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:

Connexin 30 sets synaptic strength by controlling astroglial synapse invasion

Abstract

Astrocytes play active roles in brain physiology by dynamic interactions with neurons. Connexin 30, one of the two main astroglial gap-junction subunits, is thought to be involved in behavioral and basic cognitive processes. However, the underlying cellular and molecular mechanisms are unknown. We show here in mice that connexin 30 controls hippocampal excitatory synaptic transmission through modulation of astroglial glutamate transport, which directly alters synaptic glutamate levels. Unexpectedly, we found that connexin 30 regulated cell adhesion and migration and that connexin 30 modulation of glutamate transport, occurring independently of its channel function, was mediated by morphological changes controlling insertion of astroglial processes into synaptic clefts. By setting excitatory synaptic strength, connexin 30 plays an important role in long-term synaptic plasticity and in hippocampus-based contextual memory. Taken together, these results establish connexin 30 as a critical regulator of synaptic strength by controlling the synaptic location of astroglial processes.

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

Figure 1: Cx30 controls synaptic strength by regulating synaptic glutamate levels.
Figure 2: Cx30 control of synaptic glutamate levels is caused by altered astroglial glutamate uptake.
Figure 3: Cx30-mediated effects are channel independent.
Figure 4: Cx30 regulates cell morphology, adhesion and migration.
Figure 5: Astroglial process extension toward synaptic clefts enhances glutamate clearance and decreases AMPAR-mediated synaptic transmission.
Figure 6: Astrocytes deficient in Cx30 invade synaptic clefts.

Similar content being viewed by others

References

  1. Perea, G., Navarrete, M. & Araque, A. Tripartite synapses: astrocytes process and control synaptic information. Trends Neurosci. 32, 421–431 (2009).

    Article  CAS  Google Scholar 

  2. Genoud, C. et al. Plasticity of astrocytic coverage and glutamate transporter expression in adult mouse cortex. PLoS Biol. 4, e343 (2006).

    Article  Google Scholar 

  3. Iino, M. et al. Glia-synapse interaction through Ca2+-permeable AMPA receptors in Bergmann glia. Science 292, 926–929 (2001).

    Article  CAS  Google Scholar 

  4. Oliet, S.H., Piet, R. & Poulain, D.A. Control of glutamate clearance and synaptic efficacy by glial coverage of neurons. Science 292, 923–926 (2001).

    Article  CAS  Google Scholar 

  5. Pannasch, U. & Rouach, N. Emerging role for astroglial networks in information processing: from synapse to behavior. Trends Neurosci. 36, 405–417 (2013).

    Article  CAS  Google Scholar 

  6. Elias, L.A., Wang, D.D. & Kriegstein, A.R. Gap junction adhesion is necessary for radial migration in the neocortex. Nature 448, 901–907 (2007).

    Article  CAS  Google Scholar 

  7. Theis, M., Sohl, G., Eiberger, J. & Willecke, K. Emerging complexities in identity and function of glial connexins. Trends Neurosci. 28, 188–195 (2005).

    Article  CAS  Google Scholar 

  8. Nagy, J.I., Patel, D., Ochalski, P.A. & Stelmack, G.L. Connexin30 in rodent, cat and human brain: selective expression in gray matter astrocytes, co-localization with connexin43 at gap junctions and late developmental appearance. Neuroscience 88, 447–468 (1999).

    Article  CAS  Google Scholar 

  9. Harris, A.L. Emerging issues of connexin channels: biophysics fills the gap. Q. Rev. Biophys. 34, 325–472 (2001).

    Article  CAS  Google Scholar 

  10. Roux, L., Benchenane, K., Rothstein, J.D., Bonvento, G. & Giaume, C. Plasticity of astroglial networks in olfactory glomeruli. Proc. Natl. Acad. Sci. USA 108, 18442–18446 (2011).

    Article  CAS  Google Scholar 

  11. Dere, E. et al. Connexin30-deficient mice show increased emotionality and decreased rearing activity in the open-field along with neurochemical changes. Eur. J. Neurosci. 18, 629–638 (2003).

    Article  CAS  Google Scholar 

  12. Frisch, C. et al. Mice with astrocyte-directed inactivation of connexin43 exhibit increased exploratory behaviour, impaired motor capacities, and changes in brain acetylcholine levels. Eur. J. Neurosci. 18, 2313–2318 (2003).

