Next Article in Journal
The Role of SINE-VNTR-Alu (SVA) Retrotransposons in Shaping the Human Genome
Next Article in Special Issue
Connexin 43 Mutations Lead to Increased Hemichannel Functionality in Skin Disease
Previous Article in Journal
Plasma Levels of snoRNAs are Associated with Platelet Activation in Patients with Peripheral Artery Disease
Previous Article in Special Issue
Connexin43 Is Required for the Effective Activation of Spleen Cells and Immunoglobulin Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Connexins-Based Hemichannels/Channels and Their Relationship with Inflammation, Seizures and Epilepsy

by
Laura Medina-Ceja
1,*,
Juan C. Salazar-Sánchez
1,
Jorge Ortega-Ibarra
2 and
Alberto Morales-Villagrán
2
1
Laboratory of Neurophysiology, Department of Cellular and Molecular Biology, CUCBA, University of Guadalajara, Guadalajara, Jalisco 45110, México
2
MexBio Research Innovations S.A. de C.V., El Salto, Jalisco 45696 México
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2019, 20(23), 5976; https://doi.org/10.3390/ijms20235976
Submission received: 3 October 2019 / Revised: 13 November 2019 / Accepted: 20 November 2019 / Published: 27 November 2019
(This article belongs to the Special Issue Connexin-Based Channels in Inflammatory Processes)

Abstract

:
Connexins (Cxs) are a family of 21 protein isoforms, eleven of which are expressed in the central nervous system, and they are found in neurons and glia. Cxs form hemichannels (connexons) and channels (gap junctions/electric synapses) that permit functional and metabolic coupling between neurons and astrocytes. Altered Cx expression and function is involved in inflammation and neurological diseases. Cxs-based hemichannels and channels have a relevance to seizures and epilepsy in two ways: First, this pathological condition increases the opening probability of hemichannels in glial cells to enable gliotransmitter release, sustaining the inflammatory process and exacerbating seizure generation and epileptogenesis, and second, the opening of channels favors excitability and synchronization through coupled neurons. These biological events highlight the global pathological mechanism of epilepsy, and the therapeutic potential of Cxs-based hemichannels and channels. Therefore, this review describes the role of Cxs in neuroinflammation and epilepsy and examines how the blocking of channels and hemichannels may be therapeutic targets of anti-convulsive and anti-epileptic treatments.

The goal of the present review is to clarify the importance of connexins (Cxs) in Cxs-based channels (gap junctions/electric synapses) and hemichannels (connexons), in neuroinflammation, and their relationship with seizures and epilepsy. This review is focused on classic and up-to-date studies of seizure and epilepsy models, to which neuroinflammation and Cxs-based channels and hemichannels are relevant. The first part of this review takes into account information about the molecular and cellular characteristics of Cxs-based channels/hemichannels, as well as their principal functions in neurons and glial cells. The second part describes their relationship with neuroinflammation. The third part discusses how some blockers of channels and hemichannels may be therapeutic targets of anti-convulsive and anti-epileptic treatments, in different models of seizures and epilepsy.

1. Molecular and Cellular Characteristics of Connexins-Based Hemichannels and Channels

Connexins (Cxs) are a family of 21 protein isoforms, and some studies have shown that some of these structural proteins in Cx-based channels and hemichannels have an average half-life of 2–3 h [1,2]. Of these isoforms, eleven of them are expressed in the central nervous system (CNS), and are found in neurons and glia. The principal Cxs associated with neurons are Cx45 and Cx36 [3]. In astrocytes, Cx43 is the most important Cx, but Cx26 and Cx30 are also expressed [3,4,5]; Cx29, Cx32, Cx45, and Cx47 are expressed in oligodendrocytes [6,7], and Cx32, Cx36, and Cx43 are expressed in microglia [8,9,10].
Homomeric connexons or hemichannels are formed by six of the same Cx isoforms, and with different Cx isoforms, form a heteromeric connexon or hemichannel; two hemichannels in two neighboring cells form a gap junction channel with an aqueous pore diameter of 1 nm [11], and charged surface walls that depend on the Cx type [12]. These channels, through which cells are electrically and chemically coupled, can be formed from homomeric or heteromeric hemichannels, and are therefore called homotypic or heterotypic channels, respectively. The conductance of these Cxs-based channels and hemichannels is regulated by calcium concentration [13,14], intracellular pH [15,16,17,18], some neurotransmitters, such as serotonin and dopamine [19,20], the trans-junctional voltage (relative voltage difference between coupled cells), and the membrane voltage [21,22]. The conductance is related to the type of Cxs that make up the channels and hemichannels; for example, Cx36 has a conductance of 10–15 pS, while Cx45 and Cx43 have a conductance of 27.84 ± 0.25 pS [23,24,25] and 36.8 ± 0.54 pS [25], respectively.
Experimental evidence has demonstrated the presence of Cxs-based channels and hemichannels in important brain regions susceptible to seizure generation, such as the hippocampus, amygdala, and neocortex [3,26,27,28,29,30]. Cxs-based channels are localized between the axons of principal neurons [31], between interneurons with a principally dendrodendritic localization [26], and in neurons with mixed chemical/electrical synapses [32,33]. The principal function of Cxs-based hemichannels is the release of gliotransmitters via a Ca2+-dependent mechanism [13], while channels participate in the propagation of Ca2+ waves, with long-range coupling of the astrocytic network [5]; in neurons, they participate in the synchronization and generation of oscillations in the gamma range (through interneurons dendrites) and ripple frequency (through pyramidal cell axons) [30,31].

2. Connexins-Based Hemichannels/Channels and Neuroinflammation

Neuroinflammation is the response to injury, the normal aging process, dementia, stroke, hypertension, depression, diabetes, tumors, infections, toxins, drugs, acute trauma, or neurodegenerative disease that involves the coordination of inflammatory cells, and biochemical activities that occur in the CNS [34,35,36,37]. Events associated with neuroinflammation include the activation of astrocytes and microglia, the participation of oligodendrocytes and other nervous system cells, the release of cytokines and chemokines, increased levels of prostaglandins, the infiltration of cells from the bloodstream, and the generation of reactive oxygen (ROS) and nitrogen species (NOS) (Figure 1) [38,39,40].
Neuroinflammation is regulated by components of the immune system, and by pathogen recognition receptors (PRRs) [41]. Similarly, the entry of calcium into the cell through the activity of channels and hemichannels positively regulates neuroinflammation [42], which is a common consequence of epileptic seizures, and the pathogenesis of some types of acquired and genetic epilepsy [43].
The binding of pathogen-associated molecular patterns (PAMPs) or damage-associated molecular patterns (DAMPs) to PRRs causes the activation of signaling pathways that result in neuroinflammation [44]. Some DAMPS include ATP, interleukin (IL)-1β, IL6, IL11, IL15, IL17, TNFα, uric acid, and high-mobility protein group 1 (HMGB1) [45,46]. In this sense, Cxs plays an important role in the generation and maintenance of neuroinflammation, because hemichannels have little selectivity, and their opening is considered harmful since neurotoxic damage can be caused by the induced release of gliotransmitters, ATP, glutamate, and other molecules (Figure 1) [47].
Hemichannels formed by Cxs are activated by the presence of proinflammatory cytokines (TNFα and IL1β) and ROS [48] through increases in the concentrations of intracellular Ca2+ and extracellular K+ and ATP, as well as decreases in the extracellular concentration of Ca2+ and the redox potential [49,50]. Then, under stressful conditions, Cxs-based hemichannels amplify the damage and induce neuroinflammation [51]. However, once inflammation is established, positive feedback that promotes increased opening of Cxs-based hemichannels is produced [52]. The opening of Cxs-based channels and hemichannels due to inflammation begins a positive feedback cycle that, among other things, results in excessive ATP release [53].
It is recognized that inflammation is activated by two signals. Of these signals, one is mediated by the activation of PRRs through PAMPs or DAMPs that result in the induction of the NF-κB pathway, which promotes the expression of pro-IL-1β, and genes associated with inflammatory proteins [54,55]. The second signal is attributed to ATP [56,57,58] and glutamate via N-methyl-D-aspartate (NMDA) receptors [59,60]. For example, it has been observed that, in astrocytes, the activation of exogenous ATP-induced inflammation triggers an increase in IL-1β production [61].
Under neurodegenerative conditions, neuroinflammation involves the primordial participation of CNS cells, including microglia, oligodendrocytes, and astrocytes [62,63]. Astrocytes were first described by Dieters in 1865, and named by von Lenhossék in 1895, they are currently known as the most abundant cell type in the CNS [64,65]. Astrocytes participate in a large number of functions, including those related to pro- and anti-inflammatory processes, that either contribute to inflammation [66,67,68] or respond to it [69], and in epilepsy [70]. The expression of Cx43, and in lower quantities Cx26 and Cx30, of channels and hemichannels in astrocytes is relevant to neuroinflammatory processes because it allows both the exchange of small molecules [71], and gliotransmission [72,73,74].
Neuroinflammation induces the activation of two types of reactive astrocytes, called A1 and A2, as determined by transcriptome analysis. It has been suggested that the activation of A1 by NF-κB [75,76], but not the activation of A2 has reinforcing effects on inflammation. Additionally, reactive astrocytes increase the expression of genes responsible for the formation of synapses that can, nevertheless, lead to epilepsy [77]. Another relevant feature of astrocytes is their high degree of interconnection through Cxs-based channels, which create large functional syncytial networks, that are electrically and metabolically coupled, through which network Ca2+ waves can propagate excitability [78,79].
Oligodendrocytes are small myelin-forming cells that are widely distributed in the CNS, and based on morphology, can be divided into four subtypes [80]. They mainly contribute to conduction inside neuronal circuits, but they also participate in metabolic supply and ion buffering [81,82]. Oligodendrocytes mainly express Cx29, Cx32, and Cx47; through these connections, they form channels with astrocytes and with themselves (heterotypic unions of Cx30 and Cx32, and of Cx43 and Cx47) [7,83]. These functional unions of Cxs even participate in the release of gliotransmitters, although it is not clear if oligodendrocytes have the necessary molecular machinery to fulfill this task [84]. It has also been suggested that oligodendrocytes may be associated with acquired neuroinflammation [78], but more robust information, as well as increased information on the involvement of oligodendrocytes in neuroinflammation, is needed.
Microglia are tissue-resident innate immune cells of the CNS that constitute approximately 5–10% of all CNS cells [85]. The resting state or non-activated state of microglia [86] may result from signals, such as transforming growth factor-β (TGFβ), an anti-inflammatory cytokine that is constitutively expressed in the CNS in the non-inflamed brain [87]. Microglia, particularly those that focus on debris clearing and the elimination of dead cells and synapses are essential in the maturation of neuronal circuits [88], the development of synaptic plasticity [89], and CNS homeostasis [90,91]. In the resting state, microglia exhibit their own Cxs profile, which involves the expression of Cx32 [92] and Cx36 [93,94]; however, Cx43 is not expressed by microglia [95,96,97].
Under neurological conditions such as epilepsy, microglia are rapidly activated, release proinflammatory cytokines, take on an amoeboid morphology, express a large number of surface receptors, and can detect DAMPs, which are recognized by PRRs [98,99,100,101], all of these can lead to neuronal hyperexcitability and neurodegeneration. Several studies have shown an increase in the expression of Cx32 [102], Cx36 [103], and Cx43 in microglia, which has been correlated with the development of chronic degenerative diseases. Additionally, it has been shown that, during the establishment of neurological damage, such as seizures and epilepsy, ATP is released mainly by the Cxs of astrocytes, and the pannexins of oligodendrocytes [104,105]. In vivo and in vitro studies have shown that the excessive release of ATP and glutamate is toxic to the CNS [104,105,106]; subsequently, ATP is recognized by P2x7 microglial receptors, and their activation causes potassium outflow, which is a critical event for the activation of the NLRP3 inflammasome. A central framework receptor, called nucleotide-binding oligomerization domain-like receptor family pyrin domain containing 3 (NLRP3) that initiates signaling, and three key protein organization-based proteins, results in the recruitment of procaspases and caspases (Figure 1) [107]. The assembly of the NLRP3 inflammasome results in the conversion of procaspase-1 from its pro form to its active form, which then processes the cleavage of several substrates, such as pro-IL-1β and pro-IL-18, in the mature cytokines IL-1β and IL-18 [108]; this process can aggravate inflammation. In addition, numerous studies have implicated neuroinflammation as the cause, and result, of epilepsy [109,110].

