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:

Astrocytes mediate the effect of oxytocin in the central amygdala on neuronal activity and affective states in rodents

Abstract

Oxytocin (OT) orchestrates social and emotional behaviors through modulation of neural circuits. In the central amygdala, the release of OT modulates inhibitory circuits and, thereby, suppresses fear responses and decreases anxiety levels. Using astrocyte-specific gain and loss of function and pharmacological approaches, we demonstrate that a morphologically distinct subpopulation of astrocytes expresses OT receptors and mediates anxiolytic and positive reinforcement effects of OT in the central amygdala of mice and rats. The involvement of astrocytes in OT signaling challenges the long-held dogma that OT acts exclusively on neurons and highlights astrocytes as essential components for modulation of emotional states under normal and chronic pain conditions.

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: Specific CeL astrocytes express OTRs.
Fig. 2: Astrocytic OTR activation evokes calcium transients in CeL astrocytes of rats and mice.
Fig. 3: OTR-expressing CeL astrocytes are positioned to recruit a CeL astrocyte network through gap junction.
Fig. 4: CeL astrocyte activity promotes excitatory transmission into CeL neurons.
Fig. 5: CeL astrocyte activity promotes CeL neuron firing.
Fig. 6: Astrocyte-driven CeL neuron activity modifies amygdala output.
Fig. 7: CeL astrocytes modulate CeA behavioral correlates of comfort and are required for their OTR-mediated modulation.

Similar content being viewed by others

Data and code availability

Python code (used for ex vivo calcium imaging data analysis) can be found in Supplementary Software. All data that support the findings of this study are available from the corresponding authors upon reasonable request.

References

  1. Lee, H.-J., Macbeth, A. H., Pagani, J. H. & Young, W. S. Oxytocin: the great facilitator of life. Prog. Neurobiol. 88, 127–151 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Knobloch, H. S. et al. Evoked axonal oxytocin release in the central amygdala attenuates fear response. Neuron 73, 553–566 (2012).

    Article  CAS  PubMed  Google Scholar 

  3. Chini, B., Verhage, M. & Grinevich, V. The action radius of oxytocin release in the mammalian CNS: from single vesicles to behavior. Trends Pharmacol. Sci. 38, 982–991 (2017).

    Article  CAS  PubMed  Google Scholar 

  4. Mitre, M. et al. A distributed network for social cognition enriched for oxytocin receptors. J. Neurosci. 36, 2517–2535 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Yoshida, M. et al. Evidence that oxytocin exerts anxiolytic effects via oxytocin receptor expressed in serotonergic neurons in mice. J. Neurosci. 29, 2259–2271 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Theodosis, D. T. Oxytocin-secreting neurons: a physiological model of morphological neuronal and glial plasticity in the adult hypothalamus. Front. Neuroendocrinol. 23, 101–135 (2002).

    Article  CAS  PubMed  Google Scholar 

  7. Pacholko, A. G., Wotton, C. A. & Bekar, L. K. Astrocytes—the ultimate effectors of long-range neuromodulatory networks? Front. Cell. Neurosci. 14, 1–12 (2020).

    Article  Google Scholar 

  8. Huber, D., Veinante, P. & Stoop, R. Vasopressin and oxytocin excite distinct neuronal populations in the central amygdala. Science 308, 245–248 (2005).

    Article  CAS  PubMed  Google Scholar 

  9. Viviani, D. et al. Oxytocin selectively gates fear responses through distinct outputs from the central amygdala. Science 333, 104–107 (2011).

    Article  CAS  PubMed  Google Scholar 

  10. Han, R. T. et al. Long-term isolation elicits depression and anxiety-related behaviors by reducing oxytocin-induced GABAergic transmission in central amygdala. Front. Mol. Neurosci. 11, 246 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  11. Neugebauer, V., Li, W., Bird, G. C. & Han, J. S. The amygdala and persistent pain. Neuroscientist 10, 221–234 (2004).

