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.

  • Review Article
  • Published:

Therapeutic strategies targeting connexins

Abstract

The connexin family of channel-forming proteins is present in every tissue type in the human anatomy. Connexins are best known for forming clustered intercellular channels, structurally known as gap junctions, where they serve to exchange members of the metabolome between adjacent cells. In their single-membrane hemichannel form, connexins can act as conduits for the passage of small molecules in autocrine and paracrine signalling. Here, we review the roles of connexins in health and disease, focusing on the potential of connexins as therapeutic targets in acquired and inherited diseases as well as wound repair, while highlighting the associated clinical challenges.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Connexin topology and the connexin family.
Figure 2: Canonical life cycle of connexins.
Figure 3: Links between connexins, the connexin interactome and early-stage tumorigenesis.
Figure 4: Current and potential connexin-modulating therapeutic strategies.

Similar content being viewed by others

References

  1. Revel, J. P. & Karnovsky, M. J. Hexagonal array of subunits in intercellular junctions of the mouse heart and liver. J. Cell Biol. 33, C7–C12 (1967).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Sohl, G. & Willecke, K. Gap junctions and the connexin protein family. Cardiovasc. Res. 62, 228–232 (2004). This review describes the connexin gene family, including their structure.

    PubMed  Google Scholar 

  3. Aasen, T., Mesnil, M., Naus, C. C., Lampe, P. D. & Laird, D. W. Gap junctions and cancer: communicating for 50 years. Nat. Rev. Cancer 16, 775–788 (2016). This is a comprehensive Review that summarizes the large body of work linking connexins, growth control and cancer.

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Bruzzone, R., Hormuzdi, S. G., Barbe, M. T., Herb, A. & Monyer, H. Pannexins, a family of gap junction proteins expressed in brain. Proc. Natl Acad. Sci. USA 100, 13644–13649 (2003). Here, the authors demonstrate that pannexins have channel-forming properties.

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Esseltine, J. L. & Laird, D. W. Next-generation connexin and pannexin cell biology. Trends Cell Biol. 26, 944–955 (2016).

    CAS  PubMed  Google Scholar 

  6. Srinivas, M., Verselis, V. K. & White, T. W. Human diseases associated with connexin mutations. Biochim. Biophys. Acta 1860, 192–201 (2018). An up-to-date review of diseases linked to inherited connexin gene mutations and mechanisms associated with their disruption of tissue homeostasis.

    CAS  Google Scholar 

  7. Laird, D. W., Naus, C. C. & Lampe, P. D. SnapShot: connexins and disease. Cell 170, 1260 (2017). This 'snapshot' gives a synopsis of how inherited connexin gene mutations are linked to disease.

    CAS  PubMed  Google Scholar 

  8. Robertson, J. D. The occurrence of a subunit pattern in the unit membranes of club endings in mauthner cell synapses in goldfish brains. J. Cell Biol. 19, 201–221 (1963).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Benedetti, E. L. & Emmelot, P. Hexagonal array of subunits in tight junctions separated from isolated rat liver plasma membranes. J. Cell Biol. 38, 15–24 (1968).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Benedetti, E. L. & Emmelot, P. Electron microscopic observations on negatively stained plasma membranes isolated from rat liver. J. Cell Biol. 26, 299–305 (1965).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Sheridan, J. D. Electrophysiological evidence for low-resistance intercellular junctions in the early chick embryo. J. Cell Biol. 37, 650–659 (1968).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Kanno, Y. & Loewenstein, W. R. Cell-to-cell passage of large molecules. Nature 212, 629–630 (1966).

    CAS  PubMed  Google Scholar 

  13. Loewenstein, W. R. Permeability of membrane junctions. Ann. NY Acad. Sci. 137, 441–472 (1966).

    CAS  PubMed  Google Scholar 

  14. Paul, D. L. Molecular cloning of cDNA for rat liver gap junction protein. J. Cell Biol. 103, 123–134 (1986).

    CAS  PubMed  Google Scholar 

  15. Kumar, N. M. & Gilula, N. B. Cloning and characterization of human and rat liver cDNAs coding for a gap junction protein. J. Cell Biol. 103, 767–776 (1986).

    CAS  PubMed  Google Scholar 

  16. Beyer, E. C., Paul, D. L. & Goodenough, D. A. Connexin43: a protein from rat heart homologous to a gap junction protein from liver. J. Cell Biol. 105, 2621–2629 (1987).

    CAS  PubMed  Google Scholar 

  17. Kumar, N. M. & Gilula, N. B. Molecular biology and genetics of gap junction channels. Semin. Cell Biol. 3, 3–16 (1992).

    CAS  PubMed  Google Scholar 

  18. Sohl, G. & Willecke, K. An update on connexin genes and their nomenclature in mouse and man. Cell Commun. Adhes. 10, 173–180 (2003).

    PubMed  Google Scholar 

  19. Willecke, K., Hennemann, H., Dahl, E., Jungbluth, S. & Heynkes, R. The diversity of connexin genes encoding gap junctional proteins. Eur. J. Cell Biol. 56, 1–7 (1991).

    CAS  PubMed  Google Scholar 

  20. Angelillo-Scherrer, A. et al. Connexin 37 limits thrombus propensity by downregulating platelet reactivity. Circulation 124, 930–939 (2011).

    CAS  PubMed  Google Scholar 

  21. Vaiyapuri, S. et al. Gap junctions and connexin hemichannels underpin hemostasis and thrombosis. Circulation 125, 2479–2491 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Goodenough, D. A., Goliger, J. A. & Paul, D. L. Connexins, connexons, and intercellular communication. Annu. Rev. Biochem. 65, 475–502 (1996).

    CAS  PubMed  Google Scholar 

  23. Laird, D. W. Life cycle of connexins in health and disease. Biochem. J. 394, 527–543 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Beyer, E. C. et al. Heteromeric mixing of connexins: compatibility of partners and functional consequences. Cell Commun. Adhes. 8, 199–204 (2001).

    CAS  PubMed  Google Scholar 

  25. Koval, M. Pathways and control of connexin oligomerization. Trends Cell Biol. 16, 159–166 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Cottrell, G. T. & Burt, J. M. Functional consequences of heterogeneous gap junction channel formation and its influence in health and disease. Biochim. Biophys. Acta 1711, 126–141 (2005).

    CAS  PubMed  Google Scholar 

  27. Verselis, V., White, R. L., Spray, D. C. & Bennett, M. V. Gap junctional conductance and permeability are linearly related. Science 234, 461–464 (1986).

    CAS  PubMed  Google Scholar 

  28. Goldberg, G. S., Lampe, P. D. & Nicholson, B. J. Selective transfer of endogenous metabolites through gap junctions composed of different connexins. Nat. Cell Biol. 1, 457–459 (1999).

    CAS  PubMed  Google Scholar 

  29. White, T. W. & Bruzzone, R. Multiple connexin proteins in single intercellular channels: connexin compatibility and functional consequences. J. Bioener. Biomembr. 28, 339–350 (1996).

    CAS  Google Scholar 

  30. Maeda, S. et al. Structure of the connexin 26 gap junction channel at 3.5 Å resolution. Nature 458, 597–602 (2009). This high-resolution Cx26 structure study provided much more clarity to the arrangement of the transmembrane domains lining the connexon pore.

    CAS  PubMed  Google Scholar 

  31. Bennett, B. C. et al. An electrostatic mechanism for Ca(2+)-mediated regulation of gap junction channels. Nat. Commun. 7, 8770 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Fallon, R. F. & Goodenough, D. A. Five-hour half-life of mouse liver gap-junction protein. J. Cell Biol. 90, 521–526 (1981).

    CAS  PubMed  Google Scholar 

  33. Laird, D. W., Puranam, K. L. & Revel, J. P. Turnover and phosphorylation dynamics of connexin43 gap junction protein in cultured cardiac myocytes. Biochem. J. 273, 67–72 (1991).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Beardslee, M. A., Laing, J. G., Beyer, E. C. & Saffitz, J. E. Rapid turnover of connexin43 in the adult rat heart. Circ. Res. 83, 629–635 (1998).

    CAS  PubMed  Google Scholar 

  35. Jiang, J. X., Paul, D. L. & Goodenough, D. A. Posttranslational phosphorylation of lens fiber connexin46: a slow occurrence. Invest. Ophthalmol. Vis. Sci. 34, 3558–3565 (1993).

    CAS  PubMed  Google Scholar 

  36. Kelly, J. J., Shao, Q., Jagger, D. J. & Laird, D. W. Cx30 exhibits unique characteristics including a long half-life when assembled into gap junctions. J. Cell Sci. 128, 3947–3960 (2015).

    CAS  PubMed  Google Scholar 

  37. Leybaert, L. et al. Connexins in cardiovascular and neurovascular health and disease: pharmacological implications. Pharmacol. Rev. 69, 396–478 (2017). This comprehensive review from many leaders in the gap junction field discusses the role of connexins in disease and how they are potential drug targets.

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Delmar, M. et al. Connexins and disease. Cold Spring Harb. Perspect Biol. 10, a029348 (2017).

    Google Scholar 

  39. Laird, D. W. Closing the gap on autosomal dominant connexin-26 and connexin-43 mutants linked to human disease. J. Biol. Chem. 283, 2997–3001 (2008).

    CAS  PubMed  Google Scholar 

  40. Koval, M., Molina, S. A. & Burt, J. M. Mix and match: investigating heteromeric and heterotypic gap junction channels in model systems and native tissues. FEBS Lett. 588, 1193–1204 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Musil, L. S. & Goodenough, D. A. Multisubunit assembly of an integral plasma membrane channel protein, gap junction connexin43, occurs after exit from the ER. Cell 74, 1065–1077 (1993).

    CAS  PubMed  Google Scholar 

  42. Shaw, R. M. et al. Microtubule plus-end-tracking proteins target gap junctions directly from the cell interior to adherens junctions. Cell 128, 547–560 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Johnson, R. G. et al. Gap junctions assemble in the presence of cytoskeletal inhibitors, but enhanced assembly requires microtubules. Exp. Cell Res. 275, 67–80 (2002).

    CAS  PubMed  Google Scholar 

  44. Giepmans, B. N. et al. Gap junction protein connexin-43 interacts directly with microtubules. Curr. Biol. 11, 1364–1368 (2001).

    CAS  PubMed  Google Scholar 

  45. Li, H. et al. Properties and regulation of gap junctional hemichannels in the plasma membranes of cultured cells. J. Cell Biol. 134, 1019–1030 (1996).

    CAS  PubMed  Google Scholar 

  46. Lopez, W. et al. Mechanism of gating by calcium in connexin hemichannels. Proc. Natl Acad. Sci. USA 113, E7986–E7995 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Rhett, J. M. & Gourdie, R. G. The perinexus: a new feature of Cx43 gap junction organization. Heart Rhythm 9, 619–623 (2011).

    PubMed  PubMed Central  Google Scholar 

  48. Vermij, S. H., Abriel, H. & van Veen, T. A. Refining the molecular organization of the cardiac intercalated disc. Cardiovasc. Res. 113, 259–275 (2017).

    CAS  PubMed  Google Scholar 

  49. Willebrords, J. et al. Connexins and their channels in inflammation. Crit. Rev. Biochem. Mol. Biol. 51, 413–439 (2016). This review focuses on the relationships between connexins and inflammation.

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Gaietta, G. et al. Multicolor and electron microscopic imaging of connexin trafficking. Science 296, 503–507 (2002).

    CAS  PubMed  Google Scholar 

  51. Lauf, U. et al. Dynamic trafficking and delivery of connexons to the plasma membrane and accretion to gap junctions in living cells. Proc. Natl Acad. Sci. USA 99, 10446–10451 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Amsterdam, A., Josephs, R., Lieberman, M. E. & Lindner, H. R. Organization of intramembrane particles in freeze-cleaved gap junctions of rat graafian rollicles: optical-diffraction analysis. J. Cell Sci. 21, 93–105 (1976).

    CAS  PubMed  Google Scholar 

  53. Jordan, K., Chodock, R., Hand, A. R. & Laird, D. W. The origin of annular junctions: a mechanism of gap junction internalization. J. Cell Sci. 114, 763–773 (2001).

    CAS  PubMed  Google Scholar 

  54. Falk, M. M. et al. Degradation of endocytosed gap junctions by autophagosomal and endo-/lysosomal pathways: a perspective. J. Membr. Biol. 245, 465–476 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Norris, R. P., Baena, V. & Terasaki, M. Localization of phosphorylated connexin 43 using serial section immunogold electron microscopy. J. Cell Sci. 130, 1333–1340 (2017).

    CAS  PubMed  Google Scholar 

  56. Qin, H., Shao, Q., Igdoura, S. A., Alaoui-Jamali, M. A. & Laird, D. W. Lysosomal and proteasomal degradation play distinct roles in the life cycle of Cx43 in gap junctional intercellular communication-deficient and -competent breast tumor cells. J. Biol. Chem. 278, 30005–30014 (2003).

    CAS  PubMed  Google Scholar 

  57. Leithe, E., Sirnes, S., Fykerud, T., Kjenseth, A. & Rivedal, E. Endocytosis and post-endocytic sorting of connexins. Biochim. Biophys. Acta 1818, 1870–1879 (2011).

    PubMed  Google Scholar 

  58. McLachlin, J. R., Caveney, S. & Kidder, G. M. Control of gap junction formation in early mouse embryos. Dev. Biol. 98, 155–164 (1983).

    CAS  PubMed  Google Scholar 

  59. Davies, T. C., Barr, K. J., Jones, D. H., Zhu, D. & Kidder, G. M. Multiple members of the connexin gene family participate in preimplantation development of the mouse. Dev. Genet. 18, 234–243 (1996).

    CAS  PubMed  Google Scholar 

  60. Boulay, A. C. et al. Hearing is normal without connexin30. J. Neurosci. 33, 430–434 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Kruger, O. et al. Defective vascular development in connexin 45-deficient mice. Development 127, 4179–4193 (2000).

    CAS  PubMed  Google Scholar 

  62. Delorme, B. et al. Developmental regulation of connexin 40 gene expression in mouse heart correlates with the differentiation of the conduction system. Dev. Dyn. 204, 358–371 (1995).

    CAS  PubMed  Google Scholar 

  63. Delmar, M. & Makita, N. Cardiac connexins, mutations and arrhythmias. Curr. Opin. Cardiol. 27, 236–241 (2012).

    PubMed  Google Scholar 

  64. Merrifield, P. A. & Laird, D. W. Connexins in skeletal muscle development and disease. Semin. Cell Dev. Biol. 50, 67–73 (2016).

    CAS  PubMed  Google Scholar 

  65. Rozental, R., Giaume, C. & Spray, D. C. Gap junctions in the nervous system. Brain Res. Rev. 32, 11–15 (2000).

    CAS  PubMed  Google Scholar 

  66. Di, W. L., Rugg, E. L., Leigh, I. M. & Kelsell, D. P. Multiple epidermal connexins are expressed in different keratinocyte subpopulations including connexin 31. J. Invest. Dermatol. 117, 958–964 (2001).

    CAS  PubMed  Google Scholar 

  67. Hennemann, H., Kozjek, G., Dahl, E., Nicholson, B. & Willecke, K. Molecular cloning of mouse connexins26 and -32: similar genomic organization but distinct promoter sequences of two gap junction genes. Eur. J. Cell Biol. 58, 81–89 (1992).

    CAS  PubMed  Google Scholar 

  68. Fladmark, K. E. et al. Gap junctions and growth control in liver regeneration and in isolated rat hepatocytes. Hepatology 25, 847–855 (1997).

    CAS  PubMed  Google Scholar 

  69. Stains, J. P. & Civitelli, R. Gap junctions in skeletal development and function. Biochim. Biophys. Acta 1719, 69–81 (2005).

    CAS  PubMed  Google Scholar 

  70. Chow, L. & Lye, S. J. Expression of the gap junction protein connexin-43 is increased in the human myometrium toward term and with the onset of labor. Am. J. Obstet. Gynecol. 170, 788–795 (1994).

    CAS  PubMed  Google Scholar 

  71. Lefebvre, D. L., Piersanti, M., Bai, X. H., Chen, Z. Q. & Lye, S. J. Myometrial transcriptional regulation of the gap junction gene, connexin-43. Reprod. Fertil. Dev. 7, 603–611 (1995).

    CAS  PubMed  Google Scholar 

  72. Kilarski, W. M., Rezapour, M., Backstrom, T., Roomans, G. M. & Ulmsten, U. Morphometric analysis of gap junction density in human myometrium at term. Acta Obstet. Gynecol. Scand. 73, 377–384 (1994).

    CAS  PubMed  Google Scholar 

  73. Naus, C. C. & Laird, D. W. Implications and challenges of connexin connections to cancer. Nat. Rev. Cancer 10, 435–441 (2010).

    CAS  PubMed  Google Scholar 

  74. Becker, D. L., Phillips, A. R., Duft, B. J., Kim, Y. & Green, C. R. Translating connexin biology into therapeutics. Semin. Cell Dev. Biol. 50, 49–58 (2016). This review, from the research teams that modulated Cx43 levels using antisense technologies and are currently exploiting hemichannel directed drugs, describes how these approaches can be utilized for therapeutic purposes.

    CAS  PubMed  Google Scholar 

  75. Ghatnekar, G. S. et al. Connexin43 carboxyl-terminal peptides reduce scar progenitor and promote regenerative healing following skin wounding. Regen. Med. 4, 205–223 (2009).

    CAS  PubMed  Google Scholar 

  76. Gourdie, R. G. et al. The unstoppable connexin43 carboxyl-terminus: new roles in gap junction organization and wound healing. Ann. NY Acad. Sci. 1080, 49–62 (2006).

    CAS  PubMed  Google Scholar 

  77. Scheckenbach, K. E., Crespin, S., Kwak, B. R. & Chanson, M. Connexin channel-dependent signaling pathways in inflammation. J. Vasc. Res. 48, 91–103 (2011).

    CAS  PubMed  Google Scholar 

  78. Loewenstein, W. R. & Kanno, Y. Intercellular communication and the control of tissue growth: lack of communication between cancer cells. Nature 209, 1248–1249 (1966).

    CAS  PubMed  Google Scholar 

  79. Loewenstein, W. R. & Kanno, Y. Intercellular communication and tissue growth. I. Cancerous growth. J. Cell Biol. 33, 225–234 (1967).

    CAS  PubMed  PubMed Central  Google Scholar 

  80. Yotti, L. P., Chang, C. C. & Trosko, J. E. Elimination of metabolic cooperation in Chinese hamster cells by a tumor promoter. Science 206, 1089–1091 (1979).

    CAS  PubMed  Google Scholar 

  81. Zhu, D., Caveney, S., Kidder, G. M. & Naus, C. C. Transfection of C6 glioma cells with connexin 43 cDNA: analysis of expression, intercellular coupling, and cell proliferation. Proc. Natl Acad. Sci. USA 88, 1883–1887 (1991).

    CAS  PubMed  PubMed Central  Google Scholar 

  82. Naus, C. C., Zhu, D., Todd, S. D. & Kidder, G. M. Characteristics of C6 glioma cells overexpressing a gap junction protein. Cell. Mol. Neruobiol. 12, 163–175 (1992).

    CAS  Google Scholar 

  83. Zhu, D., Kidder, G. M., Caveney, S. & Naus, C. C. Growth retardation in glioma cells cocultured with cells overexpressing a gap junction protein. Proc. Natl Acad. Sci. USA 89, 10218–10221 (1992).

    CAS  PubMed  PubMed Central  Google Scholar 

  84. McLachlan, E., Shao, Q., Wang, H. L., Langlois, S. & Laird, D. W. Connexins act as tumor suppressors in three-dimensional mammary cell organoids by regulating differentiation and angiogenesis. Cancer Res. 66, 9886–9894 (2006).

    CAS  PubMed  Google Scholar 

  85. Temme, A. et al. High incidence of spontaneous and chemically induced liver tumors in mice deficient for connexin32. Curr. Biol. 7, 713–716 (1997).

    CAS  PubMed  Google Scholar 

  86. Evert, M., Ott, T., Temme, A., Willecke, K. & Dombrowski, F. Morphology and morphometric investigation of hepatocellular preneoplastic lesions and neoplasms in connexin32-deficient mice. Carcinogenesis 23, 697–703 (2002).

    CAS  PubMed  Google Scholar 

  87. Fukumasu, H. et al. Higher susceptibility of spontaneous and NNK-induced lung neoplasms in connexin 43 deficient CD1 × AJ F1 mice: paradoxical expression of connexin 43 during lung carcinogenesis. Mol. Carcinog. 52, 497–506 (2013).

    CAS  PubMed  Google Scholar 

  88. Stewart, M. K., Bechberger, J. F., Welch, I., Naus, C. C. & Laird, D. W. Cx26 knockout predisposes the mammary gland to primary mammary tumors in a DMBA-induced mouse model of breast cancer. Oncotarget 6, 37185–37199 (2015).

    PubMed  PubMed Central  Google Scholar 

  89. Mesnil, M. et al. Negative growth control of HeLa cells by connexin genes: connexin species specificity. Cancer Res. 55, 629–639 (1995).

    CAS  PubMed  Google Scholar 

  90. Krutovskikh, V. A. et al. Differential effect of subcellular localization of communication impairing gap junction protein connexin43 on tumor cell growth in vivo. Oncogene 19, 505–513 (2000).

    CAS  PubMed  Google Scholar 

  91. Yang, J. et al. Reciprocal positive regulation between Cx26 and PI3K/Akt pathway confers acquired gefitinib resistance in NSCLC cells via GJIC-independent induction of EMT. Cell Death Dis. 6, e1829 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Laird, D. W. The gap junction proteome and its relationship to disease. Trends Cell Biol. 20, 92–101 (2010).

    CAS  PubMed  Google Scholar 

  93. Ruiz-Meana, M. et al. Mitochondrial connexin43 as a new player in the pathophysiology of myocardial ischaemia-reperfusion injury. Cardiovasc. Res. 77, 325–333 (2008).

    CAS  PubMed  Google Scholar 

  94. Sun, Y. et al. Connexin 43 interacts with Bax to regulate apoptosis of pancreatic cancer through a gap junction-independent pathway. Int. J. Oncol. 41, 941–948 (2012).

    CAS  PubMed  Google Scholar 

  95. Boengler, K. & Schulz, R. Connexin 43 and mitochondria in cardiovascular health and disease. Adv. Exp. Med. Biol. 982, 227–246 (2017).

    CAS  PubMed  Google Scholar 

  96. Dang, X., Doble, B. W. & Kardami, E. The carboxy-tail of connexin-43 localizes to the nucleus and inhibits cell growth. Mol. Cell Biochem. 242, 35–38 (2003).

    CAS  PubMed  Google Scholar 

  97. Saez, J. C. & Leybaert, L. Hunting for connexin hemichannels. FEBS Lett. 588, 1205–1211 (2014).

    CAS  PubMed  Google Scholar 

  98. Ito, A. et al. A role for heterologous gap junctions between melanoma and endothelial cells in metastasis. J. Clin. Invest. 105, 1189–1197 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  99. Plante, I., Stewart, M. K., Barr, K., Allan, A. L. & Laird, D. W. Cx43 suppresses mammary tumor metastasis to the lung in a Cx43 mutant mouse model of human disease. Oncogene 30, 1681–1692 (2011).

    CAS  PubMed  Google Scholar 

  100. Chen, Q. et al. Carcinoma-astrocyte gap junctions promote brain metastasis by cGAMP transfer. Nature 533, 493–498 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  101. Alonso, F. et al. Targeting endothelial connexin40 inhibits tumor growth by reducing angiogenesis and improving vessel perfusion. Oncotarget 7, 14015–14028 (2016).

    PubMed  PubMed Central  Google Scholar 

  102. Chan, D. K. & Chang, K. W. GJB2-associated hearing loss: systematic review of worldwide prevalence, genotype, and auditory phenotype. Laryngoscope 124, E34–E53 (2014).

    PubMed  Google Scholar 

  103. Petit, C., Levilliers, J. & Hardelin, J. P. Molecular genetics of hearing loss. Annu. Rev. Genet. 35, 589–646 (2001).

    CAS  PubMed  Google Scholar 

  104. Jamieson, S., Going, J. J., D'Arcy, R. & George, W. D. Expression of gap junction proteins connexin 26 and connexin 43 in normal human breast and in breast tumours. J. Pathol. 184, 37–43 (1998).

    CAS  PubMed  Google Scholar 

  105. Marquez-Rosado, L., Singh, D., Rincon-Arano, H., Solan, J. L. & Lampe, P. D. CASK (LIN2) interacts with Cx43 in wounded skin and their coexpression affects cell migration. J. Cell Sci. 125, 695–702 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  106. Dunn, C. A. & Lampe, P. D. Injury-triggered Akt phosphorylation of Cx43: a ZO-1-driven molecular switch that regulates gap junction size. J. Cell Sci. 127, 455–464 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Richards, T. S. et al. Protein kinase C spatially and temporally regulates gap junctional communication during human wound repair via phosphorylation of connexin43 on serine368. J. Cell Biol. 167, 555–562 (2004). This is a study of the Cx43 protein and phosphorylation level changes related to the temporal and spatial need for gap junction communication during wound healing.

    CAS  PubMed  PubMed Central  Google Scholar 

  108. Goliger, J. A. & Paul, D. L. Wounding alters epidermal connexin expression and gap junction-mediated intercellular communication. Mol. Biol. Cell 6, 1491–1501 (1995). This study shows how connexin (Cx26, Cx31.1 and Cx43) expression dramatically changes in wounded skin.

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Lampe, P. D. et al. Cellular interaction of integrin alpha3beta1 with laminin 5 promotes gap junctional communication. J. Cell Biol. 143, 1735–1747 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Saitoh, M., Oyamada, M., Oyamada, Y., Kaku, T. & Mori, M. Changes in the expression of gap junction proteins (connexins) in hamster tongue epithelium during wound healing and carcinogenesis. Carcinogenesis 18, 1319–1328 (1997).

    CAS  PubMed  Google Scholar 

  111. Coutinho, P., Qiu, C., Frank, S., Tamber, K. & Becker, D. Dynamic changes in connexin expression correlate with key events in the wound healing process. Cell Biol. Int. 27, 525–541 (2003).

    CAS  PubMed  Google Scholar 

  112. King, T. J. et al. Deficiency in the gap junction protein connexin32 alters p27Kip1 tumor suppression and MAPK activation in a tissue-specific manner. Oncogene 24, 1718–1726 (2005).

    CAS  PubMed  Google Scholar 

  113. Kretz, M. et al. Altered connexin expression and wound healing in the epidermis of connexin-deficient mice. J. Cell Sci. 116, 3443–3452 (2003).

    CAS  PubMed  Google Scholar 

  114. Qiu, C. et al. Targeting connexin43 expression accelerates the rate of wound repair. Curr. Biol. 13, 1697–1703 (2003).

    CAS  PubMed  Google Scholar 

  115. Wang, C. M., Lincoln, J., Cook, J. E. & Becker, D. L. Abnormal connexin expression underlies delayed wound healing in diabetic skin. Diabetes 56, 2809–2817 (2007).

    CAS  PubMed  Google Scholar 

  116. Davidson, J. O., Green, C. R., Nicholson, L. F., Bennet, L. & Gunn, A. J. Deleterious effects of high dose connexin 43 mimetic peptide infusion after cerebral ischaemia in near-term fetal sheep. Int. J. Mol. Sci. 13, 6303–6319 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  117. Nakano, Y. et al. Connexin43 knockdown accelerates wound healing but inhibits mesenchymal transition after corneal endothelial injury in vivo. Invest. Ophthalmol. Vis. Sci. 49, 93–104 (2008).

    PubMed  Google Scholar 

  118. Grupcheva, C. N. et al. Improved corneal wound healing through modulation of gap junction communication using connexin43-specific antisense oligodeoxynucleotides. Invest. Opthamol. Vis. Sci. 53, 1130–1138 (2012). This is a key study showing that Cx43 antisense treatment to a scrape corneal wound rat model causes a significant reduction in wound area.

    CAS  Google Scholar 

  119. Moore, K. et al. A synthetic connexin 43 mimetic peptide augments corneal wound healing. Exp. Eye Res. 115, 178–188 (2013). This study effectively shows that microencapsulated aCT1 significantly improves rat corneal surgical wound closure.

    CAS  PubMed  PubMed Central  Google Scholar 

  120. Moore, K., Ghatnekar, G., Gourdie, R. G. & Potts, J. D. Impact of the controlled release of a connexin 43 peptide on corneal wound closure in an STZ model of type I diabetes. PLOS ONE 9, e86570 (2014).

    PubMed  PubMed Central  Google Scholar 

  121. Spray, D. C. & Burt, J. M. Structure-activity relations of the cardiac gap junction channel. Am. J. Physiol. 258, C195–C205 (1990).

    CAS  PubMed  Google Scholar 

  122. Jalife, J., Morley, G. E. & Vaidya, D. Connexins and impulse propagation in the mouse heart. J. Cardiovasc. Electrophysiol. 10, 1649–1663 (1999).

    CAS  PubMed  Google Scholar 

  123. Kirchhoff, S. et al. Reduced cardiac conduction velocity and predisposition to arrhythmias in connexin40-deficient mice. Curr. Biol. 8, 299–302 (1998).

    CAS  PubMed  Google Scholar 

  124. Gutstein, D. E. et al. Conduction slowing and sudden arrhythmic death in mice with cardiac-restricted inactivation of connexin43. Circ. Res. 88, 333–339 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  125. Simon, A. M., Goodenough, D. A. & Paul, D. L. Mice lacking connexin40 have cardiac conduction abnormalities characteristic of atrioventricular block and bundle branch block. Curr. Biol. 8, 295–298 (1998).

    CAS  PubMed  Google Scholar 

  126. Severs, N. J., Bruce, A. F., Dupont, E. & Rothery, S. Remodelling of gap junctions and connexin expression in diseased myocardium. Cardiovasc. Res. 80, 9–19 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  127. Guerrero, P. A. et al. Slow ventricular conduction in mice heterozygous for a connexin43 null mutation. J. Clin. Invest. 99, 1991–1998 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  128. Morley, G. E. et al. Characterization of conduction in the ventricles of normal and heterozygous Cx43 knockout mice using optical mapping. J. Cardiovasc. Electrophysiol. 10, 1361–1375 (1999).

    CAS  PubMed  Google Scholar 

  129. Hesketh, G. G. et al. Ultrastructure and regulation of lateralized connexin43 in the failing heart. Circ. Res. 106, 1153–1163 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  130. Hichri, E., Abriel, H. & Kucera, J. P. Distribution of cardiac sodium channels in clusters potentiates ephaptic interactions in the intercalated disc. J. Physiol. 596, 563–589 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  131. Veeraraghavan, R. et al. Sodium channels in the Cx43 gap junction perinexus may constitute a cardiac ephapse: an experimental and modeling study. Pflugers Arch. 467, 2093–2105 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  132. Raisch, T. B. et al. Intercalated disk extracellular nanodomain expansion in patients with atrial fibrillation. Front. Physiol. 9, 398 (2018).

    PubMed  PubMed Central  Google Scholar 

  133. Beardslee, M. A. et al. Dephosphorylation and intracellular redistribution of ventricular connexin43 during electrical uncoupling induced by ischemia. Circ. Res. 87, 656–662 (2000).

    CAS  PubMed  Google Scholar 

  134. Murry, C. E., Jennings, R. B. & Reimer, K. A. Preconditioning with ischemia: a delay of lethal cell injury in ischemic myocardium. Circulation 74, 1124–1136 (1986).

    CAS  PubMed  Google Scholar 

  135. Jain, S. K., Schuessler, R. B. & Saffitz, J. E. Mechanisms of delayed electrical uncoupling induced by ischemic preconditioning. Circ. Res. 92, 1138–1144 (2003).

    CAS  PubMed  Google Scholar 

  136. Schulz, R. & Heusch, G. Connexin 43 and ischemic preconditioning. Cardiovasc. Res. 62, 335–344 (2004).

    CAS  PubMed  Google Scholar 

  137. Martins-Marques, T., Anjo, S. I., Pereira, P., Manadas, B. & Girao, H. Interacting network of the gap junction (GJ) protein connexin43 (Cx43) is modulated by ischemia and reperfusion in the heart. Mol. Cell. Proteomics 14, 3040–3055 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  138. Flenniken, A. M. et al. A Gja1 missense mutation in a mouse model of oculodentodigital dysplasia. Development 132, 4375–4386 (2005).

    CAS  PubMed  Google Scholar 

  139. Zhao, H. B. Hypothesis of K(+)-recycling defect is not a primary deafness mechanism for Cx26 (GJB2) deficiency. Front. Mol. Neurosci. 10, 162 (2017).

    PubMed  PubMed Central  Google Scholar 

  140. Verselis, V. K. Connexin hemichannels and cochlear function. Neurosci. Lett. https://doi.org/10.1016/j.neulet.2017.09.020 (2017).

    CAS  PubMed  Google Scholar 

  141. Mittal, R. et al. Signaling in the auditory system: implications in hair cell regeneration and hearing function. J. Cell. Physiol. 232, 2710–2721 (2017).

    CAS  PubMed  Google Scholar 

  142. Rubinos, C., Villone, K., Mhaske, P. V., White, T. W. & Srinivas, M. Functional effects of Cx50 mutations associated with congenital cataracts. Am. J. Physiol. Cell Physiol. 306, C212–C220 (2014).

    CAS  PubMed  Google Scholar 

  143. Pal, J. D. et al. Connexin46 mutations linked to congenital cataract show loss of gap junction channel function. Am. J. Physiol. Cell Physiol. 279, C596–C602 (2000).

    CAS  PubMed  Google Scholar 

  144. Kannabiran, C. & Balasubramanian, D. Molecular genetics of cataract. Indian J. Ophthalmol. 48, 5–13 (2000).

    CAS  PubMed  Google Scholar 

  145. Bergoffen, J. et al. Connexin mutations in X-linked Charcot-Marie-Tooth disease. Science 262, 2039–2042 (1993).

    CAS  PubMed  Google Scholar 

  146. Ionasescu, V., Searby, C. & Ionasescu, R. Point mutations of the connexin32 (GJB1) gene in X-linked dominant Charcot-Marie-Tooth neuropathy. Hum. Mol. Genet. 3, 355–358 (1994).

    CAS  PubMed  Google Scholar 

  147. Orthmann-Murphy, J. L. et al. Hereditary spastic paraplegia is a novel phenotype for GJA12/GJC2 mutations. Brain 132, 426–438 (2009).

    PubMed  Google Scholar 

  148. Kim, M. S., Gloor, G. B. & Bai, D. The distribution and functional properties of Pelizaeus-Merzbacher-like disease-linked Cx47 mutations on Cx47/Cx47 homotypic and Cx47/Cx43 heterotypic gap junctions. Biochem. J. 452, 249–258 (2013).

    CAS  PubMed  Google Scholar 

  149. Ressot, C. & Bruzzone, R. Connexin channels in Schwann cells and the development of the X-linked form of Charcot-Marie-Tooth disease. Brain Res. Rev. 32, 192–202 (2000).

    CAS  PubMed  Google Scholar 

  150. Lilly, E., Sellitto, C., Milstone, L. M. & White, T. W. Connexin channels in congenital skin disorders. Semin. Cell Dev. Biol. 50, 4–12 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  151. Lee, J. R. & White, T. W. Connexin-26 mutations in deafness and skin disease. Expert Rev. Mol. Med. 11, e35 (2009).

    PubMed  Google Scholar 

  152. Churko, J. M. & Laird, D. W. Gap junction remodeling in skin repair following wounding and disease. Physiology 28, 190–198 (2013).

    CAS  PubMed  Google Scholar 

  153. Mese, G. et al. The Cx26-G45E mutation displays increased hemichannel activity in a mouse model of the lethal form of keratitis-ichthyosis-deafness syndrome. Mol. Biol. Cell 22, 4776–4786 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  154. Paznekas, W. A. et al. GJA1 mutations, variants, and connexin 43 dysfunction as it relates to the oculodentodigital dysplasia phenotype. Hum. Mutat. 30, 724–733 (2009).

    CAS  PubMed  Google Scholar 

  155. Paznekas, W. A. et al. Connexin 43 (GJA1) mutations cause the pleiotropic phenotype of oculodentodigital dysplasia. Am. J. Hum. Genet. 72, 408–418 (2003). This report describes how several mutations in the gene encoding Cx43 are causal of oculodentodigital dysplasia.

    CAS  PubMed  Google Scholar 

  156. Reaume, A. G. et al. Cardiac malformation in neonatal mice lacking connexin43. Science 267, 1831–1834 (1995).

    CAS  PubMed  Google Scholar 

  157. Roberts, J. D. et al. Targeted deep sequencing reveals no evidence for somatic mosaicism in atrial fibrillation. Circ. Cardiovasc. Genet. 8, 50–57 (2015).

    CAS  PubMed  Google Scholar 

  158. Gollob, M. H. et al. Somatic mutations in the connexin 40 gene (GJA5) in atrial fibrillation. N. Engl. J. Med. 354, 2677–2688 (2006).

    CAS  PubMed  Google Scholar 

  159. Kelly, J. J., Simek, J. & Laird, D. W. Mechanisms linking connexin mutations to human diseases. Cell Tissue Res. 360, 701–721 (2015).

    CAS  PubMed  Google Scholar 

  160. Hagendorff, A., Schumacher, B., Kirchhoff, S., Luderitz, B. & Willecke, K. Conduction disturbances and increased atrial vulnerability in connexin40-deficient mice analyzed by transesophageal stimulation. Circulation 99, 1508–1515 (1999).

    CAS  PubMed  Google Scholar 

  161. Chan, D. K., Schrijver, I. & Chang, K. W. Connexin-26-associated deafness: phenotypic variability and progression of hearing loss. Genet. Med. 12, 174–181 (2010).

    CAS  PubMed  Google Scholar 

  162. Martinez, A. D., Acuna, R., Figueroa, V., Maripillan, J. & Nicholson, B. Gap-junction channels dysfunction in deafness and hearing loss. Antioxid. Redox Signal. 11, 309–322 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  163. Tekin, M., Arnos, K. S. & Pandya, A. Advances in hereditary deafness. Lancet 358, 1082–1090 (2001).

    CAS  PubMed  Google Scholar 

  164. Haapaniemi, E., Botla, S., Persson, J., Schmierer, B. & Taipale, J. CRISPR-Cas9 genome editing induces a p53-mediated DNA damage response. Nat. Med. 24, 927–930 (2018).

    CAS  PubMed  Google Scholar 

  165. Riquelme, M. A., Kar, R., Gu, S. & Jiang, J. X. Antibodies targeting extracellular domain of connexins for studies of hemichannels. Neuropharmacology 75, 525–532 (2013).

    CAS  PubMed  Google Scholar 

  166. Zhou, J. Z. et al. Osteocytic connexin hemichannels suppress breast cancer growth and bone metastasis. Oncogene 35, 5597–5607 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  167. Coutinho, P. et al. Limiting burn extension by transient inhibition of connexin43 expression at the site of injury. Br. J. Plast. Surg. 58, 658–667 (2005).

    CAS  PubMed  Google Scholar 

  168. Cronin, M., Anderson, P. N., Cook, J. E., Green, C. R. & Becker, D. L. Blocking connexin43 expression reduces inflammation and improves functional recovery after spinal cord injury. Mol. Cell Neurosci. 39, 152–160 (2008). This report reveals that Cx43 antisense application in two models of rat spinal cord injury reduces swelling, tissue disruption, astrocytic GFAP upregulation and neutrophil extravasation.

    CAS  PubMed  Google Scholar 

  169. O'Quinn, M. P., Palatinus, J. A., Harris, B. S., Hewett, K. W. & Gourdie, R. G. A peptide mimetic of the connexin43 carboxyl terminus reduces gap junction remodeling and induced arrhythmia following ventricular injury. Circ. Res. 108, 704–715 (2011). Here, aCT1 is shown to significantly reduce lateralization of Cx43 in a mouse model of heart cryoinjury.

    CAS  PubMed  PubMed Central  Google Scholar 

  170. Palatinus, J. A., Rhett, J. M. & Gourdie, R. G. The connexin43 carboxyl terminus and cardiac gap junction organization. Biochim. Biophys. Acta 1818, 1831–1843 (2012).

    CAS  PubMed  Google Scholar 

  171. Giepmans, B. N. & Moolenaar, W. H. The gap junction protein connexin43 interacts with the second PDZ domain of the zona occludens-1 protein. Curr. Biol. 8, 931–934 (1998).

    CAS  PubMed  Google Scholar 

  172. Toyofuku, T. et al. Direct association of the gap junction protein connexin-43 with ZO-1 in cardiac myocytes. J. Biol. Chem. 273, 12725–12731 (1998).

    CAS  PubMed  Google Scholar 

  173. Grek, C. L. et al. Topical administration of a connexin43-based peptide augments healing of chronic neuropathic diabetic foot ulcers: a multicenter, randomized trial. Wound Repair Regen. 23, 203–212 (2015).

    PubMed  PubMed Central  Google Scholar 

  174. Ghatnekar, G. S., Grek, C. L., Armstrong, D. G., Desai, S. C. & Gourdie, R. G. The effect of a connexin43-based peptide on the healing of chronic venous leg ulcers: a multicenter, randomized trial. J. Invest. Dermatol. 135, 289–298 (2015). This phase II clinical trial shows that aCT1 treatment causes a significantly greater reduction in the per cent ulcer area and a doubling of the incidence of complete wound closure.

    CAS  PubMed  Google Scholar 

  175. Montgomery, J., Ghatnekar, G. S., Grek, C. L., Moyer, K. E. & Gourdie, R. G. Connexin 43-based therapeutics for dermal wound healing. Int. J. Mol. Sci. 19, E1778 (2018).

    PubMed  Google Scholar 

  176. Grek, C. L. et al. A multicenter randomized controlled trial evaluating a Cx43-mimetic peptide in cutaneous scarring. J. Invest. Dermatol. 137, 620–630 (2017). This report describes the results from a phase II clinical trial on the effect of aCT1 after laparoscopic surgery and shows improvements in scar pigmentation, thickness, surface roughness and mechanical suppleness.

    CAS  PubMed  Google Scholar 

  177. Hunter, A. W., Barker, R. J., Zhu, C. & Gourdie, R. G. Zonula occludens-1 alters connexin43 gap junction size and organization by influencing channel accretion. Mol. Biol. Cell 16, 5686–5698 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  178. Rhett, J. M., Jourdan, J. & Gourdie, R. G. Connexin 43 connexon to gap junction transition is regulated by zonula occludens-1. Mol. Biol. Cell 22, 1516–1528 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  179. Mendoza-Naranjo, A. et al. Overexpression of the gap junction protein Cx43 as found in diabetic foot ulcers can retard fibroblast migration. Cell Biol. Int. 36, 661–667 (2012).

    CAS  PubMed  Google Scholar 

  180. Palatinus, J. A. & Gourdie, R. G. Diabetes increases cryoinjury size with associated effects on Cx43 gap junction function and phosphorylation in the mouse heart. J. Diabetes Res. 2016, 8789617 (2016).

    PubMed  Google Scholar 

  181. Wong, P. et al. The role of connexins in wound healing and repair: novel therapeutic approaches. Front. Physiol. 7, 596 (2016).

    PubMed  PubMed Central  Google Scholar 

  182. Deva, N. C., Zhang, J., Green, C. R. & Danesh-Meyer, H. V. Connexin43 modulation inhibits scarring in a rabbit eye glaucoma trabeculectomy model. Inflammation 35, 1276–1286 (2012).

    CAS  PubMed  Google Scholar 

  183. Ormonde, S. et al. Regulation of connexin43 gap junction protein triggers vascular recovery and healing in human ocular persistent epithelial defect wounds. J. Membr. Biol. 245, 381–388 (2012).

    CAS  PubMed  Google Scholar 

  184. Goadsby, P. J. Emerging therapies for migraine. Nat. Clin. Pract. Neurol. 3, 610–619 (2007).

    CAS  PubMed  Google Scholar 

  185. Damodaram, S., Thalakoti, S., Freeman, S. E., Garrett, F. G. & Durham, P. L. Tonabersat inhibits trigeminal ganglion neuronal-satellite glial cell signaling. Headache 49, 5–20 (2009).

    PubMed  PubMed Central  Google Scholar 

  186. Dahlof, C. G., Hauge, A. W. & Olesen, J. Efficacy and safety of tonabersat, a gap-junction modulator, in the acute treatment of migraine: a double-blind, parallel-group, randomized study. Cephalalgia 29, 7–16 (2009).

    PubMed  Google Scholar 

  187. Silberstein, S. D. et al. Tonabersat, a gap-junction modulator: efficacy and safety in two randomized, placebo-controlled, dose-ranging studies of acute migraine. Cephalalgia 29, 17–27 (2009).

    PubMed  Google Scholar 

  188. Silberstein, S. D. Tonabersat, a novel gap-junction modulator for the prevention of migraine. Cephalalgia 29, 28–35 (2009).

    PubMed  Google Scholar 

  189. Kim, Y. et al. Tonabersat prevents inflammatory damage in the central nervous system by blocking connexin43 hemichannels. Neurotherapeutics 14, 1148–1165 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  190. Mao, Y. et al. Characterisation of Peptide5 systemic administration for treating traumatic spinal cord injured rats. Exp. Brain Res. 235, 3033–3048 (2017).

    CAS  PubMed  Google Scholar 

  191. Danesh-Meyer, H. V. et al. Connexin43 mimetic peptide reduces vascular leak and retinal ganglion cell death following retinal ischaemia. Brain 135, 506–520 (2012). This study reports that Peptide5 treatment increases neuronal rescue after retinal IRI.

    PubMed  Google Scholar 

  192. Chen, Y. S. et al. Intravitreal injection of lipoamino acid-modified connexin43 mimetic peptide enhances neuroprotection after retinal ischemia. Drug Deliv. Transl Res. 5, 480–488 (2015).

    CAS  PubMed  Google Scholar 

  193. Davidson, J. O. et al. Connexin hemichannel blockade improves outcomes in a model of fetal ischemia. Ann. Neurol. 71, 121–132 (2012).

    CAS  PubMed  Google Scholar 

  194. Davidson, J. O. et al. Connexin hemichannel blockade is neuroprotective after asphyxia in preterm fetal sheep. PLOS ONE 9, e96558 (2014).

    PubMed  PubMed Central  Google Scholar 

  195. Galinsky, R. et al. Connexin hemichannel blockade improves survival of striatal GABA-ergic neurons after global cerebral ischaemia in term-equivalent fetal sheep. Sci. Rep. 7, 6304 (2017). Here, Peptide5 infused for 24 hours following global ischaemia is shown to improve survival of striatal GABAergic neurons in sheep.

    PubMed  PubMed Central  Google Scholar 

  196. Chen, Y. S., Green, C. R., Wang, K., Danesh-Meyer, H. V. & Rupenthal, I. D. Sustained intravitreal delivery of connexin43 mimetic peptide by poly(D,L-lactide-co-glycolide) acid micro- and nanoparticles—closing the gap in retinal ischaemia. Eur. J. Pharm. Biopharm. 95, 378–386 (2015).

    PubMed  Google Scholar 

  197. Guo, C. X. et al. Connexin43 mimetic peptide improves retinal function and reduces inflammation in a light-damaged albino rat model. Invest. Ophthalmol. Vis. Sci. 57, 3961–3973 (2016). In this report, Peptide5 is found to significantly preserve photoreceptoral and postphotoreceptoral neurons in bright-light-treated albino rats.

    CAS  PubMed  Google Scholar 

  198. Kim, Y. et al. Characterizing the mode of action of extracellular connexin43 channel blocking mimetic peptides in an in vitro ischemia injury model. Biochim. Biophys. Acta 1861, 68–78 (2017).

    CAS  Google Scholar 

  199. Danesh-Meyer, H. V., Zhang, J., Acosta, M. L., Rupenthal, I. D. & Green, C. R. Connexin43 in retinal injury and disease. Prog. Retin. Eye Res. 51, 41–68 (2016). Here, the authors review the evidence that Cx43 hemichannels may be acting as a pathological pore in retinal injury and disease.

    CAS  PubMed  Google Scholar 

  200. Mugisho, O. O. et al. The inflammasome pathway is amplified and perpetuated in an autocrine manner through connexin43 hemichannel mediated ATP release. Biochim. Biophys. Acta 1862, 385–393 (2018).

    CAS  Google Scholar 

  201. Tonkin, R. S. et al. Attenuation of mechanical pain hypersensitivity by treatment with Peptide5, a connexin-43 mimetic peptide, involves inhibition of NLRP3 inflammasome in nerve-injured mice. Exp. Neurol. 300, 1–12 (2018). In this study, Peptide5 treatment results in significantly reduced Cx43 and microglial and astrocyte activity in the dorsal horn of the spinal cord in chronic constriction injury mice.

    CAS  PubMed  Google Scholar 

  202. Cea, L. A. et al. De novo expression of connexin hemichannels in denervated fast skeletal muscles leads to atrophy. Proc. Natl Acad. Sci. USA 110, 16229–16234 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  203. Cea, L. A. et al. Dexamethasone-induced muscular atrophy is mediated by functional expression of connexin-based hemichannels. Biochim. Biophys. Acta 1862, 1891–1899 (2016).

    CAS  PubMed  Google Scholar 

  204. Yi, C. et al. Astroglial connexin43 contributes to neuronal suffering in a mouse model of Alzheimer's disease. Cell Death Differ. 23, 1691–1701 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  205. Obert, E. et al. Targeting the tight junction protein, zonula occludens-1, with the connexin43 mimetic peptide, alphaCT1, reduces VEGF-dependent RPE pathophysiology. J. Mol. Med. 95, 535–552 (2017). Here, aCT1 delivered via eye drops is shown to reduce light-induced retinal pigment epithelium damage.

    CAS  PubMed  Google Scholar 

  206. Duchêne, A. et al. Impact of astroglial connexins on modafinil pharmacological properties. Sleep 39, 1283–1292 (2016). In this report, flecainide enhances the wake-promoting and procognitive effects of modafinil in narcoleptic orexin knockout mice.

    PubMed  PubMed Central  Google Scholar 

  207. Liu, X. et al. The psychostimulant modafinil enhances gap junctional communication in cortical astrocytes. Neuropharmacology (2013).

  208. Cruikshank, S. J. et al. Potent block of Cx36 and Cx50 gap junction channels by mefloquine. Proc. Natl Acad. Sci. USA 101, 12364–12369 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  209. Picoli, C. et al. Human connexin channel specificity of classical and new gap junction inhibitors. J. Biomol. Screen. 17, 1339–1347 (2012).

    PubMed  Google Scholar 

  210. Jeanson, T. et al. Potentiation of amitriptyline anti-hyperalgesic-like action by astroglial connexin 43 inhibition in neuropathic rats. Sci. Rep. 6, 38766 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  211. Dhein, S. et al. A new synthetic antiarrhythmic peptide reduces dispersion of epicardial activation recovery interval and diminishes alterations of epicardial activation patterns induced by regional ischemia. A mapping study. Naunyn Schmiedebergs Arch. Pharmacol. 350, 174–184 (1994).

    CAS  PubMed  Google Scholar 

  212. Lin, X., Zemlin, C., Hennan, J. K., Petersen, J. S. & Veenstra, R. D. Enhancement of ventricular gap-junction coupling by rotigaptide. Cardiovasc. Res. 79, 416–426 (2008). Here, rotigaptide is demonstrated to increase gap junction coupling in ventricular cardiomyocytes.

    CAS  PubMed  PubMed Central  Google Scholar 

  213. Jorgensen, N. R. et al. The antiarrhythmic peptide analog rotigaptide (ZP123) stimulates gap junction intercellular communication in human osteoblasts and prevents decrease in femoral trabecular bone strength in ovariectomized rats. Endocrinology 146, 4745–4754 (2005).

    PubMed  Google Scholar 

  214. Clarke, T. C., Thomas, D., Petersen, J. S., Evans, W. H. & Martin, P. E. The antiarrhythmic peptide rotigaptide (ZP123) increases gap junction intercellular communication in cardiac myocytes and HeLa cells expressing connexin 43. Br. J. Pharmacol. 147, 486–495 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  215. Hsieh, Y. C. et al. Gap junction modifier rotigaptide decreases the susceptibility to ventricular arrhythmia by enhancing conduction velocity and suppressing discordant alternans during therapeutic hypothermia in isolated rabbit hearts. Heart Rhythm. 13, 251–261 (2016). This study demonstrates that the gap junction modifier rotigaptide protects rabbit hearts from ventricular arrhythmias.

    PubMed  Google Scholar 

  216. Ueda, N., Yamamoto, M., Honjo, H., Kodama, I. & Kamiya, K. The role of gap junctions in stretch-induced atrial fibrillation. Cardiovasc. Res. 104, 364–370 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  217. Rossman, E. I. et al. The gap junction modifier, GAP-134 [(2S,4R)-1-(2-aminoacetyl)-4-benzamido-pyrrolidine-2-carboxylic acid], improves conduction and reduces atrial fibrillation/flutter in the canine sterile pericarditis model. J. Pharmacol. Exp. Ther. 329, 1127–1133 (2009).

    CAS  PubMed  Google Scholar 

  218. Ng, F. S. et al. Enhancement of gap junction function during acute myocardial infarction modifies healing and reduces late ventricular arrhythmia susceptibility. JACC Clin. Electrophysiol. 2, 574–582 (2016).

    PubMed  PubMed Central  Google Scholar 

  219. Pedersen, C. M. et al. Rotigaptide protects the myocardium and arterial vasculature from ischaemia reperfusion injury. Br. J. Clin. Pharmacol. 81, 1037–1045 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  220. Skyschally, A., Walter, B., Schultz Hansen, R. & Heusch, G. The antiarrhythmic dipeptide ZP1609 (danegaptide) when given at reperfusion reduces myocardial infarct size in pigs. Naunyn Schmiedebergs Arch. Pharmacol. 386, 383–391 (2013).

    CAS  PubMed  Google Scholar 

  221. Boengler, K., Bulic, M., Schreckenberg, R., Schluter, K. D. & Schulz, R. The gap junction modifier ZP1609 decreases cardiomyocyte hypercontracture following ischaemia/reperfusion independent from mitochondrial connexin 43. Br. J. Pharmacol. 174, 2060–2073 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  222. Laurent, G. et al. Effects of chronic gap junction conduction-enhancing antiarrhythmic peptide GAP-134 administration on experimental atrial fibrillation in dogs. Circ. Arrhythm. Electrophysiol. 2, 171–178 (2009).

    CAS  PubMed  Google Scholar 

  223. Engstrom, T. et al. Danegaptide for primary percutaneous coronary intervention in acute myocardial infarction patients: a phase 2 randomised clinical trial. Heart https://doi.org/10.1136/heartjnl-2017-312774 (2018).

    PubMed  Google Scholar 

  224. Huang, G. Y. et al. Alteration in connexin 43 gap junction gene dosage impairs conotruncal heart development. Dev. Biol. 198, 32–44 (1998).

    CAS  PubMed  Google Scholar 

  225. Lo, C. W., Waldo, K. L. & Kirby, M. L. Gap junction communication and the modulation of cardiac neural crest cells. Trends Cardiovasc. Med. 9, 63–69 (1999).

    CAS  PubMed  Google Scholar 

  226. Huang, G. Y. et al. Gap junction-mediated cell-cell communication modulates mouse neural crest migration. J. Cell Biol. 143, 1725–1734 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  227. Li, W. E. et al. An essential role for connexin43 gap junctions in mouse coronary artery development. Development 129, 2031–2042 (2002).

    CAS  PubMed  Google Scholar 

  228. Dobrowolski, R. & Willecke, K. Connexin-caused genetic diseases and corresponding mouse models. Antioxid. Redox Signal. 11, 283–295 (2009).

    CAS  PubMed  Google Scholar 

  229. Gabriel, H. D. et al. Transplacental uptake of glucose is decreased in embryonic lethal connexin26-deficient mice. J. Cell Biol. 140, 1453–1461 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  230. Winterhager, E. et al. Connexin expression patterns in human trophoblast cells during placental development. Placenta 20, 627–638 (1999).

    CAS  PubMed  Google Scholar 

  231. Nishii, K., Shibata, Y. & Kobayashi, Y. Connexin mutant embryonic stem cells and human diseases. World J. Stem Cells 6, 571–578 (2014).

    PubMed  PubMed Central  Google Scholar 

  232. Denoyelle, F. et al. Clinical features of the prevalent form of childhood deafness, DFNB1, due to a connexin-26 gene defect: implications for genetic counselling. Lancet 353, 1298–1303 (1999).

    CAS  PubMed  Google Scholar 

  233. Cohn, E. S. & Kelley, P. M. Clinical phenotype and mutations in connexin 26 (DFNB1/GJB2), the most common cause of childhood hearing loss. Am. J. Med. Genet. 89, 130–136 (1999).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  235. Smith, F. J., Morley, S. M. & McLean, W. H. A novel connexin 30 mutation in Clouston syndrome. J. Invest. Dermatol. 118, 530–532 (2002).

    CAS  PubMed  Google Scholar 

  236. Bosen, F. et al. The Clouston syndrome mutation connexin30 A88V leads to hyperproliferation of sebaceous glands and hearing impairments in mice. FEBS Lett. 588, 1795–1801 (2014).

    CAS  PubMed  Google Scholar 

  237. Huang, D., Chen, Y. S., Green, C. R. & Rupenthal, I. D. Hyaluronic acid coated albumin nanoparticles for targeted peptide delivery in the treatment of retinal ischaemia. Biomaterials 168, 10–23 (2018).

    CAS  PubMed  Google Scholar 

  238. Xu, L. et al. Design and characterization of a human monoclonal antibody that modulates mutant connexin 26 hemichannels implicated in deafness and skin disorders. Front. Mol. Neurosci. 10, 298 (2017).

    PubMed  PubMed Central  Google Scholar 

  239. Panchin, Y. et al. A ubiquitous family of putative gap junction molecules. Curr. Biol. 10, R473–R474 (2000). This paper reports the discovery of a new family of putative channel-forming proteins called pannexins.

    CAS  PubMed  Google Scholar 

  240. Baranova, A. et al. The mammalian pannexin family is homologous to the invertebrate innexin gap junction proteins. Genomics 83, 706–716 (2004).

    CAS  PubMed  Google Scholar 

  241. Panchin, Y. V. Evolution of gap junction proteins—the pannexin alternative. J. Exp. Biol. 208, 1415–1419 (2005).

    CAS  PubMed  Google Scholar 

  242. Bond, S. R. & Naus, C. C. The pannexins: past and present. Front. Physiol. 5, 58 (2014).

    PubMed  PubMed Central  Google Scholar 

  243. Lohman, A. W. & Isakson, B. E. Differentiating connexin hemichannels and pannexin channels in cellular ATP release. FEBS Lett. 588, 1379–1388 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  244. Isakson, B. E. & Thompson, R. J. Pannexin-1 as a potentiator of ligand-gated receptor signaling. Channels 8, 118–123 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  245. Penuela, S. et al. Pannexin 1 and pannexin 3 are glycoproteins that exhibit many distinct characteristics from the connexin family of gap junction proteins. J. Cell Sci. 120, 3772–3783 (2007).

    CAS  PubMed  Google Scholar 

  246. Boassa, D. et al. Pannexin1 channels contain a glycosylation site that targets the hexamer to the plasma membrane. J. Biol. Chem. 282, 31733–31743 (2007).

    CAS  PubMed  Google Scholar 

  247. Penuela, S., Bhalla, R., Nag, K. & Laird, D. W. Glycosylation regulates pannexin intermixing and cellular localization. Mol. Biol. Cell. 20, 4313–4323 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  248. Ambrosi, C. et al. Pannexin1 and pannexin2 channels show quaternary similarities to connexons and different oligomerization numbers from each other. J. Biol. Chem. 285, 24420–24431 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  249. Wang, J. et al. The membrane protein pannexin1 forms two open-channel conformations depending on the mode of activation. Sci. Signal. 7, ra69 (2014).

    PubMed  PubMed Central  Google Scholar 

  250. Thompson, R. J. & Macvicar, B. A. Connexin and pannexin hemichannels of neurons and astrocytes. Channels 2, 81–86 (2008).

    PubMed  Google Scholar 

  251. Vanden Abeele, F. et al. Functional implications of calcium permeability of the channel formed by pannexin 1. J. Cell Biol. 174, 535–546 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  252. Bhalla-Gehi, R., Penuela, S., Churko, J. M., Shao, Q. & Laird, D. W. Pannexin1 and pannexin3 delivery, cell surface dynamics, and cytoskeletal interactions. J. Biol. Chem. 285, 9147–9160 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  253. Boyce, A. K., Wicki-Stordeur, L. E. & Swayne, L. A. Powerful partnership: crosstalk between pannexin 1 and the cytoskeleton. Front. Physiol. 5, 27 (2014).

    PubMed  PubMed Central  Google Scholar 

  254. Bao, L., Locovei, S. & Dahl, G. Pannexin membrane channels are mechanosensitive conduits for ATP. FEBS Lett. 572, 65–68 (2004). This study reports the identification of ATP as a permeant of pannexin channels.

    CAS  PubMed  Google Scholar 

  255. Chekeni, F. B. et al. Pannexin 1 channels mediate 'find-me' signal release and membrane permeability during apoptosis. Nature 467, 863–867 (2010). Here, the authors demonstrate that ATP and uridine-5′-triphosphate release through caspase-cleaved pannexin 1 channels serves a functional role in apoptosis.

    CAS  PubMed  PubMed Central  Google Scholar 

  256. Whyte-Fagundes, P. & Zoidl, G. Mechanisms of pannexin1 channel gating and regulation. Biochim. Biophys. Acta 1860, 65–71 (2018).

    CAS  Google Scholar 

  257. Sandilos, J. K. et al. Pannexin 1, an ATP release channel, is activated by caspase cleavage of its pore-associated C terminal autoinhibitory region. J. Biol. Chem. 287, 11303–11311 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  258. Qu, Y. et al. Pannexin-1 is required for ATP release during apoptosis but not for inflammasome activation. J. Immunol. 186, 6553–6561 (2011).

    CAS  PubMed  Google Scholar 

  259. Chiu, Y. H. et al. A quantized mechanism for activation of pannexin channels. Nat. Commun. 8, 14324 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  260. Gehi, R., Shao, Q. & Laird, D. W. Pathways regulating the trafficking and turnover of pannexin1 protein and the role of the C-terminal domain. J. Biol. Chem. 286, 27639–27653 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  261. Boyce, A. K. J., Epp, A. L., Nagarajan, A. & Swayne, L. A. Transcriptional and post-translational regulation of pannexins. Biochim. Biophys. Acta 1860, 72–82 (2018).

    CAS  Google Scholar 

  262. Shao, Q. et al. A germline variant in the PANX1 gene has reduced channel function and is associated with multisystem dysfunction. J. Biol. Chem. 291, 12432–12443 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  263. Penuela, S. et al. Loss of pannexin 1 attenuates melanoma progression by reversion to a melanocytic phenotype. J. Biol. Chem. 287, 29184–29193 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  264. Kim, Y. et al. Connexins and pannexins in cerebral ischemia. Biochim. Biophys. Acta 1860, 224–236 (2018).

    CAS  Google Scholar 

  265. Dong, F., Yang, X. J., Jiang, T. B. & Chen, Y. Ischemia triggered ATP release through pannexin-1 channel by myocardial cells activates sympathetic fibers. Microvasc. Res. 104, 32–37 (2016).

    CAS  PubMed  Google Scholar 

  266. Thompson, R. J. Pannexin channels and ischaemia. J. Physiol. 593, 3463–3470 (2014).

    PubMed  PubMed Central  Google Scholar 

  267. Freitas-Andrade, M., Bechberger, J. F., MacVicar, B. A., Viau, V. & Naus, C. C. Pannexin1 knockout and blockade reduces ischemic stroke injury in female, but not in male mice. Oncotarget 8, 36973–36983 (2017).

    PubMed  PubMed Central  Google Scholar 

  268. Good, M. E. et al. Endothelial cell pannexin1 modulates severity of ischemic stroke by regulating cerebral inflammation and myogenic tone. JCI Insight 3, 96272 (2018).

    PubMed  Google Scholar 

  269. Santiago, M. F. et al. Targeting pannexin1 improves seizure outcome. PLOS ONE 6, e25178 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  270. Gulbransen, B. D. et al. Activation of neuronal P2X7 receptor-pannexin-1 mediates death of enteric neurons during colitis. Nat. Med. 18, 600–604 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  271. Chen, S. P. et al. Inhibition of the P2X7-PANX1 complex suppresses spreading depolarization and neuroinflammation. Brain 140, 1643–1656 (2017).

    PubMed  PubMed Central  Google Scholar 

  272. Seror, C. et al. Extracellular ATP acts on P2Y2 purinergic receptors to facilitate HIV-1 infection. J. Exp. Med. 208, 1823–1834 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  273. Moon, P. M. et al. Deletion of Panx3 prevents the development of surgically induced osteoarthritis. J. Mol. Med. 93, 845–856 (2015).

    CAS  PubMed  Google Scholar 

  274. Thompson, R. J. et al. Activation of pannexin-1 hemichannels augments aberrant bursting in the hippocampus. Science 322, 1555–1559 (2008).

    CAS  PubMed  Google Scholar 

  275. Dossi, E. et al. Pannexin-1 channels contribute to seizure generation in human epileptic brain tissue and in a mouse model of epilepsy. Sci. Transl Med. 10, eaar3796 (2018).

    PubMed  Google Scholar 

  276. Aquilino, M. S., Whyte-Fagundes, P., Zoidl, G. & Carlen, P. L. Pannexin-1 channels in epilepsy. Neurosci. Lett. https://doi.org/10.1016/j.neulet.2017.09.004 (2017).

    CAS  PubMed  Google Scholar 

  277. Penuela, S., Harland, L., Simek, J. & Laird, D. W. Pannexin channels and their links to human disease. Biochem. J. 461, 371–381 (2014).

    CAS  PubMed  Google Scholar 

  278. Burma, N. E. et al. Blocking microglial pannexin-1 channels alleviates morphine withdrawal in rodents. Nat. Med. 23, 355–360 (2017).

    CAS  PubMed  Google Scholar 

  279. Xu, J., Chen, L. & Li, L. Pannexin hemichannels: a novel promising therapy target for oxidative stress related diseases. J. Cell. Physiol. 233, 2075–2090 (2018).

    CAS  PubMed  Google Scholar 

  280. Willebrords, J., Maes, M., Crespo Yanguas, S. & Vinken, M. Inhibitors of connexin and pannexin channels as potential therapeutics. Pharmacol. Ther. 180, 144–160 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  281. Jiang, J. X. & Penuela, S. Connexin and pannexin channels in cancer. BMC Cell Biol. 17 (Suppl. 1), 12 (2016).

    PubMed  PubMed Central  Google Scholar 

  282. Kranz, K. et al. Expression of pannexin1 in the outer plexiform layer of the mouse retina and physiological impact of its knock-out. J. Comp. Neurol. 521, 1119–1135 (2012).

    Google Scholar 

  283. Bargiotas, P. et al. Pannexins in ischemia-induced neurodegeneration. Proc. Natl Acad. Sci. USA. 108, 20772–20777 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  284. Silverman, W., Locovei, S. & Dahl, G. Probenecid, a gout remedy, inhibits pannexin 1 channels. Am. J. Physiol. Cell Physiol. 295, C761–C767 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  285. Poon, I. K. et al. Unexpected link between an antibiotic, pannexin channels and apoptosis. Nature 507, 329–334 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  286. Good, M. E. et al. Pannexin 1 channels as an unexpected new target of the anti-hypertensive drug spironolactone. Circ. Res. 122, 606–615 (2018).

    CAS  PubMed  Google Scholar 

  287. Billaud, M. et al. A molecular signature in the pannexin1 intracellular loop confers channel activation by the α1 adrenoreceptor in smooth muscle cells. Sci. Signal. 8, ra17 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  288. Schulte, J., Sepp, K. J., Wu, C., Hong, P. & Littleton, J. T. High-content chemical and RNAi screens for suppressors of neurotoxicity in a Huntington's disease model. PLOS ONE 6, e23841 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  289. Michalski, K. & Kawate, T. Carbenoxolone inhibits pannexin1 channels through interactions in the first extracellular loop. J. Gen. Physiol. 147, 165–174 (2016).

    PubMed  PubMed Central  Google Scholar 

  290. Laird, D. W. & Revel, J. P. Biochemical and immunochemical analysis of the arrangement of connexin43 in rat heart gap junction membranes. J. Cell Sci. 97, 109–117 (1990).

    CAS  PubMed  Google Scholar 

  291. Gemel, J., Lin, X., Veenstra, R. D. & Beyer, E. C. N-Terminal residues in Cx43 and Cx40 determine physiological properties of gap junction channels, but do not influence heteromeric assembly with each other or with Cx26. J. Cell Sci. 119, 2258–2268 (2006).

    CAS  PubMed  Google Scholar 

  292. John, S. A. & Revel, J. P. Connexon integrity is maintained by non-covalent bonds: intramolecular disulfide bonds link the extracellular domains in rat connexin-43. Biochem. Biophys. Res. Commun. 178, 1312–1318 (1991).

    CAS  PubMed  Google Scholar 

  293. Morley, G. E., Ek-Vitorin, J. F., Taffet, S. M. & Delmar, M. Structure of connexin43 and its regulation by pHi. J. Cardiovasc. Electrophysiol. 8, 939–951 (1997).

    CAS  PubMed  Google Scholar 

  294. Solan, J. L. & Lampe, P. D. Connexin43 phosphorylation: structural changes and biological effects. Biochem. J. 419, 261–272 (2009).

    CAS  PubMed  Google Scholar 

  295. Leithe, E., Mesnil, M. & Aasen, T. The connexin 43 C-terminus: a tail of many tales. Biochim. Biophys. Acta 1860, 48–64 (2018).

    CAS  Google Scholar 

  296. Bai, D., Yue, B. & Aoyama, H. Crucial motifs and residues in the extracellular loops influence the formation and specificity of connexin docking. Biochim. Biophys. Acta 1860, 9–21 (2018).

    CAS  Google Scholar 

  297. Sonntag, S. et al. Mouse lens connexin23 (Gje1) does not form functional gap junction channels but causes enhanced ATP release from HeLa cells. Eur. J. Cell Biol. 88, 65–77 (2009).

    CAS  PubMed  Google Scholar 

  298. Leo-Macias, A., Agullo-Pascual, E. & Delmar, M. The cardiac connexome: non-canonical functions of connexin43 and their role in cardiac arrhythmias. Semin. Cell Dev. Biol. 50, 13–21 (2016).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Given the >20,000 reports on connexins and pannexins and reference number limitations, the authors apologize for not including many primary articles and for summarizing many exciting findings by citing reviews. Work in the authors' laboratories is supported by a US National Institutes of Health grant (GM55632) to P.D.L. and the Canadian Institutes of Health Research (130530, 123228, 148630 and 148584) and Canada Research Chair Program to D.W.L.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dale W. Laird.

Ethics declarations

Competing interests

P.D.L. is a non-paid member of the scientific advisory board of FirstString Research and has been granted stock options (currently no value) for his service. D.W.L. received a small 1-year grant from Zealand Pharma in 2016 to test potential pannexin 1-modulating peptides.

Related links

PowerPoint slides

Glossary

Connexins

Tetraspanning membrane proteins that are the molecular constituent of gap junctions.

Hemichannel

A connexon that has the ability to open for the passage of small molecules.

Ischaemia–reperfusion injury

(IRI). An injury that occurs when blood supply (oxygen) returns to tissue after a period of ischaemia (hypoxia). Damage following oxygen restoration results in inflammation and oxidative damage.

Connexon

A hexamer arrangement of connexins that contains the same or different connexin proteins arranged into an oligomer.

Gap junctional intercellular communication

(GJIC). The process in which contacting cells harbouring functional gap junctions exchange small molecules.

Perinexus

A zone at the periphery of gap junctions that is enriched in connexons and hemichannels, at which Cx43–zonula occludens 1 interactions can occur.

Connexosomes

Double-membraned structures formed from the internalization of gap junction components from contacting cells.

Intercalated discs

Specialized membrane structures at the ends of cardiomyocytes that contain desmosomes, adherens junctions, sodium channels and gap junctions and allow depolarization waves to transmit from one cell to its neighbour.

Peptide mimetics

Short peptide sequences of usually 8–24 amino acids that correspond to segments of connexin proteins that can be used to modulate connexin function.

Antisense oligodeoxynucleotide

(AsODN). A short-chain nucleotide that can be designed to target connexin-encoding RNA to block protein expression.

Zonula occludens 1

(ZO1). ZO1 is a membrane-associated guanylate kinase (MAGUK) scaffolding protein that interacts with Cx43 and tight junction-associated proteins.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Laird, D., Lampe, P. Therapeutic strategies targeting connexins. Nat Rev Drug Discov 17, 905–921 (2018). https://doi.org/10.1038/nrd.2018.138

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrd.2018.138

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research