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:

Divergent transcriptional regulation of astrocyte reactivity across disorders

An Author Correction to this article was published on 28 March 2023

This article has been updated

Abstract

Astrocytes respond to injury and disease in the central nervous system with reactive changes that influence the outcome of the disorder1,2,3,4. These changes include differentially expressed genes (DEGs) whose contextual diversity and regulation are poorly understood. Here we combined biological and informatic analyses, including RNA sequencing, protein detection, assay for transposase-accessible chromatin with high-throughput sequencing (ATAC-seq) and conditional gene deletion, to predict transcriptional regulators that differentially control more than 12,000 DEGs that are potentially associated with astrocyte reactivity across diverse central nervous system disorders in mice and humans. DEGs associated with astrocyte reactivity exhibited pronounced heterogeneity across disorders. Transcriptional regulators also exhibited disorder-specific differences, but a core group of 61 transcriptional regulators was identified as common across multiple disorders in both species. We show experimentally that DEG diversity is determined by combinatorial, context-specific interactions between transcriptional regulators. Notably, the same reactivity transcriptional regulators can regulate markedly different DEG cohorts in different disorders; changes in the access of transcriptional regulators to DNA-binding motifs differ markedly across disorders; and DEG changes can crucially require multiple reactivity transcriptional regulators. We show that, by modulating reactivity, transcriptional regulators can substantially alter disorder outcome, implicating them as therapeutic targets. We provide searchable resources of disorder-related reactive astrocyte DEGs and their predicted transcriptional regulators. Our findings show that transcriptional changes associated with astrocyte reactivity are highly heterogeneous and are customized from vast numbers of potential DEGs through context-specific combinatorial transcriptional-regulator interactions.

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

Fig. 1: Astrocyte reactivity DEGs and TRs vary across disorders.
Fig. 2: Astrocyte reactivity TRs were identified using multiple approaches, including DEG analysis, chromatin accessibility changes at DNA-binding motifs and immunohistochemistry.
Fig. 3: TR regulation of astrocyte reactivity DEGs, chromatin accessibility and disorder outcome.
Fig. 4: Astrocyte reactivity DEGs and TRs diverge across CNS disorders in mice and humans.

Similar content being viewed by others

Data availability

Raw and normalized genomic data have been deposited at the NCBI Gene Expression Omnibus under the SuperSeries accession number GSE199482. Genomics data are also available through a searchable, open-access website (http://tr.astrocytereactivity.com). Source data are provided with this paper.

Change history

References

  1. Sofroniew, M. V. & Vinters, H. V. Astrocytes: biology and pathology. Acta Neuropathol. 119, 7–35 (2010).

    Article  PubMed  Google Scholar 

  2. Burda, J. E. & Sofroniew, M. V. Reactive gliosis and the multicellular response to CNS damage and disease. Neuron 81, 229–248 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Linnerbauer, M., Wheeler, M. A. & Quintana, F. J. Astrocyte crosstalk in CNS inflammation. Neuron 108, 608–622 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Escartin, C. et al. Reactive astrocyte nomenclature, definitions, and future directions. Nat. Neurosci. 24, 312–325 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Haim, L. B. & Rowitch, D. H. Functional diversity of astrocytes in neural circuit regulation. Nat. Rev. Neurosci. 18, 31–41 (2017).

    Article  CAS  PubMed  Google Scholar 

  7. Khakh, B. S. & Deneen, B. The emerging nature of astrocyte diversity. Annu. Rev. Neurosci. 42, 187–207 (2019).

    Article  CAS  PubMed  Google Scholar 

  8. Lu, T. Y. et al. Axon degeneration induces glial responses through Draper-TRAF4-JNK signalling. Nat. Commun. 8, 14355 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  9. Sofroniew, M. V. Astrocyte barriers to neurotoxic inflammation. Nat. Rev. Neurosci. 16, 249–263 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Khakh, B. S. & Sofroniew, M. V. Diversity of astrocyte functions and phenotypes in neural circuits. Nat. Neurosci. 18, 942–952 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Yu, X. et al. Context-specific striatal astrocyte molecular responses are phenotypically exploitable. Neuron 108, 1146–1162 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Anderson, M. A. et al. Astrocyte scar formation aids central nervous system axon regeneration. Nature 532, 195–200 (2016).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  13. Schreiner, B. et al. Astrocyte depletion impairs redox homeostasis and triggers neuronal loss in the adult CNS. Cell Rep. 12, 1377–1384 (2015).

    Article  CAS  PubMed  Google Scholar 

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

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  15. Buenrostro, J. D., Giresi, P. G., Zaba, L. C., Chang, H. Y. & Greenleaf, W. J. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat. Methods 10, 1213–1218 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Wanner, I. B. et al. Glial scar borders are formed by newly proliferated, elongated astrocytes that interact to corral inflammatory and fibrotic cells via STAT3-dependent mechanisms after spinal cord injury. J. Neurosci. 33, 12870–12886 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Diaz-Castro, B., Bernstein, A. M., Coppola, G., Sofroniew, M. V. & Khakh, B. S. Molecular and functional properties of cortical astrocytes during peripherally induced neuroinflammation. Cell Rep. 36, 109508 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Inoue, F., Kreimer, A., Ashuach, T., Ahituv, N. & Yosef, N. Identification and massively parallel characterization of regulatory elements driving neural induction. Cell Stem Cell 25, 713–727 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Lattke, M. et al. Extensive transcriptional and chromatin changes underlie astrocyte maturation in vivo and in culture. Nat. Commun. 12, 4335 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  20. Granja, J. M. et al. ArchR is a scalable software package for integrative single-cell chromatin accessibility analysis. Nat. Genet. 53, 403–411 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Schep, A. N., Wu, B., Buenrostro, J. D. & Greenleaf, W. J. chromVAR: inferring transcription-factor-associated accessibility from single-cell epigenomic data. Nat. Methods 14, 975–978 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Wingelhofer, B. et al. Implications of STAT3 and STAT5 signaling on gene regulation and chromatin remodeling in hematopoietic cancer. Leukemia 32, 1713–1726 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Henry, C. J. et al. Minocycline attenuates lipopolysaccharide (LPS)-induced neuroinflammation, sickness behavior, and anhedonia. J. Neuroinflamm. 5, 15 (2008).

