Next Article in Journal
Biocomposites of Epoxidized Natural Rubber/Poly(lactic acid) Modified with Natural Fillers (Part I)
Next Article in Special Issue
Modafinil Administration to Preadolescent Rat Impairs Non-Selective Attention, Frontal Cortex D2 Expression and Mesolimbic GABA Levels
Previous Article in Journal
Why Concurrent CDDP and Radiotherapy Has Synergistic Antitumor Effects: A Review of In Vitro Experimental and Clinical-Based Studies
Previous Article in Special Issue
Intricacies of GABAA Receptor Function: The Critical Role of the β3 Subunit in Norm and Pathology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Phospholipase C in GABAergic Inhibition and Its Relevance to Epilepsy

1
Department of Biological Sciences, Ulsan National Institute of Science and Technology (UNIST), Ulsan 44919, Korea
2
Korea Brain Research Institute (KBRI), Daegu 41062, Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(6), 3149; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22063149
Submission received: 29 November 2020 / Revised: 2 March 2021 / Accepted: 17 March 2021 / Published: 19 March 2021
(This article belongs to the Special Issue GABAergic Physiology in Neurological Disorders)

Abstract

:
Epilepsy is characterized by recurrent seizures due to abnormal hyperexcitation of neurons. Recent studies have suggested that the imbalance of excitation and inhibition (E/I) in the central nervous system is closely implicated in the etiology of epilepsy. In the brain, GABA is a major inhibitory neurotransmitter and plays a pivotal role in maintaining E/I balance. As such, altered GABAergic inhibition can lead to severe E/I imbalance, consequently resulting in excessive and hypersynchronous neuronal activity as in epilepsy. Phospholipase C (PLC) is a key enzyme in the intracellular signaling pathway and regulates various neuronal functions including neuronal development, synaptic transmission, and plasticity in the brain. Accumulating evidence suggests that neuronal PLC is critically involved in multiple aspects of GABAergic functions. Therefore, a better understanding of mechanisms by which neuronal PLC regulates GABAergic inhibition is necessary for revealing an unrecognized linkage between PLC and epilepsy and developing more effective treatments for epilepsy. Here we review the function of PLC in GABAergic inhibition in the brain and discuss a pathophysiological relationship between PLC and epilepsy.

1. Introduction

Epilepsy, one of the most common neurological disorders, is characterized by repeated spontaneous seizures with abnormal hyperexcitation and synchronous discharge of neurons [1]. Approximately 60 million people worldwide suffer from epilepsy with cognitive and psychiatric comorbidities [2,3]. Although several biological factors have been identified as an etiology of epilepsy, including genetic mutation, brain injury, tumor, and aging, the precise cause of epilepsy in most cases is still unknown [4,5]. One of the hypotheses explaining the pathophysiological mechanism of epilepsy is that the disruption of excitation and inhibition balance (E/I balance) could generally lead to abnormal excitability of neurons [6,7,8]. In the brain, neurons receive numerous excitatory and inhibitory synaptic inputs and once the synaptic potentials in dendrites and soma are integrated together, neurons produce axon potentials with various shapes, rates, and patterns of firing [9]. E/I balance either by increasing excitation or decreasing inhibition is associated with the hyperexcitation of neurons, which can cause epileptic seizures [7,10,11].
γ-aminobutyric acid (GABA) is a major inhibitory neurotransmitter in the brain and using GABA, GABAergic inhibitory neurons primarily regulate the excitability of neurons. GABAergic neurons produce GABA from glutamate using glutamic acid decarboxylase (GAD) and this synthesized GABA is packaged into synaptic vesicles at synaptic terminals through vesicular GABA transporters (VGATs). Synaptically released GABA binds to both presynaptic and postsynaptic GABA receptors (GABAA and GABAB) and suppresses the excitation of presynaptic terminals and postsynaptic neurons. In addition, the uptake of released GABA at GABAergic synapses is mediated by GABA transporters (GATs). Molecular and cellular abnormalities of GABA synthesis, release, uptake, and GABA receptor-mediated signaling can alter E/I balance in neurons and the dysfunctions of any of these processes may be implicated in neurological disorders including epilepsy, schizophrenia, and autism [12,13,14]. As a matter of fact, the therapeutic rationale of the majority of current antiepileptic drugs (AEDs) is to restore altered E/I balance by elevating the level of GABA at synapses and potentiating the functions of GABA receptors. Hence, a better understanding of the underlying molecular mechanisms that regulate GABAergic inhibition in the brain will be crucial to identify new drug targets, as well as to increase the efficacy and minimize the side effects of antiepileptic drugs.
Phospholipase C (PLC) is an essential enzyme in intracellular signal transduction cascades (Figure 1). PLC hydrolyzes phosphatidylinositol 4,5-bisphosphate (PIP2), generating secondary signal transducers including inositol 1,4,5-triphosphate (IP3) and diacylglycerol (DAG). IP3 increases intracellular calcium level by binding to IP3 receptors in the endoplasmic reticulum and DAG activates protein kinase C (PKC)-related signaling cascades. Throughout the body, PLC is associated with key cellular processes such as proliferation, differentiation, migration, and survival [15,16,17,18]. There are in total 13 mammalian isozymes of PLC including β (1–4), γ (1, 2), δ (1, 3, 4), ε, ζ, and η (1, 2), which are classified according to their distinct domain structures and biochemical properties. Each PLC isozyme is differentially expressed among tissues and regulates the complex cellular functions in a tissue-dependent manner. Among these PLC isozymes, PLCβ and PLCγ are major PLC enzymes abundantly expressed in the brain and play diverse roles in neuronal functions. In this review article, we summarize the molecular and cellular mechanisms of GABAergic inhibition on the regulation of E/I balance. Then we particularly focus on the role of PLC in GABAergic inhibition. Finally, we discuss the potential relationship between PLC and epilepsy.

2. Epilepsy Model and Antiepileptic Drugs

There are several epilepsy animal models to study the mechanism of epileptogenesis and evaluate the efficacy of AEDs [19,20,21,22,23] (Table 1). Animal models of epilepsy share similar pathological mechanisms as well as mimic the seizure behaviors of human patients. Therefore, drug screening using various epilepsy models is useful and essential to develop new AEDs. Chemoconvulsants such as pilocarpine and kainate (also referred as kainic acid, KA) are generally used to generate acute seizures within a relatively short period of time by systemic- or micro-injection. Pilocarpine is a muscarinic acetylcholine receptor agonist and pilocarpine-induced seizure has been well established as an epilepsy model. Systemic injection of pilocarpine induces tonic–clonic generalized seizures by activating M1 muscarinic receptors, especially in hippocampal neurons, and causes electrophysiological and morphological changes in the hippocampus [24,25,26]. KA is an L-glutamate analog and activates ionotropic KA receptors that are usually expressed in both excitatory and inhibitory synapses. Administration of KA causes neuronal depolarization and excitotoxic damage, particularly in the hippocampus, eventually leading to spontaneous seizures [27,28,29]. In another acute epilepsy model, electrical stimulation, so-called electrical kindling, is repeatedly delivered to the specific brain regions such as the hippocampus and amygdala. This electrical kindling increases NMDA receptor-dependent synaptic transmission, neuronal loss, and mossy fiber sprouting in the hippocampus, all of which are similar to the deficits in human epilepsy patients [30,31,32].
Unlike the acute seizures by chemical stimulants or electrical stimulation, genetic models of epilepsy in rats and mice can provide more direct insight into the genetic etiology of human epilepsy. The Genetic Absence Epilepsy Rat from Strasbourg (GAERS) is an absence seizure model that is characterized by a brief and nonconvulsive behavioral arrest and apparent unconsciousness with spike-and-wave discharges (SWDs) on electroencephalographic recordings [33]. The Wistar Albino Glaxo from Rijswijk (WAG/Rij) is also one of the absence seizure models, while the specific pathophysiological mechanism of these inbred strains is still not fully understood. Dilute brown agouti coat color (DBA/2) mice frequently show tonic-clonic seizures in response to a specific auditory stimulus [34,35]. Similarly, genetically epilepsy-prone rats (GEPR) exhibit sound-induced seizures with GABAergic, serotonergic, and noradrenergic deficits [36,37,38,39,40].
Currently, many AEDs have been developed for the treatment of complex seizure types [44]. Mechanisms of action of major AEDs are to decrease neuronal excitation by controlling voltage-gated ion channels and glutamatergic neurotransmission or to increase neuronal inhibition by upregulating GABA level and potentiating the responsiveness of GABA receptors. In this review, we will focus on the AEDs targeting GABAergic mechanisms (Table 2). Both benzodiazepines and barbiturates are positive allosteric modulators of GABAA receptors and act on GABAA receptors by increasing the conductance of chloride through ion channels [45,46,47], yet these drugs regulate GABAA receptors in different ways. Benzodiazepines bind to GABAA receptors only when in the presence of GABA, while barbiturates, when at high concentrations, can bind to GABAA receptors even without ambient GABA. Phenobarbital, an allosteric modulator of GABAA receptors, has been widely used for the treatment of status epilepticus and generalized tonic-clonic seizures. Upon binding to GABAA receptor subunits, it increases the influx of Cl into neurons and therefore reinforces the hyperpolarization, causing the inhibition of neuronal excitation [44,46,48]. Although effective in reducing seizures, phenobarbital has severe cognitive and behavioral adverse effects, including decreased consciousness, dizziness, nystagmus, and ataxia [49]. Vigabatrin, an irreversible inhibitor of mitochondrial enzyme GABA transaminase, blocks the catabolic process of GABA [50,51]. Tiagabine is a selective competitive inhibitor of GABA transporter GAT-1, which blocks the reuptake of GABA in the synaptic cleft [52]. Vigabatrin and tiagabine increase the level of ambient GABA at synapses so that they can facilitate GABAergic inhibition.

