Next Article in Journal
Evaluating Drug Interactions between Ritonavir and Opioid Analgesics: Implications from Physiologically Based Pharmacokinetic Simulation
Previous Article in Journal
Concomitant Treatment with Doxycycline and Rifampicin in Balb/c Mice Infected with Brucella abortus 2308 Fails to Reduce Inflammation and Motor Disability
Previous Article in Special Issue
Use of Quantitative Electroencephalography to Inform Age- and Sex-Related Differences in NMDA Receptor Function Following MK-801 Administration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

NMDA Receptor Antagonists: Emerging Insights into Molecular Mechanisms and Clinical Applications in Neurological Disorders

by
Ayodeji Olatunde Egunlusi
1,* and
Jacques Joubert
2
1
Pharmaceutical Chemistry, Faculty of Pharmacy, Rhodes University, P.O. Box 94, Makhanda 6140, South Africa
2
Pharmaceutical Chemistry, School of Pharmacy, University of the Western Cape, Private Bag X17, Bellville 7535, South Africa
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2024, 17(5), 639; https://doi.org/10.3390/ph17050639
Submission received: 12 April 2024 / Revised: 4 May 2024 / Accepted: 13 May 2024 / Published: 15 May 2024
(This article belongs to the Special Issue Pharmacological Insight into NMDA Receptor Antagonists)

Abstract

:
Neurodegenerative disorders (NDs) include a range of chronic conditions characterized by progressive neuronal loss, leading to cognitive, motor, and behavioral impairments. Common examples include Alzheimer’s disease (AD) and Parkinson’s disease (PD). The global prevalence of NDs is on the rise, imposing significant economic and social burdens. Despite extensive research, the mechanisms underlying NDs remain incompletely understood, hampering the development of effective treatments. Excitotoxicity, particularly glutamate-mediated excitotoxicity, is a key pathological process implicated in NDs. Targeting the N-methyl-D-aspartate (NMDA) receptor, which plays a central role in excitotoxicity, holds therapeutic promise. However, challenges, such as blood–brain barrier penetration and adverse effects, such as extrapyramidal effects, have hindered the success of many NMDA receptor antagonists in clinical trials. This review explores the molecular mechanisms of NMDA receptor antagonists, emphasizing their structure, function, types, challenges, and future prospects in treating NDs. Despite extensive research on competitive and noncompetitive NMDA receptor antagonists, the quest for effective treatments still faces significant hurdles. This is partly because the same NMDA receptor that necessitates blockage under pathological conditions is also responsible for the normal physiological function of NMDA receptors. Allosteric modulation of NMDA receptors presents a potential alternative, with the GluN2B subunit emerging as a particularly attractive target due to its enrichment in presynaptic and extrasynaptic NMDA receptors, which are major contributors to excitotoxic-induced neuronal cell death. Despite their low side-effect profiles, selective GluN2B antagonists like ifenprodil and radiprodil have encountered obstacles such as poor bioavailability in clinical trials. Moreover, the selectivity of these antagonists is often relative, as they have been shown to bind to other GluN2 subunits, albeit minimally. Recent advancements in developing phenanthroic and naphthoic acid derivatives offer promise for enhanced GluN2B, GluN2A or GluN2C/GluN2D selectivity and improved pharmacodynamic properties. Additional challenges in NMDA receptor antagonist development include conflicting preclinical and clinical results, as well as the complexity of neurodegenerative disorders and poorly defined NMDA receptor subtypes. Although multifunctional agents targeting multiple degenerative processes are also being explored, clinical data are limited. Designing and developing selective GluN2B antagonists/modulators with polycyclic moieties and multitarget properties would be significant in addressing neurodegenerative disorders. However, advancements in understanding NMDA receptor structure and function, coupled with collaborative efforts in drug design, are imperative for realizing the therapeutic potential of these NMDA receptor antagonists/modulators.

1. Introduction

Neurodegenerative disorders (NDs) are chronic disorders characterized by the progressive loss of neuronal cells, leading to neuronal dysfunctions. These disorders are associated with a wide range of cognitive, behavioral and motor dysfunctions, including memory loss, dyskinesia, paralysis, lack of coordination, and dysphasia. These widely studied NDs, among others, include Alzheimer’s disease (AD), Parkinson’s disease (PD), Amyotrophic lateral sclerosis (ALS), Huntington’s disease (HD), Lewy body dementia, depression, and multiple sclerosis. Although they share similarities in pathology and molecular mechanisms, each ND exhibits distinct clinical, neurobiological, and pathological characteristics influenced by risk factors such as geographical variations, age, race, sex, pre-existing pathological conditions, and xenobiotics. Nevertheless, the mechanism of degeneration in each disease remains inadequately defined [1,2,3,4]. The most common forms of NDs are AD and PD. Currently, the estimated number of people worldwide suffering from dementia is 55 million. This figure is projected to rise to 78 million by 2030 and 139 million by 2050. Regarding Parkinson’s Disease (PD), approximately 3% of the global population aged over 65 years is affected [3,5]. At this rate, there is an expected substantial increase in the economic, financial, and social burden, which could have serious consequences for the overall quality of life, especially in developing countries [2,3]. In 2021, the USA spent an estimated sum of USD 355 billion and USD 52 billion on dementia and PD, respectively. Globally, the cost associated with AD is approximated at USD 1 trillion annually, and this amount is projected to increase in the future due to the high ageing population [6,7].
Over the years, numerous studies have identified molecular and cellular mechanisms, giving rise to several etiological hypotheses. However, the mechanism of degeneration is still poorly defined. Despite many therapeutic trials derived from these hypotheses, none has been successful, as current treatments only offer symptomatic relief without halting the degenerative process or regenerating the neurons [8]. The challenge could be attributed to the multifactorial nature of these NDs, as each disorder is a result of interrelated processes that include oxidative stress, excitotoxicity, neuroinflammation, genetic mutation, endoplasmic reticulum dysfunction, protein aggregation and mitochondrial dysfunction [9,10]. Since the majority of NDs are sporadic, excitotoxicity is prominent among the proposed degenerative mechanisms. Despite the development of several molecules to address some of these mechanisms of degeneration, many have failed primarily due to their inability to cross the blood–brain barrier (BBB), a tightly spaced network of blood arteries and endothelial cells that makes ND treatments extremely complex and challenging [11,12].
Glutamate, the most vital excitatory neurotransmitter in the central nervous system (CNS), plays a crucial role in regulating various metabolic pathways. Under physiological conditions, the concentration of glutamate within the synapse is carefully regulated and maintained through neuron–astrocyte interaction, ensuring a physiological concentration in the extracellular space [13,14]. This glutamate homeostasis together with ion homeostasis is essential for preserving normal glutaminergic brain functions, including synaptic formation and signaling, neuronal plasticity, neurotransmission, learning, memory, and ageing. However, in a diseased state, this homeostatic balance is compromised, leading to an increase in glutaminergic neurotransmission and dysfunction resulting in excitotoxicity. In excitotoxicity, excessive extracellular glutamate overactivates the N-methyl-D-aspartate (NMDA) receptor, causing a significant intracellular calcium overload. This overload triggers a cascade of events that eventually leads to neuronal cell death either by apoptosis or necrosis (Figure 1) [15,16,17,18,19,20]. This form of neuronal death occurs gradually over a long period and has been implicated in the physiopathology of the most common NDs, including AD, PD, ALS and HD, especially in their early phases [21]. As such, minimizing glutamate activity, either through the synaptic clearing of excess glutamate or modulating NMDA receptors, could be therapeutically beneficial in addressing excitotoxic-mediated neuronal cell death. While the former exists under physiological conditions through astrocyte–neuron interactions, it becomes compromised in the diseased state, as observed in most NDs. This makes antagonizing the NMDA receptor a compelling strategy in slowing or halting the degenerative process and relieving symptoms associated with NDs, especially where excitotoxic-mediated death is concerned. Several NMDA receptor antagonists have been explored, but only a few, including amantadine for PD, memantine for AD, and riluzole for ALS, have been successful in clinical trials. However, these successes are not without drawbacks, as they are associated with several side effects that hinder their adherence. Moreover, many NMDA receptor antagonists have failed in clinical trials due to undesirable extrapyramidal effects and pharmacokinetic challenges.
The review aims to comprehensively explore the molecular mechanism underlying NMDA receptor antagonists at their respective binding sites. Before delving into this, we will offer insight into the structure and functions of NMDA receptors. We will also categorize these antagonists/modulators based on their binding sites and highlight the associated side-effect profiles, particularly those that have impeded their development. Furthermore, we will propose new directions or avenues for investigating NMDA receptor antagonists as potential treatments for NDs. Understanding these concepts could pave the way for innovative strategies in the treatment of neurodegenerative disorders. Additionally, current challenges and future perspectives in the field of NMDA receptor antagonists will also be discussed.

2. Structure and Functions of NMDA Receptors

The NMDA receptor is one of the ionotropic glutamate receptors that carries out excitatory neurotransmission in the CNS [22]. Under resting conditions, the NMDA receptor, primarily located in the postsynaptic site of neurons, is blocked by Mg2+. However, upon activation by glutamate or postsynaptic depolarization, it becomes highly permeable to cations, predominantly calcium ions. The NMDA receptor is divided into three subunits: GluN1, GluN2 and GluN3 subunits. The GluN2 subunits are further divided into four subtypes (GluN2 A-D), while GluN3 is subdivided into two types (GluN3 A-B). Despite the distinct biochemical and biophysical properties exhibited by GluN1 and GluN2 subunits, their combination forms the traditional heterotetrameric NMDA receptor. The resulting complex or channel is composed of one or more of the GluN2 subtypes coupled with the GluN1 subtype. This receptor, upon binding with glycine or D-serine (co-agonists) and glutamate (agonist), respectively, is necessary for optimal NMDA receptor functions or activation. This dual agonism on the NMDA receptor for its cellular function, a distinct feature that sets it apart from other neurotransmitter receptors, remains a topic of debate. However, the consensus is that glutamate is responsible for triggering NMDA receptor activation, and glycine or serine is essential for controlling the level of receptor activity. While the presynaptic function of NMDA receptors mediates neurotransmitter release and long-term plasticity, activities at the postsynaptic part of neurons are responsible for its slow current and synaptic plasticity. The binding affinity of glycine or D-serine to the GluN1 subtype of NMDA receptor is contingent on the specific brain region [7,16,23,24,25,26,27,28,29,30]. In contrast, the NMDA receptor channel formed by GluN1/GluN3 subunits is less sensitive to Ca2+ influx, not readily influenced by Mg2+ block, and can only be activated by glycine alone [28]. Thus, this type of NMDA receptor complex is less involved in Ca2+-medicated responses and is likely to have minimal impact on excitotoxicity.
Interestingly, GluN1 and GluN2 subunits share a fundamental structural similarity with other glutamate-gated ion channels or ionotropic glutamate receptors. All ionotropic glutamate receptor structures are classified into domains, and each subunit polypeptide chain consists of an amino-terminal domain (ATD), a ligand-binding domain (LBD), a transmembrane domain (TMD), and a carboxy-terminal domain (CTD). While the ATD is mainly responsible for the assembly and regulation of subunits, CTD plays a major role in receptor transport and anchoring the receptor to other intracellular molecules, enabling optimal interaction. However, the distinction lies in the presence of asparagine residue within the second transmembrane domain (M2 loop) of the GluN1 and GluN2 subunits. This region is suggested to serve as the pore-forming part of the NMDA receptor subunit and may be responsible for the ion permeability of the channel. Other transmembranes associated with the ionotropic structure are three membrane-spanning helices identified as M1, M3, and M4 [23,24,28,30,31,32,33,34]. Another distinctive feature between the NMDA receptor and other ionotropic receptors is the proximity of GluN2-ATD to the LBD of a GluN1/GluN2C receptor complex, enabling the binding of a positive allosteric modulator [34]. In terms of localization in the CNS, the GluN1 subunit is expressed ubiquitously at every developmental stage. However, this is not the case for the GluN2 subunit, as some subtypes exhibit uneven distribution, especially in the adult stage. While the GluN2A subtype is expressed widely, GluN2B, GluN2C and GluN2D are predominantly expressed in the forebrain (cortex, striatum, and hippocampus), cerebellum and midbrain, respectively. With this differential distribution, there is a possibility of targeting specific GluN2 subtypes with scaffolds that address a particular neurodegenerative disease and exhibit a good side-effect profile [7,31,33,35]. However, GluN2A and GluN2B are the main functional ion channels in the CNS. This is due to the low probability of channel opening in GluN2C and GluN2D subtype receptors [36]. Therefore, selectively targeting GluN2A or GluN2B subunits could be significantly impactful in the development of potential therapeutical agents. However, one needs to be mindful of the potential adverse effects posed by these agents.
As illustrated in Figure 2, the binding of glutamate and/or glycine to the LBD of GluN2 or GluN1 subunits, respectively, which occurs during the transition of the NMDA receptor from a resting state to an active state, causes the LBD bi-lobe to close and subsequently pulls the LBD-TMB linkers. This results in the opening of the ion channel pore, leading to the influx of Ca2+ [28,30]. This NMDA receptor-mediated Ca2+ response is responsible for mediating long-term potentiation and synaptic plasticity, the cellular basis of learning and memory, and maintaining neuronal health. The physiological functions of NMDA receptors are determined by their subunit composition, the location of subunits within the CNS, and the developmental stages of the brain (from embryo to adult) [26,34]. Noteworthy, recent studies in mice have observed a reduction in the expression and functions of the NMDA receptor, and the diffusion of the NMDA receptor, especially the GluN2B subtype, to the dendritic spine, leading to the formation of extrasynapse. This phenomenon is associated with advanced ageing, and the activation of these extrasynaptic NMDA receptors is implicated in neuronal cell death and accelerated age-related cognitive decline [37]. As such, targeting the NMDA receptor has been suggested to be therapeutically relevant and useful in addressing various neurodegenerative disorders. So far, a great deal of these targets have been developed and explored [28].