    Article  Google Scholar 

  24. Herrmann, J. E. et al. STAT3 is a critical regulator of astrogliosis and scar formation after spinal cord injury. J. Neurosci. 28, 7231–7243 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Wohlfahrt, T. et al. PU.1 controls fibroblast polarization and tissue fibrosis. Nature 566, 344–349 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  26. Sofroniew, M. V. Astrocyte reactivity: subtypes, states, and functions in CNS innate immunity. Trends Immunol. 41, 758–770 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Diaz-Castro, B., Gangwani, M. R., Yu, X., Coppola, G. & Khakh, B. S. Astrocyte molecular signatures in Huntington’s disease. Sci. Transl. Med. 11, eaaw8546 (2019).

    Article  CAS  PubMed  Google Scholar 

  28. Sun, S. Y. et al. Translational profiling identifies a cascade of damage initiated in motor neurons and spreading to glia in mutant SOD1-mediated ALS. Proc. Natl Acad. Sci. USA 112, E6993–E7002 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Sekar, S. et al. Alzheimer’s disease is associated with altered expression of genes involved in immune response and mitochondrial processes in astrocytes. Neurobiol. Aging 36, 583–591 (2015).

    Article  CAS  PubMed  Google Scholar 

  30. Habib, N. et al. Disease-associated astrocytes in Alzheimer’s disease and aging. Nat. Neurosci. 23, 701–706 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Kamphuis, W. et al. GFAP and vimentin deficiency alters gene expression in astrocytes and microglia in wild-type mice and changes the transcriptional response of reactive glia in mouse model for Alzheimer’s disease. Glia 63, 1036–1056 (2015).

    Article  PubMed  Google Scholar 

  32. Boisvert, M. M., Erikson, G. A., Shokhirev, M. N. & Allen, N. J. The aging astrocyte transcriptome from multiple regions of the mouse brain. Cell Rep. 22, 269–285 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Rojo, A. I. et al. Deficiency in the transcription factor NRF2 worsens inflammatory parameters in a mouse model with combined tauopathy and amyloidopathy. Redox Biol. 18, 173–180 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Oksanen, M. et al. NF-E2-related factor 2 activation boosts antioxidant defenses and ameliorates inflammatory and amyloid properties in human presenilin-1 mutated Alzheimer’s disease astrocytes. Glia 68, 589–599 (2020).

    Article  PubMed  Google Scholar 

  35. Laug, D. et al. Nuclear factor I-A regulates diverse reactive astrocyte responses after CNS injury. J. Clin. Invest. 129, 4408–4418 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Venkatesh, I. et al. Co-occupancy identifies transcription factor co-operation for axon growth. Nat. Commun. 12, 2555 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  37. Garcia, A. D. R., Doan, N. B., Imura, T., Bush, T. G. & Sofroniew, M. V. GFAP-expressing progenitors are the principle source of constitutive neurogenesis in adult mouse forebrain. Nat. Neurosci. 7, 1233–1241 (2004).

    Article  CAS  PubMed  Google Scholar 

  38. Sanz, E. et al. Cell-type-specific isolation of ribosome-associated mRNA from complex tissues. Proc. Natl Acad. Sci. USA 106, 13939–13944 (2009).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  39. Sumi-Ichinose, C., Ichinose, H., Metzger, D. & Chambon, P. SNF2β-BRG1 is essential for the viability of F9 murine embryonal carcinoma cells. Mol. Cell. Biol. 17, 5976–5986 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Metz, G. A. & Whishaw, I. Q. The ladder rung walking task: a scoring system and its practical application. J. Vis. Exp. https://doi.org/10.3791/1204 (2009).

  41. Dantzer, R., O'Connor, J. C., Freund, G. G., Johnson, R. W. & Kelley, K. W. From inflammation to sickness and depression: when the immune system subjugates the brain. Nat. Rev. Neurosci. 9, 46–56 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Sukoff Rizzo, S. J. et al. Evidence for sustained elevation of IL-6 in the CNS as a key contributor of depressive-like phenotypes. Transl. Psychiatry 2, e199 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Lachmann, A. et al. ChEA: transcription factor regulation inferred from integrating genome-wide ChIP-X experiments. Bioinformatics 26, 2438–2444 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Fornes, O. et al. JASPAR 2020: update of the open-access database of transcription factor binding profiles. Nucleic Acids Res. 48, D87–D92 (2020).

    CAS  PubMed  Google Scholar 

  45. Matys, V. et al. TRANSFAC: transcriptional regulation, from patterns to profiles. Nucleic Acids Res. 31, 374–378 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. 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 

  47. Krishnaswami, S. R. et al. Using single nuclei for RNA-seq to capture the transcriptome of postmortem neurons. Nat. Protoc. 11, 499–524 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Bhattacharyya, S., Sathe, A. A., Bhakta, M., Xing, C. & Munshi, N. V. PAN-INTACT enables direct isolation of lineage-specific nuclei from fibrous tissues. PLoS ONE 14, e0214677 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Batiuk, M. Y. et al. An immunoaffinity-based method for isolating ultrapure adult astrocytes based on ATP1B2 targeting by the ACSA-2 antibody. J. Biol. Chem. 292, 8874–8891 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Cusanovich, D. A. et al. Multiplex single cell profiling of chromatin accessibility by combinatorial cellular indexing. Science 348, 910–914 (2015).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  51. Deschamps, S. et al. Chromatin loop anchors contain core structural components of the gene expression machinery in maize. BMC Genom. 22, 23 (2021).