3. GABAergic Dysfunction in Epilepsy

Excitatory and inhibitory synaptic currents precisely coordinate neuronal functions at the level of the synapse and neural circuit. E/I balance in the brain is determined by several physiological factors. For instance, synapse development, transmission, and plasticity modulate E/I balance at the synapse level, while the firing properties of neurons and spatiotemporal synchronization of neuronal firing can determine E/I balance at the level of the neural circuit [66,67]. At the most fundamental level, however, E/I balance is governed by neurotransmitters glutamate and GABA, both of which are primary neurotransmitters to regulate E/I balance in the brain. The concentration of glutamate in the mammalian brain is approximately 80–100 nmol/mg protein and that of GABA is 10–30 nmol/mg protein [68]. Although the population of GABAergic neurons is also much smaller than that of excitatory neurons, which is about 25–30% of excitatory neurons, GABAergic neurons strongly control the excitation of neurons with local and long-distance innervations in the brain [69]. Previous studies have observed altered GABAergic inhibition in both epilepsy animal models and human patients. In temporal lobe epilepsy patients as well as in the pilocarpine-induced epilepsy animal model, mRNA and protein levels of GAD65 and GAD67 were markedly increased in the hippocampus, probably to facilitate GABA synthesis and protect against the long-term hyperexcitability of neurons [70,71]. Two isoforms of GAD, GAD65, and GAD67, are expressed in distinct subcellular locations during neuronal development. GAD67 is mostly distributed in neuronal cell bodies, whereas GAD65 is highly expressed in axon terminals [72]. GAD65 knockout mice showed increased seizure susceptibility and the genetic deletion of GAD67 caused severe developmental defects and early deaths in mice [70,73,74,75,76]. On the other hand, elevated GAD67 expression in the hippocampal CA3 region by using recombinant adeno-associated virus (AAV) significantly decreased the seizure generation in temporal lobe epilepsy models [77]. These results indicate that the manipulation of GABA level by targeting the GAD enzyme can be a therapeutic strategy for maintaining appropriate excitability of neurons and treating epilepsy [77]. VGAT is also responsible for synaptic inhibition at GABAergic inhibitory synapses and its dysfunction is related to epilepsy. However, the specific role and contribution of VGAT is still debated in epilepsy animal models. For instance, VGAT level was found to decrease in animal epilepsy models such as cortical dysplasia, Mongolian gerbil, and picrotoxin model. Yet other studies, in contrast, reported no change of VGAT in the kainic acid model or increased VGAT expression in the pilocarpine model [78,79,80,81,82,83,84,85]. Mutation of GAT-1 was identified in epileptic encephalopathy patients, resulting in an impaired GABA reuptake process [86]. In addition to neuronal GAT-1, astrocytic GAT-1 is also crucial to GABAergic inhibition in neural circuits and its currents were attenuated in the absence seizure model [87]. This evidence supports the idea that GABA synthesis and reuptake are essential for maintaining GABAergic inhibition and the dysregulation of these processes can induce epileptic seizures.
Activation of GABA receptors by GABA binding initiates and critically mediates the postsynaptic effects of GABAergic synaptic transmission. There are two types of GABA receptors in the brain, which are GABAA receptors and GABAB receptors. GABAA receptors are ligand-gated ion channels and function to mediate the majority of fast inhibitory synaptic transmission in the brain. Activation of GABAA receptors leads to the hyperpolarization of neurons by increasing the movement of Cl into the cytosol, which prevents the action potential for neuronal activity [88,89,90]. In humans, GABAA receptors that consist of five subunits, have 19 distinct subunits including α (1–6), β (1–3), γ (1–3), δ, ε, θ, π, and ρ (1–3). Unlike GABAA receptors, GABAB receptors are G protein-coupled receptors (GPCRs) and are responsible for the slow inhibitory function. They mediate the hyperpolarization of neurons by opening inwardly-rectifying K+ channels and inactivating the voltage-gated Ca2+ channels [91,92]. Functional GABAB receptors are assembled from two subunits, GABAB1 and GABAB2, to form heterodimers [93]. Any dysfunction of GABA receptor-mediated signaling is directly implicated in various neurological disorders and psychiatric disorders [94,95]. Importantly, genetic and functional alterations of GABA receptors are closely linked to epilepsy. Mutations of GABA receptors have been identified in human epilepsy patients. For example, mutations in the α1 subunit of the GABAA receptor were observed in human juvenile myoclonic epilepsy [96]. Missense variants of α1, α5 subunits in GABAA receptors were found in early-onset epilepsy patients [97]. Moreover, the γ2 subunit of the GABAA receptor was mutated in children with absence seizures [98]. Consistent with these human studies, genetic ablation of GABAA receptors in mice led to epileptic seizures [99]. The expression of GABAA receptors was also reduced in the rodent hippocampus following pilocarpine or kainic acid-induced seizures [100,101]. As with GABAA receptors, the mutations of GABAB receptors have been reported in epilepsy. Polymorphisms of the GABAB receptor were frequently present in temporal lobe epilepsy patients [102,103]. In animal studies, the expression of the GABAB receptor was decreased in the rat epilepsy model [104]. GABAB receptor knockout mice showed spontaneous seizures with hyperalgesia, hyperlocomotion, and memory deficit [105]. In current antiepileptic treatments, some antiepileptic drugs enhance GABAergic inhibition through the elevation of brain GABA levels or the potentiation of GABAA receptors [46,47,106]. As mentioned above, barbiturates and benzodiazepines allosterically modulate GABAA receptors and are widely used as antiepileptic drugs [107,108]. Conversely, GABAA receptor antagonists including bicuculline, picrotoxin, and pentylenetetrazol induce seizures and are used to experimentally model epilepsy [109]. Together, a deeper understanding of the GABAergic system and GABA receptor-mediated signaling in the brain is vital to the identification of etiology and therapeutic targets for epilepsy.

4. Phospholipase C Beta (PLCβ) and GABAergic Inhibition

PLCβ is activated by Gαq and Gβγ subunits of G protein-coupled receptors (GPCRs) and interacts with diverse GPCRs such as 5-HT receptors, metabotropic glutamate receptors, and muscarinic acetylcholine receptors [110,111,112]. PLCβ is abundantly expressed in the brain and plays a pivotal role in regulating neuronal activity. Several studies have investigated the role of PLCβ in neuronal cells by examining PLCβ isoform-specific knockout mice. PLCβ1 knockout mice displayed an altered anxiety level and memory impairment [113]. When PLCβ3 was genetically deleted in mice, these PLCβ3-deficient mice showed high sensitivity to an opioid agonist, indicating that PLCβ3 would be a key regulator in opioid signaling and addiction [114]. PLCβ4 knockout mice exhibited deficits in visual processing, nociceptive responses, muscular coordination, and synapse elimination in the cerebellum [115,116,117].
In addition, a large number of studies have demonstrated that PLCβ is implicated in regulating GABAergic inhibition. Acetylcholine can regulate GABAergic synaptic transmission by affecting hyperpolarization, depolarization, GABA release, and oscillatory properties of GABAergic interneurons in the hippocampus [118]. Acetylcholine binds to and activates two types of receptors, which are nicotinic acetylcholine receptor (nAChR, ligand-gated ion channel) and muscarinic acetylcholine receptor (mAChR, metabotropic receptor). mAChR is a GPCR and is classified into five subtypes. Among these five subtypes, M1, M3, and M5 receptors are coupled with the Gαq protein and can activate the PLCβ signaling pathway. Notably, the activation of muscarinic receptors is associated with the generation of epileptic seizures [119]. Activation of muscarinic M1 and M2 receptors by muscarinic agonists muscarine and oxotremorine suppressed GABA release, reducing the amplitude of evoked inhibitory postsynaptic currents (eIPSCs) in the rat auditory cortex [120]. On the contrary, the application of PLC blocker U73122 alleviated the oxotremorine-induced depression of eIPSC amplitude. This result clearly shows that the activation of muscarinic acetylcholine receptors can inhibit GABA transmission via the PLCβ pathway, eventually altering neuronal excitation. In addition, KA receptors are implicated in epilepsy and KA-induced seizures. Interestingly, KA receptors can affect GABAergic synaptic transmission in the presynaptic terminal. In the hippocampus, the activation of KA receptors attenuated GABA release through the PLCβ pathway in immature CA3 neurons of rats [121,122]. Moreover, presynaptic D2 receptors activated the PLC-IP3-calcineurin signaling cascade and decreased GABA release at striatopallidal terminals of medium-sized spiny GABAergic neurons [123]. These results reveal that several neurotransmitters act on presynaptic receptors and can activate presynaptic PLCβ. Once activated, PLCβ stimulates the downstream signaling pathway at the GABAergic presynaptic terminal, leading to the inhibition of GABA release. On the other hand, it is also likely that PLCβ is involved in enhancing GABA release. In the rat cerebellum, ethanol increased spontaneous GABA release from Purkinje cells through PLC activation, and the suppression of PLC by using PLC antagonist edelfosine prevented the ethanol-induced GABA release [124]. Similarly, 5-HT2A receptors activated PLCβ and facilitated GABA release at the thalamic reticular presynaptic terminal [125,126]. From this evidence, it appears that PLC and the downstream signaling pathway mainly suppress GABA release at the GABAergic presynaptic terminal, while they also enhance GABA release depending on cell types and interacting receptors (Figure 2).
PLCβ and its relevance to epilepsy have been strongly suggested in many studies (Table 3). Genetic mutations of PLCβ1 were found in human infantile epileptic encephalopathy [127,128,129,130]. In animal studies, PLCβ1 knockout mice exhibited status epilepticus with generalized and tonic-clonic seizures and prematurely died starting from three-weeks-old [131]. Moreover, the population of somatostatin interneurons that primarily target the distal dendrite of a neuron was reduced in the hilus of PLCβ1 knockout mice. Loss of PLCβ1 function by using transgene insertion in mice caused late-onset epileptic symptoms at six months of age with aberrant mossy fiber sprouting in the hippocampus, decreased NMDA receptor density in the stratum oriens of the CA1, and increased apoptosis [132]. In the thalamus, dysfunction of PLCβ4 in thalamocortical neurons induced abnormal burst firing, resulting in absence seizures characterized by spike-wave discharges on electroencephalography (EEG) [133]. On the other hand, the expression of PLCβ was remarkably increased in the hippocampus in the chemical-induced epilepsy model, possibly to protect against neuronal damage [134,135,136]. Furthermore, both PLCβ1 and PLCβ4 expression in hippocampal GABAergic interneurons was reduced following pilocarpine-induced seizures [136]. Collectively, these findings suggest that the function of PLCβ is potentially important for preventing excessive neuronal excitability in epilepsy.