3. Types and Molecular Mechanisms of NMDA Receptor Antagonists

Several studies have implicated excitotoxicity as a prominent factor in the pathogenesis of most neurodegenerative disorders. This makes exploring NMDA receptor antagonists a potentially viable therapeutic approach to addressing these disorders. A number of these NMDA antagonists have been explored and investigated as potential neuroprotective agents. Despite the extensive research in this area, there has been an uphill challenge in developing these antagonists into effective therapeutical tools due to the compromised physiological function of NMDA receptors, resulting in extrapyramidal side effects such as cognitive impairment, hallucination, and psychosis. Nevertheless, the distinct subunits feature in different parts of the brain, particularly the GluN2 subunit, offering hope for the design and development of subunit-type antagonists with acceptable side-effect profiles [38]. These antagonists are categorized based on their binding sites into competitive, non-competitive, and negative allosteric antagonists, and their NMDA receptor subunits/subtypes, pharmacological and side-effect profiles are highlighted in Table 1.

3.1. Competitive NMDA Receptor Antagonist

Understanding the molecular mechanism underlying neurodegenerative disorders led to the development of competitive NMDA receptor antagonists (Figure 3) such as D-CPP (D(-)3-(2-carboxypiperazin-4-yl)-propyl-l-phosphonic acid), D-CPP-ene (D(-)3-(2-carboxypiperazin-4-yl)-propenyl-l-phosphonic acid; Midafotel) D-AP5 (2-amino-5-phosphonopentanoic acid, D-AP7 (2-amino-7-phosphonoheptanoic acid), DCKA (5,7-dichlorokynurenic acid), CGP-78608 (2-amino-4-methyl-5-phosphono-3-pentenoate-l-ethyl ester, CGP-37849 (DL-(E)-2-amino-4-methyl-5-phosphono-3-pentenoic acid), CGS-19755 (l-cis-2-carboxypiperidine-4-yl)-propyl-l-phosphonic acid; Selfotel), SDZ-220-040 ((s)-α-amino-2,4,-dichloro-4-hydroxy-5-(phosphonomethyl)-[1,1-biphenyl]-3-propanoic acid), L689-560 (trans-2-carboxy-5,7-dichloro-4-phenylaminocarbonylamino-1,2,3,4-tetrahydroquinoline), L-701,324 (7-chloro-4-hydroxy-3(3-phenoxy)phenyl-2(H)quinoline), PPDA (2S, 3R)-1-(phenanthrene-2- carbonyl)p iperazine-2,3-dicarboxylic acid, NVP-AAMO77 ([[[(1S)-1-(4-bromophenyl)ethyl]amino] (1,2,3,4-tetrahydro-2,3-dioxo-5-quinoxalinyl)methyl]phosphonic acid tetrasodium salt; PEAQX) and ST3 (((S)-5-[(R)-2-amino-2-carboxyethyl]-1-[4-(3-fluoropropyl)phenyl]-4,5-dihydro-1H-pyrazole-3-carboxylic acid). These antagonists bind directly to the binding sites of glycine or glutamate at the GluN1 or GluN2 subunits, respectively. To understand the binding interaction of these antagonists, analysis of the cryo-EM structure of an intact NMDA receptor complex bound to antagonists or FRET analysis of the crystal structure of the LBD indicates an increase in the opening of the GluN1 and/or GluN2 clamshell by various degrees (13–28°) when compared to glycine or glutamate binding, respectively, occupying the same active binding site. Subsequently, these clamshell openings lead to the relaxation of the tension in the LBD-TMD linker, resulting in the closure of the ion channel pore (Figure 4) [28,39,40,41].
Functional NMDA receptor antagonists have been shown to exhibit anticonvulsant, anti-ischemic, antidepressant-like and anxiolytic-like properties [39,42,43]. Despite their strong activities in attenuating glutamate-medicated excitotoxicity, these antagonists have been marked by unfavorable side effects such as hallucination, agitation, confusion, paranoia, delirium, drowsiness, and coma. These adverse effects render them unsafe for human use, leading to their failure in clinical trials [44,45]. A notable example is D-CPP-ene, initially touted as a promising antiepileptic agent due to the absence of phencyclidine-like adverse effects observed at therapeutic doses in pre-clinical studies. It was also well tolerated in phase I clinical trials, with healthy volunteers exhibiting tolerance up to a dose of 2000 mg/day. However, D-CPP-ene was terminated at phase II due to severe adverse effects, including confusion, disorientation, gait ataxia and sedation, or worsened seizures noted at daily doses of 500–1000 mg/day [42]. Moreover, the majority of competitive NMDA antagonists except SDZ 220-581 permeate poorly across the BBB due to their hydrophilic nature [44,46,47]. Table 1 illustrates the pharmacological action and adverse effects associated with a few competitive NMDA antagonists that were investigated for neurological disorders. Despite a decade-long search for a competitive antagonist with a good safety profile and minimal side effects, none has completed clinical trials due to the associated psychotomimetic or dopaminergic transmission side effects. Generally, antagonists targeting GluN2 (A-D) are more prone to these unwanted adverse effects than the GluN1 subunits. While several non-selective NMDA or GluN2 antagonists have displayed psychotomimetic and/or dopaminergic side effects similar to MK-801, GluN1 antagonists have shown more favorable outcomes [48,49,50,51]. Therefore, it is suggested that GluN1 antagonists could potentially address various neurological disorders. However, available information is derived only from preclinical studies on disorders such as anxiety, depression, and epilepsy. Moreover, the clinical relevance of a GluN1 antagonist is highly debatable because of the wide distribution of GluN1 subunits in the central nervous system or the important functions of co-agonists in NMDA receptor physiological functions. Clinical studies for this class of antagonists are needed to confirm their therapeutic relevance. Noteworthy is the limited preclinical data for complex neurodegenerative disorders like AD and PD [28,39,41,42,46,52,53,54,55,56,57,58,59,60]. Preclinical and clinical studies are essential to bridge this knowledge gap and determine the potential efficacy of GluN1 antagonists in treating these neurodegenerative disorders.

3.2. Uncompetitive or Non-Competitive NMDA Receptor Antagonists

Decades ago, the failure of competitive NMDA receptors to effectively address neurological disorders clinically redirected research towards non-competitive NMDA receptor antagonists. During this period, the focus was on targeting this binding site to tackle neurodegenerative disorders such as depression [61], PD, and AD. The aim was to alleviate the undesirable adverse effects associated with competitive NMDA receptor antagonists. These non-competitive NMDA receptor antagonists, also known as channel blockers, include phencyclidine (PCP), dizocilpine maleate (MK-801), ketamine, and tiletamine (Figure 5). They act by binding with high affinity to the PCP binding site at the entrance of the channel gate, as shown in Figure 6, to block calcium-mediated responses. Consequently, they have demonstrated neuroprotective effects in conditions such as stroke, cardiac arrest, and neurodegenerative disorders. Moreover, they have been shown to display anti-dyskinetic and anticonvulsant effects, although variations may arise depending on the rodent strains or models employed [62,63,64,65,66,67]. These variations have led to ambiguous effects in the literature data on some of these open-channel blockers. For instance, one group found no anticonvulsant effects with ketamine in a 4-aminopyridine (4-AP) induced epileptic model of hippocampal slices, while another group was able to indicate its anticonvulsant properties in a 4-AP-induced seizure model of male Wistar rats [62]. The differences in effects could be attributed to the variation in NMDA receptor subunit complexes expressed by the cells or animals, influenced by age.
Despite their efficacy, due to their high affinity, they bind rapidly and dissociate slowly, prolonging the calcium-mediated response and resulting in unfavorable clinical outcomes [68]. Similar to most competitive NMDA receptor antagonists, they are known to induce adverse effects, such as neuropsychological, psychotomimetic, and dopaminergic transmission effects, which limit their clinical use. Some of these antagonists can induce schizophrenia-like symptoms, even in healthy volunteers. The dopaminergic transmission effects of non-competitive NMDA receptors stem from their ability to activate the dopaminergic system, subsequently increasing dopamine synthesis, release and metabolism in various parts of the brain [64,68,69]. Confounding the problem is the influence of these antagonists on the increase in turnover and release of serotonin, which are known to exacerbate schizophrenia-like symptoms [70].
The rediscovery of clinically tolerated memantine and amantadine as non-competitive NMDA receptor antagonists marked a significant breakthrough in the treatment of neurodegenerative disorders. Much like MK-801, both drugs bind to the PCP binding site of the ion channel complex. However, the efficacy and clinical use of amantadine and memantine as anti-dyskinesia (in PD) and a neuroprotective agent (in AD), respectively, are partially dependent on its weak NMDA receptor antagonist [66,71]. This clinical tolerability is attributed to the considerably shorter residence time within the channel in an open state when compared to MK-801. Interestingly, a study demonstrated the ineffectiveness of memantine at low-level NMDA receptor activation but found it to be highly efficacious in the overactivation state. This favorable kinetics makes these blockers better neuroprotective agents with minimal side-effect profiles when compared to MK-801 [71,72,73,74,75]. Additionally, both amantadine and memantine are known to attenuate epileptiform activity behaviors induced by 4-AP in a rat model. When compared to amantadine, memantine has displayed better therapeutical indices in the management of epilepsy and other neurological or psychological disorders [62].
Similar to competitive NMDA receptor antagonists, some non-competitive antagonists have been associated with undesirable adverse effects. These effects are not only attributed to their strong binding affinity to the PCP binding site but also to their promiscuity. For example, antagonists like ketamine are known to bind to other receptors, enhancing the activity or transmission of other neurotransmitters such as dopamine, serotonin, noradrenaline α-Amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA), acetylcholine, opioid, and gamma-aminobutyric acid (GABA) [76,77,78]. This can lead to unwanted side effects, including addiction, dependency, and tolerance, particularly with prolonged use.

