Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Letter
  • Published:

Optimized base editors enable efficient editing in cells, organoids and mice

Abstract

CRISPR base editing enables the creation of targeted single-base conversions without generating double-stranded breaks. However, the efficiency of current base editors is very low in many cell types. We reengineered the sequences of BE3, BE4Gam, and xBE3 by codon optimization and incorporation of additional nuclear-localization sequences. Our collection of optimized constitutive and inducible base-editing vector systems dramatically improves the efficiency by which single-nucleotide variants can be created. The reengineered base editors enable target modification in a wide range of mouse and human cell lines, and intestinal organoids. We also show that the optimized base editors mediate efficient in vivo somatic editing in the liver in adult mice.

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: Optimizing the coding sequence of BE3 improves protein expression and target base editing.
Figure 2: N-terminal NLS sequences increase the range and potency of target base editing.
Figure 3: Optimized enzymes induce efficient base editing in a wide range of cell systems.

Similar content being viewed by others

Accession codes

Primary accessions

Sequence Read Archive

  • SRP151111

References

  1. Komor, A.C., Kim, Y.B., Packer, M.S., Zuris, J.A. & Liu, D.R. Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage. Nature 533, 420–424 (2016).

    Article  CAS  Google Scholar 

  2. Nishida, K. et al. Targeted nucleotide editing using hybrid prokaryotic and vertebrate adaptive immune systems. Science 353, aaf8729 (2016).

    Article  Google Scholar 

  3. Hess, G.T. et al. Directed evolution using dCas9-targeted somatic hypermutation in mammalian cells. Nat. Methods 13, 1036–1042 (2016).

    Article  CAS  Google Scholar 

  4. Ma, Y. et al. Targeted AID-mediated mutagenesis (TAM) enables efficient genomic diversification in mammalian cells. Nat. Methods 13, 1029–1035 (2016).

    Article  CAS  Google Scholar 

  5. Gaudelli, N.M. et al. Programmable base editing of A.T to G.C in genomic DNA without DNA cleavage. Nature 551, 464–471 (2017).

    Article  CAS  Google Scholar 

  6. Cox, D.B.T. et al. RNA editing with CRISPR-Cas13. Science 358, 1019–1027 (2017).

    Article  CAS  Google Scholar 

  7. Rees, H.A. et al. Improving the DNA specificity and applicability of base editing through protein engineering and protein delivery. Nat. Commun. 8, 15790 (2017).

    Article  CAS  Google Scholar 

  8. Kim, K. et al. Highly efficient RNA-guided base editing in mouse embryos. Nat. Biotechnol. 35, 435–437 (2017).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  10. Kim, S., Bae, T., Hwang, J. & Kim, J.S. Rescue of high-specificity Cas9 variants using sgRNAs with matched 5′ nucleotides. Genome Biol. 18, 218 (2017).

    Article  Google Scholar 

  11. 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  Google Scholar 

  12. Kleinstiver, B.P. et al. Engineered CRISPR-Cas9 nucleases with altered PAM specificities. Nature 523, 481–485 (2015).

    Article  Google Scholar 

  13. Dow, L.E. et al. Inducible in vivo genome editing with CRISPR-Cas9. Nat. Biotechnol. 33, 390–394 (2015).

    Article  CAS  Google Scholar 

  14. Staahl, B.T. et al. Efficient genome editing in the mouse brain by local delivery of engineered Cas9 ribonucleoprotein complexes. Nat. Biotechnol. 35, 431–434 (2017).

    Article  CAS  Google Scholar 

  15. Komor, A.C. et al. Improved base excision repair inhibition and bacteriophage Mu Gam protein yields C:G-to-T:A base editors with higher efficiency and product purity. Sci. Adv. 3, eaao4774 (2017).