    Article  Google Scholar 

  13. 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  Google Scholar 

  14. Rampon, C. et al. Effects of environmental enrichment on gene expression in the brain. Proc. Natl. Acad. Sci. USA 97, 12880–12884 (2000).

    Article  CAS  Google Scholar 

  15. Ji, J. & Maren, S. Hippocampal involvement in contextual modulation of fear extinction. Hippocampus 17, 749–758 (2007).

    Article  Google Scholar 

  16. Tzingounis, A.V. & Wadiche, J.I. Glutamate transporters: confining runaway excitation by shaping synaptic transmission. Nat. Rev. Neurosci. 8, 935–947 (2007).

    Article  CAS  Google Scholar 

  17. Butchbach, M.E., Tian, G., Guo, H. & Lin, C.L. Association of excitatory amino acid transporters, especially EAAT2, with cholesterol-rich lipid raft microdomains: importance for excitatory amino acid transporter localization and function. J. Biol. Chem. 279, 34388–34396 (2004).

    Article  CAS  Google Scholar 

  18. Bellesi, M., Melone, M., Gubbini, A., Battistacci, S. & Conti, F. GLT-1 upregulation impairs prepulse inhibition of the startle reflex in adult rats. Glia 57, 703–713 (2009).

    Article  Google Scholar 

  19. Omrani, A. et al. Up-regulation of GLT-1 severely impairs LTD at mossy fibre–CA3 synapses. J. Physiol. (Lond.) 587, 4575–4588 (2009).

    Article  CAS  Google Scholar 

  20. Rothstein, J.D. et al. Beta-lactam antibiotics offer neuroprotection by increasing glutamate transporter expression. Nature 433, 73–77 (2005).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  22. Grifa, A. et al. Mutations in GJB6 cause nonsyndromic autosomal dominant deafness at DFNA3 locus. Nat. Genet. 23, 16–18 (1999).

    Article  CAS  Google Scholar 

  23. Schütz, M. et al. The human deafness-associated connexin 30 T5M mutation causes mild hearing loss and reduces biochemical coupling among cochlear non-sensory cells in knock-in mice. Hum. Mol. Genet. 19, 4759–4773 (2010).

    Article  Google Scholar 

  24. Colin, A. et al. Engineered lentiviral vector targeting astrocytes in vivo. Glia 57, 667–679 (2009).

    Article  Google Scholar 

  25. Qu, C., Gardner, P. & Schrijver, I. The role of the cytoskeleton in the formation of gap junctions by Connexin 30. Exp. Cell Res. 315, 1683–1692 (2009).

    Article  CAS  Google Scholar 

  26. Saab, A.S. et al. Bergmann glial AMPA receptors are required for fine motor coordination. Science 337, 749–753 (2012).

    Article  CAS  Google Scholar 

  27. Parsons, J.T., Horwitz, A.R. & Schwartz, M.A. Cell adhesion: integrating cytoskeletal dynamics and cellular tension. Nat. Rev. Mol. Cell Biol. 11, 633–643 (2010).

    Article  CAS  Google Scholar 

  28. Cornell-Bell, A.H., Thomas, P.G. & Smith, S.J. The excitatory neurotransmitter glutamate causes filopodia formation in cultured hippocampal astrocytes. Glia 3, 322–334 (1990).

    Article  CAS  Google Scholar 

  29. Ozog, M.A., Siushansian, R. & Naus, C.C. Blocked gap junctional coupling increases glutamate-induced neurotoxicity in neuron-astrocyte co-cultures. J. Neuropathol. Exp. Neurol. 61, 132–141 (2002).

    Article  CAS  Google Scholar 

  30. Princen, F. et al. Rat gap junction connexin-30 inhibits proliferation of glioma cell lines. Carcinogenesis 22, 507–513 (2001).

    Article  CAS  Google Scholar 

  31. Carmona, M.A., Murai, K.K., Wang, L., Roberts, A.J. & Pasquale, E.B. Glial ephrin-A3 regulates hippocampal dendritic spine morphology and glutamate transport. Proc. Natl. Acad. Sci. USA 106, 12524–12529 (2009).

    Article  CAS  Google Scholar 

  32. Filosa, A. et al. Neuron-glia communication via EphA4/ephrin-A3 modulates LTP through glial glutamate transport. Nat. Neurosci. 12, 1285–1292 (2009).