3. Blockers of Connexins-Based Channels/Hemichannels as Anticonvulsive and Antiepileptic Therapeutic Targets in Different Models of Seizures and Epilepsy

Metabolic, mechanical, or genetic insults, as well as an imbalance between γ-Aminobutyric Acid (GABA) and glutamate, can generate seizures through the excessive release of glutamate [111], and the overactivation of their receptors in neurons and astrocytes; this results in an increase in intercellular calcium concentrations in both types of cells [112,113,114]. However, the effect of this activation and association with Cxs-based channels and hemichannels can establish prolonged seizures and epilepsy through different paths in neurons and glial cells (Figure 1).
In neurons, it can drive hyperexcitability and synchronicity as a consequence of intense neuronal firing between pyramidal cells of the hippocampus that is sustained by the uncontrolled opening of axo-axonal Cxs-based channels, induced by pH changes (alkalinization) [31,115,116,117], or be as observed in dendritic Cxs-based channels with strong olivary coupling, after the activation of NMDA receptors in inferior olive neurons [118,119]. Additionally, these observations support the chemical modulation of Cxs-based channels and hemichannels by neurotransmitters, such as glutamate or even serotonin (5-HT) [19]. The pathological alkaline-acidosis environment generated during and after seizures facilitate the conductance of Cxs-based channels that strengthens neuronal coupling in networks, which induces increased synchronization and excitability (Figure 1) [120,121,122,123,124].
These facts can be assessed by in vitro and in vivo pharmacological experiments, in which the blockade of Cxs-based channels or hemichannels, has been shown to have anti-convulsive effects (Table 1). These studies used non-specific blockers such as carbenoxolone, quinine, and mefloquine, and the combination of these drugs with classic anti-convulsants [125,126,127,128,129,130,131,132,133]; carbenoxolone blocks Cxs indistinctly via the phosphorylation or internalization of Cxs subunits [134,135,136,137], quinine blocks not only Cx36- and Cx50-based channels, but also ion channels associated with the electric membrane properties of neurons [134,135,136,137,138,139,140,141], and mefloquine has effects on different neurotransmitter systems, acting as a GABA A antagonist or inhibiting 5-HT3 receptors [142]; it induces short action potentials and has effects on l-type calcium channels [143], in some cases increasing spontaneous excitatory currents [144].
However, studies in which more effective drugs, including mimetic peptides such as Gap 26 (sequence: VCYDKSFPISHVR) and Gap 27 (sequence: SRPTEKTIFI), were used to block Cxs-based channels or hemichannels showed similar results [145,146,147]; in this respect, Gap 27, a specific peptide to the extracellular loop domain of Cx43, Cx40, and Cx37, decreased recurrent epileptiform activity in hippocampal slices [146], and prevented neuronal death caused by seizures induced by the GABA-A receptor antagonist bicuculline [147].
The pathological cycle of hyperexcitability and synchronization can trigger epileptogenesis, and Cxs-based channels and hemichannels play an important role in this process, since some studies have shown their participation in fast ripples, a high frequency oscillation that is considered an electrophysiological marker of epileptogenicity [148,149]. Several studies have demonstrated that fast ripples are decreased by blockers of Cxs-based channels and hemichannels in the hippocampus of animals with pilocarpine-induced epilepsy [150,151], as well as in vitro and in silico [152]. These facts indicate their importance and how hypersynchronization and excitability of neurons synergistically with neuroinflammation maintain pathological conditions in networks that address cerebral damage and cell loss, which favor epileptogenesis.
In glial cells (astrocytes and microglia), after the activation of glutamate receptors (NMDA, AMPA and metabotropic glutamate receptors), gliotransmitters such as ATP, glutamate, Nicotinamide Adenine Dinucleotide (NAD), and d-serine are released through Cxs-based hemichannels, which sustain an excitatory extracellular environment through the activation of extrasynaptic NMDA receptors in neurons of the hippocampus (CA1), and together with astrocytic calcium waves in networks, propagate neuronal activity, increasing the excitability and synchronicity of surrounding neurons (Figure 1), as observed in various pharmacologic in vitro studies [146,153,154,155,156]. Additionally, as described previously, the release of gliotransmitters can modulate the inflammatory response, activate immunological cells, and synergistically trigger irreversible neuronal damage [157]. Moreover, Cxs-based hemichannels can support epileptic activity by supplying glucose to epileptic neuronal networks [158].
There are specific pharmacological blockers of Cxs-based hemichannels, such as peptide 5 (with a short incubation time or at a low concentration), Gap 19, and La3+, that have been used in a few studies to support the principal role of hemichannels in seizures, and to demonstrate their anti-convulsive effects in in vitro models [159,160], and in some animal models of seizures [160]. In this last study, the suppression of seizures was associated with a decrease in d-serine concentration. Despite the importance of Cxs-based hemichannels in seizures and epilepsy (Table 1), there have been few pharmacological studies that have demonstrated the use of these drugs as a new therapeutic strategy against seizures and epilepsy. Accordingly, it is necessary to continue researching this topic experimentally, to further understand their potential as anti-convulsive and anti-epileptic treatments.
Table 1. Anticonvulsive effects of principal blockers of Cxs-based hemichannels and channels in different in vitro and in vivo models of seizures and epilepsy.
Table 1. Anticonvulsive effects of principal blockers of Cxs-based hemichannels and channels in different in vitro and in vivo models of seizures and epilepsy.
Blocker(s) of Cxs-Based Hemichannels or ChannelsSeizure or Epilepsy ModelTechnique/Brain RegionMain ResultsCitation
Carbenoxolone (10 μM) and quinine (35 μM) administered through a piece of filter paper covering the cortical surfaceIn vivo: local application of crystalline 4-AP on the surface of the cortexElectrocorticography (ECoG) in the brains of adult Wistar rats (male and female, 30–40 days old, 200–250 g)Anticonvulsive effect of carbenoxolone (reduced the generation of seizure discharges); quinine decreased summated ictal activity and the amplitudes of seizure discharges [125]
Carbenoxolone (150 mM) and meclofenamic acid (50 mM) administered through a cannula implanted in the right motor cortexIn vivo: a model of refractory focal cortical epilepsy induced with tetanus toxin (50 ng/0.5 μL, pH 7.5) in 2% bovine serum albuminIntracranial electroencephalography (iEEG) in the right motor cortex of adult Sprague-Dawley rats (240–320 g)Reduced the percentage of seizure time[126]
