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:

A phenotypic screening platform for identifying chemical modulators of astrocyte reactivity

Abstract

Disease, injury and aging induce pathological reactive astrocyte states that contribute to neurodegeneration. Modulating reactive astrocytes therefore represent an attractive therapeutic strategy. Here we describe the development of an astrocyte phenotypic screening platform for identifying chemical modulators of astrocyte reactivity. Leveraging this platform for chemical screening, we identify histone deacetylase 3 (HDAC3) inhibitors as effective suppressors of pathological astrocyte reactivity. We demonstrate that HDAC3 inhibition reduces molecular and functional characteristics of reactive astrocytes in vitro. Transcriptional and chromatin mapping studies show that HDAC3 inhibition disarms pathological astrocyte gene expression and function while promoting the expression of genes associated with beneficial astrocytes. Administration of RGFP966, a small molecule HDAC3 inhibitor, blocks reactive astrocyte formation and promotes neuroprotection in vivo in mice. Collectively, these results establish a platform for discovering modulators of reactive astrocyte states, inform the mechanisms that control astrocyte reactivity and demonstrate the therapeutic benefits of modulating astrocyte reactivity for neurodegenerative diseases.

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 reactive astrocyte cellular platform.
Fig. 2: Extensive chromatin remodeling during astrocyte reactivity.
Fig. 3: Phenotypic screen identifies HDAC3 as a regulator of pathological reactive astrocytes.
Fig. 4: HDAC3 mediates a switch between pro-inflammatory and anti-inflammatory gene expression by decreasing RelA/p65 signaling.
Fig. 5: HDAC3 inhibition blocks reactive astrocyte formation in vivo and promotes neuroprotection.

Similar content being viewed by others

Data availability

All sequencing datasets generated in this study have been deposited in GEO (https://www.ncbi.nlm.nih.gov/geo/) under SuperSeries accession code GSE185215 with subseries for astrocyte RNA-seq, ONC RNA-seq, scRNA-seq, CUT&RUN and ATAC-seq. Previously published data for ref. 46 (PMID: 34413515) is available under GEO accession GSE148612. Previously published data for ref. 22 (PMID: 31768052) is available under accession GSE138852. Previously published data for ref. 23 (PMID: 32070434) is available under GEO accession GSE180928. Previously published data for ref. 24 (PMID: 34919646) is available under GSE157783. Source data are provided with this paper.

Code availability

Data analysis was performed with publicly available packages by following available tutorials and vignettes.

References

  1. Hasel, P. & Liddelow, S. A. Astrocytes. Curr. Biol. 31, R326–R327 (2021).

    Article  CAS  PubMed  Google Scholar 

  2. Allen, N. J. & Eroglu, C. Cell biology of astrocyte–synapse interactions. Neuron 96, 697–708 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Nagai, J. et al. Behaviorally consequential astrocytic regulation of neural circuits. Neuron 109, 576–596 (2021).

    Article  CAS  PubMed  Google Scholar 

  4. Guttenplan, K. A. et al. Neurotoxic reactive astrocytes induce cell death via saturated lipids. Nature 599, 102–107 (2021).

    Article  CAS  PubMed  Google Scholar 

  5. Di Giorgio, F. P., Carrasco, M. A., Siao, M. C., Maniatis, T. & Eggan, K. Non-cell autonomous effect of glia on motor neurons in an embryonic stem cell-based ALS model. Nat. Neurosci. 10, 608–614 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  6. Ilieva, H., Polymenidou, M. & Cleveland, D. W. Non-cell autonomous toxicity in neurodegenerative disorders: ALS and beyond. J. Cell Biol. 187, 761–772 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Meyer, K. et al. Direct conversion of patient fibroblasts demonstrates non-cell autonomous toxicity of astrocytes to motor neurons in familial and sporadic ALS. Proc. Natl Acad. Sci. USA 111, 829–832 (2014).