    Article  PubMed  Google Scholar 

  12. Tye, K. M. et al. Amygdala circuitry mediating reversible and bidirectional control of anxiety. Nature 471, 358–362 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Ji, R.-R., Donnelly, C. R. & Nedergaard, M. Astrocytes in chronic pain and itch. Nat. Rev. Neurosci. 20, 667–685 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Poisbeau, P., Grinevich, V. & Charlet, A. Oxytocin signaling in pain: cellular, circuit, system, and behavioral levels. Curr. Top. Behav. Neurosci. 35, 193–211 (2018).

  15. Martin-Fernandez, M. et al. Synapse-specific astrocyte gating of amygdala-related behavior. Nat. Neurosci. 20, 1540–1548 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Althammer, F. et al. Three-dimensional morphometric analysis reveals time-dependent structural changes in microglia and astrocytes in the central amygdala and hypothalamic paraventricular nucleus of heart failure rats. J. Neuroinflammation 17, 221 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Yizhar, O. et al. Neocortical excitation/inhibition balance in information processing and social dysfunction. Nature 477, 171–178 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Lee, H. J., Caldwell, H. K., Macbeth, A. H., Tolu, S. G. & Young, W. S. A conditional knockout mouse line of the oxytocin receptor. Endocrinology 149, 3256–3263 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Wallraff, A. et al. The impact of astrocytic gap junctional coupling on potassium buffering in the hippocampus. J. Neurosci. 26, 5438–5447 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Papouin, T., Dunphy, J. M., Tolman, M., Dineley, K. T. & Haydon, P. G. Septal cholinergic neuromodulation tunes the astrocyte-dependent gating of hippocampal NMDA receptors to wakefulness. Neuron 94, 840–854 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Robin, L. M. et al. Astroglial CB1 receptors determine synaptic D-serine availability to enable recognition memory. Neuron 98, 935–944 (2018).

    Article  CAS  PubMed  Google Scholar 

  22. Corkrum, M., Rothwell, P. E., Thomas, M. J., Kofuji, P. & Araque, A. Opioid-mediated astrocyte-neuron signaling in the nucleus accumbens. Cells 8, 586 (2019).

    Article  CAS  PubMed Central  Google Scholar 

  23. Li, Y. et al. Identity of endogenous NMDAR glycine site agonist in amygdala is determined by synaptic activity level. Nat. Commun. 4, 1760 (2013).

    Article  PubMed  Google Scholar 

  24. Decosterd, I. & Woolf, C. J. Spared nerve injury: an animal model of persistent peripheral neuropathic pain. Pain 87, 149–158 (2000).

    Article  PubMed  Google Scholar 

  25. Thompson, J. M. & Neugebauer, V. Amygdala plasticity and pain. Pain Res. Manag. 2017, 8296501 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Sieberg, C. B. et al. Neuropathic pain drives anxiety behavior in mice, results consistent with anxiety levels in diabetic neuropathy patients. Pain Rep. 3, e651 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  27. László, K. et al. Positive reinforcing effect of oxytocin microinjection in the rat central nucleus of amygdala. Behav. Brain Res. 296, 279–285 (2016).

    Article  PubMed  Google Scholar 

  28. Goodin, B. R., Ness, T. J. & Robbins, M. T. Oxytocin - a multifunctional analgesic for chronic deep tissue pain. Curr. Pharm. Des. 21, 906–913 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Gravati, M. et al. Dual modulation of inward rectifier potassium currents in olfactory neuronal cells by promiscuous G protein coupling of the oxytocin receptor. J. Neurochem. 114, 1424–1435 (2010).

    CAS  PubMed  Google Scholar 

  30. Breton, J.-D. et al. Oxytocin-induced antinociception in the spinal cord is mediated by a subpopulation of glutamatergic neurons in lamina I–II which amplify GABAergic inhibition. Mol. Pain 4, 19 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  31. Angulo, M. C., Kozlov, A. S., Charpak, S. & Audinat, E. Glutamate released from glial cells synchronizes neuronal activity in the hippocampus. J. Neurosci. 24, 6920–6927 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Fellin, T. et al. Neuronal synchrony mediated by astrocytic glutamate through activation of extrasynaptic NMDA receptors. Neuron 43, 729–743 (2004).