Quinine (200, 400 or 1000 nmol) administered through a cannula implanted in the ventricle of the brainIn vivo: a model of epilepsy induced by 300 IU of crystallized penicillinEpidural EEG in adult Wistar rats (male, 4 months)Decreased the amplitude and frequency of epileptiform spikes and attenuated convulsive behavior[127]
Carbenoxolone injection (50 nmol) administered through a cannula implanted in the entorhinal cortexIn vivo: a model of seizures induced by 4-aminopyridine (10 nmol) administered through a cannula implanted in the entorhinal cortexEpidural EEG and iEEG in the entorhinal cortex of adult Wistar rats (male, 250–350 g)Decreased the amplitude and frequency of epileptiform discharges and the number and duration of epileptiform trains[128]
Carbenoxolone, Gap 27 (mimetic amino acid residues 201–211, SRPTEKTIFII) and SLS (amino acid residues 180–195, SLSAVYTCKRDPCPHQ) peptidesIn vitro: epileptiform activity induced in organotypic hippocampal slice cultures by stimulationExtracellular recordings from the CA1 and CA3 regions of hippocampal slices from 7-day-old Wistar ratsCarbenoxolone inhibited both spontaneous and evoked seizure-like events; the Cx43 mimetic peptides selectively attenuated spontaneous recurrent epileptiform activity after prolonged (10 h) treatment[146]
Quinine injection (35 pmol) administered through a cannula implanted in the entorhinal cortexIn vivo: a model of seizures induced by 4-aminopyridine (10 nmol) administered through a cannula implanted in the entorhinal cortexEpidural EEG and iEEG in the entorhinal cortex of adult Wistar rats (male, 250–350 g)Decreased the amplitude and frequency of discharge trains and blocked seizure behavior in five of six rats[129]
Cx43 mimetic peptide (5 and 50 μM, sufficient to block hemichannels, VDCFLSRPTEKT, extracellular loop two of Cx43)In vitro: a model of epileptiform injury induced by bicuculline methochloride (BMC) (48 h exposure to 100 μM) in hippocampal slices cultures from 6- to 8-day-old Wistar rats Measurement of cell death after epileptiform activity (fluorescence signal) and immunohistochemistry for microtubule-associated protein (MAP2)Exerted a protective effect in the CA1 region during the recovery period (24 h after BMC treatment) [147]
Carbenoxolone (20 mg/kg, i.p.) once a day for 14 daysIn vivo: a model of posttraumatic epilepsy induced by ferric ions (microinjection of 10 μL of 0.1 M FeCl3 solution into the sensorimotor area)Evaluation of convulsive behavior according to the Racine scale in adult male Sprague-Dawley rats aged 6–8 weeks and weighing 220–250 gAmeliorated convulsive behavior score in rats[130]
Carbenoxolone (50 nmol) and quinine (35 pmol) administered through a guide cannula in the entorhinal cortex (0.2 μL/min for 5 min) In vivo: a pilocarpine-induced model of temporal lobe epilepsy (1.2 mg/μL pilocarpine hydrochloride in a total volume of 2 μL, intracerebroventricular (i.c.v.)iEEG in the hippocampus of epileptic adult Wistar rats (male, 190–200 g)Decreased the number of Fast Ripples (FR) events and oscillation cycles per FR event[150]
Carbenoxolone (0.2 mM)In vitro: Neocortical slicesNeocortical slices from epileptic patients (temporal and occipital regions)Strongly decreased the incidence of FR events[152]
In silico: a small network of 256 multicompartment cellsSimulated networks containing only pyramidal cells, coupled only by axonal gap junctions, and without chemical synapses or interneurons The network produced FR events via a cluster of axonal Cx-based channels (gap junctions)
Carbenoxolone (40 mg/kg, i.p.) and carbenoxolone + valproic acid (300 mg/kg, i.p.)In vivo: a kindling model of epilepsy induced by pentylenetetrazole (35 mg/kg, i.p.)Epidural EEG in Wistar rats (female, 12–15 weeks old, 200 ± 50 g)Carbenoxolone prevented generalized seizures and reduced seizure stage, seizure duration and spike frequency; no significant difference between carbenoxolone + valproic acid and valproic acid[131]
Carbenoxolone (50 mg/kg, i.p., for 3 days) and quinine (50 mg/kg, i.p. for 3 days)In vivo: a lithium/pilocarpine-induced Status epilepticus (SE) model (i.p. injection of 50 mg/kg pilocarpine 18–20 h after the i.p. injection of 127 mg/kg lithium chloride)iEEG in the hippocampus of adult Sprague-Dawley rats (male)Reduced the spectral power of FR events 10 min after SE[151]
Coadministration of valproate (VPA), phenytoin (PHT), or carbamazepine (CBZ) at subtherapeutic doses (i.p.) with carbenoxolone (60 mg/kg, i.p., 5 mL/kg) or quinine (40 mg/kg, i.p., 5 mL/kg)In vivo: maximal electroshock (MES)-induced (frequency of 60 Hz, pulse width of 0.6 ms, shock duration of 0.6 s, and a current of 90 mA) and pentylenetetrazole (PTZ)-induced (70 mg/kg, i.p.) models of seizures EEG and power spectral analysis in Wistar rats (male, 270–300 g)Quinine increased the anticonvulsant activity of VPA, PHT and CBZ to generalized tonic-clonic seizures in the MES-induced model and the anticonvulsant activity of CBZ only to generalized tonic-clonic seizures in the PTZ-induced model [132]
TAT-Gap 19 (200 mM for in vitro experiments; 12 mM intrahippocampal and 1 mM in a total volume of 1 μL i.c.v. for in vivo experiments);
TAT-Gap19 (25 or 50 mg/kg i.p., electrical stimulation for in vivo experiments)
In vitro: pilocarpine (15 μM) administration in acute brain slices;
In vivo: pilocarpine model in mice and rats (12 mM, intra hippocampal)
Limbic psychomotor seizures by corneal stimulation
In vitro: ethidium bromide uptake experiments in acute brain slices from Glial Fibrillary Acidic Protein-enhanced Green Fluorescent Protein (GFAP-eGFP) transgenic mice (both genders, 2 months old);
In vivo: video-EcoG analysis (seizure duration) or modified Racine’s scale evaluation (to score kindling-induced behavioral changes) in NMRI mice (male, 20–30 g) and, video-EcoG and Racine scale evaluation of convulsive behavior in Wistar rats (male, 250–300 g)
In vitro: dye uptake experiments demonstrated that astroglial Cx43 hemichannels open in response to pilocarpine, and this was inhibited by TAT-Gap19.
In vivo: TAT-Gap19 suppressed seizures and decreased D-serine concentrations; these effects were reversed by exogenous D-serine administration, and a similar effect was observed for the electrical stimulation model
[160]
In vitro experiments: Carbenoxolone (200 mM) quinine (100 mM) and La(NO3)3 (a blocker of Cx-based hemichannels)
In vivo experiments: carbenoxolone (100 mg/kg) and quinine (40 mg/kg)
In vitro: low-Mg2+-induced epilepsy model
In vivo: Wistar Albino Glaxo/ Rijswijk (WAG/Rij) rat model of absence epilepsy
Field potential recordings, evaluation of seizure-like events (SLEs) in hippocampal-entorhinal slices from Wistar rats (11–14 days old); epidural EEG (frontal and parietal cortex) in WAG/Rij rats (female, 11–12 months old, 195–210 g); recordings of bilaterally synchronous spike-wave discharges (SWDs)Carbenoxolone prevented the occurrence of SLEs and aggravated seizures in non-convulsive absence epilepsy; quinine did not prevent SLEs but increased the number and total time of SWDs and decreased the length of the interictal intervals; La3+ completely abolished SLEs[159]