    Article  Google Scholar 

  16. Huang, S.M. et al. Tankyrase inhibition stabilizes axin and antagonizes Wnt signalling. Nature 461, 614–620 (2009).

    Article  CAS  Google Scholar 

  17. Schoumacher, M. et al. Inhibiting Tankyrases sensitizes KRAS-mutant cancer cells to MEK inhibitors via FGFR2 feedback signaling. Cancer Res. 74, 3294–3305 (2014).

    Article  CAS  Google Scholar 

  18. Mashima, T. et al. mTOR signaling mediates resistance to tankyrase inhibitors in Wnt-driven colorectal cancer. Oncotarget 8, 47902–47915 (2017).

    Article  Google Scholar 

  19. Cancer Genome Atlas Network. Comprehensive molecular characterization of human colon and rectal cancer. Nature 487, 330–337 (2012).

  20. Matano, M. et al. Modeling colorectal cancer using CRISPR-Cas9-mediated engineering of human intestinal organoids. Nat. Med. 21, 256–262 (2015).

    Article  CAS  Google Scholar 

  21. Cadoret, A. et al. New targets of beta-catenin signaling in the liver are involved in the glutamine metabolism. Oncogene 21, 8293–8301 (2002).

    Article  CAS  Google Scholar 

  22. Annunziato, S. et al. Modeling invasive lobular breast carcinoma by CRISPR/Cas9-mediated somatic genome editing of the mammary gland. Genes Dev. 30, 1470–1480 (2016).

    Article  CAS  Google Scholar 

  23. Wang, D. et al. Adenovirus-mediated somatic genome editing of Pten by CRISPR/Cas9 in mouse liver in spite of Cas9-specific immune responses. Hum. Gene Ther. 26, 432–442 (2015).

    Article  CAS  Google Scholar 

  24. Hu, J.H. et al. Evolved Cas9 variants with broad PAM compatibility and high DNA specificity. Nature 556, 57–63 (2018).

    Article  CAS  Google Scholar 

  25. Koblan, L.W. et al. Improving cytidine and adenine base editors by expression optimization and ancestral reconstruction. Nat. Biotechnol. https://doi.org/10.1038/nbt.4172 (2018).

    Article  CAS  Google Scholar 

  26. Ryu, S.M. et al. Adenine base editing in mouse embryos and an adult mouse model of Duchenne muscular dystrophy. Nat. Biotechnol. 36, 536–539 (2018).

    Article  CAS  Google Scholar 

  27. Yin, H. et al. Structure-guided chemical modification of guide RNA enables potent non-viral in vivo genome editing. Nat. Biotechnol. 35, 1179–1187 (2017).

    Article  CAS  Google Scholar 

  28. Dow, L.E. et al. A pipeline for the generation of shRNA transgenic mice. Nat. Protoc. 7, 374–393 (2012).

    Article  CAS  Google Scholar 

  29. Han, T. et al. R-Spondin chromosome rearrangements drive Wnt-dependent tumour initiation and maintenance in the intestine. Nat. Commun. 8, 15945 (2017).

    Article  CAS  Google Scholar 

  30. O'Rourke, K.P. et al. Transplantation of engineered organoids enables rapid generation of metastatic mouse models of colorectal cancer. Nat. Biotechnol. 35, 577–582 (2017).

