR assay for double-strand breaks (DSB-ddPCR) to investigate the kinetics of Cas9-mediated generation and repair of DSBs in cells. ciCas9 is a rapidly activated, single-component Cas9 variant engineered by replacing the protein's REC2 domain with the BCL-xL protein and fusing an interacting BH3 peptide to the C terminus. ciCas9 can be tunably activated by a compound that disrupts the BCL-xL–BH3 interaction within minutes. DSB-ddPCR demonstrates time-resolved, highly quantitative, and targeted measurement of DSBs. Combining these tools facilitated an unprecedented exploration of the kinetics of Cas9-mediated DNA cleavage and repair. We find that sgRNAs targeting different sites generally induce cleavage within minutes and repair within 1 or 2 h. However, we observe distinct kinetic profiles, even for proximal sites, and this suggests that target sequence and chromatin state modulate cleavage and repair kinetics.

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

Figure 1: Development of a chemically inducible Cas9 (ciCas9).
Figure 2: DSB-ddPCR is an accurate and precise way to quantify DSBs.
Figure 3: DSB and indel kinetics with ciCas9 and DSB-ddPCR.
Figure 4: ciCas9 specificity and basal activity can be tuned.
Figure 5: ciCas9 can be activated by a variety of BCL-xL disruptors.

Similar content being viewed by others

Accession codes

Primary accessions

Sequence Read Archive

Referenced accessions

Protein Data Bank

References

  1. Richardson, C.D., Ray, G.J., DeWitt, M.A., Curie, G.L. & Corn, J.E. Enhancing homology-directed genome editing by catalytically active and inactive CRISPR-Cas9 using asymmetric donor DNA. Nat. Biotechnol. 34, 339–344 (2016).

    Article  CAS  PubMed  Google Scholar 

  2. Lin, S., Staahl, B.T., Alla, R.K. & Doudna, J.A. Enhanced homology-directed human genome engineering by controlled timing of CRISPR/Cas9 delivery. eLife 3, e04766 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  3. Slaymaker, I.M. et al. Rationally engineered Cas9 nucleases with improved specificity. Science 351, 84–88 (2016).

    Article  CAS  PubMed  Google Scholar 

  4. Kleinstiver, B.P. et al. High-fidelity CRISPR-Cas9 nucleases with no detectable genome-wide off-target effects. Nature 529, 490–495 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Suzuki, K. et al. In vivo genome editing via CRISPR/Cas9 mediated homology-independent targeted integration. Nature 540, 144–149 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Knight, S.C. et al. Dynamics of CRISPR-Cas9 genome interrogation in living cells. Science 350, 823–826 (2015).

    Article  CAS  PubMed  Google Scholar 

  8. Liu, K.I. et al. A chemical-inducible CRISPR-Cas9 system for rapid control of genome editing. Nat. Chem. Biol. 12, 980–987 (2016).

    Article  CAS  PubMed  Google Scholar 

  9. Nihongaki, Y., Kawano, F., Nakajima, T. & Sato, M. Photoactivatable CRISPR-Cas9 for optogenetic genome editing. Nat. Biotechnol. 33, 755–760 (2015).

    Article  CAS  PubMed  Google Scholar 

  10. Davis, K.M., Pattanayak, V., Thompson, D.B., Zuris, J.A. & Liu, D.R. Small molecule-triggered Cas9 protein with improved genome-editing specificity. Nat. Chem. Biol. 11, 316–318 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Zetsche, B., Volz, S.E. & Zhang, F. A split-Cas9 architecture for inducible genome editing and transcription modulation. Nat. Biotechnol. 33, 139–142 (2015).

    Article  CAS  PubMed  Google Scholar 

  12. Oakes, B.L. et al. Profiling of engineering hotspots identifies an allosteric CRISPR-Cas9 switch. Nat. Biotechnol. 34, 646–651 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Nguyen, D.P. et al. Ligand-binding domains of nuclear receptors facilitate tight control of split CRISPR activity. Nat. Commun. 7, 12009 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Feng, J. et al. A general strategy to construct small molecule biosensors in eukaryotes. eLife 4, 4004 (2015).