5. Phospholipase C Gamma (PLCγ) and GABAergic Inhibition

PLCγ is usually activated by the phosphorylation of receptor tyrosine kinases (RTKs) in response to extracellular ligands such as brain-derived neurotrophic factor (BDNF), epidermal growth factor (EGF), platelet-derived growth factor (PDGF), nerve growth factor (NGF), and fibroblast growth factor (FGF). PLCγ has Src homology (SH) domains such as SH2 and SH3. SH2 binds to the tyrosine-autophosphorylation site of RTKs and SH3 interacts with other signaling molecules including the c-Cbl family of E3 ubiquitin ligase and proline-rich motifs on SOS1 which is a guanine nucleotide exchange factor [141,142]. Two isoforms of PLCγ, PLCγ1, and PLCγ2, exhibit distinct expression patterns throughout the body. PLCγ1 is highly expressed in diverse tissues including the brain, while PLCγ2 is primarily enriched in immune cells and its expression in the nervous system is relatively weak [18,143]. Polymorphism in PLCγ1 was frequently found in bipolar patients [144,145] and excitatory neuron-specific PLCγ1 knockout mice exhibited manic-like behaviors such as hyperlocomotion, decreased anxiety, increased hedonic action, and impaired learning and memory performance [146]. Moreover, the ablation of PLCγ1 in neuronal progenitor cells resulted in impaired axon guidance during developmental stages [147].
Meanwhile, it was found that the treatment of BDNF suppressed GAT-1-mediated GABA uptake and this inhibition was abolished by TrkB and PLC inhibitors such as K252a and U73122 in rat hippocampal nerve terminals [148]. In another study, the activation of PLC did not change GABA uptake in neurons, but decreased GABA uptake in astrocytes [149]. These findings suggest that PLC can control GABAergic transmission by modulating GABA reuptake in a cell type-dependent manner. Many studies have focused on the function of the BDNF/TrkB/PLCγ pathway in controlling GABAA receptor signaling. The physiological role of BDNF in epilepsy has been well documented, but it is still unclear whether the effect of BDNF is facilitating or inhibiting epileptic seizures [150,151]. It has been shown that short-term and long-term treatment of BDNF may exert differential effects on GABAergic transmission. Acute BDNF treatment decreased both evoked and spontaneous inhibitory postsynaptic currents (IPSCs), which was caused partly by a rapid reduction in postsynaptic GABAA receptor number, but did not affect excitatory postsynaptic currents (EPSCs) in hippocampal CA1 neurons. Furthermore, BDNF-induced attenuation of IPSCs was significantly suppressed by TrkB inhibitor and PLC inhibitor, indicating that these effects by BDNF seem to be mediated by TrkB and PLC [152,153]. In the mouse cerebellum, acute application of BDNF also reduced postsynaptic GABA response in cerebellar granule cells, whereas the same BDNF treatment potentiated GABAA receptor functions via the TrkB-PLCγ pathway in cerebellar Purkinje cells and consequently enhanced the amplitude of mIPSCs [154]. Another study showed that BDNF promoted the maturation of GABAergic neurons and upregulated the expression of GABAA receptor in cultured hippocampal neurons. Given the evidence above, although direct evidence was lacking in this report, this trophic effect of BDNF might be potentially mediated by PLC [155]. Interestingly, the effect of the BDNF/TrkB/PLC signaling pathway on GABAA receptor function can change across the developmental stages. It was shown that BDNF treatment acutely potentiated postsynaptic GABAA receptor-mediated currents in the rat hippocampus at postnatal day six. However, through TrkB-PLCγ signaling, BDNF later induced a long-lasting inhibition of postsynaptic GABAA receptor at postnatal day 14 [156]. In addition, the BDNF-TrkB-PLCγ pathway rapidly increased the number of GABAA receptors in the developing rat visual cortex [157], demonstrating that PLC may be one of the critical players regulating GABAA receptor expression, function, and eventually GABAergic inhibition.
In line with the above findings, it appears that PLC is also implicated in GABAB receptor-mediated functions. GABAB receptors are located in both presynaptic and postsynaptic neurons. Presynaptic GABAB receptors inhibit adenylyl cyclase and Ca2+ channels through Gαi/o and Gβγ proteins, respectively, thus inhibiting additional GABA release. Upon activation, postsynaptic GABAB receptors function to open K+ channels and cause the hyperpolarization of postsynaptic neurons, so that it suppresses neuronal excitation. GABAB receptors can also stimulate PLC and induce BDNF secretion to facilitate the maturation of GABAergic synapses at developmental stages [158]. Released BDNF increased the expression of GABAA receptor β2, β3 subunits in the postsynaptic neuronal surface, and enhanced GABAergic response [159]. As further evidence showing the functional connection between the GABAB receptor and PLC, baclofen is a GABAB receptor agonist and the application of PLC inhibitor U73122 suppressed GABAB receptor-mediated Ca2+ increase by baclofen in human retinal pigment epithelium [160]. Moreover, GABAB receptor activation could modulate the insulin-like growth factor 1 signaling pathway via PLC activation to prevent apoptosis in cerebellar granule neurons [161].
Neuronal K+-Cl cotransporter KCC2 mediates Cl extrusion from the cytosol to maintain a low intracellular concentration of Cl and hyperpolarize postsynaptic neurons [162,163,164]. The mRNA and protein expression of KCC2 were decreased in the hippocampus following kindling-induced seizures. In addition, both KCC2 mutation and downregulated KCC2 expression were identified in epilepsy patients [165]. Importantly, BDNF application could decrease the expression of KCC2 in hippocampal neurons, but the mutation of the PLCγ interaction site in TrkB rather increased the expression of KCC2 protein by exposure to BDNF [166]. Altogether, this evidence clearly indicates that PLCγ plays a key role in the regulation of GABAergic inhibition by exerting diverse actions on both presynaptic and postsynaptic sites (Figure 3).
Most studies of PLCγ on epilepsy have focused on the BDNF/TrkB/PLCγ1 signaling cascade in epileptogenesis. It was shown that seizures induced by electrical stimulation were significantly attenuated by genetic knockout of TrkB and PLCγ1 [138,167]. In addition, blocking PLCγ1 binding to TrkB suppressed epileptogenesis in the hippocampus [137,139], implying that BDNF/TrkB/PLCγ1 signaling may contribute to the etiology of epileptogenesis by shifting E/I balance towards more excitation. However, a recent study by our group found that GABAergic neuron-specific PLCγ1 knockout mice over six months of age exhibited late-onset seizures and other behavioral abnormalities owing to impaired GABAergic inhibition in the hippocampus [140]. In spite of many unanswered questions, these studies highlight a critical role for PLC in the generation of epilepsy, while its exact role may be dependent on neuronal cell types and brain regions.

6. Concluding Remark and Future Perspectives

Although epilepsy is one of the most common neurological disorders, it is apparent that the treatment of epilepsy still remains less than optimum. Many AEDs have been developed over the past decades, most of which were effective in treating epileptic seizures. However, 30–40% of epilepsy patients are not adequately controlled by current AEDs and these patients develop refractory epilepsy [168]. An alternative treatment option is brain surgery that partially removes the epileptic brain region, but it inherently has a high-risk with serious side effects. As such, a better understanding of the fundamental etiology, as well as pathophysiological mechanism of epilepsy, is crucial to identify novel and more effective treatments. Although much needs to be further investigated, a recent hypothesis of E/I imbalance in the etiology of epilepsy advanced our understanding of the molecular, cellular, and synaptic mechanisms underlying epilepsy. Most notably, it has been demonstrated that despite their small population, GABAergic neurons and GABAergic inhibition have a powerful effect in maintaining optimal E/I balance. In this review, we briefly described the potential and critical role of PLC in regulating GABAergic inhibition and epilepsy in the brain. As documented throughout this review, it is increasingly evident that PLC plays an essential role in controlling GABAergic transmission at both presynaptic and postsynaptic sites, through modulating GABA release, reuptake, and GABA receptor-mediated signaling. In addition, evidence is accumulating that major PLC enzymes highly expressed in the brain, including PLCβ and PLCγ, are directly and indirectly linked to epileptogenesis both in human patients and animal models. As PLC can exert differential molecular actions on neurons via a cell type-specific manner, undoubtedly, further work will be required to dissect the exact role of each neuronal PLC isoform on regulating GABAergic inhibition and E/I balance in both excitatory and inhibitory neurons. Together, PLC can be a potential and new therapeutic target for epilepsy, and pharmacological manipulation of specific PLC isoform may prove therapeutically fruitful in the treatment of epilepsy.

Author Contributions

Conceptualization, H.Y.K. and J.-I.K.; writing—original draft preparation, H.Y.K.; writing—review and editing, P.-G.S. and J.-I.K.; supervision, J.-I.K. All authors have read and agreed to the current version of the manuscript.

Funding

This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (2020R1A2C1005492 to J.-I.K.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

E/I balanceExcitation and inhibition balance
PLCPhospholipase C
GABAγ-aminobutyric acid
GADGlutamic acid decarboxylase
VGATVesicular GABA transporter
AEDAntiepileptic drug
PIP2Phosphatidylinositol 4,5-bisphosphate
IP3Inositol 1,4,5-triphosphate
DAGDiacylglycerol
PKCProtein kinase C
GPCRG protein-coupled receptor
RTKReceptor tyrosine kinase
BDNFBrain-derived neurotrophic factor
TrkBTropomycin receptor kinase B
AAVAdeno-associated virus
GATGABA transporter
EPSCExcitatory postsynaptic current
IPSCInhibitory postsynaptic current
KCCK+-Cl cotransporter
nAChRNicotinic acetylcholine receptor
mGluRMetabotropic glutamate receptor
EGFEpidermal growth factor
PDGFPlatelet-derived growth factor
NGFNerve growth factor
FGFFibroblast growth factor