4. Conclusions

There is evidence regarding the relationship between Cxs-based channels and hemichannels on seizures and epilepsy in different models, in which the uncontrolled opening of these channels by pH changes as a consequence of intense neuronal firing sustains inflammation, hyperexcitability, and synchronicity that maintain this pathological condition. The use of some blockers of Cxs-based channels and hemichannels may be therapeutic agents for the treatment of seizures and epilepsy.

Author Contributions

Conceptualization, L.M.-C.; investigation, L.M.-C., J.C.S.-S., J.O.-I.; writing—original draft preparation, L.M.-C., J.C.S.-S., J.O.-I., A.M.-V.; writing—review and editing, L.M.-C., A.M.-V.

Funding

This research was funded by CONACYT-BASIC SCIENCE, grant number 250930.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Flores, C.E.; Nannapaneni, S.; Davidson, K.G.; Yasumura, T.; Bennett, M.V.; Rash, J.E.; Pereda, A.E. Trafficking of gap junction channels at a vertebrate electrical synapse in vivo. Proc. Natl. Acad. Sci. USA 2012, 109, E573–E582. [Google Scholar] [CrossRef] [PubMed]
  2. Pereda, A.E.; Curti, S.; Hoge, G.; Cachope, R.; Flores, C.E.; Rash, J.E. Gap junction-mediated electrical transmission: Regulatory mechanisms and plasticity. Biochim. Biophys. Acta 2013, 1828, 134–146. [Google Scholar] [CrossRef] [PubMed]
  3. Zancan, M.; Malysz, T.; Moura, D.J.; Morás, A.M.; Steffens, L.; Rasia-Filho, A.A. Gap junctions and expression of Cx36, Cx43 and Cx45 in the posterodorsal medial amygdala of adult rats. Histol. Histopathol. 2019, 9, 18160. [Google Scholar]
  4. Condorelli, D.F.; Mudò, G.; Trovato-Salinaro, A.; Mirone, M.B.; Amato, G.; Belluardo, N. Connexin-30 mRNA is up-regulated in astrocytes and expressed in apoptotic neuronal cells of rat brain following kainate-induced seizures. Mol. Cell. Neurosci. 2002, 21, 94–113. [Google Scholar] [CrossRef] [PubMed]
  5. Ghézali, G.; Vasile, F.; Curry, N.; Fantham, M.; Cheung, G.; Ezan, P.; Cohen-Salmon, M.; Kaminski, C.; Rouach, N. Neuronal Activity Drives Astroglial Connexin 30 in Perisynaptic Processes and Shapes Its Functions. Cereb. Cortex 2019. [Google Scholar] [CrossRef]
  6. Dermietzel, R.; Farooq, M.; Kessler, J.A.; Althaus, H.; Hertzberg, E.L.; Spray, D.C. Oligodendrocytes express gap junction proteins connexin32 and connexin45. Glia 1997, 20, 101–114. [Google Scholar] [CrossRef]
  7. Nagy, J.I.; Ionescu, A.V.; Lynn, B.D.; Rash, J.E. Connexin29 and connexin32 at oligodendrocyte and astrocyte gap junctions and in myelin of the mouse central nervous system. J. Comp. Neurol. 2003, 464, 356–370. [Google Scholar] [CrossRef]
  8. Dobrenis, K.; Chang, H.Y.; Pina-Benabou, M.H.; Woodroffe, A.; Lee, S.C.; Rozental, R.; Spray, D.C.; Scemes, E. Human and mouse microglia express connexin36, and functional gap junctions are formed between rodent microglia and neurons. J. Neurosci. Res. 2005, 82, 306–315. [Google Scholar] [CrossRef]
  9. Gajardo-Gómez, R.; Labra, V.C.; Orellana, J.A. Connexins and Pannexins: New Insights into Microglial Functions and Dysfunctions. Front. Mol. Neurosci. 2016, 9, 86. [Google Scholar] [CrossRef]
  10. Shaikh, S.B.; Uy, B.; Perera, A.; Nicholson, L.F. AGEs-RAGE mediated up-regulation of connexin43 in activated human microglial CHME-5 cells. Neurochem. Int. 2012, 60, 640–651. [Google Scholar] [CrossRef]
  11. Veenstra, R.D.; Wang, H.Z.; Beyer, E.C.; Brink, P.R. Selective dye and ionic permeability of gap junction channels formed by connexin 45. Circ. Res 1994, 75, 483–490. [Google Scholar] [CrossRef]
  12. Bukauskas, F.F. Neurons and β-cells of the pancreas express connexin36, forming gap junction channels that exhibit strong cationic selectivity. J. Membr. Biol. 2012, 245, 243–253. [Google Scholar] [CrossRef]
  13. Abudara, V.; Retamal, M.A.; Del Rio, R.; Orellana, J.A. Synaptic Functions of Hemichannels and Pannexons: A Double-Edged Sword. Front. Mol. Neurosci. 2018, 11, 435. [Google Scholar] [CrossRef]
  14. Spray, D.C.; White, R.L.; Mazet, F.; Bennett, M.V. Regulation of gap junctional conductance. Am. J. Phys. 1985, 248, H753–H764. [Google Scholar] [CrossRef]
  15. Ek Vitorín, J.F.; Pontifex, T.K.; Burt, J.M. Determinants of Cx43 Channel Gating and Permeation: The Amino Terminus. Biophys. J. 2016, 110, 127–140. [Google Scholar] [CrossRef]
  16. Perez-Velazquez, J.L.; Valiante, T.A.; Carlen, P.L. Modulation of gap junctional mechanisms during calcium-free induced field burst activity: A possible role for electrotonic coupling in epileptogenesis. J. Neurosci. 1994, 14, 4308–4317. [Google Scholar] [CrossRef]
  17. González-Nieto, D.; Gómez-Hernández, J.M.; Larrosa, B.; Gutiérrez, C.; Muñoz, M.D.; Fasciani, I.; O’Brien, J.O.; Zappalà, A.; Cicirata, F.; Barrio, L.C. Regulation of neuronal connexin-36 channels by pH. Proc. Natl. Acad. Sci. USA 2008, 105, 17169–17174. [Google Scholar] [CrossRef]
  18. Rimkute, L.; Kraujalis, T.; Snipas, M.; Palacios-Prado, N.; Jotautis, V.; Skeberdis, V.A.; Bukauskas, F.F. Modulation of Connexin-36 Gap Junction Channels by Intracellular pH and Magnesium Ions. Front. Physiol. 2018, 9, 362. [Google Scholar] [CrossRef]
  19. Szabo, T.M.; Caplan, J.S.; Zoran, M.J. Serotonin regulates electrical coupling via modulation of extrajunctional conductance: H-current. Brain Res. 2010, 1349, 21–31. [Google Scholar] [CrossRef]
  20. Vivar, C.; Traub, R.D.; Gutiérrez, R. Mixed electrical-chemical transmission between hippocampal mossy fibers and pyramidal cells. Eur. J. Neurosci. 2012, 35, 76–82. [Google Scholar] [CrossRef]
  21. Fasciani, I.; Pluta, P.; González-Nieto, D.; Martínez-Montero, P.; Molano, J.; Paíno, C.L.; Millet, O.; Barrio, L.C. Directional coupling of oligodendrocyte connexin-47 and astrocyte connexin-43 gap junctions. Glia 2018, 66, 2340–2352. [Google Scholar] [CrossRef]
  22. Bargiello, T.A.; Oh, S.; Tang, Q.; Bargiello, N.K.; Dowd, T.L.; Kwon, T. Gating of Connexin Channels by transjunctional-voltage: Conformations and models of open and closed states. Biochim. Biophys. Acta Biomembr. 2018, 1860, 22–39. [Google Scholar] [CrossRef]
  23. Teubner, B.; Degen, J.; Söhl, G.; Guldenagel, M.; Bukauskas, F.E.; Trexler, E.B.; Verselis, V.K.; De Zeeuw, C.I.; Lee, C.G.; Kozak, C.A.; et al. Functional expression of the murine connexin 36 gene coding for a neuron-specific gap junctional protein. J. Membr. Biol. 2000, 176, 249–262. [Google Scholar] [CrossRef]
  24. Srinivas, M.; Rozental, R.; Kojima, T.; Dermietzel, R.; Mehler, M.; Condorelli, D.F.; Kessler, J.A.; Spray, D.C. Functional properties of channels formed by the neuronal gap junction protein connexin 36. J. Neurosci. 1999, 19, 9848–9855. [Google Scholar] [CrossRef]
  25. Zhong, G.; Akoum, N.; Appadurai, D.A.; Hayrapetyan, V.; Ahmed, O.; Martinez, A.D.; Beyer, E.C.; Moreno, A.P. Mono-Heteromeric Configurations of Gap Junction Channels Formed by Connexin43 and Connexin45 Reduce Unitary Conductance and Determine both Voltage Gating and Metabolic Flux Asymmetry. Front. Physiol. 2017, 8, 346. [Google Scholar] [CrossRef]
  26. Fukuda, T.; Kosaka, T. Gap junctions linking the dendritic network of GABAergic interneurons in the hippocampus. J. Neurosci. 2000, 20, 1519–1528. [Google Scholar] [CrossRef]
  27. Baude, A.; Bleasdale, C.; Dalezios, Y.; Somogyi, P.; Klausberger, T. Immunoreactivity for the GABAA receptor α1 subunit, somatostatin and connexin36 distinguishes axoaxonic, basket, and bistratified interneurons of the rat hippocampus. Cereb. Cortex 2007, 17, 2094–2107. [Google Scholar] [CrossRef]
  28. Zsiros, V.; Maccaferri, G. Noradrenergic modulation of electrical coupling in GABAergic networks of the hippocampus. J. Neurosci. 2008, 28, 1804–1815. [Google Scholar] [CrossRef]
  29. Gibson, J.R.; Beierlein, M.; Connors, B.W. Functional properties of electrical synapses between inhibitory interneurons of neocortical layer 4. J. Neurophysiol. 2005, 93, 467–480. [Google Scholar] [CrossRef]
  30. Mancilla, J.G.; Lewis, T.J.; Pinto, D.J.; Rinzel, J.; Connors, B.W. Synchronization of electrically coupled pairs of inhibitory interneurons in neocortex. J. Neurosci. 2007, 27, 2058–2073. [Google Scholar] [CrossRef]
  31. Traub, R.D.; Whittington, M.A.; Buhl, E.H.; LeBeau, F.E.; Bibbig, A.; Boyd, S.; Cross, H.; Baldeweg, T. A possible role for gap junctions in generation of very fast EEG oscillations preceding the onset of, and perhaps initiating, seizures. Epilepsia 2001, 42, 153–170. [Google Scholar]