    Article  CAS  Google Scholar 

  31. 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  Google Scholar 

  32. Bae, S., Park, J. & Kim, J.S. Cas-OFFinder: a fast and versatile algorithm that searches for potential off-target sites of Cas9 RNA-guided endonucleases. Bioinformatics 30, 1473–1475 (2014).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by a project grant from the NIH/NCI (CA195787-01), a U54 grant from the NIH/NCI (U54OD020355), a project grant from the Starr Cancer Consortium (I10-0095), a Research Scholar Award from the American Cancer Society (RSG-17-202-01), and a Stand Up to Cancer Colorectal Cancer Dream Team Translational Research Grant (SU2C-AACR-DT22-17). Stand Up to Cancer is a program of the Entertainment Industry Foundation. Research grants are administered by the American Association for Cancer Research, a scientific partner of SU2C. M.P.Z. is supported in part by National Cancer Institute (NCI) grant NIH T32 CA203702. E.M.S. was supported by a Medical Scientist Training Program grant from the National Institute of General Medical Sciences of the NIH under award number T32GM07739 to the Weill Cornell/Rockefeller/Sloan-Kettering Tri-Institutional MD–PhD Program and an F31 Award from the NCI/NIH under grant number 1 F31 CA224800-01. E.R.K. is supported by an F31 NRSA predoctoral fellowship from the NCI/NIH under award number F31CA192835. F.J.S.-R. was supported by the MSKCC TROT program (5T32CA160001) and is supported as an HHMI Hanna Gray Fellow. S.W.L. is supported as the Geoffrey Beene Chair of Cancer Biology and as an Investigator of the Howard Hughes Medical Institute. D.F.T. is supported by the Helmholtz Association (VH-NG-1114) and by the German Research Foundation (DFG) project B05, SFB/TR 209 'Liver Cancer'. L.E.D. was supported by a K22 Career Development Award from the NCI/NIH (CA 181280-01). The content is solely the responsibility of the authors and does not necessarily represent the official views of the NIH. We thank H. Varmus (Weill Cornell Medicine) for providing cells.

Author information

Authors and Affiliations

Authors

Contributions

M.P.Z. and E.M.S. performed experiments, analyzed data, and wrote the paper. A.K., M.F., A.S., S.G., E.M., T.H., J.T., and F.J.S.-R. performed experiments and analyzed data. M.B. and A.Y.S. performed and analyzed in vivo experiments. D.F.T. designed and supervised in vivo experiments. E.R.K. performed computational analysis of MSKCC IMPACT data. J.S., S.W.L., and C.R.V. supplied critical previously unreported reagents. L.E.D. performed and supervised experiments, analyzed data, and wrote the paper.

Corresponding author

Correspondence to Lukas E Dow.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Optimized editing constructs show equivalent generation of viral particles and transduction of target cells

a. Concentration of viral particles (IU/ml) present in supernatants from all base editing lentiviral constructs. b. Number of genomic integrations of each lentiviral construct (prior to puro selection), as measured by a Taqman copy number assay to detect the puro resistance (Pac) gene. c. Number of live NIH/3T3 cells at day 3 of puro selection. All graphs show mean values. Error bars represent s.e.m., n = 3 biologically independent experiments; statistics calculated using a two-way ANOVA with Tukey’s correction for multiple testing. No significant differences in either a or b; p>0.05.

Supplementary Figure 2 Codon usage across different Cas9 variants

a. Plots show frequency of codons across each of the 20 amino acids in different Cas9 variants. Green represents the most commonly used codon across all human genes. Red represents codons that are present in human genes less than 50% of the time that would be expected by chance. Grey represents codons that are neither the most frequent nor underrepresented. b. Percentage of favored, disfavored, and neutral codons across different Cas9 sequences.

Supplementary Figure 3 RA increases target base editing in transfection assays and improves the ratio of desired to non-desired target editing

Frequency (%) of C>T conversion and indel formation in co-transfected HEK293T cells with BE3 or RA, and FANCF.S1 (a) or CTNNB1.S45 (b) sgRNAs. Graphs show mean values. Error bars indicate s.e.m., n = 4 biologically independent experiments, asterisks (*) indicate a significant difference (p<0.05) between groups, using a two-way ANOVA with Sidak’s correction for multiple testing. c. Frequency (%) of unwanted target modifications (indels, C>A, C>G) in BE3 or RA expressing 3T3 cells generated with the PGK-Puro lentiviral vector. Graph shows mean values +/- s.e.m., n=3 biologically independent experiments. d. Relative increase in target base editing in RA-expressing lines, compared to BE3 cells. Error bars represent s.e.m., n = 12 different target cytosines among 5 different sgRNAs, includes values from day 2 and day 6; asterisks (*) indicate a significant difference (p<0.05) between groups, using a one-way ANOVA with Tukey’s correction for multiple testing.

