Skip to main content

MINI REVIEW article

Front. Mol. Neurosci., 10 October 2017
Sec. Brain Disease Mechanisms
Volume 10 - 2017 | https://doi.org/10.3389/fnmol.2017.00320

Connexin43- and Pannexin-Based Channels in Neuroinflammation and Cerebral Neuropathies

Denis Sarrouilhe1* Catherine Dejean2 Marc Mesnil3
  • 1Laboratoire de Physiologie Humaine, Faculté de Médecine et Pharmacie, Université de Poitiers, Poitiers, France
  • 2Service Pharmacie, Pavillon Janet, Centre Hospitalier Henri Laborit, Poitiers, France
  • 3STIM Laboratory, ERL 7368-CNRS, Université de Poitiers, Pôle Biologie Santé, Poitiers, France

Connexins (Cx) are largely represented in the central nervous system (CNS) with 11 Cx isoforms forming intercellular channels. Moreover, in the CNS, Cx43 can form hemichannels (HCs) at non-junctional membrane as does the related channel-forming Pannexin1 (Panx1) and Panx2. Opening of Panx1 channels and Cx43 HCs appears to be involved in inflammation and has been documented in various CNS pathologies. Over recent years, evidence has accumulated supporting a link between inflammation and cerebral neuropathies (migraine, Alzheimer’s disease (AD), Parkinson’s disease (PD), major depressive disorder, autism spectrum disorder (ASD), epilepsy, schizophrenia, bipolar disorder). Involvement of Panx channels and Cx43 HCs has been also proposed in pathophysiology of neurological diseases and psychiatric disorders. Other studies showed that following inflammatory injury of the CNS, Panx1 activators are released and prolonged opening of Panx1 channels triggers neuronal death. In neuropsychiatric diseases, comorbidities are frequently present and can aggravate the symptoms and make therapeutic management more complex. The high comorbidity between some neuropathies can be partially understood by the fact that these diseases share a common etiology involving inflammatory pathways and Panx1 channels or Cx43 HCs. Thus, anti-inflammatory therapy opens perspectives of targets for new treatments and could have real potential in controlling a cerebral neuropathy and some of its comorbidities. The purpose of this mini review is to provide information of our knowledge on the link between Cx43- and Panx-based channels, inflammation and cerebral neuropathies.

Introduction

The innate immune system mediates inflammation for a physiological response to insult, infection, or biological stress. In most central nervous system (CNS) diseases, a common underlying factor seems to be the triggering of the inflammatory cascade with release of inflammatory cytokines (Vezzani et al., 2016). In brain, microglial cells predominantly confer innate immunity acting as resident macrophages of the CNS and represent the first line of defense against injury. However, excessive activation of microglia causes over-production of inflammatory cytokines directly affecting the CNS. Emerging evidence suggests that neurons, astrocytes and mastocytes also play important roles in neuroinflammation. Together with their strategic localization, their capacity to modulate microglial activation, their migration and activation at sites of injury demonstrate that mastocytes can initiate and/or modulate the neuroinflammatory process. Thus, understanding and control of interactions between the immune and the nervous systems might be a key for preventing most CNS diseases.

In neuropsychiatric diseases, comorbidities are frequently present and can aggravate the symptoms and make therapeutic management more complex. In migrepsy, a migraine syndrome with aura, an epileptic seizure follows migraine attack in a way suggesting that one would follow the other. A complex of comorbidities including migraine, major depression disorder (MDD) and suicide is also observable (Nye and Thadani, 2015). Depressive episodes are the most common comorbidity in epilepsy, affecting between 11% and 62% of epileptic patients (Błaszczyk and Czuczwar, 2016). A meta-analysis confirmed that psychiatric comorbidities are more common in patients with treatment-resistant epilepsy (Scott et al., 2017). In patients with epilepsy, the proportion of psychotic disorders is higher than in the non-epileptic population, but the increased risk of schizophrenia varies according to the study (Bakken et al., 2014). Interestingly, a link between epilepsy, autism spectrum disorder (ASD), depression and the brain inflammatory pathways was revealed (Mazarati et al., 2017). In the early stages of Alzheimer’s disease (AD), patients may have generalized convulsive seizures but also partial epilepsies located in the frontal or temporal lobe (Cretin et al., 2017). In addition, psychiatric comorbidities such as depression, schizophrenia and bipolar disorders may be severe, prodromal and predispose to the development of AD (Garcez et al., 2015).