    Article  CAS  Google Scholar 

  52. Weirauch, M. T. et al. Determination and inference of eukaryotic transcription factor sequence specificity. Cell 158, 1431–1443 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Kheradpour, P. & Kellis, M. Systematic discovery and characterization of regulatory motifs in ENCODE TF binding experiments. Nucleic Acids Res. 42, 2976–2987 (2014).

    Article  CAS  PubMed  Google Scholar 

  54. Courtine, G. et al. Transformation of nonfunctional spinal circuits into functional states after the loss of brain input. Nat. Neurosci. 12, 1333–1342 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Neve, L. D., Savage, A. A., Koke, J. R. & Garcia, D. M. Activating transcription factor 3 and reactive astrocytes following optic nerve injury in zebrafish. Comp. Biochem. Physiol. C 155, 213–218 (2012).

    CAS  Google Scholar 

  56. Kim, K. H., Jeong, J. Y., Surh, Y. J. & Kim, K. W. Expression of stress-response ATF3 is mediated by Nrf2 in astrocytes. Nucleic Acids Res. 38, 48–59 (2010).

    Article  CAS  PubMed  Google Scholar 

  57. Koyama, Y. et al. Endothelin-1 stimulates expression of cyclin D1 and S-phase kinase-associated protein 2 by activating the transcription factor STAT3 in cultured rat astrocytes. J. Biol. Chem. 294, 3920–3933 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Cardinaux, J. R., Allaman, I. & Magistretti, P. J. Pro-inflammatory cytokines induce the transcription factors C/EBPbeta and C/EBPdelta in astrocytes. Glia 29, 91–97 (2000).

    Article  CAS  PubMed  Google Scholar 

  59. Ko, C. Y. et al. Glycogen synthase kinase-3β-mediated CCAAT/enhancer-binding protein delta phosphorylation in astrocytes promotes migration and activation of microglia/macrophages. Neurobiol. Aging 35, 24–34 (2014).

    Article  CAS  PubMed  Google Scholar 

  60. Pardo, L. et al. Targeted activation of CREB in reactive astrocytes is neuroprotective in focal acute cortical injury. Glia 64, 853–874 (2016).

    Article  PubMed  Google Scholar 

  61. Zou, F. et al. Different functions of HIPK2 and CtBP2 in traumatic brain injury. J. Mol. Neurosci. 49, 395–408 (2013).

    Article  CAS  PubMed  Google Scholar 

  62. Robinson, K. F., Narasipura, S. D., Wallace, J., Ritz, E. M. & Al-Harthi, L. Negative regulation of IL-8 in human astrocytes depends on beta-catenin while positive regulation is mediated by TCFs/LEF/ATF2 interaction. Cytokine 136, 155252 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Yang, C. et al. β-Catenin signaling initiates the activation of astrocytes and its dysregulation contributes to the pathogenesis of astrocytomas. Proc. Natl Acad. Sci. USA 109, 6963–6968 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  64. Wu, J. F. et al. Ablation of the transcription factors E2F1-2 limits neuroinflammation and associated neurological deficits after contusive spinal cord injury. Cell Cycle 14, 3698–3712 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Beck, H., Semisch, M., Culmsee, C., Plesnila, N. & Hatzopoulos, A. K. Egr-1 regulates expression of the glial scar component phosphacan in astrocytes after experimental stroke. Am. J. Pathol. 173, 77–92 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Mayer, S. I., Rossler, O. G., Endo, T., Charnay, P. & Thiel, G. Epidermal-growth-factor-induced proliferation of astrocytes requires Egr transcription factors. J. Cell Sci. 122, 3340–3350 (2009).

    Article  CAS  PubMed  Google Scholar 

  67. Wang, H. H., Hsieh, H. L., Wu, C. Y. & Yang, C. M. Oxidized low-density lipoprotein-induced matrix metalloproteinase-9 expression via PKC-delta/p42/p44 MAPK/Elk-1 cascade in brain astrocytes. Neurotox. Res. 17, 50–65 (2010).

    Article  PubMed  Google Scholar 

  68. Gerhauser, I., Alldinger, S. & Baumgartner, W. Ets-1 represents a pivotal transcription factor for viral clearance, inflammation, and demyelination in a mouse model of multiple sclerosis. J. Neuroimmunol. 188, 86–94 (2007).

    Article  CAS  PubMed  Google Scholar 

  69. Hashimoto, K. et al. Long-term activation of c-Fos and c-Jun in optic nerve head astrocytes in experimental ocular hypertension in monkeys and after exposure to elevated pressure in vitro. Brain Res. 1054, 103–115 (2005).

    Article  CAS  PubMed  Google Scholar 

  70. Yang, C. C., Hsiao, L. D. & Yang, C. M. Galangin inhibits LPS-induced MMP-9 expression via suppressing protein kinase-dependent AP-1 and FoxO1 activation in rat brain astrocytes. J. Inflamm. Res. 13, 945–960 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Cui, M., Huang, Y., Tian, C., Zhao, Y. & Zheng, J. FOXO3a inhibits TNF-α- and IL-1β-induced astrocyte proliferation: implication for reactive astrogliosis. Glia 59, 641–654 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  72. Kizil, C. et al. Regenerative neurogenesis from neural progenitor cells requires injury-induced expression of Gata3. Dev. Cell 23, 1230–1237 (2012).