    Google Scholar 

  15. Senturk, S. et al. Rapid and tunable method to temporally control gene editing based on conditional Cas9 stabilization. Nat. Commun. 8, 14370 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Maji, B. et al. Multidimensional chemical control of CRISPR–Cas9. Nat. Chem. Biol. 13, 9–11 (2017).

    Article  CAS  PubMed  Google Scholar 

  17. Ivashkevich, A., Redon, C.E., Nakamura, A.J., Martin, R.F. & Martin, O.A. Use of the γ-H2AX assay to monitor DNA damage and repair in translational cancer research. Cancer Lett. 327, 123–133 (2012).

    Article  CAS  PubMed  Google Scholar 

  18. Olive, P.L. & Banáth, J.P. The comet assay: a method to measure DNA damage in individual cells. Nat. Protoc. 1, 23–29 (2006).

    Article  CAS  PubMed  Google Scholar 

  19. Hanada, K. et al. The structure-specific endonuclease Mus81 contributes to replication restart by generating double-strand DNA breaks. Nat. Struct. Mol. Biol. 14, 1096–1104 (2007).

    Article  CAS  PubMed  Google Scholar 

  20. Tsai, S.Q. et al. GUIDE-seq enables genome-wide profiling of off-target cleavage by CRISPR-Cas nucleases. Nat. Biotechnol. 33, 187–197 (2015).

    Article  CAS  PubMed  Google Scholar 

  21. Liang, Z., Sunder, S., Nallasivam, S. & Wilson, T.E. Overhang polarity of chromosomal double-strand breaks impacts kinetics and fidelity of yeast non-homologous end joining. Nucleic Acids Res. 44, 2769–2781 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  22. Furda, A.M., Bess, A.S., Meyer, J.N. & Van Houten, B. Analysis of DNA damage and repair in nuclear and mitochondrial DNA of animal cells using quantitative PCR. Methods Mol. Biol. 920, 111–132 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Grégoire, M.-C. et al. Quantification and genome-wide mapping of DNA double-strand breaks. DNA Repair (Amst.) 48, 63–68 (2016).

    Article  Google Scholar 

  24. Lensing, S.V. et al. DSBCapture: in situ capture and sequencing of DNA breaks. Nat. Methods 13, 855–857 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Ran, F.A. et al. In vivo genome editing using Staphylococcus aureus Cas9. Nature 520, 186–191 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Rose, J.C. et al. A computationally engineered RAS rheostat reveals RAS–ERK signaling dynamics. Nat. Chem. Biol. 13, 119–126 (2017).

    Article  CAS  PubMed  Google Scholar 

  27. Nishimasu, H. et al. Crystal structure of Cas9 in complex with guide RNA and target DNA. Cell 156, 935–949 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Moreno-Mateos, M.A. et al. CRISPRscan: designing highly efficient sgRNAs for CRISPR-Cas9 targeting in vivo. Nat. Methods 12, 982–988 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Isaac, R.S. et al. Nucleosome breathing and remodeling constrain CRISPR-Cas9 function. eLife 5, e13450 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  30. Chen, X. et al. Probing the impact of chromatin conformation on genome editing tools. Nucleic Acids Res. 44, 6482–6492 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Doench, J.G. et al. Optimized sgRNA design to maximize activity and minimize off-target effects of CRISPR-Cas9. Nat. Biotechnol. 34, 184–191 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Rose, J.C. et al. Protocol for rapidly inducible Cas9 and DSB-ddPCR. Protocol Exchange http://dx.doi.org/10.1038/protex.2017.066 (2017).

  33. Hindson, C.M. et al. Absolute quantification by droplet digital PCR versus analog real-time PCR. Nat. Methods 10, 1003–1005 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Hart, T. et al. High-resolution CRISPR screens reveal fitness genes and genotype-specific cancer liabilities. Cell 163, 1515–1526 (2015).

    Article  CAS  PubMed  Google Scholar 

  35. Ashkenazi, A., Fairbrother, W.J., Leverson, J.D. & Souers, A.J. From basic apoptosis discoveries to advanced selective BCL-2 family inhibitors. Nat. Rev. Drug Discov. 16, 273–284 (2017).

    Article  CAS  PubMed  Google Scholar 

  36. Konopleva, M. et al. Mechanisms of apoptosis sensitivity and resistance to the BH3 mimetic ABT-737 in acute myeloid leukemia. Cancer Cell 10, 375–388 (2006).