    Article  CAS  PubMed  Google Scholar 

  33. Corkrum, M. et al. Dopamine-evoked synaptic regulation in the nucleus accumbens requires astrocyte activity. Neuron 105, 1036–1047.e5 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Ma, Z., Stork, T., Bergles, D. E. & Freeman, M. R. Neuromodulators signal through astrocytes to alter neural circuit activity and behaviour. Nature 539, 428–432 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Pestana, F., Edwards-Faret, G., Belgard, T. G., Martirosyan, A. & Holt, M. G. No longer underappreciated: the emerging concept of astrocyte heterogeneity in neuroscience. Brain Sci. 10, 1–21 (2020).

    Article  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  37. Grinevich, V. & Neumann, I. D. Brain oxytocin: how puzzle stones from animal studies translate into psychiatry. Mol. Psychiatry 26, 265–279 (2021).

  38. Quintana, D. S. et al. Advances in the field of intranasal oxytocin research: lessons learned and future directions for clinical research. Mol. Psychiatry 26, 80–91 (2021).

  39. Kirsch, P. et al. Oxytocin modulates neural circuitry for social cognition and fear in humans. J. Neurosci. 25, 11489–11493 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Takata, N. et al. Optogenetic astrocyte activation evokes BOLD fMRI response with oxygen consumption without neuronal activity modulation. Glia 66, 2013–2023 (2018).

  41. Figley, C. R. & Stroman, P. W. The role(s) of astrocytes and astrocyte activity in neurometabolism, neurovascular coupling, and the production of functional neuroimaging signals. Eur. J. Neurosci. 33, 577–588 (2011).

    Article  PubMed  Google Scholar 

  42. Kastanenka, K. V. et al. A roadmap to integrate astrocytes into systems neuroscience. Glia 68, 5–26 (2019).

  43. Poskanzer, K. E. & Yuste, R. Astrocytes regulate cortical state switching in vivo. Proc. Natl Acad. Sci. USA 2016, 1–10 (2016).

    Google Scholar 

  44. Kjaerby, C., Rasmussen, R., Andersen, M. & Nedergaard, M. Does global astrocytic calcium signaling participate in awake brain state transitions and neuronal circuit function? Neurochem. Res. 42, 1810–1822 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Fenoy, A. J., Goetz, L., Chabardès, S. & Xia, Y. Deep brain stimulation: are astrocytes a key driver behind the scene? CNS Neurosci. Ther. 20, 191–201 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Monai, H. & Hirase, H. Astrocytes as a target of transcranial direct current stimulation (tDCS) to treat depression. Neurosci. Res. 126, 15–21 (2018).

    Article  PubMed  Google Scholar 

  47. Mills, S. E. E., Nicolson, K. P. & Smith, B. H. Chronic pain: a review of its epidemiology and associated factors in population-based studies. Br. J. Anaesth. 123, e273–e283 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  48. Woo, A. K. Depression and anxiety in pain. Rev. Pain 4, 8–12 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Oliveira, J. F., Sardinha, V. M., Guerra-Gomes, S., Araque, A. & Sousa, N. Do stars govern our actions? Astrocyte involvement in rodent behavior. Trends Neurosci. 38, 535–549 (2015).

    Article  CAS  PubMed  Google Scholar 

  50. Rouach, N., Koulakoff, A., Abudara, V., Willecke, K. & Giaume, C. Astroglial metabolic networks sustain hippocampal synaptic transmission. Science 322, 1551–1555 (2008).

    Article  CAS  PubMed  Google Scholar 

  51. Pannasch, U. et al. Astroglial networks scale synaptic activity and plasticity. Proc. Natl Acad. Sci. USA 108, 8467–8472 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Theis, M. et al. Accelerated hippocampal spreading depression and enhanced locomotory activity in mice with astrocyte-directed inactivation of connexin43. J. Neurosci. 23, 766–776 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Teubner, B. et al. Connexin30 (Gjb6)-deficiency causes severe hearing impairment and lack of endocochlear potential. Hum. Mol. Genet. 12, 13–21 (2003).

    Article  CAS  PubMed  Google Scholar 

  54. Shigetomi, E. et al. Imaging calcium microdomains within entire astrocyte territories and endfeet with GCaMPs expressed using adeno-associated viruses. J. Gen. Physiol. 141, 633–647 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Ting, J. T. et al. Preparation of acute brain slices using an optimized N-methyl-D-glucamine protective recovery method. J. Vis. Exp. (132), 53825 (2018).