3.3. Allosteric NMDA Receptor Antagonists

Despite numerous studies on competitive and uncompetitive NMDA receptor antagonists, addressing neurodegenerative disorders still poses enormous challenges, as current treatments only offer symptomatic relief. To date, no drug has successfully halted or slowed down the degenerative process. An alternative to competitive and noncompetitive antagonists is the allosteric modulation of NMDA receptors. The influence of GluN2 on the biophysical characteristics of NMDA channels provides a potential allosteric target [79,80,81]. For instance, the GluN2B subunit presents an attractive site for therapeutic interventions in chronic neurodegenerative diseases like PD, AD, ALS, HD, and multiple sclerosis, as well as in acute neuronal diseases like traumatic brain injury, epilepsy, and stroke. What makes this target particularly intriguing is its enrichment in perisynaptic and extrasynaptic NMDA receptors, major contributors to excitotoxic-induced neuronal cell death. However, antagonists targeting this subunit have failed in clinical trials due to their lack of benefit in PD patients. Like non-competitive antagonists, the binding of allosteric NMDA antagonists to their receptor is independent of the presence or absence of agonists (glutamate, glycine or serine) at the binding site [27,82,83,84,85,86,87,88,89,90,91,92,93,94]. Allosteric modulators (Figure 7) of NMDA receptor channels can be classified into positive and negative modulators, and several of them have been explored extensively. Both positive and negative modulators bind to the ATD part of the NMDA receptor channel to potentiate or block, respectively, Ca2+-mediated responses. As shown in Figure 8, modulators like ifenprodil, a selective GluN2B inhibitor, act by binding to the interface between the GluN1 and GluN2B ATDs, and the mobility of the GluN2 lower lobe is vital for NMDA receptor inhibition [34,80,88,89,95,96,97,98]. This was confirmed by a cross-linking study that indicated a decrease in distance between the lobes of GluN2 ATD for the ifenprodil-bound NMDA receptor and immobile GluN (1 and 2) ATD upper lobes, despite conformation changes within the ATDs [80]. Similar to ifenprodil, are radiprodil and Ro 25-6918, which have been shown to selectively inhibit GluN2 subunits to antagonize NMDA receptors. These selective GluN2B antagonists are efficacious with reduced side effects against a few neurodegenerative disorders. However, clinical trials for ifenprodil and radiprodil were terminated early due to poor bioavailability and lack of recruitment, respectively. For instance, radiprodil was initially investigated for infantile spasm syndrome due to its stronger anticonvulsant effects in younger rat pups compared to adult animals. However, it was terminated in the early stage of phase 2 clinical trials because of challenges in recruiting infant patients within the prescribed timeframe [89,99,100]. Ifenprodil was investigated as an adjunct therapy in PD patients with waning efficacy of levodopa but failed in phase 2 clinical trials as the drug failed to reduce tremor, rigidity and bradykinesia due to poor BBB permeability [101]. However, it found success in treating cerebral ischemic disease and is currently marketed in Japan and France as Cerocral®, a cerebral vasodilator [102].
Other negative allosteric modulators include DQP-1105 which has been shown to selectively block the GluN2D subtype to regulate synaptic transmission in the subthalamic nucleus, substantia nigra, striatum and spinal cord, and the GluN2C to presynaptically modulate gabaminergic synaptic transmission in the suprachiasmatic nucleus [103]. TCN-201 (sulfonamide derivative) was also identified as a promising selective GluN2A allosteric modulator but could not proceed to biological studies due to its poor solubility [27,82,96,104]. However, analogues of TCN-201 like MPX-004 and MPX-007 with enhanced solubility and increased efficacy have been designed and explored. These analogues could provide opportunities to further understand the mechanism of GluN2A NMDA receptor allosteric modulators [27,30].
Regrettably, the majority of active drugs targeting allosteric sites have not produced the desired therapeutic success in clinical trials. However, recent developments have identified a series of phenanthroic and naphthoic acid derivatives with enhanced GluN2B, GluN2A or GluN2C/GluN2D selectivity, depending on their functional moieties, and improved pharmacodynamic properties [92,96,104,105,106]. Additionally, these derivatives are amphipathic, possessing both hydrophobic and charged moieties. This characteristic promotes BBB permeation via the neurosteroid transporter, suggesting a good bioavailability for this group of compounds, particularly in addressing NDs [92,96].
Table 1. Pharmacological and side-effect profiles of developed NMDA receptor antagonists.
Table 1. Pharmacological and side-effect profiles of developed NMDA receptor antagonists.
Antagonist TypeCompoundsReceptor Subunits/SubtypesDevelopmental StagePharmacological ProfilesSide-Effect ProfilesReferences
Competitive antagonists D-CPP/D-CPP-ene (Midafotel)GluN2ATerminated at Phase 11 clinical trialsAntiepileptic and neuroprotective effects against head injury, cerebral ischemia and stroke
Alteration of acute behavioural response to cocaine.
Stimulate short-term increase in NREM (non-rapid eye movement) sleep
Hallucinations
Poor concentration Confusion
Gait ataxia
Sedation Depression
[28,39,42,54,55,68,107]
D-AP5/D-AP7Non-subunit selective Preclinical or experimental studiesBlock fear acquisition and expression
Block or interfere with acute response to psychostimulants such as cocaine amphetamine, or methamphetamine
Similar to those of D-CPP-ene/D-CPP[28,55]
DCKAGluN1Preclinical or experimental studiesAnxiolytic effect
Neuroprotective against NMDA/glycine-induced toxicity
Lack of psychotomimetic effects or side effects associated with dopaminergic transmission [28,56,108,109,110]
CGP-78608, CGP-37849 & CGP-40116GluN1Preclinical or Experimental studiesAnticonvulsant effectLack of side effects associated with dopaminergic transmission[28,45,64,68,111]
CGS-19755 (Selfotel)GluN2ATerminated at Phase III clinical trialsNeuroprotective effect against global and focal ischemia, trauma and strokePsychotomimetic side effects like Hallucination
Confusion
Paranoia
Delirium
Lack of side effects associated with dopaminergic transmission
[28,44,46,52,53]
L689-560 & L701-324GluN1Preclinical or experimental studiesAnticonvulsant effects
Anxiolytic
Antidepressant-like effect in mice
Sedation
Lack of neuronal vacuolisation and psychotomimetic potential
Ataxia at a high dose
Modest impairment of reference memory, but no negative effect on working memory
[28,49,56,112,113,114,115,116]
PPDAGluN2A, GluN2C &GluN2DPreclinical or experimental studiesPrevent the complete worsening effect of tissue-type plasminogen activator on NMDA-induced neuronal death in both cultured cortical and hippocampal neurons
Anti-allynic and anti-hyperalgesic effects in rat
Motor dysfunction at high dose[28,41,58,60,117,118,119,120,121]
NVP-AAMO77 (PEAQX)GluN2A
GluN2C & GluN2B
Preclinical or experimental studiesProduce anti-compulsive behaviour in a rat model
Impairment of contextual and temporal fear responses
Antidepressant-like effect in rodents
Affect motor coordination stamina and motivation run in a rat dyskinesia model
Motor memory impairment or learning memory deficit
[28,36,117,119,122,123,124,125,126]
SDZ-220-040GluN2BPreclinical or experimental studiesDesign to readily cross the BBB. Effectively disrupt prepulse inhibition in rats
Anticonvulsant effect
Protection against focal ischemia
Attenuate neuropathic pain
Sedation
Ataxia Psychotomimetic effects
[28,127,128,129]
Non-competitive antagonistMK-801Open-Channel blockerPreclinical or experimental studiesReverse mild stress-induced anhedonia in male Wistar rats
Neuroprotective effect in several animal models of cerebral ischaemia
Block L-Dopa-induced dyskinesia in a rat preclinical model, but only at concentrations that worsen parkinsonism
Anti-convulsant effect
Weight loss
Hypothermia
Death
Hallucination
Ataxia
Hyperlocomotion
[53,61,63,66,130,131]
MemantineOpen-channel blockerApproved for AD in humanNeuroprotective effect in AD, vascular dementia and prodromal stages of psychosis
Antidepressant-like effect
Antinociceptive effect in rats
Anticonvulsant effect
Occasional restlessness
Headache Hypertension
Drowsiness Constipation Diarrhoea
Nausea
Anorexia
Dyspnea
Slight dizziness at a high dose
[45,62,71,72,75,132,133,134]
AmantadineOpen-channel blockerApproved for PD in humanAnti-dyskinetic effect
Effectively reduce L-Dopa-induced abnormal involuntary movement
Anti-convulsant effect
Neuroprotective effect
Visual hallucination
Confusion
Blurred vision
Leg oedema
Dry mouth
Constipation
Urinary retention
[62,63,66,131,135]
PCPOpen-channel blockerPreclinical/experimental studiesAnticonvulsant effects in NMDA- or quinolate-induced seizure model
Anaesthetic effect
Hallucination
Ataxia
Hyperlocomotion
Emergency delirium
[45,136]
KetamineOpen-channel blockerApproved as an anaesthetic agentAnticonvulsant effect in NMDA- or quinolate-induced seizure model
Anaesthetic effect
Antidepressant effect in resistant major depressive disorder
Induce cognitive deficits and psychotic symptoms
Hallucination
Abuse
Psychological and physiological dependences
Possible neurotoxicity
Nystagmus
Drowsiness
Nausea and vomiting
Blood pressure elevation
Liver and bladder damage
[45,130,132,133,136,137,138,139]
TiletamineChannel blockerApproved for veterinary useAnaesthetic effect
Anticonvulsant effects in NMDA- or quinolate-induced seizure model
Robust Sedation in human and animal
Ataxia
Feeling of dissociation
Hallucination
[136]
Negative allosteric modulatorIfenprodilGluN2BPhase III clinical trials completedNeuroprotective effect in both in vitro and in vivo models of cerebral ischemia
Anticonvulsant effects in rodent
Rapid antidepressant effect
Alleviate neuropathic pain
Impair cognitive behavioural tasks[28,81,85,93,100,134,140,141,142]
RadiprodilGluN2BTerminated at Phase II clinical trialsAnticonvulsant effect in rate model (stronger in young rat pups than adult animals)
Decrease epileptic spasms in infants
Vomiting
Pyrexia
[100]
Ro25-6981GluN2BPreclinical or experimental studiesRapid antidepressant effect and counteract depressive-like behaviour in chronically stressed rodent
Neuroprotective effect against glutamate-induced toxicity in a cultured cortical neuron
Improve anxiety and compulsive behaviour in obsessive-compulsive disorder rat
Alleviate cerebral ischemia-reperfusion and oxidative damage in male Sprague Dawley rats
Antipakinsonian effect in 6-OHDA-lesioned and MPTP PD rat model
Reduced memory in early life stress mice[28,88,143,144,145,146]
DQP-1105GluN2C & GluN2DPreclinical or experimental studiesNeuroprotective effects in GluN2D-rich substantia nigra compacta dopaminergic neuronsMotor dysfunction[28,103,147,148,149,150,151]

4. Current Challenges and Future Perspectives

Over the years, translating preclinical studies of NMDA receptor antagonists into successful clinical drugs has been an uphill task. This is mainly due to the complexity of the NMDA receptors, as their optimum function is needed for several brain physiological functions. Deviation in the form of hypofunction or hyperfunction is detrimental to general well-being [30]. Despite decades of studies, the use of NMDA receptor antagonists to address these defects in many neurodegenerative disorders while still maintaining the optimum physiological function of the NMDA receptor has yet to produce the desired outcome. Not only are there conflicting results in preclinical studies, but therapeutic effects observed in animal studies have failed to translate into human studies. This suggests a limited understanding of the NMDA receptor’s physiological functions in human brains. Confounding this issue is the presence of the heterodimeric (GluN1/GluN2B or GluN1/GluN2D) heterotrimeric (GluN1/GluN2B/GluN2D) structure of NMDA receptors in certain neurons, which significantly increases the difficulty in resolving the function of different NMDA receptor subtypes [96,150]. Moreover, several of these antagonists are marked by undesirable adverse effects like psychotomimetic, dopaminergic transmission, or schizophrenia-like effects. These adverse effects are linked to either the competitive inhibition of endogenous excitatory neurotransmitters, the strong binding affinities of these antagonists, the activities at other neurotransmitter receptors, or the inability of NMDA receptors to metamodulate other neurotransmitter receptors [25]. Even though clinically tolerated NMDA antagonists (memantine and amantadine) are available, they are not without side effects. Interestingly, a few selective allosteric modulators (ifenprodil) targeting NMDA receptors have shown promising results with good side-effect profiles when compared to competitive and noncompetitive NMDA receptor antagonists. However, they are often poorly bioavailable, and their selectivity to a particular NMDA receptor subtype or subunit is relative. Not only are they difficult to study despite being knowledgeable about the binding sites [88], but they are also known to interact with other neurotransmitter receptors. For instance, ifendropil, a selective GluN2B negative allosteric modulator, has been shown to also act at adrenoceptor 1 & 2 (α1&2), sigma and (serotonin) 5-TH1A&2 receptors [85,90,152]. Similar GluN2B antagonists. like Ro-25698, with a low crossreactivity with adrenoceptors, are known to bind to 5-HT, histamine-1 (H1) and sigma receptors [152]. Thus, they can cause unwanted side effects. Nevertheless, the significance of the GluN2B subunit, particularly those at the extrasynaptic sites, in excitotoxic-induced cell death makes them a favorable therapeutic target for any potential modulators. Modulators with enhanced solubility and bioavailability must be specific and selective to the GluN2B subtype. To achieve this, a better understanding of the mechanism of allosteric binding of antagonists or modulators to NMDA subunits is needed, but remains poorly defined to date.
To improve the solubility and bioavailability of these GluN2B antagonists, researchers could explore nano-drug formulations, focusing on solid lipid nanoparticles. This form of drug delivery system would involve incorporating the drug molecules in stearic acid/poloxamer188 nanoparticles and coating them with chitosan. These nanoparticles, with an average size of 300 nm, through intranasal administration, have been shown to effectively cross the BBB in brain endothelial cell permeation and uptake studies. Moreover, the cationic nature of chitosan may aid adhesion to brain endothelial cells, improving the transport of nanoparticles. Additionally, the nose–brain delivery system provides an enriched blood vessel that enhances the bioavailabilities of loaded drugs. Similar drug delivery systems have been used to improve the BBB permeability of dopamine and riluzole, with the potential to treat PD and ALS, respectively [153,154,155]. Surprisingly, to date, no studies have explored or reported on this form of drug delivery system for ifenprodil and its derivatives, which were once touted as promising drug agents for the treatment of NDs. Exploring such an avenue could provide the much-needed breakthrough for this class of antagonists in halting or slowing the degenerative processes mediated by glutamate-induced toxicity.
Recently, fluoroethylnormemantine (FNM), a novel NMDA receptor antagonist derived from memantine, has been developed for the treatment of stress-induced maladaptive behavior associated with depression. When compared to (R, S)-ketamine, FNM has been shown to exert rapid antidepressant actions with a low side-effect profile in mice by selectively antagonizing NMDA receptors [156]. Similar to memantine, FNM binds in a non-competitive manner and an open active state to the PCP site of the NMDA receptor as observed in radioligand binding studies ([18F]-FNM) [157]. FNM is currently in phase 1 clinical trials, offering hope for the treatment of neurodegenerative disorders, including post-traumatic stress disorder, AD, major depressive disorder and treatment-resistant depression [158]. Another promising rapid antidepressant agent is esmethadone (REL-1017), a dextro isomer of methadone with little or no activity towards the opioid receptor. Like memantine, esmethadone is a low-affinity non-competitive NMDA receptor antagonist. Esmethadone is currently in a phase 3 clinical trial for the treatment of major depressive disorder [159]. The development of these antagonists (Figure 9) emphasizes the prominent effect of excitotoxicity in neurodegenerative disorders and redefines polycyclic cage structures with NMDA receptor selectivity and fast kinetic interaction. Moreover, these polycyclic cages are permeable to the BBB with a minimal side-effect profile. This offers the opportunity to explore more polycyclic cages acting at the PCP binding site in an open active NMDA receptor state. Several structural-related polycyclic cages have been shown to display neuroprotective effects against glutamate-induced toxicity and other degenerative processes [160]. However, the reported findings are based solely on experimental data, and further exploration through clinical studies on these groups of NMDA receptor antagonists is warranted.
The multifaceted nature of many neurodegenerative disorders makes designing and developing potential treatments complex and highly challenging. Factors contributing to the degenerative process are interrelated, including excitotoxicity, oxidative stress, neuroinflammation, protein aggregation, and mitochondria dysfunction. The majority of NMDA receptor antagonists are designed to target excitotoxicity, corresponding to a single-target approach. Moreover, a few of these antagonists fail to cross the BBB due to their hydrophilic nature. In recent years, the focus has been on designing and developing multifunctional agents with the potential to address glutamate-induced toxicity and other therapeutic targets, adopting a multi-target approach to drug development. These antagonists, sharing functional and structural similarities to amantadine and memantine, exhibit polycyclic cage structures that incur lipophilicity and enhanced permeation via the BBB. For example, a series of triazole-bridged aryl adamantane derivatives have been explored as multifunctional agents for the potential treatment of AD. These derivatives demonstrated potent inhibition of acetylcholinesterase enzymes, Aβ aggregation, and NMDA receptor, as well as good BBB permeability and a good safety profile in neuronal cell lines like SH-SY5Y neuroblastoma cells. This unique attribute makes them promising candidates for the treatment of AD [161]. Similarly, a series of polycyclic propargylamine and acetylene derivatives were investigated, revealing multifunctional activities, including neuroprotection, monoamine oxidase (MAO) inhibition, anti-apoptotic activities, and inhibition of NMDA receptors and voltage-gated calcium channels [162]. Another study explored some carbamate-based cholinesterase inhibitors, with structural similarities to acetylcholine, as potential multifunctional agents for AD treatments. These inhibitors exhibit diverse scaffolds, such as physostigmine, isosorbide, quinazoline, quinoline, xanthone, chalcone, flavonoid, indole-like, resveratrol and coumarin derivatives. In addition to their anti-cholinesterase activity, preclinical findings suggest multiple activities, including antioxidant properties, anti-neuroinflammation, metal chelation, neuroprotection, monoamine oxidase inhibition, neurotrophic effect and/or reduction in Aβ aggregation. Hence, they represent promising multifunctional candidates for the treatment of AD [163,164]. However, the majority of the available data for these inhibitors or antagonists stems from preclinical or experimental studies. There is a pressing need for in vivo and clinical studies to establish the therapeutical clinical efficacy of these groups of compounds.
Currently, only two FDA-approved multifunctional drugs (Namzaric® and Auvelity®) are available, each containing an NMDA receptor antagonist and another therapeutic agent, for the treatment of neurodegenerative disorders. Namzaric® is marketed for moderate to severe AD, while Auvelity® is designated for the treatment of agitation associated with AD. Despite the enhanced activity and formulation adherence of each drug over its components, they have demonstrated similar side-effect profiles to their parent drugs [165,166,167]. Therefore, there is a need for a multifunctional hybrid or hybrids capable of antagonizing NMDA receptors, providing symptomatic relief, and targeting other degenerative processes such as neuroinflammation, oxidation and mitochondrial dysfunctions. Designing and developing selective GluN2B antagonists/modulators with polycyclic moieties and multitarget properties would be highly desirable. Such a multifaceted approach, with polycyclic scaffolds that incur good bioavailability, holds significant promise in addressing neurodegenerative disorders [168].

