Next Article in Journal
Novel Variant of the SLC4A1 Gene Associated with Hereditary Spherocytosis
Next Article in Special Issue
Pharmacological Potential of 3-Benzazepines in NMDAR-Linked Pathophysiological Processes
Previous Article in Journal
Immune-Onco-Microbiome: A New Revolution for Gynecological Cancers
Previous Article in Special Issue
The Roles of Glutamate Receptors and Their Antagonists in Status Epilepticus, Refractory Status Epilepticus, and Super-Refractory Status Epilepticus
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Glutamate Receptors in Epilepsy

1
Department of Neurology, Tainan Sin-Lau Hospital, Tainan 70142, Taiwan
2
Zhengxin Neurology & Rehabilitation Center, Tainan 70459, Taiwan
3
Department of Pediatrics, Chi-Mei Medical Center, Tainan 71004, Taiwan
4
Department of Neurology, National Cheng Kung University Hospital, College of Medicine, National Cheng Kung University, Tainan 70142, Taiwan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomedicines 2023, 11(3), 783; https://doi.org/10.3390/biomedicines11030783
Submission received: 30 January 2023 / Revised: 26 February 2023 / Accepted: 2 March 2023 / Published: 4 March 2023
(This article belongs to the Special Issue Glutamate Receptor in Health and Development)

Abstract

:
Glutamate is an essential excitatory neurotransmitter in the central nervous system, playing an indispensable role in neuronal development and memory formation. The dysregulation of glutamate receptors and the glutamatergic system is involved in numerous neurological and psychiatric disorders, especially epilepsy. There are two main classes of glutamate receptor, namely ionotropic and metabotropic (mGluRs) receptors. The former stimulate fast excitatory neurotransmission, are N-methyl-d-aspartate (NMDA), α-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid (AMPA), and kainate; while the latter are G-protein-coupled receptors that mediate glutamatergic activity via intracellular messenger systems. Glutamate, glutamate receptors, and regulation of astrocytes are significantly involved in the pathogenesis of acute seizure and chronic epilepsy. Some glutamate receptor antagonists have been shown to be effective for the treatment of epilepsy, and research and clinical trials are ongoing.

1. Introduction

Epilepsy is a brain disease associated with chronic recurrent seizures. The mainstay theory is that it is caused by an imbalance between the excitatory and inhibitory conductance of nerve cells within the brain [1]. Glutamate is the major neurotransmitter for excitatory neuronal signaling, while γ-aminobutyric acid (GABA) has an inhibitory function. Glutamate plays a role in both pre- and post-synaptic excitatory neurotransmission resulting in cellular and network hyperactivity, and underlies the formation of ictogenesis and epileptogenesis. Exploring and understanding the glutamatergic mechanisms will contribute to the improvement of epilepsy management strategies. This article reviews the current literature on the functions of glutamate receptors and the modulation of glutamate transmission.

2. Subcellular Structure and Physiology of Glutamate Receptors

Glutamate mediates excitatory transmission via various pre- and post-synaptic glutamate receptors. The glutamate receptors encompass ionotropic receptors (transmembrane ligand-gated ion channels) and metabotropic receptors, which affect presynaptic glutamate release or indirectly modulate the function of ionotropic receptors. Ionotropic glutamate receptors (iGluRs) share a common tetramer containing A–D subunits [2,3]. Presynaptic neuronal depolarization drives calcium-dependent glutamate release from terminal vesicles to the cleft of the synapse. iGluR ion channels open in response to glutamate binding, allowing sodium with calcium cations to flow in, which rapidly depolarizes the postsynaptic membrane and initiates signal transduction in the postsynaptic neuron [4]. As high levels of extracellular glutamate have an excitotoxic effect in neurons, the glutamate concentration is optimally controlled via the reuptake of glutamate from the synapse by astrocytic glutamate transporters, including glutamate–aspartate transporter (GLAST) and glutamate transporter-1 (GLT-1) [5,6,7]. When glutamate is taken up by astrocytes, it is converted to glutamine by the enzyme glutamine synthetase. The glutamine is then delivered from the astrocytes to neurons via glutamine transporters. Finally, glutamine is converted to glutamate in the neurons by glutaminase to complete the glutamate–glutamine cycle [8]. Glutamate is also involved in long-term potentiation and synaptic plasticity in the normal brain [9]. Three kinds of postsynaptic ionotropic glutamate receptors have been identified, namely the α-amino-3-hydroxy-5-methyl-4-isoxazolepro-pionic acid (AMPA) receptors (AMPARs), which are responsible for fast excitatory neurotransmission, N-methyl-D-aspartate (NMDA) receptors (NMDARs), which mediate most of the slow postsynaptic excitatory potentials, and the kainate receptor (KAR). The function of KAR is not clearly known, although it may modulate pre- and post-synaptic excitatory neurotransmission [10]. Injection of kainate into the rat hippocampus mediated the inhibition of presynaptic GABA release and postsynaptic KAR activation of glutamatergic neurons, resulting in epileptogenesis [11,12].

2.1. AMPA Receptors (AMPARs)

Glutamate is the direct binding ligand that activates AMPARs. AMPARs are expressed in the postsynaptic neuronal membrane and are involved in rapid excitatory neurotransmission in the brain. They have a tetrameric configuration containing four types of subunit, GluA1 to GluA4 (also known as GluR-A to GluR-D), assembled in different combinations [13] (Figure 1). The GluA2 subunit is the key site for regulating calcium permeability [14,15]. The calcium permeability regulation of the GluA2 subunit is affected by a posttranscriptional modification. The AMPARs are calcium (Ca2+)-permeable if they contain the unedited GluA2 subunit or the GluA2-lacking structure [16,17]. Most AMPARs in the brain contain the edited GluA2 subunit and are therefore Ca2+-impermeable. Mounting evidence suggests that Ca2+-permeable AMPARs play an important role in receptor trafficking, synaptic plasticity, learning, and memory [18,19,20,21,22], as well as excitotoxic cell death [23,24]. In other words, the synaptic homeostatic plasticity comes from dynamic changes to AMPARs that occur during their biosynthesis, membrane trafficking, and degradation, which are controlled by complex regulating proteins [20]. In experimental rats following pilocarpine-induced status epilepticus (SE), the expression of calcium-impermeable GluA2-containing AMPARs in their cortex was increased, indicating a potential long-term change in neuroplasticity for neuroprotection in response to SE-induced neurotoxicity [25].

2.2. N-Methyl-D-Aspartate Receptors (NMDARs)

As glutamate release activates glutamate receptors, it mediates an inward cation current and thereby depolarizes the postsynaptic neurons. Two types of excitatory postsynaptic currents can be found. AMPARs activation mediates a fast current with a rapid increase and fades away, whereas the NMDARs mediates a slower current that lasts from tens to hundreds of milliseconds [26,27,28]. NMDARs also have a tetrameric structure made up of different combinations of GluN1, GluN2, and GluN3 subunits, with each having several isoforms [29] (Figure 1). The GluN1 subunit is essential for all functional NMDARs and has binding sites for glycine and D-serine. Glycine (B)-binding sites in both GluN1 and GluN3A subunits facilitate receptor forward trafficking to the cell surface. Mutation in the glycine-binding site of the human GluN3A subunit remarkably reduces the surface expression of NMDARs [30]. These findings indicate that glycine-binding sites play an important role in the regulation of NMDARs. The GluN2 subunits provide glutamate-binding sites and control the expression and functional properties of NMDARs [31]. In brief, several types of NMDAR modulators are noted. They include competitive antagonists acting at the agonist binding site, sodium and calcium blockers, allosteric sites different from endogenous agonist binding sites, and glycine site antagonists [32]. The binding of glutamatergic ligands is not sufficient to open the channel because the NMDAR channel is plugged by magnesium ion (Mg2+) at the resting membrane potential. When the Mg2+ is removed, the membrane is depolarized, allowing a voltage-dependent inward flow of sodium (Na+) and Ca2+ ions while potassium (K+) ions flow out of the cell [33]. Low extracellular magnesium is associated with seizure-like discharges in the brain due to the unblocking of NMDARs [34]. Calcium influx through NMDARs can activate the binding proteins and is thought to play an important role in synaptic plasticity, a cellular mechanism for learning and memory [35,36,37].

2.3. Kainate Receptors

Kainate receptors (KARs) are ionotropic glutamate receptors, in addition to AMPA and NMDARs. KARs are activated by the agonist, kainate, and play a role in postsynaptic excitatory neurotransmission. KARs are also tetrameric structures by different assemblies of Glu K1-Glu K5 to form functional KARs. Glu K1, K2, and K3 may undergo editing and splicing to form some variants such as GluK1a, GluK1b, GluK1c, and GluK2a-2c etcetera [38]. KARs exert their intracerebral expression in the amygdala, entorhinal cortex, and hippocampus [39,40,41]. They generate a stereotypical electrophysiological response when activated, which consists of a small amplitude and slow decay current, resulting from both their interaction with their auxiliary subunits and the biophysical properties of heteromeric receptors [42]. KARs mediate both canonical ionotropic and non-canonical metabotropic signaling. The KAR activation of a non-canonical signaling pathway is analogous to metabotropic glutamate receptors, which also modulate neuronal excitability and transmitter release. Among the multiple mechanisms in the regulation of neurotransmitter release by presynaptic KARs, some are considered to involve a non-canonical action of KARs [43]. KARs could act through a non-canonical mode of action through activation of second messenger signaling pathways [43,44].
In temporal lobe epilepsy, aberrant recruitment of KARs at recurrent mossy fiber synapses participates in epileptogenic neuronal activity [44]. In addition, kainic acid (KA) can be used to establish an experimental model of temporal lobe epilepsy (TLE), which is a common type of partial epilepsy characterized by neuronal loss in the CA1 and CA3 regions of the hippocampus. Intra-amygdaloid injections of KA induce psychomotor seizures and produce neuropathological lesions similar to those found in TLE [45,46]. Furthermore, intra-hippocampal injection of KA also produced similar preconditions for TLE, such as old cerebral insults, encephalitis, or SE [47]. KA1 subunits are highly expressed in CA3 pyramidal cells and KA2 subunits are highly expressed in both CA1 and CA3 pyramidal cells [48,49,50]. Therefore, KA1 and KA2 receptors have high affinity for glutamate and high expression in the CA3 region of the hippocampus. This is why this region is vulnerable to the excitotoxic damage induced by KA, and the hippocampus often becomes the epileptogenic zone in this activation model [51,52].
Although the mechanism of KAR is least known in the iGluRs, there is some progress of advanced knowledge in recent years. As mentioned earlier, KARs act by modulating neuronal excitability and regulation of synaptic network activity by complex expression patterns in the CNS [53]. The retention of KARs in the endoplasmic reticulum (ER) or trafficking of KARs from the ER to the cell surface is an important regulation mechanism. Some KAR-interacting proteins are involved in the surface trafficking. For example, NETO (neuropilin and tolloid-like) proteins possibly play a role in the trafficking of KARs. Other protein reacting substances such as PKC (protein kinase C)-mediated phosphorylation and SUMOylation (small ubiquitin-related modifier) have been reported to have impacts on surface expression and endocytosis [54]. Moreover, molecular modeling and calcium imaging studies showed that the occupancy of both GluK2 and GluK5 ligand binding domains is required for the full activation of GluK2/GluK5 heteromeric KAR channels [55]. Of note, the non-canonical signaling of KARs which activates phospholipase C and PKC via a metabotropic G-protein dependent pathway, regulating neuronal excitability by inhibiting the slow afterhyperpolarization, neurotransmitter release, and glutamate receptor trafficking could potentially contribute to increased neuronal excitability [56]. Both of the roles of the canonical and non-canonical KARs pathways in the generation of seizure activity are worth further investigation.
The biosynthesis, assembly, and cell surface trafficking of KARs are key determinants of neuronal excitability in the CNS. As KARs dysfunction is linked to ischemic, chronic pain, epilepsy, and psychiatric disease such as schizophrenia [10], knowing the action of KAR-interacting proteins in the trafficking and surface expression to develop new therapeutic strategies against these diseases is important.

2.4. Metabotropic Receptors

Metabotropic receptors (mGluRs) are G-protein-coupled non-ionotropic receptors that modulate either presynaptic glutamate release or postsynaptic ionotropic receptors. There are three major functional subgroups of mGluRs: Group I mGluRs encompass presynaptic mGluR1, postsynaptic mGluR5, and both in astrocytes. Group I mGluRs enhance excitatory function in mGluR1 and mGluR5. Presynaptic mGluR1 facilitates the vesicular release of glutamate, whereas postsynaptic mGluR5 can modulate the action of postsynaptic ionotropic receptors.
Activation of mGluR1 by a specific agonist can maintain and prolong interictal discharge [57]. mGluR1 can also be activated by increases in intracellular calcium and depolarization of hippocampal CA1 pyramidal cells, which are involved in inducing long-term potentiation (LTP) and long-term depression (LTD), a long-lasting synaptic plasticity, at multiple glutamatergic synapses [58]. Upregulation of hippocampal mGlu5 was found in brain slices from pharmaco-resistant TLE patients. The upregulation of mGluR5 in surviving neurons is probably a consequence of the hyperexcitability of the hippocampus, but not the cause of epileptic seizures in pharmaco-resistant TLE patients, as this phenomenon has been observed in both hippocampal sclerosis and non-hippocampal sclerosis patients [59]. Additionally, protein–protein interaction is an important coupling mechanism for the action and regulation of metabotropic receptors. mGlu5 binds to NMDARs via scaffolding proteins, both postsynaptic density protein 95 (PSD-95) and Homer protein, leading to phosphorylation of NMDARs. The net result is potentiation of NMDAR currents [58,60]. Group II metabotropic receptors include mGluR2/3, which are predominately located on the presynaptic terminal and can inhibit presynaptic glutamate release. mGluR2/3 expression was markedly and progressively down-regulated in both CA1 and CA3 in a pilocarpine-induced SE model [61,62]. mGluR3 is expressed on astrocytes and positively modulates GLAST and GLT-1, the glutamate transporter proteins for synaptic glutamate reuptake, and consequently counters hyperexcitability [63]. Interestingly, astrocytic mGluR3 and mGluR5 expression are upregulated in TLE, indicating that seizure-induced upregulation of two opposite mGluR subtypes in reactive astrocytes is a novel mechanism for modulation of glial function and changes in glial–neuronal communication in the course of epileptogenesis [64]. Group III metabotropic glutamate receptors include mGluR4, 6, 7, and 8, and function as inhibitory presynaptic receptors. In an animal model of TLE, mGluR1 can promote seizure susceptibility, whereas mGluR4 is upregulated to counteract the excitatory activity and seizure-associated vulnerability of hippocampal neurons [65]. In summary, metabotropic glutamate receptors modulate glutamatergic signaling by altering their expression and distribution, and play key roles in hippocampal excitability, epileptogenesis, and neuronal degeneration.