  32. Nagy, J.I. Evidence for connexin36 localization at hippocampal mossy fiber terminals suggesting mixed chemical/electrical transmission by granule cells. Brain Res. 2012, 1487, 107–122. [Google Scholar] [CrossRef]
  33. Hamzei-Sichani, F.; Davidson, K.G.V.; Yasumura, T.; Janssen, W.G.M.; Wearne, S.L.; Hof, P.R.; Traub, R.D.; Gutiérrez, R.; Ottersen, O.P.; Rash, J.E. Mixed electrical–chemical synapses in adult rat hippocampus are primarily glutamatergic and coupled by connexin-36. Front. Neuroanat. 2012, 6, 13. [Google Scholar] [CrossRef] [Green Version]
  34. Barrientos, R.M.; Kitt, M.M.; Watkins, L.R.; Maier, S.F. Neuroinflammation in the normal aging hippocampus. Neuroscience 2015, 309, 84–99. [Google Scholar] [CrossRef] [Green Version]
  35. Frank, M.G.; Weber, M.D.; Watkins, L.R.; Maier, S.F. Stress sounds the alarmin: The role of the danger-associated molecular pattern HMGB1 in stress-induced neuroinflammatory priming. Brain Behav. Immun. 2015, 48, 1–7. [Google Scholar] [CrossRef] [Green Version]
  36. Sofroniew, M.V. Astrocyte barriers to neurotoxic inflammation. Nat. Rev. Neurosci. 2015, 16, 249–263. [Google Scholar] [CrossRef] [Green Version]
  37. Kempuraj, D.; Thangavel, R.; Natteru, PA.; Selvakumar, GP.; Saeed, D.; Zahoor, H.; Zaheer, S.; Iyer, S.S.; Zaheer, A. Neuroinflammation Induces Neurodegeneration. J. Neurol. Neurosurg. Spine 2016, 1, 1003. [Google Scholar]
  38. Reddy, D.S.; Kuruba, R. Experimental Models of Status Epilepticus and Neuronal Injury for Evaluation of Therapeutic Interventions. Int. J. Mol. Sci. 2013, 14, 18284–18318. [Google Scholar] [CrossRef]
  39. Ransohoff, R.M.; Schafer, D.; Vincent, A.; Blachère, N.E.; Bar-Or, A. Neuroinflammation: Ways in Which the Immune System Affects the Brain. Neurotherapeutics 2015, 12, 896–909. [Google Scholar] [CrossRef] [Green Version]
  40. Shabab, T.; Khanabdali, R.; Moghadamtousi, S.Z.; Kadir, H.A.; Mohan, G. Neuroinflammation pathways: A general review. Int. J. Neurosci. 2017, 127, 624–633. [Google Scholar] [CrossRef]
  41. Hagberg, H.; Mallard, C.; Ferriero, D.M.; Vannucci, S.J.; Levison, S.W.; Vexler, Z.S.; Gressens, P. The role of inflammation in perinatal brain injury. Nat. Rev. Neurol. 2015, 11, 192–208. [Google Scholar] [CrossRef] [PubMed]
  42. Shao, B.Z.; Xu, Z.Q.; Han, B.Z.; Su, D.F.; Liu, C. NLRP3 inflammasome and its inhibitors: A review. Front. Pharmacol. 2015, 6, 262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Vezzani, A.; Balosso, S.; Ravizza, T. Neuroinflammatory pathways as treatment targets and biomarkers in epilepsy. Nat. Rev. Neurol. 2019, 15, 459–472. [Google Scholar] [CrossRef] [PubMed]
  44. Liu, J.; Buisman-Pijlman, F.; Hutchinson, M.R. Toll-like receptor 4: Innate immune regulator of neuroimmune and neuroendocrine interactions in stress and major depressive disorder. Front. Neurosci. 2014, 8, 309. [Google Scholar] [CrossRef] [Green Version]
  45. Jensen, C.J.; Massie, A.; De Keyser, J. Immune players in the CNS: The astrocyte. J. Neuroimmune Pharmacol. 2013, 8, 824–839. [Google Scholar] [CrossRef]
  46. Walker, A.K.; Kavelaars, A.; Heijnen, C.J.; Dantzer, R. Neuroinflammation and comorbidity of pain and depression. Pharmacol. Rev. 2014, 66, 80–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Montero, T.D.; Orellana, J.A. Hemichannels: New pathways for gliotransmitter release. Neuroscience 2015, 286, 45–59. [Google Scholar] [CrossRef]
  48. Retamal, M.A.; Froger, N.; Palacios-Prado, N.; Ezan, P.; Sáez, P.J.; Sáez, J.C.; Giaume, C. Cx43 hemichannels and gap junction channels in astrocytes are regulated oppositely by proinflammatory cytokines released from activated microglia. J. Neurosci. 2007, 27, 13781–13792. [Google Scholar] [CrossRef]
  49. Locovei, S.; Wang, J.; Dahl, G. Activation of pannexin 1 channels by ATP through P2Y receptors and by cytoplasmic calcium. FEBS Lett. 2006, 580, 239–244. [Google Scholar] [CrossRef] [Green Version]
  50. Baroja-Mazo, A.; Barberà-Cremades, M.; Pelegrín, P. The participation of plasma membrane hemichannels to purinergic signaling. Biochim. Biophys. Acta Biomembr. 2013, 1828, 79–93. [Google Scholar] [CrossRef] [Green Version]
  51. Rovegno, M.; Saez, J.C. Role of astrocyte connexin hemichannels in cortical spreading depression. Biochim. Biophys. Acta Biomembr. 2018, 1860, 216–223. [Google Scholar] [CrossRef] [PubMed]
  52. De Vuyst, E.; Decrock, E.; De Bock, M.; Yamasaki, H.; Naus, C.C.; Evans, W.H.; Leybaert, L. Connexin hemichannels and gap junction channels are differentially influenced by lipopolysaccharide and basic fibroblast growth factor. Mol. Biol. Cell. 2007, 18, 34–46. [Google Scholar] [CrossRef] [PubMed]
  53. Mallard, C.; Davidson, J.O.; Tan, S.; Green, C.R.; Bennet, L.; Robertson, N.J.; Gunn, A.J. Astrocytes and microglia in acute cerebral injury underlying cerebral palsy associated with preterm birth. Pediatr. Res. 2014, 75, 234–240. [Google Scholar] [CrossRef] [PubMed]
  54. Hung, S.C.; Choi, C.H.; Said-Sadier, N.; Johnson, L.; Atanasova, K.R.; Sellami, H.; Yilmaz, O.; Ojcius, D.M. P2X4 assembles with P2X7 and pannexin-1 in gingival epithelial cells and modulates ATP-induced reactive oxygen species production and inflammasome activation. PLoS ONE 2013, 8, e70210. [Google Scholar] [CrossRef] [Green Version]
  55. Sutterwala, F.S.; Haasken, S.; Cassel, S.L. Mechanism of NLRP3 inflammasome activation. Ann. N. Y. Acad. Sci. 2014, 1319, 82–95. [Google Scholar] [CrossRef]
  56. Gombault, A.; Baron, L.; Couillin, I. ATP release and purinergic signaling in NLRP3 inflammasome activation. Front. Immunol. 2012, 3, 414. [Google Scholar] [CrossRef] [Green Version]
  57. Murakami, T.; Ockinger, J.; Yu, J.; Byles, V.; McColl, A.; Hofer, A.M.; Horng, T. Critical role for calcium mobilization in activation of the NLRP3 inflammasome. Proc. Natl. Acad. Sci. USA 2012, 109, 11282–11287. [Google Scholar] [CrossRef] [Green Version]
  58. Riteau, N.; Baron, L.; Villeret, B.; Guillou, N.; Savigny, F.; Ryffel, B.; Rassendren, F.; Le Bert, M.; Gombault, A.; Couillin, I. ATP release and purinergic signaling: A common pathway for particle-mediated inflammasome activation. Cell Death Dis. 2012, 3, e403. [Google Scholar] [CrossRef] [Green Version]
  59. Weber, M.D.; Frank, M.G.; Tracey, K.J.; Watkins, L.R.; Maier, S.F. Stress induces the danger-associated molecular pattern HMGB-1 in the hippocampus of male Sprague Dawley rats: A priming stimulus of microglia and the NLRP3 inflammasome. J. Neurosci. 2015, 35, 316–324. [Google Scholar] [CrossRef]
  60. Zou, J.; Crews, F. Glutamate/NMDA excitotoxicity and HMGB1/TLR4 neuroimmune toxicity converge as components of neurodegeneration. AIMS Public Health 2015, 2, 77–100. [Google Scholar] [CrossRef]
  61. Silverman, W.R.; de Rivero Vaccari, J.P.; Locovei, S.; Qiu, F.; Carlsson, S.K.; Scemes, E.; Keane, R.W.; Dahl, G. The pannexin 1 channel activates the inflammasome in neurons and astrocytes. J. Biol. Chem. 2009, 284, 18143–18151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Liddelow, S.A.; Barres, B.A. Reactive astrocytes: Production, Function, and Therapeutic Potential. Immunity 2017, 46, 957–967. [Google Scholar] [CrossRef] [Green Version]
  63. Allnoch, L.; Baumgärtner, W.; Hansmann, F. Impact of Astrocyte Depletion upon Inflammation and Demyelination in a Murine Animal Model of Multiple Sclerosis. Int. J. Mol. Sci. 2019, 20, 3922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Sofroniew, M.V.; Vinters, H.V. Astrocytes: Biology and pathology. Acta Neuropathol. 2010, 119, 7–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Ludwin, S.K.; Rao, V.; Moore, C.S.; Antel, J.P. Astrocytes in multiple sclerosis. Mult. Scler. J. 2016, 22, 1114–1124. [Google Scholar] [CrossRef]
  66. Liddelow, S.A.; Guttenplan, K.A.; Clarke, L.E.; Bennett, F.C.; Bohlen, C.J.; Schirmer, L.; Bennett, M.L.; Munch, A.E.; Chung, W.S.; Peterson, T.C. Neurotoxic reactive astrocytes are induced by activated microglia. Nature 2017, 541, 481–487. [Google Scholar] [CrossRef]
  67. Herz, J.; Filiano, A.J.; Smith, A.; Yogev, N.; Kipnis, J. Myeloid Cells in the Central Nervous System. Immunity 2017, 46, 943–956. [Google Scholar] [CrossRef] [Green Version]
  68. Klein, R.S.; Hunter, C.A. Protective and Pathological Immunity during Central Nervous System Infections. Immunity 2017, 46, 891–909. [Google Scholar] [CrossRef] [Green Version]