Supplementary Figure 4 Optimizing the coding sequence of high-fidelity and PAM variant Cas9 enzymes improves protein expression

a. Giemsa stained NIH/3T3 cells following transduction with P2A-Puro lentiviruses, as indicated, and selection in puro for 6 days. Experiment was repeated 3 times with similar results. b. Flow cytometry plots showing fluorescence of GFP linked to original and optimized HF1, PAM variant, and BE3 enzymes. While most cells expressing optimized versions show much higher GFP fluorescence, a small fraction show low levels of GFP expression. This is likely due to integration-site specific effects on EF1-mediated transcription. c. Quantitation of mean GFP fluorescence intensity from original and optimized HF1, PAM variant, and BE3 enzymes. Error bars represent s.e.m., n = 3 biologically independent experiments.

Supplementary Figure 5 N-terminal Nuclear Localization Signal (NLS) sequences increase the efficiency and range of base editing

a. Schematic showing location of NLS sequences and linker size in each construct tested. To provide a fair comparison, each of the constructs shown carries the original (non-optimized) cDNA sequence. b. Frequency (%) of C>T conversion in co-transfected HEK293T cells with BE3, 2X, FNLS, FLAGlink, or BE4 CMV vectors and either FANCF.S1 or CTNNB1.S45 sgRNAs, as indicated. Graphs show mean values. Error bars represent s.e.m., n = 2–6 biologically independent experiments, as indicated; asterisks (*) indicate a significant difference (p<0.05) between groups, using a two-way ANOVA with Tukey’s correction for multiple testing. c. Frequency (%) of C>T conversion in the last edited cytosine relative to the first edited cytosine for each construct co-transfected with either FANCF.S1 or CTNNB1.S45 sgRNAs. Graphs show mean values. Error bars represent s.e.m., n=2-6 biologically independent experiments, as indicated; first number refers to FANCF.S1, the second to CTNNB1.S45. The BE3 condition for FANCF.S1 could not be calculated for more than one replicate as the other two showed zero editing at C11. Asterisks (*) indicate a significant difference (p<0.05) between groups, using a two-way ANOVA with Tukey’s correction for multiple testing.

Supplementary Figure 6 Expression of optimized base editors in stable NIH/3T3 and DLD1 cells

a. Immunoblot showing editor expression from PGK-Puro and P2A-Puro vectors in NIH/3T3 cells. b. Immunoblot showing editor expression from PGK-Puro and P2A-Puro vectors in DLD1 cells. c. Relative mRNA abundance of RA, 2X, and FNLS editors in NIH/3T3 stable cell lines. Graphs show mean values. Error bars represent s.e.m., n = 3 biologically independent experiments; no significant differences (p<0.05) between any of the groups, using a one-way ANOVA with Tukey’s correction for multiple testing. d. Immunoblot showing expression of each optimized editor in NIH/3T3s, relative to Cas9. Each blot was repeated at least two times with similar results.

Supplementary Figure 7 FNLS increases target base editing, the ratio of desired vs non-desired editing

a. Frequency (%) of C>T conversion in NIH/3T3 cells transduced with RA- or FNLS-P2A-Puro lentiviral vectors 2 days following introduction of different sgRNAs, as indicated. Editing in BE3-PGK-Puro cells (from Figure 1e) is shown for comparison. Graphs show mean values. Error bars represent s.e.m., n = 3 biologically independent experiments; asterisks (*) indicate a significant difference (p<0.05) between groups, using a two-way ANOVA with Tukey’s correction for multiple testing. b. Frequency (%) of unwanted target modifications (indels, C>A, C>G) in RA and FNLS expressing 3T3 cells generated with the P2A-Puro lentiviral vector. Graphs shows mean values +/- s.e.m.; n=3 biologically independent experiments. c. Relative change in base editing in FNLS-expressing lines, compared to RA cells. Graphs show mean values. Error bars represent s.e.m., n = 12 target cytosines across 5 different sgRNAs, includes day 2 and day 6; asterisks (*) indicate a significant difference (p<0.05) between groups, using an ANOVA with Tukey’s correction for multiple testing.

