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:

Spinal astrocytes in superficial laminae gate brainstem descending control of mechanosensory hypersensitivity

Abstract

Astrocytes are critical regulators of CNS function and are proposed to be heterogeneous in the developing brain and spinal cord. Here we identify a population of astrocytes located in the superficial laminae of the spinal dorsal horn (SDH) in adults that is genetically defined by Hes5. In vivo imaging revealed that noxious stimulation by intraplantar capsaicin injection activated Hes5+ SDH astrocytes via α1A-adrenoceptors (α1A-ARs) through descending noradrenergic signaling from the locus coeruleus. Intrathecal norepinephrine induced mechanical pain hypersensitivity via α1A-ARs in Hes5+ astrocytes, and chemogenetic stimulation of Hes5+ SDH astrocytes was sufficient to produce the hypersensitivity. Furthermore, capsaicin-induced mechanical hypersensitivity was prevented by the inhibition of descending locus coeruleus–noradrenergic signaling onto Hes5+ astrocytes. Moreover, in a model of chronic pain, α1A-ARs in Hes5+ astrocytes were critical regulators for determining an analgesic effect of duloxetine. Our findings identify a superficial SDH-selective astrocyte population that gates descending noradrenergic control of mechanosensory behavior.

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: Identification of Hes5+ cells as a population of SDH astrocytes in adult mice.
Fig. 2: Ca2+ increases in supSDH astrocytes by noxious stimuli via descending NAergic signaling.
Fig. 3: Activation of the descending NAergic pathway elicits astrocytic [Ca2+]i increases in the supSDH.
Fig. 4: Activation of α1A-ARs in Hes5+ astrocytes induces mechanical hypersensitivity.
Fig. 5: Chemogenetic stimulation of Hes5+ SDH astrocytes induces mechanical hypersensitivity.
Fig. 6: d-serine is involved in Hes5+ SDH astrocyte-mediated mechanical hypersensitivity.
Fig. 7: Hes5+ SDH astrocytes contribute to capsaicin-induced hypersensitivity via descending NAergic signaling.
Fig. 8: Conditional α1A-AR knockout in Hes5+ astrocytes enhances the analgesic effect of duloxetine on neuropathic mechanical hypersensitivity.

Similar content being viewed by others

Data availability

The data of the current study are presented in the figures. If necessary, the data that support the findings of this study are available from the corresponding author upon reasonable request. Source data are provided with this paper.

Code availability

No code was used for the study.

References

  1. Eroglu, C. & Barres, B. A. Regulation of synaptic connectivity by glia. Nature 468, 223–231 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Haydon, P. G. & Nedergaard, M. How do astrocytes participate in neural plasticity? Cold Spring Harb. Perspect. Biol. 7, a020438 (2014).

    PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  4. Bayraktar, O. A., Fuentealba, L. C., Alvarez-Buylla, A. & Rowitch, D. H. Astrocyte development and heterogeneity. Cold Spring Harb. Perspect. Biol. 7, a020362 (2014).

    PubMed  Google Scholar 

  5. Tsai, H. H. et al. Regional astrocyte allocation regulates CNS synaptogenesis and repair. Science 337, 358–362 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Vainchtein, I. D. et al. Astrocyte-derived interleukin-33 promotes microglial synapse engulfment and neural circuit development. Science 359, 1269–1273 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Molofsky, A. V. et al. Astrocyte-encoded positional cues maintain sensorimotor circuit integrity. Nature 509, 189–194 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Zhang, Y. et al. An RNA-sequencing transcriptome and splicing database of glia, neurons, and vascular cells of the cerebral cortex. J. Neurosci. 34, 11929–11947 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Ohtsuka, T. et al. Hes1 and Hes5 as Notch effectors in mammalian neuronal differentiation. EMBO J. 18, 2196–2207 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Kondo, T. & Raff, M. Basic helix–loop–helix proteins and the timing of oligodendrocyte differentiation. Development 127, 2989–2998 (2000).

    CAS  PubMed  Google Scholar 

  11. Lugert, S. et al. Quiescent and active hippocampal neural stem cells with distinct morphologies respond selectively to physiological and pathological stimuli and aging. Cell Stem Cell 6, 445–456 (2010).

    CAS  PubMed  Google Scholar 

  12. Basbaum, A. I., Bautista, D. M., Scherrer, G. & Julius, D. Cellular and molecular mechanisms of pain. Cell 139, 267–284 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Lee, Y., Messing, A., Su, M. & Brenner, M. GFAP promoter elements required for region-specific and astrocyte-specific expression. Glia 56, 481–493 (2008).

    PubMed  Google Scholar 

  14. Bruinstroop, E. et al. Spinal projections of the A5, A6 (locus coeruleus), and A7 noradrenergic cell groups in rats. J. Comp. Neurol. 520, 1985–2001 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Yoshimura, M. & Furue, H. Mechanisms for the anti-nociceptive actions of the descending noradrenergic and serotonergic systems in the spinal cord. J. Pharmacol. Sci. 101, 107–117 (2006).

