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:

Whole-brain in vivo base editing reverses behavioral changes in Mef2c-mutant mice

Abstract

Whole-brain genome editing to correct single-base mutations and reduce or reverse behavioral changes in animal models of autism spectrum disorder (ASD) has not yet been achieved. We developed an apolipoprotein B messenger RNA-editing enzyme, catalytic polypeptide-embedded cytosine base editor (AeCBE) system for converting C·G to T·A base pairs. We demonstrate its effectiveness by targeting AeCBE to an ASD-associated mutation of the MEF2C gene (c.104T>C, p.L35P) in vivo in mice. We first constructed Mef2cL35P heterozygous mice. Male heterozygous mice exhibited hyperactivity, repetitive behavior and social abnormalities. We then programmed AeCBE to edit the mutated C·G base pairs of Mef2c in the mouse brain through the intravenous injection of blood–brain barrier-crossing adeno-associated virus. This treatment successfully restored Mef2c protein levels in several brain regions and reversed the behavioral abnormalities in Mef2c-mutant mice. Our work presents an in vivo base-editing paradigm that could potentially correct single-base genetic mutations in the brain.

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: A de novo mutation in the MEF2C gene associated with ASD causes rapid degradation of the MEF2C protein.
Fig. 2: Reduced Mef2c protein levels and autistic-like behaviors in Mef2cL35P+/− mice.
Fig. 3: Design and validation of the AeCBE system.
Fig. 4: Restoration of Mef2c protein levels and abnormal inhibitory interneuron development in Mef2cL35P+/− mice after AeCBE editing.
Fig. 5: Restoration of spine density and impaired synaptic transmission in the medial PFC layer II and III neurons in Mef2cL35P+/− mice after AeCBE editing.
Fig. 6: Amelioration of autistic-like behaviors in Mef2cL35P+/− mice with AeCBE editing in vivo.

Similar content being viewed by others

Data availability

The datasets generated and analyzed during the present study are available from the corresponding authors upon reasonable request. Source data are provided with this paper.

Code availability

The present study does not contain any customized code.

References

  1. Colvert, E. et al. Heritability of autism spectrum disorder in a UK population-based twin sample. JAMA Psychiatry 72, 415–423 (2015).

    PubMed  PubMed Central  Google Scholar 

  2. Geschwind, D. H. & Flint, J. Genetics and genomics of psychiatric disease. Science 349, 1489–1494 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Skuse, D. H., Mandy, W. P. L. & Scourfield, J. Measuring autistic traits: heritability, reliability and validity of the Social and Communication Disorders Checklist. Br. J. Psychiatry 187, 568–572 (2005).

    PubMed  Google Scholar 

  4. Iossifov, I. et al. De novo gene disruptions in children on the autistic spectrum. Neuron 74, 285–299 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Sanders, S. J. et al. De novo mutations revealed by whole-exome sequencing are strongly associated with autism. Nature 485, 237–241 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Parikshak, N. N. et al. Integrative functional genomic analyses implicate specific molecular pathways and circuits in autism. Cell 155, 1008–1021 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Gilissen, C. et al. Genome sequencing identifies major causes of severe intellectual disability. Nature 511, 344–347 (2014).

    CAS  PubMed  Google Scholar 

  8. Guo, H. et al. Genome sequencing identifies multiple deleterious variants in autism patients with more severe phenotypes. Genet. Med. 21, 1611–1620 (2019).

    CAS  PubMed  Google Scholar 

  9. Shim, J. S. et al. MEF2C-Related 5q14.3 microdeletion syndrome detected by array CGH: a case report. Ann. Rehabil. Med. 39, 482–487 (2015).

    PubMed  PubMed Central  Google Scholar 

  10. Wang, J. et al. Novel MEF2C point mutations in Chinese patients with Rett (-like) syndrome or non-syndromic intellectual disability: insights into genotype–phenotype correlation. BMC Med. Genet. 19, 191 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Zhou, W.-Z. et al. Targeted resequencing of 358 candidate genes for autism spectrum disorder in a Chinese cohort reveals diagnostic potential and genotype–phenotype correlations. Hum. Mutat. 40, 801–815 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Le Meur, N. et al. MEF2C haploinsufficiency caused by either microdeletion of the 5q14.3 region or mutation is responsible for severe mental retardation with stereotypic movements, epilepsy and/or cerebral malformations. J. Med. Genet. 47, 22–29 (2010).