5. Conclusions

The significance of glutamate-induced excitotoxic death in the pathogenesis of neurodegenerative disorders is well established. With this knowledge, the ideal approach would be to use NMDA receptor antagonists to halt the degenerative process. Despite years of research, developing such agents has yielded little success, as current treatments, such as amantadine and memantine, only offer symptomatic relief. Many competitive and noncompetitive NMDA receptor antagonists have been explored but are marked by undesirable psychotomimetic side effects. These adverse effects are linked to the strong NMDA receptor-binding affinity or metamodulation of NMDA receptors by these antagonists that negatively influence their physiological functions. Interestingly, a more detailed exploration of the structure and function of NMDA receptors has led to the development of some selective negative allosteric modulators with good side-effect profiles. These modulators offer promise in addressing neurodegenerative disorders.
The distinctive biophysical features and localization of NMDA subunits provide a great opportunity for developing clinically effective drugs with optimum safety profiles. For instance, extrasynaptic neurons are rich in the GluN2B subunit and serve as key mediators of excitotoxic neuronal cell death. Therefore, selectively blocking this subunit would be therapeutically beneficial in addressing glutamate-induced cell death, especially in conditions such as AD and PD, where neurons in the brain cortex, hypothalamus, or striatum are predominantly affected. However, some of the developed antagonists fail in clinical trials due to poor bioavailability or lack of recruitment. One could explore the use of nano-drug formulations like solid lipid nanoparticles and nanostructured lipid carriers to improve the BBB permeability of these antagonists. Moreover, their selectivities toward the different GluN2 subtypes are often relative. This is majorly due to the amino acid sequence similarities existing among GluN2 subunits. For example, the overall amino acid sequence homology of GluN2A and GluN2B subunits are nearly identical, making it difficult to determine the strong subunit specificity of certain ligands [148,169]. Designing and developing agents that specifically and selectively target only GluN2B subunits is crucial for overcoming these challenges and improving the success rate in clinical trials [84]. This could be achieved by utilizing a receptor-based virtual screening method to identify amino acid residues that are unique to GluN2B subunits. Subsequently, a structure-based virtual screening technique would be employed to identify small molecules with optimum binding interactions with these targets. This targeted approach holds promise for the development of novel NMDA receptor antagonists that effectively address glutamate-induced excitotoxicity with minimal side effects.
However, the challenge is the subunit diversity in the NMDA receptor channel complex, which is highly complex and not completely understood, despite decade-long x-ray crystallography studies on these subunits [148,170]. One of the most intriguing and challenging aspects of studies involving certain NMDA receptor antagonists is the discrepancy between preclinical and clinical findings, ultimately resulting in clinical trial failures. This may, in part, be linked to the differences in LBD residues that exist between human and animal (rodent) NMDA receptor subunits. Additionally, the expression of these subunits may differ at each development stage of the animal as observed in rodents, leading to translational failure [148]. This highlights our limited understanding of the NMDA receptor structure and function, particularly in humans, which remains poorly defined to date. Further research and refinement in drug design and understanding the molecular structure and functions of the NMDA receptor are crucial to fully unlock the therapeutic potential of this strategy in treating neurodegenerative disorders. Elucidating the structure of human NMDA receptor subunits will be a step in the right direction. However, it requires the collaborative efforts of medicinal chemists, physicists, bioinformaticists, computational chemistry, and structural biology.