References

  1. Goldberg, E.M.; Coulter, D.A. Mechanisms of epileptogenesis: A convergence on neural circuit dysfunction. Nat. Rev. Neurosci. 2013, 14, 337–349. [Google Scholar] [CrossRef] [Green Version]
  2. Moshe, S.L.; Perucca, E.; Ryvlin, P.; Tomson, T. Epilepsy: New advances. Lancet 2015, 385, 884–898. [Google Scholar] [CrossRef]
  3. Beghi, E.; Giussani, G.; Nichols, E.; Abd-Allah, F.; Abdela, J.; Abdelalim, A.; Abraha, H.N.; Adib, M.G.; Agrawal, S.; Alahdab, F.; et al. Global, regional, and national burden of epilepsy, 1990–2016: A systematic analysis for the Global Burden of Disease Study 2016. Lancet Neurol. 2019, 18, 357–375. [Google Scholar] [CrossRef] [Green Version]
  4. Noebels, J. Pathway-driven discovery of epilepsy genes. Nat. Neurosci. 2015, 18, 344–350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Wang, J.; Lin, Z.J.; Liu, L.; Xu, H.Q.; Shi, Y.W.; Yi, Y.H.; He, N.; Liao, W.P. Epilepsy-associated genes. Seizure 2017, 44, 11–20. [Google Scholar] [CrossRef] [Green Version]
  6. Gatto, C.L.; Broadie, K. Genetic controls balancing excitatory and inhibitory synaptogenesis in neurodevelopmental disorder models. Front. Synaptic Neurosci. 2010, 2, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Bonansco, C.; Fuenzalida, M. Plasticity of Hippocampal Excitatory-Inhibitory Balance: Missing the Synaptic Control in the Epileptic Brain. Neural Plast. 2016, 2016, 8607038. [Google Scholar] [CrossRef] [Green Version]
  8. Sohal, V.S.; Rubenstein, J.L.R. Excitation-inhibition balance as a framework for investigating mechanisms in neuropsychiatric disorders. Mol. Psychiatry 2019, 24, 1248–1257. [Google Scholar] [CrossRef]
  9. Bean, B.P. The action potential in mammalian central neurons. Nat. Rev. Neurosci. 2007, 8, 451–465. [Google Scholar] [CrossRef] [PubMed]
  10. Fritschy, J.M. Epilepsy, E/I Balance and GABA(A) Receptor Plasticity. Front. Mol. Neurosci. 2008, 1, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Nelson, S.B.; Valakh, V. Excitatory/Inhibitory Balance and Circuit Homeostasis in Autism Spectrum Disorders. Neuron 2015, 87, 684–698. [Google Scholar] [CrossRef] [Green Version]
  12. Treiman, D.M. GABAergic mechanisms in epilepsy. Epilepsia 2001, 42 (Suppl. 3), 8–12. [Google Scholar] [CrossRef]
  13. Coghlan, S.; Horder, J.; Inkster, B.; Mendez, M.A.; Murphy, D.G.; Nutt, D.J. GABA system dysfunction in autism and related disorders: From synapse to symptoms. Neurosci. Biobehav. Rev. 2012, 36, 2044–2055. [Google Scholar] [CrossRef] [Green Version]
  14. de Jonge, J.C.; Vinkers, C.H.; Hulshoff Pol, H.E.; Marsman, A. GABAergic Mechanisms in Schizophrenia: Linking Postmortem and In Vivo Studies. Front. Psychiatry 2017, 8, 118. [Google Scholar] [CrossRef] [Green Version]
  15. Suh, P.G.; Park, J.I.; Manzoli, L.; Cocco, L.; Peak, J.C.; Katan, M.; Fukami, K.; Kataoka, T.; Yun, S.; Ryu, S.H. Multiple roles of phosphoinositide-specific phospholipase C isozymes. BMB Rep. 2008, 41, 415–434. [Google Scholar] [CrossRef]
  16. Yang, Y.R.; Follo, M.Y.; Cocco, L.; Suh, P.G. The physiological roles of primary phospholipase C. Adv. Biol. Regul. 2013, 53, 232–241. [Google Scholar] [CrossRef]
  17. Yang, Y.R.; Kang, D.S.; Lee, C.; Seok, H.; Follo, M.Y.; Cocco, L.; Suh, P.G. Primary phospholipase C and brain disorders. Adv. Biol. Regul. 2016, 61, 80–85. [Google Scholar] [CrossRef]
  18. Nakamura, Y.; Fukami, K. Regulation and physiological functions of mammalian phospholipase C. J. Biochem. 2017, 161, 315–321. [Google Scholar] [CrossRef] [Green Version]
  19. Loscher, W. Critical review of current animal models of seizures and epilepsy used in the discovery and development of new antiepileptic drugs. Seizure 2011, 20, 359–368. [Google Scholar] [CrossRef] [Green Version]
  20. Kandratavicius, L.; Balista, P.A.; Lopes-Aguiar, C.; Ruggiero, R.N.; Umeoka, E.H.; Garcia-Cairasco, N.; Bueno-Junior, L.S.; Leite, J.P. Animal models of epilepsy: Use and limitations. Neuropsychiatr. Dis. Treat. 2014, 10, 1693–1705. [Google Scholar] [CrossRef] [Green Version]
  21. Grone, B.P.; Baraban, S.C. Animal models in epilepsy research: Legacies and new directions. Nat. Neurosci. 2015, 18, 339–343. [Google Scholar] [CrossRef]
  22. Loscher, W. Animal Models of Seizures and Epilepsy: Past, Present, and Future Role for the Discovery of Antiseizure Drugs. Neurochem. Res. 2017, 42, 1873–1888. [Google Scholar] [CrossRef] [PubMed]
  23. Juvale, I.I.A.; Che Has, A.T. The evolution of the pilocarpine animal model of status epilepticus. Heliyon 2020, 6, e04557. [Google Scholar] [CrossRef] [PubMed]
  24. Turski, W.A.; Cavalheiro, E.A.; Schwarz, M.; Czuczwar, S.J.; Kleinrok, Z.; Turski, L. Limbic seizures produced by pilocarpine in rats: Behavioural, electroencephalographic and neuropathological study. Behav. Brain Res. 1983, 9, 315–335. [Google Scholar] [CrossRef]
  25. Furtado Mde, A.; Braga, G.K.; Oliveira, J.A.; Del Vecchio, F.; Garcia-Cairasco, N. Behavioral, morphologic, and electroencephalographic evaluation of seizures induced by intrahippocampal microinjection of pilocarpine. Epilepsia 2002, 43 (Suppl. 5), 37–39. [Google Scholar] [CrossRef]
  26. Curia, G.; Longo, D.; Biagini, G.; Jones, R.S.; Avoli, M. The pilocarpine model of temporal lobe epilepsy. J. Neurosci. Methods 2008, 172, 143–157. [Google Scholar] [CrossRef] [PubMed]
  27. Ben-Ari, Y. Limbic seizure and brain damage produced by kainic acid: Mechanisms and relevance to human temporal lobe epilepsy. Neuroscience 1985, 14, 375–403. [Google Scholar] [CrossRef]
  28. Levesque, M.; Avoli, M. The kainic acid model of temporal lobe epilepsy. Neurosci. Biobehav. Rev. 2013, 37, 2887–2899. [Google Scholar] [CrossRef] [Green Version]
  29. Crepel, V.; Mulle, C. Physiopathology of kainate receptors in epilepsy. Curr. Opin. Pharmacol. 2015, 20, 83–88. [Google Scholar] [CrossRef]
  30. Goddard, G.V. Development of epileptic seizures through brain stimulation at low intensity. Nature 1967, 214, 1020–1021. [Google Scholar] [CrossRef]
  31. Goddard, G.V.; McIntyre, D.C.; Leech, C.K. A permanent change in brain function resulting from daily electrical stimulation. Exp. Neurol. 1969, 25, 295–330. [Google Scholar] [CrossRef]
  32. McNamara, J.O. Kindling model of epilepsy. Adv. Neurol. 1986, 44, 303–318. [Google Scholar] [PubMed]
  33. Marescaux, C.; Vergnes, M.; Depaulis, A. Genetic absence epilepsy in rats from Strasbourg--a review. J. Neural Transm. Suppl. 1992, 35, 37–69. [Google Scholar] [CrossRef] [PubMed]
  34. Chapman, A.G.; Croucher, M.J.; Meldrum, B.S. Evaluation of anticonvulsant drugs in DBA/2 mice with sound-induced seizures. Arzneimittelforschung 1984, 34, 1261–1264. [Google Scholar] [PubMed]
  35. De Sarro, G.; Russo, E.; Citraro, R.; Meldrum, B.S. Genetically epilepsy-prone rats (GEPRs) and DBA/2 mice: Two animal models of audiogenic reflex epilepsy for the evaluation of new generation AEDs. Epilepsy Behav. 2017, 71, 165–173. [Google Scholar] [CrossRef] [PubMed]
  36. Faingold, C.L.; Gehlbach, G.; Caspary, D.M. Decreased effectiveness of GABA-mediated inhibition in the inferior colliculus of the genetically epilepsy-prone rat. Exp. Neurol. 1986, 93, 145–159. [Google Scholar] [CrossRef]
  37. Jobe, P.C.; Dailey, J.W.; Reigel, C.E. Noradrenergic and serotonergic determinants of seizure susceptibility and severity in genetically epilepsy-prone rats. Life Sci. 1986, 39, 775–782. [Google Scholar] [CrossRef]
  38. Dailey, J.W.; Mishra, P.K.; Ko, K.H.; Penny, J.E.; Jobe, P.C. Noradrenergic abnormalities in the central nervous system of seizure-naive genetically epilepsy-prone rats. Epilepsia 1991, 32, 168–173. [Google Scholar] [CrossRef]
  39. Jobe, P.C.; Mishra, P.K.; Browning, R.A.; Wang, C.; Adams-Curtis, L.E.; Ko, K.H.; Dailey, J.W. Noradrenergic abnormalities in the genetically epilepsy-prone rat. Brain Res. Bull. 1994, 35, 493–504. [Google Scholar] [CrossRef]
  40. Evans, M.S.; Cady, C.J.; Disney, K.E.; Yang, L.; Laguardia, J.J. Three brief epileptic seizures reduce inhibitory synaptic currents, GABA(A) currents, and GABA(A)-receptor subunits. Epilepsia 2006, 47, 1655–1664. [Google Scholar] [CrossRef]
  41. Powell, K.L.; Cain, S.M.; Ng, C.; Sirdesai, S.; David, L.S.; Kyi, M.; Garcia, E.; Tyson, J.R.; Reid, C.A.; Bahlo, M.; et al. A Cav3.2 T-type calcium channel point mutation has splice-variant-specific effects on function and segregates with seizure expression in a polygenic rat model of absence epilepsy. J. Neurosci. 2009, 29, 371–380. [Google Scholar] [CrossRef] [PubMed]
  42. Coenen, A.M.; Van Luijtelaar, E.L. The WAG/Rij rat model for absence epilepsy: Age and sex factors. Epilepsy Res. 1987, 1, 297–301. [Google Scholar] [CrossRef]
  43. Coenen, A.M.; Drinkenburg, W.H.; Inoue, M.; van Luijtelaar, E.L. Genetic models of absence epilepsy, with emphasis on the WAG/Rij strain of rats. Epilepsy Res. 1992, 12, 75–86. [Google Scholar] [CrossRef]
  44. Loscher, W.; Klitgaard, H.; Twyman, R.E.; Schmidt, D. New avenues for anti-epileptic drug discovery and development. Nat. Rev. Drug Discov. 2013, 12, 757–776. [Google Scholar] [CrossRef] [PubMed]
  45. Vining, E.P. Use of barbiturates and benzodiazepines in treatment of epilepsy. Neurol. Clin. 1986, 4, 617–632. [Google Scholar] [CrossRef]
  46. Rogawski, M.A.; Loscher, W. The neurobiology of antiepileptic drugs. Nat. Rev. Neurosci. 2004, 5, 553–564. [Google Scholar] [CrossRef] [Green Version]