Supplementary Figure 8 FNLS increases editing and optimized BE4Gam reduces indel frequency in human cells

a. Frequency (%) of C>T conversion in H23 and DLD1 cells transduced with BE3-PGK-Puro, FNLS or BE4GamRA-P2A-Puro lentiviral vectors 6 days following introduction of sgRNAs targeting either FANCF.S1 or CTNNB1.S45. Graphs show mean values. Error bars represent s.e.m., n = 3 biologically independent experiments (n=2 for BE4Gam in H23 cells); asterisks (*) indicate a significant difference (p<0.05) between groups, using an ANOVA with Tukey’s correction for multiple testing. In cases where cultures were not completely transduced with sgRNA (due to incomplete antibiotic selection), editing was normalized to the percentage of tdTomato positive cells, as measured by flow cytometry at the time of collection. b. Frequency (%) of indels in DLD1, PC9, and, H23 cells expressing either BE3, RA, FNLS, or BE4Gam and infected with sgRNAs targeting either FANCF.S1 or CTNNB1.S45. Graphs show mean values. Error bars represent s.e.m., n = 3 biologically independent experiments (n=2 for BE4Gam in H23 cells), asterisks (*) indicate a significant difference (p<0.05) between groups, using an ANOVA with Tukey’s correction for multiple testing.

Supplementary Figure 9 Optimized BE4Gam reduces non-desired base editing compared to FNLS

Frequency (%) of unwanted target modifications (C>A, C>G) in DLD1, PC9, and H23 cells expressing either BE3, FNLS, of BE4Gam and infected with sgRNAs targeting either FANCF.S1 or CTNNB1.S45. Graphs show mean values. Error bars represent s.e.m., n = 3 biologically independent experiments.

Supplementary Figure 10 Editing at off target sites

a. Frequency (%) of C>T conversion of any C in the editing window at two predicted off target sites for FANCF.S1 and CTNNB1.S45 in DLD1 cells expressing BE3, RA, or FNLS. Graph shows mean values. Error bars represent s.e.m., n = 3 biologically independent experiments. b. Sanger sequencing chromatograms showing detectable off target editing for the Apc.492 sgRNA (indicated by blue arrowheads) in NIH/3T3 cells. No editing was detected for either of two predicted off-target sites for Apc.1405, or the top predicted off-target site for Pik3ca.545. We were unable to amplify the Pik3ca_OT2 target region from genomic DNA. Bases highlighted green represent the target cytosine, while bases in black represent mismatches to the perfect sgRNA target site. Chromatograms are representative of three independent experiments, each with similar results.

Supplementary Figure 11 2X increases the range of editing in human and mouse cells

a. Frequency (%) of C>T conversion in NIH/3T3 cells transduced with RA- or FNLS-P2A-Puro lentiviral vectors 2 and 6 days following introduction of different sgRNAs, as indicated. Editing in BE3-PGK-Puro cells (from Figure 1e) is shown for comparison. Graphs show mean values. Error bars represent s.e.m., n = 3 biologically independent experiments; asterisks (*) indicate a significant difference (p<0.05) between groups, using a two-way ANOVA with Tukey’s correction for multiple testing. b. Frequency (%) of unwanted target modifications (indels, C>A, C>G) in RA or 2X expressing NIH/3T3 cells at Day 6. Graph shows mean values. Error bars represent s.e.m., n=3 biologically independent experiments. c and d. Frequency (%) of target C>T conversion in DLD1 cells expressing either BE3, RA, or 2X, and infected with sgRNAs targeting FANCF.S1 (c) or CTNNB1.S45 (d). e. Frequency (%) of target C>T conversion in NIH/3T3 cells expressing either BE3, BE3RA, or 2X, and infected with an sgRNA targeting (mouse) Ctnnb1.S45. Graphs show mean values. Error bars represent s.e.m., n = 3 biologically independent experiments; asterisks (*) indicate a significant difference (p<0.05) between groups, using a two-way ANOVA with Tukey’s correction for multiple testing.