  56. Serrano, A., Haddjeri, N., Lacaille, J., Robitaille, R. & Centre-ville, S. GABAergic network activation of glial cells underlies hippocampal heterosynaptic depression. J. Neurosci. 26, 5370–5382 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Jourdain, P. et al. Glutamate exocytosis from astrocytes controls synaptic strength. Nat. Neurosci. 10, 331–339 (2007).

    Article  CAS  PubMed  Google Scholar 

  58. Ikegaya, Y., Le Bon-Jego, M. & Yuste, R. Large-scale imaging of cortical network activity with calcium indicators. Neurosci. Res. 52, 132–138 (2005).

    Article  CAS  PubMed  Google Scholar 

  59. Shigetomi, E., Bowser, D. N., Sofroniew, M. V. & Khakh, B. S. Two forms of astrocyte calcium excitability have distinct effects on NMDA receptor-mediated slow inward currents in pyramidal neurons. J. Neurosci. 28, 6659–6663 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Anlauf, E. & Derouiche, A. Glutamine synthetase as an astrocytic marker: its cell type and vesicle localization. Front. Endocrinol. 4, 144 (2013).

    Article  Google Scholar 

  61. Luis-Delgado, O. E. et al. Calibrated forceps: a sensitive and reliable tool for pain and analgesia studies. J. Pain 7, 32–39 (2006).

    Article  PubMed  Google Scholar 

  62. Walf, A. A. & Frye, C. A. The use of the elevated plus maze as an assay of anxiety-related behavior in rodents. Nat. Protoc. 2, 322–328 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. King, T. et al. Unmasking the tonic-aversive state in neuropathic pain. Nat. Neurosci. 12, 1364–1366 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Virtanen, P. et al. SciPy 1.0: fundamental algorithms for scientific computing in Python. Nat. Methods 17, 261–272 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Zeisel, A. et al. Cell types in the mouse cortex and hippocampus revealed by single-cell RNA-seq. Science 347, 1138–1142 (2015).

    Article  CAS  PubMed  Google Scholar 

  66. Batiuk, M. Y. et al. Identification of region-specific astrocyte subtypes at single cell resolution. Nat. Commun. 11, 1220 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Kofuji, P. & Araque, A. G-protein-coupled receptors in astrocyte–neuron communication. Neuroscience https://doi.org/10.1016/j.neuroscience.2020.03.025 (2020).