A common part between all these brain pathologies/comorbidities could be inflammation in which gap junction proteins seem to be involved. Such an involvement is not surprising when considering that gap junctions and their structural proteins, the connexins (Cx), are very present in the CNS in which cells have to be efficiently connected for treating incoming information from the body and its environment and controlling consequently adapted physiological responses. And indeed, several CNS cell functions (electrical synapses in neurons, ionic and neurotransmitter buffering by astrocytes, propagation of astrocytic Ca2+ waves, myelin stabilization in oligodendrocytes, etc.) appeared to be supported by gap-junctional intercellular communication (GJIC; Giaume and Venance, 1998; Deans et al., 2001; Eugenin et al., 2012; Georgiou et al., 2017). Such crucial roles may explain the presence of 11 out 21 Cx isoforms in human CNS that are differently dispatched between astrocytes, oligodendrocytes, microglia and neurons (Giaume and Liu, 2012). However, more recently, it appeared that Cxs mediate communication between intracellular and extracellular compartments by forming non-juxtaposed hemichannels (HCs). In brain, this activity is mostly observed for Cx43 under various stimuli that permit the release of transmitters like glutamate and ATP (Giaume et al., 2013). Such activity is shared by pannexins (Panx) which are distant homologs of Cxs unable to form gap junctions but transmembrane channels. Interestingly, the different forms of communication permitted by Cxs and Panxs are involved in CNS inflammation with various effects that depend on the communication type. For instance, extracellular factors such as pro-inflammatory cytokines (IL-6, IL-1β, TNF-α) which are liberated by microglia in case of inflammation inhibit Cx43-mediated GJIC in astrocytes whereas their HCs are activated (Retamal et al., 2007). Cx43 HCs are indeed open during inflammation, contributing to the activation of the inflammasome pathway and its spread to neighboring cells (Kim et al., 2016). Similar action has been observed for Pannexin1 (Panx1), which activates inflammasome in astrocytes and is involved in ischemic injury (Bennett et al., 2012; Makarenkova and Shestopalov, 2014). All these observations elicit Cx43 and Panx1 as therapeutical targets whose inhibition could decrease inflammation in CNS. On this aspect, recent data identifying inhibitors of Cx43 HCs (tonabersat) or Panx1 channels (probenecid) to prevent inflammasome activation and damage in the CNS are encouraging (Jian et al., 2016; Kim et al., 2017).

The purpose of this mini review is to provide information on the link between Cx43- and Panx-based channels, inflammation and cerebral neuropathies.

Migraine with Aura

Neuroimaging and experimental studies suggest that cortical spreading depression (CSD), a slow wave of neuronal and glial depolarization, triggers migraine aura, activates the trigeminovascular system and is possibly responsible for migraine headache (Charles and Baca, 2013; Sarrouilhe et al., 2014). Recently, Karatas et al. (2013) elucidated the cellular and molecular mechanisms linking CSD induction and activation of the trigeminovascular system in mice, involving a parenchymal inflammatory process. Upon CSD induction, neuronal Panx1 channels transiently open with subsequent activation of a multiprotein complex (inflammasome) that mediates the innate inflammatory response. The initiation of the inflammatory response involves the proteolytic activation of caspase-1 and a release of high-mobility group box 1 (HMGB1) and interleukin-1β (IL-1β; Silverman et al., 2009; Karatas et al., 2013). Subsequently, in astrocytes forming the glia limitans, nuclear factor KappaB (NF-κB) is activated and translocated to the nucleus and nitric oxide synthase (iNOS) and cyclooxygenase-2 (COX2) are induced. Then, cytokines, prostanoids and NO, released in the subarachnoid space, promote an activation of the perivascular nociceptive trigeminal ending in pia mater that may induce headache. Pharmacological inhibition of neuronal Panx1 channels by carbenoxolone abolishes the inflammatory signaling cascade, perivascular nociceptive trigeminal activation and reduces CSD-induced mast cell degranulation. Even if these results suggest that neuronal Panx1 channels are a link between stressed neurons and subsequent inflammatory pathways, further investigations taking recourse to Panx1 knock-out mice are necessary to validate these findings (Karatas et al., 2013).

Alzheimer’s Disease

Several cell types would play a role in neuroinflammation observed in AD. Among them, microglial cells are known to contribute to the chronic inflammation state observed in AD. Increased levels of inflammatory mediators detected in brains from AD patients contribute to disease progression and severity (Heneka et al., 2015). In vitro treatment with active fragment of the amyloid beta peptide induces microglial Cx43 HC and Panx1 channel opening. Activated microglia releases pro-inflammatory cytokines that contribute to the amyloid peptide-induced Cx43 HCs and Panx1 channels opening in astrocytes (Orellana et al., 2011a). In reactive astrocytes, Cx43 is the main HC contributor whereas Panx1 channels are restricted to astrocyte subpopulation contacting amyloid plaques (Yi et al., 2016). Both microglia and astrocytes could release gliotransmitters (ATP/glutamate) through HCs, resulting in neuronal Cx36 HCs and Panx1 channels opening that triggers neuronal death (Orellana et al., 2011b; Koulakoff et al., 2012). Mastocyte secretory granules contain pro-inflammatory mediators released in the extracellular milieu via a Ca2+ influx upon activation. When the progression of AD is studied in a murine model of AD (APPswe/PS1dE9 mice), the number of mastocytes in cortical and hippocampal areas early increases. Moreover, patients with AD present mastocytes near their amyloid plaques. Acute treatment with amyloid beta peptide induces rapid degranulation of cultured mastocytes via a Panx1 channel-dependent Ca2+ influx and this response is prevented by Panx1 blockers. In brain mastocytes, acute treatment with amyloid beta peptide also induces activity of Panx1 channels and Cx43 HCs, an effect that is associated with enhanced histamine release. In the presence of amyloid plaques, brain mastocytes of APPswe/PS1dE9 mice show high Panx1 channel and Cx43 HC activity (Harcha et al., 2015). Thus, channel response of mastocytes to amyloid peptide treatment seems to occur earlier than microglia, astrocyte and neuron (Orellana et al., 2011b; Harcha et al., 2015). Therefore, mastocytes might be brain cells that play a critical role in the onset and progression of AD by early sensing amyloid peptide, releasing pro-inflammatory molecules and recruiting other cells to the neuroinflammatory response.