    Article  CAS  PubMed  Google Scholar 

  73. Garcia, A. D., Petrova, R., Eng, L. & Joyner, A. L. Sonic hedgehog regulates discrete populations of astrocytes in the adult mouse forebrain. J. Neurosci. 30, 13597–13608 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Du, F. et al. Hyperthermic preconditioning protects astrocytes from ischemia/reperfusion injury by up-regulation of HIF-1 alpha expression and binding activity. Biochim. Biophys. Acta 1802, 1048–1053 (2010).

    Article  CAS  PubMed  Google Scholar 

  75. Choi, K., Ni, L. & Jonakait, G. M. Fas ligation and tumor necrosis factor alpha activation of murine astrocytes promote heat shock factor-1 activation and heat shock protein expression leading to chemokine induction and cell survival. J. Neurochem. 116, 438–448 (2011).

    Article  CAS  PubMed  Google Scholar 

  76. Tzeng, S. F., Kahn, M., Liva, S. & De Vellis, J. Tumor necrosis factor-α regulation of the Id gene family in astrocytes and microglia during CNS inflammatory injury. Glia 26, 139–152 (1999).

    Article  CAS  PubMed  Google Scholar 

  77. Aronica, E. et al. Expression of Id proteins increases in astrocytes in the hippocampus of epileptic rats. Neuroreport 12, 2461–2465 (2001).

    Article  CAS  PubMed  Google Scholar 

  78. Jarosinski, K. W. & Massa, P. T. Interferon regulatory factor-1 is required for interferon-γ-induced MHC class I genes in astrocytes. J. Neuroimmunol. 122, 74–84 (2002).

    Article  CAS  PubMed  Google Scholar 

  79. Tarassishin, L. et al. Interferon regulatory factor 3 inhibits astrocyte inflammatory gene expression through suppression of the proinflammatory miR-155 and miR-155*. Glia 59, 1911–1922 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  80. Gadea, A., Schinelli, S. & Gallo, V. Endothelin-1 regulates astrocyte proliferation and reactive gliosis via a JNK/c-Jun signaling pathway. J. Neurosci. 28, 2394–2408 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Park, J. H. et al. Induction of Kruppel-like factor 4 expression in reactive astrocytes following ischemic injury in vitro and in vivo. Histochem. Cell Biol. 141, 33–42 (2014).

    Article  CAS  PubMed  Google Scholar 

  82. Jeong, K. H., Lee, K. E., Kim, S. Y. & Cho, K. O. Upregulation of Kruppel-like factor 6 in the mouse hippocampus after pilocarpine-induced status epilepticus. Neuroscience 186, 170–178 (2011).

    Article  CAS  PubMed  Google Scholar 

  83. Liu, F., Ni, J. J., Huang, J. J., Kou, Z. W. & Sun, F. Y. VEGF overexpression enhances the accumulation of phospho-S292 MeCP2 in reactive astrocytes in the adult rat striatum following cerebral ischemia. Brain Res. 1599, 32–43 (2015).

    Article  CAS  PubMed  Google Scholar 

  84. Wang, F. et al. 2-Arachidonylglycerol protects primary astrocytes exposed to oxygen-glucose deprivation through a blockade of NDRG2 signaling and STAT3 phosphorylation. Rejuv. Res. 19, 215–222 (2016).

    Article  Google Scholar 

  85. Perez-Ortiz, J. M. et al. Mechanical lesion activates newly identified NFATc1 in primary astrocytes: implication of ATP and purinergic receptors. Eur. J. Neurosci. 27, 2453–2465 (2008).

    Article  PubMed  Google Scholar 

  86. Yang, Y. et al. Hemoglobin pretreatment endows rat cortical astrocytes resistance to hemin-induced toxicity via Nrf2/HO-1 pathway. Exp. Cell. Res. 361, 217–224 (2017).

    Article  CAS  PubMed  Google Scholar 

  87. Brambilla, R. et al. Inhibition of astroglial nuclear factor κB reduces inflammation and improves functional recovery after spinal cord injury. J. Exp. Med. 202, 145–156 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. LeComte, M. D., Shimada, I. S., Sherwin, C. & Spees, J. L. Notch1-STAT3-ETBR signaling axis controls reactive astrocyte proliferation after brain injury. Proc. Natl Acad. Sci. USA 112, 8726–8731 (2015).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  89. Chen, X. L. et al. Effects of interleukin-6 and IL-6/AMPK signaling pathway on mitochondrial biogenesis and astrocytes viability under experimental septic condition. Int. Immunopharmacol. 59, 287–294 (2018).

    Article  CAS  PubMed  Google Scholar 

  90. Gutbier, S. et al. Prevention of neuronal apoptosis by astrocytes through thiol-mediated stress response modulation and accelerated recovery from proteotoxic stress. Cell Death Differ. 25, 2101–2117 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Chen, Y. et al. The basic helix-loop-helix transcription factor olig2 is critical for reactive astrocyte proliferation after cortical injury. J. Neurosci. 28, 10983–10989 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Koyama, Y. Signaling molecules regulating phenotypic conversions of astrocytes and glial scar formation in damaged nerve tissues. Neurochem. Int. 78, 35–42 (2014).

    Article  CAS  PubMed  Google Scholar 

  93. Steliga, A. et al. Transcription factor Pax6 is expressed by astroglia after transient brain ischemia in the rat model. Folia Neuropathol. 51, 203–213 (2013).

    Article  CAS  PubMed  Google Scholar 

  94. Guo, X. et al. The AMPK-PGC-1α signaling axis regulates the astrocyte glutathione system to protect against oxidative and metabolic injury. Neurobiol. Dis. 113, 59–69 (2018).