  68. Octeau, J. C. et al. Transient, consequential increases in extracellular potassium ions accompany channelrhodopsin2 excitation. Cell Rep. 27, 2249–2261 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Reeves, A. M. B., Shigetomi, E. & Khakh, B. S. Bulk loading of calcium indicator dyes to study astrocyte physiology: key limitations and improvements using morphological maps. J. Neurosci. 31, 9353–9358 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by the Centre National de la Recherche Scientifique contract UPR3212, the Université de Strasbourg contract UPR3212, IASP Early Career Research grant 2012, FP7 Career Integration grant 334455, Initiative of Excellence (IDEX) Attractiveness grant 2013, IDEX Interdisciplinary grant 2015, University of Strasbourg Institute for Advanced Study (USIAS) fellowship 2014-15, Foundation Fyssen research grant 2015, NARSAD Young Investigator grant 24821, ANR JCJC grant 19-CE16-0011-0 (to A.C.), ANR-DFG grant GR 3619/701 (to A.C. and V.G.), Alexander von Humboldt fellowship (to D.H.), Seed grant from DFG within the Collaborative Center SFB 1158 and Fyssen Foundation fellowship (to A.L.); DFG grants GR 3619/13-1, GR 3619/15-1 and GR 3619/16-1 (to V.G.); SFB 1158 (to C.P., J.S. and V.G.), SNSF-DFG grant GR 3619/8-1 (to R.S. and V.G.), the Fritz Thyssen Foundation (to V.G.), DFG Postdoctoral Fellowship AL 2466/1-1 (to F.A.), the Alexander von Humboldt Foundation (to D.H.), the Fyssen Foundation and PROCOP grant and SFB1158 seed grant for young scientists (to A.L.), post-doctoral fellowship from the Fund for Scientific Research Flanders (12V7519N) (to J.W.), Russian Science Foundation RSF (17-75-10061) and the subsidy allocated to the Kazan Federal University for the State Assignment, no. 0671-2020-0059 (to A.R.), the intramural research program of the NIMH (ZIAMH002498) (to W.S.Y.), National Institutes of Health grants R01NS094640 and R01HL090948, SFB/TRR 152 and ERC-CoG-772395 (to J.E.S.), the European Research Council (Consolidator Grant no. 683154) and the European Union’s Horizon 2020 Research and Innovation Program (Marie Sklodowska-Curie Innovative Training Networks, grant no. 722053, EU-GliaPhD) (to N.R.). The authors thank V. Lelièvre for in situ hybridization advice; R. Goutagny for in vivo optogenetics assistance; F. Magara for anxiety behavior advice; B. Kurpiers and the Interdisciplinary Neurobehavioral Core Facility of Heidelberg University for experiments performed there; S. Reibel and the Chronobiotron UMS 3415 for all animal care; and T. Splettstoesser (www.scistyle.com) for help with the preparation of Fig. 7.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: A.C.; Methodology: A.C., B.Boutrel, C.M.L., C.P., D.K., F.A., I.D., J.E.S., J.W., J.Y.C., N.R., P.D., P.P., R.S., V.G., W.S.Y. and YG; Analysis: A.C., B.Bellanger, B.B.J., C.M.L., D.K., F.A., H.S.K.B., J.W. and S.G.; In situ hybridization: D.H., F.A., H.S.K.B., H.W., J.S. and M.E.; Immunohistochemistry: A.L., D.H., F.A., J.W., M.E. and M.D.S.G.; Imaris analysis: F.A., M.K.K. and R.K.R.; Ex vivo patch-clamp electrophysiology: A.B., A.C., J.W., S.G., D.K., I.W., B.Bellanger and M.A.; Ex vivo calcium imaging: A.B., C.M.L., D.K. and J.W.; Astrocyte characterization: A.B., A.R., B.Bellanger, D.K., I.W., M.E. and S.G.; Behavior: A.C., B.B.J., D.K. and J.W.; Mice line validation: W.S.Y.; Viral vectors validation: D.B., L.B., M.D.S.G., M.E. and V.G.; Spared nerve injuries: P.I. and M.P.; Writing: A.B., A.C., D.K., F.A., J.W. and V.G.; Funding acquisition: A.C. and V.G.; Supervision: A.C. and V.G.; Project administration: A.C.

Corresponding authors

Correspondence to Valery Grinevich or Alexandre Charlet.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Neuroscience thanks Thomas Papouin, Gertrudis Perea, and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Validation of OTR expression in CeL astrocytes using several astrocytic markers.

a, Combination of FISH and IHC showing co-localization of GS (top left), NeuN (bottom left) and both markers (right) with OTR mRNA in rats. Scale bars are 100µm and 10µm. (n rats = 5, n slices = 20, n astrocytes = 1354, n neurons = 1254). b, proportion of OTR positive astrocytes and neurons in mice. n astrocytes = 897; n neurons = 688; n rats = 4 c, (top left) FISH overview for OTR mRNA (green), counterstained with polyclonal anti-ALDH1L1 antibody (red) in CeA. (bottom left) High magnification image of cells positive for both OTR mRNA and ALDH1L1 (double arrows); green arrows point OTR mRNA-positive cells; red arrows point ALDH1L1-positive cells. Scale bars: 400 (Top) and 50µm (Bottom). (Right) Quantification of ALDH1L1-positive cells positive for OTR mRNA. n astrocytes = 450; n rats = 4. d, RNAscope in situ hybridization showing GFAP (red) and OTR (green) expressing cells in mice CeA. Merged images include DAPI stain (blue); n = 3. (Bottom) Negative control probe targeting the bacterial gene DapB; n = 3. Scale bar is 10μm. st: stria terminalis. Data are expressed as mean across slices plus SEM and white circles represent individual cell data. (Statistics and numbers in Supplementary Table 1).