  69. Zamanian, J.L.; Xu, L.; Foo, L.C.; Nouri, N.; Zhou, L.; Giffard, R.G.; Barres, B.A. Genomic Analysis of Reactive Astrogliosis. J. Neurosci. 2012, 32, 6391–6410. [Google Scholar] [CrossRef] [Green Version]
  70. Heppner, F.L.; Ransohoff, R.M.; Becher, B. Immune attack: The role of inflammation in Alzheimer disease. Nat. Rev. Neurosci. 2015, 16, 358–372. [Google Scholar] [CrossRef]
  71. Sáez, J.C.; Schalper, K.A.; Retamal, M.A.; Orellana, J.A.; Shoji, K.F.; Bennett, M.V. Cell membrane permeabilization via connexin hemichannels in living and dying cells. Exp. Cell Res. 2010, 316, 2377–2389. [Google Scholar] [CrossRef] [PubMed]
  72. Bezzi, P.; Volterra, A. A neuron-glia signalling network in the active brain. Curr. Opin. Neurobiol. 2001, 11, 387–394. [Google Scholar] [CrossRef]
  73. Anderson, M.A.; Burda, J.E.; Ren, Y.; Ao, Y.; O’Shea, T.M.; Kawaguchi, R.; Coppola, G.; Khakh, B.S.; Deming, T.J.; Sofroniew, M.V. Astrocyte scar formation aids CNS axon regeneration. Nature 2016, 532, 195–200. [Google Scholar] [CrossRef] [Green Version]
  74. Kang, J.; Kang, N.; Lovatt, D.; Torres, A.; Zhao, Z.; Lin, J.; Nedergaard, M. Connexin 43 Hemichannels Are Permeable to ATP. J. Neurosci. 2008, 28, 4702–4711. [Google Scholar] [CrossRef] [PubMed]
  75. Lian, H.; Yang, L.; Cole, A.; Sun, L.; Chiang, A.C.-A.; Fowler, S.W.; Shim, D.J.; Rodriguez-Rivera, J.; Taglialatela, G.; Jankowsky, J.L.; et al. NFκB-activated Astroglial Release of Complement C3 Compromises Neuronal Morphology and Function Associated with Alzheimer’s Disease. Neuron 2015, 85, 101–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Wang, L.; Pei, S.; Han, L.; Guo, B.; Li, Y.; Duan, R.; Yao, Y.; Xue, B.; Chen, X.; Jia, Y. Mesenchymal Stem Cell-Derived Exosomes Reduce A1 Astrocytes via Downregulation of Phosphorylated NFκB P65 Subunit in Spinal Cord Injury. Cell. Physiol. Biochem. 2018, 50, 1535–1559. [Google Scholar] [CrossRef]
  77. Shen, Y.; Qin, H.; Chen, J.; Mou, L.; He, Y.; Yan, Y.; Zhou, H.; Lv, Y.; Chen, Z.; Wang, J.; et al. Postnatal activation of TLR4 in astrocytes promotes excitatory synaptogenesis in hippocampalneurons. J. Cell Biol. 2016, 215, 719–734. [Google Scholar] [CrossRef] [Green Version]
  78. Papaneophytou, C.; Georgiou, E.; Kleopa, K.A. The role of oligodendrocyte gap junctions in neuroinflammation. Channels 2019, 13, 247–263. [Google Scholar] [CrossRef]
  79. del Río-Hortega, P. Son homologables la glía de escasas radiaciones y la celuladeSchwann. Bol. Soc. Esp. Biol. 1922, 10, 25–28. [Google Scholar]
  80. Battefeld, A.; Klooster, J.; Kole, M.H. Myelinating satellite oligodendrocytes are integrated in a glial syncytium constraining neuronal high-frequency activity. Nat. Commun. 2016, 7, 11298. [Google Scholar] [CrossRef] [Green Version]
  81. Philips, T.; Rothstein, J.D. Oligodendroglia: Metabolic supporters of neurons. J. Clin. Investig. 2017, 127, 3271–3280. [Google Scholar] [CrossRef] [PubMed]
  82. Magnotti, L.M.; Goodenough, D.A.; Paul, D.L. Deletion of oligodendrocyte Cx32 and astrocyte Cx43 causeswhite matter vacuolation, astrocyte loss and early mortality. Glia 2011, 59, 1064–1074. [Google Scholar] [CrossRef] [Green Version]
  83. Vejar, S.; Oyarzún, J.E.; Retamal, M.A.; Ortiz, F.C.; Orellana, J.A. Connexin and Pannexin-Based Channels in Oligodendrocytes: Implications in Brain Health and Disease. Front. Cell. Neurosci. 2019, 13, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Georgiou, E.; Sidiropoulou, K.; Richter, J. Gene therapy targeting oligodendrocytes provides therapeutic benefit in a leukodystrophy model. Brain 2017, 140, 599–616. [Google Scholar] [CrossRef] [PubMed]
  85. Lawson, L.J.; Perry, V.H.; Dri, P.; Gordon, S. Heterogeneity in the distribution and morphology of microglia in the normal adult mouse brain. Neuroscience 1990, 39, 151–170. [Google Scholar] [CrossRef]
  86. Kettenmann, H.; Hanisch, U.K.; Noda, M.; Verkhratsky, A. Physiology of Microglia. Physiol. Rev. 2011, 91, 461–553. [Google Scholar] [CrossRef]
  87. Brionne, T.C.; Tesseur, I.; Masliah, E.; Wyss-Coray, T. Loss of TGF-beta 1 leads to increased neuronal cell death and microgliosis in mouse brain. Neuron 2003, 40, 1133–1145. [Google Scholar] [CrossRef] [Green Version]
  88. Basilico, B.; Pagani, F.; Grimaldi, A.; Cortese, B.; Di Angelantonio, S.; Weinhard, L.; Gross, C.; Limatola, C.; Maggi, L.; Ragozzino, D. Microglia shape presynaptic properties at developing glutamatergic synapses. Glia 2019, 67, 53–67. [Google Scholar] [CrossRef] [Green Version]
  89. Wu, Y.; Dissing-Olesen, L.; MacVicar, B.A.; Stevens, B. Microglia: Dynamic Mediators of Synapse Development and Plasticity. Trends Immunol. 2015, 36, 605–613. [Google Scholar] [CrossRef]
  90. Schafer, D.P.; Lehrman, E.K.; Stevens, B. “quad-partite” synapse: Microglia-synapse interactions in the developing and mature CNS. Glia 2013, 61, 24–36. [Google Scholar] [CrossRef] [Green Version]
  91. Wake, H.; Moorhouse, A.J.; Miyamoto, A.; Nabekura, J. Microglia: Actively surveying and shaping neuronal circuit structure and function. Trends Neurosci. 2013, 36, 209–217. [Google Scholar] [CrossRef] [PubMed]
  92. Ikegami, A.; Haruwaka, K.; Wake, H. Microglia: Lifelong modulator of neural circuits. Neuropathology 2019, 39, 173–180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Maezawa, I.; Jin, L.W. Rett Syndrome Microglia Damage Dendrites and Synapses by the Elevated Release of Glutamate. J. Neurosci. 2010, 30, 5346–5356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Parenti, R.; Campisi, A.; Vanella, A.; Cicirata, F. Immunocytochemical and RT-PCR analysis of connexin36 in cultures of mammalian glial cells. Arch. Ital. Biol. 2002, 140, 101–108. [Google Scholar] [PubMed]
  95. Belousov, A.B.; Nishimune, H.; Denisova, J.V.; Fontes, J.D. A potential role for neuronal connexin 36 in the pathogenesis of amyotrophic lateral sclerosis. Neurosci. Lett. 2018, 666, 1–4. [Google Scholar] [CrossRef] [PubMed]
  96. Richter, N.; Wendt, S.; Georgieva, P.B.; Hambardzumyan, D.; Nolte, C.; Kettenmann, H. Glioma-associated microglia and macrophages/monocytes display distinct electrophysiological properties and do not communicate via gap junctions. Neurosci. Lett. 2014, 583, 130–135. [Google Scholar] [CrossRef] [Green Version]
  97. Rouach, N.; Avignone, E.; Même, W.; Koulakoff, A.; Venance, L.; Blomstrand, F.; Giaume, C. Gap junctions and connexin expression in the normal and pathological central nervous system. Biol. Cell 2002, 94, 457–475. [Google Scholar] [CrossRef]
  98. Sáez, P.J.; Shoji, K.F.; Retamal, M.A.; Harcha, P.A.; Ramírez, G.; Jiang, J.X.; von Bernhardi, R.; Sáez, J.C. ATP is required and advances cytokine-induced gap junction formation in microglia in vitro. Mediat. Inflamm. 2013, 2013, 216402. [Google Scholar] [CrossRef]
  99. Gaikwad, S.; Patel, D.; Agrawal-Rajput, R. CD40 Negatively Regulates ATP-TLR4-Activated Inflammasome in Microglia. Cell. Mol. Neurobiol. 2017, 37, 351–359. [Google Scholar] [CrossRef]
  100. Bona, E.; Andersson, A.L.; Blomgren, K.; Gilland, E.; Puka-Sundvall, M.; Gustafson, K.; Hagberg, H. Chemokine and Inflammatory Cell Response to Hypoxia-Ischemia in Immature Rats. Pediatr. Res. 1999, 45, 500–509. [Google Scholar] [CrossRef] [Green Version]
  101. Benson, M.; Manzanero, J.S.; Borges, K. Complex alterations in microglial M1/M2 markers during the development of epilepsy in two mouse models. Epilepsia 2015, 56, 895–905. [Google Scholar] [CrossRef]
  102. Wolf, S.A.; Boddeke, H.W.; Kettenmann, H. Microglia in Physiology and Disease. Annu. Rev. Physiol. 2017, 79, 619–643. [Google Scholar] [CrossRef]
  103. Takeuchi, H.; Jin, S.; Wang, J.; Zhang, G.; Kawanokuchi, J.; Kuno, R.; Sonobe, Y.; Mizuno, T.; Suzumura, A. Tumor Necrosis Factor-α Induces neurotoxicity via Glutamate Release from Hemichannels of Activated Microglia in an Autocrine Manner. J. Biol. Chem. 2006, 281, 21362–21368. [Google Scholar] [CrossRef] [Green Version]
  104. Orellana, J.A.; Froger, N.; Ezan, P.; Jiang, J.X.; Bennett, M.V.; Naus, C.C.; Giaume, C.; Sáez, J.C. ATP and glutamate released via astroglial connexin 43 hemichannels mediate neuronal death through activation of pannexin 1 hemichannels. J. Neurochem. 2011, 118, 826–840. [Google Scholar] [CrossRef] [Green Version]
  105. Domercq, M.; Perez-Samartin, A.; Aparicio, D.; Alberdi, E.; Pampliega, O.; Matute, C. P2X7 receptors mediate ischemic damage to oligodendrocytes. Glia 2010, 58, 730–740. [Google Scholar] [CrossRef]