2.5. Astrocytes in Glutamate Uptake and Release

Astrocytes act on the uptake of synaptic and extracellular glutamate by astrocytic glutamate transporters. Several studies over the past 10 years show that glutamate is released from astrocytes for regulation of neuronal activity under physiological conditions [66]. When glutamate is released from excitatory neurons, it stimulates group I metabotropic glutamate receptors in astrocytes. The activation of these receptors induces elevated release of the astrocytic intracellular Ca2+ concentration, which in turn triggers glutamate release from astrocytes. The astroglial-released glutamate subsequently activates the extrasynaptic GLuN2B subunit of NMDARs in the nearby neurons, generating slow inward currents inside these neurons that synchronize their action potential firing [67,68,69,70]. The astroglial-released glutamate also stimulates presynaptic group I mGluRs and NMDARs, inducing presynaptic neurons to release more glutamate [71,72,73]. This glial–neuronal crosstalk can possibly explain why astrocytes are capable of generating seizures by impairment of homeostatic extracellular glutamate [74]. On the other hand, glutamate also potentiates neuronal inhibition. Inhibitory interneurons release GABA, which activates astrocytic GABAB receptors and induces an increased release of Ca2+ within astrocytes, which in turn triggers glutamate releases [75]. Glutamate acts on presynaptic ionotropic glutamate receptors and augments the release of more GABA from the surrounding inhibitory neurons [76]. Therefore, astrocytes act through a unique dual function involving both glutamate uptake and release to maintain glutamate homeostasis in the CNS.

2.6. Cannabidiol (CBD) and Glutamate Signaling

Cannabidiol (CBD) is a compound extracted from cannabis, which has been proven to have multiple pharmacological effects, including anti-inflammatory [77,78,79], immunomodulatory [80], neuroprotective [81,82], and anti-seizure effects [83,84]. It could induce neuroprotection through the normalization of homeostasis of glutamate [85]. Evidence shows that CBD modulates glutamate and GABA levels in brain regions such as the basal ganglia and the dorsomedial prefrontal cortex [86]. In the presynaptic nerve terminal, a transmembrane orphan G protein-coupled receptor 55 (GPR 55) can be activated by agonists to increase intracellular Ca2+ concentration and facilitate glutamate release from vesicles [83]. GRP 55 is one of the targets on which CBD acts. CBD has been shown to antagonize GRP 55, thereby decreasing glutamine release and exerting its anti-seizure effect [84,87]. Furthermore, the overexpression of cannabinoid type 1 (CB1) receptors in hilar mossy cells of the dentate gyrus and in pyramidal cells of the hippocampal regions was shown to significantly reduce the severity of kainate-induced SE [88] and the expression of CB1 receptors in excitatory neurons of the cerebral cortex, hippocampus, and amygdala of the global CB1 receptor knockout mice was able to prevent the exacerbation of kainate-induced seizures [89,90]. Clinically, CBD has been demonstrated to reduce seizure frequency in drug-resistant epilepsies, including Dravet syndrome, Lennox-Gastaut syndrome, and tuberous sclerosis complex [91,92,93,94,95]. In conclusion, glutamate plays a crucial role in the majority of excitatory transmission in the CNS via ionotropic and metabotropic glutamate receptors. The three ionotropic glutamate receptors (NMDAR, AMPAR, KAR) share similar tetramer structures, with slow excitatory transmission associated with NMDAR and fast excitatory transmission with AMPAR. The inward cation current regulation, including Ca2+ and Na+, may contribute to synaptic plasticity, learning, and memory in the normal brain. Metabotropic glutamate receptors are G-protein-coupled receptors, containing three subgroups that mediate glutamate release in the presynapse, ionotropic receptors in the postsynapse, and glutamate recycling in astrocytes. Alterations in their expression and distribution may play a role in the development of epilepsy or neurodegeneration.

3. Main Mechanism of Glutamate Receptor in Epilepsy

Activation of the NMDAR has been demonstrated to evoke burst firing in CA1 neurons and granular cells of the hippocampus [96,97]. It has shown that both NMDA and quisqualate, an AMPA agonist, evoke a distinct spike in activity. Quisqualate induces a rapid depolarization that evokes tonic firing, whereas NMDA produces bursts of fast action potentials superimposed on an underlying depolarizing shift of membrane potential [98]. From animal models of epilepsy and SE in humans, it is evident that glutamate constitutes a key role in seizures and epileptogenesis. SE-induced glutamate release results in over-stimulation of glutamate receptors, sustained prolonged seizure activity, and seizure-related excitotoxic brain injury [99,100].
Region-specific alterations of glutamate receptor subunits in a pilocarpine model of TLE suggest cellular mechanisms contributing to a generation of independent epileptogenic networks in different temporal lobe areas [101]. The role of NMDAR-mediated excitotoxicity, primarily via excessive Ca2+ influx through the extrasynaptic GluN2B-containing receptors, in the context of both acute and chronic neurological disorders, is well-established [56,102,103,104,105,106], and the carboxy terminal domain of GluN2B remains an important factor mediating excitotoxic signaling [103]. In addition, interfering peptides disrupting the coupling of GluN2B to PSD-95 and nNOS [107,108], reducing nNOS activation and nitrosative and oxidative stresses, entered clinical trials and showed effectiveness in terms of clinical outcome [109,110]. The diversity of the carboxy terminal domain among different GluN2 subunits is therefore important in determining the subunit-specific trafficking of GluN2-containing NMDARs and excitotoxicity in various neurological disorders.
A well-established mechanism of regulation of receptor surface expression is the phosphorylation/dephosphorylation of specific residues in the carboxy terminal domain of GluN2A and GluN2B subunits. As we know, a key mediator of neuronal excitotoxicity is the excessive activation of NMDARs and the calcium influx. It has been demonstrated that the activation of synaptic versus extrasynaptic NMDAR stimulation is remarkably different in the context of excitotoxicity, possibly related to the different subunit composition of the two receptors [111]. Importantly, studies also provided strong evidence that the carboxy terminal domain of GluN2 subunits and their C-terminal interacting proteins play roles in controlling the subcellular localization, synaptic targeting, and intracellular Ca2+ signaling of GluN2-containing NMDARs, and affect neuronal excitotoxicity [56,111].
With prolonged SE, receptor trafficking is noted with GABA receptors being internalized and NMDARs migrating to neuronal synapses [112,113]. Overactivation of AMPARs can elicit TLE. There is typically relatively dense expression of AMPARs in the hippocampus in this type of chronic seizure [114,115]. These findings show that glutamate receptor expression and distribution are dynamic. These SE-induced changes in receptor localization explain why drugs that target GABAergic neurotransmission (such as benzodiazepines) usually fail to control refractory SE, whereas NMDAR antagonists in combination with GABA agonists can often alleviate sustained SE. Sodium and calcium channels also exert indirect effects in regulating glutamate concentration and neuron protection. Presynaptic calcium and sodium channel activation enhance glutamate efflux into the extracellular space, and a higher concentration of glutamate subsequently leads to neuronal injury [116]. Riluzole, a neuroprotector, can lessen the symptoms of motor neuron disease by blocking persistent calcium and sodium currents, thereby modulating glutamate release via presynaptic NMDARs [117].
One of the main mechanisms for the induction of seizures is the suppression of GABA release when KA activates KAR, along with postsynaptic KAR activation of glutamatergic neurons [118,119]. Interneurons in the CA1 region of the hippocampus have KARs containing the GluK1 subunit in their axonal compartment and GluK2 in the somatodendritic compartment [120]. KARs containing the GluK2 subunit have been linked to limbic epilepsy, related to their specific expression in CA3 pyramidal neurons [49]. In animal models of TLE and in human patients, neuronal networks undergo a substantial reorganization whereby neuronal death, sprouting, and aberrant connection are prominent [121]. GluK1 mRNA is abundant in temporal lobe areas, including the amygdala, and prolonged activation of basolateral GluK1 subunit-containing KARs by ATPA (an agonist of GluK1 subunit-containing KARs) induces spontaneous epileptiform activity, which is sensitive to KAR antagonism [119,122].
NMDAR antagonists such as MK-801 or ketamine have antiseizure effects and provide neuroprotection against excitotoxicity after prolonged SE [123,124,125]. After prolonged SE, NMDARs mediate long-term remodeling of synaptic connectivity and dendritic morphology, thereby increasing excitatory signaling and promoting the onset of spontaneous recurrent seizures. Moreover, NMDAR subunits undergo posttranscriptional modifications and trafficking of GluN1 subunits from the intracellular compartment to the cell membrane surface in response to SE [113,126]. Brain-specific microRNA-134, a small non-coding RNA, is localized to the synaptodendritic compartment of rat hippocampal neurons and participates in NMDAR-dependent spine remodeling, which reinforces postsynaptic sites of excitatory transmission [127]. Inhibition of microRNA-134 is a promising mechanism for the development of antiseizure medications [128].

3.1. NMDA Receptors Mutation

Some gene mutations have been identified in NMDARs. Mutations in GRIN1 (encodes the GluN1 subunit), GRIN2B (encodes the GluN2B), and GRIN2D (Glu N2D) present with severe clinical phenotypes, including severe epilepsy with mental retardation and developmental delay (Figure 2). Four de novo missense GRIN1 mutations cause infantile-onset epilepsy with encephalopathies, as well as hyperkinetic movement disorders [129]. GRIN2B gain-of-function mutation causes West syndrome, presenting with focal epilepsy and intellectual disability [130]. Six new GRIN2D variants were reported with initial early-onset seizures (by one year old) and general growth retardation, and some had low intraocular pressure, dyskinesia, or autistic behavior [131]. GluN2A mutations or variants can be related to benign Rolandic epilepsy [132,133]. Notably, the excitatory glutamatergic neurons and synapses outweigh the amount of GABAergic neurons and synapses [134,135]. A similar epilepsy phenotype may exist in a state of hyper-NMDA function (excessive NMDAR activation, or NMDA-pathies) or hypo-NMDA function, which diminishes GABA interneuron activation and thereby causes disinhibition [136]. For this reason, either increased or decreased NMDA function will enhance neuronal excitation. Hypofunction of NMDARs is also found in the aging brain and persistent hypofunction will cause neurodegeneration with cognitive and psychotic impairment, in addition to epilepsy [137].

3.2. AMPA Receptors Mutation

Mutation of AMPARs is less common than NMDARs. AMPAR gene mutations are often associated with cognitive impairment, developmental delay, and psychiatric diseases, such as autism spectrum disorder, besides epilepsy [138] (Figure 2). Heterozygous de novo GRIA2 (encodes Glu A2 subunit) mutations in 28 patients were reported [139]. Their phenotypes are intellectual disability, neurodevelopmental abnormalities including autism spectrum disorder, Rett syndrome-like features, and seizures or developmental epileptic encephalopathy. As previously mentioned, the GluA2 subunit is the key site for regulating calcium permeability and is affected by RNA editing as well as the structural components. In GluA2 mutant mice that lack GluA2, LTP in the CA1 region of hippocampal slices was enhanced with normal neuronal excitability [140]. The increase in Ca2+ permeability is a possible mechanism for enhanced LTP. The mutant mice also showed disorders in mental development and motor coordination. These results suggest an important role for GluA2 in synaptic plasticity and behavior.

3.3. Anti-NMDA Antibody Encephalitis

Anti-NMDA receptor encephalitis has drawn much attention over the years and is a common type of autoimmune encephalitis [141,142]. The antibody causes internalization of the receptors with a selective and reversible decrease in NMDAR surface density and synaptic localization [142]. Although patients’ antibodies reduced synaptic NMDAR-mediated currents, the localization or expression of other glutamate receptors or synaptic number were not affected [143]. It can be postulated that NMDA antagonists will reduce excitatory transmission in the brain and alleviate seizures. However, a therapeutic trial revealed that D-CPP-ene, a competitive NMDA antagonist, did not help patients with intractable complex partial seizures. Instead, it worsened seizures in three of the eight patients [144]. In the presence of patients’ cerebrospinal fluid and purified IgG antibodies, extracellular glutamate levels increased due to dysfunction of the NMDA-related glutamatergic turnover or impaired turnover of receptors, and thereby NMDA-related excitotoxicity occurred [145]. In another study, patients’ CSF and purified IgG were injected into mice brains, leading to increased extracellular glutamate [146]. This hyperglutamatergic state is speculated to be due to an imbalance between the NMDA and AMPA pathways [147]. In other words, the brain circuitry shifts to an AMPA-dependent hyperglutamatergic state under the influence of antibodies on NMDARs. It is possible that NMDAR antibody-induced elevated glutamate in the human brain is reversible and does not reach the synaptic threshold for neurodegeneration through the overactivation of AMPARs [148]. It is also accordant with the relatively better prognosis of patients with anti-NMDA encephalitis in the presence of early immunotherapy and removal of causative tumors [141,149].
To summarize, both NMDA and AMPA receptors have been associated with epilepsy and SE. Mutations or variants of subunits in NMDA and AMPA channels have been linked to various clinical features, including growth retardation, early onset of epilepsy, and psychiatric disorders. The overexpression and alteration of NMDA receptors, along with the internalization of GABA receptors, are currently considered to be potential mechanisms underlying SE, based on the evidence of the anti-seizure effects of various NMDA receptor antagonists in animal studies. Additionally, antibodies in anti-NMDA receptor encephalitis have been found to decrease the density of NMDA receptors via internalization and to increase extracellular glutamate levels, leading to excitotoxicity.