  47. Sills, G.J.; Rogawski, M.A. Mechanisms of action of currently used antiseizure drugs. Neuropharmacology 2020, 168, 107966. [Google Scholar] [CrossRef]
  48. Pal, D.K. Phenobarbital for childhood epilepsy: Systematic review. Paediatr. Perinat Drug Ther. 2006, 7, 31–42. [Google Scholar] [CrossRef] [Green Version]
  49. Zhang, L.L.; Zeng, L.N.; Li, Y.P. Side effects of phenobarbital in epilepsy: A systematic review. Epileptic Disord. 2011, 13, 349–365. [Google Scholar] [CrossRef]
  50. Jung, M.J.; Lippert, B.; Metcalf, B.W.; Schechter, P.J.; Bohlen, P.; Sjoerdsma, A. The effect of 4-amino hex-5-ynoic acid (gamma-acetylenic GABA, gammma-ethynyl GABA) a catalytic inhibitor of GABA transaminase, on brain GABA metabolism in vivo. J. Neurochem. 1977, 28, 717–723. [Google Scholar] [CrossRef]
  51. Loscher, W. Effect of inhibitors of GABA transaminase on the synthesis, binding, uptake, and metabolism of GABA. J. Neurochem. 1980, 34, 1603–1608. [Google Scholar] [CrossRef]
  52. Bauer, J.; Cooper-Mahkorn, D. Tiagabine: Efficacy and safety in partial seizures—Current status. Neuropsychiatr. Dis. Treat. 2008, 4, 731–736. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Trepanier, L.A.; Van Schoick, A.; Schwark, W.S.; Carrillo, J. Therapeutic serum drug concentrations in epileptic dogs treated with potassium bromide alone or in combination with oTher. anticonvulsants: 122 cases (1992–1996). J. Am. Vet. Med. Assoc. 1998, 213, 1449–1453. [Google Scholar] [PubMed]
  54. Korinthenberg, R.; Burkart, P.; Woelfle, C.; Moenting, J.S.; Ernst, J.P. Pharmacology, efficacy, and tolerability of potassium bromide in childhood epilepsy. J. Child. Neurol. 2007, 22, 414–418. [Google Scholar] [CrossRef] [PubMed]
  55. Loscher, W.; Rogawski, M.A. How theories evolved concerning the mechanism of action of barbiturates. Epilepsia 2012, 53 (Suppl. 8), 12–25. [Google Scholar] [CrossRef] [PubMed]
  56. Schwartz-Porsche, D.; Loscher, W.; Frey, H.H. Therapeutic efficacy of phenobarbital and primidone in canine epilepsy: A comparison. J. Vet. Pharmacol. Ther. 1985, 8, 113–119. [Google Scholar] [CrossRef] [PubMed]
  57. Appleton, R.; Sweeney, A.; Choonara, I.; Robson, J.; Molyneux, E. Lorazepam versus diazepam in the acute treatment of epileptic seizures and status epilepticus. Dev. Med. Child. Neurol. 1995, 37, 682–688. [Google Scholar] [CrossRef]
  58. Ochoa, J.G.; Kilgo, W.A. The Role of Benzodiazepines in the Treatment of Epilepsy. Curr. Treat. Options Neurol. 2016, 18, 18. [Google Scholar] [CrossRef]
  59. Nicholson, M.W.; Sweeney, A.; Pekle, E.; Alam, S.; Ali, A.B.; Duchen, M.; Jovanovic, J.N. Diazepam-induced loss of inhibitory synapses mediated by PLCdelta/Ca(2+)/calcineurin signalling downstream of GABAA receptors. Mol. Psychiatry 2018, 23, 1851–1867. [Google Scholar] [CrossRef]
  60. Pinder, R.M.; Brogden, R.N.; Speight, T.M.; Avery, G.S. Sodium valproate: A review of its pharmacological properties and therapeutic efficacy in epilepsy. Drugs 1977, 13, 81–123. [Google Scholar] [CrossRef]
  61. Baldino, F., Jr.; Geller, H.M. Sodium valproate enhancement of gamma-aminobutyric acid (GABA) inhibition: Electrophysiological evidence for anticonvulsant activity. J. Pharmacol. Exp. Ther. 1981, 217, 445–450. [Google Scholar] [PubMed]
  62. Bang, F.; Birket-Smith, E.; Mikkelsen, B. Clonazepam in the treatment of epilepsy. A clinical long-term follow-up study. Epilepsia 1976, 17, 321–324. [Google Scholar] [CrossRef] [PubMed]
  63. Obeid, T.; Panayiotopoulos, C.P. Clonazepam in juvenile myoclonic epilepsy. Epilepsia 1989, 30, 603–606. [Google Scholar] [CrossRef]
  64. Livingston, J.H.; Beaumont, D.; Arzimanoglou, A.; Aicardi, J. Vigabatrin in the treatment of epilepsy in children. Br. J. Clin. Pharmacol. 1989, 27 (Suppl. 1), 109S–112S. [Google Scholar] [CrossRef] [Green Version]
  65. Angehagen, M.; Ben-Menachem, E.; Ronnback, L.; Hansson, E. Novel mechanisms of action of three antiepileptic drugs, vigabatrin, tiagabine, and topiramate. Neurochem. Res. 2003, 28, 333–340. [Google Scholar] [CrossRef] [PubMed]
  66. Xue, M.; Atallah, B.V.; Scanziani, M. Equalizing excitation-inhibition ratios across visual cortical neurons. Nature 2014, 511, 596–600. [Google Scholar] [CrossRef]
  67. Zhou, S.; Yu, Y. Synaptic E-I Balance Underlies Efficient Neural Coding. Front. Neurosci. 2018, 12, 46. [Google Scholar] [CrossRef] [Green Version]
  68. Erecinska, M.; Silver, I.A. Metabolism and role of glutamate in mammalian brain. Prog. Neurobiol. 1990, 35, 245–296. [Google Scholar] [CrossRef]
  69. Jones, E.G. GABAergic neurons and their role in cortical plasticity in primates. Cereb. Cortex 1993, 3, 361–372. [Google Scholar] [CrossRef]
  70. Esclapez, M.; Houser, C.R. Up-regulation of GAD65 and GAD67 in remaining hippocampal GABA neurons in a model of temporal lobe epilepsy. J. Comp. Neurol. 1999, 412, 488–505. [Google Scholar] [CrossRef]
  71. Sperk, G.; Wieselthaler-Holzl, A.; Pirker, S.; Tasan, R.; Strasser, S.S.; Drexel, M.; Pifl, C.; Marschalek, J.; Ortler, M.; Trinka, E.; et al. Glutamate decarboxylase 67 is expressed in hippocampal mossy fibers of temporal lobe epilepsy patients. Hippocampus 2012, 22, 590–603. [Google Scholar] [CrossRef] [Green Version]
  72. Esclapez, M.; Tillakaratne, N.J.; Kaufman, D.L.; Tobin, A.J.; Houser, C.R. Comparative localization of two forms of glutamic acid decarboxylase and their mRNAs in rat brain supports the concept of functional differences between the forms. J. Neurosci. 1994, 14, 1834–1855. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Asada, H.; Kawamura, Y.; Maruyama, K.; Kume, H.; Ding, R.; Ji, F.Y.; Kanbara, N.; Kuzume, H.; Sanbo, M.; Yagi, T.; et al. Mice lacking the 65 kDa isoform of glutamic acid decarboxylase (GAD65) maintain normal levels of GAD67 and GABA in their brains but are susceptible to seizures. Biochem. Biophys. Res. Commun. 1996, 229, 891–895. [Google Scholar] [CrossRef] [PubMed]
  74. Kash, S.F.; Johnson, R.S.; Tecott, L.H.; Noebels, J.L.; Mayfield, R.D.; Hanahan, D.; Baekkeskov, S. Epilepsy in mice deficient in the 65-kDa isoform of glutamic acid decarboxylase. Proc. Natl. Acad. Sci. USA 1997, 94, 14060–14065. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Freichel, C.; Potschka, H.; Ebert, U.; Brandt, C.; Loscher, W. Acute changes in the neuronal expression of GABA and glutamate decarboxylase isoforms in the rat piriform cortex following status epilepticus. Neuroscience 2006, 141, 2177–2194. [Google Scholar] [CrossRef] [PubMed]
  76. Asada, H.; Kawamura, Y.; Maruyama, K.; Kume, H.; Ding, R.G.; Kanbara, N.; Kuzume, H.; Sanbo, M.; Yagi, T.; Obata, K. Cleft palate and decreased brain gamma-aminobutyric acid in mice lacking the 67-kDa isoform of glutamic acid decarboxylase. Proc. Natl. Acad. Sci. USA 1997, 94, 6496–6499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Shimazaki, K.; Kobari, T.; Oguro, K.; Yokota, H.; Kasahara, Y.; Murashima, Y.; Watanabe, E.; Kawai, K.; Okada, T. Hippocampal GAD67 Transduction Using rAAV8 Regulates Epileptogenesis in EL Mice. Mol. Ther. Methods Clin. Dev. 2019, 13, 180–186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Kang, T.C.; An, S.J.; Park, S.K.; Hwang, I.K.; Bae, J.C.; Won, M.H. Changed vesicular GABA transporter immunoreactivity in the gerbil hippocampus following spontaneous seizure and vigabatrin administration. Neurosci. Lett. 2003, 335, 207–211. [Google Scholar] [CrossRef]
  79. Jiang, K.W.; Gao, F.; Shui, Q.X.; Yu, Z.S.; Xia, Z.Z. Effect of diazoxide on regulation of vesicular and plasma membrane GABA transporter genes and proteins in hippocampus of rats subjected to picrotoxin-induced kindling. Neurosci. Res. 2004, 50, 319–329. [Google Scholar] [CrossRef]
  80. Kang, T.C.; Kim, D.S.; Kwak, S.E.; Kim, J.E.; Won, M.H.; Kim, D.W.; Choi, S.Y.; Kwon, O.S. Epileptogenic roles of astroglial death and regeneration in the dentate gyrus of experimental temporal lobe epilepsy. Glia 2006, 54, 258–271. [Google Scholar] [CrossRef]
  81. Boulland, J.L.; Jenstad, M.; Boekel, A.J.; Wouterlood, F.G.; Edwards, R.H.; Storm-Mathisen, J.; Chaudhry, F.A. Vesicular glutamate and GABA transporters sort to distinct sets of vesicles in a population of presynaptic terminals. Cereb. Cortex 2009, 19, 241–248. [Google Scholar] [CrossRef] [PubMed]
  82. Saito, K.; Kakizaki, T.; Hayashi, R.; Nishimaru, H.; Furukawa, T.; Nakazato, Y.; Takamori, S.; Ebihara, S.; Uematsu, M.; Mishina, M.; et al. The physiological roles of vesicular GABA transporter during embryonic development: A study using knockout mice. Mol. Brain 2010, 3, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Zhou, F.W.; Roper, S.N. Densities of glutamatergic and GABAergic presynaptic terminals are altered in experimental cortical dysplasia. Epilepsia 2010, 51, 1468–1476. [Google Scholar] [CrossRef]
  84. Van Liefferinge, J.; Massie, A.; Portelli, J.; Di Giovanni, G.; Smolders, I. Are vesicular neurotransmitter transporters potential treatment targets for temporal lobe epilepsy? Front. Cell Neurosci. 2013, 7, 139. [Google Scholar] [CrossRef] [Green Version]
  85. Boulland, J.L.; Ferhat, L.; Tallak Solbu, T.; Ferrand, N.; Chaudhry, F.A.; Storm-Mathisen, J.; Esclapez, M. Changes in vesicular transporters for gamma-aminobutyric acid and glutamate reveal vulnerability and reorganization of hippocampal neurons following pilocarpine-induced seizures. J. Comp. Neurol. 2007, 503, 466–485. [Google Scholar] [CrossRef]