Author Contributions

The manuscript was conceptualized, designed and written by A.O.E.; J.J. reviewed and edited the final draft. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Wang, W.; Li, T.; Wang, Z.; Yin, Y.; Zhang, S.; Wang, C.; Hu, X.; Lu, S. Bibliometric Analysis of Research on Neurodegenerative Diseases and Single-Cell RNA Sequencing: Opportunities and Challenges. iScience 2023, 26, 107833. [Google Scholar] [CrossRef]
  2. Fratiglioni, L.; Qiu, C. Prevention of Common Neurodegenerative Disorders in the Elderly. Exp. Gerontol. 2009, 44, 46–50. [Google Scholar] [CrossRef]
  3. Ullah, M.F.; Ahmad, A.; Bhat, S.H.; Abu-Duhier, F.M.; Barreto, G.E.; Ashraf, G.M. Impact of Sex Differences and Gender Specificity on Behavioral Characteristics and Pathophysiology of Neurodegenerative Disorders. Neurosci. Biobehav. Rev. 2019, 102, 95–105. [Google Scholar] [CrossRef]
  4. Maselli, F.; D’Antona, S.; Utichi, M.; Arnaudi, M.; Castiglioni, I.; Porro, D.; Papaleo, E.; Gandellini, P.; Cava, C. Computational Analysis of Five Neurodegenerative Diseases Reveals Shared and Specific Genetic Loci. Comput. Struct. Biotechnol. J. 2023, 21, 5395–5407. [Google Scholar] [CrossRef]
  5. Joshi, R.; Missong, H.; Mishra, J.; Kaur, S.; Saini, S.; Kandimalla, R.; Reddy, P.H.; Babu, A.; Bhatti, G.K.; Bhatti, J.S. Nanotheranostics Revolutionizing Neurodegenerative Diseases: From Precision Diagnosis to Targeted Therapies. J. Drug Deliv. Sci. Technol. 2023, 89, 105067. [Google Scholar] [CrossRef]
  6. Arbo, B.D.; Schimith, L.E.; Goulart dos Santos, M.; Hort, M.A. Repositioning and Development of New Treatments for Neurodegenerative Diseases: Focus on Neuroinflammation. Eur. J. Pharmacol. 2022, 919, 174800. [Google Scholar] [CrossRef]
  7. Liu, W.; Li, Y.; Zhao, T.; Gong, M.; Wang, X.; Zhang, Y.; Xu, L.; Li, W.; Li, Y.; Jia, J. The Role of N-Methyl-D-Aspartate Glutamate Receptors in Alzheimer’s Disease: From Pathophysiology to Therapeutic Approaches. Prog. Neurobiol. 2023, 231, 102534. [Google Scholar] [CrossRef]
  8. Agid, Y. Neurodegenerative Disorders: Are We Wrong? Rev. Neurol. 2022, 178, 407–413. [Google Scholar] [CrossRef]
  9. Mendes, D.; Peixoto, F.; Oliveira, M.M.; Andrade, P.B.; Videira, R.A. Mitochondria Research and Neurodegenerative Diseases: On the Track to Understanding the Biological World of High Complexity. Mitochondrion 2022, 65, 67–79. [Google Scholar] [CrossRef] [PubMed]
  10. Ontario, M.L.; Siracusa, R.; Modafferi, S.; Scuto, M.; Sciuto, S.; Greco, V.; Bertuccio, M.P.; Trovato Salinaro, A.; Crea, R.; Calabrese, E.J.; et al. Potential Prevention and Treatment of Neurodegenerative Disorders by Olive Polyphenols and Hidrox. Mech. Ageing Dev. 2022, 203, 111637. [Google Scholar] [CrossRef]
  11. Awad, R.; Avital, A.; Sosnik, A. Polymeric Nanocarriers for Nose-to-Brain Drug Delivery in Neurodegenerative Diseases and Neurodevelopmental Disorders. Acta Pharm. Sin. B 2023, 13, 1866–1886. [Google Scholar] [CrossRef] [PubMed]
  12. Thukral, P.; Chowdhury, R.; Sable, H.; Kaushik, A.; Chaudhary, V. Sustainable Green Synthesized Nanoparticles for Neurodegenerative Diseases Diagnosis and Treatment. Mater. Today Proc. 2023, 73, 323–328. [Google Scholar] [CrossRef]
  13. Rose, C.R.; Ziemens, D.; Untiet, V.; Fahlke, C. Molecular and Cellular Physiology of Sodium-Dependent Glutamate Transporters. Brain Res. Bull. 2018, 136, 3–16. [Google Scholar] [CrossRef] [PubMed]
  14. van Onselen, R.; Downing, T.G. Uptake of β-N-Methylamino-L-Alanine (BMAA) into Glutamate-Specific Synaptic Vesicles: Exploring the Validity of the Excitotoxicity Mechanism of BMAA. Neurosci. Lett. 2024, 821, 137593. [Google Scholar] [CrossRef] [PubMed]
  15. Luo, L.; Li, D.; Xu, X.; Jia, Q.; Li, Z.; Xu, R.; Chen, Z.; Zhao, Y. Synthesis and Neuroprotective Effects of New Genipin Derivatives against Glutamate-Induced Oxidative Damage. Fitoterapia 2023, 169, 105616. [Google Scholar] [CrossRef] [PubMed]
  16. Neves, D.; Salazar, I.L.; Almeida, R.D.; Silva, R.M. Molecular Mechanisms of Ischemia and Glutamate Excitotoxicity. Life Sci. 2023, 328, 121814. [Google Scholar] [CrossRef]
  17. Wang, X.; Zhang, F.; Niu, L.; Yan, J.; Liu, H.; Wang, D.; Hui, J.; Dai, H.; Song, J.; Zhang, Z. High-Frequency Repetitive Transcranial Magnetic Stimulation Improves Depressive-like Behaviors in CUMS-Induced Rats by Modulating Astrocyte GLT-1 to Reduce Glutamate Toxicity. J. Affect. Disord. 2024, 348, 265–274. [Google Scholar] [CrossRef] [PubMed]
  18. Shen, Z.; Xiang, M.; Chen, C.; Ding, F.; Wang, Y.; Shang, C.; Xin, L.; Zhang, Y.; Cui, X. Glutamate Excitotoxicity: Potential Therapeutic Target for Ischemic Stroke. Biomed. Pharmacother. 2022, 151, 113125. [Google Scholar] [CrossRef] [PubMed]
  19. Velu, L.; Pellerin, L.; Julian, A.; Paccalin, M.; Giraud, C.; Fayolle, P.; Guillevin, R.; Guillevin, C. Early Rise of Glutamate-Glutamine Levels in Mild Cognitive Impairment: Evidence for Emerging Excitotoxicity. J. Neuroradiol. 2024, 51, 168–175. [Google Scholar] [CrossRef]
  20. Khesmakhi, M.V.; Salimi, Z.; Pourmotabbed, A.; Moradpour, F.; Rezayof, A.; Nedaei, S.E. The Role of Glutamate NMDA Receptors of the Mediodorsal Thalamus in Scopolamine-Induced Amnesia in Rats. Neurosci. Lett. 2024, 820, 137595. [Google Scholar] [CrossRef]
  21. Iovino, L.; Tremblay, M.E.; Civiero, L. Glutamate-Induced Excitotoxicity in Parkinson’s Disease: The Role of Glial Cells. J. Pharmacol. Sci. 2020, 144, 151–164. [Google Scholar] [CrossRef]
  22. Burada, A.P.; Vinnakota, R.; Bharti, P.; Dutta, P.; Dubey, N.; Kumar, J. Emerging Insights into the Structure and Function of Ionotropic Glutamate Delta Receptors. Br. J. Pharmacol. 2022, 179, 3612–3627. [Google Scholar] [CrossRef]
  23. Mori, H.; Mishina, M. Structure and Function of the NMDA Receptor Channel. Neuropharmacology 1995, 34, 1219–1237. [Google Scholar] [CrossRef]
  24. Murthy, S.E. Glycine-Bound NMDA Receptors Are Stretch-Activated. Trends Neurosci. 2022, 45, 794–795. [Google Scholar] [CrossRef]
  25. Olivero, G.; Grilli, M.; Marchi, M.; Pittaluga, A. Metamodulation of Presynaptic NMDA Receptors: New Perspectives for Pharmacological Interventions. Neuropharmacology 2023, 234, 109570. [Google Scholar] [CrossRef]
  26. Li, W.; Kutas, M.; Gray, J.A.; Hagerman, R.H.; Olichney, J.M. The Role of Glutamate in Language and Language Disorders—Evidence from ERP and Pharmacologic Studies. Neurosci. Biobehav. Rev. 2020, 119, 217–241. [Google Scholar] [CrossRef]
  27. Yi, F.; Mou, T.-C.; Dorsett, K.N.; Volkmann, R.A.; Menniti, F.S.; Sprang, S.R.; Hansen, K.B. Structural Basis for Negative Allosteric Modulation of GluN2A-Containing NMDA Receptors. Neuron 2016, 91, 1316–1329. [Google Scholar] [CrossRef] [PubMed]
  28. Zhou, C.; Tajima, N. Structural Insights into NMDA Receptor Pharmacology. Biochem. Soc. Trans. 2023, 51, 1713–1731. [Google Scholar] [CrossRef]
  29. Gawande, D.Y.; Shelkar, G.P.; Narasimhan, K.K.S.; Liu, J.; Dravid, S.M. GluN2D Subunit-Containing NMDA Receptors Regulate Reticular Thalamic Neuron Function and Seizure Susceptibility. Neurobiol. Dis. 2023, 181, 106117. [Google Scholar] [CrossRef]
  30. Mony, L.; Paoletti, P. Mechanisms of NMDA Receptor Regulation. Curr. Opin. Neurobiol. 2023, 83, 102815. [Google Scholar] [CrossRef] [PubMed]
  31. Wang, C.X.; Shuaib, A. NMDA/NR2BA Selective Antagonists in the Treatment of Ischemic Brain Injury. Curr. Drug Targets CNS Neurol. Disord. 2005, 4, 143–151. [Google Scholar] [CrossRef] [PubMed]
  32. Bi, X.; Rong, Y.; Chen, J.; Dang, S.; Wang, Z.; Baudry, M. Calpain-Mediated Regulation of NMDA Receptor Structure and Function. Brain Res. 1998, 790, 245–253. [Google Scholar] [CrossRef] [PubMed]
  33. Jiang, L.; Liu, N.; Zhao, F.; Huang, B.; Kang, D.; Zhan, P.; Liu, X. Discovery of GluN2A Subtype-Selective N-Methyl-d-Aspartate (NMDA) Receptor Ligands. Acta Pharm. Sin. B 2024, 14, 1987–2005. [Google Scholar] [CrossRef]
  34. Wu, E.; Zhang, J.; Zhang, J.; Zhu, S. Structural Insights into Gating Mechanism and Allosteric Regulation of NMDA Receptors. Curr. Opin. Neurobiol. 2023, 83, 102806. [Google Scholar] [CrossRef] [PubMed]
  35. Nikam, S.; Meltzer, L. NR2B Selective NMDA Receptor Antagonists. Curr. Pharm. Des. 2005, 8, 845–855. [Google Scholar] [CrossRef]
  36. Zhang, Z.; Zhang, S.; Fu, P.; Zhang, Z.; Lin, K.; Ko, J.K.-S.; Yung, K.K.-L. Roles of Glutamate Receptors in Parkinson’s Disease. Int. J. Mol. Sci. 2019, 20, 4391. [Google Scholar] [CrossRef] [PubMed]
  37. Tsai, Y.C.; Huang, S.M.; Peng, H.H.; Lin, S.W.; Lin, S.R.; Chin, T.Y.; Huang, S.M. Imbalance of Synaptic and Extrasynaptic NMDA Receptors Induced by the Deletion of CRMP1 Accelerates Age-Related Cognitive Decline in Mice. Neurobiol. Aging 2024, 135, 48–59. [Google Scholar] [CrossRef] [PubMed]
  38. Tamborini, L.; Pinto, A.; Mastronardi, F.; Iannuzzi, M.C.; Cullia, G.; Nielsen, B.; De Micheli, C.; Conti, P. 3-Carboxy-Pyrazolinalanine as a New Scaffold for Developing Potent and Selective NMDA Receptor Antagonists. Eur. J. Med. Chem. 2013, 68, 33–37. [Google Scholar] [CrossRef] [PubMed]
  39. Geter-Douglass, B.; Witkin, J.M. Dizocilpine-like Discriminative Stimulus Effects of Competitive NMDA Receptor Antagonists in Mice. Psychopharmacology 1997, 133, 43–50. [Google Scholar] [CrossRef] [PubMed]
  40. Kvist, T.; Steffensen, T.B.; Greenwood, J.R.; Mehrzad Tabrizi, F.; Hansen, K.B.; Gajhede, M.; Pickering, D.S.; Traynelis, S.F.; Kastrup, J.S.; Bräuner-Osborne, H. Crystal Structure and Pharmacological Characterization of a Novel N-Methyl-d-Aspartate (NMDA) Receptor Antagonist at the GluN1 Glycine Binding Site. J. Biol. Chem. 2013, 288, 33124–33135. [Google Scholar] [CrossRef]
  41. Jespersen, A.; Tajima, N.; Fernandez-Cuervo, G.; Garnier-Amblard, E.C.; Furukawa, H. Structural Insights into Competitive Antagonism in NMDA Receptors. Neuron 2014, 81, 366–378. [Google Scholar] [CrossRef] [PubMed]
  42. Wlaź, P.; Ebert, U.; Potschka, H.; Löscher, W. Electrical but Not Chemical Kindling Increases Sensitivity to Some Phencyclidine-like Behavioral Effects Induced by the Competitive NMDA Receptor Antagonist d-CPPene in Rats. Eur. J. Pharmacol. 1998, 353, 177–189. [Google Scholar] [CrossRef] [PubMed]
  43. Lind, G.E.; Mou, T.-C.; Tamborini, L.; Pomper, M.G.; De Micheli, C.; Conti, P.; Pinto, A.; Hansen, K.B. Structural Basis of Subunit Selectivity for Competitive NMDA Receptor Antagonists with Preference for GluN2A over GluN2B Subunits. Proc. Natl. Acad. Sci. USA 2017, 114, E6942–E6951. [Google Scholar] [CrossRef] [PubMed]
  44. Furuya, Y.; Kagaya, T.; Ogura, H.; Nishizawa, Y. Competitive NMDA Receptor Antagonists Disrupt Prepulse Inhibition without Reduction of Startle Amplitude in a Dopamine Receptor-Independent Manner in Mice. Eur. J. Pharmacol. 1999, 364, 133–140. [Google Scholar] [CrossRef] [PubMed]
  45. Driver, C.; Jackson, T.N.W.; Lagopoulos, J.; Hermens, D.F. Molecular Mechanisms Underlying the N-Methyl-d-Aspartate Receptor Antagonists: Highlighting Their Potential for Transdiagnostic Therapeutics. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2022, 119, 110609. [Google Scholar] [CrossRef] [PubMed]
  46. Takizawa, S.; Hogan, M.; Hakim, A.M. The Effects of a Competitive NMDA Receptor Antagonist (CGS-19755) on Cerebral Blood Flow and PH in Focal Ischemia. J. Cereb. Blood Flow Metab. 1991, 11, 786–793. [Google Scholar] [CrossRef] [PubMed]
  47. Willmore, C.B.; Bespalov, A.Y.; Beardsley, P.M. Competitive and Noncompetitive NMDA Antagonist Effects in Rats Trained to Discriminate Lever-Press Counts. Pharmacol. Biochem. Behav. 2001, 69, 493–502. [Google Scholar] [CrossRef] [PubMed]