Parkinson’s Disease (PD)

Chronic neuroinflammation is a characteristic of PD. Glial cell activation and increased pro-inflammatory molecules are observed in brains of PD patients and animal models. The chronic release of pro-inflammatory cytokines exacerbates the motor symptoms of PD resulting from dopaminergic neuron degeneration in the substancia nigra pars compacta (Wang et al., 2015). Cx43 upregulation has been identified in the striatum of rodent models of PD and in cultured astrocytes stimulated with rotenone (Xie et al., 2015). Moreover, gastrodin, a constituent of a Chinese herbal medicine, ameliorates PD by downregulating astrocytic Cx43 (Wang et al., 2013). The neuronal build-up of protein aggregates containing alpha-synuclein (ASN) and their release to the extracellular space are considered to be responsible for the propagation of neurodegeneration in the brain of advanced PD patients (Wang et al., 2013). A recent study showed that extracellular ASN neurotoxicity is mediated by the P2X7 receptor signaling complex. Treatment of neuroblastoma cells and rat synaptoneurosomes with exogenous ASN activated P2X7 receptors leading to Panx 1 recruitment responsible for ATP release that could lead to neurotoxicity (Wilkaniec et al., 2017).

Major Depressive Disorder

Whereas many studies suggested astrocytic gap junction dysfunction to be part of MDD etiologies (Sarrouilhe and Dejean, 2015), the respective roles of GJIC and Cx43 HCs are not elucidated (Quesseveur et al., 2015; Jeanson et al., 2016). In a functional study made in cortical and striatal mouse astrocytes, tested antidepressants were shown to exert different effects on Cx43 GJIC and HC activities (Jeanson et al., 2016). Indeed, even if they exhibit opposed effects on GJIC within a same therapeutic class, all tested drugs inhibit Cx43 HCs (Jeanson et al., 2016). Experimental and clinical data point to a role for inflammation in the development of MDD (Capuron and Miller, 2011). In MDD patients, expression of pro-inflammatory cytokines such as TNF-α and IL1-β are increased and correlate with the MDD severity. Under neuroinflammation conditions, microglia is primary stimulated and releases these cytokines that open Cx43 HCs in astrocytes while no change is observed in GJIC (Abudara et al., 2015). Moreover, several antidepressants are known to inhibit the production of these cytokines. Recently, a study made on patients with mastocytosis, a rare accumulation and activation of mast cells in various tissues, demonstrates their possible involvement in inflammation-induced depression, confirming their implication in inflammatory diseases (Georgin-Lavialle et al., 2016). On the other hand, MDD and AD, frequently co-occur and it was suggested that depression increases the risk of subsequent AD. Globally, although the data is scattered, as in the case of AD, increased mastocytes HC activity might be an early player in the inflammation pathways linked to MDD. It is clear that further research is necessary to decipher the cascade of events taking place between mastocytes activation, astrocytic gap junctions and Cx43 HC involvement, and neuronal dysfunction.

Autism Spectrum Disorder

ASD is a complex group of disorders associated with aberrant chemical synaptic transmission and plasticity (Zoghbi and Bear, 2012). Young patients with ASD have a surplus of chemical synapses due to a slowdown process involving microglia that early eliminates about half of cortical synapses (Tang et al., 2014). Few data are available about the involvement of Cx43 in ASD. Studies using conditional Cx43 knockout mice show that Cx43 is important for neurodevelopment (Wiencken-Barger et al., 2007). In postmortem brain tissues of ASD patients, Cx43 expression is increased in superior frontal cortex, a region which dysfunction may be responsible for cognition deficit observed in this neuropathy (Fatemi et al., 2008). Moreover, growing evidence indicates that the two forms of synapses interact during brain development and could contribute, together, to ASD (Miller et al., 2015; Pereda, 2015). Several studies have shown that neuroinflammation plays an important role in ASD and that mastocytes are overactivated. In ASD, pre-natal or early post-natal inflammatory and infectious processes correlate to neurodevelopmental dysfunction (Wang et al., 2014). Perinatal mastocytes activation by various triggers (infectious, stress-related, environmental, allergic) with a subsequent release of pro-inflammatory and neurotoxic molecules can contribute to brain inflammation in ASD pathogenesis through a cascade of events involving glial cells (microglia, astrocytes, oligodendrocytes), neurons, Cx43 HCs and Panx channels (Orellana et al., 2011b; Aguirre et al., 2013). It was also proposed that loss of Cx43 in the enteric glial cells contribute to brain inflammation in ASD by inducing disturbances in the gut-brain axis (Grubišc and Parpura, 2015).