    Article  CAS  Google Scholar 

  33. Murai, K.K., Nguyen, L.N., Irie, F., Yamaguchi, Y. & Pasquale, E.B. Control of hippocampal dendritic spine morphology through ephrin-A3/EphA4 signaling. Nat. Neurosci. 6, 153–160 (2003).

    Article  CAS  Google Scholar 

  34. Witcher, M.R., Kirov, S.A. & Harris, K.M. Plasticity of perisynaptic astroglia during synaptogenesis in the mature rat hippocampus. Glia 55, 13–23 (2007).

    Article  Google Scholar 

  35. Gros, D. et al. Connexin 30 is expressed in the mouse sino-atrial node and modulates heart rate. Cardiovasc. Res. 85, 45–55 (2010).

    Article  CAS  Google Scholar 

  36. Cohen-Salmon, M. et al. Connexin30 deficiency causes instrastrial fluid-blood barrier disruption within the cochlear stria vascularis. Proc. Natl. Acad. Sci. USA 104, 6229–6234 (2007).

    Article  CAS  Google Scholar 

  37. 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  Google Scholar 

  38. Scimemi, A., Tian, H. & Diamond, J.S. Neuronal transporters regulate glutamate clearance, NMDA receptor activation, and synaptic plasticity in the hippocampus. J. Neurosci. 29, 14581–14595 (2009).

    Article  CAS  Google Scholar 

  39. Melone, M., Bellesi, M. & Conti, F. Synaptic localization of GLT-1a in the rat somatic sensory cortex. Glia 57, 108–117 (2009).

    Article  Google Scholar 

  40. Barbour, B. An evaluation of synapse independence. J. Neurosci. 21, 7969–7984 (2001).

    Article  CAS  Google Scholar 

  41. Zheng, K., Scimemi, A. & Rusakov, D.A. Receptor actions of synaptically released glutamate: the role of transporters on the scale from nanometers to microns. Biophys. J. 95, 4584–4596 (2008).

    Article  CAS  Google Scholar 

  42. Haber, M., Zhou, L. & Murai, K.K. Cooperative astrocyte and dendritic spine dynamics at hippocampal excitatory synapses. J. Neurosci. 26, 8881–8891 (2006).

    Article  CAS  Google Scholar 

  43. Condorelli, D.F. et al. Connexin-30 mRNA is up-regulated in astrocytes and expressed in apoptotic neuronal cells of rat brain following kainate-induced seizures. Mol. Cell. Neurosci. 21, 94–113 (2002).

    Article  CAS  Google Scholar 

  44. Koulakoff, A., Ezan, P. & Giaume, C. Neurons control the expression of connexin 30 and connexin 43 in mouse cortical astrocytes. Glia 56, 1299–1311 (2008).

    Article  Google Scholar 

  45. Bernard, R. et al. Altered expression of glutamate signaling, growth factor, and glia genes in the locus coeruleus of patients with major depression. Mol. Psychiatry 16, 634–646 (2011).

    Article  CAS  Google Scholar 

  46. Ernst, C. et al. Dysfunction of astrocyte connexins 30 and 43 in dorsal lateral prefrontal cortex of suicide completers. Biol. Psychiatry 70, 312–319 (2011).

    Article  CAS  Google Scholar 

  47. Escartin, C. et al. Ciliary neurotrophic factor activates astrocytes, redistributes their glutamate transporters GLAST and GLT-1 to raft microdomains, and improves glutamate handling in vivo. J. Neurosci. 26, 5978–5989 (2006).

    Article  CAS  Google Scholar 

  48. Jonas, P., Major, G. & Sakmann, B. Quantal components of unitary EPSCs at the mossy fibre synapse on CA3 pyramidal cells of rat hippocampus. J. Physiol. (Lond.) 472, 615–663 (1993).

    Article  CAS  Google Scholar 

  49. Geiger, J.R.P., Roth, A., Taskin, B. & Jonas, P. Glutamate-mediated synaptic excitation of cortical interneurons. in Ionotropic Glutamate Receptors in the CNS: Handbook of Experimental Pharmacology (eds. Jonas P. & Monyer H.) 363–398 (Springer, Berlin, 1999).