  48. Avenet, P.; Léonardon, J.; Besnard, F.; Graham, D.; Depoortere, H.; Scatton, B. Antagonist Properties of Eliprodil and Other NMDA Receptor Antagonists at Rat NR1A/NR2A and NR1A/NR2B Receptors Expressed in Xenopus Oocytes. Neurosci. Lett. 1997, 223, 133–136. [Google Scholar] [CrossRef] [PubMed]
  49. Danysz, W.; Parsons, C.G.; Karcz-Kubicha, M.; Schwaier, A.; Popik, P.; Wedzony, K.; Lazarewicz, J.; Quack, G. GlycineB Antagonists as Potential Therapeutic Agents. Previous Hopes and Present Reality. Amino Acids 1998, 14, 235–239. [Google Scholar] [CrossRef]
  50. Mugnaini, M.; Dal Forno, G.; Corsi, M.; Bunnemann, B. Receptor Binding Characteristics of the Novel NMDA Receptor Glycine Site Antagonist [3H]GV150526A in Rat Cerebral Cortical Membranes. Eur. J. Pharmacol. 2000, 391, 233–241. [Google Scholar] [CrossRef]
  51. Jansen, M.; Dannhardt, G. Antagonists and Agonists at the Glycine Site of the NMDA Receptor for Therapeutic Interventions. Eur. J. Med. Chem. 2003, 38, 661–670. [Google Scholar] [CrossRef] [PubMed]
  52. Péez-Pinzón, M.A.; Steinberg, G.K. CGS 19755 (Selfotel): A Novel Neuroprotective Agent Against CNS Injury. CNS Drug Rev. 1996, 2, 257–268. [Google Scholar] [CrossRef] [PubMed]
  53. Sharkey, J.; Ritchie, I.M.; Butcher, S.P.; Kelly, J.S. Comparison of the Patterns of Altered Cerebral Glucose Utilisation Produced by Competitive and Non-Competitive NMDA Receptor Antagonists. Brain Res. 1996, 735, 67–82. [Google Scholar] [CrossRef]
  54. Bespalov, A.; Kudryashova, M.; Zvartau, E. Prolongation of Morphine Analgesia by Competitive NMDA Receptor Antagonist D-CPPene (SDZ EAA 494) in Rats. Eur. J. Pharmacol. 1998, 351, 299–305. [Google Scholar] [CrossRef] [PubMed]
  55. Carmack, S.A.; Kim, J.S.; Sage, J.R.; Thomas, A.W.; Skillicorn, K.N.; Anagnostaras, S.G. The Competitive NMDA Receptor Antagonist CPP Disrupts Cocaine-Induced Conditioned Place Preference, but Spares Behavioral Sensitization. Behav. Brain Res. 2013, 239, 155–163. [Google Scholar] [CrossRef]
  56. Riaza Bermudo-Soriano, C.; Perez-Rodriguez, M.M.; Vaquero-Lorenzo, C.; Baca-Garcia, E. New Perspectives in Glutamate and Anxiety. Pharmacol. Biochem. Behav. 2012, 100, 752–774. [Google Scholar] [CrossRef] [PubMed]
  57. Liu, P.; Zhang, T.; Chen, Q.; Li, C.; Chu, Y.; Guo, Q.; Zhang, Y.; Zhou, W.; Chen, H.; Zhou, Z.; et al. Biomimetic Dendrimer–Peptide Conjugates for Early Multi-Target Therapy of Alzheimer’s Disease by Inflammatory Microenvironment Modulation. Adv. Mater. 2021, 33, 2100746. [Google Scholar] [CrossRef] [PubMed]
  58. Jullienne, A.; Montagne, A.; Orset, C.; Lesept, F.; Jane, D.E.; Monaghan, D.T.; Maubert, E.; Vivien, D.; Ali, C. Selective Inhibition of GluN2D-Containing N-Methyl-D-Aspartate Receptors Prevents Tissue Plasminogen Activator-Promoted Neurotoxicity Both in Vitro and in Vivo. Mol. Neurodegener. 2011, 6, 68. [Google Scholar] [CrossRef] [PubMed]
  59. Mabrouk, O.S.; Mela, F.; Calcagno, M.; Budri, M.; Viaro, R.; Dekundy, A.; Parsons, C.G.; Auberson, Y.P.; Morari, M. GluN2A and GluN2B NMDA Receptor Subunits Differentially Modulate Striatal Output Pathways and Contribute to Levodopa-Induced Abnormal Involuntary Movements in Dyskinetic Rats. ACS Chem. Neurosci. 2013, 4, 808–816. [Google Scholar] [CrossRef] [PubMed]
  60. Mikics, E.; Toth, M.; Biro, L.; Bruzsik, B.; Nagy, B.; Haller, J. The Role of GluN2B-Containing NMDA Receptors in Short- and Long-Term Fear Recall. Physiol. Behav. 2017, 177, 44–48. [Google Scholar] [CrossRef]
  61. Papp, M.; Moryl, E. Antidepressant Activity of Non-Competitive and Competitive NMDA Receptor Antagonists in a Chronic Mild Stress Model of Depression. Eur. J. Pharmacol. 1994, 263, 1–7. [Google Scholar] [CrossRef]
  62. Szakács, R.; Weiczner, R.; Mihály, A.; Krisztin-Péva, B.; Zádor, Z.; Zádor, E. Non-Competitive NMDA Receptor Antagonists Moderate Seizure-Induced c-Fos Expression in the Rat Cerebral Cortex. Brain Res. Bull. 2003, 59, 485–493. [Google Scholar] [CrossRef]
  63. Tóth, Z.; Mihály, A.; Mátyás, A.; Krisztin-Péva, B. Non-Competitive Antagonists of NMDA and AMPA Receptors Decrease Seizure-Induced c-Fos Protein Expression in the Cerebellum and Protect against Seizure Symptoms in Adult Rats. Acta Histochem. 2018, 120, 236–241. [Google Scholar] [CrossRef]
  64. Bubser, M.; Keseberg, U.; Notz, P.K.; Schmidt, W.J. Differential Behavioural and Neurochemical Effects of Competitive and Non-Competitive NMDA Receptor Antagonists in Rats. Eur. J. Pharmacol. 1992, 229, 75–82. [Google Scholar] [CrossRef]
  65. Sun, W.L.; Wessinger, W.D. Characterization of the Non-Competitive Antagonist Binding Site of the NMDA Receptor in Dark Agouti Rats. Life Sci. 2004, 75, 1405–1415. [Google Scholar] [CrossRef]
  66. Flores, A.J.; Bartlett, M.J.; So, L.Y.; Laude, N.D.; Parent, K.L.; Heien, M.L.; Sherman, S.J.; Falk, T. Differential Effects of the NMDA Receptor Antagonist MK-801 on Dopamine Receptor D1- and D2-Induced Abnormal Involuntary Movements in a Preclinical Model. Neurosci. Lett. 2014, 564, 48–52. [Google Scholar] [CrossRef]
  67. Berger, M.L.; Rebernik, P. Differential Influence of 7 Cations on 16 Non-Competitive NMDA Receptor Blockers. Bioorg. Med. Chem. Lett. 2015, 25, 4131–4135. [Google Scholar] [CrossRef]
  68. Marcus, M.M.; Mathé, J.M.; Nomikos, G.G.; Svensson, T.H. Effects of Competitive and Non-Competitive NMDA Receptor Antagonists on Dopamine Output in the Shell and Core Subdivisions of the Nucleus Accumbens. Neuropharmacology 2001, 40, 482–490. [Google Scholar] [CrossRef]
  69. Blot, K.; Bai, J.; Otani, S. The Effect of Non-Competitive NMDA Receptor Antagonist MK-801 on Neuronal Activity in Rodent Prefrontal Cortex: An Animal Model for Cognitive Symptoms of Schizophrenia. J. Physiol. Paris 2013, 107, 448–451. [Google Scholar] [CrossRef]
  70. Stuchlík, A.; Vales, K. Systemic Administration of MK-801, a Non-Competitive NMDA-Receptor Antagonist, Elicits a Behavioural Deficit of Rats in the Active Allothetic Place Avoidance (AAPA) Task Irrespectively of Their Intact Spatial Pretraining. Behav. Brain Res. 2005, 159, 163–171. [Google Scholar] [CrossRef]
  71. 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]
  72. Raghavendra Rao, V.L.; Dogan, A.; Todd, K.G.; Bowen, K.K.; Dempsey, R.J. Neuroprotection by Memantine, a Non-Competitive NMDA Receptor Antagonist after Traumatic Brain Injury in Rats. Brain Res. 2001, 911, 96–100. [Google Scholar] [CrossRef]
  73. Losi, G.; Lanza, M.; Makovec, F.; Artusi, R.; Caselli, G.; Puia, G. Functional in Vitro Characterization of CR 3394: A Novel Voltage Dependent N-Methyl-D-Aspartate (NMDA) Receptor Antagonist. Neuropharmacology 2006, 50, 277–285. [Google Scholar] [CrossRef]
  74. Geldenhuys, W.J.; Van der Schyf, C.J. Rationally Designed Multi-Targeted Agents Against Neurodegenerative Diseases. Curr. Med. Chem. 2013, 20, 1662–1672. [Google Scholar] [CrossRef]
  75. Shafiei-Irannejad, V.; Abbaszadeh, S.; Janssen, P.M.L.; Soraya, H. Memantine and Its Benefits for Cancer, Cardiovascular and Neurological Disorders. Eur. J. Pharmacol. 2021, 910, 174455. [Google Scholar] [CrossRef]
  76. Lin, J.-C.; Chan, M.-H.; Lee, M.-Y.; Chen, Y.-C.; Chen, H.-H. N,N-Dimethylglycine Differentially Modulates Psychotomimetic and Antidepressant-like Effects of Ketamine in Mice. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2016, 71, 7–13. [Google Scholar] [CrossRef]
  77. Suzuki, A.; Murakami, K.; Tajima, Y.; Hara, H.; Kunugi, A.; Kimura, H. TAK-137, an AMPA Receptor Potentiator with Little Agonistic Effect, Produces Antidepressant-like Effect without Causing Psychotomimetic Effects in Rats. Pharmacol. Biochem. Behav. 2019, 183, 80–86. [Google Scholar] [CrossRef]
  78. Subramanian, S.; Haroutounian, S.; Palanca, B.J.A.; Lenze, E.J. Ketamine as a Therapeutic Agent for Depression and Pain: Mechanisms and Evidence. J. Neurol. Sci. 2022, 434, 120152. [Google Scholar] [CrossRef]
  79. Costa, B.M.; Irvine, M.W.; Fang, G.; Eaves, R.J.; Mayo-Martin, M.B.; Skifter, D.A.; Jane, D.E.; Monaghan, D.T. A Novel Family of Negative and Positive Allosteric Modulators of NMDA Receptors. J. Pharmacol. Exp. Ther. 2010, 335, 614–621. [Google Scholar] [CrossRef]
  80. Sirrieh, R.E.; MacLean, D.M.; Jayaraman, V. A Conserved Structural Mechanism of NMDA Receptor Inhibition: A Comparison of Ifenprodil and Zinc. J. Gen. Physiol. 2015, 146, 173–181. [Google Scholar] [CrossRef]
  81. Zhang, Z.; Liu, J.; Fan, C.; Mao, L.; Xie, R.; Wang, S.; Yang, M.; Yuan, H.; Yang, X.; Sun, J.; et al. The GluN1/GluN2B NMDA Receptor and Metabotropic Glutamate Receptor 1 Negative Allosteric Modulator Has Enhanced Neuroprotection in a Rat Subarachnoid Hemorrhage Model. Exp. Neurol. 2018, 301, 13–25. [Google Scholar] [CrossRef] [PubMed]
  82. Zhu, S.; Paoletti, P. Allosteric Modulators of NMDA Receptors: Multiple Sites and Mechanisms. Curr. Opin. Pharmacol. 2015, 20, 14–23. [Google Scholar] [CrossRef]
  83. Zhu, S.; Stein, R.A.; Yoshioka, C.; Lee, C.-H.; Goehring, A.; Mchaourab, H.S.; Gouaux, E. Mechanism of NMDA Receptor Inhibition and Activation. Cell 2016, 165, 704–714. [Google Scholar] [CrossRef]
  84. Quan, J.; Yang, H.; Qin, F.; He, Y.; Liu, J.; Zhao, Y.; Ma, C.; Cheng, M. Discovery of Novel Tryptamine Derivatives as GluN2B Subunit-Containing NMDA Receptor Antagonists via Pharmacophore-Merging Strategy with Orally Available Therapeutic Effect of Cerebral Ischemia. Eur. J. Med. Chem. 2023, 253, 115318. [Google Scholar] [CrossRef] [PubMed]
  85. Mony, L.; Kew, J.N.C.; 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]
  86. Zhang, L.; Huang, L.H.; Chen, L.; Hao, D.; Chen, J. Neuroprotection by Tetrahydroxystilbene Glucoside in the MPTP Mouse Model of Parkinson’s Disease. Toxicol. Lett. 2013, 222, 155–163. [Google Scholar] [CrossRef] [PubMed]
  87. Irvine, M.W.; Fang, G.; Sapkota, K.; Burnell, E.S.; Volianskis, A.; Costa, B.M.; Culley, G.; Collingridge, G.L.; Monaghan, D.T.; Jane, D.E. Investigation of the Structural Requirements for N-Methyl-D-Aspartate Receptor Positive and Negative Allosteric Modulators Based on 2-Naphthoic Acid. Eur. J. Med. Chem. 2019, 164, 471–498. [Google Scholar] [CrossRef]
  88. Gibb, A.J. Allosteric Antagonist Action at Triheteromeric NMDA Receptors. Neuropharmacology 2022, 202, 108861. [Google Scholar] [CrossRef]
  89. Markus, A.; Schreiber, J.A.; Goerges, G.; Frehland, B.; Schepmann, D.; Daniliuc, C.; Fröhlich, R.; Seebohm, G.; Wünsch, B. Phenol-Benzoxazolone Bioisosteres of GluN2B-NMDA Receptor Antagonists: Unexpected Rearrangement during Reductive Alkylation with Phenylcyclohexanone. Arch. Pharm. 2022, 355, 2200225. [Google Scholar] [CrossRef]
  90. Temme, L.; Bechthold, E.; Schreiber, J.A.; Gawaskar, S.; Schepmann, D.; Robaa, D.; Sippl, W.; Seebohm, G.; Wünsch, B. Negative Allosteric Modulators of the GluN2B NMDA Receptor with Phenylethylamine Structure Embedded in Ring-Expanded and Ring-Contracted Scaffolds. Eur. J. Med. Chem. 2020, 190, 112138. [Google Scholar] [CrossRef]
  91. 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]
  92. Sapkota, K.; Burnell, E.S.; Irvine, M.W.; Fang, G.; Gawande, D.Y.; Dravid, S.M.; Jane, D.E.; Monaghan, D.T. Pharmacological Characterization of a Novel Negative Allosteric Modulator of NMDA Receptors, UBP792. Neuropharmacology 2021, 201, 108818. [Google Scholar] [CrossRef]
  93. Hanson, J.E.; Yuan, H.; Perszyk, R.E.; Banke, T.G.; Xing, H.; Tsai, M.C.; Menniti, F.S.; Traynelis, S.F. Therapeutic Potential of N-Methyl-D-Aspartate Receptor Modulators in Psychiatry. Neuropsychopharmacology 2024, 49, 51–66. [Google Scholar] [CrossRef] [PubMed]
  94. Yao, L.; Zhou, Q. Enhancing NMDA Receptor Function: Recent Progress on Allosteric Modulators. Neural Plast. 2017, 2017, 2875904. [Google Scholar] [CrossRef] [PubMed]