    CAS  PubMed  Google Scholar 

  16. Mandela, P. & Ordway, G. A. The norepinephrine transporter and its regulation. J. Neurochem. 97, 310–333 (2006).

    CAS  PubMed  Google Scholar 

  17. Howorth, P. W., Teschemacher, A. G. & Pickering, A. E. Retrograde adenoviral vector targeting of nociresponsive pontospinal noradrenergic neurons in the rat in vivo. J. Comp. Neurol. 512, 141–157 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Mulvey, B. et al. Molecular and functional sex differences of noradrenergic neurons in the mouse locus coeruleus. Cell Rep. 23, 2225–2235 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Tervo, D. G. et al. A designer AAV variant permits efficient retrograde access to projection neurons. Neuron 92, 372–382 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Ji, R. R., Baba, H., Brenner, G. J. & Woolf, C. J. Nociceptive-specific activation of ERK in spinal neurons contributes to pain hypersensitivity. Nat. Neurosci. 2, 1114–1119 (1999).

    CAS  PubMed  Google Scholar 

  21. Ross, S. B. & Stenfors, C. DSP4, a selective neurotoxin for the locus coeruleus noradrenergic system. A review of its mode of action. Neurotox. Res. 27, 15–30 (2015).

    CAS  PubMed  Google Scholar 

  22. Roth, B. L. DREADDs for neuroscientists. Neuron 89, 683–694 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Matsumoto, M., Xie, W., Ma, L. & Ueda, H. Pharmacological switch in Aβ-fiber stimulation-induced spinal transmission in mice with partial sciatic nerve injury. Mol. Pain 4, 25 (2008).

    PubMed  PubMed Central  Google Scholar 

  24. Takata, N. et al. Astrocyte calcium signaling transforms cholinergic modulation to cortical plasticity in vivo. J. Neurosci. 31, 18155–18165 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Mothet, J. P. et al. Glutamate receptor activation triggers a calcium-dependent and SNARE protein-dependent release of the gliotransmitter d-serine. Proc. Natl Acad. Sci. USA 102, 5606–5611 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Tsuda, M. et al. JAK–STAT3 pathway regulates spinal astrocyte proliferation and neuropathic pain maintenance in rats. Brain 134, 1127–1139 (2011).

    PubMed  PubMed Central  Google Scholar 

  27. Kohro, Y. et al. A new minimally-invasive method for microinjection into the mouse spinal dorsal horn. Sci. Rep. 5, 14306 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Obata, H. Analgesic mechanisms of antidepressants for neuropathic pain. Int. J. Mol. Sci. 18, 2483 (2017).

    PubMed Central  Google Scholar 

  29. Kremer, M. et al. A dual noradrenergic mechanism for the relief of neuropathic allodynia by the antidepressant drugs duloxetine and amitriptyline. J. Neurosci. 38, 9934–9954 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Moehring, F., Halder, P., Seal, R. P. & Stucky, C. L. Uncovering the cells and circuits of touch in normal and pathological settings. Neuron 100, 349–360 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Koch, S. C., Acton, D. & Goulding, M. Spinal circuits for touch, pain, and itch. Annu. Rev. Physiol. 80, 189–217 (2018).

    CAS  PubMed  Google Scholar 

  32. Martin, R., Bajo-Graneras, R., Moratalla, R., Perea, G. & Araque, A. Circuit-specific signaling in astrocyte–neuron networks in basal ganglia pathways. Science 349, 730–734 (2015).

    CAS  PubMed  Google Scholar 

  33. Henneberger, C., Papouin, T., Oliet, S. H. & Rusakov, D. A. Long-term potentiation depends on release of d-serine from astrocytes. Nature 463, 232–236 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Pankratov, Y. & Lalo, U. Role for astroglial α1-adrenoreceptors in gliotransmission and control of synaptic plasticity in the neocortex. Front. Cell Neurosci. 9, 230 (2015).

    PubMed  PubMed Central  Google Scholar 

  35. Robin, L. M. et al. Astroglial CB1 receptors determine synaptic d-serine availability to enable recognition memory. Neuron 98, 935–944.e5 (2018).

    CAS  PubMed  Google Scholar 

  36. Foster, A. C. et al. d-serine is a substrate for neutral amino acid transporters ASCT1/SLC1A4 and ASCT2/SLC1A5, and is transported by both subtypes in rat hippocampal astrocyte cultures. PLoS ONE 11, e0156551 (2016).

    PubMed  PubMed Central  Google Scholar 

  37. Hayashi, F., Takahashi, K. & Nishikawa, T. Uptake of d- and l-serine in C6 glioma cells. Neurosci. Lett. 239, 85–88 (1997).