Epilepsy

Experimental approaches demonstrated a prominent role of glial cells, activated during infectious and non-infectious causes of inflammation, in the mechanisms of seizure precipitation and recurrence. Both causes of inflammation share common pathways with activation of microglia and astrocytes releasing pro-inflammatory mediators that perturb glioneuronal communication and have proictogenic properties (Vezzani et al., 2016). Although GJs, Cx and Panx are related to the pathophysiology of epilepsy, their precise involvement is not elucidated. Besides interneuronal GJIC, inter-glial GJIC appears important for seizure generation. However, there is evidence from most animal models and patients studies that Cx43 expression increases in glia but not in neurons, opening the question on the roles of glia in seizure generation. A systematic analysis of the literature reveals significant changes in expression of the astrocytic Cx43 and Panx1 in an in vitro mouse seizure model and an increase of Panx1-2 expression in animal and human epileptic tissues (Mylvaganam et al., 2014). Using various approaches and Panx1-deleted mice, Panx1 channels (from glia and/or neurons) were proposed to contribute to status epilepticus in vivo. The data of this study are consistent with the following model: (1) the intense neuronal activity elevates extracellular K+; (2) Panx1 channels are activated; (3) ATP is released; and (4) P2X receptors are activated leading to neuronal hyperactivity and this positive feedback mechanism amplifies seizures (Santiago et al., 2011). Antiepileptic drugs tested in an astroglia/microglia co-culture model of inflammation did not alter Cx43 expression (Dambach et al., 2014). In contrast, tonabersat, a compound active against neuronal hyperexcitability and neurogenic inflammation that was proposed in the treatment for epilepsy and as a prophylactic treatment for migraine with aura, prevents inflammatory damage in the CNS by blocking Cx43 HCs (Kim et al., 2017).

Other Pathologies Potentially Mediated by Cx43 and Panx-Based Channels Activity

A recent analysis of all available data on neuroinflammation in postmortem brains of schizophrenia patients revealed variable results in astrocytic and microglial markers, glial cell density and pro-inflammatory cytokine concentration (Trépanier et al., 2016). Moreover, prenatal exposure to inflammatory conditions (i.e., LPS-exposed dams) revealed that the release of pro-inflammatory cytokines (IL-1β/TNF-α) and ATP through the activation of astrocytic Cx43 HCs and Panx1 channels results in an astrocyte-neuron crosstalk with paracrine activation of neuronal P2X7 receptors, Panx1 channels and a subsequent increased neuronal death. All these data suggest that prenatal infections could contribute to the development of neuropsychological disorders in children including schizophrenia (Avendaño et al., 2015). However, a recent work does not support a major contribution of Panx1-3 to disease risk of schizophrenia (Gawlik et al., 2016).

In the case of postmortem bipolar disorder patients, a post-mortem study revealed in their frontal cortex an increase of excitotoxicity and neuroinflammatory markers. This upregulation might explain the neurodegenerative component of bipolar disorder, with cell death, brain atrophy and cognitive decline. Markers of astrocyte and microglial activation are also upregulated in frontal cortex from bipolar disorder patients (Rao et al., 2010). Interestingly, in light of the involvement of mastocytes in some other psychiatric neuropathies, bipolar disorder is a psychiatric symptom observed in mast cell activation syndrome (Afrin, 2014). A dysregulation of Cx expression in the astrocytic syncytium can cause an imbalance in glutamatergic tripartite synapses and was proposed to be responsible for the pathophysiology of bipolar disorder (Mitterauer, 2011).

Conclusion

Researching links between inflammation, neurological diseases and psychiatric disorders with Cx and Panx channels is at a very nascent stage but has the potential to improve our understanding of these diseases to establish effective therapeutics. For some diseases like migraine with aura or AD, a signaling cascade is proposed while for others the data are too sparse. Common inflammatory pathways can explain the high comorbidity between some neuropathies. Thus, anti-inflammatory therapy could have real potential in controlling a cerebral neuropathy and some of its comorbidities (Mazarati et al., 2017). In this line, tonabersat, a compound active against neurogenic inflammation proposed in the treatment for epilepsy and as a prophylactic treatment for migraine with aura, two frequent comorbidities, prevents inflammatory damage in the CNS by blocking Cx43 HCs (Kim et al., 2017). Cx43 HCs can open in response to injury or inflammatory factors and are thus implicated in brain neuropathies, especially through the inflammasome pathway. Modulating the Cx43 HCs opening can prevent tissue damage arising from excessive and uncontrolled inflammation (Kim et al., 2016).

Panx1 channels drive inflammation by the regulation of inflammasome, the release of pro-inflammatory cytokines and the activation and migration of leukocytes (Crespo Yanguas et al., 2017). They also facilitate neuronal cell death that potentially implicates them in neurodegenerative disorders (Shestopalov and Slepak, 2014). Blockers of Panx1 channels are not highly selective as they also block GJIC, and many deleterious side effects limit their pharmacological potential.

The development of new therapeutic tools to inhibit selectively Panx1 channels and Cx43 HCs in cell subpopulations of the CNS will undoubtedly make them a promising target for anti-inflammatory therapy that could have real potential in prevention or delay of neurological diseases, psychiatric disorders and some of their comorbidities. However, such a strategy should not only focused on Cx43 but also to other highly expressed Cxs in the CNS like Cx36 which appears to be involved in several neuronal injuries (autism, ischemia, retina, etc.; Welsh et al., 2005; Bargiotas et al., 2012; Ivanova et al., 2016). Another main goal for the future is to define the respective role of Cx HCs and Panx1 channels in neuroinflammation and cerebral neuropathies. Among the few studies in this area, it was shown in reactive astrocytes of a mouse model of AD that their respective contribution seems to depend on the local environment context (Yi et al., 2016).