  95. Subramaniam, S.; O’Connor, M.J.; Masukawa, L.M.; McGonigle, P. Polyamine Effects on the NMDA Receptor in Human Brain. Exp. Neurol. 1994, 130, 323–330. [Google Scholar] [CrossRef]
  96. Monaghan, D.T.; Irvine, M.W.; Costa, B.M.; Fang, G.; Jane, D.E. Pharmacological Modulation of NMDA Receptor Activity and the Advent of Negative and Positive Allosteric Modulators. Neurochem. Int. 2012, 61, 581–592. [Google Scholar] [CrossRef]
  97. Kane, L.T.; Costa, B.M. Identification of Novel Allosteric Modulator Binding Sites in NMDA Receptors: A Molecular Modeling Study. J. Mol. Graph. Model. 2015, 61, 204–213. [Google Scholar] [CrossRef]
  98. Perszyk, R.; Katzman, B.M.; Kusumoto, H.; Kell, S.A.; Epplin, M.P.; Tahirovic, Y.A.; Moore, R.L.; Menaldino, D.; Burger, P.; Liotta, D.C.; et al. An NMDAR Positive and Negative Allosteric Modulator Series Share a Binding Site and Are Interconverted by Methyl Groups. eLife 2018, 7, e34711. [Google Scholar] [CrossRef] [PubMed]
  99. Sun, J.Y.; Kumata, K.; Chen, Z.; Zhang, Y.D.; Chen, J.H.; Hatori, A.; Fu, H.L.; Rong, J.; Deng, X.Y.; Yamasaki, T.; et al. Synthesis and Preliminary Evaluation of Novel 11C-Labeled GluN2B-Selective NMDA Receptor Negative Allosteric Modulators. Acta Pharmacol. Sin. 2021, 42, 491–498. [Google Scholar] [CrossRef]
  100. Auvin, S.; Dozières-Puyravel, B.; Avbersek, A.; Sciberras, D.; Collier, J.; Leclercq, K.; Mares, P.; Kaminski, R.M.; Muglia, P. Radiprodil, a NR2B Negative Allosteric Modulator, from Bench to Bedside in Infantile Spasm Syndrome. Ann. Clin. Transl. Neurol. 2020, 7, 343–352. [Google Scholar] [CrossRef]
  101. Montastruc, J.L.; Rascol, O.; Senard, J.M. Glutamate Antagonists and Parkinson’s Disease: A Review of Clinical Data. Neurosci. Biobehav. Rev. 1997, 21, 477–480. [Google Scholar] [CrossRef] [PubMed]
  102. Sasaki, T.; Hashimoto, K.; Niitsu, T.; Hosoda, Y.; Oda, Y.; Shiko, Y.; Ozawa, Y.; Kawasaki, Y.; Kanahara, N.; Shiina, A.; et al. Ifenprodil Tartrate Treatment of Adolescents with Post-Traumatic Stress Disorder: A Double-Blind, Placebo-Controlled Trial. Psychiatry Res. 2022, 311, 114486. [Google Scholar] [CrossRef]
  103. Swanger, S.A.; Vance, K.M.; Acker, T.M.; Zimmerman, S.S.; DiRaddo, J.O.; Myers, S.J.; Bundgaard, C.; Mosley, C.A.; Summer, S.L.; Menaldino, D.S.; et al. A Novel Negative Allosteric Modulator Selective for GluN2C/2D-Containing NMDA Receptors Inhibits Synaptic Transmission in Hippocampal Interneurons. ACS Chem. Neurosci. 2018, 9, 306–319. [Google Scholar] [CrossRef]
  104. Rajan, R.; Schepmann, D.; Schreiber, J.A.; Seebohm, G.; Wünsch, B. Synthesis of GluN2A-Selective NMDA Receptor Antagonists with an Electron-Rich Aromatic B-Ring. Eur. J. Med. Chem. 2021, 209, 112939. [Google Scholar] [CrossRef] [PubMed]
  105. Costa, B.M.; Irvine, M.W.; Fang, G.; Eaves, R.J.; Mayo-Martin, M.B.; Laube, B.; Jane, D.E.; Monaghan, D.T. Structure-Activity Relationships for Allosteric NMDA Receptor Inhibitors Based on 2-Naphthoic Acid. Neuropharmacology 2012, 62, 1730–1736. [Google Scholar] [CrossRef]
  106. Sapkota, K.; Dore, K.; Tang, K.; Irvine, M.; Fang, G.; Burnell, E.S.; Malinow, R.; Jane, D.E.; Monaghan, D.T. The NMDA Receptor Intracellular C-Terminal Domains Reciprocally Interact with Allosteric Modulators. Biochem. Pharmacol. 2019, 159, 140–153. [Google Scholar] [CrossRef]
  107. Campbell, I.G.; Gustafson, L.M.; Feinberg, I. The Competitive NMDA Receptor Antagonist CPPene Stimulates NREM Sleep and Eating in Rats. Neuropsychopharmacology 2002, 26, 348–357. [Google Scholar] [CrossRef]
  108. Baron, B.M.; Siegel, B.W.; Slone, A.L.; Harrison, B.L.; Palfreyman, M.G.; Hurt, S.D. [3H]5,7-Dichlorokynurenic Acid, a Novel Radioligand Labels NMDA Receptor-Associated Glycine Binding Sites. Eur. J. Pharmacol. Mol. Pharmacol. 1991, 206, 149–154. [Google Scholar] [CrossRef]
  109. Corbett, R.; Dunn, R.W. Effects of 5,7 Dichlorokynurenic Acid on Conflict, Social Interaction and plus Maze Behaviors. Neuropharmacology 1993, 32, 461–466. [Google Scholar] [CrossRef]
  110. Frankiewicz, T.; Pilc, A.; Parsons, C.G. Differential Effects of NMDA–Receptor Antagonists on Long-Term Potentiation and Hypoxic/Hypoglycaemic Excitotoxicity in Hippocampal Slices. Neuropharmacology 2000, 39, 631–642. [Google Scholar] [CrossRef]
  111. Hauber, W.; Waldenmeier, M.T. The AMPA Receptor Antagonist GYKI 52466 Reverses the Anti-Cataleptic Effects of the Competitive NMDA Receptor Antagonist CGP 37849. Eur. J. Pharmacol. 1994, 256, 339–342. [Google Scholar] [CrossRef]
  112. Obrenovitch, T.P.; Hardy, A.M.; Zilkha, E. Effects of L-701,324, a High-Affinity Antagonist at the N-Methyl-D-Aspartate (NMDA) Receptor Glycine Site, on the Rat Electroencephalogram. Naunyn. Schmiedebergs. Arch. Pharmacol. 1997, 355, 779–786. [Google Scholar] [CrossRef] [PubMed]
  113. Wlaź, P.; Ebert, U.; Löscher, W. Anticonvulsant Effects of Eliprodil Alone or Combined with the Glycine(B) Receptor Antagonist L-701,324 or the Competitive NMDA Antagonist CGP 40116 in the Amygdala Kindling Model in Rats. Neuropharmacology 1999, 38, 243–251. [Google Scholar] [CrossRef]
  114. Konieczny, J.; Ossowska, K.; Schulze, G.; Coper, H.; Wolfarth, S. L-701,324, a Selective Antagonist at the Glycine Site of the NMDA Receptor, Counteracts Haloperidol-Induced Muscle Rigidity in Rats. Psychopharmacology 1999, 143, 235–243. [Google Scholar] [CrossRef]
  115. Stone, T.W. Development and Therapeutic Potential of Kynurenic Acid and Kynurenine Derivatives for Neuroprotection. Trends Pharmacol. Sci. 2000, 21, 149–154. [Google Scholar] [CrossRef] [PubMed]
  116. Liu, L.; Ji, C.-H.; Wang, Y.; Zhao, J.; Liu, Y.; Tang, W.-Q.; Gu, J.-H.; Jiang, B. Antidepressant-like Activity of L-701324 in Mice: A Behavioral and Neurobiological Characterization. Behav. Brain Res. 2021, 399, 113038. [Google Scholar] [CrossRef] [PubMed]
  117. Feng, B.; Tse, H.W.; Skifter, D.A.; Morley, R.; Jane, D.E.; Monaghan, D.T. Structure—Activity Analysis of a Novel NR2C/NR2D-Preferring NMDA Receptor Antagonist: 1-(Phenanthrene-2-Carbonyl) Piperazine-2,3-Dicarboxylic Acid. Br. J. Pharmacol. 2004, 141, 508–516. [Google Scholar] [CrossRef] [PubMed]
  118. Feng, B.; Morley, R.M.; Jane, D.E.; Monaghan, D.T. The Effect of Competitive Antagonist Chain Length on NMDA Receptor Subunit Selectivity. Neuropharmacology 2005, 48, 354–359. [Google Scholar] [CrossRef] [PubMed]
  119. Bartlett, T.E.; Bannister, N.J.; Collett, V.J.; Dargan, S.L.; Massey, P.V.; Bortolotto, Z.A.; Fitzjohn, S.M.; Bashir, Z.I.; Collingridge, G.L.; Lodge, D. Differential Roles of NR2A and NR2B-Containing NMDA Receptors in LTP and LTD in the CA1 Region of Two-Week Old Rat Hippocampus. Neuropharmacology 2007, 52, 60–70. [Google Scholar] [CrossRef]
  120. Costa, B.M.; Feng, B.; Tsintsadze, T.S.; Morley, R.M.; Irvine, M.W.; Tsintsadze, V.; Lozovaya, N.A.; Jane, D.E.; Monaghan, D.T. N-Methyl-D-Aspartate (NMDA) Receptor NR2 Subunit Selectivity of a Series of Novel Piperazine-2,3-Dicarboxylate Derivatives: Preferential Blockade of Extrasynaptic NMDA Receptors in the Rat Hippocampal CA3-CA1 Synapse. J. Pharmacol. Exp. Ther. 2009, 331, 618–626. [Google Scholar] [CrossRef]
  121. Baron, A.; Montagne, A.; Cassé, F.; Launay, S.; Maubert, E.; Ali, C.; Vivien, D. NR2D-Containing NMDA Receptors Mediate Tissue Plasminogen Activator-Promoted Neuronal Excitotoxicity. Cell Death Differ. 2010, 17, 860–871. [Google Scholar] [CrossRef] [PubMed]
  122. Kinarsky, L.; Feng, B.; Skifter, D.A.; Morley, R.M.; Sherman, S.; Jane, D.E.; Monaghan, D.T. Identification of Subunit- and Antagonist-Specific Amino Acid Residues in the N -Methyl-d-Aspartate Receptor Glutamate-Binding Pocket. J. Pharmacol. Exp. Ther. 2005, 313, 1066–1074. [Google Scholar] [CrossRef] [PubMed]
  123. Wu, L.J.; Zhuo, M. Targeting the NMDA Receptor Subunit NR2B for the Treatment of Neuropathic Pain. Neurotherapeutics 2009, 6, 693–702. [Google Scholar] [CrossRef] [PubMed]
  124. Lemay-Clermont, J.; Robitaille, C.; Auberson, Y.P.; Bureau, G.; Cyr, M. Blockade of NMDA Receptors 2A Subunit in the Dorsal Striatum Impairs the Learning of a Complex Motor Skill. Behav. Neurosci. 2011, 125, 714–723. [Google Scholar] [CrossRef] [PubMed]
  125. Santos, M.; D’Amico, D.; Dierssen, M. From Neural to Genetic Substrates of Panic Disorder: Insights from Human and Mouse Studies. Eur. J. Pharmacol. 2015, 759, 127–141. [Google Scholar] [CrossRef] [PubMed]
  126. Zhu, M.; Zhang, J.; Guo, L.; Yue, R.; Li, S.; Cui, B.; Guo, S.; Niu, Q.; Yu, Y.; Wang, H.; et al. Amorphous Selenium Inhibits Oxidative Stress Injury of Neurons in Vascular Dementia Rats by Activating NMDAR Pathway. Eur. J. Pharmacol. 2023, 955, 175874. [Google Scholar] [CrossRef] [PubMed]
  127. Urwyler, S.; Campbell, E.; Fricker, G.; Jenner, P.; Lemaire, M.; McAllister, K.H.; Neijt, H.C.; Park, C.K.; Perkins, M.; Rudin, M.; et al. Biphenyl-Derivatives of 2-Amino-7-Phosphono-Heptanoic Acid, a Novel Class of Potent Competitive N-Methyl-D-Aspartate Receptor Antagonists—II. Pharmacological Characterization in Vivo. Neuropharmacology 1996, 35, 655–669. [Google Scholar] [CrossRef] [PubMed]
  128. Urwyler, S. Biphenyl-Derivatives of 2-Amino-7-Phosphono-Heptanoic Acid, a Novel Class of Potent Competitive N-Methyl-?-Aspartate Receptor Antagonists—I. Pharmacological Characterization in Vitro. Neuropharmacology 1996, 35, 643–654. [Google Scholar] [CrossRef]
  129. Olsson, S.K.; Linderholm, K.; Erhardt, S.; Engberg, G. P.1.c.039 Increased Midbrain Dopaminergic Firing by the Competitive N-Methyl-D-Aspartate Receptor Antagonist SDZ 220–581. Eur. Neuropsychopharmacol. 2007, 17, S264–S265. [Google Scholar] [CrossRef]
  130. Bender, C.; de Olmos, S.; Bueno, A.; de Olmos, J.; Lorenzo, A. Comparative Analyses of the Neurodegeneration Induced by the Non-Competitive NMDA-Receptor-Antagonist Drug MK801 in Mice and Rats. Neurotoxicol. Teratol. 2010, 32, 542–550. [Google Scholar] [CrossRef]
  131. Mellone, M.; Gardoni, F. Modulation of NMDA Receptor at the Synapse: Promising Therapeutic Interventions in Disorders of the Nervous System. Eur. J. Pharmacol. 2013, 719, 75–83. [Google Scholar] [CrossRef]
  132. Ghasemi, M.; Phillips, C.; Fahimi, A.; McNerney, M.W.; Salehi, A. Mechanisms of Action and Clinical Efficacy of NMDA Receptor Modulators in Mood Disorders. Neurosci. Biobehav. Rev. 2017, 80, 555–572. [Google Scholar] [CrossRef]
  133. Czarnecka, K.; Chuchmacz, J.; Wójtowicz, P.; Szymański, P. Memantine in Neurological Disorders—Schizophrenia and Depression. J. Mol. Med. 2021, 99, 327–334. [Google Scholar] [CrossRef]
  134. Hedegaard, M.; Hansen, K.B.; Andersen, K.T.; Bräuner-Osborne, H.; Traynelis, S.F. Molecular Pharmacology of Human NMDA Receptors. Neurochem. Int. 2012, 61, 601–609. [Google Scholar] [CrossRef]
  135. Rascol, O.; Fabbri, M.; Poewe, W. Amantadine in the Treatment of Parkinson’s Disease and Other Movement Disorders. Lancet Neurol. 2021, 20, 1048–1056. [Google Scholar] [CrossRef]
  136. Kolesnikova, T.O.; Khatsko, S.L.; Shevyrin, V.A.; Morzherin, Y.Y.; Kalueff, A.V. Effects of a Non-Competitive N-Methyl-d-Aspartate (NMDA) Antagonist, Tiletamine, in Adult Zebrafish. Neurotoxicol. Teratol. 2017, 59, 62–67. [Google Scholar] [CrossRef]
  137. Cadinu, D.; Grayson, B.; Podda, G.; Harte, M.K.; Doostdar, N.; Neill, J.C. NMDA Receptor Antagonist Rodent Models for Cognition in Schizophrenia and Identification of Novel Drug Treatments, an Update. Neuropharmacology 2018, 142, 41–62. [Google Scholar] [CrossRef]