4. Antiseizure Medications Acting on Glutamate Receptors

It is clear that targeting the balance of excitatory and inhibitory neurotransmission is the cornerstone of antiseizure medications (ASM). Today, several ASMs acting on the glutamatergic system are available. Topiramate (TPM) has shown a broad spectrum of efficacy in drug-resistant partial and secondary generalized seizures [150,151]. In addition to blockade of voltage-dependent sodium channels and potentiation of GABA-mediated neurotransmission, it exerts antagonism in glutamate receptors. TPM inhibits neuronal excitatory pathways by selective action at AMPA and kainate receptors. TPM blocks kainate-evoked inward currents in cultured hippocampal neurons and exerts a biphasic effect on kainate-evoked currents, all of which may contribute to its anticonvulsant activity [152,153]. TPM also has a negative modulatory effect on L-type calcium channels, thus reducing the amplitude of high voltage-activated calcium currents [154].
Felbamate has a dual mechanism of action as a positive modulator of GABAA receptors and an inhibitor of glutamate currents mediated by NMDARs [155]. Actually, felbamate’s anticonvulsant effect is partly mediated by antagonism at the glycine B site of NMDARs [156]. Pregabalin selectively binds to the α2δ subunit of presynaptic voltage-gated calcium channels to block P/Q-type calcium currents, and indirectly inhibits the calcium-dependent release of glutamate from vesicles [157]. Lamotrigine is categorized as a sodium channel blocker by its action on voltage-gated sodium channels. It can also modulate P/Q-type, N-type, and R-type calcium channels on presynaptic terminals [158,159], thereby inhibiting the release of glutamate.
Riluzole modulates glutamate function by inhibition of voltage-gated sodium currents to decrease presynaptic glutamate release [160], as well as by blocking Ca2+ entry via the NMDA channel, thereby reducing neuronal excitability [161]. Riluzole also enhances the activities of astrocytic glutamate transporters glutamate-aspartate transporter (GLAST) and glutamate transporter-1 (GLT-1), which cause buffering effects on excessive extracellular glutamate [162]. In a pilocarpine-induced or electroconvulsive shock-induced seizure model, riluzole showed its effect on seizure suppression [163,164]. Although riluzole’s efficacy has not been proven in epilepsy patients, a laminar cortex model, examining neuronal excitability and network alteration, has predicted its potential therapeutic targets for treating seizures caused by loss-of-function changes in Kv1 channelopathy [165].
Memantine blocks ionotropic NMDA receptors by binding to Mg2+ binding sites when the receptor channel is open. This action inhibits the prolonged influx of Ca2 to stabilize NMDA activity [166]. In a pentylenetetrazole (PTZ)-induced seizure model, memantine prevented convulsions and morphological changes in rat brain neurons [167]. A mice model showed that memantine reduced cortical excitability by producing a more uniform direct current stimulation (DCS)-induced long-term depression-like activities throughout the cortical thickness. Additionally, a combination of memantine and DCS suppressed KA-induced seizures [168]. The application of memantine needs further investigations in patients with epilepsy.
As NMDA hypofunction is thought to be associated with schizophrenia, the potential adverse behavioral effects of NMDAR-selective antagonists in epilepsy treatment raise some concerns [169]. Ketamine administration in healthy men resulting in dose-dependent negative psychiatric symptoms, as well as a decrease in verbal and nonverbal declarative memory performance is an example [170]. In contrast, AMPAR signaling exerts fewer effects on neuroplasticity than NMDARs, and has greater potential to modulate hyperexcitability with incidence of psychosis [171,172]. It is for this reason that targeting modulation of AMPARs may be more promising than targeting NMDARs [173].
NMDAR antagonists provide anticonvulsant effects in the maximal electroshock and reflex seizure models [174]. They also exhibit seizure protection effects in PTZ-kindling-induced hippocampal astrocytosis, oxidative stress, and neuronal loss [175]. However, some NMDAR antagonists showed only weak anticonvulsant effects and did not increase the focal seizure threshold in fully kindled rats [176]. Instead, they may produce prominent behavioral side effects in kindled animals [177]. AMPAR antagonists are different from NMDAR antagonists in anticonvulsant activity. They have better anticonvulsant effects in fully kindled seizures, particularly having a synergistic effect with low-dose NMDA antagonists [178,179]. Furthermore, they have less behavioral side effects compared to those seen with NMDA antagonists in kindled animals. The pharmacological features of AMPAR antagonists in kindling rats suggest their potential in the treatment of partial seizures. In slices of lateral amygdala from humans with intractable TLE, blockade of AMPARs, but not NMDARs, abolished spontaneous interictal-like activity, indicating the role of AMPARs in the abnormal electrical network activity of the epileptic brain [180]. As AMPARs are involved in fast excitatory transmission, prolonged activation of these receptors is a critical process in the development and progression of epileptic seizures. Several animal studies have shown that AMPAR receptor antagonists can terminate self-sustained synchronized activity and neuroprotection in SE [181,182,183,184,185,186]. Perampanel is a noncompetitive AMPAR antagonist with a negative allosteric modulation effect [187] that decreases neuronal-synchronized epileptiform activity [172]. Perampanel has been shown to shorten the after-discharge duration in increased intensity of stimulation in the amygdala kindling model, whereas sodium channel blockers and an SV-2A ligand (levetiracetam) failed to show this effect [188]. Our previous study also expanded on perampanel’s AMPAR ionic mechanism in modulating neuronal excitability [189]. Moreover, in a recent study, perampanel, but not the NMDA antagonist amantadine, retarded development of spontaneous recurrent seizures in pilocarpine-induced SE and also reduced SE-induced late behavioral consequences [185]. Perampanel has been approved for the treatment of partial-onset epilepsy and generalized tonic-clonic epilepsy [190,191,192].
In conclusion, numerous ASMs have been found to have anti-seizure effects through various mechanisms, such as inhibiting presynaptic glutamate release and antagonizing ionotropic glutamate receptors. These medications include classic ASMs such as topiramate, felbamate, pregabalin, and lamotrigine, as well as NMDAR antagonists such as riluzole, memantine, ketamine, and AMPAR antagonists such as perampanel.

5. Treatment of SE by Glutamate Receptor Antagonists

Although NMDARs are targets for antiepileptic drugs, competitive NMDAR antagonists, such as D-2-amino-5-phosphonopen-tanoate (AP5) and 3-(2-carboxypiperazin-4-yl)propyl-l-phosphonate (CPP), have poor brain penetration [193,194], and their function may be dampened by the high concentrations of synaptic and extrasynaptic glutamate during epileptic seizures. Noncompetitive antagonists, such as dizocilpine (MK-801) and ketamine, have no such limitations [195]. An animal study revealed noncompetitive antagonist MK-801 to be superior to the competitive antagonist CPP and PH-sensitive site antagonist ifenprodil in terminating experimental SE [196]. Receptor trafficking is a major cause of refractory or super-refractory status epilepticus (SRSE) if treatment of SE is delayed. Intrasynaptic membrane GABAA receptors are internalized, while NMDARs accumulate and are upregulated in the post-synaptic membrane, in cases of prolonged SE [113,197]. Therefore, GABAergic medications sometimes fail to control SE. In such situations, ketamine, a noncompetitive NMDA antagonist, is an option for late, refractory SE. Ketamine has been shown to play a role in the treatment of SRSE in adults and children [198,199,200,201,202,203,204]. The seizure control rate ranges from 50% to as high as 90%. In a larger retrospective series in 68 patients with SRSE under midazolam infusion, 55 patients showed more than a 50% decrease in seizure burden and 43 patients completely ceased seizures within 24 h of starting add-on ketamine treatment [205]. Their treatment regimen did not alter intracranial pressure or cerebral blood flow. In addition, the advantage of ketamine is its lack of adverse cardiopulmonary depression effect [206,207].
Magnesium sulfate is a potential therapy for SE. As mentioned earlier, Mg2+ blocks NMDAR at a resting state, and it is only when Mg2+ is unblocked that the membrane can be depolarized, allowing a voltage-dependent inward flow of sodium (Na+) and Ca2+ ions that generate excitability. Magnesium sulfate administration in rats was associated with an increased seizure threshold and resistance to experimental seizures [208]. Intravenous magnesium treatment reduced refractory epilepsy with recurrent SE and relieved bouts of epilepsia partialis continua in adolescents with POLG-1 mutation [209]. An open-label, randomized, controlled study showed significantly lengthened durations of seizure-free periods by adding magnesium sulphate to adrenocorticotropic hormone in patients with infantile spasm [210]. In pregnant women, intravenous magnesium sulphate is the standard of care in the management of eclampsia [211].
In conclusion, several candidate medications, such as competitive and noncompetitive NMDAR antagonists, have been investigated on their ability to control SE (Table 1) in animal models based on the receptor trafficking theory, and they have shown positive effects. However, the genuine efficacy of these medications in clinical studies on patients with SE remains unknown. Further clinical trials using ketamine or magnesium sulfate are warranted to precisely determine their effects on SE and RSE.

6. Future Perspective of ASM

Kainate receptors (KARs) are among the glutamate ionotropic receptors whose mechanism in seizures is not clearly understood. This non-NMDA receptor is upregulated in astrocytes in response to SE, as increased expression of some subunits has been noted [212]. Although the significance of increased expression of the subunits is not clear, it means that selectively targeting astrocytic processes that contribute to glutamate release may hold potential for epilepsy therapy [213]. Furthermore, selurampanel (BGG492) is an AMPA/KAR antagonist and clinical trials for its use in treating partial and photosensitive epilepsy are ongoing [214]. A 12-week phase II study revealed a higher percentage of total partial seizure reduction over placebo for BGG492 150 mg given three times a day [215].
Genetic variations in patients with epilepsy must be taken into consideration in treating epilepsy. For example, genetic variations of the CYP2B6 alleles (*1 and *6) lead to different levels of ketamine plasma concentrations. Patients with *1/*1 genotypes may accelerate clearance of ketamine, yet decreased clearance was noted in patients with the *1/*6, and particularly, *6/*6 allele [216,217]. The variation of ketamine concentration may have an impact on the efficacy of treatment and possibly serious adverse effects. Regarding GRIN2A mutation-related epileptic encephalopathy, memantine was reported to have decreased the frequency and onset of seizures in a 3-year-old boy with heterozygous c.1083G  > A (p.Leu361 =) variant in GRIN2A [218]. In a de novo missense mutation (c.2434C > A; p.L812M) patient, add-on memantine reduced seizure burden and improved interictal EEG [219]. However, we have not yet known whether memantine is effective in all of these patient populations. Some factors such as genetic variant location, mechanism of gain-of-function (or loss-of-function), patient age, memantine dose, different serum and brain concentrations might influence its efficacy [220].
Overall, there are several aspects worthy of future ASM research, including clinical trials of AMPA/KAR antagonists, the use of memantine in patients with epilepsy with specific genetic variations or mutations, and the investigation of the underlying mechanisms of how these mutations impact glutamate receptors.

7. Conclusions

Knowing the pathophysiology of glutamate and glutamate receptors and their interactions is essential in facilitating the clinical management of epilepsy. Different conformations of NMDARs and AMPARs, as well as different ligand substrates, have distinct types and strengths of neurotransmission. Astrocytes play an important role in regulating glutamate, either through reuptake of synaptic glutamate or release of glutamate triggered by increased intracellular Ca2+ levels in astrocytes. In addition, receptor trafficking in prolonged seizures leads to altered receptor expression and localization, which results in drug-resistant SE. Modulation of glutamatergic signaling by various glutamate receptor antagonists has been shown to be effective for treatment of epilepsy. New mechanisms of ASM such as targeting kainate receptors and astrocytic processes are promising. Genome sequence exploration in some epileptic encephalopathies, and evidence of genetic epilepsy could help in the development of promising personalized treatment in precision medicine.

Author Contributions

Conceptualization, T.-S.C., T.-H.H., M.-C.L. and C.-W.H.; methodology, T.-S.C., T.-H.H. and C.-W.H.; validation, T.-S.C.; formal analysis, T.-H.H. and C.-W.H.; investigation, T.-S.C., T.-H.H., M.-C.L. and C.-W.H.; data curation, T.-S.C. and C.-W.H.; writing—original draft preparation, T.-S.C., T.-H.H., M.-C.L. and C.-W.H.; writing—review and editing, T.-S.C., T.-H.H., M.-C.L. and C.-W.H.; funding acquisition, C.-W.H. All authors have read and agreed to the published version of the manuscript.

Funding

This study was partly funded by a grant from Ministry of Science and Technology, Taiwan (MOST-109-2314-B-006-034-MY3, MOST-111-2314-B-006-103-MY2), and a grant from the National Cheng Kung University Hospital (NCKUH-11201005).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to express their appreciation for the comments from members of the staff of the Department of Neurology, National Cheng Kung University Hospital, Tainan, Taiwan.