Supplementary Figure 12 Base editing induced mutational activation of CTNNB1 enables outgrowth following Tankyrase and MEK inhibition

a. Schematic overview of the fluorescence-based competitive proliferation assay. Parental cells are shown in gray, transduced cells (tdTomato+) are in red, and cells bearing the target editing are highlighted in blue. Neutral competition keeps both tdTomato+ and tdTomato- cell proportions constant, whereas positive or negative selection causes the tdTomato+ population to increase or decrease, respectively. b. BE3, RA, 2X, and FNLS-expressing DLD1 cells were transduced with CTNNB1.S45 sgRNAs and treated with DMSO (left) or XAV939 1uM + Trametinib 10nM (right). Graph shows the number of tdTomato+ cells relative to the start of the assay. Bars represents measurements every 5 days (0, 5, 10, 15). Graphs show mean values. Error bars represent s.e.m., n = 3 biologically independent experiments; asterisks (*) indicate a significant difference (p<0.05) between groups, using an ANOVA with Tukey’s correction for multiple testing. c. Same as in b. but using FANCF.S1 (control) sgRNA. Note the neutral impact on relative proliferation in all the conditions, in contrast to CTNNB1.S45.

Supplementary Figure 13 Multiplexed editing in organoids, and dox-inducible base editing in mESCs

a. Images show FNLS/Apc.1405 and FNLS/Apc.1405/Pik3ca.545 transfected organoids, following selection by RSPO1 withdrawal and treatment with 25nM Trametinib for 5 days. b. Sanger sequencing chromatograms of the Pik3ca target locus, showing enrichment of the Pik3caE545K mutation following selection with Trametinib. Multiplexed editing and MEK inhibitor selection experiments were repeated on three independent occasions with similar results. c. Sanger sequencing chromatogram showing inducible base-editing in the presence of doxycycline (dox) in mouse ES cell lines transduced with either Apc.1405 or Pi3kca.545 sgRNAs. Base editing only occurs in cells expressing RA. Chromatograms representative of experiments repeated at least two times with similar results.

Supplementary Figure 14 xFNLS and xF2X show increased editing in human cell lines

a. Immunoblot showing expression levels of different base editor variants in PC9 cells. Sanger sequencing chromatograms showing editing 6 days following introduction of FANCF.S1 or CTNNB1.S45 sgRNAs (cytosines highlighted in green) in human PC9 (b) or DLD1 (c) cells expressing stably expressing FNLS, xBE3, xF2X, or xFNLS. xFNLS and xF2X enhance editing relative to xBE3 but are not as effective as FNLS containing the original Cas9 sequence. As expected, xF2X markedly increases editing at cytosine 10 of the CTNNB1 target site, as noted for 2X. Chromatograms represent a single experiment performed in parallel with both cell lines.

Supplementary Figure 15 Lentiviral vectors generated in this manuscript

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–15 (PDF 2517 kb)

Life Sciences Reporting Summary (PDF 133 kb)

Supplementary Tables

Supplementary Tables 1–6 (PDF 385 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zafra, M., Schatoff, E., Katti, A. et al. Optimized base editors enable efficient editing in cells, organoids and mice. Nat Biotechnol 36, 888–893 (2018). https://doi.org/10.1038/nbt.4194

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nbt.4194

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