    Article  CAS  PubMed  Google Scholar 

  95. Diehl, J. A., Tong, W., Sun, G. & Hannink, M. Tumor necrosis factor-alpha-dependent activation of a RelA homodimer in astrocytes. Increased phosphorylation of RelA and MAD-3 precede activation of RelA. J. Biol. Chem. 270, 2703–2707 (1995).

    Article  CAS  PubMed  Google Scholar 

  96. Yoo, K. Y. et al. Time-course alterations of Toll-like receptor 4 and NF-κB p65, and their co-expression in the gerbil hippocampal CA1 region after transient cerebral ischemia. Neurochem. Res. 36, 2417–2426 (2011).

    Article  CAS  PubMed  Google Scholar 

  97. Gupta, A. S. et al. RelB controls adaptive responses of astrocytes during sterile inflammation. Glia 67, 1449–1461 (2019).

    PubMed  PubMed Central  Google Scholar 

  98. Li, H. et al. The deficiency of NRSF/REST enhances the pro-inflammatory function of astrocytes in a model of Parkinson’s disease. Biochim. Biophys. Acta 1866, 165590 (2020).

    Article  CAS  Google Scholar 

  99. Marumo, T. et al. Notch signaling regulates nucleocytoplasmic Olig2 translocation in reactive astrocytes differentiation after ischemic stroke. Neurosci. Res. 75, 204–209 (2013).

    Article  CAS  PubMed  Google Scholar 

  100. Tanigaki, K. & Honjo, T. Two opposing roles of RBP-J in Notch signaling. Curr. Top. Dev. Biol. 92, 231–252 (2010).

    Article  CAS  PubMed  Google Scholar 

  101. Wong, J. K. et al. Attenuation of cerebral ischemic injury in Smad1 deficient mice. PLoS ONE 10, e0136967 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  102. Law, A. K. T. et al. TGF-β1 induction of the adenine nucleotide translocator 1 in astrocytes occurs through Smads and Sp1 transcription factors. BMC Neurosci. 5, 1 (2004).

    Article  PubMed  PubMed Central  Google Scholar 

  103. Chen, C. et al. Astrocyte-specific deletion of Sox2 promotes functional recovery after traumatic brain injury. Cereb. Cortex 29, 54–69 (2019).

    Article  PubMed  Google Scholar 

  104. Song, W. et al. Immunohistochemical staining of ERG and SOX9 as potential biomarkers of docetaxel response in patients with metastatic castration-resistant prostate cancer. Oncotarget 7, 83735–83743 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  105. Mao, X., Moerman-Herzog, A. M., Wang, W. & Barger, S. W. Differential transcriptional control of the superoxide dismutase-2 κB element in neurons and astrocytes. J. Biol. Chem. 281, 35863–35872 (2006).

    Article  CAS  PubMed  Google Scholar 

  106. Haroon, F. et al. Gp130-dependent astrocytic survival is critical for the control of autoimmune central nervous system inflammation. J. Immunol. 186, 6521–6531 (2011).

    Article  CAS  PubMed  Google Scholar 

  107. Khorooshi, R., Babcock, A. A. & Owens, T. NF-κB-driven STAT2 and CCL2 expression in astrocytes in response to brain injury. J. Immunol. 181, 7284–7291 (2008).

    Article  CAS  PubMed  Google Scholar 

  108. Doherty, J. et al. PI3K signaling and Stat92E converge to modulate glial responsiveness to axonal injury. PLoS Biol. 12, e1001985 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  109. Park, S. J. et al. Astrocytes, but not microglia, rapidly sense H2O2 via STAT6 phosphorylation, resulting in cyclooxygenase-2 expression and prostaglandin release. J. Immunol. 188, 5132–5141 (2012).

    Article  CAS  PubMed  Google Scholar 

  110. Lurbke, A. et al. Limited TCF7L2 expression in MS lesions. PLoS ONE 8, e72822 (2013).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  111. Chung, Y. H. et al. Enhanced expression of p53 in reactive astrocytes following transient focal ischemia. Neurol. Res. 24, 324–328 (2002).

    Article  ADS  CAS  PubMed  Google Scholar 

  112. Huang, Z. H. et al. YAP is a critical inducer of SOCS3, preventing reactive astrogliosis. Cereb. Cortex 26, 2299–2310 (2016).

    Article  PubMed  Google Scholar 

  113. Vivinetto, A. L. et al. Zeb2 is a regulator of astrogliosis and functional recovery after CNS injury. Cell Rep. 31, 107834 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Itoh, N. et al. Cell-specific and region-specific transcriptomics in the multiple sclerosis model: focus on astrocytes. Proc. Natl Acad. Sci. USA 115, E302–E309 (2018).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Zhang, Y. et al. Purification and characterization of progenitor and mature human astrocytes reveals transcriptional and functional differences with mouse. Neuron 89, 37–53 (2016).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported by the US National Institutes of Health (NS084030 to M.V.S.; and F32NS096858, K99NS105915 and R00NS105915 to J.E.B.); the Dr Miriam and Sheldon G. Adelson Medical Foundation (to M.V.S. and R.K.); the Paralyzed Veterans Research Foundation of America (to J.E.B., T.M.O. and M.V.S.); the American Australian Fellowship (to T.M.O.); the Microscopy Core Resource of UCLA Broad Stem Cell Research Center; and Wings for Life (to M.V.S., J.E.B. and T.M.O.).