Conflicts of Interest

The authors declare no conflict of interest, and 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. Scharfman, H.E. The neurobiology of epilepsy. Curr. Neurol. Neurosci. Rep. 2007, 7, 348–354. [Google Scholar] [CrossRef] [PubMed]
  2. Wo, Z.G.; Oswald, R.E. Unraveling the modular design of glutamate-gated ion channels. Trends Neurosci. 1995, 18, 161–168. [Google Scholar] [PubMed]
  3. Sobolevsky, A.I. Structure and gating of tetrameric glutamate receptors. J. Physiol. 2015, 593, 29–38. [Google Scholar] [CrossRef]
  4. Traynelis, S.F.; Wollmuth, L.P.; McBain, C.J.; Menniti, F.S.; Vance, K.M.; Ogden, K.K.; Hansen, K.B.; Yuan, H.; Myers, S.J.; Dingledine, R. Glutamate receptor ion channels: Structure, regulation, and function. Pharmacol. Rev. 2010, 62, 405–496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Lehre, K.P.; Danbolt, N.C. The number of glutamate transporter subtype molecules at glutamatergic synapses: Chemical and stereological quantification in young adult brain. J. Neurosci. 1998, 18, 8751–8757. [Google Scholar] [CrossRef] [Green Version]
  6. Krugler, P.; Schleyear, V. Developmental expression of glutamate transporter and glutamate dehydrogenase in astrocytes of the postnatal rat hippocampus. Hippocampus 2004, 14, 975–985. [Google Scholar] [CrossRef] [PubMed]
  7. Jia, M.; Njapo, S.A.; Rastogi, V.; Hedna, V.S. Taming glutamate excitotoxicity: Strategic pathway modulation for neuroprotection. CNS Drugs 2015, 29, 153–162. [Google Scholar] [CrossRef] [PubMed]
  8. Eid, T.; Gruenbaum, S.E.; Dhaher, R.; Lee, T.S.W.; Zhou, Y.; Danbolt, N.C. The glutamate-glutamine cycle in epilepsy. Adv. Neurobiol. 2016, 13, 351–400. [Google Scholar]
  9. Willard, S.S.; Koochekpour, S. Glutamate, glutamate receptors and downstream signaling pathways. Int. J. Biol. Sci. 2013, 9, 948–959. [Google Scholar] [CrossRef] [Green Version]
  10. Lerma, J.; Marques, J.M. Kainate receptors in health and disease. Neuron 2013, 80, 292–311. [Google Scholar] [CrossRef] [Green Version]
  11. Rodríguez-Moreno, A.; Herreras, O.; Lerma, J. Kainate receptors presynaptically downregulate GABAergic inhibition in the rat hippocampus. Neuron 1997, 19, 893–901. [Google Scholar] [CrossRef] [PubMed]
  12. Rodríguez-Moreno, A.; Sihra, T.S. Presynaptic kainate receptor facilitation of glutamate release involves protein kinase A in the rat hippocampus. J. Physiol. 2004, 557, 733–745. [Google Scholar] [CrossRef] [PubMed]
  13. Nakagawa, T. The biochemistry, ultrastructure, and subunit assembly mechanism of AMPA receptors. Mol. Neurobiol. 2010, 42, 161–184. [Google Scholar] [CrossRef] [Green Version]
  14. Hollmann, M.; Hartley, M.; Heinemann, S. Ca2+ permeability of KA-AMPA—Gated glutamate receptor channels depends on subunit composition. Science 1991, 252, 851–853. [Google Scholar] [CrossRef] [PubMed]
  15. Sans, N.; Vissel, B.; Petralia, R.S.; Wang, Y.X.; Chang, K.; Royle, G.A.; Wang, C.Y.; O’Gorman, S.; Heinemann, S.F.; Wenthold, R.J. Aberrant formation of glutamate receptor complexes in hippocampal neurons of mice lacking the GluR2 AMPA receptor subunit. J. Neurosci. 2003, 23, 9367–9373. [Google Scholar] [CrossRef] [Green Version]
  16. Hume, R.I.; Dingledine, R.; Heinemann, S.F. Identification of a site in glutamate receptor subunits that controls calcium permeability. Science 1991, 253, 1028–1031. [Google Scholar] [CrossRef]
  17. Sommer, B.; Köhler, M.; Sprengel, R.; Seeburg, P.H. RNA editing in brain controls a determinant of ion flow in glutamate-gated channels. Cell 1991, 67, 11–19. [Google Scholar] [CrossRef]
  18. Malinow, R.; Malenka, R.C. AMPA receptor trafficking and synaptic plasticity. Annu. Rev. Neurosci. 2002, 25, 103–126. [Google Scholar] [CrossRef] [Green Version]
  19. Cull-Candy, S.; Kelly, L.; Farrant, M. Regulation of Ca2+- permeable AMPA receptors: Synaptic plasticity and beyond. Curr. Opin. Neurobiol. 2006, 16, 288–297. [Google Scholar] [CrossRef]
  20. Anggono, V.; Huganir, R.L. Regulation of AMPA receptor trafficking and synaptic plasticity. Curr. Opin. Neurobiol. 2012, 22, 461–469. [Google Scholar] [CrossRef] [Green Version]
  21. Liu, S.J.; Zukin, R.S. Ca2+- permeable AMPA receptors in synaptic plasticity and neuronal death. Trends Neurosci. 2007, 30, 126–134. [Google Scholar] [CrossRef]
  22. Wiltgen, B.J.; Royle, G.A.; Gray, E.E.; Abdipranoto, A.; Thangthaeng, N.; Jacobs, N.; Saab, F.; Tonegawa, S.; Heinemann, S.F.; O’Dell, T.J.; et al. A role for calcium-permeable AMPA receptors in synaptic plasticity and learning. PLoS ONE 2010, 5, e12818. [Google Scholar] [CrossRef]
  23. Pellegrini-Giampietro, D.E.; Gorter, J.A.; Bennet, M.V.; Zukin, R.S. The GluR2 (GluR-B) hypothesis: Ca2+-permeable AMPA receptors in neurological disorders. Trends Neurosci. 1997, 20, 464–470. [Google Scholar] [CrossRef]
  24. Kwak, S.; Weiss, J.H. Calcium-permeable AMPA channels in neurodegenerative disease and ischemia. Curr. Opin. Neurobiol. 2006, 16, 281–287. [Google Scholar] [CrossRef]
  25. Russo, I.; Bonini, D.; Via, L.L.; Barlati, S.; Barbon, A. AMPA receptor properties are modulated in the early stages following pilocarpine-induced status epilepticus. Neuromolecular Med. 2013, 15, 324–338. [Google Scholar] [CrossRef]
  26. Hestrin, S.; Nicoll, R.A.; Perkel, D.J.; Sah, P. Analysis of excitatory synaptic action in pyramidal cells using whole-cell recording from rat hippocampal slices. J. Physiol. 1990, 422, 203–225. [Google Scholar] [CrossRef] [PubMed]
  27. Trussell, L.O.; Zhang, S.; Raman, I.M. Desensitization of AMPA receptors upon multiquantal neurotransmitter release. Neuron 1993, 10, 1185–1196. [Google Scholar] [CrossRef]
  28. Geiger, J.R.; Lübke, J.; Roth, A.; Frotscher, M.; Jonas, P. Submillisecond AMPA receptor-mediated signaling at a principal neuron-interneuron synapse. Neuron 1997, 18, 1009–1023. [Google Scholar] [CrossRef] [Green Version]
  29. Paoletti, P. Molecular basis of NMDA receptor functional diversity. Eur. J. Neurosci. 2011, 33, 1351–1365. [Google Scholar] [CrossRef] [PubMed]
  30. Skrenkova, K.; Hemelikova, K.; Kolcheva, M.; Kortus, S.; Kaniakova, M.; Krausova, B.; Horak, M. Structural features in the glycine-binding sites of the GluN1 and GluN3A subunits regulate the surface delivery of NMDA receptors. Sci. Rep. 2019, 9, 12303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Akazawa, C.; Shigemoto, R.; Bessho, Y.; Nakanishi, S.; Mizuno, N. Differential expression of five N-methyl-D-aspartate receptor subunit mRNAs in the cerebellum of developing and adult rats. J. Comp. Neurol. 1994, 347, 150–160. [Google Scholar] [CrossRef]
  32. Ogden, K.K.; Traynelis, S.F. New advances in NMDA receptor pharmacology. Trends Pharmacol. Sci. 2011, 32, 726–733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Dingledine, R.; Borges, K.; Bowie, D.; Traynelis, S.F. The glutamate receptor ion channels. Pharmacol. Rev. 1999, 51, 7–61. [Google Scholar] [PubMed]
  34. Mody, I.; Lambert, J.D.; Heinemann, U. Low extracellular magnesium induces epileptiform activity and spreading depression in rat hippocampal slices. J. Neurophysiol. 1987, 57, 869–888. [Google Scholar] [CrossRef]
  35. Gnegy, M.E. Ca2+/calmodulin signaling in NMDA-induced synaptic plasticity. Crit. Rev. Neurobiol. 2000, 14, 91–129. [Google Scholar] [CrossRef] [PubMed]
  36. Husi, H.; Ward, M.A.; Choudhary, J.S.; Blackstock, W.P.; Grant, S.G. Proteomic analysis of NMDAR-adhesion protein signaling complexes. Nat. Neurosci. 2000, 3, 661–669. [Google Scholar] [CrossRef] [Green Version]
  37. Massey, P.V.; Johnson, B.E.; Moult, P.R.; Auberson, Y.P.; Brown, M.W.; Molnar, E.; Collingridge, G.L.; Bashir, Z.I. Differential roles of NR2A and NR2B-containing NMDARs in cortical long-term potentiation and long-term depression. J. Neurosci. 2004, 24, 7821–7828. [Google Scholar] [CrossRef] [Green Version]
  38. Pinheiro, P.; Mulle, C. Kainate receptors. Cell Tissue Res. 2006, 326, 457–482. [Google Scholar] [CrossRef] [PubMed]
  39. Rogawski, M.A.; Gryder, D.; Castaneda, D.; Yonekawa, W.; Banks, M.K.; Lia, H. GluR5 kainate receptors, seizures, and the amygdala. Ann. N. Y. Acad. Sci. 2003, 985, 150–162. [Google Scholar] [CrossRef]
  40. Patel, S.; Meldrum, B.S.; Collins, J.F. Distribution of [3H]kainic acid and binding sites in the rat brain: In vivo and in vitro receptor autoradiography. Neurosci. Lett. 1986, 70, 301–307. [Google Scholar] [CrossRef]
  41. Bloss, E.B.; Hunter, R.G. Hippocampal kainate receptors. Vitam. Horm. 2010, 82, 167–184. [Google Scholar]
  42. Pressey, J.C.; Woodin, M.A. Kainate receptor regulation of synaptic inhibition in the hippocampus. J. Physiol. 2021, 599, 485–492. [Google Scholar] [CrossRef] [PubMed]
  43. Valbuena, S.; Lerma, J. Non-canonical Signaling, the Hidden Life of Ligand-Gated Ion Channels. Neuron 2016, 92, 316–329. [Google Scholar] [CrossRef] [Green Version]
  44. Mulle, C.; Crépel, V. Regulation and dysregulation of neuronal circuits by KARs. Neuropharmacology 2021, 197, 108699. [Google Scholar] [CrossRef]
  45. Ben-Ari, Y.; Lagowska, J. Epileptogenic action of intra-amygdaloid injection of kainic acid. Comptes Rendus Hebd. Séances Acad. Sci. Série. Sci. Nat. 1978, 287, 813–816. [Google Scholar]
  46. Ben-Ari, Y.; Tremblay, E.; Ottersen, O.P. Primary and secondary cerebral lesions produced by kainic acid injections in the rat. Comptes Rendus Hebd. Séances Acad. Sci. Série. Sci. Nat. 1979, 288, 991–994. [Google Scholar]
  47. Ben-Ari, Y.; Tremblay, E.; Ottersen, O.P.; Meldrum, B.S. The role of epileptic activity in hippocampal and “remote” cerebral lesions induced by kainic acid. Brain Res. 1980, 191, 79–97. [Google Scholar] [CrossRef] [PubMed]
  48. Bahn, S.; Volk, B.; Wisden, W. Kainate receptor gene expression in the developing rat brain. J. Neurosci. Off. J. Soc. Neurosci. 1994, 14, 5525–5547. [Google Scholar] [CrossRef]
  49. Werner, P.; Voigt, M.; Keinänen, K.; Wisden, W.; Seeburg, P.H. Cloning of a putative high-affinity kainate receptor expressed predominantly in hippocampal CA3 cells. Nature 1991, 351, 742–744. [Google Scholar] [CrossRef]
  50. Wisden, W.; Seeburg, P.H. A complex mosaic of high-affinity kainate receptors in rat brain. J. Neurosci. Off. J. Soc. Neurosci. 1993, 13, 3582–3598. [Google Scholar] [CrossRef] [Green Version]
  51. Ben-Ari, Y.; Cossart, R. Kainate, a double agent that generates seizures: Two decades of progress. Trends Neurosci. 2000, 23, 580–587. [Google Scholar] [CrossRef]
  52. Lévesque, M.; Langlois, J.M.P.; Lema, P.; Courtemanche, R.; Bilodeau, G.A.; Carmant, L. Synchronized gamma oscillations (30–50 Hz) in the amygdalo-hippocampal network in relation with seizure propagation and severity. Neurobiol. Dis. 2009, 35, 209–218. [Google Scholar] [CrossRef] [PubMed]
  53. Carta, M.; Fièvre, S.; Gorlewicz, A.; Mulle, C. Kainate receptors in the hippocampus. Eur. J. Neurosci. 2014, 39, 1835–1844. [Google Scholar] [CrossRef]
  54. Pahl, S.; Tapken, D.; Haering, S.C.; Hollmann, M. Traffiking of kaintae receptors. Membranes 2014, 4, 565–595. [Google Scholar] [CrossRef] [Green Version]
  55. Scholefield, C.L.; Atlason, P.T.; Jane, D.E.; Molnár, E. Assembly and trafficking of homomeric and heteromeric kainate receptors with impaired ligand binding sites. Neurochem. Res. 2019, 44, 585–599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Nair, J.D.; Wilkinson, K.A.; Henley, J.M.; Mellor, J.R. Kainate receptors and synaptic plasticity. Neuropharmacology 2021, 196, 108540. [Google Scholar] [CrossRef]
  57. Merlin, L.R.; Wong, R.S. Role of Group I metabotropic glutamate receptors in the patterning of epileptiform activities in vitro. J. Neurophysiol. 1997, 78, 539–544. [Google Scholar] [CrossRef] [Green Version]
  58. Mannaioni, G.; Marino, M.J.; Valenti, O.; Traynelis, S.F.; Conn, P.J. Metabotropic glutamate receptors 1 and 5 differentially regulate CA1 pyramidal cell function. J. Neurosci. 2001, 21, 5925–5934. [Google Scholar] [CrossRef] [Green Version]
  59. Notenboom, R.G.E.; Hampson, D.R.; Jansen, G.H.; van Rijen, P.C.; van Veelen, C.W.M.; van Nieuwenhuizen, O.; de Graan, P.N.E. Up-regulation of hippocampal metabotropic glutamate receptor 5 in temporal lobe epilepsy patients. Brain 2006, 129, 91–107. [Google Scholar] [CrossRef] [Green Version]