    Article  CAS  PubMed  Google Scholar 

  37. Wei, G. et al. Chemical genomics identifies small-molecule MCL1 repressors and BCL-xL as a predictor of MCL1 dependency. Cancer Cell 21, 547–562 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Shoemaker, A.R. et al. A small-molecule inhibitor of Bcl-XL potentiates the activity of cytotoxic drugs in vitro and in vivo. Cancer Res. 66, 8731–8739 (2006).

    Article  CAS  PubMed  Google Scholar 

  39. Leverson, J.D. et al. Exploiting selective BCL-2 family inhibitors to dissect cell survival dependencies and define improved strategies for cancer therapy. Sci. Transl. Med. 7, 279ra40 (2015).

    Article  PubMed  Google Scholar 

  40. Tao, Z.-F. et al. Discovery of a potent and selective BCL-XL inhibitor with in vivo activity. ACS Med. Chem. Lett. 5, 1088–1093 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Wendt, M.D. et al. Discovery and structure-activity relationship of antagonists of B-cell lymphoma 2 family proteins with chemopotentiation activity in vitro and in vivo. J. Med. Chem. 49, 1165–1181 (2006).

    Article  CAS  PubMed  Google Scholar 

  42. Oltersdorf, T. et al. An inhibitor of Bcl-2 family proteins induces regression of solid tumours. Nature 435, 677–681 (2005).

    Article  CAS  PubMed  Google Scholar 

  43. Lessene, G. et al. Structure-guided design of a selective BCL-X(L) inhibitor. Nat. Chem. Biol. 9, 390–397 (2013).

    Article  CAS  PubMed  Google Scholar 

  44. Tse, C. et al. ABT-263: a potent and orally bioavailable Bcl-2 family inhibitor. Cancer Res. 68, 3421–3428 (2008).

    Article  CAS  PubMed  Google Scholar 

  45. Wilson, W.H. et al. Navitoclax, a targeted high-affinity inhibitor of BCL-2, in lymphoid malignancies: a phase 1 dose-escalation study of safety, pharmacokinetics, pharmacodynamics, and antitumour activity. Lancet Oncol. 11, 1149–1159 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Geisinger, J.M., Turan, S., Hernandez, S., Spector, L.P. & Calos, M.P. In vivo blunt-end cloning through CRISPR/Cas9-facilitated non-homologous end-joining. Nucleic Acids Res. 44, e76 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  47. Voss, N.R. & Gerstein, M. 3V: cavity, channel and cleft volume calculator and extractor. Nucleic Acids Res. 38, W555–W562 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Huang, P.-S. et al. RosettaRemodel: a generalized framework for flexible backbone protein design. PLoS One 6, e24109 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Kiani, S. et al. Cas9 gRNA engineering for genome editing, activation and repression. Nat. Methods 12, 1051–1054 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Langmead, B. & Salzberg, S.L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    PubMed  PubMed Central  Google Scholar 

  52. Goreshnik, I. & Maly, D.J. A small molecule-regulated guanine nucleotide exchange factor. J. Am. Chem. Soc. 132, 938–940 (2010).

    Article  CAS  PubMed  Google Scholar 

  53. Cer, R.Z., Mudunuri, U., Stephens, R. & Lebeda, F.J. IC50-to-K i: a web-based tool for converting IC50 to K i values for inhibitors of enzyme activity and ligand binding. Nucleic Acids Res. 37, W441–W445 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank M. Dickinson for assistance generating reagents. This work was supported by the NIH (R01GM086858 (D.J.M.), R01GM109110 (D.M.F.), and F30CA189793 (J.C.R.)), and by the NSF (0954242 (D.J.M.)).

Author information

Authors and Affiliations

Authors

Contributions

J.C.R., D.M.F., and D.J.M. conceived the study and designed the experiments. B.M.T. assisted in the design of ciCas9. J.C.R. and J.J.S. performed the experiments. J.J.S. prepared samples for high-throughput sequencing and ddPCR. H.V.D. performed fluorescence BH3 competition experiments under the supervision of J.C.R. and D.J.M. W.J.V. helped develop the DSB-ddPCR assay and analyze data under the supervision of J.H.B. J.C.R. and D.M.F. prepared custom Python scripts. J.C.R., D.M.F., and D.J.M. wrote and edited the manuscript. All authors approved the final manuscript.