Author information

Authors and Affiliations

Authors

Contributions

J.E.B., T.M.O. and M.V.S. designed experiments. J.E.B., T.M.O., Y.A., S. Wan, A.M.B., J.H.K., A.R. and S. Wah conducted experiments. J.E.B., T.M.O., Y.A., K.B.S., A.C., S.D., R.K., S.M. and M.V.S. analysed data. J.E.B., T.M.O. and M.V.S. prepared the manuscript.

Corresponding authors

Correspondence to Joshua E. Burda or Michael V. Sofroniew.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature 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 figures and tables

Extended Data Fig. 1 RNA sequencing, Transcriptional Regulator Enrichment Analysis (TREA) and previously published TRs.

a. Flow diagram of RNA-seq procedure. To minimize technical differences, we used mice of similar age and genetic background for all experimental disorder models and examined the same anatomical region (thoracic spinal cord). Spinal cord tissue from different experiments were frozen until all experiments were completed. All tissue was then processed at the same time to limit the potential for technical variations. The RiboTag procedure was used to harvest ribosome-associated RNA transcripts specifically from reactive astrocytes12. RiboTag hemagglutinin (HA) was transgenically-targeted specifically to astrocytes by using well-characterized mGFAP-Cre12. RNA-sequencing and analysis were conducted under identical conditions. b. Specificity of RiboTag-HA targeting to astrocytes and not microglia or other cells. Two sets of orthogonal (3D) scans from uninjured spinal cord or after spinal cord injury (SCI) with multichannel immunofluorescence for HA targeted to astrocytes plus Sox9 and Iba1 as markers of astrocytes or microglia respectively. The same areas are shown with different fluorescent wavelength filters and different orthogonal slices that demonstrate 3D staining associated with astrocytes or microglia. HA is robustly present in Sox9-positive astrocytes but is not detectable in Iba1-postive microglia in either uninjured cord or after SCI. Absence of HA-targeting to neurons or oligodendrocytes has been demonstrated previously12. These immunohistochemical comparisons were repeated independently three times with similar results. c. Venn diagrams show that the relative proportions of shared astrocyte reactivity DEGs and TRs identified in spinal cord astrocytes after EAE, SCI or LPS do not detectably differ when using thresholds of FDR < 0.1 or FDR < 0.05. Table shows PCA analysis of DEGs identified in spinal cord astrocytes after EAE, SCI or LPS using thresholds of FDR < 0.1 or FDR < 0.05. The relative locations of the three disorders in PC space when compared to non-reactive astrocytes do not detectably differ when using thresholds of FDR < 0.1 or FDR < 0.05 as reflected in the percent of total vector length and the angles between vectors. d. Flow diagram of Transcriptional Regulator Enrichment Analysis (TREA) procedure for TR identification by upstream analysis of DEGs in reactive astrocytes. To identify TRs of astrocyte gene expression, we applied a conservative, multi-step algorithm that draws on both computationally- and biologically-derived regulator–target gene interaction data from multiple resource databases: i) ChEA:43 transcription factor regulation inferred from integrating genome-wide ChIP-X experiments, ii) JASPAR44 and iii) TRANSFAC45 transcription factor DNA-binding preferences as position weight matrices, and iv) Ingenuity Pathway Analysis Upstream Regulator Analytic (IPA®, Qiagen, Valencia, CA). Using these resource databases, TREA identifies TRs implicated in regulating DEGs by interrogating multiple forms of TR–target gene regulatory interactions that include findings from experimental studies involving techniques such as chromatin immunoprecipitation and genetic loss-of-function studies, as well as well-validated predictive computational DNA binding ‘motif’ analytics. In this manner, TR–target gene interactions considered by TREA include traditional direct TR-DNA binding mechanisms, as well as indirect forms of gene expression regulation wherein a TR may act through different types of intermediaries to effect expression of downstream target genes, including chromatin modifiers and other forms of epigenetic regulators. TRs were included if they met either of two criteria: (1) convergence across resource databases; (2) differential gene expression of the TR plus convergence with at least one resource database. Together, these databases allow for interrogation of gene expression datasets for enrichment of downstream targets for approximately 1350 TRs. Resource database output files containing statistically enriched TRs and their downstream astrocyte target gene IDs were processed for TREA using custom Python scripts available at GitHub repository (https://github.com/burdalab/TREA). Final TREA libraries containing significantly enriched TRs and associated astrocyte target genes were then generated for each condition’s DEG dataset. TREA libraries were used for all comparisons of astrocyte TRs and gene expression profiles across disorders and experimental conditions and to generate a resource database of reactive astrocyte TRs and the DEGs that they regulate across a broad spectrum of CNS disorders and conditions. This database can be accessed via an open-source website http://tr.astrocytereactivity.com and has multiple search parameters according to TR, DEG or condition. e. Published astrocyte reactivity TRs plus literature references55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113. *4 of 62 published TRs not predicted in EAE, LPS or SCI.

Source data

Extended Data Fig. 2 Single nucleus ATAC sequencing.

a. Experimental models: All ATAC-seq experiments examining LPS or SCI treatments or healthy controls used wild-type or transgenic mice of 657Bl6 background strain. Transgenic mice expressing mGFAP-Cre12 were used for astrocyte-specific deletion (cKO) of Smarca4 or Stat3. b. The same region of thoracic spinal cord at T9-T10 was harvested for all evaluations. Spinal cord tissue from different experiments were frozen until all experiments were completed. All tissue was then processed at the same time to limit technical variations. c. Two spinal cords from the same experimental group were pooled to prepare suspensions of nuclei. These suspensions were enriched for astrocyte nuclei with a Sox9 antibody and magnetic beads precipitation. Two such suspensions were prepared from a total of n = 4 mice per experimental condition. d. Two biological replicates, each consisting of nuclei from two mice, were used for single nucleus ATAC-seq. All biological replicates from all experimental conditions were sequenced at the same time to avoid batch effects. e. Box plots compare the distribution of the per-nucleus averages of unique DNA fragments per nucleus or TSS enrichment per nucleus for the seven experimental conditions listed, with each dot representing the average value of all nuclei collected from each experimental group (whiskers show range, box encompass 25–75% quartiles and the centre line indicates the median; n = 8 WT healthy mice and n = 4 mice for all other experimental conditions). f. UMAP clustering based on differential ATAC peaks across 145,973 high quality nuclei isolated across all conditions showed separation of nuclei into multiple distinct clusters. g. UMAP distribution of examples of specific DAGs used to identify the dominant cell types in different clusters. h. Heatmap of DAGs used to identify specific cell types. Each line represents the per-nucleus z-score for a given gene averaged across all nuclei in a cluster. i. UMAP and stacked bar graph show relative contribution of biological replicates for different experimental conditions. Because experiments for LPS and SCI were conducted at different times, groups of healthy control mice were collected for each experiment. All biological replicates showed similar contributions to their respective clusters, confirming that separation of treatment groups with essentially no overlap of LPS with either healthy or SCI was due to biological variation and not due to technical artifacts. * each biological replicate consisted of nuclei from two mice.

Source data

Extended Data Fig. 3 Identification of TR proteins in reactive astrocytes by immunohistochemistry.

For immunohistochemical (IHC) detection of TR protein we focused on previously unpublished, TREA-predicted astrocyte reactivity TRs for which sensitive antibodies were commercially available whose specificity was supported by western blots. a. Newly identified TRs co-localized by IHC in reactive astrocytes in thoracic spinal cord (T9-T-10) after LPS treatment. TRs are show in alphabetical sequence. b. Summary of all (newly identified and previously published) astrocyte reactivity TRs after LPS treatment identified here by at least two experimental procedures, either prediction from DEGs by TREA, or prediction based on significant change in motif access determined by ATAC-seq or by IHC or by all three. c. Newly identified TRs co-localized by IHC in reactive astrocytes in thoracic spinal cord (T9-T-10) after SCI (continued in Extended Data Figs. 4, 5). TRs are show in alphabetical sequence. Each immunohistochemical evaluation was repeated at least three times with similar results.

Extended Data Fig. 4 Immunohistochemistry and motif analysis of TR proteins in reactive astrocytes after SCI.

a. Newly identified TRs co-localized by IHC in reactive astrocytes in thoracic spinal cord (T9-T-10) after SCI (see also Extended Data Figs. 3, 5). Each immunohistochemical evaluation was repeated at least three times with similar results. b. Summary and heatmap of all (newly identified and previously published) astrocyte reactivity TRs after SCI that have detectable DNA-binding motifs and were identified here as exhibiting significant change in motif access determined by ATAC-seq.

Source data

Extended Data Fig. 5 Immunohistochemistry of TR proteins in reactive astrocytes after SCI.

a. Newly identified TRs co-localized by IHC in reactive astrocytes in thoracic spinal cord (T9-T-10) after SCI (see also Extended Data Figs. 3, 4). b. Summary of all (newly identified and previously published) astrocyte reactivity TRs after LPS treatment identified here by at least two experimental procedures, either prediction from DEGs by TREA, or prediction based on significant change in motif access determined by ATAC-seq or by IHC or by all three. c. Nine examples of previously published TRs co-localized by IHC in reactive astrocytes in thoracic spinal cord (T9-T-10) after SCI show that these exhibit similar staining patterns to TRs newly identified here. Each immunohistochemical evaluation was repeated at least three times with similar results.

Extended Data Fig. 6 Astrocyte-specific TR deletion (cKO).

a. Experimental models: Transgenic mice expressing mGFAP-Cre12 were used for astrocyte-specific deletion (cKO) of Smarca4 or Stat3. Smarca4 was selected as a newly identified TREA-predicted reactivity TR that lacks DNA binding motifs and acts as a chromatin regulator via protein-protein interactions. Stat3 is a well-established astrocyte reactivity TR that acts via DNA-binding motifs12,24 and that was predicted by both TREA and ATAC-seq motif analysis as a reactivity TR in both LPS and SCI. b. Heatmaps show that Smarca4-astro-cKO had minimal effects on overall gene expression or on the expression of highly enriched astrocyte genes under basal conditions in untreated mice, shown as mean FPKM (fragments per kilobase of transcript sequence per million mapped fragments). c. Immunohistochemistry images and graphs (mean + sem) of various staining parameters show that Smarca4-astro-cKO had no visibly detectable or quantifiably significant effects on the appearance or number of astrocytes, neurons or microglia in untreated mice. Unpaired t tests, ns non-significant WT versus Smarca4-cKO, n = 4 mice per group. d. Effects of Smarca4-astro-cKO on RNA-seq reads of Gfap and seven predicted Smarca4-regulated DEGs that are not expressed in WT untreated and are upregulated by LPS in WT but not Smarca4-astro-cKO mice. e. In situ hybridization shows predicted Smarca4-regulated gene, Cnr1, expresed in WT, but not Smarca4-astro-cKO mice after LPS. f. Multichannel immunofluorescence demonstration of Stat3 and Srebf1, or of Smarca4 and Zfp36, in the same reactive astrocytes after SCI. g. Heatmaps compare changes from healthy in differential gene expression (DEG) or differential chromatin accessibility (DAG) across the same genes after LPS or SCI in WT, Smarca4-cKO or Stat3-cKO mice, and graphs show an 86 to 96% congruence between changes in gene expression and changes in chromatin accessibility in both Smarca4- and Stat3-regulated DEGs after LPS or SCI respectively. h. Heatmaps show that Stat3-cKO significantly alters the changes from healthy normally observed after SCI in both differential gene expression (DEG) and differential chromatin accessibility (DAG) across the same 31 genes that lack Stat3-binding motifs. i. Heatmaps show that Stat3-cKO significantly alters the changes from healthy normally observed after SCI in differential gene expression (DEG) of 71 chromatin regulators. j. Multichannel immunofluorescence of Irf9 and Cxcl10 in the same reactive astrocyte after SCI. k. Graphs (mean + sem) show effects of Smarca4-astro-cKO and Stat3-astro-cKO on various microglial histopathological responses to LPS, P values are cKO+LPS versus WT+LPS, one-way ANOVA with Bonferroni’s test, ns nonsignificant; for all graphs WT with no LPS n = 18 mice, all other conditions n = 6 mice. PCA shows composite microglia histopathology score derived from histopathological quantifications in the four graphs. l. Immunohistochemistry and graphs of mean + sem (n > 6) cell counts of the neuronal marker, NeuN, shows that high dose systemic LPS sufficient to cause pronounced microglial activation and behavioural effects did not lead to detectable neuronal loss in either spinal cord or brain. One-way ANOVA with Bonferroni’s test, ns nonsignificant; for both graphs WT with no LPS n = 18 mice, all other conditions n = 4 mice. m. Distance biplot of PCA for the effects of Smarca4-astro-cKO and Stat3-astro-cKO on composite locomotor scores after SCI. The locations of values for each individual locomotor parameter, from either the longitudinal observer scored Open Field (OF) evaluations or ladder walk testing at 28 days after SCI, indicates their contributions to defining the PC space. Graph (mean + sem) shows composite SCI locomotor score derived from PCA of all OF and ladderwalk locomotor parameters recorded over 28 days of recovery; P values are cKO versus WT, one-way ANOVA with Bonferroni’s test, n = 11 mice per group. n. Immunofluorescence images and graph (mean + sem) of staining intensity show effects of Smarca4-astro-cKO and Stat3-astro-cKO on CD68 and Gfap 28 days after SCI. o. Bar graph (mean + sem) shows composite histology scores derived from PCA of Gfap, CD13, MBP and CD68 quantifications at 28 days after SCI; P values are cKO versus WT, one-way ANOVA with Bonferroni’s test, WT n = 8, Smarca4cKO n = 7, Stat3cKO n = 9 mice. Line graph shows correlation analysis of composite locomotor and histological scores. p. DEG-associated functional signalling pathway analysis shows that deletion of either Smarca4 or Stat3 alters many functions normally associated with WT astrocyte reactivity after LPS or SCI. Each immunohistochemical or in situ hybridization evaluation shown (c,e,f,j,l,n) was repeated at least three times with similar results.

Source data

Extended Data Fig. 7 DEGs and TRs in reactive astrocytes compared with astrocytes in healthy tissue.

a,b. Venn diagrams and graph compare proportions of astrocyte reactivity DEGs that derive from genes that either are, or are not, expressed in healthy astrocytes in EAE, LPS or SCI. c-f. Heatmap, Venn digrams and graphs compare proportion of TREA-predicted reactivity TRs that derive from TRs that either are, or are not, predicted in healthy astrocytes in EAE, LPS or SCI. In all three disorders much higher proportions of TRs than DEGs were recruited from those not detectably active in healthy tissue. Notably, of the TRs shared by all three disorders 87% were derived from TRs already active in healthy states, whereas 70%, 76% and 54% of disorder unique TRs respectively in EAE, SCI or LPS, were derived from TRs not detectably active in healthy states.

Source data

Extended Data Fig. 8 Comparison of DEGs and TRs across disorders.

8. a. Heatmap and graph show that in every disorder examined, TRs overlapped more with other disorders than did DEGs. b. Bar graphs show numbers of astroctye DEGs or TRs across all 15 disorders and conditions examined here. c. 3D distance plots of PCA for DEGs and TRs from disorders with WT astrocytes. Graphs on left show PCA of all 8 disorders, and on right show PCA of 5 disorders with three most divergent disorders removed. d. TRs predicted in three or more of six neurodegenerative disorders with genetic mutations or polymorphisms, compared with their predictions in disorders with WT astrocytes. No TRs are unique to multiple neurodegenerative disorders. e. DEGs upregulated in three or more of six neurodegenerative disorders with genetic mutations or polymorphisms, compared with their expression levels in disorders with WT astrocytes. No DEGs are unique to multiple neurodegenerative disorders.

Extended Data Fig. 9 Schematic of a working model of astrocyte reactivity transcriptional regulation.

Divergent non-cell autonomous, disorder-selective reactivity triggers lead to context-dependent and combinatorial TR interactions in which a core set of TRs (TRA & TRB) is active across many if not most forms of astrocyte reactivity. These core TRs can nevertheless regulate different cohorts of DEGs in different disorders and contexts via complex and interdependent combinatorial interactions that also involve disorder- or context-selective TRs (TRC - TRn). These TR interactions can be influenced by astrocyte cell autonomous factors such as mutations, polymorphisms, or by differing basal starting conditions that can vary with regional or local astrocyte heterogeneity or with other factors such as exposure to previous insults. These TR interactions give rise to the exquisitely heterogenous DEG profiles associated with different astrocyte reactivity states in different disorders and different contexts.

Source data

Extended Data Table 1 RNA Archival data sources

Supplementary information

Source data

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Burda, J.E., O’Shea, T.M., Ao, Y. et al. Divergent transcriptional regulation of astrocyte reactivity across disorders. Nature 606, 557–564 (2022). https://doi.org/10.1038/s41586-022-04739-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-022-04739-5

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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