  60. Hermans, E.; Challiss, R.A. Structural, signalling and regulatory properties of the group I metabotropic glutamate receptors: Prototypic family C G-protein-coupled receptors. Biochem. J. 2001, 359, 465–484. [Google Scholar] [CrossRef] [PubMed]
  61. Tang, F.R.; Chia, S.C.; Chen, P.M.; Gao, H.; Lee, W.L.; Yeo, T.S.; Burgunder, J.M.; Probst, A.; Sim, M.K.; Ling, E.A. Metabotropic glutamate receptor 2/3 in the hippocampus of patients with mesial temporal lobe epilepsy, and of rats and mice after pilocarpine-induced status epilepticus. Epilepsy Res. 2004, 59, 167–180. [Google Scholar] [CrossRef] [PubMed]
  62. Pacheco Otalora, L.F.; Couoh, J.; Shigamoto, R.; Zarei, M.M.; Garrido Sanabria, E.R. Abnormal mGluR2/3 expression in the perforant path termination zones and mossy fibers of chronically epileptic rats. Brain Res. 2006, 1098, 170–185. [Google Scholar] [CrossRef] [PubMed]
  63. Aronica, E.; Gorter, J.A.; Ijlst-Keizers, H.; Rozemuller, A.J.; Yankaya, B.; Leenstra, S.; Troost, D. Expression and functional role of mGluR3 and mGluR5 in human astrocytes and glioma cells: Opposite regulation of glutamate transporter proteins. Eur. J. Neurosci. 2003, 17, 2106–2118. [Google Scholar] [CrossRef] [PubMed]
  64. Aronica, E.; van Vliet, E.A.; Mayboroda, O.A.; Troost, D.; da Silva, F.H.; Gorter, J.A. Upregulation of metabotropic glutamate receptor subtype mGluR3 and mGluR5 in reactive astrocytes in a rat model of mesial temporal lobe epilepsy. Eur. J. Neurosci. 2000, 12, 2333–2344. [Google Scholar] [CrossRef]
  65. Pitsch, J.; Schoch, S.; Gueler, N.; Flor, P.J.; van der Putten, H.; Becker, A.J. Functional role of mGluR1 and mGluR4 in pilocarpine-induced temporal lobe epilepsy. Neurobiol. Dis. 2007, 26, 623–633. [Google Scholar] [CrossRef]
  66. Hamilton, N.B.; Attwell, D. Do astrocytes really exocytose neurotransmitters? Nat. Rev. Neurosci. 2010, 11, 227–238. [Google Scholar] [CrossRef]
  67. Parri, H.R.; Gould, T.M.; Crunelli, V. Spontaneous astrocytic Ca2+ oscillations in situ drive NMDAR-mediated neuronal excitation. Nat. Neurosci. 2001, 4, 803–812. [Google Scholar] [CrossRef]
  68. 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]
  69. Angulo, M.C.; Kozlov, A.S.; Charpak, S.; Audinat, E. Glutamate Released from Glial Cells Synchronizes Neuronal Activity in the Hippocampus. J. Neurosci. 2004, 24, 6920–6927. [Google Scholar] [CrossRef] [Green Version]
  70. Perea, G.; Araque, A. Properties of Synaptically Evoked Astrocyte Calcium Signal Reveal Synaptic Information Processing by Astrocytes. J. Neurosci. 2005, 25, 2192–2203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Araque, A.; Sanzgiri, R.P.; Parpura, V.; Haydon, P.G. Calcium elevation in astrocytes causes an NMDA receptor-dependent increase in the frequency of miniature synaptic currents in cultured hippocampal neurons. J. Neurosci. 1998, 18, 6822–6829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Jourdain, P.; Bergersen, L.H.; Bhaukaurally, K.; Bezzi, P.; Santello, M.; Domercq, M.; Matute, C.; Tonello, F.; Gundersen, V.; Volterra, A. Glutamate exocytosis from astrocytes controls synaptic strength. Nat. Neurosci. 2007, 10, 331–339. [Google Scholar] [CrossRef]
  73. Perea, G.; Araque, A. Astrocytes Potentiate Transmitter Release at Single Hippocampal Synapses. Science 2007, 317, 1083–1086. [Google Scholar] [CrossRef]
  74. Alcoreza, O.B.; Patel, D.C.; Tewari, B.P.; Sontheiner, H. Dysregulation of abient glutamate and glutamate receptors in epilepsy: An astrocytic perspective. Front. Neurol. 2021, 12, 652159. [Google Scholar] [CrossRef]
  75. Losi, G.; Mariotti, L.; Carmignoto, G. GABAergic interneuron to astrocyte signalling: A neglected form of cell communication in the brain. Philos. Trans. R. Soc. Lond. B Bio. Sci. 2014, 369, 20130609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Kang, J.; Jiang, L.; Goldman, S.A.; Nedergaard, M. Astrocyte-mediated potentiation of inhibitory synaptic transmission. Nat. Neurosci. 1998, 1, 683–692. [Google Scholar] [CrossRef]
  77. Esposito, G.; Scuderi, C.; Valenza, M.; Togna, G.I.; Latina, V.; De Filippis, D.; Cipriano, M.; Carratu, M.R.; Iuvone, T.; Steardo, L. Cannabidiol reduces AB-induced neuroinflammation and promotes hippocampal neurogenesis through PPARү involvement. PLoS ONE 2011, 6, e28668. [Google Scholar] [CrossRef] [PubMed]
  78. Atalay, S.; Jarocka-Karpowicz, I.; Skrzydlewska, E. Antioxidative and Anti-Inflammatory Properties of Cannabidiol. Antioxidants 2019, 9, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Pellati, F.; Borgonetti, V.; Brighenti, V.; Biagi, M.; Benvenuti, S.; Corsi, L. Cannabis sativa L. and nonpsychoacitve cannabinoids: Their chemistry and role against oxidative stress, inflammation and cancer. Biomed Res. Int. 2018, 2018, 1691428. [Google Scholar] [CrossRef] [Green Version]
  80. Kozela, E.; Pietr, M.; Juknat, A.; Rimmerman, N.; Levy, R.; Vogel, Z. Cannabinoids delta(9)-tetrahydrocannabinol and cannabidiol differentially inhibit the lipopolysaccharide-activated NF-kappaB and interferon-beta/STAT proinflammatory pathways in BV-2 microglial cells. J. Biol. Chem. 2010, 285, 1616–1626. [Google Scholar] [CrossRef] [Green Version]
  81. Schiavon, A.P.; Soares, L.M.; Bonato, J.M.; Milani, H.; Guimarães, F.S.; Weffort de Oliveira, R.M. Protective effects of cannabidiol against hippocampal cell death and cognitive impairment induced by bilateral common carotid artery occlusion in mice. Neurotox. Res. 2014, 26, 307–316. [Google Scholar] [CrossRef]
  82. Friedman, L.K.; Wongvravit, J.P. Anticonvulsant and neuroprotective effects of cannabidiol during the juvenile period. J. Neuropathol. Exp. Neurol. 2018, 77, 904–919. [Google Scholar] [CrossRef] [PubMed]
  83. Gray, R.A.; Whalley, B.J. The proposed mechanism of action of CBD in epilepsy. Epileptic Disord. 2020, 22, 10–15. [Google Scholar]
  84. Zavala-Tecuapetla, C.; Luna-Munguia, H.; López-Meraz, M.L.; Cuellar-Herrera, M. Advances and Challenges of Cannabidiol as an Anti-Seizure Strategy: Preclinical Evidence. Int. J. Mol. Sci. 2022, 23, 16181. [Google Scholar] [CrossRef] [PubMed]
  85. Premoli, M.; Aria, F.; Bonini, S.A.; Maccarinelli, G.; Gianoncelli, A.; Pina, S.D.; Tambaro, S.; Memo, M.; Mastinu, A. Cannabidiol: Recent advances and new insights for neuropsychiatric disorders treatment. Life Sci. 2019, 224, 120–127. [Google Scholar] [CrossRef]
  86. Pretzsch, C.M.; Freyberg, J.; Voinescu, B.; Lythgoe, D.; Horder, J.; Mendez, M.A.; Wichers, R.; Ajram, L.; Ivin, G.; Heasman, M.; et al. Effects of cannabidiol on brain excitation and inhibition systems; a randomised placebo-controlled single dose trial during magnetic resonance spectroscopy in adults with and without autism spectrum disorder. Neuropsychopharmacology 2019, 44, 1398–1405. [Google Scholar] [CrossRef] [Green Version]
  87. Ryberg, E.; Larsson, N.; Sjogren, S.; Hjorth, S.; Hermansson, N.O.; Leonova, J.; Elebring, T.; Nilsson, K.; Drmota, T.; Greasley, P.J. The orphan receptor GPR55 is a novel cannabinoid receptor. Br. J. Pharmacol. 2007, 152, 1092–1101. [Google Scholar] [CrossRef] [PubMed]
  88. Guggenhuber, S.; Monory, K.; Lutz, B.; Klugmann, M. AAV vector-mediated overexpression of CB1 cannabinoid receptor in pyramidal neurons of the hippocampus protects against seizure-induced excitoxicity. PLoS ONE 2010, 5, e15707. [Google Scholar] [CrossRef]
  89. Ruehle, S.; Remmers, F.; Romo-Parra, H.; Massa, F.; Wickert, M.; Wörtge, S.; Häring, M.; Kaiser, N.; Marsicano, G.; Pape, H.C.; et al. Cannabinoid CB1 receptor in dorsal telencephalic glutamatergic neurons: Distinctive sufficiency for hippocampus-dependent and amygdala-dependent synaptic and behavioral functions. J. Neurosci. 2013, 33, 10264–10277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Sugaya, Y.; Kano, M. Endocannabinoid-Mediated Control of Neural Circuit Excitability and Epileptic Seizures. Front. Neural Circuits 2022, 15, 781113. [Google Scholar] [CrossRef]
  91. Devinsky, O.; Marsh, E.; Friedman, D.; Thiele, E.; Laux, L.; Sullivan, J.; Miller, I.; Flamini, R.; Wilfong, A.; Filloux, F.; et al. Cannabidiol in patients with treatment-resistant epilepsy: An open-label interventional trial. Lancet Neurol. 2016, 15, 270–278. [Google Scholar] [CrossRef]
  92. Devinsky, O.; Cross, J.H.; Laux, L.; Marsh, E.; Miller, I.; Nabbout, R.; Scheffer, I.E.; Thiele, E.A.; Stephen Wright, S.; Cannabidiol in Dravet Syndrome Study Group. Trial of cannabidiol for drug-resistant seizures in the Dravet syndrome. N. Engl. J. Med. 2017, 376, 2011–2020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Devinsky, O.; Patel, A.D.; Cross, J.H.; Villanueva, V.; Wirrell, E.C.; Privitera, M.; Greenwood, S.M.; Roberts, C.; Checketts, D.; VanLandingham, K.E.; et al. Effect of cannabidiol on drop seizure in the Lennox-Gastaut syndrome. N. Engl. J. Med. 2018, 378, 1888–1897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Miller, I.; Scheffer, I.E.; Gunning, B.; Sanchez-Carpintero, R.; Gil-Nagel, A.; Perry, M.S.; Saneto, R.P.; Checketts, D.; Dunayevich, E.; Knappertz, V.; et al. Dose-ranging effect of adjunctive oral cannabidiol vs placebo on convulsive seizure frequency in Dravet syndrome. JAMA Neurol. 2020, 77, 613–621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Thiele, E.A.; Bebin, E.M.; Bhathal, H.; Jansen, F.E.; Kotulska, K.; Lawson, J.A.; O’Callaghan, F.J.; Wong, M.; Sahebkar, F.; Checketts, D.; et al. Add-on cannabidiol treatment for drug-resistant seizures in tuberous sclerosis complex: A placebo-controlled randomized clinical trial. JAMA Neurol. 2021, 78, 285–292. [Google Scholar] [CrossRef] [PubMed]
  96. Grunelli, V.; Forda, S.; Kelly, J.S. The reversal potential of excitatory amino acid action on granule cells of the rat dentate gyrus. J. Physiol. 1984, 351, 327–342. [Google Scholar] [CrossRef]
  97. Duchen, M.R.; Burton, N.R.; Biscore, T.J. An intracellular study of the interaction of N-methyl-D-aspartate with ketamine in the mouse hippocampal slice. Brain Res. 1985, 342, 149–153. [Google Scholar] [CrossRef]
  98. Peet, M.J.; Gregersen, H.; Mclennan, H. 2-Amino-5-phosphonovalerate and CO2+ selectively block depolarization and burst firing of rat hippocampal CA1 pyramidal neurons by N-methyl-D-aspartate. Neuroscience 1986, 17, 635–641. [Google Scholar] [CrossRef]
  99. McDonough, J.H., Jr.; Shih, T.M. Neuropharmacological mechanisms of nerve agent-induced seizure and neuropathology. Neurosci. Biobehav. Rev. 1997, 21, 559–579. [Google Scholar] [CrossRef]
  100. Dorandeu, F.; Barbier, L.; Dhote, F.; Testylier, G.; Carpentier, P. Ketamine combinations for the field treatment of soman-induced self-sustaining status epilepticus. Review of current data and perspectives. Chem. Biol. Interact. 2013, 203, 154–159. [Google Scholar] [CrossRef]
  101. Dubey, V.; Dey, S.; Dixit, A.B.; Tripathi, M.; Chandra, P.S.; Banerjee, J. Differential glutamate receptor expression and function in the hippocampus, anterior temporal lobe and neocortex in a pilocarpine model of temporal lobe epilepsy. Exp. Neurol. 2022, 347, 113916. [Google Scholar] [CrossRef] [PubMed]
  102. Vieira, M.; Yong, X.L.H.; Roche, K.W.; Anggono, V. Regulation of NMDA glutamate receptor functions by the GluN2 subunits. J. Neurochem. 2020, 154, 121–143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Martel, M.A.; Ryan, T.J.; Bell, K.F.S.; Fowler, J.H.; McMahon, A.; Al-Mubarak, B.; Komiyama, N.H.; Horsburgh, K.; Kind, P.C.; Grant, S.G.N.; et al. The subtype of GluN2 C-terminal domain determines the response to excito_toxic insults. Neuron 2012, 74, 543–556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Bellone, C.; Nicoll, R.A. Rapid bidirectional switching of synap_tic NMDA receptors. Neuron 2007, 55, 779–785. [Google Scholar] [CrossRef] [Green Version]
  105. Ryan, T.J.; Kopanitsa, M.V.; Indersmitten, T.; Nithianantharajah, J.; Afinowi, N.O.; Pettit, C.; Stanford, L.E.; Sprengel, R.; Saksida, L.M.; Bussey, T.J.; et al. Evolution of GluN2A/B cytoplasmic domains diversified vertebrate synaptic plasticity and behavior. Nat. Neurosci. 2013, 16, 25–32. [Google Scholar] [CrossRef]
  106. Parsons, M.P.; Raymond, L.A. Extrasynaptic NMDA receptor involvement in central nervous system disorders. Neuron 2014, 82, 279–293. [Google Scholar] [CrossRef] [Green Version]