    CAS  PubMed  Google Scholar 

  38. Kaplan, E. et al. ASCT1 (Slc1a4) transporter is a physiologic regulator of brain d-serine and neurodevelopment. Proc. Natl Acad. Sci. USA 115, 9628–9633 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Kronschlager, M. T. et al. Gliogenic LTP spreads widely in nociceptive pathways. Science 354, 1144–1148 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Papouin, T. et al. Synaptic and extrasynaptic NMDA receptors are gated by different endogenous coagonists. Cell 150, 633–646 (2012).

    CAS  PubMed  Google Scholar 

  41. Kuraishi, Y., Harada, Y. & Takagi, H. Noradrenaline regulation of pain-transmission in the spinal cord mediated by α-adrenoceptors. Brain Res. 174, 333–336 (1979).

    CAS  PubMed  Google Scholar 

  42. Howe, J. R. & Yaksh, T. L. Changes in sensitivity to intrathecal norepinephrine and serotonin after 6-hydroxydopamine (6-OHDA), 5,6-dihydroxytryptamine (5,6-DHT) or repeated monoamine administration. J. Pharmacol. Exp. Ther. 220, 311–321 (1982).

    CAS  PubMed  Google Scholar 

  43. Sekiguchi, K. J. et al. Imaging large-scale cellular activity in spinal cord of freely behaving mice. Nat. Commun. 7, 11450 (2016).

    PubMed  PubMed Central  Google Scholar 

  44. O’Neill, J. et al. Unravelling the mystery of capsaicin: a tool to understand and treat pain. Pharmacol. Rev. 64, 939–971 (2012).

    PubMed  PubMed Central  Google Scholar 

  45. Dubin, A. E. & Patapoutian, A. Nociceptors: the sensors of the pain pathway. J. Clin. Invest. 120, 3760–3772 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Kuner, R. & Flor, H. Structural plasticity and reorganisation in chronic pain. Nat. Rev. Neurosci. 18, 20–30 (2016).

    PubMed  Google Scholar 

  47. Li, X., Conklin, D., Ma, W., Zhu, X. & Eisenach, J. C. Spinal noradrenergic activation mediates allodynia reduction from an allosteric adenosine modulator in a rat model of neuropathic pain. Pain 97, 117–125 (2002).

    CAS  PubMed  Google Scholar 

  48. Huang, J. et al. A neuronal circuit for activating descending modulation of neuropathic pain. Nat. Neurosci. 22, 1659–1668 (2019).

    CAS  PubMed  Google Scholar 

  49. Hughes, S. W., Hickey, L., Hulse, R. P., Lumb, B. M. & Pickering, A. E. Endogenous analgesic action of the pontospinal noradrenergic system spatially restricts and temporally delays the progression of neuropathic pain following tibial nerve injury. Pain 154, 1680–1690 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Viisanen, H. & Pertovaara, A. Influence of peripheral nerve injury on response properties of locus coeruleus neurons and coeruleospinal antinociception in the rat. Neuroscience 146, 1785–1794 (2007).

    CAS  PubMed  Google Scholar 

  51. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Madisen, L. et al. A robust and high-throughput Cre reporting and characterization system for the whole mouse brain. Nat. Neurosci. 13, 133–140 (2010).

    CAS  PubMed  Google Scholar 

  53. Futatsugi, A. et al. IP3 receptor types 2 and 3 mediate exocrine secretion underlying energy metabolism. Science 309, 2232–2234 (2005).

    CAS  PubMed  Google Scholar 

  54. Lugert, S. et al. Homeostatic neurogenesis in the adult hippocampus does not involve amplification of Ascl1high intermediate progenitors. Nat. Commun. 3, 670 (2012).

    PubMed  Google Scholar 

  55. Chen, T. W. et al. Ultrasensitive fluorescent proteins for imaging neuronal activity. Nature 499, 295–300 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Armbruster, B. N., Li, X., Pausch, M. H., Herlitze, S. & Roth, B. L. Evolving the lock to fit the key to create a family of G protein-coupled receptors potently activated by an inert ligand. Proc. Natl Acad. Sci. USA 104, 5163–5168 (2007).

    PubMed  PubMed Central  Google Scholar 

  57. Zhang, Q. et al. Fusion-related release of glutamate from astrocytes. J. Biol. Chem. 279, 12724–12733 (2004).

    CAS  PubMed  Google Scholar 

  58. Watabe-Uchida, M., Zhu, L., Ogawa, S. K., Vamanrao, A. & Uchida, N. Whole-brain mapping of direct inputs to midbrain dopamine neurons. Neuron 74, 858–873 (2012).