Author Contributions

DS wrote and edited the manuscript, conceived and designed the major ideas developed in the manuscript. MM and CD have made substantial, direct and intellectual contribution to the work. All authors read and approved the final manuscript.

Conflict of Interest Statement

The authors declare that the work was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

Abudara, V., Roux, L., Dallérac, G., Matias, I., Dulong, J., Mothet, J. P., et al. (2015). Activated microglia impairs neuroglial interaction by opening Cx43 hemichannels in hippocampal astrocytes. Glia 63, 795–811. doi: 10.1002/glia.22785

PubMed Abstract | CrossRef Full Text | Google Scholar

Afrin, L. B. (2014). The presentation, diagnosis and treatment of mast cell activation syndrome. Curr. Allergy Clin. Immunol. 27, 146–160.

Google Scholar

Aguirre, A., Maturana, C. J., Harcha, P. A., and Sáez, J. C. (2013). Possible involvement of TLRs and hemichannels in stress-induced CNS dysfunction via mastocytes and glia activation. Mediators Inflamm. 2013:893521. doi: 10.1155/2013/893521

PubMed Abstract | CrossRef Full Text | Google Scholar

Avendaño, B. C., Montero, T. D., Chávez, C. E., von Bernhardi, R., and Orellana, J. A. (2015). Prenatal exposure to inflammatory conditions increases Cx43 and Panx1 unopposed channel opening and activation of astrocytes in the offspring effect on neuronal survival. Glia 63, 2058–2072. doi: 10.1002/glia.22877

PubMed Abstract | CrossRef Full Text | Google Scholar

Bakken, I. J., Revdal, E., Nesvåg, R., Brenner, E., Knudsen, G. P., Surén, P., et al. (2014). Substance use disorders and psychotic disorders in epilepsy; a population-based registry story. Epilepsy Res. 108, 1435–1443. doi: 10.1016/j.eplepsyres.2014.06.021

PubMed Abstract | CrossRef Full Text | Google Scholar

Bargiotas, P., Muhammad, S., Rahman, M., Jakob, N., Trabold, R., Fuchs, E., et al. (2012). Connexin 36 promotes cortical spreading depolarization and ischemic brain damage. Brain Res. 1479, 80–85. doi: 10.1016/j.brainres.2012.08.046

PubMed Abstract | CrossRef Full Text | Google Scholar

Bennett, M. V., Garré, J. M., Orellana, J. A., Bukauskas, F. F., Nedergaard, M., and Sáez, J. C. (2012). Connexin and pannexin hemichannels in inflammatory responses of glia and neurons. Brain Res. 1487, 3–15. doi: 10.1016/j.brainres.2012.08.042

PubMed Abstract | CrossRef Full Text | Google Scholar

Błaszczyk, B., and Czuczwar, S. J. (2016). Epilepsy coexisting with depression. Pharmacol. Reports. 68, 1084–1092. doi: 10.1016/j.pharep.2016.06.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Capuron, L., and Miller, A. H. (2011). Immune system to brain signaling: neuropsychopharmacological implications. Pharmacol. Ther. 130, 226–238. doi: 10.1016/j.pharmthera.2011.01.014

PubMed Abstract | CrossRef Full Text | Google Scholar

Charles, A. C., and Baca, S. M. (2013). Cortical spreading depression and migraine. Nat. Rev. Neurol. 9, 637–644. doi: 10.1038/nrneurol.2013.192

PubMed Abstract | CrossRef Full Text | Google Scholar

Crespo Yanguas, S., Willebrords, J., Johnstone, S. R., Maes, M., Decrock, E., De Bock, M., et al. (2017). Pannexin1 as mediator of inflammation and cell death. Biochim. Biophys. Acta 1864, 51–61. doi: 10.1016/j.bbamcr.2016.10.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Cretin, B., Philippi, N., Bousiges, O., Dibitonto, L., Sellal, F., Martin-Hunyadi, C., et al. (2017). Do we know how to diagnose epilepsy early in Alzheimer’s disease? Rev. Neurol. 173, 374–380. doi: 10.1016/j.neurol.2017.03.028

PubMed Abstract | CrossRef Full Text | Google Scholar

Dambach, H., Hinkerohe, D., Prochnow, N., Stienen, M. N., Moinfar, Z., Haase, C. G., et al. (2014). Glia and epilepsy: experimental investigation of antiepileptic drugs in an astroglia/microglia co-culture model of inflammation. Epilepsia 55, 184–192. doi: 10.1111/epi.12473

PubMed Abstract | CrossRef Full Text | Google Scholar

Deans, M. R., Gibson, J. R., Sellitto, C., Connors, B. W., and Paul, D. L. (2001). Synchronous activity of inhibitory networks in neocortex requires electrical synapses containing connexin36. Neuron 31, 477–485. doi: 10.1016/s0896-6273(01)00373-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Eugenin, E. A., Basilio, D., Sáez, J. C., Orellana, J. A., Raine, C. S., Bukauskas, F., et al. (2012). The role of gap junction channels during physiologic and pathologic conditions of the human central nervous system. J. Neuroimmune Pharmacol. 7, 499–518. doi: 10.1007/s11481-012-9352-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Fatemi, S. H., Folsom, T. D., Reutiman, T. J., and Lee, S. (2008). Expression of astrocytic markers aquaporin 4 and connexin 43 is altered in brains of subjects with autism. Synapse 62, 501–507. doi: 10.1002/syn.20519