  107. Aarts, M.; Liu, Y.; Liu, L.; Besshoh, S.; Arundine, M.; Gurd, J.W.; Wang, Y.T.; Salter, M.W.; Tymianski, M. Treatment of ischemic brain damage by per_turbing NMDA receptor- PSD-95 protein interactions. Science 2002, 298, 846–850. [Google Scholar] [CrossRef]
  108. Ittner, L.M.; Gotz, J. Amyloid-beta and tau–a toxic pas de deux in Alzheimer’s disease. Nat. Rev. Neurosci. 2011, 12, 65–72. [Google Scholar] [CrossRef]
  109. Ballarin, B.; Tymianski, M. Discovery and development of NA-1 for the treatment of acute ischemic stroke. Acta Pharmacol. Sin. 2018, 39, 661–668. [Google Scholar] [CrossRef] [Green Version]
  110. Hill, M.D.; Martin, R.H.; Mikulis, D.; Wong, J.H.; Silver, F.L.; terBrugge, K.G.; Milot, G.; Clark, W.M.; Macdonald, R.L.; Kelly, M.E.; et al. Safety and efficacy of NA-1 in pa_tients with iatrogenic stroke after endovascular aneurysm repair (ENACT): A phase 2, randomised, double-blind, placebo-controlled trial. Lancet Neurol. 2012, 11, 942–950. [Google Scholar] [CrossRef]
  111. Pagano, J.; Giona, F.; Beretta, S.; Verpelli, C.; Sala, C. N-methyl-d-aspartate receptor function in neuronal and synaptic development and signaling. Curr. Opin. Pharmacol. 2021, 56, 93–101. [Google Scholar] [CrossRef] [PubMed]
  112. Wasterlain, C.G.; Chen, J.W. Mechanistic and pharmacologic aspects of status epilepticus and its treatment with new antiepileptic drugs. Epilepsia 2008, 49, 63–73. [Google Scholar] [CrossRef] [PubMed]
  113. Naylor, D.E.; Liu, H.; Niquet, J.; Wasterlain, C.G. Rapid surface accumulation of NMDA receptors increases glutamatergic excitation during status epilepticus. Neurobiol. Dis. 2013, 54, 225–238. [Google Scholar] [CrossRef] [Green Version]
  114. Kärkkäinen, O.; Kupila, J.; Häkkinen, M.; Laukkanen, V.; Tupala, E.; Kautiainen, H.; Tiihonen, J.; Storvik, M. AMPA receptors in post-mortem brains of Cloninger type 1 and 2 alcoholics: A whole-hemisphere autoradiography study. Psychiatry Res. 2013, 214, 429–434. [Google Scholar] [CrossRef] [PubMed]
  115. Yamazaki, M.; Fukaya, M.; Hashimoto, K.; Yamasaki, M.; Tsujita, M.; Itakura, M.; Abe, M.; Natsume, R.; Takahashi, M.; Kano, M.; et al. TARPs gamma-2 and gamma-7 are essential for AMPA receptor expression in the cerebellum. Eur. J. Neurosci. 2010, 31, 2204–2220. [Google Scholar] [CrossRef]
  116. Kimura, M.; Sawada, K.; Miyagawa, T.; Kuwada, M.; Katayama, K.; Nishizawa, Y. Role of glutamate receptors and voltage-dependent calcium and sodium channels in the extracellular glutamate/aspartate accumulation and subsequent neuronal injury induced by oxygen/glucose deprivation in cultured hippocampal neurons. J. Pharmacol. Exp. Ther. 1998, 285, 178–185. [Google Scholar]
  117. Lamanauskas, N.; Nistril, A. Riluzole blockes persistent Na+ and Ca2+ currents and modulates release of glutamate via presynaptic NMDA receptors on neonatal rat hypoglossal motorneurons in vitro. Eur. J. Neurosci. 2008, 27, 2501–2514. [Google Scholar] [CrossRef] [PubMed]
  118. Falcon-Moya, R.; Sihra, T.S.; Rodríguez-Moreno, A. Kainate receptors: Role in epilepsy. Front. Mol. Neurosci. 2018, 11, 217. [Google Scholar] [CrossRef]
  119. Negrete-Díaz, J.V.; Falcón-Moya, R.; Rodríguez-Moreno, A. Kainate receptors: From synaptic activity to disease. FEBS J. 2022, 289, 5074–5088. [Google Scholar] [CrossRef]
  120. Rodrıguez-Moreno, A.; López-García, J.C.; Lerma, J. Two populations of kainate receptors with separate signaling mechanisms in hippocampal interneurons. Proc. Natl. Acad. Sci. USA 2000, 97, 1293–1298. [Google Scholar] [CrossRef] [Green Version]
  121. Ben-Ari, Y.; Crepel, V.; Represa, A. Seizures beget seizures in temporal lobe epilepsies: The boomerang effects of newly formed aberrant kainatergic synapses. Epilepsy Curr. 2008, 8, 68–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Braga, M.F.; Aroniadou-Anderjaska, V.; Li, H. The physiological role of kainate receptors in the amygdala. Mol. Neurobiol. 2004, 30, 127–141. [Google Scholar] [CrossRef] [PubMed]
  123. Mazarati, A.M.; Wasterlain, C.G. N-methyl-D-asparate receptor antagonists abolish the maintenance phase of self-sustaining status epilepticus in rat. Neurosci. Lett. 1999, 265, 187–190. [Google Scholar] [CrossRef]
  124. Borris, D.J.; Bertram, E.H.; Kapur, J. Ketamine controls prolonged status epilepticus. Epilepsy Res. 2000, 42, 117–122. [Google Scholar] [CrossRef] [Green Version]
  125. Dorandeu, F.; Dhote, F.; Barbier, L.; Baccus, B.; Testylier, G. Treatment of status epilepticus with ketamine, are we there yet? CNS Neurosci. Ther. 2013, 19, 411–427. [Google Scholar] [CrossRef]
  126. Ryley Parrish, R.; Albertson, A.J.; Buckingham, S.C.; Hablitz, J.J.; Mascia, K.L.; Davis Haselden, W.; Lubin, F.D. Status epilepticus triggers early and late alterations in brain-derived neurotrophic factor and NMDA glutamate receptor Grin2b DNA methylation levels in the hippocampus. Neuroscience 2013, 248, 602–619. [Google Scholar] [CrossRef] [Green Version]
  127. Schratt, G.M.; Tuebing, F.; Nigh, E.A.; Kane, C.G.; Sabatini, M.E.; Kiebler, M.; Greenberg, M.E. A brain-specific micro-RNA regulates dendritic spine development. Nature 2006, 439, 283–289. [Google Scholar] [CrossRef]
  128. Jimenez-Mateos, E.M.; Engel, T.; Merino-Serrais, P.; McKiernan, R.C.; Tanaka, K.; Mouri, G.; Sano, T.; O’Tuathaigh, C.; Waddington, J.L.; Prenter, S.; et al. Silencing microRNA-134 produces neuroprotective and prolonged seizure suppressive effects. Nat. Med. 2012, 18, 1087–1094. [Google Scholar] [CrossRef] [Green Version]
  129. Ohba, C.; Shiina, M.; Tohyama, J.; Haginoya, K.; Lerman-Sagie, T.; Okamoto, N.; Blumkin, L.; Lev, D.; Mukaida, S.; Nozaki, F.; et al. GRIN1 mutations cause encephalopathy with infantile-onset epilepsy, and hyperkinetic and stereotyped movement disorders. Epilepsia 2015, 56, 841–848. [Google Scholar] [CrossRef]
  130. Lemke, J.R.; Hendricks, R.; Geider, K.; Laube, B.; Schwake, M.; Harvey, R.J.; James, V.M.; Pepler, A.; Steiner, I.; Hörtnagel, K.; et al. GRIN2B mutation in West syndrome and intellectual disability with focal epilepsy. Ann. Neurol. 2014, 75, 147–154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. XiangWei, W.S.; Kannan, V.; Xu, Y.C.; Kosobucki, G.J.; Anthony, J.; Schulien, A.J.; Kusumoto, H.; Moufawad El Achkar, C.; Bhattacharya, S.; Lesca, G.; et al. Heterogeneous clinical and functional features of GRIN2D-related developmental and epileptic encephalopathy. Brain 2019, 142, 3009–3027. [Google Scholar] [CrossRef]
  132. Lemke, J.R.; Lal, D.; Reinthaler, E.M.; Steiner, I.; Nothnagel, M.; Alber, M.; Geider, K.; Laube, B.; Schwake, M.; Finsterwalder, K.; et al. Mutations in GRIN2A cause idiopathic focal epilepsy with rolandic spikes. Nat. Genet. 2013, 45, 1067–1072. [Google Scholar] [CrossRef]
  133. Xu, X.X.; Liu, X.R.; Fan, C.Y.; Jin-Xing Lai, J.X.; Shi, Y.W.; Yang, W.; Su, T.; Xu, J.Y.; Luo, J.H.; Liao, W.P. Functional Investigation of a GRIN2A Variant Associated with Rolandic Epilepsy. Neurosci. Bull. 2018, 34, 237–246. [Google Scholar] [CrossRef] [PubMed]
  134. Gu, X.; Zhou, Y.; Hu, X.; Gu, Q.; Wu, X.; Cao, M.; Ke, K.; Liu, C. Reduced numbers of cortical GABA-immunoreactive neurons in the chronic D-galactose treatment model of brain aging. Neurosci. Lett. 2013, 549, 82–86. [Google Scholar] [CrossRef] [PubMed]
  135. Megías, M.; Emri, Z.; Freund, T.F.; Gulyás, A.I. Total number and distribution of inhibitory and excitatory synapses on hippocampal CA1 pyramidal cells. Neuroscience 2001, 102, 527–540. [Google Scholar] [CrossRef] [PubMed]
  136. Gataullina, S.; Bienvenu, T.; Nabbout, R.; Huberfeld, G.; Dulac, O. Gene mutations in paediatric epilepsies cause NMDA-pathy, and phasic and tonic GABA-pathy. Dev. Med. Child Neurol. 2019, 61, 891–898. [Google Scholar] [CrossRef] [PubMed]
  137. Newcomer, J.W.; Farber, N.B.; Olney, J.W. NMDA receptor function, memory and brain aging. Dialogues Clin. Neurosci. 2000, 2, 219–232. [Google Scholar] [CrossRef]
  138. Soto, D.; Altafaj, X.; Sindreu, C.; Bayés, A. Glutamate receptor mutations in psychiatric and neurodevelopmental disorders. Commun. Integr. Biol. 2014, 7, e27887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Salpietro, V.; Dixon, C.L.; Guo, H.; Bello, O.D.; Vandrovcova, J.; Efthymiou, S.; Maroofian, R.; Heimer, G.; Burglen, L.; Valence, S.; et al. AMPA receptor GluA2 subunit defects are a cause of neurodevelopmental disorders. Nat. Commun. 2019, 10, 3094. [Google Scholar] [CrossRef] [Green Version]
  140. Jia, Z.; Agopyan, N.; Miu, P.; Xiong, Z.; Henderson, J.; Gerlai, R.; Taverna, F.A.; Velumian, A.; MacDonald, J.; Carlen, P.; et al. Enhanced LTP in mice deficient in the AMPA receptor GluR2. Neuron 1996, 17, 945–956. [Google Scholar] [CrossRef] [Green Version]
  141. Dalmau, J.; Gleichman, A.J.; Hughes, E.G.; Rossi, J.E.; Peng, X.; Lai, M.; Dessain, S.K.; Rosenfeld, M.R.; Balice-Gordon, R.; Lynch, D.R. Anti-NMDA-receptor encephalitis: Case series and analysis of the effects of antibodies. Lancet Neurol. 2008, 7, 1091–1098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Hughes, E.G.; Peng, X.; Gleichman, A.J.; Lai, M.; Zhou, L.; Tsou, R.; Parsons, T.D.; Lynch, D.R.; Dalmau, J.; Balice-Gordon, R.J. Cellular and synaptic mechanisms of anti-NMDA receptor encephalitis. J. Neurosci. 2010, 30, 5866–5875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Chen, T.S.; Lai, M.C.; Huang, H.I.; Wu, S.N.; Huang, C.W. Immunity, Ion Channels and Epilepsy. Int. J. Mol. Sci. 2022, 23, 6446. [Google Scholar] [CrossRef]
  144. Sveinbjornsdottir, S.; Sander, J.W.; Upton, D.; Thompson, P.J.; Patsalos, P.N.; Hirt, D.; Emre, M.; Lowe, D.; Duncan, J.S. The excitatory amino acid antagonist D-CPP-ene (SDZ EAA-494) in patients with epilepsy. Epilepsy Res. 1993, 16, 165–174. [Google Scholar] [CrossRef] [PubMed]
  145. Manto, M.; Laute, M.A.; Aguera, M.; Rogemond, V.; Pandolfo, M.; Honnorat, J. Effects of anti-glutamic acid decarboxylase antibodies associated with neurological diseases. Ann. Neurol. 2007, 61, 544–551. [Google Scholar] [CrossRef] [Green Version]
  146. Manto, M.; Dalmau, J.; Didelot, A.; Rogemond, V.; Honnorat, J. In vivo effects of antibodies from patients with anti-NMDA receptor encephalitis: Further evidence of synaptic glutamatergic dysfunction. Orphanet J. Rare Dis. 2010, 5, 31. [Google Scholar] [CrossRef] [Green Version]
  147. Wolf, J.A.; Moyer, J.T.; Lazarewicz, M.T.; Contreras, D.; Benoit-Marand, M.; O’Donnell, P.; Finkel, L.H. NMDA/AMPA ratio impacts state transitions and entrainment to oscillations in a computational model of the nucleus accumbens medium spiny projection neuron. J. Neurosci. 2005, 25, 9080–9095. [Google Scholar] [CrossRef] [Green Version]
  148. Corona, J.C.; Tapia, R. AMPA receptor activation, but not the accumulation of endogeneous extracellular glutamate, induces paralysis and motor neuron death in rat spinal cord in vivo. J. Neurochem. 2004, 89, 988–997. [Google Scholar] [CrossRef]
  149. Seki, M.; Suzuki, S.; Iizuka, T.; Shimizu, T.; Nihei, Y.; Suzuki, N.; Dalmau, J. Neurological response to early removal of ovarian teratoma in antiNMDA-R encephalitis. J. Neurol. Neurosurg. Psychiatry 2008, 79, 324–326. [Google Scholar] [CrossRef]
  150. Faught, E.; Wilder, B.J.; Ramsay, R.E.; Reife, R.A.; Kramer, L.D.; Pledger, G.W.; Karim, R.M. Topiramate placebo-controlled dose-ranging trial in refractory partial epilepsy using 200-, 400-, 600-mg dosages. Neurology 1996, 46, 1684–1690. [Google Scholar] [CrossRef]
  151. Privitera, M.; Fincham, R.; Penry, J.; Reife, R.; Kramer, L.; Pledger, G.; Karim, R. Topiramate placebo-controlled dose-ranging trial in refractory partial epilepsy using 600-, 800-, and 1,000-mg daily dosages. Neurology 1996, 46, 1678–1683. [Google Scholar] [CrossRef]
  152. Severt, L.; Coulter, D.A.; Sombati, S. Topiramate selectively blocks kainite currentts in cultured hippocampal neurons. Epilepsia 1995, 36, s38. [Google Scholar]
  153. Gibbs, J.W.; Sombati, S.; DeLorenzo, R.I.; Coulter, D.A. Cellular actions of topiramate: Blockade of kainate-evoked inward currents in cultured hippocampal neurons. Epilepsia 2000, 41, 10–16. [Google Scholar] [CrossRef] [PubMed]
  154. Zhang, X.; Velumian, A.A.; Jones, O.T.; Carlen, P.L. Modulation of high-voltage activated calcium channels in dentate granule cells by topiramate. Epilepsia 2000, 41, 52–60. [Google Scholar] [CrossRef]