Corresponding authors

Correspondence to Dustin J Maly or Douglas M Fowler.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Comparison of REC2 and BCL-xL domains and modeling domain replacement.

(a) The REC2 domain of Cas9 and (b) BCL-xL display similar globular structures and volumes. (c, d) Modeling the replacement of the REC2 domain with BCL-xL. Six BCL-xL models are shown (multiple colors). In order to accommodate differences in relative termini orientation between BCL-xL and the REC2 domain, a flexible linker (length = 15 residues) was used to connect residue 307 of Cas9 to BCL-xL, while the N-terminus of BCL-xL was connected directly to residue 179 of Cas9. Cas9.BCL constructs containing 15 versus 20 residue linkers exhibited no appreciable difference in activity (data not shown).

Supplementary Figure 2 A Cas9 variant in which the REC2 domain was replaced with BCL-xL retains activity.

(a, b) HEK-293T cells were transfected with Cas9.BCL, wild type Cas9, or no Cas9, in addition to one of two sgRNAs targeting the MYC locus (sgRNA3 or sgRNA5) (n =1). Indel quantification was performed using an analysis method similar to that previously described (Online Methods). These data indicated Cas9 variants in which the REC2 domain has been replaced by BCL-xL retain editing activity. (c) HEK-293T cells were transfected with plasmids encoding Cas9 or Cas9.BCL, each bearing a FLAG-tag, and cells were lysed after 24 hours. Expression was detected using an anti-FLAG antibody (Cell Signaling), and an anti-GAPDH (loading control) antibody.

Supplementary Figure 3 Schematic depiction of Cas9 constructs generated and tested in this study.

Constructs contained residues 4-198 of BCL-xL (green); the BH3 peptide sequence (magenta); a nuclear localization signal (NLS, orange); FLAG tag (yellow); and human codon optimized wild type S. pyogenes Cas9 (purple). Flexible linkers ranging in length from 5 to 30 amino acids (blue) are composed of glycine, serine, threonine, and alanine. NLS = nuclear localization sequence.

Supplementary Figure 4 Indel frequency at the AAVS1 locus as a function of ciCas9 linker length after 24 hours in the presence or absence of A3.

HEK-293T cells were transfected using Turbofectin 8.0 with (450 ng Cas9 construct, 450 ng AAVS1 sgRNA, and 100 ng pMAX-GFP) or (450 ng ciCas9 and 550 ng pMAX-GFP), or left untransfected. 24 hours after transfection, cells were treated with A3 (10 μM) or left untreated. 24 hours after drug treatment, cells were harvested, and indel were quantified as described in the Online Methods. Error bars depict the s.e.m. (n = 3 cell culture replicates).

Supplementary Figure 5 ciCas9 enables inducible genome editing in U2OS and HCT116 cells.

U2OS and HCT116 cells were transfected using Turbofectin 8.0 with (450 ng ciCas9, 450 ng AAVS1 sgRNA, and 100 ng pMAX-GFP) or (450 ng ciCas9 and 550 ng pMAX-GFP) as a no sgRNA control. 24 hours after transfection, cells were treated with A3 (10 μM final concentration) or left untreated. Cells were harvested 24 hours after drug addition and indels quantified as described in the Online Methods. Indel frequencies (%) for individual cell culture replicates (n = 3) are shown (solid line = mean). Editing via transient transfection is inefficient in these other cell lines53, as has been previously reported for U2OS and HCT116 cells using the same sgRNA54.

Supplementary Figure 6 Comparison of indel kinetics at the EMX1 locus with ciCas9, paCas9, and iCas9.