    Article  CAS  PubMed  Google Scholar 

  8. Clarke, L. E. et al. Normal aging induces A1-like astrocyte reactivity. Proc. Natl Acad. Sci. USA 115, E1896–E1905 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Liddelow, S. A. et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature 541, 481–487 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Wheeler, M. A. et al. MAFG-driven astrocytes promote CNS inflammation. Nature 578, 593–599 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Smith, H. L. et al. Astrocyte unfolded protein response induces a specific reactivity state that causes non-cell-autonomous neuronal degeneration. Neuron 105, 855–866 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Burda, J. E. et al. Divergent transcriptional regulation of astrocyte reactivity across disorders. Nature 606, 557–564 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Dooves, S. et al. Astrocytes are central in the pathomechanisms of vanishing white matter. J. Clin. Invest. 126, 1512–1524 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Absinta, M. et al. A lymphocyte-microglia-astrocyte axis in chronic active multiple sclerosis. Nature 597, 709–714 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Leng, K. et al. CRISPRi screens in human astrocytes elucidate regulators of distinct inflammatory reactive states. Nat. Neurosci. 25, 1528–1542 (2022).

  16. Yun, S. P. et al. Block of A1 astrocyte conversion by microglia is neuroprotective in models of Parkinson’s disease. Nat. Med. 24, 931–938 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Najm, F. J. et al. Drug-based modulation of endogenous stem cells promotes functional remyelination in vivo. Nature 522, 216–220 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Foo, L. C. et al. Development of a method for the purification and culture of rodent astrocytes. Neuron 71, 799–811 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Pike, S. C., Welsh, N., Linzey, M. & Gilli, F. Theiler’s virus-induced demyelinating disease as an infectious model of progressive multiple sclerosis. Front. Mol. Neurosci. 15, 1019799 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Chastain, E. M., Duncan, D. S., Rodgers, J. M. & Miller, S. D. The role of antigen presenting cells in multiple sclerosis. Biochim. Biophys. Acta 1812, 265–274 (2011).

    Article  CAS  PubMed  Google Scholar 

  21. Chung, W. S. et al. Astrocytes mediate synapse elimination through MEGF10 and MERTK pathways. Nature 504, 394–400 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Grubman, A. et al. A single-cell atlas of entorhinal cortex from individuals with Alzheimer’s disease reveals cell-type-specific gene expression regulation. Nat. Neurosci. 22, 2087–2097 (2019).

    Article  CAS  PubMed  Google Scholar 

  23. Al-Dalahmah, O. et al. Single-nucleus RNA-seq identifies Huntington disease astrocyte states. Acta Neuropathol. Commun. 8, 19 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Smajic, S. et al. Single-cell sequencing of human midbrain reveals glial activation and a Parkinson-specific neuronal state. Brain 145, 964–978 (2022).

    Article  PubMed  Google Scholar 

  25. Loven, J. et al. Selective inhibition of tumor oncogenes by disruption of super-enhancers. Cell 153, 320–334 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Whyte, W. A. et al. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell 153, 307–319 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Hnisz, D. et al. Super-enhancers in the control of cell identity and disease. Cell 155, 934–947 (2013).

    Article  CAS  PubMed  Google Scholar 

  28. Molofsky, A. V. & Deneen, B. Astrocyte development: a guide for the perplexed. Glia 63, 1320–1329 (2015).

    Article  PubMed  Google Scholar 

  29. Hartmann, K. et al. Complement 3+-astrocytes are highly abundant in prion diseases, but their abolishment led to an accelerated disease course and early dysregulation of microglia. Acta Neuropathol. Commun. 7, 83 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  30. Malvaez, M. et al. HDAC3-selective inhibitor enhances extinction of cocaine-seeking behavior in a persistent manner. Proc. Natl Acad. Sci. USA 110, 2647–2652 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Suzuki, T. et al. Identification of highly selective and potent histone deacetylase 3 inhibitors using click chemistry-based combinatorial fragment assembly. PLoS ONE 8, e68669 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Srinivasan, R. et al. New transgenic mouse lines for selectively targeting astrocytes and studying calcium signals in astrocyte processes in situ and in vivo. Neuron 92, 1181–1195 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. McQuown, S. C. et al. HDAC3 is a critical negative regulator of long-term memory formation. J. Neurosci. 31, 764–774 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Gryder, B. E. et al. Histone hyperacetylation disrupts core gene regulatory architecture in rhabdomyosarcoma. Nat. Genet. 51, 1714–1722 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Leus, N. G., Zwinderman, M. R. & Dekker, F. J. Histone deacetylase 3 (HDAC 3) as emerging drug target in NF-kappaB-mediated inflammation. Curr. Opin. Chem. Biol. 33, 160–168 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Chen, L. F. & Greene, W. C. Regulation of distinct biological activities of the NF-kappaB transcription factor complex by acetylation. J. Mol. Med. (Berl.) 81, 549–557 (2003).