  155. Rho, J.M.; Donevan, S.D.; Rogawski, M.A. Mechanism of action of the anticonvulsant felbamate: Opposing effects on N-methyl-D-aspartate and g-aminobutyric acid A receptors. Ann. Neuro. 1994, 35, 229–234. [Google Scholar] [CrossRef] [PubMed]
  156. McCabe, R.T.; Wasterlain, C.G.; Kucharczyk, N.; Sofia, R.D.; Vogel, J.R. Evidence for anticonvulsant and neuroprotectant action of felbamate mediated by strychnine-insensitive glycine receptors. J. Pharmacol. Exp. Ther. 1993, 264, 1248–1252. [Google Scholar]
  157. Dooley, D.J.; Taylor, C.P.; Donevan, S.; Feltner, D. Ca2+ channel α2δ ligands: Novel modulators of neurotransmission. Trends Pharmacol. Sci. 2007, 28, 75–82. [Google Scholar] [CrossRef]
  158. Wang, S.J.; Huang, C.C.; Hsu, K.S.; Tsai, J.J.; Gean, P.W. Inhibition of N-type calcium currents by lamotrigine in rat amygdalar neurones. Neuroreport 1996, 7, 3037–3040. [Google Scholar] [CrossRef]
  159. Dibue, M.; Kamp, M.A.; Alpdogan, S.; Tevoufouet, E.E.; Neiss, W.F.; Hescheler, J.; Schneider, T. Cav2.3 (R-type) calcium channels are critical for mediating anticonvulsive and neuroprotective properties of lamotrigine in vivo. Epilepsia 2013, 54, 1542–1550. [Google Scholar] [CrossRef]
  160. Prakriya, M.; Mennerick, S. Selective Depression of Low–Release Probability Excitatory Synapses by Sodium Channel Blockers. Neuron 2000, 26, 671–682. [Google Scholar] [CrossRef] [Green Version]
  161. Debono, M.-W.; Le Guern, J.; Canton, T.; Doble, A.; Pradier, L. Inhibition by riluzole of electrophysiological responses mediated by rat kainate and NMDA receptors expressed in Xenopus oocytes. Eur. J. Pharmacol. 1993, 235, 283–289. [Google Scholar] [CrossRef]
  162. Fumagalli, E.; Funicello, M.; Rauen, T.; Gobbi, M.; Mennini, T. Riluzole enhances the activity of glutamate transporters GLAST, GLT1 and EAAC1. Eur. J. Pharmacol. 2008, 578, 171–176. [Google Scholar] [CrossRef]
  163. Kim, J.E.; Kim, D.S.; Kwak, S.E.; Choi, H.C.; Song, H.K.; Choi, S.Y.; Kwon, O.S.; Kim, Y.I.; Kang, T.C. Anti-glutamatergic effect of riluzole: Comparison with valproic acid. Neuroscience 2007, 147, 136–145. [Google Scholar] [CrossRef]
  164. Borowicz, K.K.; Sekowski, A.; Drelewska, E.; Czuczwar, S. Rilozole enhances the antiseizure action of conventional antiepileptic drugs against pentetrazole-induced convulsions in mice. Pol. J. Pharmacol. 2004, 56, 187–193. [Google Scholar] [PubMed]
  165. Du, J.; Vegh, V.; Reutens, D.C. Persistent sodium current blockers can suppress seizures caused by loss of low-threshold D-type potassium currents: Predictions from an in silico study of Kv 1 channel disorder. Epilepsia Open 2020, 5, 86–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Chen, H.S.; Pellegrini, J.W.; Aggarwal, S.K.; Lei, S.Z.; Warach, S.; Jensen, F.E.; Lipton, S.A. Open-channel block of N-methyl-Daspartate (NMDA) responses by memantine: Therapeutic advantage against NMDA receptor-mediated neurotoxicity. J. Neurosci. 1992, 12, 4427–4436. [Google Scholar] [CrossRef] [Green Version]
  167. Zaitsev, A.V.; Kim, K.; Vasilev, D.S.; Lukomskaya, N.Y.; Lavrentyeva, V.V.; Tumanova, N.L.; Zhuravin, I.A.; Magazanik, L.G. N-methyl-D-aspartate receptor channel blockers prevent pentylenetetrazole-induced convulsions and morphological changes in rat brain neurons. J. Neurosci. Res. 2015, 93, 454–465. [Google Scholar] [CrossRef]
  168. Sun, Y.; Dhamne, S.C.; Carretero-Guillén, A.; Salvador, R.; Goldenberg, M.C.; Godlewski, B.R.; Pascual-Leone, A.; Madsen, J.R.; Stone, S.S.D.; Ruffini, G.; et al. Drug-Responsive Inhomogeneous Cortical Modulation by Direct Current Stimulation. Ann. Neurol. 2020, 88, 489–502. [Google Scholar] [CrossRef]
  169. Wada, Y.; Hasegawa, H.; Nakamura, M.; Yamaguchi, N. The NMDA receptor antagonist MK-801 has a dissociative effect on seizure activity of hippocampal-kindled cats. Pharmacol. Biochem. Behav. 1992, 43, 1269–1272. [Google Scholar] [CrossRef]
  170. Newcomer, J.W.; Farber, N.B.; Jevtovic-Todorovic, V.; Selke, G.; Melson, A.K.; Hershey, T.; Craft, S.; Olney, J.W. Ketamine-Induced NMDA Receptor Hypofunction as a Model of Memory Impairment and Psychosis. Neuropsychopharmacology 1999, 20, 106–118. [Google Scholar] [CrossRef] [PubMed]
  171. Goda, Y.; Stevens, C.F. Synaptic plasticity: The basis of particular types of learning. Curr. Biol. 1996, 6, 375–378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Rogawski, M.A. AMPA receptors as a molecular target in epilepsy therapy. Acta Neurol. Scand. Suppl. 2013, 197, 9–18. [Google Scholar] [CrossRef] [Green Version]
  173. Hunt, D.L.; Castillo, P.E. Synaptic plasticity of NMDA receptors: Mechanisms and functional implications. Curr. Opin. Neurobiol. 2012, 22, 496–508. [Google Scholar] [CrossRef] [Green Version]
  174. Zarnowski, T.; Kleinrok, Z.; Turski, W.A.; Czuczwar, S.J. The competitive NMDA antagonist, D-CPP-ene, potentiates the anticonvulsant activity of conventional antiepileptics against maximal electroshock-induced mice. Neuropharmacology 1994, 33, 619–624. [Google Scholar] [CrossRef] [PubMed]
  175. Zhu, X.; Dong, J.; Shen, K.; Bai, Y.; Zhang, Y.; Lv, X.; Chao, J.; Yao, H. NMDA receptor NR2B subunits contribute to PTZ-kindling-induced hippocampal astrocytosis and oxidative stress. Brain Res. Bull. 2015, 114, 70–78. [Google Scholar] [CrossRef] [PubMed]
  176. Löscher, W.; Hönack, D. Anticonvulsant and behavioral effects of two novel competitive N-methyl-D-aspartic acid receptor antagonists, CGP 37849 and CGP 39551, in the kindling model of epilepsy. Comparison with MK-801 and carbamazepine. J. Pharmacol. Exp. Ther. 1991, 256, 432–440. [Google Scholar] [PubMed]
  177. Dziki, M.; Hönack, D.; Löscher, W. Kindled rats are more sensitive than non-kindled rats to the behavioural effects of combined treatment with MK-801 and valproate. Eur. J. Pharmacol. 1992, 222, 273–278. [Google Scholar] [CrossRef]
  178. Rogawski, M.A.; Kurzman, P.S.; Yamaguchi, S.I.; Li, H. Role of AMPA and GluR5 kainate receptors in the development and expression of amygdala kindling in the mouse. Neuropharmacology 2001, 40, 28–35. [Google Scholar] [CrossRef]
  179. Löscher, W.; Rundfeldt, C.; Hönack, D. Low doses of NMDA receptor antagonists synergistically increase the anticonvulsant effect of the AMPA receptor antagonist NBQX in the kindling model of epilepsy. Eur. J. Neurosci. 1993, 5, 1545–1550. [Google Scholar] [CrossRef]
  180. Graebenitz, S.; Kedo, O.; Speckmann, E.J.; Gorji, A.; Panneck, H.; Hans, V.; Palomero-Gallagher, N.; Schleicher, A.; Zilles, K.; Pape, H.C. Interictal-like network activity and receptor expression in the epileptic human lateral amygdala. Brain 2011, 134, 2929–2947. [Google Scholar] [CrossRef] [Green Version]
  181. Hanada, T.; Ido, K.; Kosasa, T. Effect of perampanel, a novel AMPA antagonist, on benzodiazepine-resistant status epilepticus in a lithium-pilocarpine rat model. Pharmacol. Res. Perspect. 2014, 2, e00063. [Google Scholar] [CrossRef] [PubMed]
  182. Figueiredo, T.H.; Qashu, F.; Apland, J.P.; Aroniadou-Anderjaska, V.; Souza, A.P.; Braga, M.F.M. The GluK1 (GluR5) Kainate/{alpha}-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor antagonist LY293558 reduces soman-induced seizures and neuropathology. J. Pharmacol. Exp. Ther. 2011, 336, 303–312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Dhir, A.; Chavda, V. Pre- and post-exposure talampanel (GYKI 53773) against kainic acid seizures in neonatal rats. Pharmacol. Rep. PR. 2016, 68, 190–195. [Google Scholar] [CrossRef] [PubMed]
  184. Fritsch, B.; Stott, J.J.; Joelle Donofrio, J.; Rogawski, M.A. Treatment of early and late kainic acid-induced status epilepticus with the noncompetitive AMPA receptor antagonist GYKI 52466. Epilepsia 2010, 51, 108–117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Mohammad, H.; Sekar, S.; Wei, Z.; Moien-Afshari, F.; Taghibiglou, C. Perampanel but Not Amantadine Prevents Behavioral Alterations and Epileptogenesis in Pilocarpine Rat Model of Status Epilepticus. Mol. Neurobiol. 2019, 56, 2508–2523. [Google Scholar] [CrossRef] [PubMed]
  186. Wu, T.; Ido, K.; Osada, Y.; Kotani, S.; Tamaoka, A.; Hanada, T. The neuroprotective effect of perampanel in lithium-pilocarpine rat seizure model. Epilepsy Res. 2017, 137, 152–158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Rheims, S.; Ryvlin, P. Profile of perampanel and its potential in the treatment of partial onset seizures. Neuropsychiatr. Dis. Treat. 2013, 9, 629–637. [Google Scholar] [CrossRef] [Green Version]
  188. Wu, T.; Ido, K.; Ohgoh, M.; Hanada, T. Mode of seizure inhibition by sodium channel blockers, an SV2A ligand, and an AMPA receptor antagonist in a rat amygdala kindling model. Epilepsy Res. 2019, 154, 42–49. [Google Scholar] [CrossRef]
  189. Lai, M.C.; Tzeng, R.C.; Huang, C.W.; Wu, S.N. The Novel Direct Modulatory Effects of Perampanel, an Antagonist of AMPA Receptors, on Voltage-Gated Sodium and M-type Potassium Currents. Biomolecules 2019, 9, 638. [Google Scholar] [CrossRef] [Green Version]
  190. Krauss, G.L.; Serratosa, J.M.; Villanueva, V.; Endziniene, M.; Hong, Z.; French, J.; Yang, H.; Squillacote, D.; Edwards, H.B.; Zhu, J.; et al. Randomized phase III study 306: Adjunctive perampanel for refractory partial-onset seizures. Neurology 2012, 78, 1408–1415. [Google Scholar] [CrossRef]
  191. Nishida, T.; Lee, S.K.; Inoue, Y.; Saeki, K.; Ishikawa, K.; Kaneko, S. Adjunctive perampanel in partial-onset seizures: Asia-Pacific, randomized phase III study. Acta Neurol. Scand. 2018, 137, 392–399. [Google Scholar] [CrossRef]
  192. French, J.A.; Krauss, G.L.; Wechsler, R.T.; Wang, X.F.; DiVentura, B.; Brandt, C.; Trinka, E.J.; O’Brien, T.; Laurenza, A.; Patten, A.; et al. Perampanel for tonic-clonic seizures in idiopathic generalized epilepsy A randomized trial. Neurology 2015, 85, 950–957. [Google Scholar] [CrossRef] [Green Version]
  193. Davies, J.; Evans, R.H.; Herrling, P.L.; Jones, A.W.; Olverman, H.J.; Pook, P.; Watkins, J.C. CPP, a new potent and selective NMDA antagonist. Depression of central neuron responses, affinity for [3H]D-AP5 binding sites on brain membranes and anticonvulsant activity. Brain Res. 1986, 382, 169–173. [Google Scholar] [CrossRef] [PubMed]
  194. Kristensen, J.D.; Hartvig, P.; Karlsten, R.; Gordh, T.; Halldin, M. CSF and plasma pharmacokinetics of the NMDA receptor antagonist CPP after intrathecal, extradural and i.v. administration in anaesthetized pigs. Br. J. Anaesth. 1995, 74, 193–200. [Google Scholar] [CrossRef]
  195. Rogawski, M.A. Therapeutic potential of excitatory amino acid antagonists: Channel blockers and 2,3-benzodiazepines. Trends Pharmacol. Sci. 1993, 14, 325–331. [Google Scholar] [CrossRef] [PubMed]
  196. Yen, W.; Williamson, J.; Bertram, E.H.; Kapur, J. A comparison of three NMDA receptor antagonists in the treatment of prolonged status epilepticus. Epilepsy Res. 2004, 59, 43–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Goodkin, H.P.; Yeh, J.L.; Kapur, J. Status epilepticus increases intracellular accumulation of GABAA receptors. J. Neurosci. 2005, 25, 5511–5520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Gaspard, N.; Foreman, B.; Judd, L.M.; Brenton, J.N.; Nathan, B.R.; McCoy, B.M.; Al-Otaibi, A.; Kilbride, R.; Fernández, I.S.; Mendoza, L.; et al. Intravenous ketamine for the treatment of refractory status epilepticus: A retrospective multicenter study. Epilepsia 2013, 54, 1498–1503. [Google Scholar] [CrossRef] [Green Version]
  199. Kramer, A.H. Early ketamine to treat refractory status epilepticus. Neurocrit. Care 2012, 16, 299–305. [Google Scholar] [CrossRef]
  200. Synowiec, A.S.; Singh, D.S.; Yenugadhati, V.; Valeriano, J.P.; Schramke, C.J.; Kelly, K.M. Ketamine use in the treatment of refractory status epilepticus. Epilepsy Res. 2013, 105, 183–188. [Google Scholar] [CrossRef]
  201. Sabharwal, V.; Ramsay, E.; Martinez, R.; Shumate, R.; Khan, F.; Dave, H.; Iwuchukwu, I.; McGrade, H. Propofol-ketamine combination therapy for effective control of super-refractory status epilepticus. Epilepsy Behav. 2015, 52, 264–266. [Google Scholar] [CrossRef] [PubMed]
  202. Rosati, A.; L’Erario, M.; Ilvento, L.; Cecchi, C.; Pisano, T.; Mirabile, L.; Guerrini, R. Efficacy and safety of ketamine in refractory status epilepticus in children. Neurology 2012, 79, 2355–2358. [Google Scholar] [CrossRef] [PubMed]