(a) Indel kinetics are shown as reported for paCas9 co-transfected with EMX1 sgRNA in HEK-293T cells. Indel frequencies were determined using T7E1 assay, which the authors note has a detection limit of 1%. Error bars depict the s.e.m. (n=4 cell culture replicates) (b) HEK-293T cells were transfected with ciCas9 and an identical EMX1 sgRNA. After 24 hours, cells were treated with A3 and harvested at the indicated time points. Indel frequencies were quantified using high-throughput sequencing, as described in the Online Methods, and are also plotted in Fig. 3a (right panel). Indels were detectable by 2 hours, the first time point tested, and this increase in indel frequency was significant relative to both the zero minute and two hour no drug time points (one-sided t-tests: n = 3, p = 0.00045 and p = 0.00056, respectively, n = 3 cell culture replicates). (c) Reproduction of previously reported EMX1* sgRNA kinetics with iCas9 in HEK-293 cells. No statistics were reported. (d) HEK-293T cells were transfected with ciCas9 and an identical EMX1* sgRNA. After 24 hours, cells were treated with A3 and harvested at the indicated time points. Indel frequencies were determined using high-throughput sequencing, as described in the Online Methods. Indels were detectable with ciCas9 at 30 min and significant relative to the zero minute time point (one-sided t-test, n = 3, p = 0.016). Error bars depict the s.e.m. (n = 3 cell culture replicates). The *EMX1 sgRNA used in these experiments has an additional transcribed G preceding the 20 bp target sequence, but it is otherwise identical to the EMX1 sgRNA used throughout this study.

Supplementary Figure 7 DSB-ddPCR is accurate and precise.

Standard curves show the relationship between DSB frequency as measured by the DSB-ddPCR assay and expected DSB frequency at the AAVS1, MYC and EMX1 loci. Technical duplicates are shown and, in most cases, overlap.

Supplementary Figure 8 Proteinase K effectively digests ciCas9 prior to DNA isolation for DSB-ddPCR and high-throughput sequencing.

HEK-293T cells were transfected with 450 ng ciCas9, 450 ng EMX1 sgRNA, and 100 ng pMAX-GFP per well in 12-well plates. 24 hours after transfection, cells were incubated with A3 (10 μM) for two hours, then harvested. Non-transfected wells were harvested as a control. After resuspension in 1:1 mixture of PBS and Buffer AL (DNeasy Blood & Tissue Kit, Qiagen), samples were split into thirds. One third was processed according to the DNA isolation protocol as described in Online Methods, including 1 hour proteinase K digestion at 56 ºC and DNA purification via spin-column (Proteinase K +, DNeasy Column +)—similar digestions are often used to remove Cas9 from DNA for sequencing, genome-wide DSB detection, or T7E1, SURVEYOR, or in vitro Cas9 cleavage assays8,28,31,55. One third was digested with proteinase K for 1 hour at 56 ºC only (Proteinase K +, DNeasy Column -), and one third of the cells underwent only the first step in processing—resuspension in PBS and Buffer AL (DNeasy kit)—but were not treated with Proteinase K or run on DNeasy column. Protein was then isolated from each sample by chloroform/methanol precipitation, resuspended in SDS-PAGE denaturing sample buffer, boiled, and western blotted. ciCas9 expression was evaluated using anti-FLAG antibody (Sigma), and anti-GAPDH used as a loading control. ciCas9 expression was undetectable after proteinase K digestion. n = 3 cell culture replicates for ciCas9, n = 1 for no transfection control.

Supplementary Figure 9 Evaluation of background signal throughout DSB-ddPCR timecourse experiments.

DSB frequency curves from Fig. 3a are plotted with corresponding no sgRNA and no transfection controls. Error bars depict the s.e.m. (n = 3 cell culture replicates).

Supplementary Figure 10 Determination of initial DSB generation rate for MYC sgRNA4 and sgRNA5.

Initial DSB generation rates were calculated by performing a linear regression over early time points (0-30 min, Fig. 3b) MYC sgRNA4 (slope = 0.66, standard error = 0.094; r2 = 0.88) and sgRNA5 (slope 1.2, standard error = 0.11; r2 = 0.94). The difference in initial rates is significant (analysis of covariance (ANCOVA), p = 0.0035). Linear regressions and statistical calculations were performed in Prism software. Individual cell culture replicates are plotted at each time point for each sgRNA, n = 3. Dashed lines show the fitted regression lines.

Supplementary Figure 11 High-resolution indel timecourse corresponding to DSB timecourses in (Fig 3b).

Statistically significant indels were detectable at 2 hours for both MYC sgRNA4 (p = 0.0017) and MYC sgRNA5 (p=0.00079) relative to t = 0 min. The increase in indel frequency was also significant relative to no drug (“- A3”) at 2 hours (MYC sgRNA4 p = 0.00071; MYC sgRNA5 p = 0.00062). All p-values were calculated with one-sided t-tests, n = 3, error bars depict the s.e.m. The left panel shows the full time course. The right panel shows the first two hours.