  106. Engel, T.; Alves, M.; Sheedy, C.; Henshall, D.C. ATPergic signalling during seizures and epilepsy. Neuropharmacology 2016, 104, 140–153. [Google Scholar] [CrossRef]
  107. Baron, L.; Gombault, A.; Fanny, M.; Villeret, B.; Savigny, F.; Guillou, N.; Panek, C.; Le Bert, M.; Lagente, V.; Rassendren, F.; et al. The NLRP3 inflammasome is activated by nanoparticles through ATP, ADP and adenosine. Cell Death Dis. 2015, 6, e1629. [Google Scholar] [CrossRef] [Green Version]
  108. Heneka, M.T.; McManus, R.M.; Latz, E. Inflammasome signalling in brain function and neurodegenerative disease. Nat. Rev. Neurosci. 2018, 19, 610–621. [Google Scholar] [CrossRef]
  109. Voet, S.; Srinivasan, S.; Lamkanfi, M.; van Loo, G. Inflammasomes in neuroinflammatory and neurodegenerative diseases. EMBO Mol. Med. 2019, 11, e10248. [Google Scholar] [CrossRef]
  110. Rong, S.; Wan, D.; Fan, Y.; Liu, S.; Sun, K.; Huo, J.; Zhang, P.; Li, X.; Xie, X.; Wang, F.; et al. Amentoflavone Affects Epileptogenesis and Exerts Neuroprotective Effects by Inhibiting NLRP3 Inflammasome. Front. Pharmacol. 2019, 10, 856. [Google Scholar] [CrossRef] [Green Version]
  111. Tapia, R.; Medina-Ceja, L.; Peña, F. Action of 4-aminopyridine on extracellular amino acids in hippocampus and entorhinal cortex: A dual microdialysis and electroencehalographic study in awake rats. Brain Res. Bull. 2000, 53, 255–262. [Google Scholar]
  112. Condorelli, D.F.; Conti, F.; Gallo, V.; Kirchhoff, F.; Seifert, G.; Steinhauser, C.; Verkhratsky, A.; Yuan, X. Expression and functional analysis of glutamate receptors in glial cells. Adv. Exp. Med. Biol. 1999, 468, 49–67. [Google Scholar]
  113. Seifert, G.; Steinhauser, C. Ionotropic glutamate receptors in astrocytes. Prog. Brain Res. 2001, 132, 287–299. [Google Scholar]
  114. Verkhratsky, A.; Kirchhoff, F. Glutamate-mediated neuronal-glial transmission. J. Anat. 2007, 210, 651–660. [Google Scholar] [CrossRef]
  115. Hamzei-Sichani, F.; Kamasawa, N.; Janssen, W.G.; Yasumura, T.; Davidson, K.G.; Hof, P.R.; Wearne, S.L.; Stewart, M.G.; Young, S.R.; Whittington, M.A.; et al. Gap junctions on hippocampal mossy fiber axons demonstrated by thin-section electron microscopy and freeze fracture replica immunogold labeling. Proc. Natl. Acad. Sci. USA 2007, 104, 12548–12553. [Google Scholar] [CrossRef] [Green Version]
  116. Traub, R.D.; Draguhn, A.; Whittington, M.A.; Baldeweg, T.; Bibbig, A.; Buhl, E.H.; Schmitz, D. Axonal gap junctions between principal neurons: A novel source of network oscillations, and perhaps epileptogenesis. Rev. Neuroci. 2002, 13, 1–30. [Google Scholar] [CrossRef] [Green Version]
  117. Traub, R.D.; Whittington, M.A.; Gutiérrez, R.; Draguhn, A. Electrical coupling between hippocampal neurons: Contrasting roles of principal cell gap junctions and interneuron gap junctions. Cell Tissue Res. 2018, 373, 671–691. [Google Scholar] [CrossRef]
  118. Mathy, A.; Clark, B.A.; Häusser, M. Synaptically induced long-term modulation of electrical coupling in the inferior olive. Neuron 2014, 81, 1290–1296. [Google Scholar] [CrossRef] [Green Version]
  119. Turecek, J.; Yuen, G.S.; Han, V.Z.; Zeng, X.H.; Bayer, K.U.; Welsh, J.P. NMDA receptor activation strengthens weak electrical coupling in mammalian brain. Neuron 2014, 81, 1375–1388. [Google Scholar] [CrossRef] [Green Version]
  120. Siesjö, B.K.; von Hanwehr, R.; Nergelius, G.; Nevander, G.; Ingvar, M. Extra- and intracellular pH in the brain during seizures and in the recovery period following the arrest of seizure activity. J. Cereb. Blood Flow Metab. 1985, 5, 47–57. [Google Scholar] [CrossRef]
  121. Schalper, K.A.; Sánchez, H.A.; Lee, S.C.; Altenberg, G.A.; Nathanson, M.H.; Sáez, J.C. Connexin 43 hemichannels mediate the Ca2+ influx induced by extracellular alkalinization. Am. J. Physiol. Cell Physiol. 2010, 299, C1504–C1515. [Google Scholar] [CrossRef] [Green Version]
  122. Palacios-Prado, N.; Briggs, S.W.; Skeberdis, V.A.; Pranevicius, M.; Bennett, M.V.; Bukauskas, F.F. pH-dependent modulation of voltage gating in connexin45 homotypic and connexin45/connexin43 heterotypic gap junctions. Proc. Natl. Acad. Sci. USA 2010, 107, 9897–9902. [Google Scholar] [CrossRef] [Green Version]
  123. Pogoda, K.; Kameritsch, P.; Retamal, M.A.; Vega, J.L. Regulation of gap junction channels and hemichannels by phosphorylation and redox changes: A revision. BMC Cell Biol. 2016, 17 (Suppl. 1), 11. [Google Scholar] [CrossRef] [Green Version]
  124. Garciarena, C.D.; Malik, A.; Swietach, P.; Moreno, A.P.; Vaughan-Jones, R.D. Distinct moieties underlie biphasic H+ gating of connexin43 channels, producing a pH optimum for intercellular communication. FASEB J. 2018, 32, 1969–1981. [Google Scholar] [CrossRef] [Green Version]
  125. Gajda, Z.; Szupera, Z.; Blazsó, G.; Szente, M. Quinine, a blocker of neuronal cx36 channels, suppresses seizure activity in rat neocortex in vivo. Epilepsia 2005, 46, 1581–1591. [Google Scholar] [CrossRef]
  126. Nilsen, K.E.; Kelso, A.R.; Cock, H.R. Antiepileptic effect of gap-junction blockers in a rat model of refractory focal cortical epilepsy. Epilepsia 2006, 47, 1169–1175. [Google Scholar] [CrossRef]
  127. Bostanci, M.O.; Bağirici, F. Anticonvulsive effects of carbenoxolone on penicillin-induced epileptiform activity: An in vivo study. Neuropharmacology 2007, 52, 362–367. [Google Scholar] [CrossRef]
  128. Medina-Ceja, L.; Cordero-Romero, A.; Morales-Villagrán, A. Antiepileptic effect of carbenoxolone on seizures induced by 4-aminopyridine: A study in the rat hippocampus and entorhinal cortex. Brain Res. 2008, 1187, 74–81. [Google Scholar] [CrossRef]
  129. Medina-Ceja, L.; Ventura-Mejía, C. Differential effects of trimethylamine and quinine on seizures induced by 4-aminopyridine administration in the entorhinal cortex of vigilant rats. Seizure 2010, 19, 507–513. [Google Scholar] [CrossRef] [Green Version]
  130. Chen, W.; Gao, Z.; Ni, Y.; Dai, Z. Carbenoxolone pretreatment and treatment of posttraumatic epilepsy. Neural Regen. Res. 2013, 8, 169–176. [Google Scholar]
  131. Sefil, F.; Arık, A.E.; Acar, M.D.; Bostancı, M.Ö.; Bagirici, F.; Kozan, R. Interaction between carbenoxolone and valproic acid on pentylenetetrazole kindling model of epilepsy. Int. J. Clin. Exp. Med. 2015, 8, 10508–10514. [Google Scholar]
  132. Franco-Pérez, J.; Manjarrez-Marmolejo, J.; Rodríguez-Balderas, C.; Castro, N.; Ballesteros-Zebadua, P. Quinine and carbenoxolone enhance the anticonvulsant activity of some classical antiepileptic drugs. Neurol. Res. 2018, 40, 26–33. [Google Scholar] [CrossRef]
  133. Szente, M.; Gajda, Z.; Said, A.K.; Hermesz, E. Involvement of electrical coupling in the in vivo ictal epileptiform activity induced by 4-aminopyridine in the neocortex. Neuroscience 1067. [Google Scholar] [CrossRef]
  134. Davidson, J.S.; Baumgarten, I.M. Glycyrrhetinic acid derivatives: A novel class of inhibitors of gap-junctional intercellular communication. Structure-activity relationships. J. Pharmacol. Exp. Ther. 1988, 246, 1104–1107. [Google Scholar]
  135. Goldberg, G.S.; Moreno, A.P.; Bechberger, J.F.; Hearn, S.S.; Shivers, R.R.; MacPhee, D.J.; Zhang, Y.C.; Naus, C.C. Evidence that disruption of connexon particle arrangements in gap junction plaques is associated with inhibition of gap junctional communication by a glycyrrhetinic acid derivative. Exp. Cell Res. 1996, 222, 48–53. [Google Scholar] [CrossRef]
  136. Guan, X.; Wilson, S.; Schlender, K.K.; Ruch, R.J. Gap-junction disassembly and connexin 43 dephosphorylation induced by 18 beta-glycyrrhetinic acid. Mol. Carcinog. 1996, 16, 157–164. [Google Scholar] [CrossRef]
  137. Yin, X.; Feng, L.; Ma, D.; Yin, P.; Wang, X.; Hou, S.; Hao, Y.; Zhang, J.; Xin, M.; Feng, J. Roles of astrocytic connexin-43, hemichannels, and gap junctions in oxygen-glucose deprivation/reperfusion injury induced neuroinflammation and the possible regulatory mechanisms of salvianolic acid B and carbenoxolone. J. Neuroinflamm. 2018, 15, 97. [Google Scholar] [CrossRef]
  138. Walden, J.; Speckmann, E.J. Effects of quinine on membrane potential and membrane currents in identified neurons of Helix pomatia. Neurosci. Lett. 1981, 27, 139–143. [Google Scholar] [CrossRef]
  139. Cherubini, E.; North, R.A.; Surprenant, A. Quinine blocks a calcium-activated potassium conductance in mammalian enteric neurones. Br. J. Pharmacol. 1984, 83, 3–5. [Google Scholar] [CrossRef] [Green Version]