  86. Carvill, G.L.; McMahon, J.M.; Schneider, A.; Zemel, M.; Myers, C.T.; Saykally, J.; Nguyen, J.; Robbiano, A.; Zara, F.; Specchio, N.; et al. Mutations in the GABA Transporter SLC6A1 Cause Epilepsy with Myoclonic-Atonic Seizures. Am. J. Hum. Genet. 2015, 96, 808–815. [Google Scholar] [CrossRef] [Green Version]
  87. Pirttimaki, T.; Parri, H.R.; Crunelli, V. Astrocytic GABA transporter GAT-1 dysfunction in experimental absence seizures. J. Physiol. 2013, 591, 823–833. [Google Scholar] [CrossRef] [PubMed]
  88. Jacob, T.C.; Moss, S.J.; Jurd, R. GABA(A) receptor trafficking and its role in the dynamic modulation of neuronal inhibition. Nat. Rev. Neurosci. 2008, 9, 331–343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Sigel, E.; Steinmann, M.E. Structure, function, and modulation of GABA(A) receptors. J. Biol. Chem. 2012, 287, 40224–40231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Engin, E.; Benham, R.S.; Rudolph, U. An Emerging Circuit Pharmacology of GABAA Receptors. Trends Pharmacol. Sci. 2018, 39, 710–732. [Google Scholar] [CrossRef]
  91. Bettler, B.; Kaupmann, K.; Mosbacher, J.; Gassmann, M. Molecular structure and physiological functions of GABA(B) receptors. Physiol. Rev. 2004, 84, 835–867. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Gassmann, M.; Bettler, B. Regulation of neuronal GABA(B) receptor functions by subunit composition. Nat. Rev. Neurosci. 2012, 13, 380–394. [Google Scholar] [CrossRef]
  93. Terunuma, M. Diversity of structure and function of GABAB receptors: A complexity of GABAB-mediated signaling. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2018, 94, 390–411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Kim, Y.S.; Yoon, B.E. Altered GABAergic Signaling in Brain Disease at Various Stages of Life. Exp. Neurobiol. 2017, 26, 122–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Fatemi, S.H.; Folsom, T.D.; Thuras, P.D. Deficits in GABA(B) receptor system in schizophrenia and mood disorders: A postmortem study. Schizophr. Res. 2011, 128, 37–43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Cossette, P.; Liu, L.; Brisebois, K.; Dong, H.; Lortie, A.; Vanasse, M.; Saint-Hilaire, J.M.; Carmant, L.; Verner, A.; Lu, W.Y.; et al. Mutation of GABRA1 in an autosomal dominant form of juvenile myoclonic epilepsy. Nat. Genet. 2002, 31, 184–189. [Google Scholar] [CrossRef] [PubMed]
  97. Butler, K.M.; Moody, O.A.; Schuler, E.; Coryell, J.; Alexander, J.J.; Jenkins, A.; Escayg, A. De novo variants in GABRA2 and GABRA5 alter receptor function and contribute to early-onset epilepsy. Brain 2018, 141, 2392–2405. [Google Scholar] [CrossRef]
  98. Wallace, R.H.; Marini, C.; Petrou, S.; Harkin, L.A.; Bowser, D.N.; Panchal, R.G.; Williams, D.A.; Sutherland, G.R.; Mulley, J.C.; Scheffer, I.E.; et al. Mutant GABA(A) receptor gamma2-subunit in childhood absence epilepsy and febrile seizures. Nat. Genet. 2001, 28, 49–52. [Google Scholar] [CrossRef] [PubMed]
  99. DeLorey, T.M.; Handforth, A.; Anagnostaras, S.G.; Homanics, G.E.; Minassian, B.A.; Asatourian, A.; Fanselow, M.S.; Delgado-Escueta, A.; Ellison, G.D.; Olsen, R.W. Mice lacking the beta3 subunit of the GABAA receptor have the epilepsy phenotype and many of the behavioral characteristics of Angelman syndrome. J. Neurosci. 1998, 18, 8505–8514. [Google Scholar] [CrossRef] [Green Version]
  100. Bouilleret, V.; Loup, F.; Kiener, T.; Marescaux, C.; Fritschy, J.M. Early loss of interneurons and delayed subunit-specific changes in GABA(A)-receptor expression in a mouse model of mesial temporal lobe epilepsy. Hippocampus 2000, 10, 305–324. [Google Scholar] [CrossRef]
  101. Houser, C.R.; Esclapez, M. Downregulation of the alpha5 subunit of the GABA(A) receptor in the pilocarpine model of temporal lobe epilepsy. Hippocampus 2003, 13, 633–645. [Google Scholar] [CrossRef]
  102. Prosser, H.M.; Gill, C.H.; Hirst, W.D.; Grau, E.; Robbins, M.; Calver, A.; Soffin, E.M.; Farmer, C.E.; Lanneau, C.; Gray, J.; et al. Epileptogenesis and enhanced prepulse inhibition in GABA(B1)-deficient mice. Mol. Cell Neurosci. 2001, 17, 1059–1070. [Google Scholar] [CrossRef] [PubMed]
  103. Gambardella, A.; Manna, I.; Labate, A.; Chifari, R.; La Russa, A.; Serra, P.; Cittadella, R.; Bonavita, S.; Andreoli, V.; LePiane, E.; et al. GABA(B) receptor 1 polymorphism (G1465A) is associated with temporal lobe epilepsy. Neurology 2003, 60, 560–563. [Google Scholar] [CrossRef] [PubMed]
  104. Chandler, K.E.; Princivalle, A.P.; Fabian-Fine, R.; Bowery, N.G.; Kullmann, D.M.; Walker, M.C. Plasticity of GABA(B) receptor-mediated heterosynaptic interactions at mossy fibers after status epilepticus. J. Neurosci. 2003, 23, 11382–11391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Schuler, V.; Luscher, C.; Blanchet, C.; Klix, N.; Sansig, G.; Klebs, K.; Schmutz, M.; Heid, J.; Gentry, C.; Urban, L.; et al. Epilepsy, hyperalgesia, impaired memory, and loss of pre- and postsynaptic GABA(B) responses in mice lacking GABA(B(1)). Neuron 2001, 31, 47–58. [Google Scholar] [CrossRef] [Green Version]
  106. Brodie, M.J.; Besag, F.; Ettinger, A.B.; Mula, M.; Gobbi, G.; Comai, S.; Aldenkamp, A.P.; Steinhoff, B.J.; Sibley, D.R. Epilepsy, Antiepileptic Drugs, and Aggression: An Evidence-Based Review. Pharmacol. Rev. 2016, 68, 563–602. [Google Scholar] [CrossRef]
  107. Greenfield, L.J., Jr. Molecular mechanisms of antiseizure drug activity at GABAA receptors. Seizure 2013, 22, 589–600. [Google Scholar] [CrossRef] [Green Version]
  108. Campo-Soria, C.; Chang, Y.; Weiss, D.S. Mechanism of action of benzodiazepines on GABAA receptors. Br. J. Pharmacol. 2006, 148, 984–990. [Google Scholar] [CrossRef] [Green Version]
  109. Nutt, D.J.; Cowen, P.J.; Batts, C.C.; Grahame-Smith, D.G.; Green, A.R. Repeated administration of subconvulsant doses of GABA antagonist drugs. I. Effect on seizure threshold (kindling). Psychopharmacology (Berlin) 1982, 76, 84–87. [Google Scholar] [CrossRef]
  110. Allen, J.A.; Yadav, P.N.; Roth, B.L. Insights into the regulation of 5-HT2A serotonin receptors by scaffolding proteins and kinases. Neuropharmacology 2008, 55, 961–968. [Google Scholar] [CrossRef] [Green Version]
  111. Gerber, U.; Gee, C.E.; Benquet, P. Metabotropic glutamate receptors: Intracellular signaling pathways. Curr. Opin. Pharmacol. 2007, 7, 56–61. [Google Scholar] [CrossRef] [PubMed]
  112. Felder, C.C. Muscarinic acetylcholine receptors: Signal transduction through multiple effectors. FASEB J. 1995, 9, 619–625. [Google Scholar] [CrossRef] [PubMed]
  113. McOmish, C.E.; Burrows, E.L.; Howard, M.; Hannan, A.J. PLC-beta1 knockout mice as a model of disrupted cortical development and plasticity: Behavioral endophenotypes and dysregulation of RGS4 gene expression. Hippocampus 2008, 18, 824–834. [Google Scholar] [CrossRef] [PubMed]
  114. Xie, W.; Samoriski, G.M.; McLaughlin, J.P.; Romoser, V.A.; Smrcka, A.; Hinkle, P.M.; Bidlack, J.M.; Gross, R.A.; Jiang, H.; Wu, D. Genetic alteration of phospholipase C beta3 expression modulates behavioral and cellular responses to mu opioids. Proc. Natl. Acad. Sci. USA 1999, 96, 10385–10390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Jiang, H.; Lyubarsky, A.; Dodd, R.; Vardi, N.; Pugh, E.; Baylor, D.; Simon, M.I.; Wu, D. Phospholipase C beta 4 is involved in modulating the visual response in mice. Proc. Natl. Acad. Sci. USA 1996, 93, 14598–14601. [Google Scholar] [CrossRef] [Green Version]
  116. Hashimoto, K.; Miyata, M.; Watanabe, M.; Kano, M. Roles of phospholipase Cbeta4 in synapse elimination and plasticity in developing and mature cerebellum. Mol. Neurobiol. 2001, 23, 69–82. [Google Scholar] [CrossRef]
  117. Miyata, M.; Kashiwadani, H.; Fukaya, M.; Hayashi, T.; Wu, D.; Suzuki, T.; Watanabe, M.; Kawakami, Y. Role of thalamic phospholipase C[beta]4 mediated by metabotropic glutamate receptor type 1 in inflammatory pain. J. Neurosci. 2003, 23, 8098–8108. [Google Scholar] [CrossRef] [Green Version]
  118. Lawrence, J.J. Cholinergic control of GABA release: Emerging parallels between neocortex and hippocampus. Trends Neurosci. 2008, 31, 317–327. [Google Scholar] [CrossRef]
  119. Hamilton, S.E.; Loose, M.D.; Qi, M.; Levey, A.I.; Hille, B.; McKnight, G.S.; Idzerda, R.L.; Nathanson, N.M. Disruption of the m1 receptor gene ablates muscarinic receptor-dependent M current regulation and seizure activity in mice. Proc. Natl. Acad. Sci. USA 1997, 94, 13311–13316. [Google Scholar] [CrossRef] [Green Version]
  120. Salgado, H.; Bellay, T.; Nichols, J.A.; Bose, M.; Martinolich, L.; Perrotti, L.; Atzori, M. Muscarinic M2 and M1 receptors reduce GABA release by Ca2+ channel modulation through activation of PI3K/Ca2+ -independent and PLC/Ca2+ -dependent PKC. J. Neurophysiol. 2007, 98, 952–965. [Google Scholar] [CrossRef]
  121. Rodriguez-Moreno, A.; Lerma, J. Kainate receptor modulation of GABA release involves a metabotropic function. Neuron 1998, 20, 1211–1218. [Google Scholar] [CrossRef]
  122. Caiati, M.D.; Sivakumaran, S.; Cherubini, E. In the developing rat hippocampus, endogenous activation of presynaptic kainate receptors reduces GABA release from mossy fiber terminals. J Neurosci. 2010, 30, 1750–1759. [Google Scholar] [CrossRef]
