Next Article in Journal
An Algae-Made RBD from SARS-CoV-2 Is Immunogenic in Mice
Next Article in Special Issue
The Role of Zinc and NMDA Receptors in Autism Spectrum Disorders
Previous Article in Journal
Paeonol Protects against Methotrexate Hepatotoxicity by Repressing Oxidative Stress, Inflammation, and Apoptosis—The Role of Drug Efflux Transporters
Previous Article in Special Issue
Pharmacological Comparative Characterization of REL-1017 (Esmethadone-HCl) and Other NMDAR Channel Blockers in Human Heterodimeric N-Methyl-D-Aspartate Receptors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting NMDA Receptor Complex in Management of Epilepsy

by
Shravan Sivakumar
1,
Mehdi Ghasemi
1,* and
Steven C. Schachter
2,3,4,*
1
Department of Neurology, University of Massachusetts Chan Medical School, Worcester, MA 01655, USA
2
Department of Neurology, Beth Israel Deaconess Medical Center, Boston, MA 02215, USA
3
Harvard Medical School, Boston, MA 02114, USA
4
Consortia for Improving Medicine with Innovation & Technology (CIMIT), Boston, MA 02114, USA
*
Authors to whom correspondence should be addressed.
Pharmaceuticals 2022, 15(10), 1297; https://doi.org/10.3390/ph15101297
Submission received: 6 August 2022 / Revised: 17 October 2022 / Accepted: 20 October 2022 / Published: 21 October 2022
(This article belongs to the Special Issue NMDA Receptor-Based Therapeutics)

Abstract

:
N-methyl-D-aspartate receptors (NMDARs) are widely distributed in the central nervous system (CNS) and play critical roles in neuronal excitability in the CNS. Both clinical and preclinical studies have revealed that the abnormal expression or function of these receptors can underlie the pathophysiology of seizure disorders and epilepsy. Accordingly, NMDAR modulators have been shown to exert anticonvulsive effects in various preclinical models of seizures, as well as in patients with epilepsy. In this review, we provide an update on the pathologic role of NMDARs in epilepsy and an overview of the NMDAR antagonists that have been evaluated as anticonvulsive agents in clinical studies, as well as in preclinical seizure models.

1. Introduction

Epilepsy is a common neurological disorder, affecting about 1% of the general population, approximately 50 million people worldwide [1]. Despite the widespread use of anti-seizure medications (ASMs) over the past few decades in the management of epilepsy, about one-third of patients with epilepsy show no response to anti-seizure medication [2,3]. The International League Against Epilepsy (ILAE) defines drug-resistant epilepsy (DRE) as the “failure of adequate trials of 2 tolerated, appropriately chosen and used AED schedules (whether as monotherapies or in combination) to achieve sustained seizure freedom” [4]. The incidence of DRE was recently found to be about 19.6% of total epilepsy cases [5]. This calls for the need for further understanding of the mechanisms implicated in epileptogenesis, which could bolster treatment options.
Several pathways have been implicated to play a role in epileptogenesis [6], among which the desynchrony between neuronal excitation and inhibition is widely speculated to majorly contribute. Previous research on seizures from temporal lobe epilepsy (TLE) has demonstrated that glutamate levels rise in the extracellular fluid and that glutamate can directly activate N-methyl-D-aspartate receptors (NMDARs) and cause neuroexcitatory toxicity [7]. An overwhelming body of evidence exists centered around the role of NMDARs in several neurological disorders, including epilepsy. More recently, a sub-type of autoimmune encephalitis has been found to be associated with ~20% of epilepsy cases. This has further highlighted the role of the NMDAR complex in epileptogenesis [8]. These factors have formed the basis of research directed at both the preventive and therapeutic roles of NMDAR-modulating therapy in the management of neurological disorders.
The NMDAR complex is characterized by excitatory neurotransmitters (glutamate receptors) located on the synapses of interneurons regulating the balance between neuronal excitation and inhibition, and it has been implicated to play a role in epilepsy [9,10]. Several animal models have consistently shown that blocking NMDARs is effective in both the prevention and reversal of neurological disorders, including epilepsy [11]. A high sensitivity to modulation, as well as a proclivity for negative side effects, including neurotoxicity, has made it challenging for the development of newer agents. In this review, we highlight the role of NMDAR-guided therapy by providing an overview of the key pathophysiology, linking the role of the NMDAR complex in epilepsy with an emphasis on drugs currently in use, as well as on-going preclinical studies utilizing NMDAR-modulating therapy.

2. NMDAR Complex

The NMDAR is one of the ionotropic glutamate receptors (iGluR) serving as a target for action of the major excitatory neurotransmitter glutamate at the presynaptic terminal and post-synaptic membrane in the central nervous system (CNS) [12].
The NMDAR complex contributes to normal brain functioning, which begins in-utero. By providing neuronal excitation that promotes survival and efficient connectivity, NMDARs are involved in neurodevelopment [13]. This developmental period extends from the third trimester of pregnancy to the first several years of postnatal life in humans. NMDAR transmission is involved in the connectivity between hippocampal and prefrontal circuits [14]. NMDAR activation in hippocampal pathways controls an activity-dependent synaptic modification called long-term potentiation (LTP), contributing to learning and new memory formation [15,16].
In addition, the NMDAR complex is involved in spatial learning [17]. Furthermore, NMDAR transmission is involved in persistent neuronal firing, which has been implicated in underlying neurocognitive disorders and particularly in aversive mental states [18].
The receptor complex consists of a heterotetrametric structure in which an NR1 (GluN1) subunit is ubiquitous, and there are varying combinations of NR2 (GluN2) or NR3 (GluN3) subunits, with multiple binding sites for glutamate, polyamine, Mg2+, and glycine (Figure 1). The varying combinations of subunit binding sites determine the pharmacological regulation of the NMDAR. The glutamate binding site is situated on the NR2 subunit, and, similarly, the glycine binding site is situated on the NR1 subunit. These binding sites exhibit varying neuroanatomic expressions.
In the resting state, the channel is blocked by Mg2+ and remains equally permeable to Na+ and Ca2+ ions [19,20]. Membrane depolarization relieves the Mg2+ blockage, and the resulting neuronal excitation in-term mediates the NMDAR responses contributing to the neurotoxicity from excess Ca2+ ions. Neurotoxicity resulting from overexcitation has been widely speculated to play a major role in epilepsy [21].

2.1. NMDA Trafficking

NMDAR delivery to synapses and intracellular trafficking both depend on PDZ proteins [22]. NMDARs are not evenly distributed once they reach the neural surface, showing a higher concentration in postsynaptic densities and a lower one in extra-synaptic compartments [23]. Surface NMDARs are dynamically anchored in the postsynaptic density (PSD) region via an interaction between GluN2 subunits and proteins with PDZ-binding domains [24]. However, there are still questions about where receptor membrane trafficking occurs. The dysregulation of NMDAR trafficking has been implicated in several neuropsychiatric disorders in the past, and increasing evidence points to their involvement in epilepsy [25].
The trafficking of NMDARs to membranes was noted through an increase in synaptic and/or presynaptic NR1 subunits in a rat model of status epilepticus (SE) [26]. This increase in NMDA expression coincides with the loss of synaptic inhibitions through the internalization of (GABA)A receptors implicated in the propagation of seizures to SE [27]. Furthermore, GRIN2A mutations can impact NMDAR trafficking overall by altering the levels of GLuN2A proteins and by altering GLuN2A membrane trafficking [28]. The defective interaction of protein binding sites involved in vesicular trafficking (SNX27) due to the phosphorylation of GluN2A has also been implicated in NMDAR trafficking defects [29]. In animal models and human epileptic brain tissue, G-protein-coupled receptor 40 (GPR40) affected N-methyl-D-aspartate (NMDA) receptor-mediated synaptic transmission through the regulation of NR2A and NR2B expressions on the surface of neurons [30]. Furthermore, the endocytosis of NMDARs and the binding of GPR40 with NR2A and NR2B were regulated through GPR40 [30]. Alterations in the interactions involved in NMDAR trafficking could open new avenues of therapeutic targets to alleviate neuronal overexcitation in epilepsy.

2.2. NMDA Modulation (Glycine and Other Sites)

The binding of a co-agonist at the glycine-mediated site (GMS) is necessary for NMDAR action, in addition to glutamate. The modulation of the GMS of the NMDAR is low, given the low saturation in vivo despite the high CSF concentrations of glycine [31]. D-serine serves as a major endogenous co-agonist of the NMDAR, and D-serine levels were recently found to be upregulated in intractable epilepsy [32,33,34]. It is possible that endogenous glycine does not fully stimulate NMDARs as suggested by the selective potentiation of the convulsant activity of NMDA by D-serine [35]. However, the therapeutic effect of glycine modulation is severely limited given the requirements of a high dose and a narrow therapeutic window and its severe adverse effects, such as oxidative damage, neurotoxicity, and nephrotoxicity [36].
When compared to other NMDAR subtypes, NR2B-containing receptors appear to contribute more favorably to pathogenic processes, such as epilepsy caused by excessive glutamatergic pathway activation. This makes them more of a preferential target for modulation [37]. A common mechanism involved in the allosteric modulation of NMDARs is through proton selectivity by shifting the pKa of the proton sensor [38,39]. This mechanism is involved in selective allosteric inhibition via ifenprodil, polyamines, and extracellular zinc at NR2A-containing receptors [40,41]. The allosteric modulation of NMDARs through a novel synthetic analogue of 24(S)-hydroxycholesterol-SGE-301 prevented the NMDAR dysfunction in patients with autoimmune encephalitis from NMDAR antibodies in cultured neurons [42]. A potential mechanism implicated was the prolonged decay time of NMDAR-dependent spontaneous excitatory postsynaptic currents suggesting a prolonged open time of the channel. More recently, miR-219, a microRNA, was implicated to play a regulatory role in the modulation of excitatory neurotransmission in epilepsy [43]. The upregulation of NR1 subunits was noted through an inverse relationship between miR-219 and NMDA-NR1 expression in the amygdala and hippocampus of patients with intractable mesial temporal lobe epilepsy [43].

2.3. NMDA mGluR and AMPA Interactions

Metabotropic glutamate receptors (mGluRs), which are a subtype of glutamate receptors, are members of G-protein-coupled receptors (GPCRs) involved in intracellular secondary messenger systems modulating neuronal excitability, which is of relevance in epilepsy [44]. mGluR5 responses have been found to be regulated by the activation of NMDARs via a protein kinase C (PKC) pathway [45,46]. Prior work has shown that mGluR5-positive modulators can attenuate the behavioral effects of NMDAR antagonists, PCP and MK-801 [47,48]. Despite the fact that there have not yet been any large clinical trials focusing on mGluR5 in epilepsy, selective group I mGluR antagonists were explored for their anticonvulsant effects in rodent models of epilepsy by Chapman et al., 2000, and Yan et al., 2005 [49,50]. However, the limitations behind their potential usefulness as anticonvulsant drugs would be due to the dominant effects of mGluR1 in cerebellar function and motor control [51]. Accordingly, patients who express autoantibodies against mGluR1 [52] or Homer-3 (a scaffolding protein for mGluRs) [53] exhibit signs of cerebellar dysfunction, such as ataxia.
The increased phosphorylation of NMDA and AMPA receptor subunits in rat models has been implicated in the regulation of synaptic plasticity and memory consolidation via the activation of ERK1/2 signaling [54]. In humans, Anti-GluA1 and Anti-GluA2 antibodies that target AMPAR subunits have been found in patients with epilepsy caused by autoimmune limbic encephalitis [55,56]. However, AMPAR autoantibodies were found to not have any interaction with NMDARs [57]. More recently, a combination of NMDAR and AMPAR antagonists in a mice model demonstrated that these receptor interactions could potentially contribute to delayed epileptogenesis through granule cell dispersion [58]. Though the response was a delay rather than the prevention of epileptogenesis, further clinical trials are warranted to study this interaction.