  50. Singer, A., Schuss, Z. & Holcman, D. Narrow escape and leakage of Brownian particles. Phys. Rev. E 78, 051111 (2008).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank R. Nicoll, A. Koulakoff, E. Brouillet, O. Chever, F. Mammano and D. Schmitz for discussions and for technical assistance. This work was supported by grants from the Human Frontier Science Program Organization (Career Development Award), French Research Agency (Programme Jeunes chercheurs and Programme Blanc), City of Paris (Programme Emergence) and INSERM to N.R., from CEA and CNRS to C.E. and N.D., from International Brain Research Organization to V.A., from the French Research Ministry and Deutsche Forschungsgemeinschaft to U.P., from Labex Memolife to G.D., from the doctoral school ED3C, Paris 6 University to G.G. and from the Max-Planck Society to D.F.

Author information

Authors and Affiliations

Authors

Contributions

G.G., C.E., P.E., M.C.-S. and K.B. contributed equally to this work. Conception and experimental design: N.R., U.P., D.H., G.D., D.F., G.K., C.E., K.B., M.C.-S.; Methodology and data acquisition: U.P., N.R., D.F., G.D., D.H., G.K., C.E., N.D., K.B., M.C.-S., G.G., V.A., P.E.; Analysis and interpretation of data: U.P., N.R., D.F., G.D., D.H., G.K., C.E., K.B., M.C.-S., G.G., V.A., P.E., A.D., J.H.R.L.; Manuscript writing and revision: N.R., U.P., G.D., D.H., G.K., C.E., K.B., J.H.R.L.

Corresponding author

Correspondence to Nathalie Rouach.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Cx30 controls glutamatergic transmission of CA1 pyramidal cells.

(a) Application of CNQX (100 nM) for 5 min in hippocampal slices from Cx30+/+ mice (n = 5 cells) mimics the decrease in mEPSC frequency and amplitude from pyramidal cells of Cx30−/− mice (Frequency, P = 0.0020, t(4) = 7.162; amplitude P= 0.0065, t(4) = 5.194, paired t test). (b,c) Frequency and amplitude cumulative probability of pyramidal cell mIPSCs (n = 11 cells for Cx30+/+ and Cx30−/− mice, b) and pyramidal cell mEPSCs before onset of Cx30 expression (P8-10 mice, Cx30+/+ n = 11 cells; Cx30−/− n = 13 cells, c) are not changed (mIPSC frequency: P = 0.8186, KS = 2; mIPSC amplitude: P = 0.5596, KS = 0.25; mEPSC frequency: P> 0.9999, KS = 0.1000; mEPSC amplitude: P = 0.3291, KS = 0.3, Kolmogorov-Smirnov). Data are expressed as mean ±s.e.m. ** P < 0.01.

Supplementary Figure 2 Normal hippocampal structure in Cx30−/− mice.

(a,b) Immunostaining of hippocampal slices for neurons with NeuN and astrocytes with S100 reveals no alteration in both cell types numbers in Cx30−/− mice (n = 10 and 8 fields, respectively) compared to Cx30+/+ mice (n = 10 and 8 fields, respectively; NeuN: P = 0.2485, t(18) = 1.193; S100: P = 0.9098, t(14) = 0.1154, unpaired t test). Scale bar, 400 μm for NeuN, 100 μm for S100. (c,d) Synapse density in defined volume fractions and PSD area, analyzed by EM, are comparable in Cx30−/− and Cx30+/+ mice (Synapse density: Cx30+/+ n = 6 volume fractions, Cx30−/− n = 9 volume fractions, P = 0.8251, t(13) = 0.2254, unpaired t test; PSD area: Cx30+/+ n = 167 spines; Cx30−/− n = 256 spines, P = 0.9298, U = 21607, Mann-Whitney). (e) Immunoblot analysis of hippocampal extracts from Cx30+/+ (n = 3 mice) and Cx30−/− mice (n = 3 mice) shows no difference in protein expression for synaptophysin (Syn) and PSD-95. Tubulin (Tub). Full-length blots are presented in Supplementary Figure S8d. Data are expressed as mean ±s.e.m.

Supplementary Figure 3 Cx30 regulates LTP induction, but does not affect Schaffer collateral afferent responses, intrinsic membrane properties and excitability of CA1 pyramidal cells and gliotransmitter release.