PubMed Abstract | CrossRef Full Text | Google Scholar

Garcez, M. L., Falchetti, A. C., Mina, F., and Budni, J. (2015). Alzheimer’s disease associated with psychiatric comorbidities. An. Acad. Bras. Cienc. 87, 1461–1473. doi: 10.1590/0001-3765201520140716

PubMed Abstract | CrossRef Full Text | Google Scholar

Gawlik, M., Wagner, M., Pfuhlmann, B., and Stöber, G. (2016). The role of Pannexin gene variants in schizophrenia: systematic analysis of phenotypes. Eur. Arch. Psychiatry Clin. Neurosci. 266, 433–437. doi: 10.1007/s00406-015-0619-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Georgin-Lavialle, S., Moura, D. S., Salvador, A., Chauvet-Gelinier, J.-C., Launay, J.-M., Damaj, G., et al. (2016). Mast cells’ involvement in inflammation pathways linked to depression: evidence in mastocytosis. Mol. Psychiatry 21, 1511–1516. doi: 10.1038/mp.2015.216

PubMed Abstract | CrossRef Full Text | Google Scholar

Georgiou, E., Sidiropoulou, K., Richter, J., Papaneophytou, C., Sargiannidou, I., Kagiava, A., et al. (2017). Gene therapy targeting oligodendrocytes provides therapeutic benefit in a leukodystrophy model. Brain 140, 599–616. doi: 10.1093/brain/aww351

PubMed Abstract | CrossRef Full Text | Google Scholar

Giaume, C., Leybaert, L., Naus, C. C., and Saez, J. C. (2013). Connexin and pannexin hemichannels in brain glial cells: properties, pharmacology, and roles. Front. Pharmacol. 4:88. doi: 10.3389/fphar.2013.00088

PubMed Abstract | CrossRef Full Text | Google Scholar

Giaume, C., and Liu, X. (2012). From a glial syncytium to a more restricted and specific glial networking. J. Physiol. Paris 106, 34–39. doi: 10.1016/j.jphysparis.2011.09.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Giaume, C., and Venance, L. (1998). Intercellular calcium signaling and gap junctional communication in astrocytes. Glia 24, 50–64. doi: 10.1002/(sici)1098-1136(199809)24:1<50::aid-glia6>3.3.co;2-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Grubišc, V., and Parpura, V. (2015). The second brain in autism spectrum disorder:could connexin43 expressed in enteric glial cells play a role? Front. Cell. Neurosci. 9:242. doi: 10.3389/fncel.2015.00242

PubMed Abstract | CrossRef Full Text | Google Scholar

Harcha, P. A., Vargas, A., Yi, C., Koulakoff, A. A., Giaume, C., and Sáez, J. C. (2015). Hemichannels are required for amyloid β-peptide-induced degranulation and are activated in brain mast cells of APPswe/PS1dE9 mice. J. Neurosci. 35, 9526–9538. doi: 10.1523/JNEUROSCI.3686-14.2015

PubMed Abstract | CrossRef Full Text | Google Scholar

Heneka, M. T., Carson, M. J., El Khoury, J., Landreth, G. E., Brosseron, F., Feinstein, D. L., et al. (2015). Neuroinflammation in Alzheimer’s disease. Lancet Neurol. 14, 388–405. doi: 10.1016/S1474-4422(15)70016-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Ivanova, E., Yee, C. W., Baldoni, R., and Sagdullaev, B. T. (2016). Aberrant activity in retinal degeneration impairs central visual processing and relies on Cx36-containing gap junctions. Exp. Eye Res. 150, 81–89. doi: 10.1016/j.exer.2015.05.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Jeanson, T., Pondaven, A., Ezan, P., Mouthon, F., Charvériat, M., and Giaume, C. (2016). Antidepressants impact connexin43 channel functions in astrocytes. Front. Cell. Neurosci. 9:495. doi: 10.3389/fncel.2015.00495

PubMed Abstract | CrossRef Full Text | Google Scholar

Jian, Z., Ding, S., Deng, H., Wang, J., Yi, W., Wang, L., et al. (2016). Probenecid protects against oxygen-glucose deprivation injury in primary astrocytes by regulating inflammasome activity. Brain Res. 1643, 123–129. doi: 10.1016/j.brainres.2016.05.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Karatas, H., Erdener, S. E., Gursoy-Ozdemir, Y., Lule, S., Eren-Koçak, E., Sen, Z. D., et al. (2013). Spreading depression triggers headache by activating neuronal panx1 channels. Science 339, 1092–1095. doi: 10.1126/science.1231897

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, Y., Davidson, J. O., Gunn, K. C., Phillips, A. R., Green, C. R., and Gunn, A. J. (2016). Role of hemichannels in CNS inflammation and the inflammasome pathway. Adv. Protein Chem. Struct. Biol. 104, 1–37. doi: 10.1016/bs.apcsb.2015.12.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, Y., Griffin, J. M., Nor, M. N. M., Zhang, J., Freestone, P. S., Danesh-Meyer, H. V., et al. (2017). Tonabersat prevents inflammatory damage in the central nervous system by blocking connexin43 hemichannels. Neurotherapeutics doi: 10.1007/s13311-017-0536-9 [Epub ahead of print].