    Article  CAS  PubMed  Google Scholar 

  37. Shin, H. M. et al. Inhibitory action of novel aromatic diamine compound on lipopolysaccharide-induced nuclear translocation of NF-κB without affecting IκB degradation. FEBS Lett. 571, 50–54 (2004).

    Article  CAS  PubMed  Google Scholar 

  38. Tonelli, C., Chio, I. I. C. & Tuveson, D. A. Transcriptional regulation by Nrf2. Antioxid. Redox Signal. 29, 1727–1745 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Vargas, M. R., Johnson, D. A., Sirkis, D. W., Messing, A. & Johnson, J. A. Nrf2 activation in astrocytes protects against neurodegeneration in mouse models of familial amyotrophic lateral sclerosis. J. Neurosci. 28, 13574–13581 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Calkins, M. J., Vargas, M. R., Johnson, D. A. & Johnson, J. A. Astrocyte-specific overexpression of Nrf2 protects striatal neurons from mitochondrial complex II inhibition. Toxicol. Sci. 115, 557–568 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Gan, L., Vargas, M. R., Johnson, D. A. & Johnson, J. A. Astrocyte-specific overexpression of Nrf2 delays motor pathology and synuclein aggregation throughout the CNS in the α-synuclein mutant (A53T) mouse model. J. Neurosci. 32, 17775–17787 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Sigfridsson, E. et al. Astrocyte-specific overexpression of Nrf2 protects against optic tract damage and behavioural alterations in a mouse model of cerebral hypoperfusion. Sci. Rep. 8, 12552 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Zhao, W. et al. Astrocytic Nrf2 expression protects spinal cord from oxidative stress following spinal cord injury in a male mouse model. J. Neuroinflammation 19, 134 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Draheim, T. et al. Activation of the astrocytic Nrf2/ARE system ameliorates the formation of demyelinating lesions in a multiple sclerosis animal model. Glia 64, 2219–2230 (2016).

    Article  CAS  PubMed  Google Scholar 

  45. Bowers, M. E., Xia, B., Carreiro, S. & Ressler, K. J. The class I HDAC inhibitor RGFP963 enhances consolidation of cued fear extinction. Learn. Mem. 22, 225–231 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Hasel, P., Rose, I. V. L., Sadick, J. S., Kim, R. D. & Liddelow, S. A. Neuroinflammatory astrocyte subtypes in the mouse brain. Nat. Neurosci. 24, 1475–1487 (2021).

    Article  CAS  PubMed  Google Scholar 

  47. Zamanian, J. L. et al. Genomic analysis of reactive astrogliosis. J. Neurosci. 32, 6391–6410 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Sun, L. et al. Loss of HDAC11 ameliorates clinical symptoms in a multiple sclerosis mouse model. Life Sci. Alliance 1, e201800039 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Zhang, S., Fujita, Y., Matsuzaki, R. & Yamashita, T. Class I histone deacetylase (HDAC) inhibitor CI-994 promotes functional recovery following spinal cord injury. Cell Death Dis. 9, 460 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Kuboyama, T. et al. HDAC3 inhibition ameliorates spinal cord injury by immunomodulation. Sci. Rep. 7, 8641 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Janczura, K. J. et al. Inhibition of HDAC3 reverses Alzheimer’s disease-related pathologies in vitro and in the 3xTg-AD mouse model. Proc. Natl Acad. Sci. USA 115, E11148–E11157 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Zhang, M. J. et al. The HDAC3 inhibitor RGFP966 ameliorated ischemic brain damage by downregulating the AIM2 inflammasome. FASEB J. 34, 648–662 (2020).

    Article  CAS  PubMed  Google Scholar 

  53. Smith, C. A., Vianna, J. R. & Chauhan, B. C. Assessing retinal ganglion cell damage. Eye (Lond.) 31, 209–217 (2017).

    Article  CAS  PubMed  Google Scholar 

  54. Guttenplan, K. A. et al. Neurotoxic reactive astrocytes drive neuronal death after retinal injury. Cell Rep. 31, 107776 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Kanamori, A., Naka, M., Fukuda, M., Nakamura, M. & Negi, A. Latanoprost protects rat retinal ganglion cells from apoptosis in vitro and in vivo. Exp. Eye Res. 88, 535–541 (2009).