3. NMDAR Alterations and Their Role in Human Epilepsy

The alterations in NMDARs in epilepsy have been extensively investigated in the past through the use of a variety of techniques, such as gene expression, immunoblotting, and binding affinity techniques.
Through an in situ hybridization technique, Bayer et al., in 1995, showed that a loss of NR1-positive cells was associated with overall neuronal loss involving pyramidal cells [59]. Furthermore, NR2 subunit mRNA levels were increased in patients with hippocampal sclerosis (HS) [60]. In the dentate gyrus, there appears to be an increase in NR2 immunoreactivity that is associated with abnormal mossy fiber sprouting in this region [61]. It has been consistently demonstrated that the inhibition of the glutamate binding site (NR2 subunit) decreases granule cell hyperexcitability in cases showing mossy fiber sprouting in hippocampi [62,63,64]. More direct evidence from pyramidal neurons in human cortical slice preparations from patients with mesial temporal lobe epilepsy showed that an increased endogenous activity of NMDARs was associated with neuronal hyperexcitability [65]. More recently, in focal epilepsies, through the use of [(18)F]GE-179, a ligand that selectively binds to open NMDAR ion channels, McGinnity CJ et al. demonstrated NMDA channel overactivity through the use of positron emission tomography (PET) [66]. An alteration in NMDAR subunit (NR2B) composition in cortical dysplasia tissue has been shown to contribute to functional abnormalities due to decreased Mg2+ sensitivity of the receptor, which results in neuronal hyperexcitability [67]. In addition, increases in the levels of NR2B and 2D subunit mRNAs and functional NR2B-containing receptors (using a ligand-binding method) were noted in tuberous sclerosis [68]. Dysplastic neurons showed increased expressions of NR2B and 2C subunit mRNAs, whereas only NR2D mRNA was upregulated in giant cells, suggesting that dysplastic neurons and giant cells contribute differently to epileptogenesis in the tuberous sclerosis complex [68]. These studies emphasize that various alterations in the different subunits of NMDARs (especially NR1 and NR2 subunits) in different brain regions could be responsible for seizure development accordingly among the several types of epilepsies.
In human patients with symptomatic epilepsies, the regulation of the GluN2B subunit of the NMDAR complex via NRG1-ErbB4-Src signaling pathways was identified as a potential modulating target through the use of the immunoblotting technique [69]. In patients with intractable temporal lobe epilepsies, D-serine and NMDAR1 expressions were significantly increased [34]. These observations highlight the importance of neurochemical targeting, which can be further explored in the future to guide anti-NMDAR complex therapies.
Moreover, a variety of animals models of seizures and epilepsy have demonstrated alterations in NMDAR expressions and protein levels, although the results vary depending on the NMDAR subunits, brain regions, and animal species assessed. This has been comprehensively reviewed in the literature [33,69,70,71].

3.1. Genetic Mutations of the NDMA Receptor

Various genetic expressions of the subunits of the NMDAR (GluN1, GluN2A-2D, and GluN3A-3B) can contribute to the development of distinct clinical phenotypes [72,73]. The genetic mutations in patients with epilepsy are classified broadly as loss-of-function, no-change, and gain-of-function mutations [73].
The GluN1 subunit, encoded by GRIN1, is typically involved in loss-of-function mutations contributing to structural changes resulting in a wide range of epilepsies of variable semiology (spasms, tonic and atonic seizures, hypermotor seizures, focal dyscognitive seizures, febrile seizures, generalized seizures, status epilepticus, myoclonic seizures, etc.) [74]. Reportedly, up to half of GRIN1 mutations are loss-of-function mutations, with the rest being gain-of-function mutations. Hence, the co-existence of both hypo-functioning and hyper-functioning NMDARs within the same disease phenotype invariably contributes to electrophysiological imbalance [10,12]. However, this relation of NMDAR function to the pathogenesis of epilepsy still needs further clarification. GRIN2A mutations, which can alter the GluN2A receptor, occur more so than any other of the NMDAR subunit mutations [75]. About 70% of GRIN2A variants are likely to lead to the development of epilepsy, whereas 30% of individuals with GRIN2B variants have epilepsy [76,77]. GluN2 subunit mutations may control epileptiform abnormalities arising from the hippocampus [78]. GRIN2D mutations are associated with treatment refractory epileptic encephalopathy [79].
Various alterations in subunits in different brain regions are, thus, held accountable for seizure development in different types of epilepsies. Identifying receptor mutations has, hence, been implicated to contribute to a personalized medicine approach in epilepsy treatment. Modeling NMDAR dysfunction with neurons derived from human induced pluripotent stem cells (iPSCs), as well as identifying signaling pathways, has been suggested to further contribute to the development of drugs targeting the NMDAR complex of gene regulatory variants [80].

3.2. Anti-NMDAR Encephalitis

Autoimmune encephalitis from anti-NMDAR antibodies has gained significant attention over the past decade [81,82] to the extent of being regarded as its own entity of epilepsy [83,84]. Anti-NMDA-NR2A/B antibodies are present in ~20% of patients with epilepsy [8]. The autoantibodies against NMDARs can cause a reversible loss of NMDAR function on the surface of neurons [85]. Several mechanisms are involved:
(a)
Internalization of NMDAR,
(b)
Disruption of interaction of NMDAR with EphB2R,
(c)
Decreased long-term potential (LTP), which can lower the threshold for seizures [86,87,88,89].
Reductions in excitatory neurotransmission caused by NMDAR antibodies in an vitro rat model may play a role in NMDA encephalitis [90]. These mechanisms invariably result in a state of excitatory and inhibitory chemical imbalance contributing to seizures in a significant number of patients. These are typically improved by the prompt use of immunomodulatory therapy [91]. A marked decrease in anti-NMDA-NR1 antibody titers was seen following the administration of steroids, intravenous immunoglobulin G (IVIG), or plasmapheresis/plasma exchange, which are part of the first-line management strategies, and following the administration of steroid-sparing agents: rituximab, cyclophosphamide, or both [85,92,93].

4. NMDAR Modulators Currently in Use

4.1. Ketamine

Ketamine acts as a noncompetitive NMDA antagonist and has low affinity for the NMDAR at the phencyclidine site within the ionotropic channel [94]. It was first synthesized in 1962. Conventionally used as an anesthetic agent due to its rapid onset of action and short half-life, ketamine induces a state of “dissociative anesthesia” resulting from an overall CNS depressant effect rather than an inhibitory effect [66,95]. The anticonvulsant effects of ketamine can augment the seizure protective effect of benzodiazepine loading [96,97]. This therapeutic benefit may result from the upregulation of NMDARs during prolonged seizures at the same time when there is an overall decrease in sensitivity to GABA agonists [98,99,100].
Animal models have consistently shown the synergistic anticonvulsant effects of ketamine therapy in combination with other antiepileptic medications demonstrating both dose- and time-dependent adjuvant effects [101,102]. A multicenter retrospective study provided preliminary data on the safety and efficacy of intravenous ketamine use in the treatment of refractory status epilepticus, which is defined as seizure activity that does not respond to two antiepileptic drugs at appropriate doses [103]. High-dose and the early initiation of ketamine infusions (2.2 mg/kg/h) were associated with a decrease in seizure burden in patients with super-refractory status epilepticus (SRSE) [104]. The favorable effect on the hemodynamic profile, which is due to ketamine’s sympathomimetic properties, unique benefits of conscious sedation, and overall efficacy in treating refractory seizures, makes ketamine an agent to consider using in the management of patients with severe acute traumatic brain injury [105,106,107]. In critically ill patients where polypharmacy is a concern, ketamine use can limit the need for the further escalation of sedation and thereby help to avoid the need for intubation [108,109]. Further prospective randomized control trials are required to establish a consensus statement on dosing ketamine in refractory epilepsy management in adult and pediatric populations.
The adverse effects of ketamine infusions include neuropsychiatric symptoms, such as hallucinations and delirium, and arrhythmias. There has been a rare report of new-onset seizures in a pediatric patient [100]. Ketamine has been found to induce dose-dependent neuronal injury in animal models through apoptosis, particularly in the frontal cortex and hippocampal regions [110,111]. This, at the same time, is offset by the neuroprotective effect conferred by ketamine through an increase in regional cerebral blood flow volumes, thereby limiting the damage from an incomplete cerebral ischemia [112,113]. More recently, the therapeutic effect of ketamine has been explored in targeted therapy on spreading depolarizations, which represent a unique pathophysiology contributing to secondary injury progression in severe acute brain injury [114,115,116]. Furthermore, an interplay between seizures and spreading depolarizations has been suggested [116,117]. This further emphasizes the fact that NMDARs are involved in this epiphenomenon of secondary brain injury progression. The prevention of secondary brain injury spread has potential to improve outcomes in critically ill patients with sub-arachnoid hemorrhage, traumatic brain injury, and malignant ischemic strokes.

4.2. Memantine

Memantine is a noncompetitive, open-channel NMDAR antagonist that blocks NMDAR ion channels by binding to or near Mg2+ binding sites preferentially when the receptor channel is open, thereby inhibiting the prolonged influx of Ca2+ ions with near-normal physiological NMDA activity [118]. The efficacy of memantine as an anticonvulsant for both monotherapy and in combination with other AEDs was explored in preclinical work in the early 1980s [119,120,121].
In a lithium–pilocarpine mice model, memantine prevented cognitive impairments post-status epilepticus [122]. In a pentylenetetrazole (PTZ) model of seizures, memantine prevented convulsions and the development of morphological changes [123]. Memantine had a significant neuroprotective effect on hippocampal and cortical neurons in culture against glutamate and NMDA excitotoxicity [124]. A mice model showed evidence of a decrease in the frequency of induced seizures with the addition of memantine [125]. However, the addition of memantine to the AED regimen in GRIN2B-mutation-related encephalopathy did not result in any significant decrease in the frequency of seizures in a group of six patients [126].
Given the paucity of clinical data, more clinical trials testing the anticonvulsant effects of memantine as an add-on therapy on seizure types in patients with epilepsy are needed.

4.3. Amantadine

Amantadine was originally used in the management of influenza. It acts by increasing the release and inhibiting the reuptake of dopamine in the brain. In addition to the several other pharmacological actions of amantadine, its role in NMDAR blockage by increasing the rate of channel closure was demonstrated in the early 1990s [127].
The safety and efficacy of amantadine as an add-on therapy in pediatric refractory generalized seizures was first evaluated in a small case series of four patients by Shahar et al. in 1993 [128,129]. In a larger retrospective review of 13 patients, a target amantadine dose of 4 to 7 mg/kg/day was utilized in conjunction with other AEDs. A total of 58% of patients had at least a 50% seizure reduction, with the majority (nearly 86%) of responders sustaining a seizure reduction greater than 90% at 12 months following treatment initiation [130]. Of note, no renal, hepatic, or hematologic toxicity was noted in this study, and other adverse effects included vomiting, behavioral changes, headache, dizziness, and weight loss.
Its favorable pharmacological profile, including the potential improved effects on cognitive recovery [131,132,133], make amantadine a promising agent of choice to explore in critically ill patients with refractory epilepsy. Further larger, multicenter trials are needed to study the efficacy of amantadine add-on therapy in refractory epilepsy.

4.4. Magnesium Sulphate

The potential for convulsions to occur in states of Mg2+ depletion prompted some researchers to investigate the metabolism of this metal in epilepsies [134]. Experiments have shown that Mg2+ blocks Ca2+ within the NMDAR channel, which, in turn, is relieved by cellular depolarization [135]. The anticonvulsant effects of the systemic administration of Mg2+ have been studied in animal models [136]. An increase in brain Mg2+ concentrations in rat brains has been found to be associated with an increased seizure threshold and resistance to NMDA-stimulated hippocampal seizures [137]. The inhibition of NMDARs is central to Mg2+ anticonvulsant effects [138].
In two patients with juvenile-onset Alpers’ syndrome, intravenous Mg2+ treatment reduced refractory epilepsy with recurring status epilepticus and bouts of epilepsia partialis continua [139]. Furthermore, in infants with infantile spasms, the addition of magnesium sulphate to adrenocorticotropic hormone (ACTH) showed significantly improved durations of seizure-free periods in an open-label, randomized, controlled study [140]. In humas, systemic administrations of magnesium sulphate are the standard of care in the management of eclampsia in pregnancy [141]. Concerns remain over the possibilities of hypermagnesemia toxicity in eclampsia treatment [142].

4.5. Felbamate

Felbamate is a propanediol dicarbamate derivative that was first approved as an ASM by the U.S. Food and Drug Administration (FDA) in 1993 [143]. Several mechanisms, including the blockage of voltage-gated sodium channels and dual actions on both excitatory (NMDA) and inhibitory (GABA) mechanisms, have been postulated to contribute to the anticonvulsant effect of felbamate [144].
In addition, the specific selectivity of felbamate to the NR2B subunit of NMDA has been noted. NR2B mutations have been implicated to play a role in epileptogenesis in mouse models [145]. Felbamate’s efficacy as add-on therapy is well-established for intractable partial seizures, infantile spasms, and Lennox-Gastaut syndrome [146,147]. Serious adverse effects of aplastic anemia and liver failure have been reported, which has limited the use of felbamate as add-on therapy in partial epilepsy [148].

4.6. Remacemide

Remacemide is a noncompetitive, low-affinity, NMDAR antagonist. A rapid and reversible inhibition of the NMDA current was first observed in rat forebrain membranes [149]. Its anticonvulsant effects in animal models of epilepsy were observed in a dose range of 6–60 mg/kg [150]. Remacemide hydrochloride was shown to have therapeutic activity in patients with medically refractive epilepsy at a dose of 600 mg/day [151]. Chadwick et al. [152] and Jones et al. [153] evaluated the role of the remacemide QID regimen as an add-on in medically refractory epilepsy.
It is worth noting that in patients with newly diagnosed epilepsy, when compared to other AEDs, such as carbamazepine, remacemide had no benefit as monotherapy [154]. A double-blind, parallel-group trial comparing remacemide with carbamazepine in partial or generalized tonic–clonic seizures showed a better cognitive and psychomotor profile at the cost of inferior seizure recurrence with the use of remacemide [155]. Some common adverse effects observed are dizziness, somnolence, and gastrointestinal symptoms. Diplopia and fatigue have been observed when used as adjunctive therapy with conventional AEDs [156].