Extended Data Fig. 2 Validation of astrocyte-specific approaches.

a, (left) C1V1-mCherry (OxyOpto) expressing oxytocin neuron of the PVN. Scale bar is 20µm; n = 11. (right) Schematic representation of the whole cell patch clamp of the OxytOpto-expressing cells of the PVN. b, (left) λ542nm light exposure (yellow line) induce a depolarizing current, (right) enabling precise spiking of OxytOpto-expressing cells. c, Typical confocal image of CeL astrocytes co-labeled with SR101 and OGB1; n = 46. Scale bar 20μm. d, (left) Pseudo-color pictures of an SR101 positive cell identified as an astrocyte compared to neurons identified under oblique infrared light, Scale bar 10μm. (Middle) Electrophysiological properties of patched SR101+ (red, n = 82) and SR101- cells (grey, n = 20). Medians, quartiles and whiskers can be found in Supplementary Table 2. (Right) Typical responses to 20pA current steps of a SR101+ (red) and a SR101- cells (black). e, CeL SR101 positive cell filled with biocytin through whole cell patch-clamp (green) lacks NeuN signal (Blue); n = 3. Scale bar 50 μm. f, Illustration of calcium imaging data analysis method. (left) Fluorescence traces were splited into a “baseline” and a “post-stimulation” trace. Calcium transients were detected according to a prominence threshold and their frequency was quantified before and after the drug application. All data were averaged across astrocytes per slice, and this results was used as statistical unit. All data were compared (before vs. after drug application) and the results were expressed in ratio (baseline/drug effect), a ratio of 1 meaning neither an increase nor a decrease of the measured parameter. ns = 7, na = 36. g, (top left) Immunohistochemical staining for glutamine synthase (GS; blue), Cre-GFP (green) and Cre recombinase (red). (bottom left) Overview of the CeA, displaying correct viral targeting of the CeL subdivision. (right) Quantification of efficiency and specificity of the transduction. Efficiency: Over 1001 GS positive cells, 561 were also GFP positive, indicating an efficiency of 56±4.9%. Specificity: we counted a total of n = 977 GFP-positive cells, 940 of which were positive for GS (96.2±2.1%). None of the GFP or Cre signals were detected in NeuN positive cells (0 out of n = 850, 4 mice, not shown). Finally, 99.82 ± 0.2% of GS-positive astrocytes containing GFP signal were Cre-positive (n = 1001). n mice = 4. Scale bars: 400 (Bottom) and 20µm (Top). h, Immunohistochemical analysis of OTR expression in NeuN-positive cells of the CeL revealed no difference in OTR levels between control and GFAP-Cre injected animals; 85.0 ± 1.6% (n = 688, n = 4 mice) and 86.7 ± 1.6% (n = 660, n = 4 mice) respectively. Scale bars: 20µm. Data in f-h are expressed as mean across slices plus SEM and white circles represent individual cell data. ***p < 0.001, two-sided Mann-Whitney U test. (Statistics and numbers in Supplementary Table 2).

Extended Data Fig. 3 Astro-astrocytic communication in the CeL do not depends on purinergic nor NMDAR-dependent signalling.

a, Illustration of the 3D reconstruction and the quantification of the number of contact between astrocytes. (left) Pictures of the raw fluorescence obtained after anti-GFAP immunohistochemistry, (middle) based on this staining, we performed a 3D reconstruction of astrocytes and (right) evaluate the number of astrocyte-astrocyte contact using the given formula. With n = upper limit of summation, i = index of summation, xi = typical element (that is astrocytic contacts with 1µm or less proximity). With this approach, we quantified all astrocyte contacts without discrimination between individual astrocyte entities. b, (left) Typical ΔF traces following TGOT+TTX application in presence of PPADS (50 µM) + Suramin (75 µM) + A438079 (1 µM). (right) Proportion of responding astrocytes, AUC of ΔF traces and Ca2+ transients frequency normalized to baseline values following application of TGOT+TTX (0.4 μM) in presence of PPADS (50 µM) + Suramin (75 µM) + A438079 (1 µM) (ns = 10, na = 112). c, (left) Typical ΔF traces following TGOT+TTX application in presence of CPT (5 µM) or SCH 58261 (100 nM). (right) Proportion of responding astrocytes, AUC of ΔF traces and Ca2+ transients frequency normalized to baseline values following application of TGOT+TTX (0.4 μM) in presence of CPT (ns = 7, na = 22) or SCH (ns = 8, na = 32). d, (left) Typical ΔF traces following TGOT+TTX application in presence of AP5 (50 µM). (right) Proportion of responding astrocytes, AUC of ΔF traces and Ca2+ transients frequency normalized to baseline values following application of TGOT+TTX (0.4 μM) in presence of AP5 (ns = 6, na = 55). Data are expressed as mean across slices plus SEM and white circles represent individual cell data. White circles indicate average across astrocytes per slice. (Statistics and numbers in Supplementary Table 3).