(a) LTP, induced by a pairing protocol (arrow, 2 Hz stimulation and depolarization of the postsynaptic cell at 0 mV for 1 min) bypassing insufficient postsynaptic activation is normal in CA1 pyramidal cells from Cx30−/− mice (n = 4 cells, Cx30+/+ n = 4 cells, P = 0.3687, t(8) = 0.9525, comparison 15-25 min after the tetanus, unpaired t test). Sample traces represent averaged EPSC responses before and 15-25 min after the stimulation. Scale bar 10 pA, 20 ms. (b) Afferent responses of Schaffer collaterals are unchanged in Cx30−/− mice, as assessed by input-output curves showing presynaptic fiber volley amplitude as a function of stimulation intensity in CA1 stratum radiatum of Cx30−/− mice (n = 10 slices) and Cx30+/+ mice (n = 9 slices, genotype: P = 0.9236, F(1,102) = 0.009254; stimulation intensity: P < 0.0001, F(5,102) = 40.88, two-way ANOVA). (c) Representative sample traces of action potentials elicited by current injections (100 pA, 500 ms) in CA1 pyramidal neurons from Cx30+/+ and Cx30−/− slices. Scale bar, 20 mV, 50 ms. Higher magnification of a representative action potential. Scale bar, 10 mV, 5 ms. Measured membrane properties are indicated. (d) Intrinsic membrane properties of CA1 pyramidal neurons from Cx30−/− mice (n = 18 cells) are similar to the ones of Cx30+/+ mice (n = 12 cells, membrane potential: P = 0.4072, U = 88, Mann-Whitney; spike amplitude: P = 0.4431, t(28) = 0.7780, unpaired t test; spike threshold: P = 0.1245, t(28) = 1.583, unpaired t test; AHP: P = 0.5660, U = 94, Mann-Whitney; number of spikes: P = 0.3511, t(28) = 0.9482, unpaired t test). Afterhyperpolarization (AHP). (e) Endogenous activation of adenosine-1 receptors (inhibited by DPCPX), NMDA receptors (inhibited by CPP) or metabotropic glutamate receptors (inhibited by LY341495) is not changed in CA1 pyramidal neurons from Cx30−/− mice (DPCPX, n = 6 cells; CPP, n = 4 cells and LY341495, n = 4 cells; Cx30+/+ mice: DPCPX, n = 9 cells, CPP, n = 5 cells, LY341495, n = 4 cells; DPCPX: P = 0.6809, U = 23; CPP: P = 0.2857, U = 5; LY341495: P = 0.6571, U = 6, Mann-Whitney). Data are expressed as mean ±s.e.m.

Supplementary Figure 4 Cx30 alters synaptically evoked glutamate transporter currents in astrocytes without modifying their intrinsic membrane properties.

(a) Simultaneous recordings of synaptically evoked fEPSPs (smaller inset) and GLT currents in astrocytes (1). Kynurenic acid (Kyn, 5 mM) inhibits fEPSPs (inset) and reduces the glial potassium current. The remaining glial current (2) consists of a fast inward current and a slow component of smaller amplitude, which can be isolated by application of TBOA (200 μM) (3). Subtraction of the slow component isolates the GLT current (2-3). Scale bars, 2.5 pA, 25 ms for GLT current; 0.2 mV, 2 ms for fEPSP. In hippocampal slices from Cx30+/+ (n = 10 slices) and Cx30−/− mice (n = 10 slices), amplitudes of evoked fiber volleys (b) and pharmacologically isolated potassium currents (c) were similar (b: P = 0.7517, t(18) = 0.3212; c: P = 0.1091, t(11.6) = 1.735, unpaired t test with Welch's correction), whereas GLT currents (d) and fEPSP slope/fiber volley ratios (e) were altered (d: P = 0.0438, t(18) = 2.168; e: P = 0.0372, t(18) = 2.25, unpaired t test). (f) Resting membrane potential and membrane resistance in astrocytes from Cx30+/+ (n = 20 cells) and Cx30−/− mice (n = 16 cells) were similar (Membrane potential: P = 0.8917, t(34) = 0.1372, unpaired t test; membrane resistance: P = 0.3894, U = 132.5, Mann-Whitney). (g) Current-voltage (I/V) plots of CA1 astrocytic currents evoked by 150 ms voltage steps (+180 – +40 mV) (Cx30+/+ n = 28 cells, Cx30−/− n = 25 cells, genotype: P = 0.6685, F(1,1173) = 0.1835, voltage step: P < 0.0001, F(22,1173) = 415.9, two-way ANOVA). Scale bar, 1 nA, 25 ms. Data are expressed as mean ± s.e.m. * P < 0.05.

Supplementary Figure 5 Cx30 does not alter the distribution of GLT1 in raft domains, and GLT1 upregulation has no effect on synaptic transmission in wild-type mice.