Supplementary Figure 12 ciCas9 exhibits improved editing specificity relative to wild-type Cas9.

HEK-293T cells were transfected using Turbofectin 8.0 with (450 ng ciCas9, 450 ng VEGFA sgRNA2 or VEGFA sgRNA3, and 100 ng pMAX-GFP) or (450 ng ciCas9 and 550 ng pMAX-GFP) as a no sgRNA control. 24 hours after transfection cells were treated with A3 (10 μM final concentration) or left untreated. Cells were harvested 24 hours after drug addition and indels quantified as described in the Online Methods. (a) On- and Off-target editing for VEGFA sgRNA2, and (b) the corresponding specificity ratios. (c) On- and Off-target editing for VEGFA sgRNA3, and (d) the corresponding specificity ratios. VEGFA sgRNA3. Indel frequencies (%) for individual cell culture replicates (n = 3) are shown (solid line = mean). For specificity ratios (on-target/off-target indel frequency ratios) error bars depict the s.e.m (n = 3 cell culture replicates).

Supplementary Figure 13 A ciCas9 variant (e-ciCas9) with specificity enhancing mutations to minimize off-target activity.

Three mutations (K810A, K1003A, R1060A) were introduced to generate enhanced specificity ciCas9 (e-ciCas9). HEK-293T cells were co-transfected with MYC sgRNA4 (MYCg4) or VEGFA sgRNA2 (VEGFAg2), and either ciCas9 or e-ciCas9. Cells were then incubated either in the presence or absence of A3 for 24 hours before harvesting. Indel frequencies at the on-target site and a prominent off-target site were determined via high throughput sequencing. (a) On-target and (b) off-target 4 (OT4) indel frequencies (%) for MYC sgRNA4. (c) Specificity ratios for MYC sgRNA4 on-target and OT4 editing. (d) On-target and (e) off-target 1 (OT1) indel frequencies (%) for VEGFA sgRNA2. (f) Specificity ratios for VEGFA sgRNA2 on-target and OT1 editing. Error bars = s.e.m., n = 3 cell culture replicates.

Supplementary Figure 14 Alterations in BH3/BCL-xL affinity do not affect ciCas9 variant expression.

HEK-293T cells were transfected with 450 ng ciCas9 or ciCas9(F22), 450 ng EMX1 sgRNA, and 100 ng pMAX-GFP per well in 12-well plates using Turbofectin 8.0. Non-transfected wells were included as a control. Cells were harvested and lysed 48 hours after transfection and ciCas9 variant expression was determined by western blot using antibodies targeting FLAG (Sigma) and GAPDH (loading control). Blots representative of two independent experiments.

Supplementary Figure 15 A3 does not affect DSBs, indels, or cell viability in the context of wild type Cas9-mediated editing.

HEK-293T cells were transfected with 225 ng of plasmid encoding wild type Cas9, 225 ng of plasmid encoding AAVS1 sgRNA, and 50 ng of plasmid encoding pMAX-GFP per well in 24-well plates using Turbofectin 8.0. 24 hours after transfection cells were treated with A3 (10 μM final concentration) or left untreated. (a-b) Cells were harvested at 0 and 24 hours after drug addition. (a) Indels were quantified via high-throughput sequencing, and (b) DSB frequency was determined via ddPCR. In parallel, (c) cell viability and (d) cell death relative to WT Cas9 (– A3) at 0 hours were determined using the LIVE/DEAD Viability/Cytotoxicity Kit (ThermoFisher). Error bars depict s.e.m. (n = 3 cell culture replicates).

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–15, Supplementary Tables 1–6 and Supplementary Note 1. (PDF 2135 kb)

Life Sciences Reporting Summary (PDF 129 kb)

Supplementary Dataset 1

Supplementary Dataset 1 (XLSX 58 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Rose, J., Stephany, J., Valente, W. et al. Rapidly inducible Cas9 and DSB-ddPCR to probe editing kinetics. Nat Methods 14, 891–896 (2017). https://doi.org/10.1038/nmeth.4368

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nmeth.4368

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