Extended Data Fig. 4 Specific optogenetic stimulation of CeL astrocytes elicit calcium transients in a Ca2+-dependent manner.

a, (left) Immunohistochemistry image shows CeL cells transfected with rAAV-Gfap-C1V1(t/t)-mCherry (AstrOpto) with co-labeling for ALDH1L1. White arrow shows one cell expanded in insets. Scale bars are 25 and 10 µm (insets). (right) Quantification of the efficiency and specificity of transduction of C1V1 in CeL astrocytes. PGFAP- C1V1-mCherry rAAV vector was injected into rat CeA (bilaterally, 200nl). Specificity: Over 1090 mCherry-positive cells, 98.8±0.7% were positive for ALDH1L1. None of the analyzed cells were positive for NeuN. Efficiency: Over 1090 ALDH1L1 positive cells, 62.3±3.5% were also mCherry positive. n = 4 rats. b, (left) Typical ΔF traces following AstrOpto activation. λ542nm (1s @ 0.5Hz during 3 minutes) with (top 3 traces) or without (bottom 3 traces) extracellular calcium. (right) Proportion of responding astrocytes, AUC of ΔF traces and Ca2+ transients frequency normalized to baseline values following AstrOpto stimulation with (ns = 12, na = 49) or without Ca2+ in the ACSF (ns = 12, na = 53). Data are expressed as mean across slices plus SEM. White circles indicate average across astrocytes per slice. **p<0.01, ***p<0.001, two-sided Mann-Whitney U test. (Statistics and numbers in Supplementary Table 4).

Extended Data Fig. 5 Astro-neuronal communication depends on NMDAR but not purinergic transmission.

Figure 5. a, DAAO (0.15 IU/ml, incubation time > 1h30; n = 7) and AP5 (50 µM; n = 6) prevent the effect of photoactivation of astrocytes (AstrOpto) on IPSC frequency in CeM neurons. b, Two consecutive TGOT application effects on IPSCs frequencies in CeM neurons after DAAO (0.15 U/ml, incubation time > 1h30) followed by D-Serine (20 min, 100 μM) incubation of the same cells (n = 7). c, dOVT (1 µM; n = 6) prevents the effect of TGOT on IPSCs frequencies in CeM neurons. d, Effect of double (20 min apart) application of TGOT on IPSCs frequencies in CeM (0.4 µM n = 7). e, Effect of DCKA (10 µM, n = 15) and DNQX (25 µM; n = 10) on TGOT-induced increase in IPSC frequency in CeM neurons. f, Purinergic antagonists do not prevent TGOT effect on IPSC frequency in CeM neurons. PPADS (50 µM; n = 9), CPT (5 µM; n = 11), SCH (100 nM; n = 10). Data are expressed as averages plus SEM and white circles represent individual cell data. #p < 0.05, ##p < 0.01, ###p < 0.001, Friedman and Dunn’s Multiple comparisons, **p < 0.01, ***p < 0.001, two-sided unpaired t-test or Mann-Whitney U test. (Statistics and numbers in Supplementary Table 5).

