Skip to content
Licensed Unlicensed Requires Authentication Published by De Gruyter January 16, 2014

Establishment of neurogenic microenvironment in the neurovascular unit: the connexin 43 story

  • Alla B. Salmina EMAIL logo , Andrey V. Morgun , Nataliya V. Kuvacheva , Olga L. Lopatina , Yulia K. Komleva , Nataliya A. Malinovskaya and Elena A. Pozhilenkova

Abstract

Connexins (Cx) play an important role in the coordination of intercellular communication, and autocrine and paracrine regulation of cells within the neurovascular unit (NVU). Gap junctional mechanisms control proliferation and differentiation processes underlying neurogenesis and angiogenesis in the brain. Cx43 possesses some unique properties [the ability to form either intercellular channels permeable for regulatory molecules and ions or hemichannels open to the extracellular space to provide release of cell metabolites; functional coupling with nicotinamide adenine dinucleotide (NAD+)-consuming and NAD+-dependent enzymatic processes] which may be of great importance for the fate of the stem cells. Dynamic changes in Cx43 expression are associated with different stages of brain cells development either at embryonic or adult periods of ontogenesis. This review summarizes recent data on Cx43-controlled neurogenesis in the context of NVU development and functioning. Understanding the molecular mechanisms of gap junctional intercellular communication will support translational studies focused on the development of regeneration-based approaches for the therapy of central nervous system pathology.


Corresponding author: Alla B. Salmina, Department of Biochemistry, Medical, Pharmaceutical and Toxicological Chemistry, Krasnoyarsk State Medical University named after Prof. V.F. Voino-Yasenetsky, Krasnoyarsk 660022, Russia; and Research Institute of Molecular Medicine and Pathobiochemistry, Krasnoyarsk State Medical University named after Prof. V.F. Voino-Yasenetsky, Krasnoyarsk 660022, Russia, e-mail:

During writing this review, the authors were supported by the grant given by the Russian Foundation for Basic Research and the Krasnoyarsk Regional Foundation for Supporting Research and Technological Activity (Russian Federation), project N 13-04-98091.

References

Abbott, N.J., Rönnbäck, L., and Hansson, E. (2006). Astrocyte-endothelial interactions at the blood-brain barrier. Nat. Rev. Neurosci. 7, 41–53.10.1038/nrn1824Search in Google Scholar PubMed

Alvarez-Maubecin, V., Garcia-Hernandez, F., Williams, J.T., and Van Bockstaele, E.J. (2000). Functional coupling between neurons and glia. J. Neurosci. 20, 4091–4098.10.1523/JNEUROSCI.20-11-04091.2000Search in Google Scholar

Aksoy, P., Escande, C., White, T.A., Thompson, M., Soares, S., Benech, J.C. and Chini, E.N. (2006). Regulation of SIRT 1 mediated NAD dependent deacetylation: a novel role for the multifunctional enzyme CD38. Biochem. Biophys. Res. Commun. 349, 353–359.10.1016/j.bbrc.2006.08.066Search in Google Scholar PubMed

Azarashvili, T., Baburina, Y., Grachev, D., Krestinina, O., Evtodienko, Y., Stricker, R., and Reiser, G. (2011). Calcium-induced permeability transition in rat brain mitochondria is promoted by carbenoxolone through targeting connexin43. Am. J. Physiol. Cell. Physiol. 300, C707–C720.10.1152/ajpcell.00061.2010Search in Google Scholar PubMed

Bao, B., Jiang, J., Yanase, T., Nishi, Y., and Morgan, J.R. (2011). Connexon-mediated cell adhesion drives microtissue self-assembly. FASEB J. 25, 255–264.10.1096/fj.10-155291Search in Google Scholar PubMed PubMed Central

Barbe, M.T., Monyer, H., and Bruzzone, R. (2006). Cell-cell communication beyond connexins: the pannexin channels. Physiology (Bethesda) 21, 103–114.10.1152/physiol.00048.2005Search in Google Scholar PubMed

Barker, A.J. and Ullian, E.M. (2008). New roles for astrocytes in developing synaptic circuits. Communicative Integrative Biol. 1, 207–211.10.4161/cib.1.2.7284Search in Google Scholar PubMed PubMed Central

Behrens, J., Kameritsch, P., Wallner, S., Pohl, U., and Pogoda, K. (2010). The carboxyl tail of Cx43 augments p38 mediated cell migration in a gap junction-independent manner. Eur. J. Cell Biol. 89, 828–838.10.1016/j.ejcb.2010.06.003Search in Google Scholar PubMed

Boengler, K., Ruiz-Meana, M., Gent, S., Ungefud, E., Soetkamp, D., Miro-Casa, E., Cabestrero, A., Fernandez-Sanz, C., Semenzato, M., Di Lisa, F., et al. (2012). Mitochodnrial connexin 43 impacts on respiratory complex I activity and mitochondrial oxygen consumption. J. Cell Mol. Med. 16, 1649–1655.10.1111/j.1582-4934.2011.01516.xSearch in Google Scholar PubMed PubMed Central

Boengler, K., Ungefug, E., Heusch, G., Leybaert, L., and Schulz, R. (2013). Connexin 43 impacts on mitochondrial potassium uptake. Front. Pharmacol. 4, 1–5, doi: 10.3389/fphar.2013.00073.10.3389/fphar.2013.00073Search in Google Scholar PubMed PubMed Central

Bonkowski, D., Katyshev, V., Balabanov, R.D., Borisov, A., and Dore-Duffy, P. (2011). The CNS microvascular pericyte: pericyte-astrocyte crosstalk in the regulation of tissue survival. Fluids Barriers CNS 8, 8.10.1186/2045-8118-8-8Search in Google Scholar PubMed PubMed Central

Boucher, S. and Bennett, S.A. (2003). Differential connexin expression, gap junction intercellular coupling, and hemichannel formation in NT2/D1 human neural progenitors and terminally differentiated hNT neurons. J. Neurosci. Res. 72, 393–404.10.1002/jnr.10575Search in Google Scholar PubMed

Bruzzone, S., Franco, L., Guida, L., Zocchi, E., Contini, P., Bisso, A., Usai, C., and De Flora, A. (2001). A self-restricted CD38-connexin 43 cross-talk affects NAD+ and cyclic ADP-ribose metabolism and regulates intracellular calcium in 3T3 fibroblasts. J. Biol. Chem. 276, 48300–48308.10.1074/jbc.M107308200Search in Google Scholar PubMed

Burns, A.R., Phillips, S.C., and Sokoya, E.M. (2012). Pannexin protein expression in the rat middle cerebral artery. J. Vsc. Res. 49, 101–110.10.1159/000332329Search in Google Scholar PubMed PubMed Central

Cao, X., Li, L.P., Qin, X.H., Li, S.J., Zhang, M., Wang, Q., Hu, H.H., Fang, Y.Y., Gao, Y.B., Li, X.W., et al. (2013). Astrocytic adenosine 5′-triphosphate release regulates the proliferation of neural stem cells in the adult hippocampus. Stem Cells. 31, 1633–1643.10.1002/stem.1408Search in Google Scholar PubMed

Chambers, R.A., Potenza, M.N., Hoffman, R.E., and Miranker, W. (2004). Stimulated apoptosis/neurogenesis regulates learning and memory capabilities of adaptive neural networks. Neuropsychopharmacol. 29, 747–758.10.1038/sj.npp.1300358Search in Google Scholar PubMed

Cheng, A., Tang, H., Cai, J., Zhu, M., Zhang, X., Rao, M., and Mattson, M.P. (2004). Gap junctional communication is required to maintain mouse cortical neural progenitor cells in a proliferative state. Dev. Biol. 272, 203–216.10.1016/j.ydbio.2004.04.031Search in Google Scholar PubMed

Chojnacki, A.K., Mak, G.K., and Weiss, S. (2009). Identity crisis for adult periventricular neural stem cells: subventricular zone astrocytes, ependymal cells or both? Nature Rev. Neurosci. 10, 153–163.Search in Google Scholar

Cina, C., Bechberger, J.F., Ozog, M.A., and Naus, C.C. (2007). Expression of connexins in embryonic mouse neocortical development. J. Comp. Neurol. 504, 298–313.10.1002/cne.21426Search in Google Scholar PubMed

Cina, C., Maass, K., Theis, M., Willecke, K., Bechberger, J.F., and Naus, C.C. (2009). Involvement of the cytoplasmic C-terminal domain of connexin43 in neuronal migration. J. Neurosci. 29, 2009–2021.10.1523/JNEUROSCI.5025-08.2009Search in Google Scholar PubMed PubMed Central

Cole, A.R. (2012). GSK3 as a sensor determining cell fate in the brain. Front. Mol. Neurosci. 5, 4.10.3389/fnmol.2012.00004Search in Google Scholar PubMed PubMed Central

Cunningham, L.A., Candelario, K., and Li, L. (2012). Roles for HIF-1α in neural stem cell function and the regenerative response to stroke. Behav. Brain Res. 227, 410–417.10.1016/j.bbr.2011.08.002Search in Google Scholar

Da Silva, P.G., Benton, J.L., Beltz, B.S., and Allodi, S. (2012). Adult neurogenesis: ultrastructure of a neurogenic niche and neurovascular relationships. PLoS One 7, e39267.Search in Google Scholar

Danesh-Meyer, H.V., Kerr, N.M., Zhang, J., Eady, E.K., O’Carroll, S.J., Nicholson, J.F.B., Johnson, C.S., and Green, C.R. (2011). Connexin43 mimetic peptide reduces vascular leak and retinal ganglion cell death following retinal ischemia. Brain 135, 506–520.10.1093/brain/awr338Search in Google Scholar

De Bock, M., Wang, N., Decrock, E., Bol, M., Gadicherla, A.K., Culot, M., Cecchelli, R., Bultynck, G., and Leybaert, L. (2013). Endothelial calcium dynamics, connexin channels and blood-brain barrier function. Prog. Neurobiol. 108, 1–20.10.1016/j.pneurobio.2013.06.001Search in Google Scholar

Deng, W., Aimone, J.B., and Gage, F.H. (2010). New neurons and new memories: how does adult hippocampal neurogenesis affect learning and memory. Nat. Rev. Neurosci. 11, 339–350.10.1038/nrn2822Search in Google Scholar

Dermietzel, R., Traub, O., Hwang, T.K., Beyer, E., Bennett, M.V., Spray, D.C., and Willecke, K. (1989). Differential expression of three gap junction proteins in developing and mature brain tissues. Proc. Natl. Acad. Sci. USA 86, 10148–52.10.1073/pnas.86.24.10148Search in Google Scholar

Dermietzel, R., Gao, Y., Scemes, E., Vieira, D., Urban, M., Kremer, M., Bennett, M.V., and Spray, D.C. (2000). Connexin43 null mice reveal that astrocytes express multiple connexins. Brain Res. Brain Res. Rev. 32, 45–56.10.1016/S0165-0173(99)00067-3Search in Google Scholar

Dore-Duffy, P. (2008). Pericytes: pluripotent cells of the blood brain barrier. Curr. Pharm. Des. 14, 1581–1593.10.2174/138161208784705469Search in Google Scholar

Drewes, L.R. (2012). Making connexons in the neurovascular unit. J. Cerebral Blood Flow Metab. 32, 1455–1456.10.1038/jcbfm.2012.44Search in Google Scholar

Dudek, F.E., Gribkoff, V.K., Olson, J.E., and Hertzberg, E.L. (1988). Reduction of dye coupling in glial cultures by microinjection of antibodies against the liver gap junction polypeptide. Brain Res. 439, 275–80.10.1016/0006-8993(88)91484-9Search in Google Scholar

Duval, N., Gomès, D., Calaora, V., Calabrese, A., Meda, P., and Bruzzone, R. (2002). Cell coupling and Cx43 expression in embryonic mouse neural progenitor cells. J. Cell Sci. 115(Pt 16), 3241–3251.10.1242/jcs.115.16.3241Search in Google Scholar PubMed

Ebong, E.E. and Depaola, N. (2013). Specificity in the participation of connexin proteins in flow-induced endothelial gap junction communication. Pflugers. Arch. 465, 1293–1302.10.1007/s00424-013-1245-9Search in Google Scholar PubMed

Elias, L.A. and Kriegstein, A.R. (2008). Gap junctions: multifaceted regulators of embryonic cortical development. Trends Neurosci. 31, 243–250.10.1016/j.tins.2008.02.007Search in Google Scholar PubMed PubMed Central

Elias, L.A., Wang, D.D., and Kriegstein, A.R. (2007). Gap junction adhesion is necessary for radial migration in the neocortex. Nature 448, 901–907.10.1038/nature06063Search in Google Scholar PubMed

Elias, L.A.B., Turmaine, M., Parnavelas, J.G., and Kriegstein, A.R. (2010). Cx43 mediates the tangential to radial migratory switch in ventrally derived cortical interneurons. J. Neurosci. 30, 7072–7077.10.1523/JNEUROSCI.5728-09.2010Search in Google Scholar PubMed PubMed Central

Elzarrad, M.K., Haroon, A., Willecke, K., Dobrowolski, R., Gillespie, M.N., and Al-Mehdi, A.B. (2008). Connexin-43 upregulation in micrometastases and tumor vasculature and its role in tumor cell attachment to pulmonary endothelium. BMC Med. 6, 20.10.1186/1741-7015-6-20Search in Google Scholar PubMed PubMed Central

Evans, W.H., Bultynck, G., and Leybaert, L. (2012). Manipulating connexin communication channels: use of peptidomimetics and the translational outputs. J. Membr. Biol. 245, 437–449.10.1007/s00232-012-9488-5Search in Google Scholar PubMed PubMed Central

Ezan, P., André, P., Cisternino, S., Saubaméa, B., Boulay, A.C., Doutremer, S., Thomas, M.A., Quenech’du, N., Giaume, C., and Cohen-Salmon, M. (2012). Deletion of astroglial connexins weakens the blood-brain barrier. J. Cereb. Blood Flow Metab. 32, 1457–1467.10.1038/jcbfm.2012.45Search in Google Scholar PubMed PubMed Central

Faustmann, P.M., Haase, C.G., Romberg, S., Hinkerohe, D., Szlachta, D., Smikalla, D., Krause, D., and Dermietzel, R. (2003). Microglia activation influences dye coupling and Cx43 expression of the astrocytic network. Glia 42, 101–108.10.1002/glia.10141Search in Google Scholar PubMed

Feine, I., Pinkas, I., Salomon, Y., and Schertz, A. (2012). Local oxidative stress expansion through endothelial cells – a key role for gap junction intercellular communication. PLoS One 7, e41633.10.1371/journal.pone.0041633Search in Google Scholar PubMed PubMed Central

Fields, R.D. and Burnstock, G. (2006). Purinergic signalling in neuron-glia interactions. Nat. Rev. Neurosci. 7, 423–436.10.1038/nrn1928Search in Google Scholar PubMed PubMed Central

Figueroa, X. F., Lillo, M. A., Gaete, P. S., Riquelme, M., and Sáez, J. C. (2013). Diffusion of nitric oxide across cell membranes of the vascular wall requires specific connexin-based channels. Neuropharmacology. 75, 471–478. doi: 10.1016/ j.neuropharm.2013.02.022.10.1016/j.neuropharm.2013.02.022Search in Google Scholar

Francis, R., Xu., X., Park, H., Wei, C.J., Chang, S., Chatterjee, B., and Lo, C. (2011). Connexin43 modulates cell polarity and directional cell migration by regulating microtubule dynamics. PLoS One 6, e26379.10.1371/journal.pone.0026379Search in Google Scholar

Freeman, M.R. (2010). Specification and morphogenesis of astrocytes. Science 330, 774–778.10.1126/science.1190928Search in Google Scholar

Freitas, A.S., Xavier, A.L., Furtado, C.M., Hedin-Pereira, C., Fróes, M.M., and Menezes, J.R. (2012). Dye coupling and connexin expression by cortical radial glia in the early postnatal subventricular zone. Dev. Neurobiol. 72, 1482–1497.10.1002/dneu.22005Search in Google Scholar

Frisch, C., Theis, M., De Souza Silva, M.A., Dere, E., Söhl, G., Teubner, B., Namestkova, K., Willecke, K., and Huston, J.P. (2003). Mice with astrocyte-directed inactivation of connexin43 exhibit increased exploratory behaviour, impaired motor capacities, and changes in brain acetylcholine levels. Eur. J. Neurosci. 18, 2313–2318.10.1046/j.1460-9568.2003.02971.xSearch in Google Scholar

Fushiki, S., Perez Velazquez, J.L., Zhang, L., Bechberger J.F., Carlen, P.L., and Naus, C.C. (2003). Changes in neuronal migration in neocortex of connexin43 null mutant mice, J. Neuropathol. Exp. Neurol. 62, 304–314.10.1093/jnen/62.3.304Search in Google Scholar

Gärtner, C., Ziegelhöffer, B., Kostelka, M., Stepan, H., Mohr, F.W., and Dhein, S. (2012). Knock-down of endothelial connexins impairs angiogenesis. Pharmacol. Res. 65, 347–357.10.1016/j.phrs.2011.11.012Search in Google Scholar

Giaume, C., Koulakoff, A., Roux, L., Holcman, D., and Rouach, N. (2010). Astroglial networks: a step further in neuroglial and gliovascular interactions. Nat. Rev. Neurosci. 11, 87–99.10.1038/nrn2757Search in Google Scholar

Giaume, C., Leybaert, L., Naus, C.C., and Sáez, J.C. (2013). Connexin and pannexin hemichannels in brain glial cells: properties, pharmacology, and roles. Front. Pharmacol. 4, 88.10.3389/fphar.2013.00088Search in Google Scholar

Giepmans, B.N., Verlaan, I., Hengeveld, T., Janssen, H., Calafat, J., Falk, M.M., and Moolenaar, W.H. (2001). Gap junction protein connexin-43 interacts directly with microtubules. Curr. Biol. 11, 1364–1368.10.1016/S0960-9822(01)00424-9Search in Google Scholar

Glaser, T., Resende, R.R., and Ulrich, H. (2013). Implications of purinergic receptor-mediated intracellular calcium transients in neural differentiation. Cell Commun. Signal. 11, 12.10.1186/1478-811X-11-12Search in Google Scholar PubMed PubMed Central

Goldberg, J.S. and Hirschi, K.K. (2009). Diverse roles of the vasculature within the neural stem cell niche. Regen. Med. 4, 879–897.10.2217/rme.09.61Search in Google Scholar PubMed PubMed Central

Han, Y., Yu, H.X., Sun, M.L., Wang, Y., Xi, W., Yu, Y.Q. (2013). Astrocyte-restricted disruption of connexin-43 impairs neuronal plasticity in mouse barrel cortex. Eur J Neurosci. 2013. doi: 10.1111/ejn.12394.10.1111/ejn.12394Search in Google Scholar PubMed

Hatakeyama, T., Dai, P., Harada, Y., Hino, H., Tsukahara, F., Maru, Y., Otsuji, E., and Takamatsu, T. (2013). Connexin43 functions as a novel interacting partner of heat shock cognate protein 70. Sci. Rep. 3, 2719.10.1038/srep02719Search in Google Scholar PubMed PubMed Central

Hawkins, B.T. and Davis, T.P. (2005). The blood-brain barrier/neurovascular unit in health and disease. Pharmacol. Rev. 57, 173–85.10.1124/pr.57.2.4Search in Google Scholar PubMed

Higashida, H., Salmina, A.B., Olovyannikova, R.Y., Hashii, M., Yokoyama, S., Koizumi, K., Jin, D., Liu, H.X., Lopatina, O., Amina, S., et al. (2007). Cyclic ADP-ribose as a universal calcium signal molecule in the nervous system. Neurochem. Int. 51, 192–199.10.1016/j.neuint.2007.06.023Search in Google Scholar PubMed

Higashida, H, Yokoyama, S., Kikuchi, M., and Munesue, T. (2012). CD38 and its role in oxytocin secretion and social behavior. Horm. Behav. 61, 351–358.10.1016/j.yhbeh.2011.12.011Search in Google Scholar PubMed

Iacobas, D.A., Iacobas, S., Urban-Maldonado, M., and Spray, D.C. (2005). Sensitivity of the brain transcriptome to connexin ablation. Biochim. Biophys. Acta. 1711, 183–196.10.1016/j.bbamem.2004.12.002Search in Google Scholar PubMed

Iacobas, D.A., Iacobas, S., and Spray, D.C. (2007a). Connexin-dependent transcellular transcriptomic networks in mouse brain. Prog. Biophys. Mol. Biol. 94, 169–185.10.1016/j.pbiomolbio.2007.03.015Search in Google Scholar PubMed

Iacobas, D.A., Suadicani, S.O., Iacobas, S., Chrisman, C., Cohen, M.A., Spray, D.C., Scemes, E. (2007b) Gap junction and purinergic P2 receptor proteins as a functional unit: insights from transcriptomics. J. Membr. Biol. 217, 83–91.10.1007/s00232-007-9039-7Search in Google Scholar PubMed

Iacobas D.A., Urban-Maldonado M., Iacobas S., Scemes E., and Spray D.C. (2008). Array analysis of gene expression in connexin-43 null astrocytes. Cell Commun. Adhes. 15, 195–206.10.1080/15419060802014222Search in Google Scholar

Iacobas, S., Iacobas, D.A., Spray, D.C., and Scemes, E. (2012). The connexin43-dependent transcriptome during brain development: importance of genetic background. Brain Res. 1487, 131–139.10.1016/j.brainres.2012.05.062Search in Google Scholar

Ii, M., Nishimura, H., Sekiguchi, H., Kamei, N., Yokoyama, A., Horii, M., and Asahara, T. (2009). Concurrent vasculogenesis and neurogenesis from adult neural stem cells. Circ. Res. 105, 860–868.10.1161/CIRCRESAHA.109.199299Search in Google Scholar

Imbeault, S., Gauvin, L.G., Toeg, H.D., Pettit, A., Sorbara, C.D., Migahed, L., DesRoches, R., Menzies, A.S., Nishii, K., Paul, D.L., et al. (2009). The extracellular matrix controls gap junction protein expression and function in postnatal hippocampal neural progenitor cells. BMC Neurosci. 10, 13.10.1186/1471-2202-10-13Search in Google Scholar

Jaeger, A., Baake, J., Weiss, D.G., and Kriehuber, R. (2013). Glycogen synthase kinase-3beta regulates differentiation-induced apoptosis of human neural progenitor cells. Int. J. Dev. Neurosci. 31, 61–68.10.1016/j.ijdevneu.2012.10.005Search in Google Scholar

Jeyaraman, M.M., Fandrich, R.R., and Kardami, E. (2013). Together and apart: inhibition of DNA synthesis by connexin-43 and its relationship to transforming growth factor β. Front. Pharmacol. 4, 90.10.3389/fphar.2013.00090Search in Google Scholar

Johnstone, S.R., Best, A.K., Wright, C.S., Isakson, B.E., Errington, R.J., and Martin, P.E. (2010). Enhanced connexin 43 expression delays intra-mitotic duration and cell cycle traverse independently of gap junction channel function. J. Cell Biochem. 110, 772–782.10.1002/jcb.22590Search in Google Scholar

Johnstone, S.R., Kroncke, B.M., Straub, A.C., Best, A.K., Dunn, C.A., Mitchell, L.A., Peskova Y., Nakamoto, R.K., Koval, M., Lo, C.W., et al. (2012). MAPK phosphorylation of connexin 43 promotes binding of cyclin E and smooth muscle cell proliferation. Circ. Res. 111, 201–211.10.1161/CIRCRESAHA.112.272302Search in Google Scholar

Kalvelyte, A., Imbrasaite, A., Bukauskiene, A., Verselis, V.K., and Bukauskas, F.F. (2003). Connexins and apoptotic transformation. Biochem. Pharmacol. 66, 1661–1672.10.1016/S0006-2952(03)00540-9Search in Google Scholar

Kameritsch, P., Khandoga, N., Pohl, U., and Pogoda, K. (2013). Gap junctional communication promotes apoptosis in a connexin-type-dependent manner. Cell Death Dis. 4, e584.10.1038/cddis.2013.105Search in Google Scholar PubMed PubMed Central

Kar, R., Riquelme, M.A., Werner, S., Jiang, J.X. (2013). Connexin 43 channels protect osteocytes against oxidative stress-induced cell death. J. Bone Miner. Res. 28, 1611–1621.10.1002/jbmr.1917Search in Google Scholar PubMed PubMed Central

Kaufman, J., Gordon, C., Bergamaschi, R., Wang, H.Z., Cohen, I.S., Valiunas, V., and Brink, P.R. (2013). The effects of the histone deacetylase inhibitor 4-phenylbutyrate on gap junction conductance and permeability. Front. Pharmacol. 4, 111.10.3389/fphar.2013.00111Search in Google Scholar PubMed PubMed Central

Khan, Z., Akhtar, M., Asklund, T., Juliusson, B., Almqvist, P.M., and Ekström, T.J. (2007). HDAC inhibition amplifies gap junction communication in neural progenitors: potential for cell-mediated enzyme prodrug therapy. Exp. Cell Res. 313, 2958–2967.10.1016/j.yexcr.2007.05.004Search in Google Scholar PubMed

Klotz, L.O. (2012). Posttranscriptional regulation of connexin-43 expression. Arch. Biochem. Biophys. 524, 23–29.10.1016/j.abb.2012.03.012Search in Google Scholar PubMed

Kojima, T., Hirota, Y., Ema, M., Takahashi, S., Miyoshi, I., Okano, H., and Sawamoto, K. (2010). Subventricular zone-derived neural progenitor cells migrate along a blood vessel scaffold toward the post-stroke striatum. Stem Cells. 28, 545–554.10.1002/stem.306Search in Google Scholar PubMed

Koulakoff, A., Ezan, P., and Giaume, C. (2008). Neurons control the expression of connexin 30 and connexin 43 in mouse cortical astrocytes. Glia 56, 1299–1311.10.1002/glia.20698Search in Google Scholar PubMed

Kowiański, P., Lietzau, G., Steliga, A., Waśkow, M., and Moryś, J. (2013). The astrocytic contribution to neurovascular coupling – still more questions than answers? Neurosci. Res. 75, 171–183.Search in Google Scholar

Kunze, A., Congreso, M.R., Hartmann, C., Wallraff-Beck, A., Hüttmann, K., Bedner, P., Requardt, R., Seifert, G., Redecker, C., Willecke, K., et al. (2009). Connexin expression by radial glia-like cells is required for neurogenesis in the adult dentate gyrus. Proc. Natl. Acad. Sci. USA 106, 11336–1141.10.1073/pnas.0813160106Search in Google Scholar PubMed PubMed Central

Lacar, B., Young, S.Z., Platel, J.C., and Bordey, A. (2011). Gap junction-mediated calcium waves define communication networks among murine postnatal neural progenitor cells. Eur. J. Neurosci. 34, 1895–18905.10.1111/j.1460-9568.2011.07901.xSearch in Google Scholar PubMed PubMed Central

Lacar, B., Herman, P., Platel, J.C., Kubera, C., Hyder, F., and Bordey, A. (2012a). Neural progenitor cells regulate capillary blood flow in the postnatal subventricular zone. J. Neurosci. 32, 16435–16448.10.1523/JNEUROSCI.1457-12.2012Search in Google Scholar PubMed PubMed Central

Lacar, B., Herman, P., Hartman, N.W., Hyder, F., and Bordey, A. (2012b). S phase entry of neural progenitor cells correlates with increased blood flow in the young subventricular zone. PLoS One 7, e31960.10.1371/journal.pone.0031960Search in Google Scholar PubMed PubMed Central

Lan, Y., Liu, B., Yao, H., Li, F., Weng, T., Yang, G., Li, W., Cheng, X., Mao, N. and Yang, X. (2007). Essential role of endothelial Smad4 in vascular remodeling and integrity. Mol. Cell Biol. 27, 7683–7692.10.1128/MCB.00577-07Search in Google Scholar

Lange, S., Trost, A., Tempfer, H., Bauer, H.C., Bauer, H., Rohde, E., Reitsamer, H.A., Franklin, R.J., Aigner, L., and Rivera, F.J. (2013). Brain pericyte plasticity as a potential drug target in CNS repair. Drug Discov. Today 18, 456–463.10.1016/j.drudis.2012.12.007Search in Google Scholar

Lemcke, H. and Kuznetsov, S.A. (2013). Involvement of connexin43 in the EGF/EGFR signalling during self-renewal and differentiation of neural progenitor cells. Cell Signal 25, 2676–2684.10.1016/j.cellsig.2013.08.030Search in Google Scholar

Lemcke, H., Nittel, M.L., Weiss, D.G., and Kuznetsov, S.A. (2013). Neuronal differentiation requires a biphasic modulation of gap junctional intercellular communication caused by dynamic changes of connexin43 expression. Eur. J. Neurosci. 38, 2218–2228.10.1111/ejn.12219Search in Google Scholar

Leung, D.S., Unsicker, K., and Reuss, B. (2002). Expression and developmental regulation of gap junction connexins cx26, cx32, cx43 and cx45 in the rat midbrain-floor. Int. J. Dev. Neurosci. 20, 63–75.10.1016/S0736-5748(01)00056-9Search in Google Scholar

Levin, M. (2007). Gap junctional communication in morphogenesis. Prog. Biophys. Mol. Biol. 94, 186–206.10.1016/j.pbiomolbio.2007.03.005Search in Google Scholar PubMed PubMed Central

Li, K., Chi, Y., Gao, K., Yan, Q., Matsue, H., Takeda, M., Kitamura, M., and Yao, J. (2013). Connexin43 hemichannel-mediated regulation of connexin43. PLoS One 8, e58057.10.1371/journal.pone.0058057Search in Google Scholar PubMed PubMed Central

Liebmann, M., Stahr, A., Guenther, M., Witte, O.W., and Frahm, C. (2013). Astrocytic Cx43 and Cx30 differentially modulate adult neurogenesis in mice. Neurosci. Lett. 545, 40–45.10.1016/j.neulet.2013.04.013Search in Google Scholar PubMed

Liebner, S., Czupalla, C.J., and Wolburg, H. (2011). Current concepts of blood-brain barrier development. Int. J. Dev. Biol. 55, 467–476.10.1387/ijdb.103224slSearch in Google Scholar PubMed

Liu, X., Bolteus, A.J., Balkin, D.M., Henschel, O., and Bordey, A. (2006). GFAP-expressing cells in the postnatal subventricular zone display a unique glial phenotype intermediate between radial glia and astrocytes. Glia 54, 394–410.10.1002/glia.20392Search in Google Scholar PubMed

Liu, X., Hashimoto-Torii, K., Torii, M., Ding, C., and Rakic, P. (2010). Gap junctions/hemichannels modulate interkinetic nuclear migration in the forebrain precursors. J. Neurosci. 30, 4197–4209.10.1523/JNEUROSCI.4187-09.2010Search in Google Scholar PubMed PubMed Central

Liu, Y., Namba, T., Liu, J., Suzuki, R., Shioda, S., and Seki, T. (2010). Glial fibrillary acidic protein-expressing neural progenitors give rise to immature neurons via early intermediate progenitors expressing both glial fibrillary acidic protein and neuronal markers in the adult hippocampus. Neurosci. 166, 241–251.10.1016/j.neuroscience.2009.12.026Search in Google Scholar PubMed

Liu, X., Sun, L., Torii, M., and Rakic, P. (2012). Connexin 43 controls the multipolar phase of neuronal migration to the cerebral cortex. Proc. Natl. Acad. Sci. USA 109, 8280–8285.10.1073/pnas.1205880109Search in Google Scholar PubMed PubMed Central

Lok, J., Gupta, P., Guo, S., Kim, W.J., Whalen, M.J., van Leyen, K., and Lo, E.H. (2007). Cell-cell signaling in the neurovascular unit. Neurochem. Res. 32, 2032–2045.10.1007/s11064-007-9342-9Search in Google Scholar PubMed

Mäe, M., Armulik, A., and Betsholtz, C. (2011). Getting to know the cast – cellular interactions and signaling at the neurovascular unit. Curr. Pharm. Des. 17, 2750–2754.10.2174/138161211797440113Search in Google Scholar PubMed

Malmersjö, S., Rebellato, P., Smedler, E., Planert, H., Kanatani, S., Liste, I., Nanou, E., Sunner, H., Abdelhady, S., Zhang, S., et al. (2013a). Neural progenitors organize in small-world networks to promote cell proliferation. Proc. Natl. Acad. Sci. USA 110, E1524–E1532.10.1073/pnas.1220179110Search in Google Scholar PubMed PubMed Central

Malmersjö, S., Rebellato, P., Smedler, E., and Uhlén, P. (2013b). Small-world networks of spontaneous Ca(2+) activity. Commun. Integr. Biol. 6, e24788.10.4161/cib.24788Search in Google Scholar PubMed PubMed Central

Martin, P.E. and Evans, W.H. (2004). Incorporation of connexins into plasma membranes and gap junctions. Cardiovasc Res. 62, 378–387.10.1016/j.cardiores.2004.01.016Search in Google Scholar PubMed

Matyash, M., Matyash, V., Nolte, C., Sorrentino, V., and Kettenmann, H. (2002). Requirement of functional ryanodine receptor type 3 for astrocyte migration. FASEB J. 16, 84–86.10.1096/fj.01-0380fjeSearch in Google Scholar PubMed

Messemer, N., Kunert, C., Grohmann, M., Sobottka, H., Nieber, K., Zimmermann, H., Franke, H., Nörenberg, W., Straub, I., Schaefer, M., et al. (2013). P2X7 receptors at adult neural progenitor cells of the mouse subventricular zone. Neuropharmacology 73, 122–137.10.1016/j.neuropharm.2013.05.017Search in Google Scholar PubMed

Morita, M., Saruta, C., Kozuka, N., Okubo, Y., Itakura, M., Takahashi, M., and Kudo, Y. (2007). Dual regulation of astrocyte gap junction hemichannels by growth factors and a pro-inflammatory cytokine via the mitogen-activated protein kinase cascade. Glia 55, 508–515.10.1002/glia.20471Search in Google Scholar PubMed

Moriyama, Y., Takagi, N., Itokawa, C., and Tanonaka, K. (2013). Injection of neural progenitor cells attenuates decrease in level of connexin 43 in brain capillaries after cerebral ischemia. Neurosci. Lett. 543,152–156.10.1016/j.neulet.2013.03.053Search in Google Scholar PubMed

Morrens, J., Van Den Broeck, W., and Kempermann, G. (2012). Glial cells in adult neurogenesis. Glia 60, 159–174.10.1002/glia.21247Search in Google Scholar PubMed

Murugesan, N., Demarest, T.G., Madri, J.A., and Pachter, J.S. (2012). Brain regional angiogenic potential at the neurovascular unit during normal aging. Neurobiol. Aging 33, 1004.e1–16.10.1016/j.neurobiolaging.2011.09.022Search in Google Scholar PubMed PubMed Central

Nagasawa, K., Chiba, H., Fujita, H., Kojima, T., Saito, T., Endo, T., and Sawada, N. (2006). Possible involvement of gap junctions in the barrier function of tight junctions of brain and lung endothelial cells. J. Cell Physiol. 208, 123–132.10.1002/jcp.20647Search in Google Scholar PubMed

Nagy, J.I. and Li, W.E. (2000). A brain slice model for in vitro analyses of astrocytic gap junction and connexin 43 regulation: actions of ischemia, glutamate and elevated potassium. Eur. J. Neurosci. 12, 4567–72.Search in Google Scholar

Nelson, B.R., Hodge, R.D., Bedogni, F., and Hevner, R.F. (2013). Dynamic interactions between intermediate neurogenic progenitors and radial glia in embryonic mouse neocortex: potential role in Dll1-Notch signaling. J. Neurosci. 33, 9122–9139.10.1523/JNEUROSCI.0791-13.2013Search in Google Scholar PubMed PubMed Central

Ogundele, O.M., Omoaghe, A.O., Ajonijebu, D.C., Ojo, A.A., Fabiyi, T.D., Olajide, O.J., Falode, D.T., and Ademiyi, P.A. (2013). Glia activation and its role in oxidative stress. Metab. Brain Dis. Nov. 13. doi: 10.1007/s11011-013-9446-7.10.1007/s11011-013-9446-7Search in Google Scholar PubMed

Okamoto, T., Akiyama, M., Takeda, M., Gabazza, E.C., Hayashi, T., and Sizuki, K. (2009). Connexin32 is expressed in vascular endothelioal cells and participates in gap-junction intercellular5 communication. Biochem. Biophys. Res. Commun. 382, 264–268.10.1016/j.bbrc.2009.02.148Search in Google Scholar PubMed

Okamoto, T., Akiyama, M., Takeda, M., Akita, N., Yoshida, K., Hayashi, T., and Suzuki, K. (2011). Connexin32 protects against vascular inflammation by modulating inflammatory cytokine expression by endothelial cells. Exp. Cell Res. 317, 348–355.10.1016/j.yexcr.2010.10.018Search in Google Scholar PubMed

Orellana, J.A., Froger, N., Ezan, P., Jiang, J.X., Bennett, M.V., Naus, C.C., Giaume, C., and Sáez, J.C. (2011). ATP and glutamate released via astroglial connexin 43 hemichannels mediate neuronal death through activation of pannexin 1 hemichannels. J. Neurochem. 118, 826–840.10.1111/j.1471-4159.2011.07210.xSearch in Google Scholar PubMed PubMed Central

Orellana, J.A., Sáez, P.J., Cortés-Campos, C., Elizondo, R.J., Shoji, K.F., Contreras-Duarte, S., Figueroa, V., Velarde, V., Jiang, J.X., Nualart, F., et al. (2012). Glucose increases intracellular free Ca2+ in tanycytes via ATP released through connexin 43 hemichannels. Glia 60, 53–68.10.1002/glia.21246Search in Google Scholar PubMed PubMed Central

Oyamada, M., Takebe, K., Endo, A., Hara, S., and Oyamada Y. (2013). Connexin expression and gap-junctional intercellular communication in ES cells and iPS cells. Front. Pharmacol. 4, 85.10.3389/fphar.2013.00085Search in Google Scholar PubMed PubMed Central

Palaez, D., Huang, C.Y., and Cheung, H.S. (2013). Isolation of pluripotent neural crest-derived stem cells from adult human tissues by connexin-43 enrichment. Stem Cells Dev. 22, 2906–2914.10.1089/scd.2013.0090Search in Google Scholar PubMed

Panchin, Y.V. (2005). Evolution of gap junction proteins–the pannexin alternative. J. Exp. Biol. 208, 1415–1419.10.1242/jeb.01547Search in Google Scholar PubMed

Parekkadan, B., Berdichevsky,Y., Irimia, D., Leeder, A., Yarmush, G., Toner, M., Levine, J.B., and Yarmush, M.L. (2008). Cell-cell interaction modulates neuroectodermal specification of embryonic stem cells. Neurosci. Lett. 438, 190–195.10.1016/j.neulet.2008.03.094Search in Google Scholar PubMed PubMed Central

Parpura, V., Heneka, M.T., Montana, V., Oliet, S.H., Schousboe, A., Haydon, P.G., Stout, R.F. Jr., Spray, D.C., Reichenbach, A., Pannicke, T., et al. (2012). Glial cells in (patho)physiology. J. Neurochem. 121, 4–27.10.1111/j.1471-4159.2012.07664.xSearch in Google Scholar PubMed PubMed Central

Pelligrino, D. A., Vetri, F., and Xu, H. L. (2011). Purinergic mechanisms in gliovascular coupling. Semin. Cell Dev. Biol. 22, 229–236.10.1016/j.semcdb.2011.02.010Search in Google Scholar PubMed PubMed Central

Penuela, S., Gehi, R., and Laird, D.W. (2013). The biochemistry and functions of pannexin channels. Biochimica et Biophys. Acta. 1828, 15–22.10.1016/j.bbamem.2012.01.017Search in Google Scholar PubMed

Plane, J.M., Andjelkovic, A.V., Keep, R.F., and Parent, J.M. (2010). Intact and injured endothelial cells differentially modulate postnatal murine forebrain neural stem cells. Neurobiol. Dis. 37, 218–227.10.1016/j.nbd.2009.10.008Search in Google Scholar PubMed PubMed Central

Porlan, E., Perez-Villalba, A., Delgado, A.C., and Ferrón, S.R. (2013). Paracrine regulation of neural stem cells in the subependymal zone. Arch. Biochem. Biophys. 534, 11–19.10.1016/j.abb.2012.10.001Search in Google Scholar PubMed

Rafalski, V.A. and Brunet, A. (2011) Energy metabolism in adult neural stem cell fate. Prog. Neurobiol. 93, 182–203.Search in Google Scholar

Rela, L. and Szczupak, L. (2004). Gap junctions: their importance for the dynamics of neural circuits. Mol. Neurobiol. 30, 341–357.10.1385/MN:30:3:341Search in Google Scholar

Retamal, M.A., Froger, N., Palacios-Prado, N., Ezan, P., Sáez, P.J., Sáez, J.C., and Giaume, C. (2007a). Cx43 hemichannels and gap junction channels in astrocytes are regulated oppositely by proinflammatory cytokines released from activated microglia. J. Neurosci. 27, 13781–13792.10.1523/JNEUROSCI.2042-07.2007Search in Google Scholar

Retamal, M.A., Schalper, K.A., Shoji, K.F., Bennett, M.V., and Sáez J.C. (2007b). Opening of connexin 43 hemichannels is increased by lowering intracellular redox potential. Proc. Natl. Acad. Sci. USA 104, 8322–8327.10.1073/pnas.0702456104Search in Google Scholar

Risau, W., Esser, S., and Engelhardt, B. (1998). Differentiation of blood-brain barrier endothelial cells. Pathol. Biol. (Paris). 46, 171–175.Search in Google Scholar

Rodriguez, E.M., Guerra, M.M., Vio, K., Gonzalez, C., Ortloff, A., Batiz, L.F., Rodriguez, S., Jara, M.C., Munoz, R.I., Ortega, E., et al. (2012). A cell junction pathology of neural stem cells leads to abnormal neurogenesis and hydrocephalus. Biol. Res. 45, 231–241.10.4067/S0716-97602012000300005Search in Google Scholar

Roitbak, T., Li, L., and Cunningham, L.A. (2008). Neural stem/progenitor cells promote endothelial cell morphogenesis and protect endothelial cells against ischemia via HIF-1alpha-regulated VEGF signaling. J. Cereb. Blood Flow Metab. 28, 1530–1542.10.1038/jcbfm.2008.38Search in Google Scholar

Roitbak, T., Surviladze, Z., and Cunningham, L.A. (2011). Continuous expression of HIF-1α in neural stem/progenitor cells. Cell Mol. Neurobiol. 31, 119–133.10.1007/s10571-010-9561-5Search in Google Scholar

Rottlaender, D., Boengler, K., Wolny, M., Michels, G., Endres-Becker, J., Motloch, L.J., Schwaiger, A., Buechert, A., Schulz, R., Heusch, G, et al. (2010). Connexin 43 acts as a cytoprotective mediator of signal transduction by stimulating mitochondrial KATP channels in mouse cardiomyocytes. J. Clin. Invest. 120, 1441–1453.10.1172/JCI40927Search in Google Scholar

Rottlaender, D., Boengler, K., Wolny, M., Schwaiger, A., Motloch, L.J., Ovize, M., Schulz, R., Heusch, G., and Hoppe, U.C. (2012). Glycogen synthase kinase 3β transfers cytoprotective signaling through connexin 43 onto mitochondrial ATP-sensitive K+ channels. Proc. Natl. Acad. Sci. USA 109, E242–E251.10.1073/pnas.1107479109Search in Google Scholar

Rouach, N., Koulakoff, A., and Giaume, C. (2004). Neurons set the tone of gap junctional communication in astrocytic networks. Neurochem. Int. 45, 265–272.10.1016/j.neuint.2003.07.004Search in Google Scholar

Rozental, R., Srinivas, M., Gökhan, S., Urban, M., Dermietzel, R., Kessler, J.A., Spray, D.C., and Mehler, M.F. (2000). Temporal expression of neuronal connexins during hippocampal ontogeny. Brain Res. Brain Res. Rev. 32, 57–71.10.1016/S0165-0173(99)00096-XSearch in Google Scholar

Salmina, A.B. (2009). Neuron-glia interactions as therapeutic targets in neurodegeneration. J Alzheimers. Dis. 16, 485–502.10.3233/JAD-2009-0988Search in Google Scholar

Salmina, A.B., Inzhutova, A.I., Morgun, A.V., Okuneva, O.S., Malinovskaia, N.A., Lopatina, O.L., Petrova, M.M., Taranushenko, T.E., Fursov, A.A., and Kuvacheva, N.V. (2012). [NAD+-converting enzymes in neuronal and glial cells: CD38 as a novel target for neuroprotection]. Vestn. Ross. Akad. Med. Nauk. 10, 29–37.10.15690/vramn.v67i10.413Search in Google Scholar

Santiago, M.F., Alcami, P., Striedinger, K.M., Spray, D.C., and Scemes, E. (2010). The carboxyl-terminal domain of connexin43 is a negative modulator of neuronal differentiation. J. Biol. Chem. 285, 11836–11845.10.1074/jbc.M109.058750Search in Google Scholar

Scemes, E. (2000). Components of astrocytic intercellular calcium signaling. Mol Neurobiol. 22, 167–79.10.1385/MN:22:1-3:167Search in Google Scholar

Scemes, E. (2008). Modulation of astrocyte P2Y1 receptors by the carboxyl terminal domain of the gap junction protein Cx43. Glia 56, 145–153.10.1002/glia.20598Search in Google Scholar PubMed PubMed Central

Scemes, E., Duval, N., and Meda, P. (2003). Reduced expression of P2Y1 receptors in connexin43-null mice alters calcium signaling and migration of neural progenitor cells. J. Neurosci. 23, 11444–11452.10.1523/JNEUROSCI.23-36-11444.2003Search in Google Scholar

Sengillo, J.D., Winkler, E.A., Walker, C.T., Sullivan, J.S., Johnson, M., and Zlokovic, B.V. (2013). Deficiency in mural vascular cells coincides with blood-brain barrier disruption in Alzheimer’s disease. Brain Pathol. 23, 303–310.10.1111/bpa.12004Search in Google Scholar PubMed PubMed Central

Sheng, W.S., Hu, S., Feng, A., and Rock, R.B. (2013). Reactive oxygen species from human astrocytes induced functional impairment and oxidative damage. Neurochem Res. 38, 2148–2159.10.1007/s11064-013-1123-zSearch in Google Scholar PubMed PubMed Central

Sild, M. and Van Horn, M.R. (2013). Astrocytes use a novel transporter to fill gliotransmitter vesicles with D-serine: evidence for vesicular synergy. J. Neurosci. 33, 10193–10194.10.1523/JNEUROSCI.1665-13.2013Search in Google Scholar PubMed PubMed Central

Simard, M., Arcuino, G., Takano, T., Liu, Q.S., and Nedergaard, M. (2003). Signaling at the gliovascular interface. J. Neurosci. 23, 9254–9262.10.1523/JNEUROSCI.23-27-09254.2003Search in Google Scholar

Sokoya, E.M., Burns, A.R., Setiawan, C.T., Coleman, H.A., Parkington, H.C., and Tare, M. (2006). Evidence for the involvement of myoendothelial gap junctions in EDHF-mediated relaxation in the rat middle cerebral artery. Am. J. Physiol. Heart Circ. Physiol. 291, H385–H393.10.1152/ajpheart.01047.2005Search in Google Scholar PubMed

Song, E.K., Rah, S.Y., Lee, Y.R., Yoo, C.H., Kim, Y.R., Yeom, J.H., Park, K.H., Kim, J.S., Kim, U.H., and Han, M.K. (2011). Connexin-43 hemichannels mediate cyclic ADP-ribose generation and its Ca2+-mobilizing activity by NAD+/cyclic ADP-ribose transport. J. Biol. Chem. 286, 44480–44490.10.1074/jbc.M111.307645Search in Google Scholar PubMed PubMed Central

Spray, D.C. and Iacobas, D.A. (2007). Organizational principles of the connexin-related brain transcriptome. J. Membr. Biol. 218, 39–47.10.1007/s00232-007-9049-5Search in Google Scholar PubMed

Stanimirovic, D.B. and Friedman, A. (2012). Pathophysiology of the neurovascular unit: disease cause or consequence? J Cereb Blood Flow Metab. 32, 1207–1221.Search in Google Scholar

Stehberg, J., Moraga-Amaro, R., Salazar, C., Becerra, A., Echeverría, C., Orellana, J.A., Bultynck, G., Ponsaerts, R., Leybaert, L., Simon, F., et al. (2012). Release of gliotransmitters through astroglial connexin 43 hemichannels is necessary for fear memory consolidation in the basolateral amygdala. FASEB J. 26, 3649–3657.10.1096/fj.11-198416Search in Google Scholar PubMed

Stone, J., Itin, A., Alon, T., Pe’er, J., Gnessin, H., Chan-Ling, T., and Keshet, E. (1995). Development of retinal vasculature is mediated by hypoxia-induced vascular endothelial growth factor (VEGF) expression by neuroglia. J. Neurosci. 15, 4738–4747.10.1523/JNEUROSCI.15-07-04738.1995Search in Google Scholar

Stout, C.E., Costantin, J.L., Naus, C.C., and Charles, A.C. (2002). Intercellular calcium signaling in astrocytes via ATP release through connexin hemichannels. J. Biol.Chem. 277, 10482–10488.10.1074/jbc.M109902200Search in Google Scholar PubMed

Striedinger, K., Meda, P., and Scemes, E. (2007). Exocytosis of ATP from astrocyte progenitors modulates spontaneous Ca2+ oscillations and cell migration. Glia 55, 652–662.10.1002/glia.20494Search in Google Scholar PubMed PubMed Central

Suadicani, S.O. Flores, C.E., Urban-Maldonado, M. Beelitz, M., and Scemes, E. (2004). Gap junction channels coordinate the propagation of intercellular Ca2+ signals generated by P2Y receptor activation. Glia 48, 217–229.10.1002/glia.20071Search in Google Scholar PubMed PubMed Central

Sun, J.D., Liu, Y., Yuan, Y.H., Li, J., and Chen, N.H. (2012). Gap junction dysfunction in the prefrontal cortex induces depressive-like behaviors in rats. Neuropsychopharmacology 37, 1305–1320.10.1038/npp.2011.319Search in Google Scholar PubMed PubMed Central

Suyama, S., Sunabori, T., Kanki, H., Sawamoto, K., Gachet, C., Koizumi, S., and Okano, H. (2012). Purinergic signaling promotes proliferation of adult mouse subventricular zone cells. J. Neurosci. 32, 9238–9247.10.1523/JNEUROSCI.4001-11.2012Search in Google Scholar PubMed PubMed Central

Swayne, L.A., Sorbara, C.D., and Bennett, S.A. (2010). Pannexin 2 is expressed by postnatal hippocampal neural progenitors and modulated neuronal commitment. J. Biol. Chem. 285, 24977–24986.10.1074/jbc.M110.130054Search in Google Scholar PubMed PubMed Central

Tam, S.J. and Watts, R.J. (2010). Connecting vascular and nervous system development: angiogenesis and the blood-brain barrier. Annu. Rev. Neurosci. 33, 379–408.10.1146/annurev-neuro-060909-152829Search in Google Scholar PubMed

Tam, S.J., Richmond, D.L., Kaminker, J.S., Modrusan, Z., Martin-McNulty, B., Cao, T.C., Weimer, R.M., Carano, R.A., van Bruggen, N., and Watts, R.J. (2012). Death receptors DR6 and TROY regulate brain vascular development. Dev. Cell. 22, 403–417.10.1016/j.devcel.2011.11.018Search in Google Scholar PubMed

Tanigami, H., Okamoto, T., Yasue, Y., and Shimaoka, M. (2012). Astroglial integrins in the development and regulation of neurovascular units. Pain Res. Treat. Article ID 964652, 10 pages. doi:10.1155/2012/964652.10.1155/2012/964652Search in Google Scholar PubMed PubMed Central

Tavazoie, M., Van der Veken, L., Silva-Vargas, V., Louissaint, M., Colonna, L., Zaidi, B., Garcia-Verdugo, J.M., and Doetsch, F. (2008). A specialized vascular niche for adult neural stem cells. Cell Stem Cell. 3, 279–288.10.1016/j.stem.2008.07.025Search in Google Scholar PubMed PubMed Central

Theis, M. and Giaume, C. (2012) Connexin-based intercellular communication and astrocyte heterogeneity. Brain Res. 1487, 88–98.10.1016/j.brainres.2012.06.045Search in Google Scholar PubMed

Theodoric, N., Bechberger, J.F., Naus, C.C., and Sin, W.C. (2012). Role of gap junction protein connexin43 in astrogliosis induced by brain injury. PLoS One 7, e47311.10.1371/journal.pone.0047311Search in Google Scholar PubMed PubMed Central

Todd, K.L., Kristan, W.B. Jr., and French, K.A. (2010). Gap junction expression is required for normal chemical synapse formation. J. Neurosci. 30, 15277–15285.10.1523/JNEUROSCI.2331-10.2010Search in Google Scholar PubMed PubMed Central

Toyofuku, T., Yabuki, M., Otsu, K., Kuzuya, T., Hori, M., and Tada, M. (1998). Direct association of the gap junction protein connexin-43 with ZO-1 in cardiac myocytes. J. Biol. Chem. 273, 12725–12731.10.1074/jbc.273.21.12725Search in Google Scholar PubMed

Valle-Casuso, J.C., González-Sánchez., A., Medina, J.M., and Tabernero, A. (2012). HIF-1 and c-Src mediate increased glucose uptake induced by endothelin-1 and connexin43 in astrocytes. PLoS One 7, e32448.10.1371/journal.pone.0032448Search in Google Scholar PubMed PubMed Central

Valiunas, V. (2013). Cyclic nucleotide permeability through unopposed connexin hemichannels. Front. Pharmacol. 4, 75.10.3389/fphar.2013.00075Search in Google Scholar PubMed PubMed Central

van der Heyden, M.A., Rook, M.B., Hermans, M.M., Rijksen, G., Boonstra, J., Defize, L.H., and Destrée, O.H. (1998). Identification of connexin43 as a functional target for Wnt signalling. J. Cell Sci. 111, 1741–1749.10.1242/jcs.111.12.1741Search in Google Scholar PubMed

Vangilder, R.L., Rosen, C.L., Barr, T.L., and Huber, J.D. (2011). Targeting the neurovascular unit for treatment of neurological disorders. Pharmacol. Ther. 130, 239–247.10.1016/j.pharmthera.2010.12.004Search in Google Scholar PubMed PubMed Central

Véliz, L.P., González, F.G., Duling, B.R., Sáez, J.C., and Boric, M.P. (2008). Functional role of gap junctions in cytokine-induced leukocyte adhesion to endothelium in vivo. Am. J. Physiol. Heart Circ. Physiol. 295, H1056–H1066.10.1152/ajpheart.00266.2008Search in Google Scholar PubMed PubMed Central

Walton, N.M., Shin, R., Tajinda, K., Heusner, C.L., Kogan, J.H., Miyake, S., Chen, Q., Tamura, K., and Matsumoto, M. (2012). Adult neurogenesis transiently generates oxidative stress. PLoS One 7, e35264.10.1371/journal.pone.0035264Search in Google Scholar PubMed PubMed Central

Wang, H.H., Kung, C.I., Tseng, Y.Y., Lin, Y.C., Chen, C.H., Tsai, C.H., and Yeh, H.I. (2008). Activation of endothelial cells to pathological status by down-regulation of connexin43. Cardiovasc. Res. 79, 509–518.10.1093/cvr/cvn112Search in Google Scholar PubMed

Wang, W., Osenbroch, P., Skinnes, R., Esbensen, Y., Bjoras, M., and Eide, L. (2011). Mitochondrial DNA integrity is essential for mitochondrial maturation during differentiation of neural stem cells. Stem Cell 28, 2195–2204.10.1002/stem.542Search in Google Scholar PubMed

Wang, X., Bukoreshtliev, N.V., and Gerdes, H.H. (2012). Developing neurons from transient nanotubes facilitating electrical coupling and calcium signaling with distant astrocytes. Plos One 7, e47429.Search in Google Scholar

Wang, H.H., Su, C.H., Wu, Y.J., Li, J.Y., Tseng, Y.M., Lin, Y.C., Hsieh, C.L., Tsai, C.H., and Yeh, H.I. (2013). Reduction of connexin43 in human endothelial progenitor cells impairs the angiogenic potential. Angiogenesis 16, 553–560.10.1007/s10456-013-9335-zSearch in Google Scholar PubMed

Weidenfeller, C., Svendsen, C.N., and Shusta, E.V. (2007). Differentiating embryonic neural progenitor cells induce blood-brain barrier properties. J. Neurochem. 101, 555–565.10.1111/j.1471-4159.2006.04394.xSearch in Google Scholar PubMed PubMed Central

Wicki-Stordeur, L.E., and Swayne, L.A. (2012). Large pore ion and metabolite-permeable channel regulation of postnatal ventricular zone neural stem and progenitor cells: interplay between aquaporins, connexins, and pannexins? Stem Cells Int. 2012, 454180.10.1155/2012/454180Search in Google Scholar PubMed PubMed Central

Wicki-Stordeur, L.E., Dzugalo, A.D., Swansburg, R.M., Suits, J.M., and Swayne, L.A. (2012). Pannexin 1 regulates postnatal neural stem and progenitor cell proliferation. Neural. Dev. 7, 11.10.1186/1749-8104-7-11Search in Google Scholar PubMed PubMed Central

Wiencken-Barger, A.E., Djukic, B., Casper, K.B., and McCarthy, K.D. (2007). A role for connexin43 during neurodevelopment. Glia 55, 675–686.10.1002/glia.20484Search in Google Scholar PubMed

Wilson, A.C., Clemente, L., Liu, T., Bowen, R.L., Meethal, S.V., and Atwood, C.S. (2008). Reproductive hormones regulate the selective permeability of the blood-brain barrier. Biochim. Biophys. Acta 1782, 401–407.10.1016/j.bbadis.2008.02.011Search in Google Scholar PubMed

Winkler, E.A., Bell, R.D., and Zlokovic, B.V. (2011). Central nervous system pericytes in health and disease. Nat. Neurosci. 14, 1398–1405.10.1038/nn.2946Search in Google Scholar PubMed PubMed Central

Wuestefeld, R., Chen, J., Meller, K., Brand-Saberi, B., and Theiss, C. (2012). Impact of vegf on astrocytes: analysis of gap junctional intercellular communication, proliferation, and motility. Glia. 60, 936–947.10.1002/glia.22325Search in Google Scholar PubMed

Xiong, Y., Mahmood, A., and Chopp, M. (2010). Angiogenesis, neurogenesis and brain recovery of function following injury. Curr. Opin. Investig. Drugs 11, 298–308.Search in Google Scholar

Yamashita, M. (2013). From neuroepithelial cells to neurons: changes in the physiological properties of neuroepithelial stem cells. Arch. Biochem. Biophys. 534, 64–70.10.1016/j.abb.2012.07.016Search in Google Scholar PubMed

Yue, J., Wei, W., Lam, C.M., Zhao, Y.J., Dong, M., Zhang, L.R., Zhang, L.H., and Lee, H.C. (2009). CD38/cADPR/Ca2+ pathway promotes cell proliferation and delays nerve growth factor-induced differentiation in PC12 cells. J. Biol. Chem. 284, 29335–29342.10.1074/jbc.M109.049767Search in Google Scholar PubMed PubMed Central

Received: 2013-10-18
Accepted: 2013-12-13
Published Online: 2014-01-16
Published in Print: 2014-02-01

©2014 by Walter de Gruyter Berlin Boston

Downloaded on 2.6.2024 from https://www.degruyter.com/document/doi/10.1515/revneuro-2013-0044/html
Scroll to top button