  123. Jijon-Lorenzo, R.; Caballero-Floran, I.H.; Recillas-Morales, S.; Cortes, H.; Avalos-Fuentes, J.A.; Paz-Bermudez, F.J.; Erlij, D.; Floran, B. Presynaptic Dopamine D2 Receptors Modulate [(3)H]GABA Release at StriatoPallidal Terminals via Activation of PLC->IP3->Calcineurin and Inhibition of AC->cAMP->PKA Signaling Cascades. Neuroscience 2018, 372, 74–86. [Google Scholar] [CrossRef] [PubMed]
  124. Kelm, M.K.; Weinberg, R.J.; Criswell, H.E.; Breese, G.R. The PLC/IP 3 R/PKC pathway is required for ethanol-enhanced GABA release. Neuropharmacology 2010, 58, 1179–1186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Goitia, B.; Rivero-Echeto, M.C.; Weisstaub, N.V.; Gingrich, J.A.; Garcia-Rill, E.; Bisagno, V.; Urbano, F.J. Modulation of GABA release from the thalamic reticular nucleus by cocaine and caffeine: Role of serotonin receptors. J. Neurochem. 2016, 136, 526–535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Millan, M.J.; Marin, P.; Bockaert, J.; Mannoury la Cour, C. Signaling at G-protein-coupled serotonin receptors: Recent advances and future research directions. Trends Pharmacol. Sci. 2008, 29, 454–464. [Google Scholar] [CrossRef]
  127. Kurian, M.A.; Meyer, E.; Vassallo, G.; Morgan, N.V.; Prakash, N.; Pasha, S.; Hai, N.A.; Shuib, S.; Rahman, F.; Wassmer, E.; et al. Phospholipase C beta 1 deficiency is associated with early-onset epileptic encephalopathy. Brain 2010, 133, 2964–2970. [Google Scholar] [CrossRef] [Green Version]
  128. Poduri, A.; Chopra, S.S.; Neilan, E.G.; Elhosary, P.C.; Kurian, M.A.; Meyer, E.; Barry, B.J.; Khwaja, O.S.; Salih, M.A.; Stodberg, T.; et al. Homozygous PLCB1 deletion associated with malignant migrating partial seizures in infancy. Epilepsia 2012, 53, e146–e150. [Google Scholar] [CrossRef] [Green Version]
  129. Schoonjans, A.S.; Meuwissen, M.; Reyniers, E.; Kooy, F.; Ceulemans, B. PLCB1 epileptic encephalopathies; Review and expansion of the phenotypic spectrum. Eur. J. Paediatr. Neurol. 2016, 20, 474–479. [Google Scholar] [CrossRef]
  130. Desprairies, C.; Valence, S.; Maurey, H.; Helal, S.I.; Weckhuysen, S.; Soliman, H.; Mefford, H.C.; Spentchian, M.; Heron, D.; Leguern, E.; et al. Three novel patients with epileptic encephalopathy due to biallelic mutations in the PLCB1 gene. Clin. Genet. 2020, 97, 477–482. [Google Scholar] [CrossRef]
  131. Kim, D.; Jun, K.S.; Lee, S.B.; Kang, N.G.; Min, D.S.; Kim, Y.H.; Ryu, S.H.; Suh, P.G.; Shin, H.S. Phospholipase C isozymes selectively couple to specific neurotransmitter receptors. Nature 1997, 389, 290–293. [Google Scholar] [CrossRef]
  132. Bohm, D.; Schwegler, H.; Kotthaus, L.; Nayernia, K.; Rickmann, M.; Kohler, M.; Rosenbusch, J.; Engel, W.; Flugge, G.; Burfeind, P. Disruption of PLC-beta 1-mediated signal transduction in mutant mice causes age-dependent hippocampal mossy fiber sprouting and neurodegeneration. Mol. Cell Neurosci. 2002, 21, 584–601. [Google Scholar] [CrossRef]
  133. Cheong, E.; Zheng, Y.; Lee, K.; Lee, J.; Kim, S.; Sanati, M.; Lee, S.; Kim, Y.S.; Shin, H.S. Deletion of phospholipase C beta4 in thalamocortical relay nucleus leads to absence seizures. Proc. Natl. Acad. Sci. USA 2009, 106, 21912–21917. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Liu, J.X.; Tang, Y.C.; Liu, Y.; Tang, F.R. mGluR5-PLCbeta4-PKCbeta2/PKCgamma pathways in hippocampal CA1 pyramidal neurons in pilocarpine model of status epilepticus in mGluR5+/+ mice. Epilepsy Res. 2008, 82, 111–123. [Google Scholar] [CrossRef] [PubMed]
  135. Lee, S.B.; Oh, Y.J.; Chung, J.K.; Jeong, J.H.; Lee, S.D.; Park, D.K.; Park, K.H.; Ko, J.S.; Kim, D.S. Altered PLCbeta-1 expression in the gerbil hippocampal complex following spontaneous seizure. BMB Rep. 2011, 44, 566–571. [Google Scholar] [CrossRef] [Green Version]
  136. Liu, J.X.; Hu, M.; Chen, X.L.; Xu, J.H.; Yang, P.B.; Zhang, J.S.; Liu, Y. Reduced expression of Phospholipase C beta in hippocampal interneuron during pilocarpine induced status epilepticus in mice. Neurochem. Int. 2014, 68, 10–17. [Google Scholar] [CrossRef]
  137. He, X.P.; Pan, E.; Sciarretta, C.; Minichiello, L.; McNamara, J.O. Disruption of TrkB-mediated phospholipase Cgamma signaling inhibits limbic epileptogenesis. J. Neurosci. 2010, 30, 6188–6196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. He, X.P.; Wen, R.; McNamara, J.O. Impairment of kindling development in phospholipase Cgamma1 heterozygous mice. Epilepsia 2014, 55, 456–463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Gu, B.; Huang, Y.Z.; He, X.P.; Joshi, R.B.; Jang, W.; McNamara, J.O. A Peptide Uncoupling BDNF Receptor TrkB from Phospholipase Cgamma1 Prevents Epilepsy Induced by Status Epilepticus. Neuron 2015, 88, 484–491. [Google Scholar] [CrossRef] [Green Version]
  140. Kim, H.Y.; Yang, Y.R.; Hwang, H.; Lee, H.E.; Jang, H.J.; Kim, J.; Yang, E.; Kim, H.; Rhim, H.; Suh, P.G.; et al. Deletion of PLCgamma1 in GABAergic neurons increases seizure susceptibility in aged mice. Sci. Rep. 2019, 9, 17761. [Google Scholar] [CrossRef] [Green Version]
  141. Rellahan, B.L.; Graham, L.J.; Tysgankov, A.Y.; DeBell, K.E.; Veri, M.C.; Noviello, C.; Bonvini, E. A dynamic constitutive and inducible binding of c-Cbl by PLCgamma1 SH3 and SH2 domains (negatively) regulates antigen receptor-induced PLCgamma1 activation in lymphocytes. Exp. Cell Res. 2003, 289, 184–194. [Google Scholar] [CrossRef]
  142. Kim, M.J.; Chang, J.S.; Park, S.K.; Hwang, J.I.; Ryu, S.H.; Suh, P.G. Direct interaction of SOS1 Ras exchange protein with the SH3 domain of phospholipase C-gamma1. Biochemistry 2000, 39, 8674–8682. [Google Scholar] [CrossRef] [PubMed]
  143. Jang, H.J.; Yang, Y.R.; Kim, J.K.; Choi, J.H.; Seo, Y.K.; Lee, Y.H.; Lee, J.E.; Ryu, S.H.; Suh, P.G. Phospholipase C-gamma1 involved in brain disorders. Adv. Biol. Regul. 2013, 53, 51–62. [Google Scholar] [CrossRef] [PubMed]
  144. Turecki, G.; Grof, P.; Cavazzoni, P.; Duffy, A.; Grof, E.; Ahrens, B.; Berghofer, A.; Muller-Oerlinghausen, B.; Dvorakova, M.; Libigerova, E.; et al. Evidence for a role of phospholipase C-gamma1 in the pathogenesis of bipolar disorder. Mol. Psychiatry 1998, 3, 534–538. [Google Scholar] [CrossRef] [Green Version]
  145. Lovlie, R.; Berle, J.O.; Stordal, E.; Steen, V.M. The phospholipase C-gamma1 gene (PLCG1) and lithium-responsive bipolar disorder: Re-examination of an intronic dinucleotide repeat polymorphism. Psychiatr. Genet. 2001, 11, 41–43. [Google Scholar] [CrossRef] [PubMed]
  146. Yang, Y.R.; Jung, J.H.; Kim, S.J.; Hamada, K.; Suzuki, A.; Kim, H.J.; Lee, J.H.; Kwon, O.B.; Lee, Y.K.; Kim, J.; et al. Forebrain-specific ablation of phospholipase Cgamma1 causes manic-like behavior. Mol. Psychiatry 2017, 22, 1473–1482. [Google Scholar] [CrossRef] [PubMed]
  147. Kang, D.S.; Yang, Y.R.; Lee, C.; Park, B.; Park, K.I.; Seo, J.K.; Seo, Y.K.; Cho, H.; Lucio, C.; Suh, P.G. Netrin-1/DCC-mediated PLCgamma1 activation is required for axon guidance and brain structure development. EMBO Rep. 2018, 19. [Google Scholar] [CrossRef] [PubMed]
  148. Vaz, S.H.; Cristovao-Ferreira, S.; Ribeiro, J.A.; Sebastiao, A.M. Brain-derived neurotrophic factor inhibits GABA uptake by the rat hippocampal nerve terminals. Brain Res. 2008, 1219, 19–25. [Google Scholar] [CrossRef] [PubMed]
  149. Gomeza, J.; Casado, M.; Gimenez, C.; Aragon, C. Inhibition of high-affinity gamma-aminobutyric acid uptake in primary astrocyte cultures by phorbol esters and phospholipase C. Biochem. J. 1991, 275 Pt 2, 435–439. [Google Scholar] [CrossRef] [Green Version]
  150. Binder, D.K.; Croll, S.D.; Gall, C.M.; Scharfman, H.E. BDNF and epilepsy: Too much of a good thing? Trends Neurosci. 2001, 24, 47–53. [Google Scholar] [CrossRef]
  151. Scharfman, H.E. Brain-derived neurotrophic factor and epilepsy—A missing link? Epilepsy Curr. 2005, 5, 83–88. [Google Scholar] [CrossRef]
  152. Tanaka, T.; Saito, H.; Matsuki, N. Inhibition of GABAA synaptic responses by brain-derived neurotrophic factor (BDNF) in rat hippocampus. J. Neurosci. 1997, 17, 2959–2966. [Google Scholar] [CrossRef]
  153. Brunig, I.; Penschuck, S.; Berninger, B.; Benson, J.; Fritschy, J.M. BDNF reduces miniature inhibitory postsynaptic currents by rapid downregulation of GABA(A) receptor surface expression. Eur. J. Neurosci. 2001, 13, 1320–1328. [Google Scholar] [CrossRef] [PubMed]
  154. Cheng, Q.; Yeh, H.H. PLCgamma signaling underlies BDNF potentiation of Purkinje cell responses to GABA. J. Neurosci. Res. 2005, 79, 616–627. [Google Scholar] [CrossRef] [PubMed]
  155. Yamada, M.K.; Nakanishi, K.; Ohba, S.; Nakamura, T.; Ikegaya, Y.; Nishiyama, N.; Matsuki, N. Brain-derived neurotrophic factor promotes the maturation of GABAergic mechanisms in cultured hippocampal neurons. J. Neurosci. 2002, 22, 7580–7585. [Google Scholar] [CrossRef] [Green Version]
  156. Mizoguchi, Y.; Ishibashi, H.; Nabekura, J. The action of BDNF on GABA(A) currents changes from potentiating to suppressing during maturation of rat hippocampal CA1 pyramidal neurons. J. Physiol. 2003, 548, 703–709. [Google Scholar] [CrossRef] [PubMed]
  157. Mizoguchi, Y.; Kanematsu, T.; Hirata, M.; Nabekura, J. A rapid increase in the total number of cell surface functional GABAA receptors induced by brain-derived neurotrophic factor in rat visual cortex. J. Biol. Chem. 2003, 278, 44097–44102. [Google Scholar] [CrossRef] [Green Version]