PubMed Abstract | CrossRef Full Text | Google Scholar

Koulakoff, A., Mei, X., Orellana, J. A., Sáez, J. C., and Giaume, C. (2012). Glial connexin expression and function in the context of Alzheimer’s disease. Biochim. Biophys. Acta 1818, 2048–2057. doi: 10.1016/j.bbamem.2011.10.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Makarenkova, H. P., and Shestopalov, V. I. (2014). The role of pannexin hemichannels in inflammation and regeneration. Front. Physiol. 5:63. doi: 10.3389/fphys.2014.00063

PubMed Abstract | CrossRef Full Text | Google Scholar

Mazarati, A. M., Lewis, M. L., and Pittman, Q. J. (2017). Neurobehavioral comorbidities of epilepsy: role of inflammation. Epilepsia 58, 48–56. doi: 10.1111/epi.13786

PubMed Abstract | CrossRef Full Text | Google Scholar

Miller, A. C., Voelker, L. H., Shah, A. N., and Moens, C. B. (2015). Neurobeachin is required postsynaptically for electrical and chemical synapse formation. Curr. Biol. 25, 16–28. doi: 10.1016/j.cub.2014.10.071

PubMed Abstract | CrossRef Full Text | Google Scholar

Mitterauer, B. J. (2011). Downregulation and upregulation of glial connexins may cause synaptic imbalances responsible for the pathophysiology of bipolar disorder. CNS Neurosci. Ther. 17, 281–293. doi: 10.1111/j.1755-5949.2010.00178.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Mylvaganam, S., Ramani, M., Krawczyk, M., and Carlen, P. L. (2014). Roles of gap junctions, connexins and pannexins in epilepsy. Front. Physiol. 5:172. doi: 10.3389/fphys.2014.00172

PubMed Abstract | CrossRef Full Text | Google Scholar

Nye, B. L., and Thadani, V. M. (2015). Migraine and epilepsy: review of the literature. Headache 55, 359–380. doi: 10.1111/head.12536

PubMed Abstract | CrossRef Full Text | Google Scholar

Orellana, J. A., Giaume, C., and Saez, J. C. (2011a). “Role of connexin hemichannels in neurodegeneration,” in Neurodegenerative Diseases-Processes, Prevention, Protection and Monitoring, ed. R. C.-C. Chang (Rijeka: InTech), 235–254.

Google Scholar

Orellana, J. A., Shoji, K. F., Abudara, V., Ezan, P., Amigou, E., Sáez, P. J., et al. (2011b). Amyloid β-induced death in neurons involves glial and neuronal hemichannels. J. Neurosci. 31, 4962–4977. doi: 10.1523/JNEUROSCI.6417-10.2011

PubMed Abstract | CrossRef Full Text | Google Scholar

Pereda, A. E. (2015). Neurobiology: all synapses are created equal. Curr. Biol. 25, R38–R41. doi: 10.1016/j.cub.2014.11.029

PubMed Abstract | CrossRef Full Text | Google Scholar

Quesseveur, G., Portal, B., Basile, J. A., Ezan, P., Mathou, A., Halley, H., et al. (2015). Attenuated levels of hippocampal connexin43 and its phosphorylation correlate with antidepressant- and anxiolytic-like activities in mice. Front. Cell. Neurosci. 9:490. doi: 10.3389/fncel.2015.00490

PubMed Abstract | CrossRef Full Text | Google Scholar

Rao, J. S., Harry, G. J., Rapoport, S. I., and Kim, H. W. (2010). Increased excitotoxicity and neuroinflammatory markers in postmortem frontal cortex from bipolar disorder patients. Mol. Psychiatry 15, 384–392. doi: 10.1038/mp.2009.47

PubMed Abstract | CrossRef Full Text | Google Scholar

Retamal, M. A., Froger, N., Palacios-Prado, N., Ezan, P., Séez, P. J., Saez, J. C., et al. (2007). Cx43 hemichannels and gap junction channels in astrocytes are regulated oppositely by proinflammatory cytokines released from activated microglia. J. Neurosci. 27, 13781–13792. doi: 10.1523/JNEUROSCI.2042-07.2007

PubMed Abstract | CrossRef Full Text | Google Scholar

Santiago, M. F., Veliskova, J., Patel, N. K., Lutz, S. E., Caille, D., Charollais, A., et al. (2011). Targeting pannexin1 improves seizure outcome. PLoS One 6:e25178. doi: 10.1371/journal.pone.0025178

PubMed Abstract | CrossRef Full Text | Google Scholar

Sarrouilhe, D., and Dejean, C. (2015). Gap junctions: a new therapeutic target in major depressive disorder? Rev. Neurol. 171, 762–767. doi: 10.1016/j.neurol.2015.07.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Sarrouilhe, D., Dejean, C., and Mesnil, M. (2014). Involvement of gap junction channels in the pathophysiology of migraine with aura. Front. Physiol. 5:78. doi: 10.3389/fphys.2014.00078

PubMed Abstract | CrossRef Full Text | Google Scholar

Scott, A. J., Sharpe, L., Hunt, C., and Gandy, M. (2017). Anxiety and depressive disorders in people with epilepsy: a meta-analysis. Epilepsia 58, 973–982. doi: 10.1111/epi.13769