    CAS  PubMed  Google Scholar 

  59. Jonsson, G., Hallman, H., Ponzio, F. & Ross, S. DSP4 (N-(2-chloroethyl)-N-ethyl-2-bromobenzylamine)—a useful denervation tool for central and peripheral noradrenaline neurons. Eur. J. Pharmacol. 72, 173–188 (1981).

    CAS  PubMed  Google Scholar 

  60. Lyons, W. E., Fritschy, J. M. & Grzanna, R. The noradrenergic neurotoxin DSP-4 eliminates the coeruleospinal projection but spares projections of the A5 and A7 groups to the ventral horn of the rat spinal cord. J. Neurosci. 9, 1481–1489 (1989).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Hylden, J. L. & Wilcox, G. L. Intrathecal morphine in mice: a new technique. Eur. J. Pharmacol. 67, 313–316 (1980).

    CAS  PubMed  Google Scholar 

  62. Yoshihara, K. et al. Astrocytic Ca2+ responses in the spinal dorsal horn by noxious stimuli to the skin. J. Pharmacol. Sci. 137, 101–104 (2018).

    CAS  PubMed  Google Scholar 

  63. Thevenaz, P., Ruttimann, U. E. & Unser, M. A pyramid approach to subpixel registration based on intensity. IEEE Trans. Image Process. 7, 27–41 (1998).

    CAS  PubMed  Google Scholar 

  64. Shiratori-Hayashi, M. et al. STAT3-dependent reactive astrogliosis in the spinal dorsal horn underlies chronic itch. Nat. Med. 21, 927–931 (2015).

    CAS  PubMed  Google Scholar 

  65. Horii, T. et al. Efficient generation of conditional knockout mice via sequential introduction of lox sites. Sci. Rep. 7, 7891 (2017).

    PubMed  PubMed Central  Google Scholar 

  66. Hayashida, K., Obata, H., Nakajima, K. & Eisenach, J. C. Gabapentin acts within the locus coeruleus to alleviate neuropathic pain. Anesthesiology 109, 1077–1084 (2008).

    CAS  PubMed  Google Scholar 

  67. Chaplan, S. R., Bach, F. W., Pogrel, J. W., Chung, J. M. & Yaksh, T. L. Quantitative assessment of tactile allodynia in the rat paw. J. Neurosci. Methods 53, 55–63 (1994).

    CAS  PubMed  Google Scholar 

  68. Masuda, T. et al. Dorsal horn neurons release extracellular ATP in a VNUT-dependent manner that underlies neuropathic pain. Nat. Commun. 7, 12529 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Tsuda, M. et al. Reduced pain behaviors and extracellular signal-related protein kinase activation in primary sensory neurons by peripheral tissue injury in mice lacking platelet-activating factor receptor. J. Neurochem. 102, 1658–1668 (2007).

    CAS  PubMed  Google Scholar 

  70. Furusho, A. et al. Development of a highly-sensitive two-dimensional HPLC system with narrowbore reversed-phase and microbore enantioselective columns and application to the chiral amino acid analysis of the mammalian brain. Chromatography 39, 83–90 (2018).

    CAS  Google Scholar 

  71. Hamase, K. et al. Determination of trace amounts of chiral amino acids in complicated biological samples using two-dimensional high-performance liquid chromatography with an innovative “Shape-Fitting” peak identification/quantification method. Chromatography 39, 147–152 (2018).

    CAS  Google Scholar 

  72. Kim, S. H. & Chung, J. M. An experimental model for peripheral neuropathy produced by segmental spinal nerve ligation in the rat. Pain 50, 355–363 (1992).

    CAS  Google Scholar 

Download references

Acknowledgements

We thank E. Sakaguchi for starting the initial experiment for rAAV, and the University of Pennsylvania vector core for providing pZac2.1, pAAV2/5, pAAV2/9 and pAd DeltaF6 plasmids. This work was supported by JSPS KAKENHI grant numbers JP19H05658, 19K22500 (to M.T.), JP16K18885, JP18K14821 (to Y.K.), JP25117013 (to K.I.), by the Core Research for Evolutional Science and Technology (CREST) program from the Japan Agency for Medical Research and Development (AMED) under grant number JP20gm0910006 (to M.T.), by the Practical Research Project for Allergic Diseases and Immunology (Research on Allergic Diseases and Immunology) from AMED under grant number JP17ek0410034 (to M.T.), the Platform Project for Supporting Drug Discovery and Life Science Research (Basis for Supporting Innovative Drug Discovery and Life Science Research (BINDS)) from AMED under grant number JP20am0101091 (to M.T.), JP20am0101120 (to R. Kobayashi, T.H. and I.H.), by Naito Foundation (to Y.K. and M.T.), and by The Nakatomi Foundation (to M.T.). The authors appreciate Shiseido Co., Ltd., for their technical support for the measurement of d-serine. T.M. and K.Y. were research fellows of the JSPS (15J03522 and 19J21063, respectively).

Author information

Authors and Affiliations

Authors

Contributions

Y.K. designed and performed most of the behavioral experiments, analyzed the data and wrote the manuscript. T.M. and K.Y. performed the in vivo imaging experiments, the Ca2+ imaging experiments using spinal cord slices, analyzed the data and wrote the manuscript. K. Kohno performed fluorescence in situ hybridization. K. Koga provided assistance for the experiments involving chemogenetic stimulation of LC neurons. R. Katsuragi and K. Momokino performed the behavioral assays and immunohistochemistry. T. Oka and S.M. performed the Ca2+ imaging in spinal cord slices. T.Y. performed immunohistochemistry of the brain. R.T. assisted with the experiments. S.O., A.F. and K. Hamase performed the d-serine measurements. T. Oti and H.S. performed the in situ experiments. K. Hayashida measured the NE levels in the SDH. R. Kobayashi, T.H. and I.H. generated Adra1aflox/flox mice. H.T.-S. assisted with the experiments and designed a custom chamber for in vivo spinal cord imaging. K. Mikoshiba contributed the role of IP3R2. V.T. provided Hes5-CreERT2 mice. K.I. supervised the project. M.T. conceived this project, designed experiments, supervised the overall project and wrote the manuscript. All the authors read and discussed the manuscript.

Corresponding author

Correspondence to Makoto Tsuda.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Neuroscience thanks Rebecca Seal 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 Distribution of Hes5+ cells in the SDH, trigeminal nucleus, dorsal root ganglia, sciatic nerve and brain.

a, Immunostaining of tdTomato+ cells by cell-type markers (NeuN for neurons, IBA1 for microglia and APC for oligodendrocytes) in L4-SDH sections of Hes5-CreERT2;Rosa-tdTomato mice. The data are representative of three independent experiments. b, Double immunostaining of SDH tdTomato+ cells by SOX9 and IB4 (a marker of non-peptidergic C fibres whose central terminals are located seletively in lamina IIi). Percentage of tdTomato+ astrocytes per total tdTomato+ cells and tdTomato+ astrocytes per total astrocytes (labeled by SOX9) in laminae I−IIi and III−V of Hes5-CreERT2;Rosa-tdTomato mice treated with tamoxifen during adulthood (right, n = 3 mice). c, tdTomato expression in the brainstem including the trigemical nucleus of Hes5-CreERT2;Rosa-tdTomato mice. d, Double immunostaining of tdTomato+ cells with SOX9 or NeuN in the spinal trigeminal nucleus caudalis (Sp5c) of Hes5-CreERT2;Rosa-tdTomato mice. e, Percentage of tdTomato+ cells with each marker in the Sp5c (n = 3 mice per group). f, g, tdTomato expression in a dorsal root ganglion (DRG, f) and sciatic nerve (g) of Hes5-CreERT2;Rosa-tdTomato mice. NeuN (f) and SOX10 (a marker for Schwann cells) (g) were used to counterstain DRG and sciatic nerve, respectively. h, i, tdTomato expression in the sagittal (h) and coronal (i) brain sections of Hes5-CreERT2;Rosa-tdTomato mice. jm, Double immunostaining of tdTomato+ cells for SOX9 (an astrocytic marker) in cortex (j), hippocampus (k), thalamus (l) and hypothalamus (m) of Hes5-CreERT2;Rosa-tdTomato mice. n, Double immunostaining of tdTomato+ cells for NeuN in cortex (left) and hippocampus (right). The data in fn were representative of three independent experiments. Data show the mean ± s.e.m. Scale bars, 20 μm (a, d), 100 μm (jn), 200 μm (b, f, g), 1 mm (c, i) and 2 mm (h).

Extended Data Fig. 2 supSDH astrocytic Ca2+ responses to several stimuli using in vivo imaging.

a, Schematic illustration of in vivo imaging of the superficial SDH in mice using two-photon microscopy. The chamber was attached at the L3−4 vertebral level. b, Immunohistochemical identification of GCaMP6m-expressing cells using cell-type markers (SOX9 and GFAP, NeuN, and IBA1). Note that GCaMP6m-expressing cells were immunolabeled by SOX9 and GFAP but not by NeuN or IBA1, indicating that GCaMP6m is expressed selectively in astrocytes. The data are representative of three independent experiments. c, Peak amplitude of astrocytic [Ca2+]i in the left SDH was high during the first 15 min (but not 55–60 min) after intraplantar injection of capsaicin (n = 59 ROIs, 3 mice). A von Frey filament (vF; 0.6 or 2.0 g) was applied to the left hindpaw before (pre) and 60 min after the injection. These light mechanical stimuli did not elicit astrocytic [Ca2+]i increases. Kruskal-Wallis test with Dunn’s multiple comparisons test. d, Representative Ca2+ traces in individual supSDH astrocytes in mice injected with capsaicin into the ipsilateral (left) and contralateral side (right). Synchronous and persistent Ca2+ responses were observed in the ipsilateral but not in the contralateral side. Data show the mean ± s.e.m. Scale bar, 20 µm. See source data for statistical parameters.

Source data

Extended Data Fig. 3 Pharmacological responsiveness of the supSDH astrocytes to 5-HT and NE, and generation of Adra1aflox/flox mice.

a, Amplitudes of Ca2+ increases evoked by bath application of 5-HT (10 or 100 μM) and NE (10 μM) to spinal cord slices (n = 86 cells, 4 slices, 4 mice). b, Effect of adrenergic agonists (n = 68 cells, 3 slices, 3 mice for phenylephrine, n = 60 cells, 3 slices, 3 mice for clonidine, n = 64 cells, 3 slices, 3 mice for isoproterenol) on astrocytic Ca2+ levels. Note that the α1-AR agonist phenylephrine produced Ca2+ increases, but the α2-AR agonist clonidine and the β-AR agonist isoproterenol did not. Data show the mean ± s.e.m. c, Schematic illustration to generate a conditional allele at the Adra1a locus. d, Sequences of ssODNs with 5′- and 3′-homology arms flanking loxP and a restriction site. Asterisks indicates a phosphorothioate bond. e, Immunofluorescence of α1A-ARs in Adra1afloxflox and Hes5-CreERT2;Adra1afloxflox mice with tamoxifen administration. The data are representative of three independent experiments. Scale bar, 20 µm.

Extended Data Fig. 4 Manipulation of descending LC-NAergic pathway and assessments for Hes5+ astrocyte-specific α1A-AR-knockdown.

a, Co-expression of tyrosine hydroxylate (TH) immunoreactivity (green) and tdTomato expression in brainstem NAergic nuclei. To determine the main region of the descending NAergic pathway projecting to the SDH in mice, we injected the retrograde virus AAV2-retro expressing Cre under the control of the enhanced synapsin promoter (AAV2-retro-ESYN-Cre) in the SDH of Rosa-tdTomato mice. Note that a number of tdTomato+ LC-NAergic neurons were observed in the ventral part of the LC. The data are representative of three independent experiments. b, NE transporter immunofluorescence in the SDH 3 days after saline or DSP-4 treatments. Fluorescence intensity ratio of NE transporter in the SDH of saline or DSP-4 treated mice (n = 3 mice per group, two-tailed unpaired t-test). c, NE content in the SDH of saline or DSP-4 treated mice (n = 5 mice for saline, n = 6 mice for DSP-4; two-tailed unpaired t-test). d, Schematic illustration of experimental approach. Microinjection of AAV2-retro-ESYN-Cre into the SDH of wild-type mice, and subsequently, AAV-hM3DqFLEX into the bilateral LC. e, Immunolabeling of HA-tag (green) in the brainstem of mice injected with AAV2-retro-ESYN-Cre in to the SDH and AAV-hM3DqFLEX into the LC. f, Immunolabeling of HA-tag (green), mCherry (magenta) and TH (white) in the LC of mice transduced with hM3Dq into LC-NAergic neurons. In e and f, the images were representative of three independent experiments. g, Schematic timeline for tamoxifen treatments and behavioral experiments. Mice were injected with tamoxifen once a day for 10 days (days 12 to 21 after rAAV injection) and behavioral experiments (assessment of mechanical hypersensitivity after intrathecal injection of phenylephrine or intraplantar injection of capsaicin) were conducted 7 days after the last tamoxifen treatment. h, Amplitude of Ca2+ increases evoked by phenylephrine (0.3 μM) and ionomycin (1 μM) in primary cultured astrocytes (n = 23 cells for shScramble, n = 17 cells for shAdra1a; two-tailed Mann-Whitney test and two-tailed unpaired t-test). The shAdra1a suppressed the α1A-AR-mediated Ca2+ responses without affecting ionomycin-induced Ca2+ increases. i, Amplitude of SDH astrocytic Ca2+ increases evoked by phenylephrine (30 µM; n = 43 cells, 5 slices, 4 mice for shScramble, n = 47 cells, 5 slices, 4 mice for shAdra1a; two-tailed Mann-Whitney test). j, Immunofluorescence (left) and fluorescence intensity (right) of α1A-AR in Hes5-CreERT2;AAV-shScrambleFLEX (n = 292 cells, 9 slices, 3 mice) and shAdra1aFLEX mice (n = 270 cells, 9 slices, 3 mice). ROIs were manually selected based on DAPI (blue) /mCherry (magenta) signals (middle). Two-tailed Mann-Whitney test. Scale bars, 20 μm (j middle), 100 μm (j left), 200 μm (a, b, f) and 1 mm (e). Data show the mean ± s.e.m. See source data for statistical parameters.