    Article  CAS  PubMed  Google Scholar 

  56. Hafemeister, C. & Satija, R. Normalization and variance stabilization of single-cell RNA-seq data using regularized negative binomial regression. Genome Biol. 20, 296 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Barbar, L., Rusielewicz, T., Zimmer, M., Kalpana, K. & Fossati, V. Isolation of human CD49f+ astrocytes and in vitro iPSC-based neurotoxicity assays. STAR Protoc. 1, 100172 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Rolfe, A. J., Bosco, D. B., Broussard, E. N. & Ren, Y. In vitro phagocytosis of myelin debris by bone marrow-derived macrophages. J. Vis. Exp. 30, 56322 (2017).

    Google Scholar 

  59. Lager, A. M. et al. Rapid functional genetics of the oligodendrocyte lineage using pluripotent stem cells. Nat. Commun. 9, 3708 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  60. Xie, Z. et al. Gene set knowledge discovery with Enrichr. Curr. Protoc. 1, e90 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Kuleshov, M. V. et al. Enrichr: a comprehensive gene set enrichment analysis web server 2016 update. Nucleic Acids Res. 44, W90–W97 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Chen, E. Y. et al. Enrichr: interactive and collaborative HTML5 gene list enrichment analysis tool. BMC Bioinformatics 14, 128 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  63. Subramanian, A. et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl Acad. Sci. USA 102, 15545–15550 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Michiels, T., Dejong, V., Rodrigus, R. & Shaw-Jackson, C. Protein 2A is not required for Theiler’s virus replication. J. Virol. 71, 9549–9556 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Yu, F., Sankaran, V. G. & Yuan, G. C. CUT&RUNTools 2.0: a pipeline for single-cell and bulk-level CUT&RUN and CUT&Tag data analysis. Bioinformatics 38, 252–254 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  66. Yu, G., Wang, L. G. & He, Q. Y. ChIPseeker: an R/Bioconductor package for ChIP peak annotation, comparison and visualization. Bioinformatics 31, 2382–2383 (2015).

    Article  CAS  PubMed  Google Scholar 

  67. Zhu, L. J. et al. ChIPpeakAnno: a Bioconductor package to annotate ChIP-seq and ChIP-chip data. BMC Bioinformatics 11, 237 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  68. Lawrence, M. et al. Software for computing and annotating genomic ranges. PLoS Comput. Biol. 9, e1003118 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Heinz, S. et al. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol. Cell 38, 576–589 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Perriot, S. et al. Human induced pluripotent stem cell-derived astrocytes are differentially activated by multiple sclerosis-associated cytokines. Stem Cell Rep. 11, 1199–1210 (2018).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This study was supported by grants from the National Multiple Sclerosis Society (TA-2105-37619 to B.L.L.C.), the National Institutes of Health (R35NS116842 to P.J.T., F30HD096784 and T32GM007250 to K.C.A, K08EY029362 to A.R.S, and F31NS124282, T32NS077888 and T32GM007250 to E.F.C.), the Hartwell Foundation (to B.L.L.C.), as well as institutional support from the Case Western Reserve University (CWRU) School of Medicine and philanthropic support from sTF5 Care and the R. Blane & Claudia Walter, Long, Goodman, Geller and Weidenthal families and the Research Institute for Children’s Health. Additional support was provided by the Small Molecule Drug Development and Genomics core facilities of the CWRU Comprehensive Cancer Center (P30CA043703), the CWRU Light Microscopy Imaging Center (Shared Instrumentation Grant NIH ORIP S10OD024981) and the University of Chicago Genomics Facility (P30CA014599). We thank D. Adams, Y. Fedorov, W. Harte, M. Scavuzzo, M. Madhavan, D. Kassel and M. Elitt for technical assistance and/or discussion and C. Lilliehook for editorial support.

Author information

Authors and Affiliations

Authors

Contributions

B.L.L.C. and P.J.T. conceived and managed the overall study. B.L.L.C., J.D.K. and A.M.S. performed, quantified and analyzed in vitro studies. H.E.S., A.M.S. and A.K. helped to generate astrocyte cultures. E.C.F. generated TMEV virus. B.L.L.C. performed and analyzed all scRNA-seq experiments. B.L.L.C. performed the small-molecule screen and validations. K.C.A. performed SE analysis. E.F.C. performed CUT&RUN and ATAC-seq. B.L.L.C. performed CUT&RUN, ATAC-seq and bulk RNA-seq data analysis. B.L.L.C., A.M.S. and A.K. performed and analyzed in vivo LPS studies. M.K., E.G. and R.H.M. performed in vivo LPC studies and generated EM images. A.D.J., A.R.S., J.A.S. and B.M.S. performed and analyzed ONC studies. Y.M.-H. analyzed EM images. B.L.L.C. assembled all figures. B.L.L.C. and P.J.T. wrote the paper with input from all authors.

Corresponding authors

Correspondence to Benjamin L. L. Clayton or Paul J. Tesar.

Ethics declarations

Competing interests

B.L.L.C. and P.J.T. are listed as inventors on pending patent claims filed by Case Western Reserve University covering methods and compositions for treating neurodegenerative disorders. The other 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 Purity and function of astrocyte cultures.

a, Overview diagram of astrocyte isolation and enrichment protocol (images generated by BioRender). b, Phase contrast image showing prototypical astrocyte morphology from a single astrocyte culture. Scale bar is 50 µm. Immunofluorescence images showing expression of canonical astrocyte markers AQP4, GLT-1 (SLC1A2) and ALDH1L1 in a single astrocyte culture. Scale bar is 100 µm. c, Bulk RNA-seq data showing high expression of astrocyte marker genes and no expression of marker genes for other CNS cell types. Data are from 3 biological replicates (independent astrocyte isolations). d, Glutamate uptake by physiological astrocytes in culture. Data presented as the mean ± s.e.m. for technical replicates (open circles) from 3 biological replicates (independent astrocyte isolations). eh, ATAC-seq tracks at marker genes for CNS cell types. Chromatin is only open at astrocyte genes. i, Bulk RNA-seq data showing decreased expression of Meg10 and Mertk in agreement with decreased phagocytosis of myelin debris by reactive astrocytes. Data presented as mean ± s.e.m. for n = 3 biological replicates. P-value generated by paired two-tailed t-test. j, The log2 fold-change (log2(FC)) of secreted cytokines in reactive vs physiological astrocytes conditioned media. Data presented as mean ± s.e.m for n = 3 biological replicates.

Source data

Extended Data Fig. 2 In vitro reactive astrocytes correspond to in vivo counterparts.

a, Enrichment map of gene ontology terms for genes upregulated in reactive vs physiological astrocytes. b, Enrichment map of gene ontology terms for genes upregulated in physiological vs reactive astrocytes. c, Representative images of physiological astrocytes, astrocytes exposed to TIC cytokines and astrocytes infected with TMEV and stained for double stranded RNA to show viral infection (dsRNA in green) and GBP2 (red). d, Quantification of the percentage of astrocytes that are positive for the viral marker dsRNA. Data presented as mean ± s.e.m for n = 3 biological replicates, with p-value calculated by a one-way ANOVA. e, Quantification of the percentage of astrocytes that are GBP2 positive. Data presented as mean ± s.e.m for n = 3 biological replicates, with p-value calculated by a one-way ANOVA with Dunnett correction for multiple comparisons. f, Quantitative PCR results comparing the expression of pathological reactive astrocyte markers Gbp2 and Psmb8 by physiological astrocytes, astrocytes exposed to TIC cytokines and astrocytes infected with TMEV. P-value calculated by one-way ANOVA with Dunnett correction for multiple comparisons. g, UMAP plots from integrated single-cell RNA-seq analysis of in vitro physiological and reactive astrocytes from this study with in vivo astrocytes from LPS or vehicle treated mice46. h, Frequency of cells distributed across different astrocyte clusters. i, Expression of pathological reactive astrocyte genes and astrocyte marker genes in each single-cell cluster. j, Gene ontology results showing enriched terms for genes enriched in clusters 2, 6 and 7 that contain mainly astrocytes from reactive cultures and LPS treated mice. P-values generated by Benjamini-Hochberg false discovery rate.

Source data

Extended Data Fig. 3 Epigenomics of astrocyte reactivity.

a, Biological (Bio) replicate ATAC-seq tracks for Fig. 2c. b, Depiction of gained, lost, and shared H3K27ac or super-enhancer peaks during the transition from physiological to reactive astrocytes. c, Biological replicate H3K27ac CUT&RUN tracks in Fig. 2i. d, Gene ontology analysis of genes targeted by gained super-enhancers in reactive astrocytes that are also upregulated in reactive astrocytes compared to physiological in bulk RNA-seq analysis. P-values generated by Benjamini-Hochberg false discovery rate. e, Bulk RNA-seq volcano plot of genes with a gained H3K27ac CUT&RUN peak (+/−5 kb of the transcription start site (TSS)) in reactive astrocytes. log2(FC) and p-adj values were generated from bulk RNA-seq analysis with DESEQ2 f, Bulk RNA-seq volcano plot of genes with a shared H3K27ac CUT&RUN peak (+/−5 kb of the transcription start site (TSS)) in both reactive and physiological astrocytes. log2(FC) and p-adj values were generated from bulk RNA-seq analysis with DESEQ2 g, Bulk RNA-seq volcano plot of genes with a lost H3K27ac CUT&RUN peak (+/−5 kb of the transcription start site (TSS)) in reactive astrocytes. log2(FC) and p-adj values were generated from bulk RNA-seq analysis with DESEQ2. h, Gene ontology analysis of genes targeted by gained H3K27ac CUT&RUN peak in reactive astrocytes that are also upregulated in reactive astrocytes compared to physiological in bulk RNA-seq analysis. P-values generated by Benjamini-Hochberg false discovery rate.

Source data

Extended Data Fig. 4 High-throughput chemical screen quality control and validation of HDAC3.

a, Example images of DMSO vehicle-treated reactive and physiological control wells from the phenotypic screen. Scale bar is 100 µm. b, Z′ standard and robust scores for primary screen plates. c, Percent GBP2 positive astrocytes in DMSO vehicle treated reactive and physiological control wells on each primary screen plate. Data as the mean ± s.e.m., n= 16 wells per group on each primary screen plate. d, Dose curve of primary screen hits. Data are percent of GBP2 positive cells normalized to DMSO vehicle treated reactive astrocyte control wells. n = 2 biological replicates (independent astrocyte isolations). Black data points represent toxic doses where total cell number decreased by >50% compared to vehicle treated reactive astrocyte control wells. e, Dose curve analysis of hits from primary screen with Psmb8 positivity by in situ hybridization as a secondary endpoint. Data are percent of Psmb8 mRNA positive normalized to DMSO vehicle treated reactive astrocyte control wells with an n = 1 biological replicate (independent astrocyte isolation). Black data points represent toxic doses where total cell number decreased by >50% compared to vehicle treated reactive astrocyte control wells. f, Ranked inhibition against each HDAC isozyme for validated HDAC inhibitors in the primary screen. Highlighted is HDAC as the only shared target between all HDAC inhibitor hits. Ranked efficiency was pulled from target data provided by Selleck Chemical. g, Dose curve and IC50 value for the HDAC3 specific inhibitor RGFP966 to block astrocyte reactivity with an n = 5 biological replicates. h, Dose curve and IC50 value for the HDAC3 specific inhibitor T247 to block astrocyte reactivity with an n = 2 biological replicates. ik, Representative images and quantification of wild-type (WT) and HDAC3 knockout (KO) astrocyte cultures exposed to TIC cytokines. Scale bar is 100 µm. Data are mean ± s.e.m., n= 3 independent experiments, p-value by a paired t-test. l, GBP2 and PSMB8 qPCR results for human iPSC derived physiological or reactive astrocyte cultures treated with vehicle or 5 μM RGFP966. Data are mean ± s.e.m., n= 4 technical replicates.

Source data

Extended Data Fig. 5 Histone acetylation does not predict HDAC3 inhibition induced gene expression changes in reactive astrocytes.