PubMed Abstract | CrossRef Full Text | Google Scholar

Shestopalov, V. I., and Slepak, V. Z. (2014). Molecular pathways of pannexin1-mediated neurotoxicity. Front. Physiol. 5:23. doi: 10.3389/fphys.2014.00023

PubMed Abstract | CrossRef Full Text | Google Scholar

Silverman, W. R., de Rivero Vaccari, J. P., Locovei, S., Qiu, F., Carlsson, S. K., Scemes, E., et al. (2009). The pannexin 1 channel activates the inflammasome in neurons and astrocytes. J. Biol. Chem. 284, 18143–18151. doi: 10.1074/jbc.M109.004804

PubMed Abstract | CrossRef Full Text | Google Scholar

Tang, G., Gudsnuk, K., Kuo, S.-H., Cotrina, M. L., Rosoklija, G., Sosunov, A., et al. (2014). Loss of mTOR-dependent macroautophagy causes autistic-like synaptic pruning deficits. Neuron 83, 1131–1143. doi: 10.1016/j.neuron.2014.07.040

PubMed Abstract | CrossRef Full Text | Google Scholar

Trépanier, M. O., Hopperton, K. E., Mizrahi, R., Mechawar, N., and Bazinet, R. P. (2016). Postmortem evidence of cerebral inflammation in schizophrenia: a systematic review. Mol. Psychiatry 21, 1009–1026. doi: 10.1038/mp.2016.90

PubMed Abstract | CrossRef Full Text | Google Scholar

Vezzani, A., Fujinami, R. S., White, H. S., Preux, P. M., Blümcke, I., Sander, J. W., et al. (2016). Infections, inflammation and epilepsy. Acta Neuropathol. 131, 211–234. doi: 10.1007/s00401-015-1481-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, Q., Liu, Y., and Zhou, J. (2015). Neuroinflammation in Parkinson’s disease and its potential as therapeutic target. Transl. Neurodegener. 4:19. doi: 10.1186/s40035-015-0042-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, Q. M., Luo, A. Z., and Kong, X. (2014). Neuroinflammation and autism. N. A. J. Med. Sci. 7, 118–122. doi: 10.7156/najms.2014.0703118

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, Y., Wu, Z., Liu, X., and Fu, Q. (2013). Gastrodin ameliorates Parkinson’s disease by downregulating connexin43. Mol. Med. Rep. 8, 585–590. doi: 10.3892/mmr.2013.1535

PubMed Abstract | CrossRef Full Text | Google Scholar

Welsh, J. P., Ahn, E. S., and Placantonakis, D. G. (2005). Is autism due to brain desynchronization? Int. J. Dev. Neurosci. 23, 253–263. doi: 10.1016/j.ijdevneu.2004.09.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Wiencken-Barger, A. E., Djukic, B., Casper, K. B., and McCarthy, K. D. (2007). A role for Connexin43 during neurodevelopment. Glia 55, 675–686. doi: 10.1002/glia.20484

PubMed Abstract | CrossRef Full Text | Google Scholar

Wilkaniec, A., Gąssowska, M., Czapsli, G. A., Cieślik, M., Sulkowsli, G., and Adamczyk, A. (2017). P2X7 receptor-pannexin 1 interaction mediates extracellular α-synuclein-induced ATP release in neuroblastoma SH-SY5Y cells. Purinergic Signal. 13, 347–361. doi: 10.1007/s11302-017-9567-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Xie, H.-Y., Cui, Y., Deng, F., and Feng, J.-C. (2015). Connexin: a potential novel target for protecting the central nervous system? Neural Regen. Res. 10, 659–666. doi: 10.4103/1673-5374.155444

PubMed Abstract | CrossRef Full Text | Google Scholar

Yi, C., Mei, X., Ezan, P., Mato, S., Matias, I., Giaume, C., et al. (2016). Astroglial connexin43 contributes to neuronal suffering in a mouse model of Alzheimer’s disease. Cell Death Differ. 23, 1691–1701. doi: 10.1038/cdd.2016.63

PubMed Abstract | CrossRef Full Text | Google Scholar

Zoghbi, H. Y., and Bear, M. F. (2012). Synaptic dysfunction in neurodevelopmental disorders associated with autism and intellectual disabilities. Cold Spring Harb. Perspect. Biol 4:a009886. doi: 10.1101/cshperspect.a009886

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: connexin, pannexin, hemichannels, inflammation, neuropathies, comorbidity, gap junctions, central nervous system

Citation: Sarrouilhe D, Dejean C and Mesnil M (2017) Connexin43- and Pannexin-Based Channels in Neuroinflammation and Cerebral Neuropathies. Front. Mol. Neurosci. 10:320. doi: 10.3389/fnmol.2017.00320

Received: 24 July 2017; Accepted: 20 September 2017;
Published: 10 October 2017.

Edited by:

Juan Andrés Orellana, Pontificia Universidad Católica de Chile, Chile

Reviewed by:

Valery I. Shestopalov, University of Miami, United States
Colin Green, University of Auckland, New Zealand

Copyright © 2017 Sarrouilhe, Dejean and Mesnil. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Denis Sarrouilhe, denis.sarrouilhe@univ-poitiers.fr

Download