Source data

Extended Data Fig. 5 Functional expression of hM3Dq in Hes5+ SDH astrocytes.

a, Schematic illustration of the experimental protocol and timeline in Hes5-CreERT2 mice using the Cre-On system. AAV-hM3DqFLEX was microinjected unilaterally into the SDH of Hes5-CreERT2 mice. Tamoxifen treatment induced the expression of hM3Dq in Cre-positive (Hes5+) cells. b, Double immunostaining of hM3Dq (HA-tag) and cell-type markers (NeuN, IBA1 and APC) in the SDH of Hes5-CreERT2;AAV-hM3DqFLEX mice. The data are representative of three independent experiments. c, Representative images of the Ca2+ response (left) and an example of traces and average data (right) for the effect of CNO treatment on the fluorescence intensity of ROIs in wild-type and Hes5-CreERT2;AAV-hM3DqFLEX mice, both of which expressed GCaMP6m in SDH astrocytes. CNO induced astrocytic Ca2+ increases in spinal cord slices taken from Hes5-CreERT2;AAV-hM3DqFLEX mice but not from wild-type;AAV-hM3DqFLEX mice. The data are representative of three independent experiments. Scale bars, 20 μm (b) and 100 μm (c).

Extended Data Fig. 6 Chemogenetic stimulation of SDH astrocytes and Hes5-negative astrocytes in deeper laminae of the SDH.

a, Schematic illustration of the experimental protocol. AAV-hM3DqFLEX was microinjected unilaterally into the SDH of Gfap-Cre mice. b, Immunofluorescence of hM3Dq (HA-tag) in the SDH of Gfap-Cre and wild-type mice injected with AAV-hM3DqFLEX. Note that hM3Dq expression was induced only in Gfap-Cre;AAV-hM3DqFLEX mice and that its localization was observed in the SDH, including superficial laminae. c, Immunohistochemical identification of hM3Dq (HA-tag)-expressing cells using cell-type markers (SOX9, GFAP, NeuN, IBA1 and APC) in the SDH of Gfap-Cre;AAV-hM3DqFLEX mice. Arrows indicate hM3Dq-positive cells. hM3Dq was selectively expressed in astrocytes. The data in b and c were representative of three independent experiments. d, hM3Dq-induced mechanical hypersensitivity in Gfap-Cre;AAV-hM3DqFLEX mice (saline, n = 4 mice; CNO, n = 5 mice); repeated measures two-way ANOVA with Bonferroni’s multiple comparison test and Kruskal-Wallis test with Dunn’s multiple comparisons test. The letter indicates the following P value versus the ipsilateral side of the saline group (a, P < 0.0001). e, Schematic illustration of the experimental protocol and timeline in Hes5-CreERT2 mice using the Cre-Off system. AAV-gfaABC1D-hM3DqFLEX was microinjected unilaterally into the SDH of Hes5-CreERT2 mice, in which the hM3Dq sequence can be inverted by tamoxifen to stop hM3Dq expression in Cre-positive (Hes5+) cells. By contrast, Cre-negative (Hes5-negative) astrocytes (located in the deeper laminae) express hM3Dq. f, c-Fos+ neurons in laminae I–IIi and III–V of vehicle or CNO-treated Gfap-Cre;AAV-hM3DqFLEX mice with or without Aβ fiber stimulation (vehicle without stimuli, n = 6 mice; CNO without stimuli, n = 5 mice; vehicle with stimuli, n = 6 mice; CNO with stimuli, n = 6 mice; one-way ANOVA with Tukey’s multiple comparisons test). Insets: c-Fos immunofluorescence in laminae I–IIi. Scale bars, 50 μm (c), 100 μm (f insets) and 200 μm (b, f). Dashed lines (b, f) indicate the boundary between the gray and white matters. Data show the mean ± s.e.m. See source data for statistical parameters.

Source data

Extended Data Fig. 7 Involvement of IP3R2 and D-serine signaling on NMDA receptors for astrogliogenic hypersensitivity.

a, Immunolabeling of IP3R2 in the SDH of wild-type and Ip3r2–/– mice. b, Immunohistochemical identification of IP3R2-expressing cells using cell-type markers (SOX9, GFAP, NeuN and APC) in the SDH of wild-type mice. Note that almost all IP3R2-positive cells in the SDH were astrocytes. The data in a and b were representative of three independent experiments. c, Representative images of Ca2+ response (left) and an example of traces and average data (right) for the effect of CNO (100 μM) treatment on the fluorescence intensity of ROIs in Gfap-Cre;Ip3r2–/–;AAV-hM3DqFLEX mice in which SDH astrocytes expressed GCaMP6m. CNO did not induce Ca2+ increases in Gfap-Cre;Ip3r2–/– mice. The data are representative of three independent experiments. d, PWT of Gfap-Cre;Ip3r2+/+ and Gfap-Cre;Ip3r2–/– mice injected with AAV-hM3DqFLEX before and after CNO administration (n = 4 mice per group; repeated measures two-way ANOVA with Bonferroni’s multiple comparisons test). The letters indicate the following P values versus the ipsilateral side of Gfap-Cre;Ip3r2+/+ mice (a, P = 0.0005; b, P < 0.0001; c, P = 0.0478). e, c-Fos+ neurons in laminae I–IIi and III–V of CNO-treated Gfap-Cre;Ip3r2–/–;AAV-hM3DqFLEX mice with (n = 6 mice) or without (n = 5 mice) Aβ fiber stimulation (two-tailed unpaired t-test). CNO did not increase the number of SDH neurons positive for c-Fos in laminae I–IIi and III-V. fj, Effect of forced expression of dnSNARE (f) and pharmacological blockade of receptors for NMDA (g), AMPA (h), ATP (i) and of a glycine site on NMDA receptors (j) on the CNO-induced mechanical hypersensitivity in Gfap-Cre;AAV-hM3DqFLEX mice (dnSNARE, MK-801, CNQX and DCK, n = 5 mice per group; PPADS, n = 4 mice per group); repeated measures two-way ANOVA with Bonferroni’s multiple comparisons test. The letters indicate the following P values versus the ipsilateral side of hM3Dq group (f: a, P = 0.0002), versus the ipsilateral side of CNO + Saline treated group (g: a, P = 0.0011), and versus the ipsilateral side of CNO + Vehicle treated group (j: a, P = 0.0002). k, PWL in wild-type mice 30 min after intrathecal injection of saline or D-serine (100 nmol) (n = 6 mice per group, two-tailed unpaired t-test). l, c-Fos+ neurons in laminae I–IIi and III–V of wild-type mice 30 min after intrathecal injection of D-serine (100 nmol) with (n = 5 mice) or without (n = 4 mice) Aβ fiber stimulation; two-tailed unpaired t-test. m, Schematic illustration of experimental protocol (left). Extracellular D-serine levels (% D-serine) in supernatant of spinal cord slices (taken from Hes5-CreERT2;AAV-hM3DqFLEX mice) 2 min after vehicle or CNO (100 μM) application (right, n = 11 mice per group, two-tailed unpaired t-test). Scale bar, 20 μm (b) and 100 μm (a, c, l). Dashed lines (a and l) indicate the boundary between the gray and white matters. Data show the mean ± s.e.m. See source data for statistical parameters.

Source data

Extended Data Fig. 8 α1A-AR-knockdown of Hes5-negative astrocytes in deeper laminae of the SDH.

Schematic illustration of the Cre-OFF system to knockdown α1A-ARs in Cre-negative (Hes5-negative) astrocytes. AAV-gfaABC1D-shRNA (shScramble or shAdra1a)FLEX was microinjected unilaterally into the SDH of Hes5-CreERT2 mice. Assessment of mechanical hypersensitivity after intraplantar injection of capsaicin was conducted 7 days after the last tamoxifen treatment.

Extended Data Fig. 9 Distribution of Hes5+ SDH astrocytes after peripheral nerve injury.

a, Schematic timeline for tamoxifen treatments and fixation. Mice were injected with tamoxifen once a day for 5 days [days 5 to 9 or days 12 to 16 after peripheral nerve injury (PNI)] and were fixed 7 days after the last tamoxifen treatment. b, tdTomato expression in the SDH of Hes5-CreERT2;Rosa-tdTomato mice after PNI. c, Immunostaining of SOX9 in the ipsilateral SDH of Hes5-CreERT2;Rosa-tdTomato mice after PNI. Note that cell-type specificity and the expression pattern of tdTomato+ cells were not changed after PNI. In b and c, the data are representative of three independent experiments. Scale bars, 200 μm (c) and 500 μm (b).

Extended Data Fig. 10 Effect of intrathecal injection of NE and phenylephrine at higher doses on mechano- and thermosensory behaviors.

a, b, PWT (a) or PWL (b) in wild-type mice 30 min after intrathecal injection of noradrenaline (n = 6 mice per group; two-tailed paired t test and two-tailed Wilcoxon matched-pairs signed rank test). c, d, PWT (c) or PWL (d) in wild-type mice 30 min after injection of phenylephrine (n = 6 mice per group; two-tailed paired t test). Data show the mean ± s.e.m. See source data for statistical parameters.

Source data

Supplementary information

Supplementary Information

Supplementary Tables 1 and 2.

Reporting Summary

Source data

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 7

Statistical source data.

Source Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kohro, Y., Matsuda, T., Yoshihara, K. et al. Spinal astrocytes in superficial laminae gate brainstem descending control of mechanosensory hypersensitivity. Nat Neurosci 23, 1376–1387 (2020). https://doi.org/10.1038/s41593-020-00713-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-020-00713-4

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