(a,b) Representative sucrose gradient fractionation profile for GLT1 and flotilin-1 (Flot-1) in Cx30+/+ (a) and Cx30−/− hippocampal extracts (b). Full-length blots are presented in Supplementary Figure S8e. Raft domains are recovered in low-density fractions 3-5, which are enriched in flotilin-1, a component of lipid rafts. Fractions 8, 9, and 10 correspond to detergent-soluble material, and fraction 11 is made of detergent insoluble, non-raft, pellet material. An aliquot of total homogenate (T) used to load the discontinuous gradient is also shown on the same gel. Protein levels were measured in each fraction (right). (c) Increasing GLT1 density in astrocytes by ceftriaxone treatment does not alter excitatory synaptic transmission in wild-type mice (n = 12 slices) compared to saline treated controls (n = 11 slices, treatment: P = 0.4579, F(1,126) = 0.5545, fiber volley: P < 0.0001 F(5,126) = 56.80, two-way ANOVA), as investigated by input-output curves. Mice received daily 200 mg/kg ceftriaxone or saline for 7 days, as previously described18-20. Experiments were performed on day 7. Data are expressed as mean ± s.e.m.

Supplementary Figure 6 Cx30−/− astrocytes are not hypertrophic or reactive.

Hippocampal astrocytes of Cx30−/− mice exhibit no difference in cell soma size and overall volume, as shown by the cytoplasmic marker S100. Scale bars 10 μm, 50 μm. (b) Sulforhodamine 101 (SR101) staining confirmed the S100 staining results. Scale bars 10 μm, 30 μm. Soma size quantification for both markers is shown in (c) (S100: Cx30−/− n = 80 cells, Cx30+/+ n = 80 cells, P = 0.8681, U = 3151, Mann-Whitney; SR101: Cx30−/− n = 32 cells, Cx30+/+ n = 20 cells, P = 0.2110, t(28.34) = 1.280, unpaired t test). Immunostaining for glutamine synthetase (d), vimentin (e) or signal transducer and activator of transcription 3 (Stat3) (f) is not different between Cx30+/+ (n = 3 mice) and Cx30−/− mice (n = 3 mice). As a positive control for vimentin expression, hippocampal sections from Cx30−/− mice sacrificed 7 days after intraperitoneal kainate injection (30 mg/kg) were used. Scale bar, 10 μm. Positive control for Stat3 labeling consists in mice injected in the striatum with a lentiviral vector encoding for the cytokine ciliary neurotrophic factor (CNTF) and, 4 weeks later with quinolinate (Quin). Mice were sacrificed after 2 weeks and display robust astrocyte reactivity in the striatum. Scale bar, 30 μm. (g,h) Electron micrographs illustrating representative astrocytic elements in Cx30−/− mice (n = 3 mice) (g) and in wild-type mice with CNTF-activated astrocytes (n = 3 mice) (h). Note the enhanced filament density in the astrocyte (a) in h. Scale bar 0.2 μm. Data are expressed as mean ±s.e.m.

Supplementary Figure 7 Serial electron microscopy images and three-dimensional reconstruction of a dendritic spine and surrounding astroglial processes in Cx30+/+ and Cx30−/− mice.

Serial electron microscopy images from a series illustrating an asymmetric synapse (axonal bouton (b), dendritic spine (s)) contacted by astroglial processes (green) in Cx30+/+ (a) and Cx30−/− mice (c), respectively. Scale bar, 0.2 μm. The corresponding reconstructions of the entire dendritic spine (grey) with PSD (red) and the surrounding astrocytic elements (green) are illustrated for Cx30+/+ (b) and Cx30−/− mice (d), showing a closer association of Cx30−/− astrocytic processes to dendritic spine and a shorter distance to PSD.

Supplementary Figure 8 Full-length pictures of the blots presented in the main and supplementary figures.

(a) Figure 2f. (b) Figure 2g. (c) Figure 3c. (d) Supplementary Figure 2e. (e) Supplementary Figure 5a,b. Rectangles over the full length gels indicate where the images for the main and supplementary figures were cropped.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–8 (PDF 2976 kb)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Pannasch, U., Freche, D., Dallérac, G. et al. Connexin 30 sets synaptic strength by controlling astroglial synapse invasion. Nat Neurosci 17, 549–558 (2014). https://doi.org/10.1038/nn.3662

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.3662

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