  158. Fiorentino, H.; Kuczewski, N.; Diabira, D.; Ferrand, N.; Pangalos, M.N.; Porcher, C.; Gaiarsa, J.L. GABA(B) receptor activation triggers BDNF release and promotes the maturation of GABAergic synapses. J. Neurosci. 2009, 29, 11650–11661. [Google Scholar] [CrossRef] [PubMed]
  159. Kuczewski, N.; Fuchs, C.; Ferrand, N.; Jovanovic, J.N.; Gaiarsa, J.L.; Porcher, C. Mechanism of GABAB receptor-induced BDNF secretion and promotion of GABAA receptor membrane expression. J. Neurochem. 2011, 118, 533–545. [Google Scholar] [CrossRef]
  160. Cheng, Z.Y.; Wang, X.P.; Schmid, K.L.; Han, X.G. GABAB1 and GABAB2 receptor subunits co-expressed in cultured human RPE cells regulate intracellular Ca2+ via Gi/o-protein and phospholipase C pathways. Neuroscience 2014, 280, 254–261. [Google Scholar] [CrossRef]
  161. Tu, H.; Xu, C.; Zhang, W.; Liu, Q.; Rondard, P.; Pin, J.P.; Liu, J. GABAB receptor activation protects neurons from apoptosis via IGF-1 receptor transactivation. J. Neurosci. 2010, 30, 749–759. [Google Scholar] [CrossRef] [PubMed]
  162. Blaesse, P.; Airaksinen, M.S.; Rivera, C.; Kaila, K. Cation-chloride cotransporters and neuronal function. Neuron 2009, 61, 820–838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Mahadevan, V.; Woodin, M.A. Regulation of neuronal chloride homeostasis by neuromodulators. J. Physiol. 2016, 594, 2593–2605. [Google Scholar] [CrossRef]
  164. Liu, R.; Wang, J.; Liang, S.; Zhang, G.; Yang, X. Role of NKCC1 and KCC2 in Epilepsy: From Expression to Function. Front. Neurol. 2019, 10, 1407. [Google Scholar] [CrossRef]
  165. Duy, P.Q.; David, W.B.; Kahle, K.T. Identification of KCC2 Mutations in Human Epilepsy Suggests Strategies for Therapeutic Transporter Modulation. Front. Cell Neurosci. 2019, 13, 515. [Google Scholar] [CrossRef]
  166. Rivera, C.; Voipio, J.; Thomas-Crusells, J.; Li, H.; Emri, Z.; Sipila, S.; Payne, J.A.; Minichiello, L.; Saarma, M.; Kaila, K. Mechanism of activity-dependent downregulation of the neuron-specific K-Cl cotransporter KCC2. J. Neurosci. 2004, 24, 4683–4691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. He, X.P.; Kotloski, R.; Nef, S.; Luikart, B.W.; Parada, L.F.; McNamara, J.O. Conditional deletion of TrkB but not BDNF prevents epileptogenesis in the kindling model. Neuron 2004, 43, 31–42. [Google Scholar] [CrossRef] [Green Version]
  168. Chen, Z.; Brodie, M.J.; Liew, D.; Kwan, P. Treatment Outcomes in Patients with Newly Diagnosed Epilepsy Treated with Established and New Antiepileptic Drugs: A 30-Year Longitudinal Cohort Study. JAMA Neurol. 2018, 75, 279–286. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The principal PLC signaling cascades and functions in the brain. External ligands such as neurotransmitters and neurotrophic factors bind to and activate the upstream transmembrane receptors of PLC. PLCβ is activated by Gαq and Gβγ subunits of G protein-coupled receptors (GPCRs), whereas the activation of PLCγ is triggered by the phosphorylation of receptor tyrosine kinases (RTKs). Activation of PLC hydrolyzes phospholipid PIP2 into IP3 and DAG and these second messengers mediate diverse neuronal functions.
Figure 1. The principal PLC signaling cascades and functions in the brain. External ligands such as neurotransmitters and neurotrophic factors bind to and activate the upstream transmembrane receptors of PLC. PLCβ is activated by Gαq and Gβγ subunits of G protein-coupled receptors (GPCRs), whereas the activation of PLCγ is triggered by the phosphorylation of receptor tyrosine kinases (RTKs). Activation of PLC hydrolyzes phospholipid PIP2 into IP3 and DAG and these second messengers mediate diverse neuronal functions.
Ijms 22 03149 g001
Figure 2. The functions of PLCβ in GABAergic inhibition. (a) PLCβ is activated by diverse GPCRs such as KA receptor, muscarinic receptor, and D2 receptor. Activation of PLCβ suppresses presynaptic GABA release, resulting in increased excitation of postsynaptic neurons. (b) However, in other neuronal types, activation of PLCβ by ethanol or serotonin increases presynaptic GABA release and consequently decreases the excitation of postsynaptic neurons.
Figure 2. The functions of PLCβ in GABAergic inhibition. (a) PLCβ is activated by diverse GPCRs such as KA receptor, muscarinic receptor, and D2 receptor. Activation of PLCβ suppresses presynaptic GABA release, resulting in increased excitation of postsynaptic neurons. (b) However, in other neuronal types, activation of PLCβ by ethanol or serotonin increases presynaptic GABA release and consequently decreases the excitation of postsynaptic neurons.
Ijms 22 03149 g002
Figure 3. The functions of PLCγ in GABAergic inhibition. (a) Activation of PLCγ by the BDNF/TrkB pathway prevents GABA reuptake by GAT-1, leading to the accumulation of extracellular GABA. (b) Activation of PLCγ increases the surface expression of GABAA receptors. In addition, PLCγ inhibits Cl extrusion via the regulation of KCC2 function. Activation of GABAB receptors induces BDNF secretion through PLCγ activation and consequently increases the expression of GABAA receptors in the postsynaptic membrane, therefore decreasing the excitation of neurons.
Figure 3. The functions of PLCγ in GABAergic inhibition. (a) Activation of PLCγ by the BDNF/TrkB pathway prevents GABA reuptake by GAT-1, leading to the accumulation of extracellular GABA. (b) Activation of PLCγ increases the surface expression of GABAA receptors. In addition, PLCγ inhibits Cl extrusion via the regulation of KCC2 function. Activation of GABAB receptors induces BDNF secretion through PLCγ activation and consequently increases the expression of GABAA receptors in the postsynaptic membrane, therefore decreasing the excitation of neurons.
Ijms 22 03149 g003
Table 1. The animal models of epilepsy.
Table 1. The animal models of epilepsy.
TypeEpilepsy ModelMechanismSymptomsReference
GeneticGenetic Absence Epilepsy Rat from Strasbourg (GAERS)Inbred strain
Mutation of Cacna1h gene encoding CaV3.2
T-type calcium channel
Spike-and-wave discharges (SWD) in EEG[33,41]
WAG/RijPolygenic gene mutationSpike-and-wave discharges (SWD) in EEG[42,43]
DBA/2Mutation of Asp2 geneAudiogenic seizures[34,35]
Genetically epilepsy-prone rats (GEPR)GABAergic, serotonergic, noradrenergic deficitsAudiogenic, generalized tonic-clonic seizures[33]
ElectricalKindlingLower threshold by repeated stimulationTemporal lobe epilepsy[30,31,32]
ChemicalPilocarpineMuscarinic acetylcholine receptor agonistGeneralized tonic–clonic seizures[24,25,26]
Kainic acidL-glutamate analogTemporal lobe epilepsy[27,28,29]
Table 2. Current antiepileptic drugs with GABAergic effects.
Table 2. Current antiepileptic drugs with GABAergic effects.
DrugMechanismEpilepsy TypesReference
Potassium bromideGABA potentiationGeneralized tonic-clonic seizures,
myoclonic seizures
[53,54]
PhenobarbitalPotentiation of GABAA receptorPartial and generalized convulsive seizures[55]
PrimidoneGABA potentiationPartial and generalized convulsive seizures[56]
DiazepamPotentiation of GABAA receptorStatus epilepticus[57,58,59]
ValproateMultiple mechanisms with glutamate inhibition,
blockade of sodium and T-type calcium channels,
inhibition of GABA transaminase and re-uptake
Partial and generalized
convulsive seizures,
absence seizures
[60,61]
ClonazepamPotentiation of GABAA receptorJuvenile myoclonic epilepsy[62,63]
BenzodiazepinesPotentiation of GABAA receptorPartial and generalized convulsive seizures, Lennox–Gastaut syndrome,
myoclonic seizures
[58]
VigabatrinInhibition of GABA transaminaseInfantile spasms, complex partial seizures[64,65]
TiagabineInhibition of GABA transporterPartial seizures[52]
Table 3. The functions of PLC in epilepsy.
Table 3. The functions of PLC in epilepsy.
PLC IsozymeAnimal or Human StudyPhenotypeReference
PLCβ1Genetic knockout miceEarly-onset epileptic encephalopathy[127]
Mongolian gerbils miceIncreased PLCβ1 expression after seizures[135]
Genetic knockout miceMalignant migrating partial seizures
in infancy
[128]
Pilocarpine-induced status epilepticus in miceDecreased PLCβ1 expression in hippocampal interneurons after seizures[136]
Homozygous deletions or nonsense variant in humanInfantile epileptic encephalopathy[130]
PLCβ4Genetic knockout miceAbsence seizures[133]
PLCγ1TrkB mutation mice in PLCγ1 binding domainDecreased pilocarpine-induced
status epilepticus
[137]
Heterozygote knockout miceDecreased kindling-induced seizures[138,139]
GABAergic neuron-specific
knockout mice
Late-onset seizures[140]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, H.Y.; Suh, P.-G.; Kim, J.-I. The Role of Phospholipase C in GABAergic Inhibition and Its Relevance to Epilepsy. Int. J. Mol. Sci. 2021, 22, 3149. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22063149

AMA Style

Kim HY, Suh P-G, Kim J-I. The Role of Phospholipase C in GABAergic Inhibition and Its Relevance to Epilepsy. International Journal of Molecular Sciences. 2021; 22(6):3149. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22063149

Chicago/Turabian Style

Kim, Hye Yun, Pann-Ghill Suh, and Jae-Ick Kim. 2021. "The Role of Phospholipase C in GABAergic Inhibition and Its Relevance to Epilepsy" International Journal of Molecular Sciences 22, no. 6: 3149. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22063149

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