Extended Data Fig. 6 SNI procedure does not modify the effect of OTR activation on CeL microcircuit activity.

a, 30 days post surgeries time course of mechanical pain threshold evolution across sham (n = 23) and SNI (n = 22) rats. Data are expressed as mean across animals plus SEM. b, (left) Typical ΔF traces following TGOT+TTX application in SNI rats. (right) Proportion of responding astrocytes, AUC of ΔF traces and Ca2+ transients frequency normalized to baseline values following application of TGOT+TTX (0.4 μM) in sham (ns = 16, na = 74) or SNI (ns = 18 et na = 136) rats. Data are expressed as mean across slices plus SEM. White circles indicate average across astrocytes per slice. c, TGOT effect on CeM neurons IPSCs frequencies is unchanged between Sham (n = 9) or SNI (n = 9) rats. Data are expressed as average plus SEM and white circles represent individual cell data. d, Two consecutive TGOT application effects on IPSCs frequencies in CeM neurons after DAAO (0.15 U/ml, incubation time > 1h30) followed by D-Serine (20 min, 100 μM) incubation of the same cells from acute brain slices of SNI animals (n = 11). Data are expressed as averages plus SEM and white circles represent individual cell data. e, Verification of cannulae implantation sites. (left) Schematic representation of the stereotaxic injection viewed in a coronal slice; n rats = 67, n mice = 52. (middle, right) Infrared and fluorescence microscopy picture showing the injection site of fluorescent latex beads in the CeL on horizontal slices. Scale is 1mm. ** p<0.01, *** P<0.001, two-sided paired t-test or Wilcoxon test, #p<0.05 Friedman test followed by Dunn’s multiple comparison. (Statistics and numbers in Supplementary Table 6).

Extended Data Fig. 7 Effects of CeL astrocytes and OTR activities manipulations on locomotion and contralateral hindpaw sensitivity.

a,e, Experimental strategy for the specific expression of C1V1 in mice CeL astrocytes (a, AstrOpto) or the specific deletion of OTRs in mice CeL astrocytes (e, GFAP OTR KO). The treatments applied are color coded as the legend key indicate. Control indicate a vehicle injection. b,f, Mechanical pain threshold was assessed on the non-injured paw of SNI (bottom, gray box) and its equivalent in Sham (top) groups. TGOT or its vehicle, or astrocytes light-evoked activation of C1V1, were administered in the CeL and mechanical pain threshold assessed again at different time points. Rats (n): Sham control n = 12, TGOT n = 10, C1V1 n = 6; SNI control n = 12, TGOT n = 10, C1V1, n = 9. Mice (n): Sham WT n = 6, OTR cKO n = 8; SNI WT n = 5, OTR cKO n = 13. c,g, Locomotion was assessed through measurement of the distance travelled during the length of the elevated plus maze experiment, after administration of the different treatments. Rats (n): Sham control n = 10, TGOT n = 8, C1V1 n = 9; SNI control n = 20, TGOT n = 9, C1V1, n = 7. Mice (n): Sham WT Veh n = 7, WT TGOT n = 6; Sham OTR cKO Veh n = 5, OTR cKO TGOT n = 7; SNI WT Veh n = 7, WT TGOT n = 8; SNI OTR cKO Veh n = 5, OTR cKO TGOT n = 7. d,h, Locomotion was assessed through measurement of the distance travelled during the time of the conditioned place preference experiment, after administration of the different treatments. Rats (n): Sham control n = 13, TGOT n = 6, C1V1 n = 5; SNI control n = 10, TGOT n = 8, C1V1, n = 8. Mice (n): Sham WT Veh n = 5, WT TGOT n = 8; Sham OTR cKO Veh n = 5, OTR cKO TGOT n = 4; SNI WT Veh n = 5, WT TGOT n = 5; SNI OTR cKO Veh n = 4, OTR cKO TGOT n = 6. Data are expressed as mean across animals plus SEM. (Animals number and Statistics in Supplementary Table 7).

Supplementary information

Reporting Summary

Supplementary Tables

Tables 1–7: Numerical values and statistical analysis of data presented in Figs.1–7 and Extended Data Figs. 1–7. Table 8: List of reagents used in this study

Supplementary Software

Calcium imaging analysis using Python code

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wahis, J., Baudon, A., Althammer, F. et al. Astrocytes mediate the effect of oxytocin in the central amygdala on neuronal activity and affective states in rodents. Nat Neurosci 24, 529–541 (2021). https://doi.org/10.1038/s41593-021-00800-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-021-00800-0

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing