Next Article in Journal
Formylation as a Chemical Tool to Modulate the Performance of Photosensitizers Based on Boron Dipyrromethene Dimers
Previous Article in Journal
Short Double-Stranded DNA (≤40-bp) Affects Repair Pathway Choice
Previous Article in Special Issue
An Integrated Multi-Omic Network Analysis Identifies Seizure-Associated Dysregulated Pathways in the GAERS Model of Absence Epilepsy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Potential Roles for the GluN2D NMDA Receptor Subunit in Schizophrenia

1
Department of Psychiatry, School of Clinical Sciences, Faculty of Medical, Nursing and Health Sciences, Monash University, Clayton, VIC 3168, Australia
2
Department of Neuroscience, Faculty of Medical, Nursing and Health Sciences, Monash University, Melbourne, VIC 3004, Australia
3
Mental Health Program, Monash Health, Clayton, VIC 3168, Australia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(14), 11835; https://doi.org/10.3390/ijms241411835
Submission received: 12 May 2023 / Revised: 19 July 2023 / Accepted: 22 July 2023 / Published: 23 July 2023
(This article belongs to the Special Issue Genetics and Molecular Mechanisms of Epilepsy and Drug Resistance)

Abstract

:
Glutamate N-methyl-D-aspartate receptor (NMDAR) hypofunction has been proposed to underlie schizophrenia symptoms. This theory arose from the observation that administration of NMDAR antagonists, which are compounds that inhibit NMDAR activity, reproduces behavioural and molecular schizophrenia-like phenotypes, including hallucinations, delusions and cognitive impairments in healthy humans and animal models. However, the role of specific NMDAR subunits in these schizophrenia-relevant phenotypes is largely unknown. Mounting evidence implicates the GluN2D subunit of NMDAR in some of these symptoms and pathology. Firstly, genetic and post-mortem studies show changes in the GluN2D subunit in people with schizophrenia. Secondly, the psychosis-inducing effects of NMDAR antagonists are blunted in GluN2D-knockout mice, suggesting that the GluN2D subunit mediates NMDAR-antagonist-induced psychotomimetic effects. Thirdly, in the mature brain, the GluN2D subunit is relatively enriched in parvalbumin (PV)-containing interneurons, a cell type hypothesized to underlie the cognitive symptoms of schizophrenia. Lastly, the GluN2D subunit is widely and abundantly expressed early in development, which could be of importance considering schizophrenia is a disorder that has its origins in early neurodevelopment. The limitations of currently available therapies warrant further research into novel therapeutic targets such as the GluN2D subunit, which may help us better understand underlying disease mechanisms and develop novel and more effective treatment options.

1. Schizophrenia—An Overview

Schizophrenia is a severe, debilitating, chronic neuropsychiatric disorder with a lifetime prevalence of 0.72% [1]. It has a complex and heterogeneous presentation making it difficult to diagnose and to identify a consistent underlying aetiology or pathology. People with schizophrenia demonstrate high inter-individual variability with respect to symptoms, disease course, outcome and response to treatment. Schizophrenia symptoms are divided into three main classes: positive, negative and cognitive deficits. Of the three, positive symptoms are the most easily identifiable and are defined as psychotic features that are not usually present in healthy people, and include hallucinations, delusions and disorganised speech and behaviour [2,3]. Negative symptoms refer to a reduction or disruption in normal emotions and behaviours manifesting as social and emotional withdrawal, apathy and avolition [2,4]. Cognitive symptoms vary in severity amongst people with schizophrenia and include elements such as deficits in verbal memory, working memory, attention, executive functioning, cognitive flexibility and processing speed [5].
Schizophrenia typically has an early onset, with most people being diagnosed in their late adolescence—early adulthood years. The onset of the first psychotic episode is usually preceded by a prodromal period during which symptoms gradually emerge and this period can last several years. An early onset combined with long-term deficits in social, educational and occupational function make this disorder one of the leading causes of chronic disability with significant impacts on the quality of life of patients and their families and caregivers [6,7]. Additionally, this disorder places a substantial economic burden on health care systems with estimates of annual associated costs reaching up to USD 150 billion in the United States and AUD 4.9 billion in Australia [8,9].
The causes of schizophrenia are not fully understood but are thought to be multifactorial, involving a complex interplay between multiple genetic variants and environmental factors. Genome-wide association studies have identified multiple common variants of small effect spanning over 250 genetic loci, suggesting that schizophrenia is a polygenic disorder in most cases [10,11]. Genes associated with schizophrenia risk are involved in various functions, including the regulation of the postsynaptic membrane, synaptic transmission and neurodevelopmental pathways, including glutamate pathways [11,12,13]. Non-genetic factors that increase lifetime risk for schizophrenia include obstetric complications, advanced paternal age, living in an urban setting, childhood trauma or adversity, cannabis use and first-generation migration [14,15,16,17,18,19,20,21]. Although there have been advances in our understanding of risk factors associated with schizophrenia, the aetiological complexity has made it a challenge to identify the underlying disease mechanisms and find effective cures and preventative strategies.
Currently available pharmacological treatments, chiefly conventional and atypical anti-psychotics, and psychotherapy, have proven clinical utility and can help manage positive symptoms in some people [22,23]. However, antipsychotic medications have limited efficacy and are poorly tolerated, with substantial side effects in approximately 30% of people. Moreover, they usually offer little benefit in improving negative and cognitive symptoms, and these symptom types therefore remain a pressing, unmet medical need [22,24]. The prevalence, burden and current lack of effective treatments for schizophrenia highlight a need to improve our understanding of the underlying mechanisms and neurobiology of the disorder in order to identify and develop better treatments.
Although the predominant literature has focused on dopaminergic dysfunction in schizophrenia [25,26,27,28], and many of the currently available anti-psychotics target dopaminergic receptors [29,30,31,32], the limitations of these treatments suggests that other pathological processes may be involved. A theory gaining increasing support for the aetiology of schizophrenia is the hypofunction of one of the major glutamate receptor subtypes, the N-methyl-D-aspartate receptor (NMDAR) [33,34,35,36].

2. Glutamatergic Signalling in the Central Nervous System

Glutamate plays a key role in mediating the homeostatic balance between excitation and inhibition in the brain, cortico-cortical neurotransmission, neuronal development, neurodegeneration, nervous system plasticity and learning and memory [37]. Glutamate carries out its actions through its receptors which are divided into two groups: the ionotropic and metabotropic receptors [38,39]. The ionotropic receptors, namely the NMDA, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) and kainate receptors, are integral membrane proteins composed of four large subunits (>900 residues) and act as ligand-gated cation channels [40]. The metabotropic receptors are G-protein-coupled and activate intracellular biochemical cascades [39,41]. As this review is focused on NMDARs, specifically the GluN2D subunit and its relevance to schizophrenia, we direct readers to the following review articles [42,43,44] for further discussion of the role of AMPA, kainate and metabotropic receptors in schizophrenia.

NMDA Receptor Structure and Function

NMDARs are widely distributed and can be found on both neuronal and non-neuronal cells. These receptors play a key role in many physiological processes, including neurodevelopment, synaptogenesis and locomotion, and due to their role as critical mediators of activity-dependent synaptic plasticity, they are especially important for learning, memory formation and other forms of cognition [45,46,47,48,49,50,51]. NMDARs are unique in that they require the concomitant binding of glycine and glutamate along with membrane depolarisation for activation and are thus referred to as ‘molecular coincidence detectors’ [52,53,54]. At resting membrane potential, extracellular magnesium (Mg2+) can be found in the ion channel pore, blocking NMDARs in a voltage-dependent manner. Partial depolarisation of the neuron relieves the blockade, opening the channel. The subsequent Ca2+ influx into the neuron triggers a cascade of events that can influence both local, acute functional synaptic plasticity and, via changes in gene expression, sustained neural plasticity [55].
The heteromeric composition of NMDARs enables different pharmacological, biophysical and functional properties for the receptor, and these heteromers vary in distribution and expression, both regionally and temporally, throughout development. The different subunits that make up the NMDAR are termed GluN1, GluN2 and GluN3. The GluN1 subunit is encoded by a single gene but has eight splice variants [56]. There are four different GluN2 subunits (A–D), encoded by four different genes and two different Glun3 subunits (A and B) [39,57]. A functional NMDAR is typically composed of two GluN1 subunits and two subunits from among the GluN2A-D and GluN3A-B subunits [45,58]. The obligatory GluN1 subunit is ubiquitously expressed throughout the brain and over the lifespan [39,59]. Of the GluN2 subunits, which have a more varied and complex temporal and spatial expression, GluN2A and GluN2B are the predominant subtypes found in the adult human brain, whilst the GluN2C and GluN2D subunits are more highly expressed in the developing brain [45,60]. The subunit composition of NMDARs influences its functional properties, including agonist affinity, Mg2+ block, decay kinetics and modulation by polyamines [61,62]. Given the importance of NMDAR subunits in mediating normal brain function, it is not surprising that the dysfunction of these subunits has been linked to various neurological diseases, including schizophrenia. The NMDAR hypofunction model is one of the most commonly adopted and supported models of schizophrenia and is often employed to study the aetiology and pathology of the disorder as well as for the development of novel treatment strategies.

3. NMDA Receptor Hypothesis of Schizophrenia

From the late 1950s, phencyclidine (PCP) and ketamine have been reported to induce positive, negative and cognitive phenotypes such as psychosis-like dissociative states and neurocognitive disturbances in healthy individuals like those observed in people with schizophrenia [63,64,65,66,67,68]. Furthermore, these drugs exacerbate symptoms, including psychosis in individuals with schizophrenia [66,69,70]. The NMDAR hypothesis was first proposed soon after, in the late 1970s–1980s, when it was found that PCP and ketamine carry out their actions via NMDAR blockade [64,71,72,73]. The extensive literature drawing parallels between NMDAR antagonism and many of the unique features of schizophrenia have led some to postulate that these antagonists might act via the same mechanisms that become dysfunctional in this disorder [74,75,76]. In support of this, studies in rodents have shown that knockdown of the obligatory GluN1 subunit of the NMDAR can result in phenotypes relevant to schizophrenia, including hyperlocomotion (proposed to be indicative of a striatal hyperdopaminergic state which underlies positive symptoms such as psychosis), anhedonia, social deficits, impaired learning and abnormal neuronal oscillations [77,78,79,80,81,82].
Although NMDAR hypofunction has been linked to schizophrenia symptoms, the precise underlying mechanisms are still unclear. One hypothesis is that it is primarily the dysfunction of NMDARs on GABAergic interneurons, rather than more widespread NMDAR dysfunction, which contributes to the molecular, physiological and behavioural characteristics of schizophrenia [35,83]. GABAergic interneurons are stimulated by postsynaptic NMDAR activation and, in turn, synapse onto excitatory glutamatergic pyramidal cells in a negative feedback loop. GABAergic interneurons connect to hundreds of pyramidal cells in this manner, enabling them to coordinate synchronised network activity throughout the brain, including the hippocampus. The activity of these glutamatergic pyramidal cells, in turn, drives downstream striatal dopaminergic neurons (Figure 1) [84,85]. In this way, any disruption to normal NMDAR function would lead to downstream GABAergic and dopaminergic dysregulation. In line with this theory, studies conducted in healthy humans and in rodent models show that administration of NMDAR antagonists counterintuitively results in the hyperactivity of cortical pyramidal neurons and enhanced cortical glutamate release by disinhibition [86,87,88,89,90,91,92,93,94,95]. This is corroborated by brain imaging data in both humans and rodents showing that the administration of an NMDAR antagonist results in net excitation and decreased coordinated burst-firing in the prefrontal cortex, which is linked to cognitive deficits [89,96,97,98,99]. This cortical disinhibition suggests that NMDAR antagonists may be preferentially targeting the NMDARs on GABAergic interneurons. Additionally, NMDAR antagonists have been shown to reproduce schizophrenia-like dopaminergic dysfunction, including an increase in striatal D2 receptors and dopamine release [100,101,102,103]. This striatal hyperdopaminergic state has been linked to the positive symptoms of schizophrenia. In this way, NMDAR hypofunction might be resulting in a cortical excitatory/inhibitory (E/I) imbalance, affecting the ability of interneurons to synchronise firing across neural networks, and driving the changes in the dopaminergic system, ultimately culminating in the clinical symptoms of schizophrenia.
In recent years, there has been mounting evidence of the dysfunction and/or dysregulation of NMDAR-mediated neurotransmission in schizophrenia. Several schizophrenia candidate genes that affect NMDAR signalling have been identified, including GRIN1, GRIN2A, GRIN2B and GRIN2D, which are genes encoding the GluN1, GluN2A, GluN2B and GluN2D subunits of the NMDAR, respectively, as well as the gene encoding D-amino acid oxidase (DAAO), which regulates the availability of D-serine, a co-agonist at NMDARs [11,104,105,106]. Findings from other genetic and post-mortem patient studies have further suggested that NMDAR signalling is disrupted in schizophrenia [33,107,108,109].
The contribution of the GluN1, GluN2A and GluN2B subunits to schizophrenia pathology and symptoms has already been well researched and reported (see [34,110,111,112]). This review will instead focus on the relatively understudied GluN2D subunit, as its unique developmental expression pattern, cellular specificity and electrophysiological properties suggest that it may play a role in the NMDAR hypothesis of schizophrenia. To this end, we firstly outline some of these unique physiological features, and then describe the changes observed in this subunit in people with schizophrenia. Finally, we discuss how alterations in the expression and function of the GluN2D subunit may contribute to schizophrenia.

4. GluN2D Subunit

4.1. GluN2D Receptor Subunit Expression and Distribution

The GluN2D (common names: NMDAR2D, NR2D, GluRε4) subunit is encoded by the GRIN2D gene, which consists of 13 exons, spanning 49.3kB, and is located on chromosome 19q13.1-qter in the human genome [39]. The GRIN2D gene has two potential splice isoforms (NR2D-1 and NR2D-2), the longest of which contains 1356 amino acids [39,113]. In rodents, GluN2D has been extensively characterized; expression levels are first detected between embryonic day (E)15 and 18 during late-embryogenesis, with levels peaking by post-natal day (P)7–10 [114,115]. The GluN2D mRNA and protein expression levels decrease gradually after the early neonatal phase until late adolescence (P40–50), when they reach their relatively low steady-state expression level [61,114,115,116]. During the embryonic and early neonatal phases, the expression of GluN2D is widespread and detected in several regions, including the spinal cord, midbrain nuclei, diencephalon (thalamus, hypothalamus), certain basal ganglia nuclei (substantia nigra and subthalamic nucleus), retina, olfactory bulb, cerebellum, cerebral cortex and hippocampus [60,61,113,114,117]. The ubiquitous distribution of the GluN2D subunit during the early phases of life makes this subunit particularly interesting amongst the NMDAR subunits as it suggests that GluN2D plays a critical role in modulating circuit connectivity and function during neurodevelopment. This could be of significance when considering a disorder such as schizophrenia that is thought to have its origins in early development [118,119].
As the rodent ages and GluN2D expression levels reduce, it becomes more localized to distinct cell subtypes [61,114,115]. This is especially apparent in the hippocampus and cortex, where multiple studies have shown that GluN2D is enriched in parvalbumin (PV)-containing GABAergic interneurons in mature rodents, whereas its expression and activity on glutamatergic pyramidal cells decreases [117,120,121,122,123]. Electrophysiological analyses from the adult mouse medial prefrontal cortex (mPFC) showed that a GluN2C/D positive allosteric modulator, CIQ(+), increased the intrinsic excitability of interneurons and enhanced excitatory postsynaptic currents (EPSCs) from interneurons, whilst not having any effect on the surrounding pyramidal cells [121]. Another study found that while CIQ(+) enhanced EPSCs in WT mice, this effect was missing in GluN2D KO mice [122]. The sustained expression of GluN2D-containing receptors on interneurons over development may suggest that this subunit is required to maintain proper inhibitory drive and control overall circuit function [124,125].
In humans, the spatial and temporal expression of GluN2D is thought to be consistent with that reported in rodents; however, it is yet to be well characterised. In human foetal brains, GluN2D mRNA is abundantly expressed and is one of the predominant NMDAR subunits expressed between gestational weeks 8 and 20 [126]. In contrast, a study of neurologically normal, adult human post-mortem brains reported only moderate expression of GluN2D mRNA in the prefrontal, parietal and motor cortices, where instead the major subtypes expressed were the GluN2A and GluN2B subunits [127]. This study also reported that whilst GluN2D expression was low within most neurons in the hippocampus, expression was moderately intense within a small subset of hippocampal neurons, particularly in the hilus, a region containing many interneurons [127]. A recent study which used laser microdissection to isolate pyramidal neurons and PV+ interneurons from the human post-mortem dorsolateral PFC (dlPFC), combined with RNA sequencing and microarray, provided evidence that the GluN2D subunit is indeed particularly enriched in PV+ interneurons, but not pyramidal neurons [121]. It should be noted, however, that whilst the interneurons that were isolated for this study expressed markers specific to PV+ interneurons, such as KCNS3, they also expressed markers common to other interneuron types, such as somatostatin, and thus might not be delimited to PV+ interneurons alone [121]. Another post-mortem in situ hybridization study revealed that in control human brains, the GluN2D subunit is preferentially expressed in layers II, IIIa-c, IV, Va and VIa of the cerebral cortex and at much higher levels in the prefrontal cortex compared with the parieto-temporal or cerebellar cortices [128].
Multiple studies examining the subunit composition of NMDARs in the central nervous system suggest that while the majority of GluN2D subunits are associated with the GluN1 subunit, the GluN2D subunit also forms heteromeric assemblies with GluN2A and/or GluN2B subunits in different brain regions and neuronal subpopulations [129,130,131]. For example, the GluN2D subunit is shown to primarily form a binary complex with GluN1 in the thalamus of young (post-natal day 7) and adult rats, whereas this binary complex is not detected at all in the rat midbrain, where instead the GluN2D subunit forms triheteromeric assemblies with GluN1 and GluN2A or GluN2B [129]. The varying spatial expression of the different subtypes of GluN2D-containing NMDARs suggest distinct functional roles.

4.2. GluN2D Receptor Subunit Function

Glutamate displays 5–6 times greater potency at GluN2D-containing NMDARs than GluN2A- or GluN2B-containing NMDARs [39,132]. Similarly, GluN1 agonists such as glycine are most potent when the GluN2 subunit in the NMDAR is GluN2D [133]. GluN2D-contaning NMDARs also have a weak Mg2+ block that is 10-fold lower than that of GluN2A- or GluN2B-containing receptors and are also reported to have a slightly lower Ca2+ permeability [134,135]. The resistance to Mg2+ block suggests that neurons expressing the GluN2D subunit may be more responsive to synaptic glutamate release. The non-competitive NMDAR antagonist ketamine is more potent at, and shows approximately five-fold selectivity for GluN2D-containing receptors, compared with GluN2A or GluN2B subunits [136]. This, combined with the fact that GluN2D-containing receptors are predominantly expressed on interneurons, may suggest that the GluN2D-containing receptors might be involved in the cortical disinhibition induced by certain NMDAR antagonists such as ketamine [121,137]. Another unique feature of GluN2D-containing receptors is that they have the slowest deactivation rate of the GluN2 receptor subtypes, exhibiting 10-fold slower receptor deactivation than GluN2B and GluN2C receptors and 100-fold slower than GluN2A receptors [39,62,138]. Application of glutamate to NMDARs containing the GluN2A subunit was found to generate a macroscopic current with a deactivation time of tens of milliseconds compared with several seconds for GluN2D-containing NMDARs [138]. Importantly, this prolonged decay time is matched with an almost complete lack of desensitization, giving this subunit the capacity to shuttle large quantities of ions across the plasma membrane in order to sustain depolarization for a long enough period to initiate burst-firing [61,138,139]. This ability to generate burst-firing would be especially important for the interneurons on which they are expressed, possibly at least partially contributing to their rhythmic, oscillatory nature [137]. Thus, the unique electrophysiological properties of the GluN2D subunit, including resistance to Mg2+ blockade and a remarkably long channel decay latency, might prove important for the integrative functions of the interneurons where this subunit is predominantly expressed. However, it should be noted that this receptor has a very low open channel probability of 1–4% and, as mentioned previously, lower Ca2+ permeability than the other GluN subtypes and is also considered a low-conductance NMDAR, which might imply that it is less effective at depolarizing the postsynaptic membrane and triggering downstream signalling events [39,140].
GluN2D-containing receptors play a role in both presynaptic and postsynaptic neurotransmission. Studies have reported GluN2D-containing NMDARs on interneurons in the hippocampus and neocortex where they play a key role in postsynaptic signalling [122,141]. GluN2D-containing receptors are thought to enable interneurons to synchronize and coordinate the firing of large groups of cortical pyramidal neurons. A recent study showed that tonic activation of these GluN2D-containing NMDARs on developing cortical interneurons is required for proper intrinsic excitability, dendritic arborization, GABAergic synaptogenesis and inhibitory tone onto excitatory pyramidal cells [141]. GluN2D-containing NMDARs have also been reported on the dendrites of neurons in the subthalamic nucleus, on dopaminergic projection neurons in the substantia nigra pars compacta and in the dorsal horn of the spinal cord, where they contribute to the modulation of the indirect pathway, mediate dopamine release to the striatum and play a role in pain perception, respectively [142,143]. GluN2D-containing NMDARs expressed at presynaptic sites are thought to play a modulatory role in the hippocampus and cerebellum. In the hippocampus, GluN2D, along with GluN2B and postsynaptic metabotropic glutamate receptors, have been shown to be critical for the induction of spike-time-dependent LTD [144]. This is a form of synaptic plasticity that occurs when the timing and order of spikes from the pre- and postsynaptic neurons leads to a long-lasting weakening of the synaptic connection between them, and is important for refining synapses and neural circuits both during development and learning and memory in adults [144,145]. In the cerebellum, presynaptic GluN2D-containing NMDARs help fine tune the release of GABA onto Purkinje cells [146].
Interestingly, it has been reported that the human GRIN2D gene contains four estrogen-responsive elements which are highly preserved in the rat, suggesting that the GluN2D subunit might be under neuroendocrine control [147]. In line with this, a study using ovariectomised rats found an upregulation of GluN2D mRNA in the hypothalamus following 17β-estradiol treatment [147]. Given the well-established links between fluctuating levels of 17β-estradiol and schizophrenia onset and symptom severity [148], it is intriguing to consider that one of the actions by which 17β-estradiol may exert its effects is via regulation of GluN2D and other NMDAR subunits [149].
Overall, considering its expression peaks during development, and that it is predominantly located on interneurons where its unique properties likely contribute to the integrative function of the interneurons, it is highly likely that the GluN2D receptor subunit plays an important role in the NMDAR hypothesis of schizophrenia.

5. Alterations to GluN2D in Schizophrenia

There is evidence to suggest that the GluN2D subunit is altered in schizophrenia. A study of approximately 200 Japanese people with schizophrenia found single nucleotide polymorphisms (SNPs) in the gene for the GluN2D receptor that might contribute to schizophrenia susceptibility [150]. They report that specific combinations of four SNPs within the GRIN2D gene were significantly associated with schizophrenia. These specific combinations were found in three pairs of SNPs: INT10SNP–EX13SNP2, EX13SNP2–EX13SNP3 and EX6SNP–EX13SNP2 [150]. A recent mutation-screening study also identified an ultra-rare, loss-of-function splice-site mutation (c.1412G>A) in the exonic region of the GRIN2D gene, which may lead to the creation of a truncated, nonfunctional GluN2D receptor, thereby contributing to schizophrenia risk [151]. This study additionally found four missense mutations in schizophrenia patients in the GRIN2D gene, and although the actual functional impact of these amino acid substitutions was not examined, in silico analysis classified each of the four variants as disease-causing based on their predicted effect on protein function [151]. An in situ hybridization study on post-mortem human tissue reported a 53% increase in the proportion of GluN2D mRNA expression in the PFC of people with schizophrenia (see Table 1) [128]. This increase in the proportion of GluN2D expression was not seen in anti-psychotic-treated control brains, proving that the change was specific to schizophrenia and could not be attributed to treatment with anti-psychotics alone [128]. However, in another in situ hybridization study, no change in the mRNA levels of the GluN2D subunit was observed in the dlPFC of people with schizophrenia [152]. Furthermore, in a post-mortem Western blot study, no change in the protein expression of the GluN2D subunit was observed in the dlPFC and anterior cingulate cortex in elderly people with schizophrenia when compared with a control group [153]. In one human post-mortem study, laser microdissection was used to isolate a population of glutamatergic relay neurons and another mixed glial and GABAergic interneuron population (including PV+ interneurons) from the medial dorsal thalamus of people with schizophrenia and controls [154]. This study revealed decreased expression of the GluN2D transcript in glutamatergic relay neurons in the medial dorsal thalamus, whilst no changes were observed in the mixed glia and interneuron population in schizophrenia [154]. These relay neurons project to the PFC, suggesting that alterations to GluN2D in schizophrenia might contribute to the disruption of the thalamocortical circuit thought to be involved in the attention and sensory processing deficits and negative symptoms seen in schizophrenia [155,156,157]. Another study of a small sample size (n = 19) found that GluN2D gene expression was increased in the right cerebellum of post-mortem human schizophrenia brains compared to controls [158]. However, this study included people with schizophrenia who had been taking anti-psychotic medications chronically, which could have contributed to any differences observed compared to controls. These discrepant results highlight the need for further studies examining the region- and cell-specific expression and function of the GluN2D subunit in the brains of people with schizophrenia. In the following sections, the expected consequence of mutations in GRIN2D and changes in GluN2D expression are discussed.

6. Consequences of Loss of GluN2D Function

Strategies to explore the influence of GluN2D function include using animal models and pharmacological tools to study behavioural and physiological measures which have relevance to schizophrenia.

6.1. Genetic Models

GluN2D-knockout (KO) mice are viable, reproduce and grow normally, and have no overt changes in neuronal histology [159]. Moreover, mRNA levels of the other NMDAR subunits are unaffected in these mice [159]. These mice, however, exhibit unique behavioural phenotypes, including diminished spontaneous motor movements in open-field tests (Table 2) [159,160]. GluN2D-KO mice also display deficits in spatial learning and memory, as well as impaired contextual fear memory, but show no deficits in the novel object recognition task [137,161,162]. Most studies report no abnormalities in motor function as measured by the rotarod test, nor any differences in anxiety when compared with WT mice during the light–dark compartment test and elevated maze test [159,163]. However, Miyamoto et al. report reduced sensitivity to stress induced by the elevated-plus maze, the light–dark compartment test and forced swim tests in these KO mice [160]. The discrepancies in these findings might be due to the differences in the way the tests were performed, which could be sensitive to subtly different anxiety- or fear-related behaviours. This study also reported a downregulation of dopamine, serotonin, norepinephrine and their metabolites in the hippocampus and striatum of GluN2D-KO mice [160]. GluN2D-KO mice are reported to have a normal prepulse inhibition response, which is known to be deficient in schizophrenia [137,164]. This suggests that they might not model all aspects of schizophrenia. Interestingly, the hyperlocomotion-inducing effects of PCP and ketamine are reduced in GluN2D-KO mice, indicating that the GluN2D subunit plays an important role in mediating the effects of these drugs [137,163,165]. Additionally, the motor impairments, locomotor sensitization and increase in extracellular dopamine seen following PCP or ketamine administration in WT mice is not seen in GluN2D-KO mice [163,165,166]. These data suggest that the GluN2D subunit mediates the hyperdopaminergic-inducing effects of the NMDAR antagonists, PCP and ketamine. Autoradiography studies revealed that the increase in neuronal activity induced by sub-anaesthetic ketamine in the medial prefrontal cortex, entorhinal cortex, presubiculum and caudate putamen in WT mice is not seen in GluN2D-KO mice [98,137,167].
GluN2D-KO mice were reported to have similar basal neural oscillatory power between frequency ranges 30 and 200 Hz when compared with WT mice [168]. However, whilst the administration of the NMDAR antagonists MK-801, ketamine and memantine increased oscillatory power in WT mice, they had very little effect on GluN2D-KO mice, especially at the high gamma frequency range (65–140 Hz) [168]. Similarly, Sapkota et al. reported that ketamine increased gamma frequency (30–140 Hz) oscillatory power in WT mice but elicited a much less pronounced increase in GluN2D-KO mice, which was more apparent at the high frequency (>60 Hz) range [137]. This suggests a role for the GluN2D subunit in modulating the high-frequency neural network oscillations induced by NMDAR antagonists. The study by Sapkota et al. also found reduced PV cell density in the substantia nigra and basolateral/lateral amygdala and a trend towards a reduction in the mPFC and hippocampus in GluN2D-KO mice compared with WT mice [137].
These studies suggest the GluN2D subunit expressing NMDA receptors are critical in mediating many of the effects of NMDAR antagonists, like ketamine, on the behaviour and electrophysiology that is relevant to schizophrenia, and thus may, in turn, suggest a role for the GluN2D subunit in the development of schizophrenia. There is evidence to suggest that GluN2D-containing NMDARs can contribute to neuronal networks that underlie cognition and which are found to be disrupted in schizophrenia. GluN2D-mediated signalling could be promising as a potential therapeutic target for specific symptoms observed in individuals with schizophrenia.

6.2. Pharmacological Manipulations

Although currently, to the best of our knowledge, there are no GluN2D-selective drugs available, competitive antagonists with 3–10-fold higher selectivity for GluN2C/GluN2D compared with GluN2A/GluN2B-containing NMDARs have been developed. One such compound, (2R*,3S*)-1-(phenanthrenyl-2-carbonyl)piperazine-2,3-dicarboxylic acid (PPDA), resulted in more potent inhibition of LTD than LTP in rat hippocampal slices, suggesting a role for the GluN2D (and GluN2C) NMDAR subunits in hippocampal LTD [169]. Another study using PPDA showed that GluN2D-containing NMDARs also contribute to extrasynaptic currents in rat CA1 neurons [170]. A recent study used an analogue of PPDA, (2R*,3S*)-1-(9-bromophenanthrene-3-carbonyl)piperazine-2,3-dicarboxylic acid (UBP145), in combination with GluN2D-KO mice, to show that GluN2D-containing NMDARs play a role in short-term potentiation (STP) as well as LTP in the mouse hippocampus [171]. This study reported that UBP145 partially inhibited LTP and the slow component of STP in WT but not GluN2D-KO mice [171]. Whilst the function of STP is lesser known than LTP, it has been hypothesized that it might play a role in forms of short-term memory, such as working memory which is known to be disrupted in schizophrenia [172]. These data together suggest that GluN2D-containing NMDARs contribute to synaptic plasticity and thus cognitive processes in complex ways. Therefore, it is possible that dysregulation of these receptors is involved in the cognitive dysfunction seen in schizophrenia.

7. How Might Alterations to the GluN2D Subunit Contribute to Schizophrenia?

Several characteristics of the GluN2D subunit, including peak expression early in development, its localization to PV+ interneurons in the cortex and hippocampus and, additionally, reports of alterations to this subunit in post-mortem tissue from people with schizophrenia, and evidence of GRIN2D being a schizophrenia candidate gene, suggest that it may be involved in schizophrenia pathology. Below, we propose a mechanism by which dysfunction of the GluN2D receptor could contribute to schizophrenia.

7.1. GluN2D Subunit and Parvalbumin-Positive GABAergic Interneurons

As previously discussed, inhibitory interneurons have been identified as the key locus or point of convergence of the glutamatergic, GABAergic and dopaminergic hypotheses of schizophrenia and are also implicated in the cognitive deficits seen in schizophrenia. Interestingly, in situ hybridization, electrophysiology and immunohistochemistry studies have revealed that GluN2D-containing NMDARs are specifically enriched in the PV-expressing subclass of interneurons in the hippocampus and PFC, two regions that underlie learning and memory function (refer to Section 4.1) [120,121,122,141]. This makes the GluN2D subunit particularly intriguing in the context of schizophrenia as several studies have shown that the hypofunction of NMDARs at fast-spiking PV-containing interneurons is sufficient to produce schizophrenia-like symptoms, including cognitive dysfunction [84,173,174,175].
Despite there being more than 20 different classes of GABAergic interneurons [176], it is the interneurons containing the calcium-binding protein, PV, that have been proposed to be especially important in schizophrenia [177,178]. Not only are PV-containing interneurons crucial for regulating cortical inhibition via the pyramidal neurons they innervate, but also for the generation of synchronous gamma-frequency oscillations [178,179,180,181,182,183]. Gamma oscillations are synchronous electrophysiological brain rhythms in the gamma frequency range (30–80 Hz) that are crucial for information processing and appropriate cortical function and underpin a wide range of cognitive processes, including those disrupted in schizophrenia like working memory [184,185,186]. Abnormal gamma-frequency synchrony is a major pathological characteristic of schizophrenia and underlies cognitive deficits [186]. For example, a recent study found lower-amplitude gamma oscillations in people with schizophrenia while they were performing a working memory task [187]. Similarly, in another cohort of people with schizophrenia, gamma-band activity was reduced during the sensory processing state during an auditory task, but baseline gamma power during the resting state was increased when compared with healthy controls [188,189]. Additionally, NMDA antagonists like ketamine and PCP also induce these same gamma oscillatory disturbances in healthy humans and rodent models of schizophrenia [95,190,191,192,193,194,195]. Thus, it is possible that dysfunction of the GluN2D subunit, which is especially enriched in PV cells in the cortex and hippocampus, might contribute to the abnormal gamma oscillations and the associated cognitive deficits seen in schizophrenia. Indeed, studies have found that genetic ablation of the GluN2D subunit from mice significantly reduces NMDAR antagonist-induced high-frequency gamma oscillations [137,168]. Alterations of the GluN2D subunit could affect the ability of PV cells to mediate the inhibition of excitatory pyramidal cells in a synchronised manner, disrupting the E/I balance and resulting in abnormal neural oscillations. Indeed, a recent study has found that conditional deletion of the GluN2D subunit from PV interneurons resulted in hyperexcitation of the medial PFC and disruptions to the feedforward inhibitory circuit [196]. Behaviourally, these mice exhibited hyperlocomotion, increased anxiety-like behaviour and impaired short-term memory and cognitive flexibility [196]. This study also found that GluN2D ablation from PV interneurons resulted in a downregulation of genes involved in GABAergic and dopaminergic synapse function such as GAD67 and TH, as well as schizophrenia susceptibility genes such as Disc1, ErbB4 and their downstream targets [196], thus suggesting a critical role for the GluN2D subunit in PV cells in modulating schizophrenia-relevant changes in neural circuitry, signalling and behaviours. The selective expression of the GluN2D subunit in PV+ interneurons suggests it may be a potential therapeutic target that could reverse interneuronal hypofunction and the currently untreated cognitive impairments that result from these deficits. Positive modulation of the GluN2D-subunit-containing NMDA receptor indeed increases the firing rates and restores the GABAergic network stability and reversed working memory deficits in a mouse model of schizophrenia symptoms [121,122,197]. However, as the GluN2D subunit is so critical to development, it may prove a challenge to reverse any deficits. A recent study showed that tonic activation of GluN2C/GluN2D-containing receptors during development is needed for proper cortical PV interneuron morphological maturation and complexity, circuit integration and maintaining proper inhibitory tone onto excitatory pyramidal cells [141]. The study found that blockade of GluN2C/GluN2D-containing receptors between post-natal days 7 and 9 was sufficient to cause long-lasting cortical inhibitory network deficits, as seen in schizophrenia [141].
Cortical pyramidal neurons innervated by PV cells stimulate dopaminergic neurons in the midbrain which project to the associative striatum. The associative striatum includes the rostral and dorsal part of the caudate nuclei and is implicated in the pathophysiology of schizophrenia [198,199,200,201]. The associative striatum is rich in dopamine receptors and dopamine afferents and receptors and is thus thought to be the primary site of action of antipsychotics. As such, any disruption to the GluN2D-containing NMDARs on PV cells could also indirectly affect midbrain dopaminergic neurons and lead to enhanced dopamine release in the striatum which has been linked to the positive symptoms of schizophrenia [202]. Following treatment with ketamine, Yamamoto et al. found an increase in locomotor activity and nitric oxide (NO) synthesis in the dendrites of medium spiny neurons in the dorsal striatum and PFC in WT but not GluN2D-KO mice [166]. Postsynaptic neuronal NO synthesis is functionally coupled to the stimulation of NMDARs. The failure of ketamine to induce an increase in striatal NO synthesis in GluN2D-KO mice provides support for the role of GluN2D-containing receptors in the corticostriatal neuronal circuit. Similarly, PCP induced an increase in the number of Fos-positive (marker of neuronal activity) cells in the striatum, prefrontal cortex, thalamus and subthalamic nuclei of WT but not GluN2D-KO mice [165]. This suggests that the frontal cortex–dorsal striatum pathway is likely activated by the inhibition of GluN2D-containing NMDARs.

7.2. GluN2D Subunit and Dopaminergic Neurons

Multiple studies have found that the GluN2D subunit forms functional NMDAR channels in the substantia nigra pars compacta dopaminergic neurons [131,203,204,205]. The substantia nigra plays an essential role in modulating motor movement and reward functions. Interestingly, GluN2D-KO mice have a hypolocomotor phenotype and the hyperlocomotor effects of PCP and ketamine are reduced in GluN2D-KO mice [137,163,166]. The amount of dopamine release in the forebrain following PCP treatment is also reduced in GluN2D-KO mice [163]. This suggests that GluN2D-containing NMDARs might play either a direct or indirect role in modulating dopaminergic function and, consequently, locomotor activity. In people with schizophrenia, there are reports of long-term deficits in basic motor function and control [206,207,208]. Thus, it is possible that dysfunction of the GluN2D subunit could affect burst-firing in these dopaminergic neurons, disrupting their function in the nigrostriatal circuitry, which is hypothesised to underlie motor symptoms in schizophrenia [209]. Impairments in reward processing have also been reported in people with schizophrenia and may underlie some of the negative symptoms in schizophrenia, such as anhedonia and a lack of motivation [210,211]. A study by White et al. reported that glutamate in the substantia nigra plays a role in reward processing and found that glutamatergic dysfunction in the substantia nigra could contribute to reward-processing deficits [212]. Thus, any alterations to GluN2D-containing NMDARs in the substantia nigra might also play a role in reward-processing deficits. Indeed, Yamamoto et al. reported that ablation of the GluN2D subunit in mice resulted in an anhedonic state, indicated by a reduction in sucrose preference [166]. Additionally, loss of GluN2D-containing NMDARs resulted in anxiety- and depressive-like behaviours in mice which was linked to disruptions to the modulation of neural activity by GluN2D-containing NMDARs in the bed nucleus of the stria terminalis [213]. Together, these studies suggest a role for GluN2D-containing NMDARs in mediating emotional behaviours that are known to be affected in schizophrenia.

8. Limitations

As this is a narrative rather than a systematic review, there is potential for bias in the selection of publications which may, in turn, bias the conclusions of this review. A scoping review with clearly developed, predetermined inclusion and exclusion criteria and keywords might have led to more replicable and verifiable conclusions. Furthermore, due to the heterogeneity of the available literature on the GluN2D subunit, we are unable to make a conclusive statement on the role of the GluN2D subunit in schizophrenia. Instead, our review of the literature suggests that further research needs to be undertaken to fully understand the contribution of the GluN2D subunit to this disorder. A major contributor to the observed heterogeneity may be that the changes in GluN2D in schizophrenia are cell-specific. Here, recent advances in single-cell and/or single-nuclei RNA sequencing will enable a more detailed understanding of the role of GluN2D in schizophrenia. Additionally, we were not able to adequately investigate any potential sex-specific differences in the expression of the GluN2D subunit or its function or effects in humans or animal models as sex was not always included as a variable in the studies examined here. This stresses the importance of future work to always include sex as a variable.

9. Conclusions

Precipitating factors, including any combination of genetic predisposition and environmental factors like maternal infection or obstetric complications, can lead to NMDAR hypofunction disproportionately at fast-spiking PV-containing interneurons during development. This results in pathological phenotypes including impaired oscillatory activity and neuronal synchrony, cortical disinhibition and dopaminergic dysfunction, ultimately giving rise to the various symptoms of schizophrenia. Disruption of the GluN2D subunit and alteration to GluN2D neurotransmission could be a molecular pathway contributing to the symptomatology of schizophrenia. This is of importance as it may provide new insights into the aetiology of this disorder and might even lead to the development of novel drugs for the treatment of specific schizophrenia symptoms, including cognitive dysfunction.

Author Contributions

Conceptualization, C.V., R.A.H., N.C.J. and S.S.; methodology, C.V.; software, C.V.; investigation, C.V.; resources, R.A.H. and S.S.; writing—original draft preparation, C.V.; writing—review and editing, C.V., R.A.H., N.C.J., M.R.H. and S.S.; supervision, R.A.H., N.C.J. and S.S. All authors have read and agreed to the published version of the manuscript.

Funding

C.V. is funded by a Monash-University-allocated research training program (RTP) scholarship provided by the Australian Government. R.H. is supported by a National Health and Medical Research Council ideas grant GNT2000893.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

No new data were created for this review.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saha, S.; Chant, D.; Welham, J.; McGrath, J. A systematic review of the prevalence of schizophrenia. PLoS Med. 2005, 2, e141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. McCutcheon, R.A.; Reis Marques, T.; Howes, O.D. Schizophrenia-An Overview. JAMA Psychiatry 2020, 77, 201–210. [Google Scholar] [CrossRef] [PubMed]
  3. Kahn, R.; Sommer, I.; Murray, R.; Meyer-Lindenberg, A.; Weinberger, D.; Cannon, T.; O’Donovan, M.; Correl, C.; Kane, J.; van Os, J. Schizophrenia. Nature Reviews Disease Primers. Nov 2015, 12, 15067. [Google Scholar]
  4. Galderisi, S.; Mucci, A.; Buchanan, R.W.; Arango, C. Negative symptoms of schizophrenia: New developments and unanswered research questions. Lancet Psychiatry 2018, 5, 664–677. [Google Scholar] [CrossRef]
  5. McCutcheon, R.A.; Keefe, R.S.; McGuire, P.K. Cognitive impairment in schizophrenia: Aetiology, pathophysiology, and treatment. Mol. Psychiatry, 2023; online ahead of print. [Google Scholar] [CrossRef]
  6. Vigo, D.; Thornicroft, G.; Atun, R. Estimating the true global burden of mental illness. Lancet Psychiatry 2016, 3, 171–178. [Google Scholar] [CrossRef] [Green Version]
  7. Velligan, D.I.; Brain, C.; Bouérat Duvold, L.; Agid, O. Caregiver burdens associated with treatment-resistant schizophrenia: A quantitative caregiver survey of experiences, attitudes, and perceptions. Front. Psychiatry 2019, 10, 584. [Google Scholar] [CrossRef] [Green Version]
  8. Cloutier, M.; Aigbogun, M.S.; Guerin, A.; Nitulescu, R.; Ramanakumar, A.V.; Kamat, S.A.; DeLucia, M.; Duffy, R.; Legacy, S.N.; Henderson, C.; et al. The Economic Burden of Schizophrenia in the United States in 2013. J. Clin. Psychiatry 2016, 77, 764–771. [Google Scholar] [CrossRef]
  9. Neil, A.L.; Carr, V.J.; Mihalopoulos, C.; Mackinnon, A.; Morgan, V.A. Costs of psychosis in 2010: Findings from the second Australian National Survey of Psychosis. Aust. N. Zealand J. Psychiatry 2014, 48, 169–182. [Google Scholar] [CrossRef]
  10. Ripke, S.; Neale, B.M.; Corvin, A.; Walters, J.T.; Farh, K.-H.; Holmans, P.A.; Lee, P.; Bulik-Sullivan, B.; Collier, D.A.; Huang, H. Biological insights from 108 schizophrenia-associated genetic loci. Nature 2014, 511, 421–427. [Google Scholar]
  11. Trubetskoy, V.; Pardiñas, A.F.; Qi, T.; Panagiotaropoulou, G.; Awasthi, S.; Bigdeli, T.B.; Bryois, J.; Chen, C.-Y.; Dennison, C.A.; Hall, L.S. Mapping genomic loci implicates genes and synaptic biology in schizophrenia. Nature 2022, 604, 502–508. [Google Scholar] [CrossRef]
  12. Walsh, T.; McClellan, J.M.; McCarthy, S.E.; Addington, A.M.; Pierce, S.B.; Cooper, G.M.; Nord, A.S.; Kusenda, M.; Malhotra, D.; Bhandari, A.; et al. Rare structural variants disrupt multiple genes in neurodevelopmental pathways in schizophrenia. Science 2008, 320, 539–543. [Google Scholar] [CrossRef] [Green Version]
  13. Fromer, M.; Roussos, P.; Sieberts, S.K.; Johnson, J.S.; Kavanagh, D.H.; Perumal, T.M.; Ruderfer, D.M.; Oh, E.C.; Topol, A.; Shah, H.R.; et al. Gene expression elucidates functional impact of polygenic risk for schizophrenia. Nat. Neurosci. 2016, 19, 1442–1453. [Google Scholar] [CrossRef] [Green Version]
  14. Ursini, G.; Punzi, G.; Chen, Q.; Marenco, S.; Robinson, J.F.; Porcelli, A.; Hamilton, E.G.; Mitjans, M.; Maddalena, G.; Begemann, M.; et al. Convergence of placenta biology and genetic risk for schizophrenia. Nat. Med. 2018, 24, 792–801. [Google Scholar] [CrossRef]
  15. Malaspina, D.; Harlap, S.; Fennig, S.; Heiman, D.; Nahon, D.; Feldman, D.; Susser, E.S. Advancing paternal age and the risk of schizophrenia. Arch. Gen. Psychiatry 2001, 58, 361–367. [Google Scholar] [CrossRef] [Green Version]
  16. Fountoulakis, K.N.; Gonda, X.; Siamouli, M.; Panagiotidis, P.; Moutou, K.; Nimatoudis, I.; Kasper, S. Paternal and maternal age as risk factors for schizophrenia: A case–control study. Int. J. Psychiatry Clin. Pract. 2018, 22, 170–176. [Google Scholar] [CrossRef] [PubMed]
  17. Cantor-Graae, E.; Selten, J.-P. Schizophrenia and migration: A meta-analysis and review. Am. J. Psychiatry 2005, 162, 12–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Varese, F.; Smeets, F.; Drukker, M.; Lieverse, R.; Lataster, T.; Viechtbauer, W.; Read, J.; Van Os, J.; Bentall, R.P. Childhood adversities increase the risk of psychosis: A meta-analysis of patient-control, prospective-and cross-sectional cohort studies. Schizophr. Bull. 2012, 38, 661–671. [Google Scholar] [CrossRef] [PubMed]
  19. Toulopoulou, T.; Picchioni, M.; Mortensen, P.B.; Petersen, L. IQ, the urban environment, and their impact on future schizophrenia risk in men. Schizophr. Bull. 2017, 43, 1056–1063. [Google Scholar] [CrossRef] [Green Version]
  20. McKenzie, K.; Murray, A.; Booth, T. Do urban environments increase the risk of anxiety, depression and psychosis? An epidemiological study. J. Affect. Disord. 2013, 150, 1019–1024. [Google Scholar] [CrossRef] [Green Version]
  21. Vaucher, J.; Keating, B.J.; Lasserre, A.M.; Gan, W.; Lyall, D.M.; Ward, J.; Smith, D.J.; Pell, J.P.; Sattar, N.; Paré, G. Cannabis use and risk of schizophrenia: A Mendelian randomization study. Mol. Psychiatry 2018, 23, 1287–1292. [Google Scholar] [CrossRef] [Green Version]
  22. Leucht, S.; Corves, C.; Arbter, D.; Engel, R.R.; Li, C.; Davis, J.M. Second-generation versus first-generation antipsychotic drugs for schizophrenia: A meta-analysis. Lancet 2009, 373, 31–41. [Google Scholar] [CrossRef]
  23. Kane, J.M.; Correll, C.U. Pharmacologic treatment of schizophrenia. Dialogues Clin. Neurosci. 2022, 12, 345–357. [Google Scholar] [CrossRef]
  24. Keefe, R.S.; Bilder, R.M.; Davis, S.M.; Harvey, P.D.; Palmer, B.W.; Gold, J.M.; Meltzer, H.Y.; Green, M.F.; Capuano, G.; Stroup, T.S. Neurocognitive effects of antipsychotic medications in patients with chronic schizophrenia in the CATIE Trial. Arch. Gen. Psychiatry 2007, 64, 633–647. [Google Scholar] [CrossRef] [PubMed]
  25. Gründer, G.; Cumming, P. The dopamine hypothesis of schizophrenia: Current status. In The Neurobiology of Schizophrenia; Elsevier: Amsterdam, The Netherlands, 2016; pp. 109–124. [Google Scholar]
  26. Brisch, R.; Saniotis, A.; Wolf, R.; Bielau, H.; Bernstein, H.-G.; Steiner, J.; Bogerts, B.; Braun, K.; Jankowski, Z.; Kumaratilake, J. The role of dopamine in schizophrenia from a neurobiological and evolutionary perspective: Old fashioned, but still in vogue. Front. Psychiatry 2014, 5, 47. [Google Scholar]
  27. Millard, S.J.; Bearden, C.E.; Karlsgodt, K.H.; Sharpe, M.J. The prediction-error hypothesis of schizophrenia: New data point to circuit-specific changes in dopamine activity. Neuropsychopharmacology 2022, 47, 628–640. [Google Scholar] [CrossRef]
  28. McCutcheon, R.; Beck, K.; Jauhar, S.; Howes, O.D. Defining the locus of dopaminergic dysfunction in schizophrenia: A meta-analysis and test of the mesolimbic hypothesis. Schizophr. Bull. 2018, 44, 1301–1311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Seeman, P.; Lee, T.; Chau-Wong, M.; Wong, K. Antipsychotic drug doses and neuroleptic/dopamine receptors. Nature 1976, 261, 717–719. [Google Scholar] [CrossRef] [PubMed]
  30. Van Rossum, J. The significance of dopamine-receptor blockade for the mechanism of action of neuroleptic drugs. Arch. Int. Pharm. 1996, 160, 492–494. [Google Scholar]
  31. Seeman, M.V. History of the dopamine hypothesis of antipsychotic action. World J. Psychiatry 2021, 11, 355. [Google Scholar] [CrossRef]
  32. Li, P.; L Snyder, G.; E Vanover, K. Dopamine targeting drugs for the treatment of schizophrenia: Past, present and future. Curr. Top. Med. Chem. 2016, 16, 3385–3403. [Google Scholar] [CrossRef] [Green Version]
  33. Balu, D.T. The NMDA receptor and schizophrenia: From pathophysiology to treatment. In Advances in Pharmacology; Elsevier: Amsterdam, The Netherlands, 2016; Volume 76, pp. 351–382. [Google Scholar]
  34. Catts, V.S.; Lai, Y.L.; Weickert, C.S.; Weickert, T.W.; Catts, S.V. A quantitative review of the postmortem evidence for decreased cortical N-methyl-D-aspartate receptor expression levels in schizophrenia: How can we link molecular abnormalities to mismatch negativity deficits? Biol. Psychol. 2016, 116, 57–67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Nakazawa, K.; Sapkota, K. The origin of NMDA receptor hypofunction in schizophrenia. Pharmacol. Ther. 2020, 205, 107426. [Google Scholar] [CrossRef] [PubMed]
  36. Snyder, M.A.; Gao, W.-J. NMDA hypofunction as a convergence point for progression and symptoms of schizophrenia. Front. Cell. Neurosci. 2013, 7, 31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Zhou, Y.; Danbolt, N.C. Glutamate as a neurotransmitter in the healthy brain. J. Neural Transm. 2014, 121, 799–817. [Google Scholar] [CrossRef] [Green Version]
  38. Willard, S.S.; Koochekpour, S. Glutamate, glutamate receptors, and downstream signaling pathways. Int. J. Biol. Sci. 2013, 9, 948. [Google Scholar] [CrossRef] [Green Version]
  39. Traynelis, S.F.; Wollmuth, L.P.; McBain, C.J.; Menniti, F.S.; Vance, K.M.; Ogden, K.K.; Hansen, K.B.; Yuan, H.; Myers, S.J.; Dingledine, R. Glutamate receptor ion channels: Structure, regulation, and function. Pharm. Rev. 2010, 62, 405–496. [Google Scholar] [CrossRef] [Green Version]
  40. Hansen, K.B.; Wollmuth, L.P.; Bowie, D.; Furukawa, H.; Menniti, F.S.; Sobolevsky, A.I.; Swanson, G.T.; Swanger, S.A.; Greger, I.H.; Nakagawa, T. Structure, function, and pharmacology of glutamate receptor ion channels. Pharmacol. Rev. 2021, 73, 1469–1658. [Google Scholar] [CrossRef]
  41. Reiner, A.; Levitz, J. Glutamatergic signaling in the central nervous system: Ionotropic and metabotropic receptors in concert. Neuron 2018, 98, 1080–1098. [Google Scholar] [CrossRef] [Green Version]
  42. Yonezawa, K.; Tani, H.; Nakajima, S.; Nagai, N.; Koizumi, T.; Miyazaki, T.; Mimura, M.; Takahashi, T.; Uchida, H. AMPA receptors in schizophrenia: A systematic review of postmortem studies on receptor subunit expression and binding. Schizophr. Res. 2022, 243, 98–109. [Google Scholar] [CrossRef]
  43. Rubio, M.D.; Drummond, J.B.; Meador-Woodruff, J.H. Glutamate receptor abnormalities in schizophrenia: Implications for innovative treatments. Biomol. Ther. 2012, 20, 1. [Google Scholar] [CrossRef] [Green Version]
  44. Maksymetz, J.; Moran, S.P.; Conn, P.J. Targeting metabotropic glutamate receptors for novel treatments of schizophrenia. Mol. Brain 2017, 10, 1–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Monyer, H.; Sprengel, R.; Schoepfer, R.; Herb, A.; Higuchi, M.; Lomeli, H.; Burnashev, N.; Sakmann, B.; Seeburg, P.H. Heteromeric NMDA receptors: Molecular and functional distinction of subtypes. Science 1992, 256, 1217–1221. [Google Scholar] [CrossRef] [PubMed]
  46. Bye, C.M.; McDonald, R.J. A specific role of hippocampal NMDA receptors and arc protein in rapid encoding of novel environmental representations and a more general long-term consolidation function. Front. Behav. Neurosci. 2019, 13, 8. [Google Scholar] [CrossRef] [Green Version]
  47. Talpalar, A.E.; Kiehn, O. Glutamatergic mechanisms for speed control and network operation in the rodent locomotor CpG. Front. Neural Circuits 2010, 4, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Lambot, L.; Rodriguez, E.C.; Houtteman, D.; Li, Y.; Schiffmann, S.N.; Gall, D.; de Kerchove d’Exaerde, A. Striatopallidal neuron NMDA receptors control synaptic connectivity, locomotor, and goal-directed behaviors. J. Neurosci. 2016, 36, 4976–4992. [Google Scholar] [CrossRef]
  49. Li, N.; Lee, B.; Liu, R.-J.; Banasr, M.; Dwyer, J.M.; Iwata, M.; Li, X.-Y.; Aghajanian, G.; Duman, R.S. mTOR-dependent synapse formation underlies the rapid antidepressant effects of NMDA antagonists. Science 2010, 329, 959–964. [Google Scholar] [CrossRef] [Green Version]
  50. Hou, G.; Zhang, Z.-W. NMDA receptors regulate the development of neuronal intrinsic excitability through cell-autonomous mechanisms. Front. Cell. Neurosci. 2017, 11, 353. [Google Scholar] [CrossRef] [Green Version]
  51. Frangeul, L.; Kehayas, V.; Sanchez-Mut, J.V.; Fièvre, S.; Krishna, K.K.; Pouchelon, G.; Telley, L.; Bellone, C.; Holtmaat, A.; Gräff, J. Input-dependent regulation of excitability controls dendritic maturation in somatosensory thalamocortical neurons. Nat. Commun. 2017, 8, 2015. [Google Scholar] [CrossRef] [Green Version]
  52. Karakas, E.; Furukawa, H. Crystal structure of a heterotetrameric NMDA receptor ion channel. Science 2014, 344, 992–997. [Google Scholar] [CrossRef] [Green Version]
  53. Furukawa, H.; Singh, S.K.; Mancusso, R.; Gouaux, E. Subunit arrangement and function in NMDA receptors. Nature 2005, 438, 185–192. [Google Scholar] [CrossRef]
  54. Lerma, J.; Zukin, R.S.; Bennett, M. Glycine decreases desensitization of N-methyl-D-aspartate (NMDA) receptors expressed in Xenopus oocytes and is required for NMDA responses. Proc. Natl. Acad. Sci. USA 1990, 87, 2354–2358. [Google Scholar] [CrossRef] [PubMed]
  55. Greer, P.L.; Greenberg, M.E. From synapse to nucleus: Calcium-dependent gene transcription in the control of synapse development and function. Neuron 2008, 59, 846–860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Sugihara, H.; Moriyoshi, K.; Ishii, T.; Masu, M.; Nakanishi, S. Structures and properties of seven isoforms of the NMDA receptor generated by alternative splicing. Biochem. Biophys. Res. Commun. 1992, 185, 826–832. [Google Scholar] [CrossRef] [PubMed]
  57. Paoletti, P.; Bellone, C.; Zhou, Q. NMDA receptor subunit diversity: Impact on receptor properties, synaptic plasticity and disease. Nat. Rev. Neurosci. 2013, 14, 383–400. [Google Scholar] [CrossRef] [PubMed]
  58. Schorge, S.; Colquhoun, D. Studies of NMDA receptor function and stoichiometry with truncated and tandem subunits. J. Neurosci. 2003, 23, 1151–1158. [Google Scholar] [CrossRef]
  59. Henson, M.A.; Roberts, A.C.; Salimi, K.; Vadlamudi, S.; Hamer, R.M.; Gilmore, J.H.; Jarskog, L.F.; Philpot, B.D. Developmental regulation of the NMDA receptor subunits, NR3A and NR1, in human prefrontal cortex. Cereb. Cortex 2008, 18, 2560–2573. [Google Scholar] [CrossRef] [Green Version]
  60. Tolle, T.; Berthele, A.; Zieglgansberger, W.; Seeburg, P.H.; Wisden, W. The differential expression of 16 NMDA and non-NMDA receptor subunits in the rat spinal cord and in periaqueductal gray. J. Neurosci. 1993, 13, 5009–5028. [Google Scholar] [CrossRef] [Green Version]
  61. Monyer, H.; Burnashev, N.; Laurie, D.J.; Sakmann, B.; Seeburg, P.H. Developmental and regional expression in the rat brain and functional properties of four NMDA receptors. Neuron 1994, 12, 529–540. [Google Scholar] [CrossRef]
  62. Vicini, S.; Wang, J.F.; Li, J.H.; Zhu, W.J.; Wang, Y.H.; Luo, J.H.; Wolfe, B.B.; Grayson, D.R. Functional and pharmacological differences between recombinant N-methyl-D-aspartate receptors. J. Neurophysiol. 1998, 79, 555–566. [Google Scholar] [CrossRef] [Green Version]
  63. Krystal, J.H.; Karper, L.P.; Seibyl, J.P.; Freeman, G.K.; Delaney, R.; Bremner, J.D.; Heninger, G.R.; Bowers, M.B., Jr.; Charney, D.S. Subanesthetic effects of the noncompetitive NMDA antagonist, ketamine, in humans. Psychotomimetic, perceptual, cognitive, and neuroendocrine responses. Arch. Gen. Psychiatry 1994, 51, 199–214. [Google Scholar] [CrossRef]
  64. Javitt, D. Negative schizophrenic symptomatology and the PCP (phencyclidine) model of schizophrenia. Hillside J. Clin. Psychiatry 1987, 9, 12–35. [Google Scholar]
  65. Luby, E.D.; Cohen, B.D.; Rosenbaum, G.; Gottlieb, J.S.; Kelley, R. Study of a new schizophrenomimetic drug—Sernyl. AMA Arch. Neurol. Psychiatry 1959, 81, 363–369. [Google Scholar] [CrossRef]
  66. Luby, E.D.; Gottlieb, J.S.; Cohen, B.D.; Rosenbaum, G.; Domino, E.F. Model psychoses and schizophrenia. Am. J. Psychiatry 1962, 119, 61–67. [Google Scholar] [CrossRef] [PubMed]
  67. Rosenbaum, G.; Cohen, B.D.; Luby, E.D.; Gottlieb, J.S.; Yelen, D. Comparison of sernyl with other drugs: Simulation of schizophrenic performance with sernyl, LSD-25, and amobarbital (amytal) sodium; I. Attention, motor function, and proprioception. AMA Arch. Gen. Psychiatry 1959, 1, 651–656. [Google Scholar] [CrossRef] [PubMed]
  68. Cohen, B.D.; Luby, E.D.; Rosenbaum, G.; Gottlieb, J.S. Combined sernyl and sensory deprivation. Compr. Psychiatry 1960, 1, 345–348. [Google Scholar] [CrossRef]
  69. Itil, T.; Keskiner, A.; Kiremitci, N.; Holden, J. Effect of phencyclidine in chronic schizophrenics. Can. Psychiatr. Assoc. J. 1967, 12, 209–212. [Google Scholar] [CrossRef] [Green Version]
  70. Lahti, A.C.; Koffel, B.; LaPorte, D.; Tamminga, C.A. Subanesthetic doses of ketamine stimulate psychosis in schizophrenia. Neuropsychopharmacology 1995, 13, 9–19. [Google Scholar] [CrossRef] [PubMed]
  71. Anis, N.A.; Berry, S.C.; Burton, N.R.; Lodge, D. The dissociative anaesthetics, ketamine and phencyclidine, selectively reduce excitation of central mammalian neurones by N-methyl-aspartate. Br. J. Pharm. 1983, 79, 565–575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Kim, J.; Kornhuber, H.; Schmid-Burgk, W.; Holzmüller, B. Low cerebrospinal fluid glutamate in schizophrenic patients and a new hypothesis on schizophrenia. Neurosci. Lett. 1980, 20, 379–382. [Google Scholar] [CrossRef]
  73. Javitt, D.C.; Zukin, S.R. Recent advances in the phencyclidine model of schizophrenia. Am. J. Psychiatry 1991, 148, 1301–1308. [Google Scholar]
  74. Javitt, D.C. Glutamatergic theories of schizophrenia. Isr. J. Psychiatry Relat. Sci. 2010, 47, 4–16. [Google Scholar]
  75. Kantrowitz, J.T.; Javitt, D.C. N-methyl-d-aspartate (NMDA) receptor dysfunction or dysregulation: The final common pathway on the road to schizophrenia? Brain Res. Bull. 2010, 83, 108–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Kantrowitz, J.; Javitt, D.C. Glutamatergic transmission in schizophrenia: From basic research to clinical practice. Curr. Opin. Psychiatry 2012, 25, 96–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Mohn, A.R.; Gainetdinov, R.R.; Caron, M.G.; Koller, B.H. Mice with reduced NMDA receptor expression display behaviors related to schizophrenia. Cell 1999, 98, 427–436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Halene, T.B.; Ehrlichman, R.S.; Liang, Y.; Christian, E.P.; Jonak, G.J.; Gur, T.L.; Blendy, J.A.; Dow, H.C.; Brodkin, E.S.; Schneider, F. Assessment of NMDA receptor NR1 subunit hypofunction in mice as a model for schizophrenia. Genes Brain Behav. 2009, 8, 661–675. [Google Scholar] [CrossRef] [PubMed]
  79. Hakami, T.; Jones, N.C.; Tolmacheva, E.A.; Gaudias, J.; Chaumont, J.; Salzberg, M.; O’Brien, T.J.; Pinault, D. NMDA receptor hypofunction leads to generalized and persistent aberrant γ oscillations independent of hyperlocomotion and the state of consciousness. PLoS ONE 2009, 4, e6755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Anderson, P.M.; Pinault, D.; O’Brien, T.J.; Jones, N.C. Chronic administration of antipsychotics attenuates ongoing and ketamine-induced increases in cortical γ oscillations. Int. J. Neuropsychopharmacol. 2014, 17, 1895–1904. [Google Scholar] [CrossRef]
  81. Hudson, M.; Rind, G.; O’brien, T.; Jones, N. Reversal of evoked gamma oscillation deficits is predictive of antipsychotic activity with a unique profile for clozapine. Transl. Psychiatry 2016, 6, e784. [Google Scholar] [CrossRef]
  82. Belforte, J.E.; Zsiros, V.; Sklar, E.R.; Jiang, Z.; Yu, G.; Li, Y.; Quinlan, E.M.; Nakazawa, K. Postnatal NMDA receptor ablation in corticolimbic interneurons confers schizophrenia-like phenotypes. Nat. Neurosci. 2010, 13, 76–83. [Google Scholar] [CrossRef] [Green Version]
  83. Cohen, S.M.; Tsien, R.W.; Goff, D.C.; Halassa, M.M. The impact of NMDA receptor hypofunction on GABAergic neurons in the pathophysiology of schizophrenia. Schizophr. Res. 2015, 167, 98–107. [Google Scholar] [CrossRef] [Green Version]
  84. Nakao, K.; Jeevakumar, V.; Jiang, S.Z.; Fujita, Y.; Diaz, N.B.; Pretell Annan, C.A.; Eskow Jaunarajs, K.L.; Hashimoto, K.; Belforte, J.E.; Nakazawa, K. Schizophrenia-like dopamine release abnormalities in a mouse model of NMDA receptor hypofunction. Schizophr. Bull. 2019, 45, 138–147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Akgül, G.; McBain, C.J. Diverse roles for ionotropic glutamate receptors on inhibitory interneurons in developing and adult brain. J. Physiol. 2016, 594, 5471–5490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Homayoun, H.; Moghaddam, B. NMDA receptor hypofunction produces opposite effects on prefrontal cortex interneurons and pyramidal neurons. J. Neurosci. 2007, 27, 11496–11500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Li, Q.; Clark, S.; Lewis, D.V.; Wilson, W.A. NMDA receptor antagonists disinhibit rat posterior cingulate and retrosplenial cortices: A potential mechanism of neurotoxicity. J. Neurosci. 2002, 22, 3070–3080. [Google Scholar] [CrossRef] [Green Version]
  88. Suzuki, Y.; Jodo, E.; Takeuchi, S.; Niwa, S.; Kayama, Y. Acute administration of phencyclidine induces tonic activation of medial prefrontal cortex neurons in freely moving rats. Neuroscience 2002, 114, 769–779. [Google Scholar] [CrossRef]
  89. Jackson, M.E.; Homayoun, H.; Moghaddam, B. NMDA receptor hypofunction produces concomitant firing rate potentiation and burst activity reduction in the prefrontal cortex. Proc. Natl. Acad. Sci. USA 2004, 101, 8467–8472. [Google Scholar] [CrossRef]
  90. Moghaddam, B.; Adams, B.; Verma, A.; Daly, D. Activation of glutamatergic neurotransmission by ketamine: A novel step in the pathway from NMDA receptor blockade to dopaminergic and cognitive disruptions associated with the prefrontal cortex. J. Neurosci. 1997, 17, 2921–2927. [Google Scholar] [CrossRef] [Green Version]
  91. Okada, M.; Fukuyama, K.; Ueda, Y. Lurasidone inhibits NMDA receptor antagonist-induced functional abnormality of thalamocortical glutamatergic transmission via 5-HT7 receptor blockade. Br. J. Pharmacol. 2019, 176, 4002–4018. [Google Scholar] [CrossRef]
  92. Widman, A.J.; McMahon, L.L. Disinhibition of CA1 pyramidal cells by low-dose ketamine and other antagonists with rapid antidepressant efficacy. Proc. Natl. Acad. Sci. USA 2018, 115, E3007–E3016. [Google Scholar] [CrossRef] [Green Version]
  93. Schobel, S.A.; Chaudhury, N.H.; Khan, U.A.; Paniagua, B.; Styner, M.A.; Asllani, I.; Inbar, B.P.; Corcoran, C.M.; Lieberman, J.A.; Moore, H. Imaging patients with psychosis and a mouse model establishes a spreading pattern of hippocampal dysfunction and implicates glutamate as a driver. Neuron 2013, 78, 81–93. [Google Scholar] [CrossRef] [Green Version]
  94. Murray, J.D.; Anticevic, A.; Gancsos, M.; Ichinose, M.; Corlett, P.R.; Krystal, J.H.; Wang, X.-J. Linking microcircuit dysfunction to cognitive impairment: Effects of disinhibition associated with schizophrenia in a cortical working memory model. Cereb. Cortex. 2014, 24, 859–872. [Google Scholar] [CrossRef] [Green Version]
  95. Shaw, A.D.; Saxena, N.; Jackson, L.E.; Hall, J.E.; Singh, K.D.; Muthukumaraswamy, S.D. Ketamine amplifies induced gamma frequency oscillations in the human cerebral cortex. Eur. Neuropsychopharmacol. 2015, 25, 1136–1146. [Google Scholar] [CrossRef] [PubMed]
  96. Vollenweider, F.; Leenders, K.; Scharfetter, C.; Antonini, A.; Maguire, P.; Missimer, J.; Angst, J. Metabolic hyperfrontality and psychopathology in the ketamine model of psychosis using positron emission tomography (PET) and [18F] fluorodeoxyglucose (FDG). Eur. Neuropsychopharmacol. 1997, 7, 9–24. [Google Scholar] [CrossRef]
  97. Breier, A.; Malhotra, A.K.; Pinals, D.A.; Weisenfeld, N.I.; Pickar, D. Association of ketamine-induced psychosis with focal activation of the prefrontal cortex in healthy volunteers. Am. J. Psychiatry 1997, 154, 805–811. [Google Scholar] [PubMed] [Green Version]
  98. Miyamoto, S.; Leipzig, J.N.; Lieberman, J.A.; Duncan, G.E. Effects of ketamine, MK-801, and amphetamine on regional brain 2-deoxyglucose uptake in freely moving mice. Neuropsychopharmacology 2000, 22, 400–412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Duncan, G.E.; Moy, S.S.; Knapp, D.J.; Mueller, R.A.; Breese, G.R. Metabolic mapping of the rat brain after subanesthetic doses of ketamine: Potential relevance to schizophrenia. Brain Res. 1998, 787, 181–190. [Google Scholar] [CrossRef]
  100. Nair, V.D.; Savelli, J.E.; Mishra, R.K. Modulation of dopamine D 2 receptor expression by an NMDA receptor antagonist in rat brain. J. Mol. Neurosci. 1998, 11, 121–126. [Google Scholar] [CrossRef]
  101. Micheletti, G.; Lannes, B.; Haby, C.; Borrelli, E.; Kempf, E.; Warter, J.-M.; Zwiller, J. Chronic administration of NMDA antagonists induces D2 receptor synthesis in rat striatum. Mol. Brain Res. 1992, 14, 363–368. [Google Scholar] [CrossRef]
  102. Kegeles, L.S.; Abi-Dargham, A.; Zea-Ponce, Y.; Rodenhiser-Hill, J.; Mann, J.J.; Van Heertum, R.L.; Cooper, T.B.; Carlsson, A.; Laruelle, M. Modulation of amphetamine-induced striatal dopamine release by ketamine in humans: Implications for schizophrenia. Biol. Psychiatry 2000, 48, 627–640. [Google Scholar] [CrossRef]
  103. Smith, G.S.; Schloesser, R.; Brodie, J.D.; Dewey, S.L.; Logan, J.; Vitkun, S.A.; Simkowitz, P.; Hurley, A.; Cooper, T.; Volkow, N.D. Glutamate modulation of dopamine measured in vivo with positron emission tomography (PET) and 11C-raclopride in normal human subjects. Neuropsychopharmacology 1998, 18, 18–25. [Google Scholar] [CrossRef] [Green Version]
  104. Sun, J.; Jia, P.; Fanous, A.H.; Van Den Oord, E.; Chen, X.; Riley, B.P.; Amdur, R.L.; Kendler, K.S.; Zhao, Z. Schizophrenia gene networks and pathways and their applications for novel candidate gene selection. PLoS ONE 2010, 5, e11351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Greenwood, T.A.; Light, G.A.; Swerdlow, N.R.; Radant, A.D.; Braff, D.L. Association analysis of 94 candidate genes and schizophrenia-related endophenotypes. PLoS ONE 2012, 7, e29630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Balu, D.T.; Coyle, J.T. Neuroplasticity signaling pathways linked to the pathophysiology of schizophrenia. Neurosci. Biobehav. Rev. 2011, 35, 848–870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Hashimoto, K.; Fukushima, T.; Shimizu, E.; Komatsu, N.; Watanabe, H.; Shinoda, N.; Nakazato, M.; Kumakiri, C.; Okada, S.; Hasegawa, H.; et al. Decreased serum levels of D-serine in patients with schizophrenia: Evidence in support of the N-methyl-D-aspartate receptor hypofunction hypothesis of schizophrenia. Arch. Gen. Psychiatry 2003, 60, 572–576. [Google Scholar] [CrossRef] [Green Version]
  108. Weickert, C.S.; Fung, S.J.; Catts, V.S.; Schofield, P.R.; Allen, K.M.; Moore, L.T.; Newell, K.A.; Pellen, D.; Huang, X.F.; Catts, S.V.; et al. Molecular evidence of N-methyl-D-aspartate receptor hypofunction in schizophrenia. Mol. Psychiatry 2013, 18, 1185–1192. [Google Scholar] [CrossRef] [Green Version]
  109. Labrie, V.; Lipina, T.; Roder, J.C. Mice with reduced NMDA receptor glycine affinity model some of the negative and cognitive symptoms of schizophrenia. Psychopharmacology 2008, 200, 217–230. [Google Scholar] [CrossRef]
  110. Zhao, X.; Li, H.; Shi, Y.; Tang, R.; Chen, W.; Liu, J.; Feng, G.; Shi, J.; Yan, L.; Liu, H. Significant association between the genetic variations in the 5′ end of the N-methyl-D-aspartate receptor subunit gene GRIN1 and schizophrenia. Biol. Psychiatry 2006, 59, 747–753. [Google Scholar] [CrossRef]
  111. Poltavskaya, E.G.; Fedorenko, O.Y.; Kornetova, E.G.; Loonen, A.J.; Kornetov, A.N.; Bokhan, N.A.; Ivanova, S.A. Study of early onset schizophrenia: Associations of GRIN2A and GRIN2B polymorphisms. Life 2021, 11, 997. [Google Scholar] [CrossRef]
  112. Catts, V.S.; Derminio, D.S.; Hahn, C.-G.; Weickert, C.S. Postsynaptic density levels of the NMDA receptor NR1 subunit and PSD-95 protein in prefrontal cortex from people with schizophrenia. Npj Schizophrenia 2015, 1, 1–8. [Google Scholar] [CrossRef] [Green Version]
  113. Ishii, T.; Moriyoshi, K.; Sugihara, H.; Sakurada, K.; Kadotani, H.; Yokoi, M.; Akazawa, C.; Shigemoto, R.; Mizuno, N.; Masu, M. Molecular characterization of the family of the N-methyl-D-aspartate receptor subunits. J. Biol. Chem. 1993, 268, 2836–2843. [Google Scholar] [CrossRef] [PubMed]
  114. Akazawa, C.; Shigemoto, R.; Bessho, Y.; Nakanishi, S.; Mizuno, N. Differential expression of five N-methyl-D-aspartate receptor subunit mRNAs in the cerebellum of developing and adult rats. J. Comp. Neurol. 1994, 347, 150–160. [Google Scholar] [CrossRef]
  115. Wenzel, A.; Villa, M.; Mohler, H.; Benke, D. Developmental and regional expression of NMDA receptor subtypes containing the NR2D subunit in rat brain. J. Neurochem. 1996, 66, 1240–1248. [Google Scholar] [CrossRef] [PubMed]
  116. Dunah, A.W.; Yasuda, R.P.; Wang, Y.h.; Luo, J.; Dávila-García, M.I.; Gbadegesin, M.; Vicini, S.; Wolfe, B.B. Regional and ontogenic expression of the NMDA receptor subunit NR2D protein in rat brain using a subunit-specific antibody. J. Neurochem. 1996, 67, 2335–2345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Standaert, D.G.; Landwehrmeyer, G.B.; Kerner, J.A.; Penney Jr, J.B.; Young, A.B. Expression of NMDAR2D glutamate receptor subunit mRNA in neurochemically identified interneurons in the rat neostriatum, neocortex and hippocampus. Mol. Brain Res. 1996, 42, 89–102. [Google Scholar] [CrossRef] [PubMed]
  118. Cannon, M.; Murray, R.M. Neonatal origins of schizophrenia. Arch. Dis. Child. 1998, 78, 1–3. [Google Scholar] [CrossRef] [Green Version]
  119. Birnbaum, R.; Weinberger, D.R. Genetic insights into the neurodevelopmental origins of schizophrenia. Nat. Rev. Neurosci. 2017, 18, 727–740. [Google Scholar] [CrossRef] [PubMed]
  120. Engelhardt, J.v.; Bocklisch, C.; Tönges, L.; Herb, A.; Mishina, M.; Monyer, H. GluN2D-containing NMDA receptors-mediate synaptic currents in hippocampal interneurons and pyramidal cells in juvenile mice. Front. Cell Neurosci. 2015, 9, 95. [Google Scholar] [CrossRef] [Green Version]
  121. Garst-Orozco, J.; Malik, R.; Lanz, T.A.; Weber, M.L.; Xi, H.; Arion, D.; Enwright III, J.F.; Lewis, D.A.; O’Donnell, P.; Sohal, V.S. GluN2D-mediated excitatory drive onto medial prefrontal cortical PV+ fast-spiking inhibitory interneurons. PLoS ONE 2020, 15, e0233895. [Google Scholar] [CrossRef]
  122. Perszyk, R.E.; DiRaddo, J.O.; Strong, K.L.; Low, C.M.; Ogden, K.K.; Khatri, A.; Vargish, G.A.; Pelkey, K.A.; Tricoire, L.; Liotta, D.C.; et al. GluN2D-Containing N-methyl-d-Aspartate Receptors Mediate Synaptic Transmission in Hippocampal Interneurons and Regulate Interneuron Activity. Mol. Pharm. 2016, 90, 689–702. [Google Scholar] [CrossRef] [Green Version]
  123. Alsaad, H.A.; DeKorver, N.W.; Mao, Z.; Dravid, S.M.; Arikkath, J.; Monaghan, D.T. In the telencephalon, GluN2C NMDA receptor subunit mRNA is predominately expressed in glial cells and GluN2D mRNA in interneurons. Neurochem. Res. 2019, 44, 61–77. [Google Scholar] [CrossRef]
  124. Camp, C.R.; Yuan, H. GRIN2D/GluN2D NMDA receptor: Unique features and its contribution to pediatric developmental and epileptic encephalopathy. Eur. J. Paediatr. Neurol. 2020, 24, 89–99. [Google Scholar] [CrossRef] [PubMed]
  125. Yamasaki, M.; Okada, R.; Takasaki, C.; Toki, S.; Fukaya, M.; Natsume, R.; Sakimura, K.; Mishina, M.; Shirakawa, T.; Watanabe, M. Opposing role of NMDA receptor GluN2B and GluN2D in somatosensory development and maturation. J. Neurosci. 2014, 34, 11534–11548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Ritter, L.M.; Unis, A.S.; Meador-Woodruff, J.H. Ontogeny of ionotropic glutamate receptor expression in human fetal brain. Dev. Brain Res. 2001, 127, 123–133. [Google Scholar] [CrossRef] [PubMed]
  127. Scherzer, C.R.; Landwehrmeyer, G.B.; Kerner, J.A.; Counihan, T.J.; Kosinski, C.M.; Standaert, D.G.; Daggett, L.P.; Velicelebi, G.; Penney, J.B.; Young, A.B. Expression of N-methyl-D-aspartate receptor subunit mRNAs in the human brain: Hippocampus and cortex. J. Comp. Neurol. 1998, 390, 75–90. [Google Scholar] [CrossRef]
  128. Akbarian, S.; Sucher, N.J.; Bradley, D.; Tafazzoli, A.; Trinh, D.; Hetrick, W.P.; Potkin, S.G.; Sandman, C.A.; Bunney, W.E., Jr.; Jones, E.G. Selective alterations in gene expression for NMDA receptor subunits in prefrontal cortex of schizophrenics. J. Neurosci. 1996, 16, 19–30. [Google Scholar] [CrossRef] [Green Version]
  129. Dunah, A.W.; Luo, J.; Wang, Y.-H.; Yasuda, R.P.; Wolfe, B.B. Subunit Composition ofN-Methyl-D-aspartate Receptors in the Central Nervous System that Contain the NR2D Subunit. Mol. Pharmacol. 1998, 53, 429–437. [Google Scholar] [CrossRef] [Green Version]
  130. Brickley, S.G.; Misra, C.; Mok, M.S.; Mishina, M.; Cull-Candy, S.G. NR2B and NR2D subunits coassemble in cerebellar Golgi cells to form a distinct NMDA receptor subtype restricted to extrasynaptic sites. J. Neurosci. 2003, 23, 4958–4966. [Google Scholar] [CrossRef] [Green Version]
  131. Jones, S.; Gibb, A.J. Functional NR2B-and NR2D-containing NMDA receptor channels in rat substantia nigra dopaminergic neurones. J. Physiol. 2005, 569, 209–221. [Google Scholar] [CrossRef]
  132. Erreger, K.; Geballe, M.T.; Kristensen, A.; Chen, P.E.; Hansen, K.B.; Lee, C.J.; Yuan, H.; Le, P.; Lyuboslavsky, P.N.; Micale, N. Subunit-specific agonist activity at NR2A-, NR2B-, NR2C-, and NR2D-containing N-methyl-D-aspartate glutamate receptors. Mol. Pharmacol. 2007, 72, 907–920. [Google Scholar] [CrossRef]
  133. Chen, P.E.; Geballe, M.T.; Katz, E.; Erreger, K.; Livesey, M.R.; O’toole, K.K.; Le, P.; Lee, C.J.; Snyder, J.P.; Traynelis, S.F. Modulation of glycine potency in rat recombinant NMDA receptors containing chimeric NR2A/2D subunits expressed in Xenopus laevis oocytes. J. Physiol. 2008, 586, 227–245. [Google Scholar] [CrossRef] [PubMed]
  134. Kuner, T.; Schoepfer, R. Multiple structural elements determine subunit specificity of Mg2+ block in NMDA receptor channels. J. Neurosci. 1996, 16, 3549–3558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Clarke, R.J.; Johnson, J.W. NMDA receptor NR2 subunit dependence of the slow component of magnesium unblock. J Neurosci 2006, 26, 5825–5834. [Google Scholar] [CrossRef]
  136. Kotermanski, S.E.; Johnson, J.W. Mg2+ imparts NMDA receptor subtype selectivity to the Alzheimer’s drug memantine. J. Neurosci. 2009, 29, 2774–2779. [Google Scholar] [CrossRef] [Green Version]
  137. Sapkota, K.; Mao, Z.; Synowicki, P.; Lieber, D.; Liu, M.; Ikezu, T.; Gautam, V.; Monaghan, D.T. GluN2D N-Methyl-d-Aspartate Receptor Subunit Contribution to the Stimulation of Brain Activity and Gamma Oscillations by Ketamine: Implications for Schizophrenia. J. Pharm. Exp. 2016, 356, 702–711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Wyllie, D.J.; Behe, P.; Colquhoun, D. Single-channel activations and concentration jumps: Comparison of recombinant NR1a/NR2A and NR1a/NR2D NMDA receptors. J. Physiol. 1998, 510 Pt 1, 1–18. [Google Scholar] [CrossRef] [PubMed]
  139. Krupp, J.J.; Vissel, B.; Heinemann, S.F.; Westbrook, G.L. Calcium-dependent inactivation of recombinant N-methyl-D-aspartate receptors is NR2 subunit specific. Mol. Pharmacol. 1996, 50, 1680–1688. [Google Scholar]
  140. Cull-Candy, S.; Brickley, S.; Farrant, M. NMDA receptor subunits: Diversity, development and disease. Curr. Opin. Neurobiol. 2001, 11, 327–335. [Google Scholar] [CrossRef]
  141. Hanson, E.; Armbruster, M.; Lau, L.A.; Sommer, M.E.; Klaft, Z.-J.; Swanger, S.A.; Traynelis, S.F.; Moss, S.J.; Noubary, F.; Chadchankar, J. Tonic activation of GluN2C/GluN2D-containing NMDA receptors by ambient glutamate facilitates cortical interneuron maturation. J. Neurosci. 2019, 39, 3611–3626. [Google Scholar] [CrossRef] [Green Version]
  142. Swanger, S.A.; Vance, K.M.; Pare, J.-F.; Sotty, F.; Fog, K.; Smith, Y.; Traynelis, S.F. NMDA receptors containing the GluN2D subunit control neuronal function in the subthalamic nucleus. J. Neurosci. 2015, 35, 15971–15983. [Google Scholar] [CrossRef] [Green Version]
  143. 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] [Green Version]
  144. Andrade-Talavera, Y.; Duque-Feria, P.; Paulsen, O.; Rodríguez-Moreno, A. Presynaptic spike timing-dependent long-term depression in the mouse hippocampus. Cereb. Cortex. 2016, 26, 3637–3654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Feldman, D.E. The spike-timing dependence of plasticity. Neuron 2012, 75, 556–571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Dubois, C.J.; Lachamp, P.M.; Sun, L.; Mishina, M.; Liu, S.J. Presynaptic GluN2D receptors detect glutamate spillover and regulate cerebellar GABA release. J. Neurophysiol. 2016, 115, 271–285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Watanabe, T.; Inoue, S.; Hiroi, H.; Orimo, A.; Muramatsu, M. NMDA receptor type 2D gene as target for estrogen receptor in the brain. Mol. Brain Res. 1999, 63, 375–379. [Google Scholar] [CrossRef]
  148. Riecher-Rössler, A.; Häfner, H.; Stumbaum, M.; Maurer, K.; Schmidt, R. Can estradiol modulate schizophrenic symptomatology? Schizophr. Bull. 1994, 20, 203–214. [Google Scholar] [CrossRef] [Green Version]
  149. McCarthny, C.R.; Du, X.; Wu, Y.C.; Hill, R.A. Investigating the interactive effects of sex steroid hormones and brain-derived neurotrophic factor during adolescence on hippocampal NMDA receptor expression. Int. J. Endocrinol. 2018, 2018, 7231915. [Google Scholar] [CrossRef] [Green Version]
  150. Makino, C.; Shibata, H.; Ninomiya, H.; Tashiro, N.; Fukumaki, Y. Identification of single-nucleotide polymorphisms in the human N-methyl-D-aspartate receptor subunit NR2D gene, GRIN2D, and association study with schizophrenia. Psychiatr. Genet. 2005, 15, 215–221. [Google Scholar] [CrossRef]
  151. Yu, Y.; Lin, Y.; Takasaki, Y.; Wang, C.; Kimura, H.; Xing, J.; Ishizuka, K.; Toyama, M.; Kushima, I.; Mori, D. Rare loss of function mutations in N-methyl-D-aspartate glutamate receptors and their contributions to schizophrenia susceptibility. Transl. Psychiatry 2018, 8, 1–9. [Google Scholar] [CrossRef] [Green Version]
  152. Beneyto, M.; Meador-Woodruff, J.H. Lamina-specific abnormalities of NMDA receptor-associated postsynaptic protein transcripts in the prefrontal cortex in schizophrenia and bipolar disorder. Neuropsychopharmacology 2008, 33, 2175–2186. [Google Scholar] [CrossRef] [Green Version]
  153. Kristiansen, L.; Beneyto, M.; Haroutunian, V.; Meador-Woodruff, J. Changes in NMDA receptor subunits and interacting PSD proteins in dorsolateral prefrontal and anterior cingulate cortex indicate abnormal regional expression in schizophrenia. Mol. Psychiatry 2006, 11, 737–747. [Google Scholar] [CrossRef] [Green Version]
  154. Sodhi, M.S.; Simmons, M.; McCullumsmith, R.; Haroutunian, V.; Meador-Woodruff, J.H. Glutamatergic gene expression is specifically reduced in thalamocortical projecting relay neurons in schizophrenia. Biol. Psychiatry 2011, 70, 646–654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Andreasen, N.C. A unitary model of schizophrenia: Bleuler’s fragmented phrene as schizencephaly. Arch. Gen. Psychiatry 1999, 56, 781–787. [Google Scholar] [CrossRef]
  156. Anticevic, A.; Cole, M.W.; Repovs, G.; Murray, J.D.; Brumbaugh, M.S.; Winkler, A.M.; Savic, A.; Krystal, J.H.; Pearlson, G.D.; Glahn, D.C. Characterizing thalamo-cortical disturbances in schizophrenia and bipolar illness. Cereb. Cortex. 2014, 24, 3116–3130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Chen, P.; Ye, E.; Jin, X.; Zhu, Y.; Wang, L. Association between thalamocortical functional connectivity abnormalities and cognitive deficits in schizophrenia. Sci. Rep. 2019, 9, 2952. [Google Scholar] [CrossRef] [PubMed]
  158. Schmitt, A.; Koschel, J.; Zink, M.; Bauer, M.; Sommer, C.; Frank, J.; Treutlein, J.; Schulze, T.; Schneider-Axmann, T.; Parlapani, E. Gene expression of NMDA receptor subunits in the cerebellum of elderly patients with schizophrenia. Eur. Arch. Psychiatry Clin. Neurosci. 2010, 260, 101–111. [Google Scholar] [CrossRef] [Green Version]
  159. Ikeda, K.; Araki, K.; Takayama, C.; Inoue, Y.; Yagi, T.; Aizawa, S.; Mishina, M. Reduced spontaneous activity of mice defective in the ε4 subunit of the NMDA receptor channel. Mol. Brain Res. 1995, 33, 61–71. [Google Scholar] [CrossRef] [PubMed]
  160. Miyamoto, Y.; Yamada, K.; Noda, Y.; Mori, H.; Mishina, M.; Nabeshima, T. Lower sensitivity to stress and altered monoaminergic neuronal function in mice lacking the NMDA receptor ε4 subunit. J. Neurosci. 2002, 22, 2335–2342. [Google Scholar] [CrossRef]
  161. Obiang, P.; Macrez, R.; Jullienne, A.; Bertrand, T.; Lesept, F.; Ali, C.; Maubert, E.; Vivien, D.; Agin, V. GluN2D subunit-containing NMDA receptors control tissue plasminogen activator-mediated spatial memory. J. Neurosci. 2012, 32, 12726–12734. [Google Scholar] [CrossRef]
  162. Ide, S.; Ikekubo, Y.; Mishina, M.; Hashimoto, K.; Ikeda, K. Cognitive impairment that is induced by (R)-ketamine is abolished in NMDA GluN2D receptor subunit knockout mice. Int. J. Neuropsychopharmacol. 2019, 22, 449–452. [Google Scholar] [CrossRef] [Green Version]
  163. Hagino, Y.; Kasai, S.; Han, W.; Yamamoto, H.; Nabeshima, T.; Mishina, M.; Ikeda, K. Essential role of NMDA receptor channel ε4 subunit (GluN2D) in the effects of phencyclidine, but not methamphetamine. PLoS ONE 2010, 5, e13722. [Google Scholar] [CrossRef]
  164. Takeuchi, T.; Kiyama, Y.; Nakamura, K.; Tsujita, M.; Matsuda, I.; Mori, H.; Munemoto, Y.; Kuriyama, H.; Natsume, R.; Sakimura, K. Roles of the glutamate receptor ε2 and δ2 subunits in the potentiation and prepulse inhibition of the acoustic startle reflex. Eur. J. Neurosci. 2001, 14, 153–160. [Google Scholar] [CrossRef] [PubMed]
  165. Yamamoto, H.; Kamegaya, E.; Sawada, W.; Hasegawa, R.; Yamamoto, T.; Hagino, Y.; Takamatsu, Y.; Imai, K.; Koga, H.; Mishina, M. Involvement of the N-methyl-D-aspartate receptor GluN2D subunit in phencyclidine-induced motor impairment, gene expression, and increased Fos immunoreactivity. Mol. Brain 2013, 6, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Yamamoto, T.; Nakayama, T.; Yamaguchi, J.; Matsuzawa, M.; Mishina, M.; Ikeda, K.; Yamamoto, H. Role of the NMDA receptor GluN2D subunit in the expression of ketamine-induced behavioral sensitization and region-specific activation of neuronal nitric oxide synthase. Neurosci. Lett. 2016, 610, 48–53. [Google Scholar] [CrossRef] [PubMed]
  167. Duncan, G.E.; Miyamoto, S.; Leipzig, J.N.; Lieberman, J.A. Comparison of brain metabolic activity patterns induced by ketamine, MK-801 and amphetamine in rats: Support for NMDA receptor involvement in responses to subanesthetic dose of ketamine. Brain Res. 1999, 843, 171–183. [Google Scholar] [CrossRef] [PubMed]
  168. Mao, Z.; He, S.; Mesnard, C.; Synowicki, P.; Zhang, Y.; Chung, L.; Wiesman, A.I.; Wilson, T.W.; Monaghan, D.T. NMDA receptors containing GluN2C and GluN2D subunits have opposing roles in modulating neuronal oscillations; potential mechanism for bidirectional feedback. Brain Res. 2020, 1727, 146571. [Google Scholar] [CrossRef]
  169. Hrabetova, S.; Serrano, P.; Blace, N.; Heong, W.T.; Skifter, D.A.; Jane, D.E.; Monaghan, D.T.; Sacktor, T.C. Distinct NMDA receptor subpopulations contribute to long-term potentiation and long-term depression induction. J. Neurosci. 2000, 20, RC81. [Google Scholar] [CrossRef]
  170. Lozovaya, N.A.; Grebenyuk, S.E.; Tsintsadze, T.S.; Feng, B.; Monaghan, D.T.; Krishtal, O.A. Extrasynaptic NR2B and NR2D subunits of NMDA receptors shape ‘superslow’afterburst EPSC in rat hippocampus. J. Physiol. 2004, 558, 451–463. [Google Scholar] [CrossRef]
  171. Eapen, A.V.; Fernández-Fernández, D.; Georgiou, J.; Bortolotto, Z.A.; Lightman, S.; Jane, D.E.; Volianskis, A.; Collingridge, G.L. Multiple roles of GluN2D-containing NMDA receptors in short-term potentiation and long-term potentiation in mouse hippocampal slices. Neuropharmacology 2021, 201, 108833. [Google Scholar] [CrossRef]
  172. Ingram, R.; Kang, H.; Lightman, S.; Jane, D.E.; Bortolotto, Z.A.; Collingridge, G.L.; Lodge, D.; Volianskis, A. Some distorted thoughts about ketamine as a psychedelic and a novel hypothesis based on NMDA receptor-mediated synaptic plasticity. Neuropharmacology 2018, 142, 30–40. [Google Scholar] [CrossRef]
  173. Bygrave, A.M.; Kilonzo, K.; Kullmann, D.M.; Bannerman, D.M.; Kätzel, D. Can N-methyl-D-aspartate receptor hypofunction in schizophrenia be localized to an individual cell type? Front. Psychiatry 2019, 10, 835. [Google Scholar] [CrossRef] [Green Version]
  174. Bygrave, A.M.; Masiulis, S.; Nicholson, E.; Berkemann, M.; Barkus, C.; Sprengel, R.; Harrison, P.J.; Kullmann, D.M.; Bannerman, D.M.; Kätzel, D. Knockout of NMDA-receptors from parvalbumin interneurons sensitizes to schizophrenia-related deficits induced by MK-801. Transl. Psychiatry 2016, 6, e778. [Google Scholar] [CrossRef] [Green Version]
  175. Korotkova, T.; Fuchs, E.C.; Ponomarenko, A.; von Engelhardt, J.; Monyer, H. NMDA receptor ablation on parvalbumin-positive interneurons impairs hippocampal synchrony, spatial representations, and working memory. Neuron 2010, 68, 557–569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. DeFelipe, J.; López-Cruz, P.L.; Benavides-Piccione, R.; Bielza, C.; Larrañaga, P.; Anderson, S.; Burkhalter, A.; Cauli, B.; Fairén, A.; Feldmeyer, D. New insights into the classification and nomenclature of cortical GABAergic interneurons. Nat. Rev. Neurosci. 2013, 14, 202–216. [Google Scholar] [CrossRef] [Green Version]
  177. Kaar, S.J.; Angelescu, I.; Marques, T.R.; Howes, O.D. Pre-frontal parvalbumin interneurons in schizophrenia: A meta-analysis of post-mortem studies. J. Neural Transm. 2019, 126, 1637–1651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Chung, D.W.; Fish, K.N.; Lewis, D.A. Pathological basis for deficient excitatory drive to cortical parvalbumin interneurons in schizophrenia. Am. J. Psychiatry 2016, 173, 1131–1139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Nakazawa, K.; Zsiros, V.; Jiang, Z.; Nakao, K.; Kolata, S.; Zhang, S.; Belforte, J.E. GABAergic interneuron origin of schizophrenia pathophysiology. Neuropharmacology 2012, 62, 1574–1583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Uhlhaas, P.J.; Singer, W. Abnormal neural oscillations and synchrony in schizophrenia. Nat. Rev. Neurosci. 2010, 11, 100–113. [Google Scholar] [CrossRef]
  181. Gonzalez-Burgos, G.; Cho, R.Y.; Lewis, D.A. Alterations in cortical network oscillations and parvalbumin neurons in schizophrenia. Biol. Psychiatry 2015, 77, 1031–1040. [Google Scholar] [CrossRef] [Green Version]
  182. Antonoudiou, P.; Tan, Y.L.; Kontou, G.; Upton, A.L.; Mann, E.O. Parvalbumin and somatostatin interneurons contribute to the generation of hippocampal gamma oscillations. J. Neurosci. 2020, 40, 7668–7687. [Google Scholar] [CrossRef]
  183. Kriener, B.; Hu, H.; Vervaeke, K. Parvalbumin interneuron dendrites enhance gamma oscillations. Cell Rep. 2022, 39, 110948. [Google Scholar] [CrossRef]
  184. Buzsaki, G.; Wang, X.J. Mechanisms of gamma oscillations. Annu. Rev. Neurosci. 2012, 35, 203–225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Alekseichuk, I.; Turi, Z.; de Lara, G.A.; Antal, A.; Paulus, W. Spatial working memory in humans depends on theta and high gamma synchronization in the prefrontal cortex. Curr. Biol. 2016, 26, 1513–1521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Uhlhaas, P.J.; Singer, W. Oscillations and neuronal dynamics in schizophrenia: The search for basic symptoms and translational opportunities. Biol. Psychiatry 2015, 77, 1001–1009. [Google Scholar] [CrossRef] [PubMed]
  187. Chen, C.-M.A.; Stanford, A.D.; Mao, X.; Abi-Dargham, A.; Shungu, D.C.; Lisanby, S.H.; Schroeder, C.E.; Kegeles, L.S. GABA level, gamma oscillation, and working memory performance in schizophrenia. NeuroImage Clin. 2014, 4, 531–539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Spencer, K.M.; Niznikiewicz, M.A.; Nestor, P.G.; Shenton, M.E.; McCarley, R.W. Left auditory cortex gamma synchronization and auditory hallucination symptoms in schizophrenia. BMC Neurosci. 2009, 10, 1–13. [Google Scholar] [CrossRef] [PubMed]
  189. Spencer, K.M. Baseline gamma power during auditory steady-state stimulation in schizophrenia. Front. Hum. Neurosci. 2012, 5, 190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Hunt, M.J.; Kasicki, S. A systematic review of the effects of NMDA receptor antagonists on oscillatory activity recorded in vivo. J. Psychopharmacol. 2013, 27, 972–986. [Google Scholar] [CrossRef] [PubMed]
  191. Kocsis, B.; Brown, R.E.; McCarley, R.W.; Hajos, M. Impact of ketamine on neuronal network dynamics: Translational modeling of schizophrenia-relevant deficits. CNS Neurosci. Ther. 2013, 19, 437–447. [Google Scholar] [CrossRef]
  192. Hamm, J.P.; Gilmore, C.S.; Clementz, B.A. Augmented gamma band auditory steady-state responses: Support for NMDA hypofunction in schizophrenia. Schizophr. Res. 2012, 138, 1–7. [Google Scholar] [CrossRef] [Green Version]
  193. Sokolenko, E.; Hudson, M.R.; Nithianantharajah, J.; Jones, N.C. The mGluR2/3 agonist LY379268 reverses NMDA receptor antagonist effects on cortical gamma oscillations and phase coherence, but not working memory impairments, in mice. J. Psychopharmacol. 2019, 33, 1588–1599. [Google Scholar] [CrossRef]
  194. Lemercier, C.E.; Holman, C.; Gerevich, Z. Aberrant alpha and gamma oscillations ex vivo after single application of the NMDA receptor antagonist MK-801. Schizophr. Res. 2017, 188, 118–124. [Google Scholar] [CrossRef]
  195. Schuelert, N.; Dorner-Ciossek, C.; Brendel, M.; Rosenbrock, H. A comprehensive analysis of auditory event-related potentials and network oscillations in an NMDA receptor antagonist mouse model using a novel wireless recording technology. Physiol. Rep. 2018, 6, e13782. [Google Scholar] [CrossRef] [PubMed]
  196. Gawande, D.Y.; Narasimhan, K.K.S.; Shelkar, G.P.; Pavuluri, R.; Stessman, H.A.; Dravid, S.M. GluN2D Subunit in Parvalbumin Interneurons Regulates Prefrontal Cortex Feedforward Inhibitory Circuit and Molecular Networks Relevant to Schizophrenia. Biol. Psychiatry, 2023; In Press. [Google Scholar]
  197. Suryavanshi, P.S.; Ugale, R.; Yilmazer-Hanke, D.; Stairs, D.J.; Dravid, S. GluN2C/GluN2D subunit-selective NMDA receptor potentiator CIQ reverses MK-801-induced impairment in prepulse inhibition and working memory in Y-maze test in mice. Br. J. Pharmacol. 2014, 171, 799–809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. de la Fuente-Sandoval, C.; León-Ortiz, P.; Azcárraga, M.; Stephano, S.; Favila, R.; Díaz-Galvis, L.; Alvarado-Alanis, P.; Ramírez-Bermúdez, J.; Graff-Guerrero, A. Glutamate levels in the associative striatum before and after 4 weeks of antipsychotic treatment in first-episode psychosis: A longitudinal proton magnetic resonance spectroscopy study. JAMA Psychiatry 2013, 70, 1057–1066. [Google Scholar] [CrossRef] [PubMed]
  199. Levitt, J.J.; Nestor, P.G.; Levin, L.; Pelavin, P.; Lin, P.; Kubicki, M.; McCarley, R.W.; Shenton, M.E.; Rathi, Y. Reduced structural connectivity in frontostriatal white matter tracts in the associative loop in schizophrenia. Am. J. Psychiatry 2017, 174, 1102–1111. [Google Scholar] [CrossRef]
  200. Egerton, A.; Chaddock, C.A.; Winton-Brown, T.T.; Bloomfield, M.A.; Bhattacharyya, S.; Allen, P.; McGuire, P.K.; Howes, O.D. Presynaptic striatal dopamine dysfunction in people at ultra-high risk for psychosis: Findings in a second cohort. Biol. Psychiatry 2013, 74, 106–112. [Google Scholar] [CrossRef]
  201. Kegeles, L.S.; Abi-Dargham, A.; Frankle, W.G.; Gil, R.; Cooper, T.B.; Slifstein, M.; Hwang, D.-R.; Huang, Y.; Haber, S.N.; Laruelle, M. Increased synaptic dopamine function in associative regions of the striatum in schizophrenia. Arch. Gen. Psychiatry 2010, 67, 231–239. [Google Scholar] [CrossRef] [Green Version]
  202. McCutcheon, R.A.; Jauhar, S.; Pepper, F.; Nour, M.M.; Rogdaki, M.; Veronese, M.; Turkheimer, F.E.; Egerton, A.; McGuire, P.; Mehta, M.M. The topography of striatal dopamine and symptoms in psychosis: An integrative positron emission tomography and magnetic resonance imaging study. Biol. Psychiatry Cogn. Neurosci. Neuroimaging 2020, 5, 1040–1051. [Google Scholar] [CrossRef]
  203. 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]
  204. Huang, Z.; Gibb, A.J. Mg2+ block properties of triheteromeric GluN1–GluN2B–GluN2D NMDA receptors on neonatal rat substantia nigra pars compacta dopaminergic neurones. J. Physiol. 2014, 592, 2059–2078. [Google Scholar] [CrossRef]
  205. Sitzia, G.; Mantas, I.; Zhang, X.; Svenningsson, P.; Chergui, K. NMDA receptors are altered in the substantia nigra pars reticulata and their blockade ameliorates motor deficits in experimental parkinsonism. Neuropharmacology 2020, 174, 108136. [Google Scholar] [CrossRef]
  206. Byrial, P.; Nyboe, L.; Thomsen, P.H.; Clausen, L. Motor impairments in early onset schizophrenia. Early Interv. Psychiatry 2022, 16, 481–491. [Google Scholar] [CrossRef]
  207. Nadesalingam, N.; Chapellier, V.; Lefebvre, S.; Pavlidou, A.; Stegmayer, K.; Alexaki, D.; Gama, D.B.; Maderthaner, L.; von Känel, S.; Wüthrich, F. Motor abnormalities are associated with poor social and functional outcomes in schizophrenia. Compr. Psychiatry 2022, 115, 152307. [Google Scholar] [CrossRef] [PubMed]
  208. Peralta, V.; Campos, M.S.; De Jalón, E.G.; Cuesta, M.J. Motor behavior abnormalities in drug-naïve patients with schizophrenia spectrum disorders. Mov. Disord. 2010, 25, 1068–1076. [Google Scholar] [CrossRef]
  209. van Hooijdonk, C.F.; van der Pluijm, M.; Bosch, I.; van Amelsvoort, T.A.; Booij, J.; de Haan, L.; Selten, J.-P.; van de Giessen, E. The substantia nigra in the pathology of schizophrenia: A review on post-mortem and molecular imaging findings. Eur. Neuropsychopharmacol. 2023, 68, 57–77. [Google Scholar] [CrossRef]
  210. Abram, S.V.; Roach, B.J.; Holroyd, C.B.; Paulus, M.P.; Ford, J.M.; Mathalon, D.H.; Fryer, S.L. Reward processing electrophysiology in schizophrenia: Effects of age and illness phase. NeuroImage Clin. 2020, 28, 102492. [Google Scholar] [CrossRef]
  211. Strauss, G.P.; Waltz, J.A.; Gold, J.M. A review of reward processing and motivational impairment in schizophrenia. Schizophr. Bull. 2014, 40 (Suppl. 2), S107–S116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. White, D.M.; Kraguljac, N.V.; Reid, M.A.; Lahti, A.C. Contribution of substantia nigra glutamate to prediction error signals in schizophrenia: A combined magnetic resonance spectroscopy/functional imaging study. Npj Schizophrenia 2015, 1, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Salimando, G.J.; Hyun, M.; Boyt, K.M.; Winder, D.G. BNST GluN2D-containing NMDA receptors influence anxiety-and depressive-like behaviors and modulatecell-specific excitatory/inhibitory synaptic balance. J. Neurosci. 2020, 40, 3949–3968. [Google Scholar] [CrossRef]
Figure 1. Mechanisms by which dysfunction of GluN2D-containing NMDARs could potentially result in symptoms of schizophrenia. This simplified circuit shows that, in healthy individuals, GluN2D-containing NMDARs stimulate GABAergic interneurons on which they are expressed, which, in turn, mediate inhibition and coordinate the synchronized firing of networks of excitatory pyramidal neurons. These pyramidal neurons, in turn, stimulate dopaminergic (DA) neurons in the midbrain (via the ventral striatum and ventral pallidum, not shown) which project to the associative striatum. In healthy individuals, this circuit is under homeostatic balance, leading to gamma oscillations and healthy cognition. In schizophrenia, hypofunction of the GluN2D-containing NMDARs could lead to GABAergic interneurons increasing excitation of pyramidal neurons by reducing inhibition (disinhibition) onto pyramidal neurons. The resulting aberrant increase in cortical excitation would lead to abnormal neuronal oscillations and impaired cognition. Moreover, this hyperactivity might lead to an overdrive in midbrain DA neurons and enhanced striatal DA, which has been linked to the positive symptoms of schizophrenia.
Figure 1. Mechanisms by which dysfunction of GluN2D-containing NMDARs could potentially result in symptoms of schizophrenia. This simplified circuit shows that, in healthy individuals, GluN2D-containing NMDARs stimulate GABAergic interneurons on which they are expressed, which, in turn, mediate inhibition and coordinate the synchronized firing of networks of excitatory pyramidal neurons. These pyramidal neurons, in turn, stimulate dopaminergic (DA) neurons in the midbrain (via the ventral striatum and ventral pallidum, not shown) which project to the associative striatum. In healthy individuals, this circuit is under homeostatic balance, leading to gamma oscillations and healthy cognition. In schizophrenia, hypofunction of the GluN2D-containing NMDARs could lead to GABAergic interneurons increasing excitation of pyramidal neurons by reducing inhibition (disinhibition) onto pyramidal neurons. The resulting aberrant increase in cortical excitation would lead to abnormal neuronal oscillations and impaired cognition. Moreover, this hyperactivity might lead to an overdrive in midbrain DA neurons and enhanced striatal DA, which has been linked to the positive symptoms of schizophrenia.
Ijms 24 11835 g001
Table 1. Summary of post-mortem studies of GluN2D NMDAR subunit expression in schizophrenia.
Table 1. Summary of post-mortem studies of GluN2D NMDAR subunit expression in schizophrenia.
ReferenceOutcome MeasureBrain Region Direction of Change
[128] mRNAPrefrontal cortex and cerebellum ↑ in PFC
↔ in cerebellum
[152]Receptor binding and mRNA Dorsolateral prefrontal cortex
[153]ProteinDorsolateral prefrontal cortex and anterior cingulate cortex
[154] mRNAMedial dorsal thalamus↓ in glutamatergic relay neurons
↔ in mixed glial-interneuronal cells
[158] Receptor binding and mRNACerebellum↑ mRNA in the right cerebellum
↔ mRNA in left cerebellum and receptor binding
PFC = prefrontal cortex, mRNA = messenger ribonucleic acid. ↑ = increased expression, ↓ = decreased expression, ↔ = no change.
Table 2. Summary of schizophrenia-relevant behavioural phenotypes in GluN2D-knockout mice.
Table 2. Summary of schizophrenia-relevant behavioural phenotypes in GluN2D-knockout mice.
ReferenceSexPharmacological Manipulation and DoseBehavioural Domain
Examined
Behavioural Tests UsedMain Findings
[159]Not specifiedN/ALocomotion, anxiety, novelty preferenceOpen field test, novelty preference test, light–dark compartment test, elevated plus-maze↓ spontaneous locomotion,
↓ novelty preference,
↔ change in anxiety
[160]Not specifiedN/ALocomotion, anxietyOpen field test, light–dark compartment test, elevated plus-maze, forced swim test↓ spontaneous locomotion,
↓ anxiety
[163]MaleAcute and chronic PCP (3 mg/kg)LocomotionOpen field test↓ PCP-induced hyperlocomotion
[161]MaleN/ALocomotion, contextual fear memory, spatial memoryOpen field test, fear conditioning test, Y-maze↓ spontaneous locomotion,
↓ contextual fear memory,
↓ spatial memory
[165]Male and femaleAcute PCP (3 mg/kg)Motor functionRotarod↓ PCP-induced motor impairment
[137]MaleAcute ketamine (30 mg/kg)Locomotion, spatial memoryOpen field test, Morris water maze↓ ketamine-induced hyperlocomotion,
↓ spatial memory acquisition
[166]MaleSubchronic ketamine (25 mg/kg)LocomotionOpen field test↓ ketamine-induced hyperlocomotion
[162]Not specifiedAcute (RS)-ketamine (10 or 20 mg/kg),
(R)-ketamine (10 or 20 mg/kg), (S)-ketamine (10 or 20 mg/kg)
Novel object recognition taskNovel object recognition task↓ (R)-ketamine-induced novel object recognition deficits
N/A = not applicable, PCP = phencyclidine. ↓ = decreased, ↔ = no change.
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

Vinnakota, C.; Hudson, M.R.; Jones, N.C.; Sundram, S.; Hill, R.A. Potential Roles for the GluN2D NMDA Receptor Subunit in Schizophrenia. Int. J. Mol. Sci. 2023, 24, 11835. https://doi.org/10.3390/ijms241411835

AMA Style

Vinnakota C, Hudson MR, Jones NC, Sundram S, Hill RA. Potential Roles for the GluN2D NMDA Receptor Subunit in Schizophrenia. International Journal of Molecular Sciences. 2023; 24(14):11835. https://doi.org/10.3390/ijms241411835

Chicago/Turabian Style

Vinnakota, Chitra, Matthew R. Hudson, Nigel C. Jones, Suresh Sundram, and Rachel A. Hill. 2023. "Potential Roles for the GluN2D NMDA Receptor Subunit in Schizophrenia" International Journal of Molecular Sciences 24, no. 14: 11835. https://doi.org/10.3390/ijms241411835

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