4.7. Riluzole

2-Amino-6trifluoromethoxy benzothiazole (riluzole) was first noted to have anticonvulsant effects in the 1980s [157]. Several mechanisms involving glutamate modulation were noted. The presynaptic release of glutamate was reduced by the inhibition of voltage-gated Na+ currents in hippocampal neurons [158]. Riluzole can prevent Ca2+ entry via the NMDA channel, thereby blocking NMDAR activation [159]. A modulatory effect of riluzole on glutamate clearance has also been noted on the glutamate transporters expressed on neurons and glia [160].
Preclinical data have shown anticonvulsant effects in a variety of seizure models; in the rat dentate gyrus of both pilocarpine- and GBL-induced seizure models, riluzole was found to be more effective in reducing seizure activity than VPA [161]. Riluzole reduced seizure duration in a rat electroconvulsive shock model of epilepsy [162]. Furthermore, the inhibition of spontaneous glutamine transport by riluzole was demonstrated in hippocampal neurons [163]. In a CA3 in vitro rat slice model, the sodium channel blockage of riluzole resulted in decreased hippocampal epileptiform activity [164]. Despite the evidence from animal models, the effects of riluzole in human patients with epilepsy are yet to be evaluated in any clinical study.

4.8. Dizocilpine or MK-801

MK-801 is a noncompetitive NMDA antagonist that was first demonstrated to have antiepileptic effects in a model of induced seizures through the low-frequency kindling technique by Minabe et al. in 1992 [165]. MK-801 is a special NMDAR antagonist due to its well-demonstrated effects in both use-dependent and voltage-dependent manners via the blockade of ion permeation [166,167,168,169]. A large number of preclinical investigations have demonstrated its anticonvulsant properties in several models, including seizures induced by NMDA, quinolinic acid, lindane, 4-aminopyridine, caffeine, picrotoxin, bicuculline, cocaine, kainic acid, strychnine, pentylenetetrazole (PTZ), or hyperbaric oxygen in rodents [170,171,172,173,174,175,176,177,178,179,180,181]. However, more recent evidence based on the stargazer mice trial suggests that MK-801 has a paradoxical pro-seizure effect [182].
In tetramethylenedisulfotetramine (TMDT)-induced tonic–clonic seizures, the combination of diazepam and MK-801 had a synergistic anticonvulsant effect [183]. Status epilepticus was aborted, and mortality was eliminated with the combination of diazepam and dizocilpine in a rat model of SE induced by very high doses of lithium and pilocarpine [178], as well as in a model of soman-induced SE [184,185]. These observations suggest the possibility of clinical benefits of combinations of MK-801 and other anti-seizure medications. Due to concerns regarding the schizophrenia-like behaviors frequently noted in animal models, MK-801 has not been explored in human subjects [186].

4.9. Dextromethorphan

Initially introduced as an anti-tussive agent, owing to its several mechanisms of actions, dextromethorphan has found potential use as both an analgesic and anticonvulsive agent [187,188,189]. Dextromethorphan was noted to have noncompetitive NMDAR blocking effects with an efficacy similar to that of controlled substances, such as phencyclidine and ketamine, at high doses.
Dextromethrophan, added to existing AEDs at doses of 40 mg and 50 mg every 6 h (160 and 200 mg/day, respectively) for 8 weeks, resulted in significant improvements in seizure control in patients with drug-resistant, localization-related epilepsies [189]. Furthermore, a randomized, open-label trial of dextromethorphan in Rett’s syndrome showed evidence of significant dose-dependent improvements in clinical seizures [190]. This preliminary evidence needs further validation in larger cohorts.
The adverse effects of dextromethorphan include nystagmus, slurred speech, light-headedness, and fatigue at high doses [191]. In addition, sudden tonic–clonic movements and confusion have been reported to be caused by the toxicity associated with dextromethorphan in a case report [192].

4.10. Ifenprodil

Ifenprodil (4-[2-(4-benzylpiperidin-1-yl)-1-hydroxypropyl]phenol) is a selective antagonist of the GluN2B subtype of NMDARs [193]. The anticonvulsant effect of ifenprodil has been investigated in several animal models of epilepsy, including induced seizures by NMDA, spermine, lindane, and PTZ in rodents [194,195,196,197,198]. Furthermore, both age-dependent and activation-dependent anticonvulsant actions of ifenprodil were noted [199]. In five human patients with refractory epilepsy caused by malformations of cortical development (MCD), ifenprodil had specific antiepileptic effects by reducing pyramidal cell neural excitability [200]. More recently, in temporal lobe epilepsy medicated by the GluN2B subtype of NMDARs, intraperitoneal ifenprodil administration (20 mg/kg) resulted in the suppression of a number of chronic seizures and an anti-ictogenic effect [201].
When used in combination with other AEDs, the threshold for seizures was increased without influencing the anticonvulsant actions of other drugs (carbamazepine, diphenylhydantoin, phenobarbital, and valproate) [202].

5. Preclinical Studies with Newer NMDAR Modulators

A variety of NMDR modulators have been investigated in preclinical studies using various seizure models [11]. More recent studies have also tested newer agents that modulate NMDARs (Table 1). In a mouse model on clonic seizures induced by PTC, the involvement of the NMDAR pathway in the anticonvulsant effect of licofelone (dual 5-lipoxygenase/cyclooxygenase inhibitor) was demonstrated by combining the noncompetitive NMDAR antagonist MK-801 with licofelone (5 mg/kg) [203]. In addition, a lower dose of licofelone did not have an anticonvulsant effect [203]. Recently, the use of GNE-0723 (a positive allosteric modulator of GluN2A-subunit-containing NMDARs) reduced low-frequency oscillatory and epileptiform activities in J20 mice [204]. The GluN2B-selective antagonist Ro 25-6981 suppressed the tonic phase of generalized tonic–clonic seizures in a PTZ model of infantile rats [205].

6. Adverse Effects of NMDAR Antagonist in Clinical Settings

The inhibition of the major excitatory neurotransmitter glutamate is bound to have adverse effects that can potentially limit its potential for clinical application. In a clinical investigation, the competitive NMDAR antagonist D-CPP-ene increased seizures in three out of eight patients with epilepsy, raising the possibility that a sudden decline in NMDAR function could lead to an imbalance between the excitatory and inhibitory systems [207]. A selective blockade of NMDARs without affecting normal function remains a necessity for acceptability in clinical practice. An ideal agent should hence serve the role of an “uncompetitive” antagonist by relying on prior receptor activation by the agonist [208].
Adverse effects, such as hallucinations, lightheadedness, dizziness, fatigue, headaches, out-of-body sensation, and sensory changes, have been reported with NMDA antagonists. Early memory impairments and schizophrenia-like symptoms have been linked to NMDAR hypofunction with the use of antagonists [209]. An early blockade of NMDARs in rat brains has been found to trigger apoptotic neurodegeneration [210]. Brain growth, long-term potentiation, neuronal migration, and synaptic pruning are all significantly influenced by NMDAR activity [211]. Accordingly, the in utero use of NMDA antagonists may disrupt brain development by contributing directly to the disconnection of circuits between the hippocampus and frontal cortex [212].
Impairments in learning ability and memory as the result of NMDAR antagonist therapy, especially in early life, have been extensively studied in rat models [213,214]. These were notably more evident upon direct NMDAR antagonist injection into the amygdala and hippocampus [215]. In human subjects, following ketamine infusions, disruptions in frontal and hippocampal responses contributing to working memory were demonstrated through the use of fMRI imaging [216]. However, it is worth noting that NMDAR blockade impaired learning consolidation without having any effect on the memory retrieval of previously learned tasks [217,218].
Furthermore, in addition to intravenous administration, as extensively previously discussed, other routes of administration can affect the appearance of specific cytotoxic side effects. For example, spinal cord pathology has been noted with intrathecal administration [219]. The intrathecal administration of ketamine in a therapeutically appropriate concentration and dosage had a deleterious effect on rabbits’ central nervous systems [220]. In dogs, the subarachnoid administration of ketamine was found to be associated with histological spinal cord alterations, including gliosis, axonal edema, central chromatolysis, lymphocyte infiltration, and fibrous thickening of the dura mater [221]. Additionally, a terminally ill patient with cancer who had continuous intrathecal ketamine infusions at a rate of 5 mg/day for a period of 3 weeks was reported by Karpinski et al. to have post-mortem CNS histological abnormalities of subpial spinal cord vacuolation [222].

7. Conclusions and Perspectives

Overwhelming evidence now indicates that the NMDAR complex plays a critical role in seizure disorders. This comes from both preclinical and clinical studies that have found alterations in NMDAR expression and function in seizure models, as well as in patients with epilepsy. Accordingly, various NMDAR modulators have been tested in various animal models of seizures, and in these studies, they have shown efficacy in suppressing different types of seizures. Although few NMDAR antagonists have been evaluated in clinical trials in epileptic patients, the results are promising overall and have opened a new avenue for the treatment of epilepsy. However, long-term therapy with a newer class of NMDAR antagonists warrants further longitudinal studies, especially to assess their safety in this patient population.

Author Contributions

S.S. outlined the performed rigorous literature search, designed the table, and wrote the manuscript. M.G. conceived and designed the review, outlined the performed rigorous literature search, designed the figure, and edited the manuscript. S.C.S. conceived and designed the review and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

M.G. is supported by a clinical research training scholarship in ALS funded by The ALS Association and The American Brain Foundation, in collaboration with the American Academy of Neurology, and an NIH-funded Wellstone fellowship training grant (NIH 5P50HD060848-15) for research on FSHD.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fiest, K.M.; Sauro, K.M.; Wiebe, S.; Patten, S.B.; Kwon, C.S.; Dykeman, J.; Pringsheim, T.; Lorenzetti, D.L.; Jetté, N. Prevalence and incidence of epilepsy: A systematic review and meta-analysis of international studies. Neurology 2017, 88, 296–303. [Google Scholar] [CrossRef] [PubMed]
  2. Moshé, S.L.; Perucca, E.; Ryvlin, P.; Tomson, T. Epilepsy: New advances. Lancet 2015, 385, 884–898. [Google Scholar] [CrossRef]
  3. Pérez-Pérez, D.; Frías-Soria, C.L.; Rocha, L. Drug-resistant epilepsy: From multiple hypotheses to an integral explanation using preclinical resources. Epilepsy Behav. 2021, 121, 106430. [Google Scholar] [CrossRef]
  4. Kwan, P.; Arzimanoglou, A.; Berg, A.T.; Brodie, M.J.; Allen Hauser, W.; Mathern, G.; Moshé, S.L.; Perucca, E.; Wiebe, S.; French, J. Definition of drug resistant epilepsy: Consensus proposal by the ad hoc Task Force of the ILAE Commission on Therapeutic Strategies. Epilepsia 2010, 51, 1069–1077. [Google Scholar] [CrossRef] [PubMed]
  5. Sultana, B.; Panzini, M.A.; Veilleux Carpentier, A.; Comtois, J.; Rioux, B.; Gore, G.; Bauer, P.R.; Kwon, C.S.; Jetté, N.; Josephson, C.B.; et al. Incidence and Prevalence of Drug-Resistant Epilepsy: A Systematic Review and Meta-analysis. Neurology 2021, 96, 805–817. [Google Scholar] [CrossRef]
  6. Gan, J.; Qu, Y.; Li, J.; Zhao, F.; Mu, D. An evaluation of the links between microRNA, autophagy, and epilepsy. Rev. Neurosci. 2015, 26, 225–237. [Google Scholar] [CrossRef]
  7. Albrecht, J.; Zielińska, M. Mechanisms of Excessive Extracellular Glutamate Accumulation in Temporal Lobe Epilepsy. Neurochem. Res. 2017, 42, 1724–1734. [Google Scholar] [CrossRef]
  8. Levite, M. GLUTAMATE RECEPTOR ANTIBODIES IN NEUROLOGICAL DISEASES: Anti-AMPA-GluR3 antibodies, Anti-NMDA-NR1 antibodies, Anti-NMDA-NR2A/B antibodies, Anti-mGluR1 antibodies or Anti-mGluR5 antibodies are present in subpopulations of patients with either: Epilepsy, Encephalitis, Cerebellar Ataxia, Systemic Lupus Erythematosus (SLE) and Neuropsychiatric SLE, Sjogren’s syndrome, Schizophrenia, Mania or Stroke. These autoimmune anti-glutamate receptor antibodies can bind neurons in few brain regions, activate glutamate receptors, decrease glutamate receptor’s expression, impair glutamate-induced signaling and function, activate Blood Brain Barrier endothelial cells, kill neurons, damage the brain, induce behavioral/psychiatric/cognitive abnormalities and Ataxia in animal models, and can be removed or silenced in some patients by immunotherapy. J. Neural Transm. 2014, 121, 1029–1075. [Google Scholar] [CrossRef]
  9. Hendry, S.H.; Schwark, H.D.; Jones, E.G.; Yan, J. Numbers and proportions of GABA-immunoreactive neurons in different areas of monkey cerebral cortex. J. Neurosci. 1987, 7, 1503–1519. [Google Scholar] [CrossRef]
  10. Hanada, T. Ionotropic Glutamate Receptors in Epilepsy: A Review Focusing on AMPA and NMDA Receptors. Biomolecules 2020, 10, 464. [Google Scholar] [CrossRef]
  11. Ghasemi, M.; Schachter, S.C. The NMDA receptor complex as a therapeutic target in epilepsy: A review. Epilepsy Behav. 2011, 22, 617–640. [Google Scholar] [CrossRef] [PubMed]
  12. Paoletti, P. Molecular basis of NMDA receptor functional diversity. Eur. J. Neurosci. 2011, 33, 1351–1365. [Google Scholar] [CrossRef] [PubMed]
  13. Mennerick, S.; Zorumski, C.F. Neural activity and survival in the developing nervous system. Mol. Neurobiol. 2000, 22, 41–54. [Google Scholar] [CrossRef]
  14. Li, M.; Long, C.; Yang, L. Hippocampal-Prefrontal Circuit and Disrupted Functional Connectivity in Psychiatric and Neurodegenerative Disorders. BioMed Res. Int. 2015, 2015, 810548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Collingridge, G.L.; Bliss, T.V. Memories of NMDA receptors and LTP. Trends Neurosci. 1995, 18, 54–56. [Google Scholar] [CrossRef]
  16. Rison, R.A.; Stanton, P.K. Long-term potentiation and N-methyl-D-aspartate receptors: Foundations of memory and neurologic disease? Neurosci. Biobehav. Rev. 1995, 19, 533–552. [Google Scholar] [CrossRef]
  17. Morris, R.G.; Anderson, E.; Lynch, G.S.; Baudry, M. Selective impairment of learning and blockade of long-term potentiation by an N-methyl-D-aspartate receptor antagonist, AP5. Nature 1986, 319, 774–776. [Google Scholar] [CrossRef]
  18. Yang, S.; Seo, H.; Wang, M.; Arnsten, A.F.T. NMDAR Neurotransmission Needed for Persistent Neuronal Firing: Potential Roles in Mental Disorders. Front. Psychiatry 2021, 12, 654322. [Google Scholar] [CrossRef]
  19. Lau, C.G.; Takeuchi, K.; Rodenas-Ruano, A.; Takayasu, Y.; Murphy, J.; Bennett, M.V.; Zukin, R.S. Regulation of NMDA receptor Ca2+ signalling and synaptic plasticity. Biochem. Soc. Trans. 2009, 37, 1369–1374. [Google Scholar] [CrossRef] [Green Version]
  20. Jahr, C.E.; Stevens, C.F. Calcium permeability of the N-methyl-D-aspartate receptor channel in hippocampal neurons in culture. Proc. Natl. Acad. Sci. USA 1993, 90, 11573–11577. [Google Scholar] [CrossRef]
  21. Barker-Haliski, M.; White, H.S. Glutamatergic Mechanisms Associated with Seizures and Epilepsy. Cold Spring Harb. Perspect. Med. 2015, 5, a022863. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Kim, E.; Sheng, M. PDZ domain proteins of synapses. Nat. Rev. Neurosci. 2004, 5, 771–781. [Google Scholar] [CrossRef]
  23. Ladépêche, L.; Dupuis, J.P.; Groc, L. Surface trafficking of NMDA receptors: Gathering from a partner to another. Semin. Cell Dev. Biol. 2014, 27, 3–13. [Google Scholar] [CrossRef] [PubMed]
  24. Bard, L.; Sainlos, M.; Bouchet, D.; Cousins, S.; Mikasova, L.; Breillat, C.; Stephenson, F.A.; Imperiali, B.; Choquet, D.; Groc, L. Dynamic and specific interaction between synaptic NR2-NMDA receptor and PDZ proteins. Proc. Natl. Acad. Sci. USA 2010, 107, 19561–19566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Lau, C.G.; Zukin, R.S. NMDA receptor trafficking in synaptic plasticity and neuropsychiatric disorders. Nat. Rev. Neurosci. 2007, 8, 413–426. [Google Scholar] [CrossRef] [PubMed]
  26. Wasterlain, C.G.; Naylor, D.E.; Liu, H.; Niquet, J.; Baldwin, R. Trafficking of NMDA receptors during status epilepticus: Therapeutic implications. Epilepsia 2013, 54 (Suppl. S6), 78–80. [Google Scholar] [CrossRef] [PubMed]
  27. Mele, M.; Costa, R.O.; Duarte, C.B. Alterations in GABA(A)-Receptor Trafficking and Synaptic Dysfunction in Brain Disorders. Front. Cell. Neurosci. 2019, 13, 77. [Google Scholar] [CrossRef]
  28. Addis, L.; Virdee, J.K.; Vidler, L.R.; Collier, D.A.; Pal, D.K.; Ursu, D. Epilepsy-associated GRIN2A mutations reduce NMDA receptor trafficking and agonist potency—Molecular profiling and functional rescue. Sci. Rep. 2017, 7, 66. [Google Scholar] [CrossRef] [Green Version]
  29. Mota Vieira, M.; Nguyen, T.A.; Wu, K.; Badger, J.D.; Collins, B.M.; Anggono, V.; Lu, W.; Roche, K.W. An Epilepsy-Associated GRIN2A Rare Variant Disrupts CaMKIIα Phosphorylation of GluN2A and NMDA Receptor Trafficking. Cell Rep. 2020, 32, 108104. [Google Scholar] [CrossRef]
  30. Yang, Y.; Tian, X.; Xu, D.; Zheng, F.; Lu, X.; Zhang, Y.; Ma, Y.; Li, Y.; Xu, X.; Zhu, B.; et al. GPR40 modulates epileptic seizure and NMDA receptor function. Sci. Adv. 2018, 4, eaau2357. [Google Scholar] [CrossRef]
  31. Bergeron, R.; Meyer, T.M.; Coyle, J.T.; Greene, R.W. Modulation of N-methyl-D-aspartate receptor function by glycine transport. Proc. Natl. Acad. Sci. USA 1998, 95, 15730–15734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Mothet, J.P.; Le Bail, M.; Billard, J.M. Time and space profiling of NMDA receptor co-agonist functions. J. Neurochem. 2015, 135, 210–225. [Google Scholar] [CrossRef] [PubMed]
  33. Zhu, S.; Paoletti, P. Allosteric modulators of NMDA receptors: Multiple sites and mechanisms. Curr. Opin. Pharmacol. 2015, 20, 14–23. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, X.; Hu, B.; Lu, L.; Xu, D.; Sun, L.; Lin, W. D-serine and NMDA Receptor 1 Expression in Patients with Intractable Epilepsy. Turk. Neurosurg. 2021, 31, 76–82. [Google Scholar] [CrossRef]
  35. Singh, L.; Oles, R.J.; Tricklebank, M.D. Modulation of seizure susceptibility in the mouse by the strychnine-insensitive glycine recognition site of the NMDA receptor/ion channel complex. Br. J. Pharmacol. 1990, 99, 285–288. [Google Scholar] [CrossRef] [Green Version]
  36. Meftah, A.; Hasegawa, H.; Kantrowitz, J.T. D-Serine: A Cross Species Review of Safety. Front. Psychiatry 2021, 12, 726365. [Google Scholar] [CrossRef]
  37. Mony, L.; Kew, J.N.; Gunthorpe, M.J.; Paoletti, P. Allosteric modulators of NR2B-containing NMDA receptors: Molecular mechanisms and therapeutic potential. Br. J. Pharmacol. 2009, 157, 1301–1317. [Google Scholar] [CrossRef] [Green Version]
  38. Zhang, J.-B.; Chang, S.; Xu, P.; Miao, M.; Wu, H.; Zhang, Y.; Zhang, T.; Wang, H.; Zhang, J.; Xie, C.; et al. Structural Basis of the Proton Sensitivity of Human GluN1-GluN2A NMDA Receptors. Cell Rep. 2018, 25, 3582–3590.e3584. [Google Scholar] [CrossRef] [Green Version]
  39. Regan, M.C.; Zhu, Z.; Yuan, H.; Myers, S.J.; Menaldino, D.S.; Tahirovic, Y.A.; Liotta, D.C.; Traynelis, S.F.; Furukawa, H. Structural elements of a pH-sensitive inhibitor binding site in NMDA receptors. Nat. Commun. 2019, 10, 321. [Google Scholar] [CrossRef] [Green Version]
  40. Schreiber, J.A.; Schepmann, D.; Frehland, B.; Thum, S.; Datunashvili, M.; Budde, T.; Hollmann, M.; Strutz-Seebohm, N.; Wünsch, B.; Seebohm, G. A common mechanism allows selective targeting of GluN2B subunit-containing N-methyl-D-aspartate receptors. Commun. Biol. 2019, 2, 420. [Google Scholar] [CrossRef]
  41. Low, C.M.; Zheng, F.; Lyuboslavsky, P.; Traynelis, S.F. Molecular determinants of coordinated proton and zinc inhibition of N-methyl-D-aspartate NR1/NR2A receptors. Proc. Natl. Acad. Sci. USA 2000, 97, 11062–11067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Mannara, F.; Radosevic, M.; Planagumà, J.; Soto, D.; Aguilar, E.; García-Serra, A.; Maudes, E.; Pedreño, M.; Paul, S.; Doherty, J.; et al. Allosteric modulation of NMDA receptors prevents the antibody effects of patients with anti-NMDAR encephalitis. Brain 2020, 143, 2709–2720. [Google Scholar] [CrossRef] [PubMed]
  43. Hamamoto, O.; Tirapelli, D.P.d.C.; Lizarte Neto, F.S.; Freitas-Lima, P.; Saggioro, F.P.; Cirino, M.L.d.A.; Assirati, J.A., Jr.; Serafini, L.N.; Velasco, T.R.; Sakamoto, A.C.; et al. Modulation of NMDA receptor by miR-219 in the amygdala and hippocampus of patients with mesial temporal lobe epilepsy. J. Clin. Neurosci. 2020, 74, 180–186. [Google Scholar] [CrossRef] [PubMed]
  44. Ure, J.; Baudry, M.; Perassolo, M. Metabotropic glutamate receptors and epilepsy. J. Neurol. Sci. 2006, 247, 1–9. [Google Scholar] [CrossRef]
  45. Alagarsamy, S.; Rouse, S.T.; Junge, C.; Hubert, G.W.; Gutman, D.; Smith, Y.; Conn, P.J. NMDA-induced phosphorylation and regulation of mGluR5. Pharmacol. Biochem. Behav. 2002, 73, 299–306. [Google Scholar] [CrossRef]
  46. Chen, H.-H.; Liao, P.-F.; Chan, M.-H. mGluR5 positive modulators both potentiate activation and restore inhibition in NMDA receptors by PKC dependent pathway. J. Biomed. Sci. 2011, 18, 19. [Google Scholar] [CrossRef] [Green Version]
  47. Pietraszek, M.; Gravius, A.; Schäfer, D.; Weil, T.; Trifanova, D.; Danysz, W. mGluR5, but not mGluR1, antagonist modifies MK-801-induced locomotor activity and deficit of prepulse inhibition. Neuropharmacology 2005, 49, 73–85. [Google Scholar] [CrossRef]
  48. Henry, S.A.; Lehmann-Masten, V.; Gasparini, F.; Geyer, M.A.; Markou, A. The mGluR5 antagonist MPEP, but not the mGluR2/3 agonist LY314582, augments PCP effects on prepulse inhibition and locomotor activity. Neuropharmacology 2002, 43, 1199–1209. [Google Scholar] [CrossRef]
  49. Chapman, A.G.; Nanan, K.; Williams, M.; Meldrum, B.S. Anticonvulsant activity of two metabotropic glutamate group I antagonists selective for the mGlu5 receptor: 2-methyl-6-(phenylethynyl)-pyridine (MPEP), and (E)-6-methyl-2-styryl-pyridine (SIB 1893). Neuropharmacology 2000, 39, 1567–1574. [Google Scholar] [CrossRef]
  50. Yan, Q.J.; Rammal, M.; Tranfaglia, M.; Bauchwitz, R.P. Suppression of two major Fragile X Syndrome mouse model phenotypes by the mGluR5 antagonist MPEP. Neuropharmacology 2005, 49, 1053–1066. [Google Scholar] [CrossRef]
  51. Kano, M.; Watanabe, T. Type-1 metabotropic glutamate receptor signaling in cerebellar Purkinje cells in health and disease. F1000Research 2017, 6, 416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Sillevis Smitt, P.; Kinoshita, A.; De Leeuw, B.; Moll, W.; Coesmans, M.; Jaarsma, D.; Henzen-Logmans, S.; Vecht, C.; De Zeeuw, C.; Sekiyama, N.; et al. Paraneoplastic cerebellar ataxia due to autoantibodies against a glutamate receptor. N. Engl. J. Med. 2000, 342, 21–27. [Google Scholar] [CrossRef] [PubMed]
  53. Höftberger, R.; Sabater, L.; Ortega, A.; Dalmau, J.; Graus, F. Patient with homer-3 antibodies and cerebellitis. JAMA Neurol. 2013, 70, 506–509. [Google Scholar] [CrossRef] [Green Version]
  54. Sarantis, K.; Antoniou, K.; Matsokis, N.; Angelatou, F. Exposure to novel environment is characterized by an interaction of D1/NMDA receptors underlined by phosphorylation of the NMDA and AMPA receptor subunits and activation of ERK1/2 signaling, leading to epigenetic changes and gene expression in rat hippocampus. Neurochem. Int. 2012, 60, 55–67. [Google Scholar] [CrossRef]
  55. Laurido-Soto, O.; Brier, M.R.; Simon, L.E.; McCullough, A.; Bucelli, R.C.; Day, G.S. Patient characteristics and outcome associations in AMPA receptor encephalitis. J. Neurol. 2019, 266, 450–460. [Google Scholar] [CrossRef]
  56. Höftberger, R.; van Sonderen, A.; Leypoldt, F.; Houghton, D.; Geschwind, M.; Gelfand, J.; Paredes, M.; Sabater, L.; Saiz, A.; Titulaer, M.J.; et al. Encephalitis and AMPA receptor antibodies: Novel findings in a case series of 22 patients. Neurology 2015, 84, 2403–2412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Peng, X.; Hughes, E.G.; Moscato, E.H.; Parsons, T.D.; Dalmau, J.; Balice-Gordon, R.J. Cellular plasticity induced by anti-α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptor encephalitis antibodies. Ann. Neurol. 2015, 77, 381–398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Schidlitzki, A.; Twele, F.; Klee, R.; Waltl, I.; Römermann, K.; Bröer, S.; Meller, S.; Gerhauser, I.; Rankovic, V.; Li, D.; et al. A combination of NMDA and AMPA receptor antagonists retards granule cell dispersion and epileptogenesis in a model of acquired epilepsy. Sci. Rep. 2017, 7, 12191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Mathern, G.W.; Pretorius, J.K.; Mendoza, D.; Leite, J.P.; Chimelli, L.; Born, D.E.; Fried, I.; Assirati, J.A.; Ojemann, G.A.; Adelson, P.D.; et al. Hippocampal N-methyl-D-aspartate receptor subunit mRNA levels in temporal lobe epilepsy patients. Ann. Neurol. 1999, 46, 343–358. [Google Scholar] [CrossRef]
  60. Mathern, G.W.; Pretorius, J.K.; Kornblum, H.I.; Mendoza, D.; Lozada, A.; Leite, J.P.; Chimelli, L.M.; Fried, I.; Sakamoto, A.C.; Assirati, J.A.; et al. Human hippocampal AMPA and NMDA mRNA levels in temporal lobe epilepsy patients. Brain 1997, 120 Pt 11, 1937–1959. [Google Scholar] [CrossRef]
  61. Mathern, G.W.; Leite, J.P.; Babb, T.L.; Pretorius, J.K.; Kuhlman, P.A.; Mendoza, D.; Fried, I.; Sakamoto, A.C.; Assirati, J.A.; Adelson, P.D. Aberrant hippocampal mossy fiber sprouting correlates with greater NMDAR2 receptor staining. Neuroreport 1996, 7, 1029–1035. [Google Scholar] [CrossRef] [PubMed]
  62. Franck, J.E.; Pokorny, J.; Kunkel, D.D.; Schwartzkroin, P.A. Physiologic and morphologic characteristics of granule cell circuitry in human epileptic hippocampus. Epilepsia 1995, 36, 543–558. [Google Scholar] [CrossRef] [PubMed]
  63. Isokawa, M.; Levesque, M.F. Increased NMDA responses and dendritic degeneration in human epileptic hippocampal neurons in slices. Neurosci. Lett. 1991, 132, 212–216. [Google Scholar] [CrossRef]
  64. Masukawa, L.M.; Higashima, M.; Hart, G.J.; Spencer, D.D.; O’Connor, M.J. NMDA receptor activation during epileptiform responses in the dentate gyrus of epileptic patients. Brain Res. 1991, 562, 176–180. [Google Scholar] [CrossRef]
  65. Banerjee, J.; Banerjee Dixit, A.; Tripathi, M.; Sarkar, C.; Gupta, Y.K.; Chandra, P.S. Enhanced endogenous activation of NMDA receptors in pyramidal neurons of hippocampal tissues from patients with mesial temporal lobe epilepsy: A mechanism of hyper excitation. Epilepsy Res. 2015, 117, 11–16. [Google Scholar] [CrossRef]
  66. McGinnity, C.J.; Koepp, M.J.; Hammers, A.; Riaño Barros, D.A.; Pressler, R.M.; Luthra, S.; Jones, P.A.; Trigg, W.; Micallef, C.; Symms, M.R.; et al. NMDA receptor binding in focal epilepsies. J. Neurol. Neurosurg. Psychiatry 2015, 86, 1150–1157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. André, V.M.; Flores-Hernández, J.; Cepeda, C.; Starling, A.J.; Nguyen, S.; Lobo, M.K.; Vinters, H.V.; Levine, M.S.; Mathern, G.W. NMDA receptor alterations in neurons from pediatric cortical dysplasia tissue. Cereb. Cortex 2004, 14, 634–646. [Google Scholar] [CrossRef] [PubMed]
  68. White, R.; Hua, Y.; Scheithauer, B.; Lynch, D.R.; Henske, E.P.; Crino, P.B. Selective alterations in glutamate and GABA receptor subunit mRNA expression in dysplastic neurons and giant cells of cortical tubers. Ann. Neurol. 2001, 49, 67–78. [Google Scholar] [CrossRef]
  69. Zhu, J.M.; Li, K.X.; Cao, S.X.; Chen, X.J.; Shen, C.J.; Zhang, Y.; Geng, H.Y.; Chen, B.Q.; Lian, H.; Zhang, J.M.; et al. Increased NRG1-ErbB4 signaling in human symptomatic epilepsy. Sci. Rep. 2017, 7, 141. [Google Scholar] [CrossRef]
  70. Jeon, A.R.; Kim, J.E. PDI Knockdown Inhibits Seizure Activity in Acute Seizure and Chronic Epilepsy Rat Models via S-Nitrosylation-Independent Thiolation on NMDA Receptor. Front. Cell. Neurosci. 2018, 12, 438. [Google Scholar] [CrossRef]
  71. Liu, S.; Liu, C.; Xiong, L.; Xie, J.; Huang, C.; Pi, R.; Huang, Z.; Li, L. Icaritin Alleviates Glutamate-Induced Neuronal Damage by Inactivating GluN2B-Containing NMDARs Through the ERK/DAPK1 Pathway. Front. Neurosci. 2021, 15, 525615. [Google Scholar] [CrossRef] [PubMed]
  72. Endele, S.; Rosenberger, G.; Geider, K.; Popp, B.; Tamer, C.; Stefanova, I.; Milh, M.; Kortüm, F.; Fritsch, A.; Pientka, F.K. Mutations in GRIN2A and GRIN2B encoding regulatory subunits of NMDA receptors cause variable neurodevelopmental phenotypes. Nat. Genet. 2010, 42, 1021–1026. [Google Scholar] [CrossRef] [PubMed]
  73. Xu, X.X.; Luo, J.H. Mutations of N-Methyl-D-Aspartate Receptor Subunits in Epilepsy. Neurosci. Bull. 2018, 34, 549–565. [Google Scholar] [CrossRef]
  74. Fry, A.E.; Fawcett, K.A.; Zelnik, N.; Yuan, H.; Thompson, B.A.N.; Shemer-Meiri, L.; Cushion, T.D.; Mugalaasi, H.; Sims, D.; Stoodley, N.; et al. De novo mutations in GRIN1 cause extensive bilateral polymicrogyria. Brain 2018, 141, 698–712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Elmasri, M.; Hunter, D.W.; Winchester, G.; Bates, E.E.; Aziz, W.; Van Der Does, D.M.; Karachaliou, E.; Sakimura, K.; Penn, A.C. Common synaptic phenotypes arising from diverse mutations in the human NMDA receptor subunit GluN2A. Commun. Biol. 2022, 5, 174. [Google Scholar] [CrossRef]
  76. Myers, S.J.; Yuan, H.; Kang, J.Q.; Tan, F.C.K.; Traynelis, S.F.; Low, C.M. Distinct roles of GRIN2A and GRIN2B variants in neurological conditions. F1000Research 2019, 8, F1000. [Google Scholar] [CrossRef] [Green Version]
  77. XiangWei, W.; Jiang, Y.; Yuan, H. De Novo Mutations and Rare Variants Occurring in NMDA Receptors. Curr. Opin. Physiol. 2018, 2, 27–35. [Google Scholar] [CrossRef]
  78. Punnakkal, P.; Dominic, D. NMDA Receptor GluN2 Subtypes Control Epileptiform Events in the Hippocampus. Neuromol. Med. 2018, 20, 90–96. [Google Scholar] [CrossRef]
  79. Camp, C.R.; Yuan, H. GRIN2D/GluN2D NMDA receptor: Unique features and its contribution to pediatric developmental and epileptic encephalopathy. Eur. J. Paediatr. Neurol. 2020, 24, 89–99. [Google Scholar] [CrossRef]
  80. Zhang, W.; Ross, P.J.; Ellis, J.; Salter, M.W. Targeting NMDA receptors in neuropsychiatric disorders by drug screening on human neurons derived from pluripotent stem cells. Transl. Psychiatry 2022, 12, 243. [Google Scholar] [CrossRef]
  81. Spatola, M.; Dalmau, J. Seizures and risk of epilepsy in autoimmune and other inflammatory encephalitis. Curr. Opin. Neurol. 2017, 30, 345–353. [Google Scholar] [CrossRef] [PubMed]
  82. Dubey, D.; Pittock, S.J.; Kelly, C.R.; McKeon, A.; Lopez-Chiriboga, A.S.; Lennon, V.A.; Gadoth, A.; Smith, C.Y.; Bryant, S.C.; Klein, C.J.; et al. Autoimmune encephalitis epidemiology and a comparison to infectious encephalitis. Ann. Neurol. 2018, 83, 166–177. [Google Scholar] [CrossRef]
  83. Leypoldt, F.; Armangue, T.; Dalmau, J. Autoimmune encephalopathies. Ann. N. Y. Acad. Sci. 2015, 1338, 94–114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. 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] [Green Version]
  85. Dalmau, J.; Lancaster, E.; Martinez-Hernandez, E.; Rosenfeld, M.R.; Balice-Gordon, R. Clinical experience and laboratory investigations in patients with anti-NMDAR encephalitis. Lancet Neurol. 2011, 10, 63–74. [Google Scholar] [CrossRef] [Green Version]
  86. Ladépêche, L.; Planagumà, J.; Thakur, S.; Suárez, I.; Hara, M.; Borbely, J.S.; Sandoval, A.; Laparra-Cuervo, L.; Dalmau, J.; Lakadamyali, M. NMDA Receptor Autoantibodies in Autoimmune Encephalitis Cause a Subunit-Specific Nanoscale Redistribution of NMDA Receptors. Cell Rep. 2018, 23, 3759–3768. [Google Scholar] [CrossRef]
  87. Planagumà, J.; Haselmann, H.; Mannara, F.; Petit-Pedrol, M.; Grünewald, B.; Aguilar, E.; Röpke, L.; Martín-García, E.; Titulaer, M.J.; Jercog, P.; et al. Ephrin-B2 prevents N-methyl-D-aspartate receptor antibody effects on memory and neuroplasticity. Ann. Neurol. 2016, 80, 388–400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Planagumà, J.; Leypoldt, F.; Mannara, F.; Gutiérrez-Cuesta, J.; Martín-García, E.; Aguilar, E.; Titulaer, M.J.; Petit-Pedrol, M.; Jain, A.; Balice-Gordon, R.; et al. Human N-methyl D-aspartate receptor antibodies alter memory and behaviour in mice. Brain 2014, 138, 94–109. [Google Scholar] [CrossRef] [PubMed]
  89. Wright, S.; Hashemi, K.; Stasiak, L.; Bartram, J.; Lang, B.; Vincent, A.; Upton, A.L. Epileptogenic effects of NMDAR antibodies in a passive transfer mouse model. Brain 2015, 138, 3159–3167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Wright, S.K.; Rosch, R.E.; Wilson, M.A.; Upadhya, M.A.; Dhangar, D.R.; Clarke-Bland, C.; Wahid, T.T.; Barman, S.; Goebels, N.; Kreye, J.; et al. Multimodal electrophysiological analyses reveal that reduced synaptic excitatory neurotransmission underlies seizures in a model of NMDAR antibody-mediated encephalitis. Commun. Biol. 2021, 4, 1106. [Google Scholar] [CrossRef]
  91. Wandinger, K.P.; Saschenbrecker, S.; Stoecker, W.; Dalmau, J. Anti-NMDA-receptor encephalitis: A severe, multistage, treatable disorder presenting with psychosis. J. Neuroimmunol. 2011, 231, 86–91. [Google Scholar] [CrossRef] [PubMed]
  92. Irani, S.R.; Vincent, A. NMDA receptor antibody encephalitis. Curr. Neurol. Neurosci. Rep. 2011, 11, 298–304. [Google Scholar] [CrossRef] [PubMed]
  93. Finke, C.; Kopp, U.A.; Prüss, H.; Dalmau, J.; Wandinger, K.P.; Ploner, C.J. Cognitive deficits following anti-NMDA receptor encephalitis. J. Neurol. Neurosurg. Psychiatry 2012, 83, 195–198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Mealing, G.A.; Lanthorn, T.H.; Murray, C.L.; Small, D.L.; Morley, P. Differences in degree of trapping of low-affinity uncompetitive N-methyl-D-aspartic acid receptor antagonists with similar kinetics of block. J. Pharmacol. Exp. Ther. 1999, 288, 204–210. [Google Scholar] [PubMed]
  95. Marland, S.; Ellerton, J.; Andolfatto, G.; Strapazzon, G.; Thomassen, O.; Brandner, B.; Weatherall, A.; Paal, P. Ketamine: Use in anesthesia. CNS Neurosci. Ther. 2013, 19, 381–389. [Google Scholar] [CrossRef]
  96. Ghasemi, M.; Shafaroodi, H.; Nazarbeiki, S.; Meskar, H.; Heydarpour, P.; Ghasemi, A.; Talab, S.S.; Ziai, P.; Bahremand, A.; Dehpour, A.R. Voltage-dependent calcium channel and NMDA receptor antagonists augment anticonvulsant effects of lithium chloride on pentylenetetrazole-induced clonic seizures in mice. Epilepsy Behav. 2010, 18, 171–178. [Google Scholar] [CrossRef] [PubMed]
  97. 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]
  98. 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] [PubMed]
  99. Borsato, G.S.; Siegel, J.L.; Rose, M.Q.; Ojard, M.; Feyissa, A.M.; Quinones-Hinojosa, A.; Jackson, D.A.; Rogers, E.R.; Freeman, W.D. Ketamine in seizure management and future pharmacogenomic considerations. Pharmacogenom. J. 2020, 20, 351–354. [Google Scholar] [CrossRef]
  100. Meaden, C.W.; Barnes, S. Ketamine Implicated in New Onset Seizure. Clin. Pract. Cases Emerg. Med. 2019, 3, 401–404. [Google Scholar] [CrossRef]
  101. Borowicz, K.K.; Łuszczki, J.; Czuczwar, S.J. Interactions between non-barbiturate injectable anesthetics and conventional antiepileptic drugs in the maximal electroshock test in mice--an isobolographic analysis. Eur. Neuropsychopharmacol. 2004, 14, 163–172. [Google Scholar] [CrossRef]
  102. Martin, B.S.; Kapur, J. A combination of ketamine and diazepam synergistically controls refractory status epilepticus induced by cholinergic stimulation. Epilepsia 2008, 49, 248–255. [Google Scholar] [CrossRef] [Green Version]
  103. 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] [PubMed] [Green Version]
  104. 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 status epilepticus. Neurology 2020, 95, e2286. [Google Scholar] [CrossRef]
  105. Godoy, D.A.; Badenes, R.; Pelosi, P.; Robba, C. Ketamine in acute phase of severe traumatic brain injury “an old drug for new uses?”. Crit. Care 2021, 25, 19. [Google Scholar] [CrossRef] [PubMed]
  106. Hurth, K.P.; Jaworski, A.; Thomas, K.B.; Kirsch, W.B.; Rudoni, M.A.; Wohlfarth, K.M. The Reemergence of Ketamine for Treatment in Critically Ill Adults. Crit. Care Med. 2020, 48, 899–911. [Google Scholar] [CrossRef]
  107. Flower, O.; Hellings, S. Sedation in Traumatic Brain Injury. Emerg. Med. Int. 2012, 2012, 637171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Umunna, B.-P.; Tekwani, K.; Barounis, D.; Kettaneh, N.; Kulstad, E. Ketamine for continuous sedation of mechanically ventilated patients. J. Emerg. Trauma Shock 2015, 8, 11–15. [Google Scholar] [CrossRef] [PubMed]
  109. Luz, M.; Brandão Barreto, B.; de Castro, R.E.V.; Salluh, J.; Dal-Pizzol, F.; Araujo, C.; De Jong, A.; Chanques, G.; Myatra, S.N.; Tobar, E.; et al. Practices in sedation, analgesia, mobilization, delirium, and sleep deprivation in adult intensive care units (SAMDS-ICU): An international survey before and during the COVID-19 pandemic. Ann. Intensive Care 2022, 12, 9. [Google Scholar] [CrossRef] [PubMed]
  110. Hertle, D.N.; Dreier, J.P.; Woitzik, J.; Hartings, J.A.; Bullock, R.; Okonkwo, D.O.; Shutter, L.A.; Vidgeon, S.; Strong, A.J.; Kowoll, C.; et al. Effect of analgesics and sedatives on the occurrence of spreading depolarizations accompanying acute brain injury. Brain 2012, 135, 2390–2398. [Google Scholar] [CrossRef]
  111. Zou, X.; Patterson, T.A.; Sadovova, N.; Twaddle, N.C.; Doerge, D.R.; Zhang, X.; Fu, X.; Hanig, J.P.; Paule, M.G.; Slikker, W.; et al. Potential neurotoxicity of ketamine in the developing rat brain. Toxicol. Sci. 2009, 108, 149–158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Fujikawa, D.G. Neuroprotective effect of ketamine administered after status epilepticus onset. Epilepsia 1995, 36, 186–195. [Google Scholar] [CrossRef]
  113. Långsjö, J.W.; Maksimow, A.; Salmi, E.; Kaisti, K.; Aalto, S.; Oikonen, V.; Hinkka, S.; Aantaa, R.; Sipilä, H.; Viljanen, T.; et al. S-ketamine anesthesia increases cerebral blood flow in excess of the metabolic needs in humans. Anesthesiology 2005, 103, 258–268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Carlson, A.P.; Abbas, M.; Alunday, R.L.; Qeadan, F.; Shuttleworth, C.W. Spreading depolarization in acute brain injury inhibited by ketamine: A prospective, randomized, multiple crossover trial. J. Neurosurg. 2018, 130, 1513–1519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Dreier, J.P.; Major, S.; Pannek, H.W.; Woitzik, J.; Scheel, M.; Wiesenthal, D.; Martus, P.; Winkler, M.K.; Hartings, J.A.; Fabricius, M.; et al. Spreading convulsions, spreading depolarization and epileptogenesis in human cerebral cortex. Brain 2012, 135, 259–275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Kramer, D.R.; Fujii, T.; Ohiorhenuan, I.; Liu, C.Y. Interplay between Cortical Spreading Depolarization and Seizures. Stereotact. Funct. Neurosurg. 2017, 95, 1–5. [Google Scholar] [CrossRef] [PubMed]
  117. Foreman, B.; Lee, H.; Okonkwo, D.O.; Strong, A.J.; Pahl, C.; Shutter, L.A.; Dreier, J.P.; Ngwenya, L.B.; Hartings, J.A. The Relationship Between Seizures and Spreading Depolarizations in Patients with Severe Traumatic Brain Injury. Neurocrit. Care 2022, 37, 31–48. [Google Scholar] [CrossRef] [PubMed]
  118. 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-D-aspartate (NMDA) responses by memantine: Therapeutic advantage against NMDA receptor-mediated neurotoxicity. J. Neurosci. 1992, 12, 4427–4436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Chojnacka-Wójcik, E.; Tatarczyńska, E.; Maj, J. The influence of memantine on the anticonvulsant effects of the antiepileptic drugs. Pol. J. Pharmacol. Pharm. 1983, 35, 511–515. [Google Scholar] [PubMed]
  120. McLean, M.J. In vitro electrophysiological evidence predicting anticonvulsant efficacy of memantine and flunarizine. Pol. J. Pharmacol. Pharm. 1987, 39, 513–525. [Google Scholar] [PubMed]
  121. Urbańska, E.; Dziki, M.; Czuczwar, S.J.; Kleinrok, Z.; Turski, W.A. Antiparkinsonian drugs memantine and trihexyphenidyl potentiate the anticonvulsant activity of valproate against maximal electroshock-induced seizures. Neuropharmacology 1992, 31, 1021–1026. [Google Scholar] [CrossRef]
  122. Kalemenev, S.V.; Zubareva, O.E.; Sizov, V.V.; Lavrent’eva, V.V.; Lukomskaya, N.Y.; Kim, K.K.; Zaitsev, A.V.; Magazanik, L.G. Memantine attenuates cognitive impairments after status epilepticus induced in a lithium-pilocarpine model. Dokl. Biol. Sci. 2016, 470, 224–227. [Google Scholar] [CrossRef] [PubMed]
  123. 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] [PubMed]
  124. Czuczwar, S.J.; Turski, W.A.; Kleinrok, Z. Interactions of excitatory amino acid antagonists with conventional antiepileptic drugs. Metab. Brain Dis. 1996, 11, 143–152. [Google Scholar] [CrossRef] [PubMed]
  125. 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]
  126. Platzer, K.; Yuan, H.; Schütz, H.; Winschel, A.; Chen, W.; Hu, C.; Kusumoto, H.; Heyne, H.O.; Helbig, K.L.; Tang, S.; et al. GRIN2B encephalopathy: Novel findings on phenotype, variant clustering, functional consequences and treatment aspects. J. Med. Genet. 2017, 54, 460–470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Kornhuber, J.; Bormann, J.; Hübers, M.; Rusche, K.; Riederer, P. Effects of the 1-amino-adamantanes at the MK-801-binding site of the NMDA-receptor-gated ion channel: A human postmortem brain study. Eur. J. Pharmacol. Mol. Pharmacol. 1991, 206, 297–300. [Google Scholar] [CrossRef]
  128. Sreedharan, M.; Devadathan, K.; Pathan, H.K.; Chalipat, S.; Mohammed, K.P.A. Amantadine for the Treatment of Refractory Absence Seizures in Children. J. Pediatr. Neurosci. 2018, 13, 131–136. [Google Scholar] [CrossRef]
  129. Shahar, E.M.; Brand, N. Effect of add-on amantadine therapy for refractory absence epilepsy. J. Pediatr. 1992, 121, 819–821. [Google Scholar] [CrossRef]
  130. Perry, M.S.; Bailey, L.J.; Kotecha, A.C.; Malik, S.I.; Hernandez, A.W. Amantadine for the treatment of refractory absence seizures in children. Pediatr. Neurol. 2012, 46, 243–245. [Google Scholar] [CrossRef]
  131. Barra, M.E.; Edlow, B.L.; Brophy, G.M. Pharmacologic Therapies to Promote Recovery of Consciousness. Semin. Neurol. 2022, 42, 335–347. [Google Scholar] [CrossRef] [PubMed]
  132. Leclerc, A.M.; Riker, R.R.; Brown, C.S.; May, T.; Nocella, K.; Cote, J.; Eldridge, A.; Seder, D.B.; Gagnon, D.J. Amantadine and Modafinil as Neurostimulants Following Acute Stroke: A Retrospective Study of Intensive Care Unit Patients. Neurocrit. Care 2021, 34, 102–111. [Google Scholar] [CrossRef] [PubMed]
  133. Sawyer, E.; Mauro, L.S.; Ohlinger, M.J. Amantadine enhancement of arousal and cognition after traumatic brain injury. Ann. Pharmacother. 2008, 42, 247–252. [Google Scholar] [CrossRef]
  134. Suter, C.; Klingman, W.O. Neurologic Manifestations of Magnesium Depletion States. Neurology 1955, 5, 691. [Google Scholar] [CrossRef] [PubMed]
  135. Kampa, B.M.; Clements, J.; Jonas, P.; Stuart, G.J. Kinetics of Mg2+ unblock of NMDA receptors: Implications for spike-timing dependent synaptic plasticity. J. Physiol. 2004, 556, 337–345. [Google Scholar] [CrossRef] [PubMed]
  136. Kruse, H.D.; Orent, E.R.; McCollum, E.V. Studies on Magnesium Deficiency in Animals: I. Symptomatology Resulting from Magnesium Deprivation. J. Biol. Chem. 1932, 96, 519–539. [Google Scholar] [CrossRef]
  137. 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] [PubMed]
  138. Hallak, M.; Berman, R.F.; Irtenkauf, S.M.; Janusz, C.A.; Cotton, D.B. Magnesium sulfate treatment decreases N-methyl-D-aspartate receptor binding in the rat brain: An autoradiographic study. J. Soc. Gynecol. Investig. 1994, 1, 25–30. [Google Scholar] [CrossRef]
  139. 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]
  140. 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]
  141. 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]
  142. McCubbin, J.; Sibai, B.; Abdella, T.; Anderson, G. Cardiopulmonary arrest due to acute maternal hypermagnesaemia. Lancet 1981, 317, 1058. [Google Scholar] [CrossRef]
  143. Sachdeo, R.; Kramer, L.D.; Rosenberg, A.; Sachdeo, S. Felbamate monotherapy: Controlled trial in patients with partial onset seizures. Ann. Neurol. 1992, 32, 386–392. [Google Scholar] [CrossRef] [PubMed]
  144. Rho, J.M.; Donevan, S.D.; Rogawski, M.A. Mechanism of action of the anticonvulsant felbamate: Opposing effects on N-methyl-D-aspartate and gamma-aminobutyric acidA receptors. Ann. Neurol. 1994, 35, 229–234. [Google Scholar] [CrossRef] [PubMed]
  145. Harty, T.P.; Rogawski, M.A. Felbamate block of recombinant N-methyl-D-aspartate receptors: Selectivity for the NR2B subunit. Epilepsy Res. 2000, 39, 47–55. [Google Scholar] [CrossRef] [Green Version]
  146. French, J.; Smith, M.; Faught, E.; Brown, L. Practice advisory: The use of felbamate in the treatment of patients with intractable epilepsy: Report of the Quality Standards Subcommittee of the American Academy of Neurology and the American Epilepsy Society. Neurology 1999, 52, 1540–1545. [Google Scholar] [CrossRef] [Green Version]
  147. Zupanc, M.L.; Roell Werner, R.; Schwabe, M.S.; O’Connor, S.E.; Marcuccilli, C.J.; Hecox, K.E.; Chico, M.S.; Eggener, K.A. Efficacy of felbamate in the treatment of intractable pediatric epilepsy. Pediatr. Neurol. 2010, 42, 396–403. [Google Scholar] [CrossRef]
  148. Thakkar, K.; Billa, G.; Rane, J.; Chudasama, H.; Goswami, S.; Shah, R. The rise and fall of felbamate as a treatment for partial epilepsy—Aplastic anemia and hepatic failure to blame? Expert Rev. Neurother. 2015, 15, 1373–1375. [Google Scholar] [CrossRef] [Green Version]
  149. Subramaniam, S.; Donevan, S.D.; Rogawski, M.A. Block of the N-methyl-D-aspartate receptor by remacemide and its des-glycine metabolite. J. Pharmacol. Exp. Ther. 1996, 276, 161–168. [Google Scholar]
  150. Davies, J.A. Remacemide hydrochloride: A novel antiepileptic agent. Gen. Pharmacol. 1997, 28, 499–502. [Google Scholar] [CrossRef]
  151. Devinsky, O.; Vazquez, B.; Faught, E.; Leppik, I.E.; Pellock, J.M.; Schachter, S.; Alderfer, V.; Holdich, T.A. A double-blind, placebo-controlled study of remacemide hydrochloride in patients with refractory epilepsy following pre-surgical assessment. Seizure 2002, 11, 371–376. [Google Scholar] [CrossRef] [PubMed]
  152. Chadwick, D.W.; Betts, T.A.; Boddie, H.G.; Crawford, P.M.; Lindstrom, P.; Newman, P.K.; Soryal, I.; Wroe, S.; Holdich, T.A. Remacemide hydrochloride as an add-on therapy in epilepsy: A randomized, placebo-controlled trial of three dose levels (300, 600 and 1200 mg/day) in a Q.I.D. regimen. Seizure 2002, 11, 114–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Jones, M.W.; Blume, W.T.; Guberman, A.; Lee, M.A.; Pillay, N.; Weaver, D.F.; Veloso, F.; Holdich, T.A. Remacemide hydrochloride as an add-on therapy in epilepsy: A randomized, placebo-controlled trial of three dose levels (300, 600 and 800 mg/day) in a B.I.D. regimen. Seizure 2002, 11, 104–113. [Google Scholar] [CrossRef] [Green Version]
  154. Brodie, M.J.; Wroe, S.J.; Dean, A.D.; Holdich, T.A.; Whitehead, J.; Stevens, J.W. Efficacy and Safety of Remacemide versus Carbamazepine in Newly Diagnosed Epilepsy: Comparison by Sequential Analysis. Epilepsy Behav. 2002, 3, 140–146. [Google Scholar] [CrossRef]
  155. Wesnes, K.A.; Edgar, C.; Dean, A.D.P.; Wroe, S.J. The cognitive and psychomotor effects of remacemide and carbamazepine in newly diagnosed epilepsy. Epilepsy Behav. 2009, 14, 522–528. [Google Scholar] [CrossRef]
  156. Bialer, M.; Johannessen, S.I.; Kupferberg, H.J.; Levy, R.H.; Loiseau, P.; Perucca, E. Progress report on new antiepileptic drugs: A summary of the fourth Eilat conference (EILAT IV). Epilepsy Res. 1999, 34, 1–41. [Google Scholar] [CrossRef]
  157. Mizoule, J.; Meldrum, B.; Mazadier, M.; Croucher, M.; Ollat, C.; Uzan, A.; Legrand, J.J.; Gueremy, C.; Le Fur, G. 2-Amino-6-trifluoromethoxy benzothiazole, a possible antagonist of excitatory amino acid neurotransmission—I: Anticonvulsant properties. Neuropharmacology 1985, 24, 767–773. [Google Scholar] [CrossRef]
  158. 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]
  159. 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]
  160. 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]
  161. 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] [PubMed]
  162. Rothan, H.A.; Amini, E.; Faraj, F.L.; Golpich, M.; Teoh, T.C.; Gholami, K.; Yusof, R. NMDA receptor antagonism with novel indolyl, 2-(1,1-Dimethyl-1,3-dihydro-benzo[e]indol-2-ylidene)-malonaldehyde, reduces seizures duration in a rat model of epilepsy. Sci. Rep. 2017, 7, 45540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Erickson, J.D. Functional identification of activity-regulated, high-affinity glutamine transport in hippocampal neurons inhibited by riluzole. J. Neurochem. 2017, 142, 29–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Diao, L.; Hellier, J.L.; Uskert-Newsom, J.; Williams, P.A.; Staley, K.J.; Yee, A.S. Diphenytoin, riluzole and lidocaine: Three sodium channel blockers, with different mechanisms of action, decrease hippocampal epileptiform activity. Neuropharmacology 2013, 73, 48–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Minabe, Y.; Emori, K.; Shibata, R.; Kurachi, M. Antiepileptic effects of MK-801, a noncompetitive NMDA-receptor antagonist, in the low-frequency kindling model of epilepsy. Jpn. J. Psychiatry Neurol. 1992, 46, 755–761. [Google Scholar] [CrossRef]
  166. Song, X.; Jensen, M.; Jogini, V.; Stein, R.A.; Lee, C.H.; McHaourab, H.S.; Shaw, D.E.; Gouaux, E. Mechanism of NMDA receptor channel block by MK-801 and memantine. Nature 2018, 556, 515–519. [Google Scholar] [CrossRef]
  167. Dravid, S.M.; Erreger, K.; Yuan, H.; Nicholson, K.; Le, P.; Lyuboslavsky, P.; Almonte, A.; Murray, E.; Mosely, C.; Barber, J.; et al. Subunit-specific mechanisms and proton sensitivity of NMDA receptor channel block. J. Physiol. 2007, 581, 107–128. [Google Scholar] [CrossRef]
  168. Xi, D.; Zhang, W.; Wang, H.-X.; Stradtman, G.G., III; Gao, W.-J. Dizocilpine (MK-801) induces distinct changes of N-methyl-d-aspartic acid receptor subunits in parvalbumin-containing interneurons in young adult rat prefrontal cortex. Int. J. Neuropsychopharmacol. 2009, 12, 1395–1408. [Google Scholar] [CrossRef] [Green Version]
  169. Sato, K.; Morimoto, K.; Okamoto, M. Anticonvulsant action of a non-competitive antagonist of NMDA receptors (MK-801) in the kindling model of epilepsy. Brain Res. 1988, 463, 12–20. [Google Scholar] [CrossRef]
  170. Kubova, H.; Mares, P. Effects of MK-801 (dizocilpine) and ketamine on strychnine-induced convulsions in rats: Comparison with benzodiazepines and standard anticonvulsants. Physiol. Res. 1994, 43, 313–320. [Google Scholar]
  171. Parsons, C.; Quack, G.; Bresink, I.; Baran, L.; Przegalinski, E.; Kostowski, W.; Krzascik, P.; Hartmann, S.; Danysz, W. Comparison of the potency, kinetics and voltage-dependency of a series of uncompetitive NMDA receptor antagonists in vitro with anticonvulsive and motor impairment activity in vivo. Neuropharmacology 1995, 34, 1239–1258. [Google Scholar] [CrossRef]
  172. NIa, L.; Rukoiatkina, N.; Gorbunova, L.; Gmiro, V.; Magazanik, L. Role of NMDA and AMPA glutamate receptors in the mechanism of korazol-induced convulsions in mice. Ross. Fiziol. Zhurnal Im. IM Sechenova 2003, 89, 292–301. [Google Scholar]
  173. Brackett, R.L.; Pouw, B.; Blyden, J.F.; Nour, M.; Matsumoto, R.R. Prevention of cocaine-induced convulsions and lethality in mice: Effectiveness of targeting different sites on the NMDA receptor complex. Neuropharmacology 2000, 39, 407–418. [Google Scholar] [CrossRef]
  174. Kulkarni, S.K.; Ticku, M.K. Interaction between GABAergic anticonvulsants and the NMDA receptor antagonist MK 801 against MES-and picrotoxin-induced convulsions in rats. Life Sci. 1989, 44, 1317–1323. [Google Scholar] [CrossRef]
  175. O’Neill, S.K.; Bolger, G.T. Anticonvulsant activity of MK-801 and nimodipine alone and in combination against pentylenetetrazole and strychnine. Pharmacol. Biochem. Behav. 1989, 32, 595–600. [Google Scholar] [CrossRef]
  176. Vezzani, A.; Serafini, R.; Stasi, M.; Caccia, S.; Conti, I.; Tridico, R.; Samanin, R. Kinetics of MK-801 and its effect on quinolinic acid-induced seizures and neurotoxicity in rats. J. Pharmacol. Exp. Ther. 1989, 249, 278–283. [Google Scholar]
  177. Itzhak, Y.; Stein, I. Sensitization to the toxic effects of cocaine in mice is associated with the regulation of N-methyl-D-aspartate receptors in the cortex. J. Pharmacol. Exp. Ther. 1992, 262, 464–470. [Google Scholar]
  178. Tetz, L.M.; Rezk, P.E.; Ratcliffe, R.H.; Gordon, R.K.; Steele, K.E.; Nambiar, M.P. Development of a rat pilocarpine model of seizure/status epilepticus that mimics chemical warfare nerve agent exposure. Toxicol. Ind. Health 2006, 22, 255–266. [Google Scholar] [CrossRef]
  179. Thorat, S.; Kulkarni, S. Antagonism of caffeine-induced convulsions by ethanol and dizocilpine (MK-801) in mice. Methods Find. Exp. Clin. Pharmacol. 1991, 13, 413–417. [Google Scholar]
  180. Wardley-Smith, B.; Wann, K. Effects of four drugs on 4-aminopyridine seizures: A comparison with their effects on HPNS. Undersea Biomed. Res. 1991, 18, 413–419. [Google Scholar]
  181. Chavko, M.; Braisted, J.; Harabin, A. Effect of MK-801 on seizures induced by exposure to hyperbaric oxygen: Comparison with AP-7. Toxicol. Appl. Pharmacol. 1998, 151, 222–228. [Google Scholar] [CrossRef] [PubMed]
  182. Maheshwari, A.; Nahm, W.; Noebels, J. Paradoxical proepileptic response to NMDA receptor blockade linked to cortical interneuron defect in stargazer mice. Front. Cell. Neurosci. 2013, 7, 156. [Google Scholar] [CrossRef] [PubMed]
  183. Shakarjian, M.P.; Ali, M.S.; Velíšková, J.; Stanton, P.K.; Heck, D.E.; Velíšek, L. Combined diazepam and MK-801 therapy provides synergistic protection from tetramethylenedisulfotetramine-induced tonic-clonic seizures and lethality in mice. Neurotoxicology 2015, 48, 100–108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Schultz, M.K.; Wright, L.K.M.; de Araujo Furtado, M.; Stone, M.F.; Moffett, M.C.; Kelley, N.R.; Bourne, A.R.; Lumeh, W.Z.; Schultz, C.R.; Schwartz, J.E.; et al. Caramiphen edisylate as adjunct to standard therapy attenuates soman-induced seizures and cognitive deficits in rats. Neurotoxicol. Teratol. 2014, 44, 89–104. [Google Scholar] [CrossRef] [PubMed]
  185. Niquet, J.; Lumley, L.; Baldwin, R.; Rossetti, F.; Schultz, M.; de Araujo Furtado, M.; Suchomelova, L.; Naylor, D.; Franco-Estrada, I.; Wasterlain, C.G. Early polytherapy for benzodiazepine-refractory status epilepticus. Epilepsy Behav. 2019, 101, 106367. [Google Scholar] [CrossRef] [Green Version]
  186. Kovacic, P.; Somanathan, R. Clinical physiology and mechanism of dizocilpine (MK-801): Electron transfer, radicals, redox metabolites and bioactivity. Oxid. Med. Cell. Longev. 2010, 3, 13–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Kreutzwiser, D.; Tawfic, Q.A. Expanding Role of NMDA Receptor Antagonists in the Management of Pain. CNS Drugs 2019, 33, 347–374. [Google Scholar] [CrossRef]
  188. Weinbroum, A.A.; Ben-Abraham, R. Dextromethorphan and dexmedetomidine: New agents for the control of perioperative pain. Eur. J. Surg. 2003, 167, 563–569. [Google Scholar] [CrossRef]
  189. Kimiskidis, V.K.; Mirtsou-Fidani, V.; Papaioannidou, P.G.; Niopas, I.; Georgiadis, G.; Constadinidis, T.C.; Kazis, A.D. A phase I clinical trial of dextromethorphan in intractable partial epilepsy. Methods Find. Exp. Clin. Pharmacol. 1999, 21, 673–678. [Google Scholar] [CrossRef]
  190. Smith-Hicks, C.L.; Gupta, S.; Ewen, J.B.; Hong, M.; Kratz, L.; Kelley, R.; Tierney, E.; Vaurio, R.; Bibat, G.; Sanyal, A.; et al. Randomized open-label trial of dextromethorphan in Rett syndrome. Neurology 2017, 89, 1684–1690. [Google Scholar] [CrossRef]
  191. Hollander, D.; Pradas, J.; Kaplan, R.; McLeod, H.L.; Evans, W.E.; Munsat, T.L. High-dose dextromethorphan in amyotrophic lateral sclerosis: Phase I safety and pharmacokinetic studies. Ann. Neurol. 1994, 36, 920–924. [Google Scholar] [CrossRef] [PubMed]
  192. Majlesi, N.; Lee, D.C.; Ali, S.S. Dextromethorphan abuse masquerading as a recurrent seizure disorder. Pediatr. Emerg. Care 2011, 27, 210–211. [Google Scholar] [CrossRef] [PubMed]
  193. Chenard, B.L.; Menniti, F.S. Antagonists selective for NMDA receptors containing the NR2B subunit. Curr. Pharm. Des. 1999, 5, 381–404. [Google Scholar]
  194. Pontecorvo, M.J.; Karbon, E.W.; Goode, S.; Clissold, D.B.; Borosky, S.A.; Patch, R.J.; Ferkany, J.W. Possible cerebroprotective and in vivo NMDA antagonist activities of sigma agents. Brain Res. Bull. 1991, 26, 461–465. [Google Scholar] [CrossRef]
  195. Singh, L.; Oles, R.; Vass, C.; Woodruff, G. A slow intravenous infusion of N-methyl-DL-aspartate as a seizure model in the mouse. J. Neurosci. Methods 1991, 37, 227–232. [Google Scholar] [CrossRef]
  196. Doyle, K.; Shaw, G. Investigation of the involvement of the N-methyl-D-aspartate receptor macrocomplex in the development of spermine-induced CNS excitation in vivo. Br. J. Pharmacol. 1996, 117, 1803. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Tsuda, M.; Suzuki, T.; Misawa, M. Age-related decrease in the antiseizure effect of ifenprodil against pentylenetetrazole in mice. Dev. Brain Res. 1997, 104, 201–204. [Google Scholar] [CrossRef]
  198. Maroso, M.; Balosso, S.; Ravizza, T.; Liu, J.; Aronica, E.; Iyer, A.M.; Rossetti, C.; Molteni, M.; Casalgrandi, M.; Manfredi, A.A. Toll-like receptor 4 and high-mobility group box-1 are involved in ictogenesis and can be targeted to reduce seizures. Nat. Med. 2010, 16, 413–419. [Google Scholar] [CrossRef]
  199. Mareš, P. Age and activation determines the anticonvulsant effect of ifenprodil in rats. Naunyn Schmiedebergs Arch. Pharmacol. 2014, 387, 753–761. [Google Scholar] [CrossRef]
  200. Wang, X.; He, X.; Li, T.; Shu, Y.; Qi, S.; Luan, G. Anti-epileptic effect of ifenprodil on neocortical pyramidal neurons in patients with malformations of cortical development. Exp. Ther. Med. 2017, 14, 5757–5766. [Google Scholar] [CrossRef] [Green Version]
  201. Gorlewicz, A.; Pijet, B.; Orlova, K.; Kaczmarek, L.; Knapska, E. Epileptiform GluN2B-driven excitation in hippocampus as a therapeutic target against temporal lobe epilepsy. Exp. Neurol. 2022, 354, 114087. [Google Scholar] [CrossRef] [PubMed]
  202. Zarnowski, T.; Kleinrok, Z.; Turski, W.; Czuczwar, S. The NMDA antagonist procyclidine, but not ifenprodil, enhances the protective efficacy of common antiepileptics against maximal electroshock-induced seizures in mice. J. Neural Transm. Gen. Sect. 1994, 97, 1–12. [Google Scholar] [CrossRef] [PubMed]
  203. Gholizadeh, R.; Abdolmaleki, Z.; Bahremand, T.; Ghasemi, M.; Gharghabi, M.; Dehpour, A.R. Involvement of N-Methyl-D-Aspartate Receptors in the Anticonvulsive Effects of Licofelone on Pentylenetetrazole-Induced Clonic Seizure in Mice. J. Epilepsy Res. 2021, 11, 14–21. [Google Scholar] [CrossRef] [PubMed]
  204. Hanson, J.E.; Ma, K.; Elstrott, J.; Weber, M.; Saillet, S.; Khan, A.S.; Simms, J.; Liu, B.; Kim, T.A.; Yu, G.-Q.; et al. GluN2A NMDA Receptor Enhancement Improves Brain Oscillations, Synchrony, and Cognitive Functions in Dravet Syndrome and Alzheimer’s Disease Models. Cell Rep. 2020, 30, 381–396.e4. [Google Scholar] [CrossRef] [PubMed]
  205. Mareš, P.; Kozlová, L.; Mikulecká, A.; Kubová, H. The GluN2B-Selective Antagonist Ro 25-6981 Is Effective against PTZ-Induced Seizures and Safe for Further Development in Infantile Rats. Pharmaceutics 2021, 13, 1482. [Google Scholar] [CrossRef]
  206. Mares, P.; Tsenov, G.; Kubova, H. Anticonvulsant Action of GluN2A-Preferring Antagonist PEAQX in Developing Rats. Pharmaceutics 2021, 13, 415. [Google Scholar] [CrossRef]
  207. Sveinbjornsdottir, S.; Sander, J.; Upton, D.; Thompson, P.; Patsalos, P.; Hirt, D.; Emre, M.; Lowe, D.; Duncan, J. The excitatory amino acid antagonist D-CPP-ene (SDZ EAA-494) in patients with epilepsy. Epilepsy Res. 1993, 16, 165–174. [Google Scholar] [CrossRef]
  208. Lipton, S.A. Failures and successes of NMDA receptor antagonists: Molecular basis for the use of open-channel blockers like memantine in the treatment of acute and chronic neurologic insults. NeuroRx 2004, 1, 101–110. [Google Scholar] [CrossRef]
  209. 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]
  210. Ikonomidou, C.; Bosch, F.; Miksa, M.; Bittigau, P.; Vöckler, J.; Dikranian, K.; Tenkova, T.I.; Stefovska, V.; Turski, L.; Olney, J.W. Blockade of NMDA receptors and apoptotic neurodegeneration in the developing brain. Science 1999, 283, 70–74. [Google Scholar] [CrossRef]
  211. Johnston, M.V. Neurotransmitters and vulnerability of the developing brain. Brain Dev. 1995, 17, 301–306. [Google Scholar] [CrossRef]
  212. Deutsch, S.I.; Mastropaolo, J.; Rosse, R.B. Neurodevelopmental consequences of early exposure to phencyclidine and related drugs. Clin. Neuropharmacol. 1998, 21, 320–332. [Google Scholar] [PubMed]
  213. Mondadori, C.; Weiskrantz, L.; Buerki, H.; Petschke, F.; Fagg, G.E. NMDA receptor antagonists can enhance or impair learning performance in animals. Exp. Brain Res. 1989, 75, 449–456. [Google Scholar] [CrossRef] [PubMed]
  214. Bye, C.M.; McDonald, R.J. A Specific Role of Hippocampal NMDA Receptors and Arc Protein in Rapid Encoding of Novel Environmental Representations and a More General Long-Term Consolidation Function. Front. Behav. Neurosci. 2019, 13, 8. [Google Scholar] [CrossRef] [Green Version]
  215. Castellano, C.; Cestari, V.; Ciamei, A. NMDA receptors and learning and memory processes. Curr. Drug Targets 2001, 2, 273–283. [Google Scholar] [CrossRef]
  216. Honey, G.D.; Honey, R.A.; O’Loughlin, C.; Sharar, S.R.; Kumaran, D.; Suckling, J.; Menon, D.K.; Sleator, C.; Bullmore, E.T.; Fletcher, P.C. Ketamine disrupts frontal and hippocampal contribution to encoding and retrieval of episodic memory: An fMRI study. Cereb. Cortex 2005, 15, 749–759. [Google Scholar] [CrossRef]
  217. Hadj Tahar, A.; Blanchet, P.J.; Doyon, J. Motor-learning impairment by amantadine in healthy volunteers. Neuropsychopharmacology 2004, 29, 187–194. [Google Scholar] [CrossRef] [Green Version]
  218. Rowland, L.M.; Astur, R.S.; Jung, R.E.; Bustillo, J.R.; Lauriello, J.; Yeo, R.A. Selective cognitive impairments associated with NMDA receptor blockade in humans. Neuropsychopharmacology 2005, 30, 633–639. [Google Scholar] [CrossRef]
  219. Yaksh, T.L.; Tozier, N.; Horais, K.A.; Malkmus, S.; Rathbun, M.; LaFranco, L.; Eisenach, J. Toxicology Profile of N -Methyl-d-aspartate Antagonists Delivered by Intrathecal Infusion in the Canine Model. Anesthesiology 2008, 108, 938–949. [Google Scholar] [CrossRef] [Green Version]
  220. Vranken, J.H.; Troost, D.; de Haan, P.; Pennings, F.A.; van der Vegt, M.H.; Dijkgraaf, M.G.; Hollmann, M.W. Severe toxic damage to the rabbit spinal cord after intrathecal administration of preservative-free S(+)-ketamine. Anesthesiology 2006, 105, 813–818. [Google Scholar] [CrossRef]
  221. Gomes, L.M.; Garcia, J.B.; Ribamar, J.S., Jr.; Nascimento, A.G. Neurotoxicity of subarachnoid preservative-free S(+)-ketamine in dogs. Pain Physician 2011, 14, 83–90. [Google Scholar] [PubMed]
  222. Karpinski, N.; Dunn, J.; Hansen, L.; Masliah, E. Subpial vacuolar myelopathy after intrathecal ketamine: Report of a case. Pain 1997, 73, 103–105. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of N-methyl-D-aspartate receptor (NMDAR) complex and binding sites for GluN1 and GluN2 for some NMDAR antagonists. Ifenprodil and eliprodil mainly bind to GluN2B subunit. Mg2+, dizocilpine (MK-801), ketamine, and memantine act as noncompetitive antagonists with binding sites inside the ion channel pore region.
Figure 1. Schematic representation of N-methyl-D-aspartate receptor (NMDAR) complex and binding sites for GluN1 and GluN2 for some NMDAR antagonists. Ifenprodil and eliprodil mainly bind to GluN2B subunit. Mg2+, dizocilpine (MK-801), ketamine, and memantine act as noncompetitive antagonists with binding sites inside the ion channel pore region.
Pharmaceuticals 15 01297 g001
Table 1. Anticonvulsive effects of newer NMDAR antagonists in pre-clinical studies.
Table 1. Anticonvulsive effects of newer NMDAR antagonists in pre-clinical studies.
SubstanceEffect on NMDARsSeizure ModelEffectRef.
GNE-0723Positive allosteric modulator of GluN2AMouse model of Dravet syndrome↓ Low-frequency oscillatory and epileptiform activities[204]
Ro 25-6981Selective GluN2B antagonistPTZ model in infantile (12-day-old, P12) and juvenile (25-day-old, P25) rats↓ PTZ-induced seizures in infantile, but not juvenile, rats[205]
PEAQXSelective GluN2A antagonistPTZ-induced generalized seizuresAge-dependent differences in anticonvulsant effects in PTZ-induced seizures and epilepsy after discharge[206]
DDBMBoth GluN1 and GluN2 antagonistRat ECS model of epilepsy↓ Seizure behaviors in rats[162]
DDBM, l indolyl, [2-(1,1-Dimethyl-1,3-dihydro-benzo[e]indol-2-ylidene)-malonaldehyde]; ECS, electroconvulsive shock; PTZ, pentylenetetrazole.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sivakumar, S.; Ghasemi, M.; Schachter, S.C. Targeting NMDA Receptor Complex in Management of Epilepsy. Pharmaceuticals 2022, 15, 1297. https://doi.org/10.3390/ph15101297

AMA Style

Sivakumar S, Ghasemi M, Schachter SC. Targeting NMDA Receptor Complex in Management of Epilepsy. Pharmaceuticals. 2022; 15(10):1297. https://doi.org/10.3390/ph15101297

Chicago/Turabian Style

Sivakumar, Shravan, Mehdi Ghasemi, and Steven C. Schachter. 2022. "Targeting NMDA Receptor Complex in Management of Epilepsy" Pharmaceuticals 15, no. 10: 1297. https://doi.org/10.3390/ph15101297

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