    CAS  PubMed  Google Scholar 

  13. Leifer, D. et al. MEF2C, a MADS/MEF2-family transcription factor expressed in a laminar distribution in cerebral cortex. Proc. Natl Acad. Sci. USA 90, 1546–1550 (1993).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Okamoto, S., Krainc, D., Sherman, K. & Lipton, S. A. Antiapoptotic role of the p38 mitogen-activated protein kinase-myocyte enhancer factor 2 transcription factor pathway during neuronal differentiation. Proc. Natl Acad. Sci. USA 97, 7561–7566 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Shalizi, A. et al. A calcium-regulated MEF2 sumoylation switch controls postsynaptic differentiation. Science 311, 1012–1017 (2006).

    CAS  Google Scholar 

  16. Flavell, S. W. et al. Activity-dependent regulation of MEF2 transcription factors suppresses excitatory synapse number. Science 311, 1008–1012 (2006).

    CAS  PubMed  Google Scholar 

  17. Harrington, A. J. et al. MEF2C regulates cortical inhibitory and excitatory synapses and behaviors relevant to neurodevelopmental disorders. eLife 5, e20059 (2016).

    PubMed Central  Google Scholar 

  18. Harrington, A. J. et al. MEF2C hypofunction in neuronal and neuroimmune populations produces MEF2C haploinsufficiency syndrome-like behaviors in mice. Biol. Psychiatry 88, 488–499 (2020).

    CAS  PubMed Central  Google Scholar 

  19. Barbosa, A. C. et al. MEF2C, a transcription factor that facilitates learning and memory by negative regulation of synapse numbers and function. Proc. Natl Acad. Sci. USA 105, 9391–9396 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    Google Scholar 

  22. Cai, Y. et al. In vivo genome editing rescues photoreceptor degeneration via a Cas9/RecA-mediated homology-directed repair pathway. Sci. Adv. 5, eaav3335 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. 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).

    CAS  PubMed Central  Google Scholar 

  25. Rees, H. A. & Liu, D. R. Base editing: precision chemistry on the genome and transcriptome of living cells. Nat. Rev. Genet. 19, 770–788 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Koblan, L. W. et al. In vivo base editing rescues Hutchinson–Gilford progeria syndrome in mice. Nature 589, 608–614 (2021).

    CAS  PubMed Central  Google Scholar 

  27. Newby, G. A. & Liu, D. R. In vivo somatic cell base editing and prime editing. Mol. Ther. 29, 3107–3124 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Arbab, M. et al. Base editing rescue of spinal muscular atrophy in cells and in mice. Science 380, eadg6518 (2023).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Cheng, T. L. et al. Expanding C-T base editing toolkit with diversified cytidine deaminases. Nat. Commun. 10, 3612 (2019).

    PubMed  PubMed Central  Google Scholar 

  30. Duan, Y. et al. Brain-wide Cas9-mediated cleavage of a gene causing familial Alzheimer’s disease alleviates amyloid-related pathologies in mice. Nat. Biomed. Eng. 6, 168–180 (2022).

    CAS  PubMed  Google Scholar 

  31. Zweier, M. et al. Mutations in MEF2C from the 5q14.3q15 microdeletion syndrome region are a frequent cause of severe mental retardation and diminish MECP2 and CDKL5 expression. Hum. Mutat. 31, 722–733 (2010).

    CAS  PubMed  Google Scholar 

  32. Bienvenu, T., Diebold, B., Chelly, J. & Isidor, B. Refining the phenotype associated with MEF2C point mutations. Neurogenetics 14, 71–75 (2013).

    PubMed  Google Scholar 

  33. Srivastava, S. et al. Clinical whole exome sequencing in child neurology practice. Ann. Neurol. 76, 473–483 (2014).

    PubMed  Google Scholar 

  34. Vrečar, I. et al. Further clinical delineation of the MEF2C haploinsufficiency syndrome: report on new cases and literature review of severe neurodevelopmental disorders presenting with seizures, absent speech, and involuntary movements. J. Pediatr. Genet. 6, 129–141 (2017).

    PubMed  PubMed Central  Google Scholar 

  35. Feigin, V. L. et al. Burden of neurological disorders across the US from 1990–2017: a global burden of disease study. JAMA Neurol. 78, 165–176 (2021).

    PubMed  Google Scholar 

  36. Tu, S. et al. NitroSynapsin therapy for a mouse MEF2C haploinsufficiency model of human autism. Nat. Commun. 8, 1488 (2017).

    PubMed Central  Google Scholar 

  37. Lyons, G. E., Micales, B. K., Schwarz, J., Martin, J. F. & Olson, E. N. Expression of mef2 genes in the mouse central nervous system suggests a role in neuronal maturation. J. Neurosci. 15, 5727–5738 (1995).

    CAS  PubMed Central  Google Scholar 

  38. Durand, S. et al. NMDA receptor regulation prevents regression of visual cortical function in the absence of Mecp2. Neuron 76, 1078–1090 (2012).

    CAS  PubMed Central  Google Scholar 

  39. He, L.-J. et al. Conditional deletion of Mecp2 in parvalbumin-expressing GABAergic cells results in the absence of critical period plasticity. Nat. Commun. 5, 5036 (2014).

    CAS  Google Scholar 

  40. Tong, D.-L. et al. The critical role of ASD-related gene CNTNAP3 in regulating synaptic development and social behavior in mice. Neurobiol. Dis. 130, 104486 (2019).

    CAS  PubMed  Google Scholar 

  41. Foss-Feig, J. H. et al. Searching for cross-diagnostic convergence: neural mechanisms governing excitation and inhibition balance in schizophrenia and autism spectrum disorders. Biol. Psychiatry 81, 848–861 (2017).

    PubMed  PubMed Central  Google Scholar 

  42. Sohal, V. S. & Rubenstein, J. L. R. Excitation–inhibition balance as a framework for investigating mechanisms in neuropsychiatric disorders. Mol. Psychiatry 24, 1248–1257 (2019).

    PubMed Central  Google Scholar 

  43. Walton, R. T., Christie, K. A., Whittaker, M. N. & Kleinstiver, B. P. Unconstrained genome targeting with near-PAMless engineered CRISPR-Cas9 variants. Science 368, 290–296 (2020).

    CAS  PubMed Central  Google Scholar 

  44. Wang, X. et al. Efficient base editing in methylated regions with a human APOBEC3A–Cas9 fusion. Nat. Biotechnol. 36, 946–949 (2018).

    CAS  PubMed  Google Scholar 

  45. Liu, Y. et al. A Cas-embedding strategy for minimizing off-target effects of DNA base editors. Nat. Commun. 11, 6073 (2020).

    CAS  PubMed Central  Google Scholar 

  46. Kim, Y. B. et al. Increasing the genome-targeting scope and precision of base editing with engineered Cas9-cytidine deaminase fusions. Nat. Biotechnol. 35, 371–376 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Koblan, L. W. et al. Improving cytidine and adenine base editors by expression optimization and ancestral reconstruction. Nat. Biotechnol. 36, 843–846 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Zettler, J., Schütz, V. & Mootz, H. D. The naturally split Npu DnaE intein exhibits an extraordinarily high rate in the protein trans-splicing reaction. FEBS Lett. 583, 909–914 (2009).

    CAS  PubMed  Google Scholar 

  49. Clement, K. et al. CRISPResso2 provides accurate and rapid genome editing sequence analysis. Nat. Biotechnol. 37, 224–226 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Malinin, N. L. et al. Defining genome-wide CRISPR–Cas genome-editing nuclease activity with GUIDE-seq. Nat. Protoc. 16, 5592–5615 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. 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).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Yang, K. et al. SENP1 in the retrosplenial agranular cortex regulates core autistic-like symptoms in mice. Cell Rep. 37, 109939 (2021).

    CAS  PubMed  Google Scholar 

  54. Guo, B. et al. Anterior cingulate cortex dysfunction underlies social deficits in Shank3 mutant mice. Nat. Neurosci. 22, 1223–1234 (2019).

    CAS  Google Scholar 

  55. Yu, B. et al. Reversal of social recognition deficit in adult mice with MECP2 duplication via normalization of MeCP2 in the medial prefrontal cortex. Neurosci. Bull. 36, 570–584 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Sun, L. et al. Visualization and correction of social abnormalities-associated neural ensembles in adult MECP2 duplication mice. Sci. Bull. 65, 1192–1202 (2020).

    CAS  Google Scholar 

  57. Li, S. et al. Docking sites inside Cas9 for adenine base editing diversification and RNA off-target elimination. Nat. Commun. 11, 5827 (2020).

    PubMed Central  Google Scholar 

  58. Zhang, S. et al. TadA reprogramming to generate potent miniature base editors with high precision. Nat. Commun. 14, 413 (2023).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Koblan, L. W. et al. Efficient C*G-to-G*C base editors developed using CRISPRi screens, target-library analysis, and machine learning. Nat. Biotechnol. 39, 1414–1425 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Chen, L. et al. Re-engineering the adenine deaminase TadA-8e for efficient and specific CRISPR-based cytosine base editing. Nat. Biotechnol. 41, 663–672 (2023).

    CAS  PubMed  Google Scholar 

  61. McPartland, J. C., Webb, S. J., Keehn, B. & Dawson, G. Patterns of visual attention to faces and objects in autism spectrum disorder. J. Autism Dev. Disord. 41, 148–157 (2011).

    PubMed  PubMed Central  Google Scholar 

  62. Weigelt, S., Koldewyn, K. & Kanwisher, N. Face identity recognition in autism spectrum disorders: a review of behavioral studies. Neurosci. Biobehav. Rev. 36, 1060–1084 (2012).

    PubMed  Google Scholar 

  63. Ferguson, J. N. et al. Social amnesia in mice lacking the oxytocin gene. Nat. Genet. 25, 284–288 (2000).

    CAS  PubMed  Google Scholar 

  64. Hörnberg, H. et al. Rescue of oxytocin response and social behaviour in a mouse model of autism. Nature 584, 252–256 (2020).

    PubMed  PubMed Central  Google Scholar 

  65. Dantzer, R., Bluthe,R. M., Koob, G. F. & Le Moal, M. Modulation of social memory in male rats by neurohypophyseal peptides. Psychopharmacology 91, 363–368 (1987).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank the ASD family for their participation in this study. We thank A. Gitler and members of the NPC for their critical comments on the manuscript. This work is supported by the Strategic Priority Research Program of the Chinese Academy of Sciences no. XDB32060202 (Z.-L.Q.), Fundamental Research Funds for the Central Universities (Z.-L.Q.), Natural Science Foundation of China grants to Z.-L.Q. (nos. 31625013, 82021001), T.-L.C. (nos. 32371144, 31600826), B.Y. (no. 32000726), and W.-J.H. (no. 82201632), Program of Shanghai Academic Research Leader no. 19XD1404300 (Z.-L.Q.), Shanghai Municipal Science and Technology Major Project (no. 2018SHZDZX05) (Z.-L.Q.), National Key R&D Program of China no. 2019YFA0111000 (T.-L.C.), Natural Science Foundation of Shanghai no. 20ZR1403100 (T.-L.C.), Shanghai Municipal Science and Technology no. 20JC1419500 (T.-L.C.), the Lingang Laboratory (no. LG-QS-202203-08) to T.-L.C, and Innovative Research Team of High-Level Local Universities in Shanghai (no. SHSMU-ZDCX20211100) (Z.-L.Q.). Z.-L.Q. is supported by the Guang Ci Professorship Program of Ruijin Hospital Shanghai Jiao Tong University School of Medicine.

Author information

Authors and Affiliations

Authors

Contributions

T.-L.C. and Z.-L.Q. conceptualized the study. S.-Q.Z., C.Z. and J.-L.C. performed the base editing and plasmid construction. W.-K.L. and S.-J.Z. carried out the biochemical experiments. W.-K.L. and S.-Q.Z. carried out the cell experiments. W.-K.L., Y.-T.Y., Z.-Y.X., S.-F.S. and Z.-F.C. performed the mouse immunohistochemistry and behavioral experiments. W.-K.L., B.-Q.X., S.-Q.Z., C.Z. and Y.-L.T. carried out the laser dissections. Y.-H.S. and W.-K.L. performed the electrophysiology experiments. S.-Q.Z. and T.-L.C. carried out the GUIDE-seq. W.-K.L., S.-Q.Z. and W.-J.H. prepared the NGS samples. B.Y. carried out the NGS data analysis. W.-K.L. and J.-C.W. carried out the analysis of the animal data. W.-L.P. and W.-K.L. carried out the viral injections and analysis. W.-K.L., T.-L.C. and Z.-L.Q. wrote the manuscript.

Corresponding authors

Correspondence to Tian-Lin Cheng or Zi-Long Qiu.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Neuroscience 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 Expression of MEF2C in the human brain and design of mouse Mef2c shRNA constructs.

(a) Annotation of the genomic and transcript information of L35P (based on the hg38 database). (b) Human brain transcriptomic data for MEF2C was obtained from the Human Brain Transcriptome (HBT), which ranges from embryonic development (periods 3-7) to postnatal development (periods 8–12) to adulthood (periods 13–15). Brain region abbreviations as follows, Hippocampus (HIP), neocortex (NCX), cerebellar cortex (CBC), mediodorsal nucleus of the thalamus (MD), striatum (STR), and amygdala (AMY). (c) Western blotting of MEF2C WT, L35F, L35P, L35I, or L35A expressed in HEK293 cells. (d) Quantifying relative MEF2C protein levels normalized to GAPDH (n = 7 per group). WT versus L35F: P < 0.0001; WT versus L35P: P < 0.0001; WT versus L35I: P = 0.0005; WT versus L35A: P = 0.0003, One-way ANOVA. (e) Schematic illustration of two Mef2c shRNA designed based on BLOCK-iT™ RNAi Designer (http://rnaidesigner.thermofisher.com/rnaiexpress/). (f) Quantifying Mef2c mRNA level of cultured primary cortical neurons infected with lentiviral vectors expressing scrambled or two Mef2c shRNA (P < 0.0001, n = 4 per group, One-way ANOVA). (g) Representative western blotting of cultured primary cortical neurons infected with lentiviral vectors expressing scrambled or Mef2c shRNA-1 for 72 h. Cells were harvested at DIV5. GAPDH was used as the internal control. (h) Quantitative analysis of relative Mef2c protein expression. Neurons infected with scrambled shRNA were normalized to 100% (P = 0.0327, n = 3, unpaired two-tailed Student’s t-test). n is biological repeat numbers of independent experiments. Statistical values represent the mean ± s.e.m. *P < 0.05, ***P < 0.001, ****P < 0.0001.

Source data

Extended Data Fig. 2 MEF2C-WT, but not MEF2C-L35P, could rescue impaired neuronal dendritic and axonal morphological development in Mef2c knockdown neurons.

(a) Representative image of immunofluorescent staining for GFP (green), MAP2 (red), and DAPI (blue) of cultured E14.5 mouse cortical neurons. Mouse cortical neurons were transfected with DsRed shRNA (Ctrl, n = 39), shMef2c (n = 36), shMef2c with human MEF2C-WT (n = 40) or MEF2C-L35P (n = 32) respectively at DIV1 via Lipofectamine 3000. Neuronal cells were immunofluorescent stained at DIV10. Scale bars, 60μm. Quantitative analysis of neuronal total dendritic length (b) and branch numbers (c) of neurons transfected with Ctrl, shMef2c, shMef2c + MEF2C-WT, and shMef2c + MEF2C-L35P. (b: all P < 0.0001; c: shMef2c+WT versus shMef2c + L35P: P = 0.0002) (d) Representative immunofluorescent staining images for GFP (green), MAP2 (red), and DAPI (blue) of cultured E14.5 mouse cortical neurons. Neuronal cells were transfected with DsRed shRNA (Ctrl, n = 34), shMef2c (n = 37), shMef2c with human MEF2C-WT (n = 33) or MEF2C-L35P (n = 34) respectively at DIV1. Neuronal cells were immunofluorescent stained at DIV4. Scale bars, 60μm. Quantitative analysis of neuronal axonal (e) and total neurite length (f) of neurons transfected with Ctrl, shMef2c, shMef2c + MEF2C-WT, and shMef2c + MEF2C-L35P. n is total neuron numbers collected from more than 3 independent experiments. (e, f: all P < 0.0001) Statistical values represent the mean ± s.e.m. ***P < 0.001, ****P < 0.0001, (b, c, e, f) One-way ANOVA.

Extended Data Fig. 3 Decreased protein expression and aberrant inhibitory interneuron development in Mef2c L35P+/- mice.

(a) Design of Mef2c L35P knock-in mice by CRISPR/Cas9 mediated homologous recombination. (b) Representative images of immunohistochemical staining for PV (red), Mef2c (green) and DAPI (blue) in the retrosplenial cortex (white dotted box indicates corresponding brain region) of Mef2c WT and L35P+/- mice. Scale bars, 200μm. Quantitative analysis of PV-positive neuronal cell density in the retrosplenial cortex (c), dentate gyrus (d), the somatosensory cortex (e), and visual cortex (f) of Mef2c WT (n = 11 brain slices from 4 mice) and L35P+/- mice (n = 12 brain slices from 4 mice). (c: P < 0.0001; d: P < 0.0001; e: P = 0.0004; f: P < 0.0001) (g) Representative images of immunohistochemical staining for SST (red), Mef2c (green), and DAPI (blue) in the retrosplenial cortex (White dotted box indicates corresponding brain region) of Mef2c WT and L35P+/- mice. Scale bars, 200μm. Quantification of SST-positive neuronal cell density in the retrosplenial cortex (h), hippocampus (i), and somatosensory cortex (j) of Mef2c WT (n = 8 brain slices from 3 mice) and L35P+/- mice (n = 7 brain slices from 3 mice). Statistical values represent the mean ± s.e.m. ***P < 0.001, **** P < 0.0001, (c, d, e, f, h, i, j) unpaired two-tailed Student’s t-test.

Extended Data Fig. 4 General development and cognitive ability of Mef2c L35P+/- mice.

(a) Body weights of Mef2c WT and L35P+/- mice of both sex from postnatal one week to nine weeks. (WT, n = 29 male mice, n = 17 female mice; L35P+/-, n = 17 male mice, n = 21 female mice). (b) Representative heat maps of Mef2c WT and L35P+/- mice in the novel object recognition task (the circle indicates a familiar object; the white box indicates a stranger object). Quantification of cumulative sniffing time of Mef2c WT (c) and L35P+/- mice (d) in the task (WT, n = 14; L35P+/-, n = 12; c: P = 0.0018; d: P = 0.0235). (e) Representative heat maps of Mef2c WT and L35P+/- mice in the elevated plus maze (open arm indicates top and bottom arms; closed arm indicates left and right arms). Quantification of cumulative duration exploring in the open arm (f) and closed arm (g) of Mef2c WT and L35P+/- mice (n = 14; f: P = 0.0003; g: P = 0.0006). (h) Representative locomotion track traces of Mef2c WT and L35P+/- mice in the light/dark exploration test. (i) Quantitative analysis of cumulative duration of Mef2c WT and L35P+/- mice in the light area (n = 8 per group). (j) Quantitative analysis of cumulative self-grooming and scratching time of Mef2c WT and L35P+/- mice (n = 14, j: P = 0.0004). (k) Quantitative analysis of primary latency spent locating the target hole of Mef2c WT and L35P+/- mice in the Barnes maze (n = 14 per group). Quantification of cumulative duration exploring in the target (l) and opposite zone (m) of Mef2c WT and L35P+/- mice in the apparatus (n = 14 per group). n represents mice number. Statistical values represent the mean ± s.e.m. *p < 0.05, **p < 0.01, ***p < 0.001, (a, k) Two-way ANOVA test, (c,d) paired two-tailed Student’s t-test, (f, g, i, j, l, m) unpaired two-tailed Student’s t-test.

Extended Data Fig. 5 Female Mef2c L35P+/- mice displayed abnormal social novelty ability and object recognition.

(a) Representative heat maps for locomotion of female Mef2c WT and L35P+/- mice on the social approach and social novelty session in the three-chamber test. Quantification of cumulative duration of interacting with a mouse and empty cage of Mef2c WT (b) and L35P+/- mice (c) (n = 8, b: P < 0.0001; c: P = 0.0024). (d) The preference index of social approach test for female Mef2c WT and L35P+/- mice. Quantification of cumulative duration of interacting with novel stranger mouse and familiar mouse of female Mef2c WT (e) and L35P+/- (f) mice (n = 8, e: P = 0.0002). (g) The preference index of social novelty test for female Mef2c WT and L35P+/- mice. (P = 0.046) (h) Representative heat maps of female Mef2c WT and L35P+/- mice in the novel object recognition task (the circle indicates a familiar object; the white box indicates a stranger object). Quantification of cumulative sniffing time of Mef2c WT (i) and L35P+/- mice (j) in the task (n = 8, i: P = 0.0095; j: P = 0.163). n represents mice number. Statistical values represent the mean ± s.e.m. *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001, (b, c, e, f, i, j) paired two-tailed Student’s t-test, (d, g) unpaired two-tailed Student’s t-test.

Extended Data Fig. 6 Female Mef2c L35P+/- mice exhibited hyperactive locomotion without anxiety behavior.

(a) Representative locomotion track traces of female Mef2c WT and L35P+/- mice in the open field test. Quantitative analysis of total distance (b), mean velocity (c), and duration in the center zone (d) of Mef2c WT and L35P + /- mice in the open field test (n = 8, b: P = 0.0054; c: P = 0.0054). (e) Representative locomotion track traces of female Mef2c WT and L35P+/- mice in the light/dark exploration test. (f) Quantitative analysis of cumulative duration of female Mef2c WT and L35P+/- mice in the light area (n = 8 per group). Representative heat maps of Mef2c WT and L35P+/- mice in the elevated plus maze (open arms are top and bottom arms; closed arms are left and right). (g) Quantification of cumulative duration exploring in the open arm (h) and closed arm (i) of female Mef2c WT and L35P+/- mice (n = 8, h: P = 0.0114; i: P = 0.038). (j) Representative locomotion track traces of female Mef2c WT and L35P+/- mice in the Barnes maze. (k) Quantitative analysis of primary latency spent locating the target hole of female Mef2c WT and L35P+/- mice in the Barnes maze (n = 8 per group). Quantification of cumulative duration exploring in the target (l) and opposite zone (m) of female Mef2c WT and L35P+/- mice in the apparatus (n = 8 per group). n represents mice number. Statistical values represent the mean ± s.e.m. * P < 0.05, ** P < 0.01, (b, c, d, f, h, i, l, m) unpaired two-tailed Student’s t-test, (k) Two-way ANOVA test.

Extended Data Fig. 7 Design of AeCBE.

(a) Illustration of various designs for classic CBE and APOBEC3A-embedding CBE with different embedding positions. (b) Screening for base editing efficiency of various APOBEC3A-embedding CBEs using sgRNA indicated in the HEK293 cells. (c) Representative FACS gating strategy for GFP, mCherry, and double positive samples in Fig. 3c. Quantitative analysis of base editing efficiency (C-to-T conversion) of bystander sites C103(d), C108(e), and C112(f) after different CBE applications in cultured mouse cortical neurons (n = 3 per group; n represents independent experiments, One-way ANOVA test). Statistical values represent the mean ± s.e.m.

Extended Data Fig. 8 Immunohistochemistry of SpCas9 and NeuN i in the Mef2c L35P+/- mice and evaluation of Off-targeting effects by AeCBE in vivo.

(a) Immunohistochemistry of SpCas9 and NeuN in the prefrontal cortex of AAV-AeCBE injected mice. (b) Quantification of SpCas9 and NeuN double-positive neurons in the total NeuN cells (n = 7 slices from 3 mice). (c) Immunostaining of SpCas9 and NeuN in the hippocampus of AAV-AeCBE injected mice. (d) Quantification of SpCas9 and NeuN double-positive neurons in the total NeuN cells (n = 12 slices from 4 mice). (e) Immunohistochemistry images of Mef2c and NeuN in the hippocampus of AAV-AeCBE injected mice. (f) Quantification of Mef2c and NeuN double-positive neurons in the total NeuN cells (n = 12 slices from 4 mice). (g) Off-target sites detected by GUIDE-seq, with sgRNA targeted to Mef2c L35P region. More reads indicated more likely to be targeted by SgC8 and SpG-Cas9. (h) Venn diagrams of off-target sites detected by the in-silico method and the GUIDE-seq method. (i) Off-target editing analysis at all 11 candidate sites by next-generation sequencing in vivo. (n = 3 mice, OT2-C7: P < 0.0001; OT4-C8: P = 0.0214; OT5-C8: P = 0.0243). Statistical values represent the mean ± s.e.m. * P < 0.05, *** P < 0.001, unpaired two-tailed Student’s t-test.

Extended Data Fig. 9 In vivo CBE gene editing did not restore decreased PV-positive inhibitory interneuron in the visual and somatosensory cortex of Mef2c L35P+/- mice.

(a) Representative images of immunohistochemical staining for PV (red) and DAPI (blue) in the Dentate Gyrus. (b) Quantification of PV-positive neurons in each group. WT-EGFP versus L35P-EGFP: P < 0.0001; L35P-EGFP: P = 0.0007 (c) Representative images of immunohistochemical staining for PV (red) and DAPI (blue) in the visual cortex and somatosensory cortex (White dotted box indicates corresponding brain region) of Mef2c WT mice injected with AAV-EGFP, Mef2c L35P+/- mice injected with either EGFP or AeCBE. Scale bars, 500μm. Quantitative analysis of PV-positive neuronal cell density in the visual (d) and somatosensory cortex (e) of three groups (n = 10 brain slices from 4 mice per group, d: P < 0.0001; e: P < 0.0001). Statistical values represent the mean ± s.e.m. *** P < 0.001, **** P < 0.0001, (b, d, e) unpaired two-tailed Student’s t-test.

Extended Data Fig. 10 No restoration of abnormal behaviors in Mef2c L35P+/- male mice injection with non-targeting sgRNA(sgNC) and AeCBE.

(a) Representative heat maps for locomotion of male Mef2c WT and L35P+/- mice injected with AAV-sgNC-AeCBE in the three-chamber test. Quantification of cumulative duration of Mef2c WT (b) and L35P+/- mice (c) (n = 11; b, c: P < 0.0001). (d) The preference index of social approach for Mef2c WT and L35P+/- mice (P = 0.462). Quantification of cumulative duration of Mef2c WT (e) and L35P+/- mice (f) (n = 11; e: P = 0.0003). (g) The preference index of social novelty for Mef2c WT and L35P+/- mice (P = 0.0134). Mean cumulative sniffing time (black curve) and individual mice (gray curve) in the social intruder test for Mef2c WT (h) and L35P+/- mice (i) with AAV-sgNC-AeCBE. (h: all P < 0.0001) (j) Cumulative sniffing time in the first trial of the social intruder test for Mef2c WT and L35P+/- mice (n = 11, P < 0.0001). (k) The social recognition index of the social intruder test for Mef2c WT and L35P+/- mice (n = 11, P < 0.0001). (l) Representative locomotion track traces of Mef2c WT and L35P+/- male mice in the light/dark exploration test. (m) Quantitative analysis of cumulative duration of Mef2c WT and L35P+/- mice in the light area (n = 9). (n) Quantitative analysis of cumulative time spent self-grooming and scratching of Mef2c WT and L35P+/- mice with AAV-sgNC-AeCBE. (n = 11, P = 0.0005). (o) Representative locomotion track traces of Mef2c WT and L35P+/- mice in the open field test. Quantitative analysis of duration in the center zone (p), mean velocity (q), and total distance (r) in the open field test (n = 11, q: P = 0.0112; r: P = 0.0099). n represents mouse number. Statistical values represent the mean ± s.e.m. (b, c, e, f) paired two-tailed Student’s t-test, (d, g, j, k, m, n, p, q, r) unpaired two-tailed Student’s t-test, (h, i) One-way ANOVA test.

Supplementary information

Source data

Source Data Fig. 1

Unprocessed western blots.

Source Data Fig. 2

Unprocessed western blots.

Source Data Fig. 4

Unprocessed western blots.

Source Data Extended Data Fig. 1

Unprocessed western blots.

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

Li, WK., Zhang, SQ., Peng, WL. et al. Whole-brain in vivo base editing reverses behavioral changes in Mef2c-mutant mice. Nat Neurosci 27, 116–128 (2024). https://doi.org/10.1038/s41593-023-01499-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-023-01499-x

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