  138. Kalmoe, M.C.; Janski, A.M.; Zorumski, C.F.; Nagele, P.; Palanca, B.J.; Conway, C.R. Ketamine and Nitrous Oxide: The Evolution of NMDA Receptor Antagonists as Antidepressant Agents. J. Neurol. Sci. 2020, 412, 116778. [Google Scholar] [CrossRef]
  139. Kolcheva, M.; Ladislav, M.; Netolicky, J.; Kortus, S.; Rehakova, K.; Krausova, B.H.; Hemelikova, K.; Misiachna, A.; Kadkova, A.; Klima, M.; et al. The Pathogenic N650K Variant in the GluN1 Subunit Regulates the Trafficking, Conductance, and Pharmacological Properties of NMDA Receptors. Neuropharmacology 2023, 222, 109297. [Google Scholar] [CrossRef]
  140. Kew, J.N.C.; Trube, G.; Kemp, J.A. State-dependent NMDA Receptor Antagonism by Ro 8-4304, a Novel NR2B Selective, Non-competitive, Voltage-independent Antagonist. Br. J. Pharmacol. 1998, 123, 463–472. [Google Scholar] [CrossRef]
  141. Volgraf, M.; Sellers, B.D.; Jiang, Y.; Wu, G.; Ly, C.Q.; Villemure, E.; Pastor, R.M.; Yuen, P.W.; Lu, A.; Luo, X.; et al. Discovery of GluN2A-Selective NMDA Receptor Positive Allosteric Modulators (PAMs): Tuning Deactivation Kinetics via Structure-Based Design. J. Med. Chem. 2016, 59, 2760–2779. [Google Scholar] [CrossRef]
  142. Jimenez, E.C. Peptide Antagonists of NMDA Receptors: Structure-Activity Relationships for Potential Therapeutics. Peptides 2022, 153, 170796. [Google Scholar] [CrossRef]
  143. Löschmann, P.A.; De Groote, C.; Smith, L.; Wüllner, U.; Fischer, G.; Kemp, J.A.; Jenner, P.; Klockgether, T. Antiparkinsonian Activity of Ro 25-6981, a NR2B Subunit Specific NMDA Receptor Antagonist, in Animal Models of Parkinson’s Disease. Exp. Neurol. 2004, 187, 86–93. [Google Scholar] [CrossRef]
  144. Stan, T.L.; Alvarsson, A.; Branzell, N.; Sousa, V.C.; Svenningsson, P. NMDA Receptor Antagonists Ketamine and Ro25-6981 Inhibit Evoked Release of Glutamate in Vivo in the Subiculum. Transl. Psychiatry 2014, 4, e395. [Google Scholar] [CrossRef]
  145. Lesuis, S.L.; Lucassen, P.J.; Krugers, H.J. Early Life Stress Impairs Fear Memory and Synaptic Plasticity; a Potential Role for GluN2B. Neuropharmacology 2019, 149, 195–203. [Google Scholar] [CrossRef]
  146. Gao, X.; Chen, F.; Xu, X.; Liu, J.; Dong, F.; Liu, Y. Ro25-6981 Alleviates Neuronal Damage and Improves Cognitive Deficits by Attenuating Oxidative Stress via the Nrf2/ARE Pathway in Ischemia/Reperfusion Rats. J. Stroke Cerebrovasc. Dis. 2023, 32, 106971. [Google Scholar] [CrossRef]
  147. Acker, T.M.; Yuan, H.; Hansen, K.B.; Vance, K.M.; Ogden, K.K.; Jensen, H.S.; Burger, P.B.; Mullasseril, P.; Snyder, J.P.; Liotta, D.C.; et al. Mechanism for Noncompetitive Inhibition by Novel GluN2C/D N-Methyl-D-Aspartate Receptor Subunit-Selective Modulators. Mol. Pharmacol. 2011, 80, 782–795. [Google Scholar] [CrossRef]
  148. Wyllie, D.J.A.; Livesey, M.R.; Hardingham, G.E. Influence of GluN2 Subunit Identity on NMDA Receptor Function. Neuropharmacology 2013, 74, 4–17. [Google Scholar] [CrossRef]
  149. Pearlstein, E.; Gouty-Colomer, L.A.; Michel, F.J.; Cloarec, R.; Hammond, C. Glutamatergic Synaptic Currents of Nigral Dopaminergic Neurons Follow a Postnatal Developmental Sequence. Front. Cell. Neurosci. 2015, 9, 210. [Google Scholar] [CrossRef]
  150. Morris, P.G.; Mishina, M.; Jones, S. Altered Synaptic and Extrasynaptic NMDA Receptor Properties in Substantia Nigra Dopaminergic Neurons from Mice Lacking the GluN2D Subunit. Front. Cell. Neurosci. 2018, 12, 354. [Google Scholar] [CrossRef]
  151. Pálfi, E.; Lévay, G.; Czurkó, A.; Lendvai, B.; Kiss, T. Acute Blockade of NR2C/D Subunit-Containing N-Methyl-D-Aspartate Receptors Modifies Sleep and Neural Oscillations in Mice. J. Sleep Res. 2021, 30, e13257. [Google Scholar] [CrossRef]
  152. Chizh, B.A.; Headley, P.M.; Tzschentke, T.M. NMDA Receptor Antagonists as Analgesics: Focus on the NR2B Subtype. Trends Pharmacol. Sci. 2001, 22, 636–642. [Google Scholar] [CrossRef]
  153. He, H.; Yao, J.; Zhang, Y.; Chen, Y.; Wang, K.; Lee, R.J.; Yu, B.; Zhang, X. Solid Lipid Nanoparticles as a Drug Delivery System to across the Blood-Brain Barrier. Biochem. Biophys. Res. Commun. 2019, 519, 385–390. [Google Scholar] [CrossRef]
  154. Inês Teixeira, M.; Lopes, C.M.; Gonçalves, H.; Catita, J.; Margarida Silva, A.; Rodrigues, F.; Helena Amaral, M.; Costa, P.C. Riluzole-Loaded Lipid Nanoparticles for Brain Delivery: Preparation, Optimization and Characterization. J. Mol. Liq. 2023, 388, 122749. [Google Scholar] [CrossRef]
  155. Ortega Martínez, E.; Morales Hernández, M.E.; Castillo-González, J.; González-Rey, E.; Ruiz Martínez, M.A. Dopamine-Loaded Chitosan-Coated Solid Lipid Nanoparticles as a Promise Nanocarriers to the CNS. Neuropharmacology 2024, 249, 109871. [Google Scholar] [CrossRef]
  156. Chen, B.K.; Luna, V.M.; Shannon, M.E.; Hunsberger, H.C.; Mastrodonato, A.; Stackmann, M.; McGowan, J.C.; Rubinstenn, G.; Denny, C.A. Fluoroethylnormemantine, a Novel NMDA Receptor Antagonist, for the Prevention and Treatment of Stress-Induced Maladaptive Behavior. Biol. Psychiatry 2021, 90, 458–472. [Google Scholar] [CrossRef]
  157. Salabert, A.S.; Fonta, C.; Fontan, C.; Adel, D.; Alonso, M.; Pestourie, C.; Belhadj-Tahar, H.; Tafani, M.; Payoux, P. Radiolabeling of [18F]-Fluoroethylnormemantine and Initial in Vivo Evaluation of This Innovative PET Tracer for Imaging the PCP Sites of NMDA Receptors. Nucl. Med. Biol. 2015, 42, 643–653. [Google Scholar] [CrossRef]
  158. ReST Therapeutics. First-In-Human (FIH), Single Ascending Dose (SAD) Study of FluoroEthylNorMemantine (FENM). ClinicalTrials.gov. Available online: https://clinicaltrials.gov/study/NCT05921929?tab=table (accessed on 8 April 2024).
  159. Shram, M.J.; Henningfield, J.E.; Apseloff, G.; Gorodetzky, C.W.; De Martin, S.; Vocci, F.L.; Sapienza, F.L.; Kosten, T.R.; Huston, J.; Buchhalter, A.; et al. The Novel Uncompetitive NMDA Receptor Antagonist Esmethadone (REL-1017) Has No Meaningful Abuse Potential in Recreational Drug Users. Transl. Psychiatry 2023, 13, 192. [Google Scholar] [CrossRef]
  160. Egunlusi, A.O.; Malan, S.F.; Palchykov, V.A.; Joubert, J. Calcium Modulating Effect of Polycyclic Cages: A Suitable Therapeutic Approach Against Excitotoxic-Induced Neurodegeneration. Mini-Rev. Med. Chem. 2024, 24, 1277–1292. [Google Scholar] [CrossRef]
  161. Gutti, G.; Leifeld, J.; Kakarla, R.; Bajad, N.G.; Ganeshpurkar, A.; Kumar, A.; Krishnamurthy, S.; Klein-Schmidt, C.; Tapken, D.; Hollmann, M.; et al. Discovery of Triazole-Bridged Aryl Adamantane Analogs as an Intriguing Class of Multifunctional Agents for Treatment of Alzheimer’s Disease. Eur. J. Med. Chem. 2023, 259, 115670. [Google Scholar] [CrossRef]
  162. Zindo, F.T.; Barber, Q.R.; Joubert, J.; Bergh, J.J.; Petzer, J.P.; Malan, S.F. Polycyclic Propargylamine and Acetylene Derivatives as Multifunctional Neuroprotective Agents. Eur. J. Med. Chem. 2014, 80, 122–134. [Google Scholar] [CrossRef]
  163. Zhang, H.; Wang, Y.; Wang, Y.; Li, X.; Wang, S.; Wang, Z. Recent Advance on Carbamate-Based Cholinesterase Inhibitors as Potential Multifunctional Agents against Alzheimer’s Disease. Eur. J. Med. Chem. 2022, 240, 114606. [Google Scholar] [CrossRef]
  164. Knez, D.; Diez-Iriepa, D.; Chioua, M.; Gottinger, A.; Denic, M.; Chantegreil, F.; Nachon, F.; Brazzolotto, X.; Skrzypczak-Wiercioch, A.; Meden, A.; et al. 8-Hydroxyquinolylnitrones as Multifunctional Ligands for the Therapy of Neurodegenerative Diseases. Acta Pharm. Sin. B 2023, 13, 2152–2175. [Google Scholar] [CrossRef]
  165. Guo, J.; Wang, Z.; Liu, R.; Huang, Y.; Zhang, N.; Zhang, R. Memantine, Donepezil, or Combination Therapy—What Is the Best Therapy for Alzheimer’s Disease? A Network Meta-Analysis. Brain Behav. 2020, 10, e01831. [Google Scholar] [CrossRef]
  166. Iosifescu, D.V.; Jones, A.; O’Gorman, C.; Streicher, C.; Feliz, S.; Fava, M.; Tabuteau, H. Efficacy and Safety of AXS-05 (Dextromethorphan-Bupropion) in Patients With Major Depressive Disorder. J. Clin. Psychiatry 2022, 83, 22–27. [Google Scholar] [CrossRef]
  167. Akbar, D.; Rhee, T.G.; Ceban, F.; Ho, R.; Teopiz, K.M.; Cao, B.; Subramaniapillai, M.; Kwan, A.T.H.; Rosenblat, J.D.; McIntyre, R.S. Dextromethorphan-Bupropion for the Treatment of Depression: A Systematic Review of Efficacy and Safety in Clinical Trials. CNS Drugs 2023, 37, 867–881. [Google Scholar] [CrossRef]
  168. Yu, G.; Shi, Y.; Cong, S.; Wu, C.; Liu, J.; Zhang, Y.; Liu, H.; Liu, X.; Deng, H.; Tan, Z.; et al. Synthesis and Evaluation of Butylphthalide-Scutellarein Hybrids as Multifunctional Agents for the Treatment of Alzheimer’s Disease. Eur. J. Med. Chem. 2024, 265, 116099. [Google Scholar] [CrossRef]
  169. McKay, S.; Ryan, T.J.; McQueen, J.; Indersmitten, T.; Marwick, K.F.M.; Hasel, P.; Kopanitsa, M.V.; Baxter, P.S.; Martel, M.A.; Kind, P.C.; et al. The Developmental Shift of NMDA Receptor Composition Proceeds Independently of GluN2 Subunit-Specific GluN2 C-Terminal Sequences. Cell Rep. 2018, 25, 841–851.e4. [Google Scholar] [CrossRef]
  170. Sun, W.; Hansen, K.B.; Jahr, C.E. Allosteric Interactions between NMDA Receptor Subunits Shape the Developmental Shift in Channel Properties. Neuron 2017, 94, 58–64.e3. [Google Scholar] [CrossRef]
Figure 1. Neurodegenerative process associated with glutamate-induced excitotoxic cell death.
Figure 1. Neurodegenerative process associated with glutamate-induced excitotoxic cell death.
Pharmaceuticals 17 00639 g001
Figure 2. Schematic illustration of NMDA receptor activation.
Figure 2. Schematic illustration of NMDA receptor activation.
Pharmaceuticals 17 00639 g002
Figure 3. Structures of selected competitive NMDA receptor antagonists.
Figure 3. Structures of selected competitive NMDA receptor antagonists.
Pharmaceuticals 17 00639 g003
Figure 4. Schematic representation of competitive antagonism at NMDA receptor.
Figure 4. Schematic representation of competitive antagonism at NMDA receptor.
Pharmaceuticals 17 00639 g004
Figure 5. Structures of channel blockers or non-competitive NMDA receptor antagonists.
Figure 5. Structures of channel blockers or non-competitive NMDA receptor antagonists.
Pharmaceuticals 17 00639 g005
Figure 6. Schematic representation of NMDA receptor antagonism by channel blockers or non-competitive antagonists.
Figure 6. Schematic representation of NMDA receptor antagonism by channel blockers or non-competitive antagonists.
Pharmaceuticals 17 00639 g006
Figure 7. Negative allosteric modulators of NMDA receptors.
Figure 7. Negative allosteric modulators of NMDA receptors.
Pharmaceuticals 17 00639 g007
Figure 8. Illustration of negative allosteric modulators at NMDA receptor binding site.
Figure 8. Illustration of negative allosteric modulators at NMDA receptor binding site.
Pharmaceuticals 17 00639 g008
Figure 9. Non-competitive NMDA receptor antagonist presently in clinical trials.
Figure 9. Non-competitive NMDA receptor antagonist presently in clinical trials.
Pharmaceuticals 17 00639 g009
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Egunlusi, A.O.; Joubert, J. NMDA Receptor Antagonists: Emerging Insights into Molecular Mechanisms and Clinical Applications in Neurological Disorders. Pharmaceuticals 2024, 17, 639. https://doi.org/10.3390/ph17050639

AMA Style

Egunlusi AO, Joubert J. NMDA Receptor Antagonists: Emerging Insights into Molecular Mechanisms and Clinical Applications in Neurological Disorders. Pharmaceuticals. 2024; 17(5):639. https://doi.org/10.3390/ph17050639

Chicago/Turabian Style

Egunlusi, Ayodeji Olatunde, and Jacques Joubert. 2024. "NMDA Receptor Antagonists: Emerging Insights into Molecular Mechanisms and Clinical Applications in Neurological Disorders" Pharmaceuticals 17, no. 5: 639. https://doi.org/10.3390/ph17050639

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