a, Tukey box and whisker plot showing the average log2 fold-change (log2(FC)), from bulk RNA-seq, of the top 100 upregulated genes targeted by gained super-enhancers, all genes targeted by shared super-enhancers, and the top 100 downregulated genes targeted by lost super-enhancers in pathological reactive astrocytes. Data are presented for n = 3 biological replicates. Two-tailed p-value is generated with a one-sample Wilcoxon signed ranked test comparing to a hypothetical median of log2(FC) = 0 which would designate no difference in expression between reactive and physiological astrocytes. b, Quantification of CCL5 ELISAs performed on astrocyte conditioned media. Data presented as mean ± s.e.m for an n = 3 biological replicates with p-values compared to reactive plus vehicle control and calculated by one-way ANOVA with Dunnett correction for multiple comparisons. c, Tukey box and whisker plot showing the level of H3K27ac, from CUT&RUN, at gained, shared and lost super-enhancer in reactive astrocytes treated with RGFP966. Data are presented for n = 3 biological replicates. Two-tailed p-value is generated with a one-sample Wilcoxon signed ranked test comparing to a hypothetical median of log2(FC) = 0. d,e, Representative images and quantification of physiological astrocytes treated with vehicle, 5 μM RGFP966, or 5 μM JSH-23 and then stained for acetyl-RelA/p65 (K310). Scale bar is 100 µm. Data are mean ± s.e.m, n = 3 biological replicates, p-value by one-way ANOVA with Dunnett multiple comparison correction. f, Uncropped western blots for Fig. 3g. g, Transcription factor motifs enriched in H3K27ac CUT&RUN peaks from vehicle treated reactive astrocytes versus RGFP966 (RGFP) treated reactive astrocytes. P-values generated by HOMER. h, Transcription factor motifs enriched in H3K27ac CUT&RUN peaks from RGFP treated reactive astrocytes versus vehicle treated reactive astrocytes. P-values generated by HOMER. i, Normalized NF-κB luciferase activity in Jurkat reporter cells treated with the validated hits from the primary drug screen. Data presented as percentage of NF-κB activity from a single independent experiment.

Source data

Extended Data Fig. 6 HDAC3 inhibition modulates pro- and anti-inflammatory gene expression in reactive astrocytes.

a, Biological (Bio) replicate tracks for RelA/p65 CUT&RUN in Fig. 4k. b, Enriched gene ontology terms for genes significantly upregulated in reactive astrocytes treated with RGFP. P-values generated by Benjamini-Hochberg false discovery rate. c, Enriched transcription factors targeting genes significantly upregulated in reactive astrocytes treated with RGFP. d, Biological replicate tracks for H3K27ac CUT&RUN in Fig. 4p.

Source data

Extended Data Fig. 7 In vivo pharmacology of RGFP966 and in vivo genetic validation of HDAC3 as a mediator of astrocyte reactivity.

a, Brain concentration of RGFP966 (RGFP) 4 hrs after treatment with vehicle (Veh) or 10 mg/kg RGFP. Data presented for n = 2 biological replicates (mice). Concentration of RGFP in brain from vehicle treated mice was below quantifiable levels (BQL). bd, Representative images and quantification of immunohistochemistry for AcH4 in the cortex of mice treated with vehicle or 10 mg/kg RGFP and then exposed to systemic LPS injections to induce neuroinflammation. Scale bar is 100 µm. Data are mean ± s.e.m., n = 4 biological replicates (mice), p-value by unpaired two-tailed t-test e,f, Representative in situ hybridization images and quantification of untreated and LPS-exposed mice probed for the pan-reactive astrocyte marker Gfap (blue) and the reactive astrocyte marker Gbp2 (red) in the corpus callosum. Scale bar is 50 µm. Data are mean ± s.e.m., n = 4 biological replicates (mice), p-value by unpaired two-tailed t-test. gm, Representative images and quantification of immunohistochemistry for GFAP (red) and IBA-1 (blue) in the cortex of mice treated with vehicle or 10 mg/kg RGFP and exposed to systemic LPS or saline vehicle. Scale bar is 100 µm. Data are mean ± s.e.m., n = 3 or 4 biological replicates (mice), p-value calculated by one-way ANOVA and Tukey multiple comparison correction. n, Diagram of astrocyte specific HDAC3 knockout mouse breeding. oq, Representative in situ hybridization images and quantification of untreated wild-type (WT) and HDAC3 knockout (KO) mice and then probed for the pan-astrocyte marker Slc1a3 (green) and the reactive astrocyte marker C3 (red). Scale bar is 50 µm. Data are mean ± s.e.m., n = 3 biological replicates (mice), p-value by unpaired two-tailed t-test. ru, Representative in situ hybridization images and quantification of wild-type and HDAC3 knockout (KO) mice exposed to systemic LPS and then probed for the pan-astrocyte marker Slc1a3 (green) and the reactive astrocyte marker C3 (red). Scale bar is 50 µm. Data are mean ± s.e.m., n = 4 biological replicates (mice), p-value by unpaired two-tailed t-test.

Source data

Extended Data Fig. 8 RGFP966 has no effect on generalized gliosis in the toxin-based injury model LPC.

a,b, Representative in situ hybridization images and quantification of naïve and LPC lesioned mice probed for the pan-astrocyte marker Slc1a3 (blue) and the reactive astrocyte marker C3 (red). Scale bar is 50 µm. Data are mean ± s.e.m., n = 3 or 4 biological replicates (mice), p-value by unpaired two-tailed t-test. c,d, Representative images and quantification of LPC lesions from vehicle or RGFP966 (RGFP) treated mice stained for IBA-1. Scale bar is 100 µm. Data are mean ± s.e.m., n = 4 biological replicates (mice). eg, Representative images and quantification of LPC lesions from RGFP966 or vehicle treated mice stained for IBA-1 (orange), CD86 (blue) and CD206 (green). Scale bar is 100 µm. Data are mean ± s.e.m., n = 4 or 6 biological replicates (mice). h,i, Representative images, and quantification of LPC lesion from vehicle or RGFP treated mice stained for GFAP. Scale bar is 100 µm. Data are mean ± s.e.m., n = 4 biological replicates (mice).

Source data

Extended Data Fig. 9 RGFP decreases gene expression associated with reactive astrocytes in ONC.

a, Bulk RNA-seq volcano plot of genes up and downregulated in ONC retina versus naïve retina. log2(FC) and p-adj values were generated from bulk RNA-seq analysis with DESEQ2. b, Gene ontology terms for enriched for genes upregulated in ONC retina versus naïve retina. P-values generated by Benjamini-Hochberg false discovery rate. c, Tukey box and whisker plot depicting how the expression of genes upregulated in ONC retina change between crushed and naïve retina in mice treated with vehicle and mice treated with 10 mg/kg RGFP. Data are presented as log2(FC) for n = 7 biological replicates. P-value generated by a paired two-tailed t-test. d, Tukey box and whisker plot depicting the log2(FC) expression of ONC reactive astrocyte genes (same as in Fig. 5k) between naïve retina treated with 10 mg/kg RGFP and naïve retina treated with vehicle. Data are presented as log2(FC) for n = 7 biological replicates. P-value generated with a one-sample Wilcoxon signed ranked test comparing to a hypothetical median of log2(FC) = 0, which would designate no difference in expression. e,f, Representative images and quantification of retinal ganglion cells (RGCs) double stained with BRN3A (green) and beta-3 tubulin (red) in retina from naive mice treated with vehicle or 10 mg/kg RGFP966. Scale bar is 20 µm. Data are mean ± s.e.m., n = 9 biological replicates (mice), p-value by unpaired two-tailed t-test.

Source data

Supplementary information

Reporting Summary

Supplementary Table 1

Average expression of physiological and reactive astrocytes from scRNA-seq.

Supplementary Table 2

Differential gene expression by cluster in scRNA-seq integrated with publicly available data.

Supplementary Table 3

Gene ontology enrichment analysis.

Supplementary Table 4

log2(FC) for genes targeted by gained ATAC peak in reactive astrocytes.

Supplementary Table 5

Transcription factor motif analysis.

Supplementary Table 6

log2(FC) of gained, lost and shared SE targeted genes in reactive versus physiological astrocytes.

Supplementary Table 7

Primary chemical screen results.

Supplementary Table 8

log2(FC) for RelA/p65 target genes.

Supplementary Table 9

log2(FC) for ONC retina versus naïve retina.

Source data

Source Data Fig. 1

Source data for graphs.

Source Data Fig. 2

Source data for graphs.

Source Data Fig. 3

Source data for graphs.

Source Data Fig. 4

Source data for graphs.

Source Data Fig. 5

Source data for graphs.

Source Data Extended Data Fig. 1

Source data for graphs.

Source Data Extended Data Fig. 2

Source data for graphs.

Source Data Extended Data Fig. 3

Source data for graphs.

Source Data Extended Data Fig. 4

Source data for graphs.

Source Data Extended Data Fig. 5

Source data for graphs.

Source Data Extended Data Fig. 6

Source data for graphs.

Source Data Extended Data Fig. 7

Source data for graphs.

Source Data Extended Data Fig. 8

Source data for graphs.

Source Data Extended Data Fig. 9

Source data for graphs.

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

Clayton, B.L.L., Kristell, J.D., Allan, K.C. et al. A phenotypic screening platform for identifying chemical modulators of astrocyte reactivity. Nat Neurosci 27, 656–665 (2024). https://doi.org/10.1038/s41593-024-01580-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-024-01580-z

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