  203. Marrero-Rosado, B.M.; de Araujo Furtado, M.; Kundrick, E.R.; Walker, K.A.; Stone, M.F.; Schultz, C.R.; Nguyen, D.A.; Lumley, L.A. Ketamine as adjunct to midazolam treatment following soman-induced status epilepticus reduces seizure severity, epileptogenesis, and brain pathology in plasma carboxylesterase knockout mice. Epilepsy Behav. 2020, 111, 107229. [Google Scholar] [CrossRef]
  204. Santoro, J.D.; Filippakis, A.; Chitnis, T. Ketamine use in refractory status epilepticus associated with anti-NMDA receptor antibody encephalitis. Epilepsy Behav. Rep. 2019, 12, 100326. [Google Scholar] [CrossRef]
  205. Alkhachroum, A.; Der-Nigoghossian, C.A.; Mathews, E.; Massad, N.; Letchinger, R.; Doyle, K.; Chiu, W.T.; Kromm, J.; Rubinos, C.; Velazquez, A.; et al. Ketamine to treat super-refractory epilepticus. Neurology 2020, 95, e2286–e2294. [Google Scholar] [CrossRef] [PubMed]
  206. Liebe, J.; Li, S.; Lord, A.; Colic, L.; Kuause, A.L.; Batra, A.; Kretzschmar, M.A.; Sweeney-Reed, C.M.; Behnisch, G.; Schott, B.H.; et al. Factors influencing the cardiovascular response to subanesthetic ketamine: A randomized, placebo-controlled trial. Int. J. Neuropsychopharmacol. 2017, 20, 909–918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Perks, A.; Cheema, S.; Mohanraj, R. Anaesthesia and epilepsy. Br. J. Anaesth. 2012, 108, 562–571. [Google Scholar] [CrossRef] [Green Version]
  208. Hallak, M. Effect of parenteral magnesium sulfate administration on excitatory amino acid receptors in the rat brain. Magnes. Res. 1998, 11, 117–131. [Google Scholar]
  209. Visser, N.A.; Braun, K.P.J.; Leijten, F.S.S.; van Nieuwenhuizen, O.; Wokke, J.H.J.; van den Bergh, W.M. Magnesium treatment for patients with refractory status epilepticus due to POLG1-mutations. J. Neurol. 2011, 258, 218–222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Zou, L.P.; Wang, X.; Dong, C.H.; Chen, C.H.; Zhao, W.; Zhao, R.Y. Three-week combination treatment with ACTH + magnesium sulfate versus ACTH monotherapy for infantile spasms: A 24-week, randomized, open-label, follow-up study in China. Clin. Ther. 2010, 32, 692–700. [Google Scholar] [CrossRef] [PubMed]
  211. Euser, A.G.; Cipolla, M.J. Magnesium sulfate for the treatment of eclampsia: A brief review. Stroke 2009, 40, 1169–1175. [Google Scholar] [CrossRef] [PubMed]
  212. Vargas, J.R.; Takahashi, D.K.; Thomson, K.E.; Wilcox, K.S. The expression of kainate receptor subunits in hippocampal astrocytes after experimentally induced status epilepticus. J. Neuropathol. Exp. Neurol. 2013, 72, 919–932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Gibbons, M.B.; Smeal, R.M.; Takahashi, D.K.; Vargas, J.R.; Wilcox, K.S. Contributions of astrocytes to epileptogenesis following status epilepticus: Opportunities for preventive therapy? Neurochem. Int. 2013, 63, 660–669. [Google Scholar] [CrossRef] [Green Version]
  214. Faught, E. BGG492 (selurampanel), an AMPA/kainate receptor antagonist drug for epilepsy. Expert Opin. Investig. Drugs 2014, 23, 107–113. [Google Scholar] [CrossRef] [PubMed]
  215. Elger, C.E.; Hong, S.B.; Brandt, C.; Mancione, L.; Han, J.; Strohmaier, C. BGG492 as an adjunctive treatment in patient with partial-onset seizure: A 12-week, randomized, double-blind, placebo-controlled, phase II dose-titration study with an open-label extension. Epilepsia 2017, 7, 1217–1226. [Google Scholar] [CrossRef] [Green Version]
  216. Dinis-Oliveira, R.J. Metabolism and metabolomics of ketamine: A toxicological approach. Forensic Sci. Res. 2017, 2, 2–10. [Google Scholar] [CrossRef] [Green Version]
  217. Li, Y.; Jackson, K.A.; Slon, B.; Hardy, J.R.; Franco, M.; William, L.; Poon, P.; Coller, J.K.; Hutchinson, M.R.; Currow, D.C.; et al. CYP2B6*6 allele and age substantially reduce steady-state ketamine clearance in chronic pain patients: Impact on adverse effects. Br. J. Clin. Pharmacol. 2015, 80, 276–284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Mir, A.; Qahtani, M.; Bashir, S. GRIN2A-related severe epileptic encephalopathy treated with memantine: An example of precision medicine. J. Pediatr. Genet. 2020, 9, 252–257. [Google Scholar] [CrossRef]
  219. Pierson, T.M.; Yuan, H.; Marsh, E.D.; Fuentes-Fajardo, K.; Adams, D.R.; Markello, T.; Golas, G.; Simeonov, D.R.; Holloman, C.; Tankovic, A.; et al. GRAIN2A mutation and early- onset epileptic encephalopathy: Personalized therapy with memantine. Ann. Clin. Transl. Neurol. 2014, 1, 190–198. [Google Scholar] [CrossRef]
  220. Platzer, K.; Yuan, H.; Schütz, H.; Winschel, A.; Chen, W.; Hu, C.; Kusumoto, H.; Heyne, H.O.; Helbig, K.L.; Tang, S.; et al. GRAIN2B encephalopathy: Novel findings on phontype, variant clustering, functional consequences and treatment aspects. J. Med. Genet. 2017, 54, 460–470. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Mechanism and types of ionotropic and metabotropic glutamate receptors with associated proteins. (A) Ionotropic glutamate receptors are composed of four subunits and are structured similarly to a central ion channel pore, and are activated by NMDAR (GluN1–3 subunits), AMPAR (GluA1–4 subunits), and KAR (GluK1–5 subunits). These channels regulate Ca2+ permeability and play roles in receptor trafficking, synaptic plasticity, learning, memory, and even cell death. (B) Metabotropic glutamate receptors are G-protein-coupled receptors that modulate synaptic transmission and neuronal excitability. They are divided into three groups: Group I includes mGluR1 and 5; Group II includes mGluR2 and 3; Group III includes mGluR4, 6, 7, and 8. (C) Group I metabotropic receptors enhance the excitatory function in neurons. Presynaptic mGluR1 facilitates vesicular release of glutamate, and postsynaptic mGluR5 modulates the action of postsynaptic NMDARs via PSD-95 and Homer protein, leading to phosphorylation of NMDARs and potentiation of NMDAR currents. (D) Group II and Group III metabotropic receptors mainly inhibit excitatory function. mGluR2 and 3 and mGluR4, 6, 7, and 8 inhibit presynaptic glutamate release. mGluR3 is also expressed on astrocytes, positively modulating the glutamate transporter proteins GLAST and GLT-1, enhancing synaptic glutamate reuptake, and thus countering the hyperexcitability of neurons. (E) Glutamate in the synaptic cleft is taken up by astrocytes around the synapses, and is transformed to glutamine by glutamine synthetase; then, glutamine is transformed to glutamate and stored in vesicles of the presynaptic terminals to maintain neuronal communication (Figure created with BioRender.com, accessed on 23 February 2023).
Figure 1. Mechanism and types of ionotropic and metabotropic glutamate receptors with associated proteins. (A) Ionotropic glutamate receptors are composed of four subunits and are structured similarly to a central ion channel pore, and are activated by NMDAR (GluN1–3 subunits), AMPAR (GluA1–4 subunits), and KAR (GluK1–5 subunits). These channels regulate Ca2+ permeability and play roles in receptor trafficking, synaptic plasticity, learning, memory, and even cell death. (B) Metabotropic glutamate receptors are G-protein-coupled receptors that modulate synaptic transmission and neuronal excitability. They are divided into three groups: Group I includes mGluR1 and 5; Group II includes mGluR2 and 3; Group III includes mGluR4, 6, 7, and 8. (C) Group I metabotropic receptors enhance the excitatory function in neurons. Presynaptic mGluR1 facilitates vesicular release of glutamate, and postsynaptic mGluR5 modulates the action of postsynaptic NMDARs via PSD-95 and Homer protein, leading to phosphorylation of NMDARs and potentiation of NMDAR currents. (D) Group II and Group III metabotropic receptors mainly inhibit excitatory function. mGluR2 and 3 and mGluR4, 6, 7, and 8 inhibit presynaptic glutamate release. mGluR3 is also expressed on astrocytes, positively modulating the glutamate transporter proteins GLAST and GLT-1, enhancing synaptic glutamate reuptake, and thus countering the hyperexcitability of neurons. (E) Glutamate in the synaptic cleft is taken up by astrocytes around the synapses, and is transformed to glutamine by glutamine synthetase; then, glutamine is transformed to glutamate and stored in vesicles of the presynaptic terminals to maintain neuronal communication (Figure created with BioRender.com, accessed on 23 February 2023).
Biomedicines 11 00783 g001
Figure 2. Structures and possible mutations of ionotropic glutamate receptors with associated clinical features. (A) Ionotropic glutamate receptors such as NMDA and AMPA receptors share a common tetramer containing A–D subunits in different combinations of subunits. In NMDA receptors, the GluN1 subunit is essential for all functional NMDARs and has binding sites for glycine and D-serine. The GluN2 subunits provide glutamate-binding sites and control the expression and functional properties of NMDARs. In AMPA receptors, the GluA2 subunit is the key site for regulating calcium permeability. (B) Mutations in the subunits have been found to cause clinical syndromes. GRIN1 missense mutations can cause infantile-onset epilepsy with encephalopathies. GluN2A mutations are related to benign Rolandic epilepsy. GRIN2B gain-of-function mutation causes West syndrome. Some GRIN2D variants were reported with initial early-onset seizures. (C) The heterozygous mutations of GRIA2 gene encoding GluA2 subunit were found to have phenotypes of intellectual disability, neurodevelopmental abnormalities, and seizure (Figure created with BioRender.com, accessed on 23 February 2023).
Figure 2. Structures and possible mutations of ionotropic glutamate receptors with associated clinical features. (A) Ionotropic glutamate receptors such as NMDA and AMPA receptors share a common tetramer containing A–D subunits in different combinations of subunits. In NMDA receptors, the GluN1 subunit is essential for all functional NMDARs and has binding sites for glycine and D-serine. The GluN2 subunits provide glutamate-binding sites and control the expression and functional properties of NMDARs. In AMPA receptors, the GluA2 subunit is the key site for regulating calcium permeability. (B) Mutations in the subunits have been found to cause clinical syndromes. GRIN1 missense mutations can cause infantile-onset epilepsy with encephalopathies. GluN2A mutations are related to benign Rolandic epilepsy. GRIN2B gain-of-function mutation causes West syndrome. Some GRIN2D variants were reported with initial early-onset seizures. (C) The heterozygous mutations of GRIA2 gene encoding GluA2 subunit were found to have phenotypes of intellectual disability, neurodevelopmental abnormalities, and seizure (Figure created with BioRender.com, accessed on 23 February 2023).
Biomedicines 11 00783 g002
Table 1. Treatment of SE by various glutamate receptor (NMDAR) antagonists.
Table 1. Treatment of SE by various glutamate receptor (NMDAR) antagonists.
MedicationMechanismEfficacyRemarks
KetamineNoncompetitive NMDA antagonism-Treatment of SRSE in adults and children, with control rate ranging from 50% to as high as 90%-Does not alter intracranial pressure or cerebral blood flow
-Lack of adverse cardiopulmonary depression effect
Magnesium sulfateNMDA antagonism-Associated with an increased seizure threshold and reduced seizures in animal studies; effective add-on therapy in infantile spasm-Intravenous magnesium sulphate is the standard management in eclamptic seizures.
Dizocilpine (MK-801)Noncompetitive NMDA antagonism-In SE animal models, MK-801 is superior to the competitive antagonist CPP and noncompetitive antagonist ifenprodil
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, T.-S.; Huang, T.-H.; Lai, M.-C.; Huang, C.-W. The Role of Glutamate Receptors in Epilepsy. Biomedicines 2023, 11, 783. https://doi.org/10.3390/biomedicines11030783

AMA Style

Chen T-S, Huang T-H, Lai M-C, Huang C-W. The Role of Glutamate Receptors in Epilepsy. Biomedicines. 2023; 11(3):783. https://doi.org/10.3390/biomedicines11030783

Chicago/Turabian Style

Chen, Tsang-Shan, Tzu-Hsin Huang, Ming-Chi Lai, and Chin-Wei Huang. 2023. "The Role of Glutamate Receptors in Epilepsy" Biomedicines 11, no. 3: 783. https://doi.org/10.3390/biomedicines11030783

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