  140. Srinivas, M.; Hopperstad, M.G.; Spray, D.C. Quinine blocks specific gap junction channel subtypes. Proc. Natl. Acad. Sci. USA 2001, 98, 10942–10947. [Google Scholar] [CrossRef] [Green Version]
  141. Zou, L.; Xue, Y.; Jones, M.; Heinbockel, T.; Ying, M.; Zhan, X. The Effects of Quinine on Neurophysiological Properties of Dopaminergic Neurons. Neurotox. Res. 2018, 34, 62–73. [Google Scholar] [CrossRef]
  142. Thompson, A.J.; Lummis, S.C. Antimalarial drugs inhibit human 5-HT(3) and GABA(A) but not GABA(C) receptors. Br. J. Pharmacol. 2008, 153, 1686–1696. [Google Scholar] [CrossRef] [Green Version]
  143. Coker, S.J.; Batey, A.J.; Lightbown, I.D.; Díaz, M.E.; Eisner, D.A. Effects of mefloquine on cardiac contractility and electrical activity in vivo, in isolated cardiac preparations, and in single ventricular myocytes. Br. J. Pharmacol. 2000, 129, 323–330. [Google Scholar] [CrossRef] [Green Version]
  144. Heshmati, M.; Golden, S.A.; Pfau, M.L.; Christoffel, D.J.; Seeley, E.L.; Cahill, M.E.; Khibnik, L.A.; Russo, S.J. Mefloquine in the nucleus accumbens promotes social avoidance and anxiety-like behavior in mice. Neuropharmacology 2016, 101, 351–357. [Google Scholar] [CrossRef] [Green Version]
  145. Evans, W.H.; Bultynck, G.; Leybaert, L. Manipulating connexin communication channels: Use of peptidomimetics and the translational outputs. J. Membr. Biol. 2012, 245, 437–449. [Google Scholar] [CrossRef] [Green Version]
  146. Samoilova, M.; Wentlandt, K.; Adamchik, Y.; Velumian, A.A.; Carlen, P.L. Connexin 43 mimetic peptides inhibit spontaneous epileptiform activity in organotypic hippocampal slice cultures. Exp. Neurol. 2008, 210, 762–775. [Google Scholar] [CrossRef]
  147. Yoon, J.J.; Green, C.R.; O’Carroll, S.J.; Nicholson, L.F. Dose dependent protective effect of connexin43 mimetic peptide against neurodegeneration in an ex vivo model of epileptiform lesion. Epilepsy Res. 2010, 92, 153–162. [Google Scholar] [CrossRef]
  148. Bragin, A.; Wilson, C.L.; Engel, J., Jr. Chronic epileptogenesis requires development of a network of pathologically interconnected neuron clusters: A hypothesis. Epilepsia 2000, 41 (Suppl. 6), S144–S152. [Google Scholar] [CrossRef] [Green Version]
  149. Bragin, A.; Mody, I.; Wilson, C.L.; Engel, J., Jr. Local generation of fast ripples in epileptic brain. J. Neurosci. 2002, 22, 2012–2021. [Google Scholar] [CrossRef] [Green Version]
  150. Ventura-Mejía, C.; Medina-Ceja, L. Decreased fast ripples in the hippocampus of rats with spontaneous recurrent seizures treated with carbenoxolone and quinine. BioMed Res. Int. 2014, 2014, 282490. [Google Scholar] [CrossRef]
  151. Ran, X.; Xiang, J.; Song, P.P.; Jiang, L.; Liu, B.K.; Hu, Y. Effects of gap junctions blockers on fast ripples and connexin in rat hippocampi after status epilepticus. Epilepsy Res. 2018, 146, 28–35. [Google Scholar] [CrossRef]
  152. Simon, A.; Traub, R.D.; Vladimirov, N.; Jenkins, A.; Nicholson, C.; Whittaker, R.G.; Schofield, I.; Clowry, G.J.; Cunningham, M.O.; Whittington, M.A. Gap junction networks can generate both ripple-like and fast ripple-like oscillations. Eur. J. Neurosci. 2014, 39, 46–60. [Google Scholar] [CrossRef] [Green Version]
  153. Gigout, S.; Louvel, J.; Kawasaki, H.; D’Antuono, M.; Armand, V.; Kurcewicz, I.; Olivier, A.; Laschet, J.; Turak, B.; Devaux, B.; et al. Effects of gap junction blockers on human neocortical synchronization. Neurobiol. Dis. 2006, 22, 496–508. [Google Scholar] [CrossRef]
  154. Nikolic, L.; Shen, W.; Nobili, P.; Virenque, A.; Ulmann, L.; Audinat, E. Blocking TNFα-driven astrocyte purinergic signaling restores normal synaptic activity during epileptogenesis. Glia 2018, 66, 2673–2683. [Google Scholar] [CrossRef]
  155. Wang, Y.; Denisova, J.V.; Kang, K.S.; Fontes, J.D.; Zhu, B.T.; Belousov, A.B. Neuronal gap junctions are required for NMDA receptor-mediated excitotoxicity: Implications in ischemic stroke. J. Neurophysiol. 2010, 104, 3551–3556. [Google Scholar] [CrossRef]
  156. Fellin, T.; Pascual, O.; Gobbo, S.; Pozzan, T.; Haydon, P.G.; Carmignoto, G. Neuronal synchrony mediated by astrocytic glutamate through activation of extrasynaptic NMDA receptors. Neuron 2004, 43, 729–743. [Google Scholar] [CrossRef] [Green Version]
  157. Xing, L.; Yang, T.; Cui, S.; Chen, G. Connexin hemichannels in astrocytes: Role in CNS disorders. Front. Mol. Neurosci. 2019, 12, 23. [Google Scholar] [CrossRef] [Green Version]
  158. Rouach, N.; Koulakoff, A.; Abudara, V.; Willecke, K.; Giaume, C. Astroglial metabolic networks sustain hippocampal synaptic transmission. Science 2008, 322, 1551–1555. [Google Scholar] [CrossRef]
  159. Vincze, R.; Péter, M.; Szabó, Z.; Kardos, J.; Héja, L.; Kovács, Z. Connexin 43 Differentially Regulates Epileptiform Activity in Models of Convulsive and Non-convulsive Epilepsies. Front. Cell. Neurosci. 2019, 13, 173. [Google Scholar] [CrossRef]
  160. Walrave, L.; Pierre, A.; Albertini, G.; Aourz, N.; De Bundel, D.; Van Eeckhaut, A.; Vinken, M.; Giaume, C.; Leybaert, L.; Smolders, I. Inhibition of astroglial connexin 43 hemichannels with TAT-Gap19 exerts anticonvulsant effects in rodents. Glia 2018, 66, 1788–1804. [Google Scholar] [CrossRef] [Green Version]
Figure 1. General representation of the connexins (Cxs)-based channels and hemichannels participation in seizures and epilepsy, through synergistic actions of neuroinflammation and hyper excitability and synchronization. Different types of insults can produce seizures that modify intracellular pH, which opens hemichannels in glial cells (activated microglia and astrocytes), with the consequent calcium-dependent release of gliotransmitters (Glutamate Glu, D-serine, NAD and ATP) and proinflammatory molecules (NLRP-3 inflammasome: Caspase 1, IL-1β, IL18) that lead to neuroinflammation and open Cxs-based hemichannels. This last action is also facilitated by reactive oxygen species (ROS) and reactive nitrogen species (NOS) produced during seizures. In the same way, astrocytes can reinforce seizures through the opening of Cxs-based channels that activate a calcium wave in an astroglial syncytium that propagates excitability. In neurons, the opening of channels increases abnormal high excitability and synchronization that reinforces the cycle of seizures generation.
Figure 1. General representation of the connexins (Cxs)-based channels and hemichannels participation in seizures and epilepsy, through synergistic actions of neuroinflammation and hyper excitability and synchronization. Different types of insults can produce seizures that modify intracellular pH, which opens hemichannels in glial cells (activated microglia and astrocytes), with the consequent calcium-dependent release of gliotransmitters (Glutamate Glu, D-serine, NAD and ATP) and proinflammatory molecules (NLRP-3 inflammasome: Caspase 1, IL-1β, IL18) that lead to neuroinflammation and open Cxs-based hemichannels. This last action is also facilitated by reactive oxygen species (ROS) and reactive nitrogen species (NOS) produced during seizures. In the same way, astrocytes can reinforce seizures through the opening of Cxs-based channels that activate a calcium wave in an astroglial syncytium that propagates excitability. In neurons, the opening of channels increases abnormal high excitability and synchronization that reinforces the cycle of seizures generation.
Ijms 20 05976 g001

Share and Cite

MDPI and ACS Style

Medina-Ceja, L.; Salazar-Sánchez, J.C.; Ortega-Ibarra, J.; Morales-Villagrán, A. Connexins-Based Hemichannels/Channels and Their Relationship with Inflammation, Seizures and Epilepsy. Int. J. Mol. Sci. 2019, 20, 5976. https://doi.org/10.3390/ijms20235976

AMA Style

Medina-Ceja L, Salazar-Sánchez JC, Ortega-Ibarra J, Morales-Villagrán A. Connexins-Based Hemichannels/Channels and Their Relationship with Inflammation, Seizures and Epilepsy. International Journal of Molecular Sciences. 2019; 20(23):5976. https://doi.org/10.3390/ijms20235976

Chicago/Turabian Style

Medina-Ceja, Laura, Juan C. Salazar-Sánchez, Jorge Ortega-Ibarra, and Alberto Morales-Villagrán. 2019. "Connexins-Based Hemichannels/Channels and Their Relationship with Inflammation, Seizures and Epilepsy" International Journal of Molecular Sciences 20, no. 23: 5976. https://doi.org/10.3390/ijms20235976

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop