Next Article in Journal
Injectable Tannin-Containing Hydroxypropyl Chitin Hydrogel as Novel Bioactive Pulp Capping Material Accelerates Repair of Inflamed Dental Pulp
Previous Article in Journal
Oxylipins in Aqueous Humor of Primary Open-Angle Glaucoma Patients
Previous Article in Special Issue
Role of NMDA Receptor in High-Pressure Neurological Syndrome and Hyperbaric Oxygen Toxicity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Dysfunction of the NMDA Receptor in the Pathophysiology of Schizophrenia and/or the Pathomechanisms of Treatment-Resistant Schizophrenia

Department of Neuropsychiatry, Division of Neuroscience, Graduate School of Medicine, Mie University, Tsu 514-8507, Japan
*
Author to whom correspondence should be addressed.
Biomolecules 2024, 14(9), 1128; https://doi.org/10.3390/biom14091128
Submission received: 15 August 2024 / Revised: 29 August 2024 / Accepted: 2 September 2024 / Published: 6 September 2024
(This article belongs to the Special Issue NMDA Receptor in Health and Diseases: 2nd Edition)

Abstract

:
For several decades, the dopamine hypothesis contributed to the discovery of numerous typical and atypical antipsychotics and was the sole hypothesis for the pathophysiology of schizophrenia. However, neither typical nor atypical antipsychotics, other than clozapine, have been effective in addressing negative symptoms and cognitive impairments, which are indices for the prognostic and disability outcomes of schizophrenia. Following the development of atypical antipsychotics, the therapeutic targets for antipsychotics expanded beyond the blockade of dopamine D2 and serotonin 5-HT2A receptors to explore the partial agonism of the D2 receptor and the modulation of new targets, such as D3, 5-HT1A, 5-HT7, and metabotropic glutamate receptors. Despite these efforts, to date, psychiatry has not successfully developed antipsychotics with antipsychotic properties proven to be superior to those of clozapine. The glutamate hypothesis, another hypothesis regarding the pathophysiology/pathomechanism of schizophrenia, was proposed based on clinical findings that N-methyl-D-aspartate glutamate receptor (NMDAR) antagonists, such as phencyclidine and ketamine, induce schizophrenia-like psychotic episodes. Large-scale genome-wide association studies (GWASs) revealed that approximately 30% of the risk genes for schizophrenia (the total number was over one hundred) encode proteins associated with glutamatergic transmission. These findings supported the validation of the glutamate hypothesis, which was inspired by the clinical findings regarding NMDAR antagonists. Additionally, these clinical and genetic findings suggest that schizophrenia is possibly a syndrome with complicated pathomechanisms that are affected by multiple biological and genetic vulnerabilities. The glutamate hypothesis has been the most extensively investigated pathophysiology/pathomechanism hypothesis, other than the dopamine hypothesis. Studies have revealed the possibility that functional abnormalities of the NMDAR play important roles in the pathophysiology/pathomechanism of schizophrenia. However, no antipsychotics derived from the glutamatergic hypothesis have yet been approved for the treatment of schizophrenia or treatment-resistant schizophrenia. Considering the increasing evidence supporting the potential pro-cognitive effects of glutamatergic agents and the lack of sufficient medications to treat the cognitive impairments associated with schizophrenia, these previous setbacks cannot preclude research into potential novel glutamate modulators. Given this background, to emphasize the importance of the dysfunction of the NMDAR in the pathomechanism and/or pathophysiology of schizophrenia, this review introduces the increasing findings on the functional abnormalities in glutamatergic transmission associated with the NMDAR.

1. Introduction

Antipsychotic medication for schizophrenia is one of the most successful therapies in psychiatry. Despite these efforts, based on analyses of regional and national prevalence, disability-adjusted life years (DALYs), and years lived with disability (YLDs) for numerous diseases/disorders from 1990 to 2019, the Global Burden of Disease Study 2019 ranked schizophrenia 20th among the top 25 leading causes of YLDs (depressive and anxiety disorders were ranked 2nd and 8th, respectively). The age-standardized DALY rate among overall mental disorders was over 12% [1]. These data raise the concern that, despite the successful development of numerous antipsychotics, these efforts in psychiatry and psychopharmacology may be making only a small contribution to improving the quality of life of patients with schizophrenia. Although the impact of schizophrenia affects a smaller proportion of the global population than depressive and anxiety disorders, the weight of disability during the acute psychotic state of schizophrenia was speculated to be the highest across in the Global Burden of Disease Study 2019 [1,2]. Notably, unlike depressive or anxiety disorders, variations in schizophrenia are not observed across regions or sexes, with disability continuing into old age [1].
Following the successful discovery of typical antipsychotics, such as phenothiazines and butyrophenones, the scientific basis for these medications was summarized in the dopamine hypothesis of schizophrenia [3,4,5,6,7,8]. The dopamine hypothesis has been repeatedly/continuously revised, and the majority of psychiatrists have been convinced that appropriate pharmacological therapies have contributed to re-integrating patients with schizophrenia into society [9]. However, in the 1980s, psychiatrists recognized that the major disability in schizophrenia consists of not only positive symptoms, which are the primary targets of typical antipsychotics but also cognitive impairments and negative symptoms, which are less responsive to typical antipsychotics [10,11,12]. Typical antipsychotics relieve positive symptoms, such as hallucinations and delusions, via the inhibition of the dopamine D2 receptor. However, typical antipsychotics are not only ineffective for treating negative symptoms and cognitive impairments but also the potential causes of their exacerbation [13,14]. Psychiatry and psychopharmacology improved the dopamine hypothesis to develop more effective antipsychotics and dialed other monoamine receptor targets into the basic dopamine D2 mechanism, including the antagonism of serotonin receptors, to develop atypical antipsychotics (the revised dopamine hypothesis) [15,16].
Despite the increasing number of approved antipsychotics, several meta-analyses that compared typical and atypical antipsychotics revealed no significant differences in their efficacy, with the exception of clozapine, which is significantly more effective than other conventional antipsychotics [13,17]. Based on this historical background and the insufficient efficacy of conventional antipsychotics for the treatment of negative symptoms and cognitive impairments, alternative approaches to developing novel antipsychotics involving targets other than monoamine receptors have been attempted. Unfortunately, psychiatry has yet to produce novel antipsychotics with major targets other than monoaminergic transmission. However, the successful development of several novel types of antipsychotics is becoming more realistic, including KarXT (muscarinic acetylcholine receptor agonist) [18,19,20,21], ulotaront (agonist of trace-amine-associated receptors) [22,23], and iclepertin (glycine-transporter inhibitor) [24,25].
Large-scale genome-wide association studies (GWAS) identified over 100 risk genes for schizophrenia. Notably, approximately 30% of them encoded proteins that are associated with glutamatergic transmission [26,27,28]. For some time, it has been well-known that patients with schizophrenia can be resistant to treatment with typical and atypical antipsychotic medications [29,30]. Clozapine, the sole approved antipsychotic for patients with treatment-resistant schizophrenia (TRS) [31,32,33], has a lower affinity for the dopamine D2 receptor in comparison to conventional typical/atypical antipsychotics and enhances glutamatergic transmission [34,35,36,37,38]. These accumulating findings from psychopharmacological, neurobiological, and molecular biology studies suggest that the core of the pathophysiology and/or pathomechanism in a proportion of patients with schizophrenia probably involves the dysfunction of N-methyl-D-aspartate glutamate receptors (NMDARs). Despite intensive research, glutamatergic antipsychotics, according to the glutamate hypothesis, have not been approved for the treatment of schizophrenia to date. However, considering the increasing clinical and preclinical findings supporting the potential pro-cognitive effects of glutamatergic agents and the lack of sufficient treatment for the negative symptoms or cognitive impairments associated with schizophrenia, these previous setbacks cannot preclude research into potential glutamate modulators as novel antipsychotics. Given this background, to emphasize the importance of the glutamate hypothesis, this review introduces the increasing findings regarding functional abnormalities in glutamatergic transmission associated with the NMDAR in schizophrenia and/or TRS. Additionally, this review also mentions a new challenge in psychiatry, namely ultra-treatment-resistant schizophrenia (UTRS) [29,30,39]. UTRS is a recently proposed concept characterizing the specific clinical features of clozapine resistance [29,30,39]. However, the actual pathomechanism of UTRS is still unknown, as the actual pathophysiology of clozapine currently remains to be clarified.

2. NMDAR Function and Modulation

Glutamate is the most predominant excitatory transmitter in the brain. Glutamate receptors are categorized as ionotropic or G-protein-coupled metabotropic receptors. Ionotropic glutamate receptors are further divided into three subtypes based on their ligands, namely N-methyl-D-aspartate (NMDAR), α-amino-3-hydroxy-5-methylisoxaole-4-propionate (AMPAR), and kainic acid (KAR) receptors. Ionotropic glutamate receptors are tetramers (composed of four subunits). The majority of AMPARs and KARs are mainly permeable to sodium and potassium, but calcium permeability is energetically unfavorable. However, among ionotropic glutamate receptors, NMDARs can allow the inflow of sodium, potassium, and calcium, depending on the plasma-membrane potential, and exhibit slower and incomplete desensitization in comparison to AMPARs and KARs [40]. Furthermore, the penetration of calcium through NMDAR affects several signalings as a second messenger [41]. Because of these cation permeabilities, ionotropic glutamate receptor hyperactivation contributes to neurotoxicity. Therefore, the appropriate modulation/activation of ionotropic glutamate receptors, including the NMDAR, requires pharmacological manipulation that does not induce NMDAR neurotoxicity. A schematic of the NMDAR structure is displayed in Figure 1 to explain its various pharmacological modulatory sites.
The NMDAR is composed of a heterotetrameric complex with two GluN1 and two GluN2 subunits or two GluN1, one GluN2, and one GluN3 subunit. Several ligand binding sites for regulating ion channel activity and their selective ligands have been identified, such as the glutamate binding site in GluN2 (L-glutamate), the glycine modulatory binding site in GluN1/GluN3 (D-serine, glycine, and kynurenic acid), the Zn2+ binding site in GluN2A/GluN2B (Zn2+, ifenprodil, and Ro25-6981), the redox-regulating glutathione binding site (in GluN1/GluN2A), the spermine binding site in GluN2B, and the phencyclidine binding site in the transmembrane domain (Mg2+, MK801, ketamine, and phencyclidine) [41,42,43,44,45]. The activation of the NMDAR requires three events, namely the removal of Mg2+ from the transmembrane domain (Mg2+ blocks the cation channel pore) due to postsynaptic depolarization (usually due to activated AMPAR), the binding of glycine/D-serine to the glycine modulatory site in GluN1, and the binding of L-glutamate to the glutamine binding site in GluN2 [40]. In other words, the NMDAR has unique gating modes, namely ligand gated and voltage gated.
To date, numerous enhancers of ionotropic glutamate receptors have been synthesized [46,47]. The expression patterns of NMDARs seem to be regulated by the GluN2 subunit [48,49,50,51]. GluN2A is highly expressed in the hippocampus/cortex and modestly expressed in the midbrain, cerebellum, striatum, and brainstem [48], and GluN2B is predominantly expressed in the hippocampus, cortex, and striatum [49]. GluN2C is mainly expressed in the cerebellum and thalamus [50], and GluN2D is primarily expressed in the brainstem and forebrain [51]. Based on these expression patterns, positive allosteric modulators (PAMs) of GluN1/GluN2A were developed [46].
A positive allosteric modulator (PAM) is a ligand that enhances the function of a receptor in the presence of an agonist but is unable to directly activate the receptor without the agonist. PAMs can increase NMDAR activity by enhancing the binding of endogenous agonists to NMDARs. Therefore, PAMs are considered to have broader therapeutic applications than agonists due to the lower risk of neurotoxicity induced by NMDAR hyperactivation. First-generation NMDAR PAMs have been pursued as therapeutics, as NMDAR hypofunction has been identified in multiple brain diseases. Several NMDAR PAMs have entered clinical trials, providing human data on the effects of NMDAR potentiation. CAD-9303 has been studied in patients with schizophrenia (NCT04306146). In addition, AGE-718 is being studied in people with Huntington’s disease (NCT05358821 and NCT05107128), Parkinson’s disease (NCT05318937), and Alzheimer’s disease (NCT04602624).
Based on the expression patterns of NMDARs and the advantages of PAMs, positive allosteric modulators of GluN1/GluN2A have been developed, such as thienopyrimidin-4-ones, thiazole pyrimidinone, benzofuran analogs, and benzohydrazide analogs [46]. In order to proceed to clinical trials, expanding the structural diversity of lead compounds, enhancing pharmacodynamic features by optimizing the balance between water solubility and lipophilicity to improve bioavailability, and improving selectivity toward individual GluN2 subunits should be achieved [46,47,52].

3. Clinical Findings Supporting the Glutamate Hypothesis

3.1. Initial Clinical Findings

The most impactful initial clinical findings were observed in individuals who abused the NMDA antagonists phencyclidine and ketamine [53,54]. Both phencyclidine and ketamine produce not only positive symptoms (including auditory hallucinations) but also negative symptoms and cognitive impairments for up to 2 weeks [54]. Furthermore, the intake of phencyclidine and ketamine in patients with schizophrenia exacerbates its symptoms [55,56]. It was reported that 6.5% of patients who fulfilled the diagnostic criteria for schizophrenia at the first episode of schizophrenia-like psychosis were positive for anti-NMDAR antibodies [57]. A meta-analysis study of seven reports also demonstrated that anti-NMDAR antibodies were detected in 8% of patients with schizophrenia (115/1441) [58]. Furthermore, a recent OPTIMISE project also reported that approximately 5% of first-episode psychosis patients without antipsychotic exposure were positive for anti-NMDAR antibodies [59]. Notably, the serum levels of anti-NMDAR antibodies in first-episode patients with schizophrenia were positively correlated with the severity of schizophrenia using PANS scores [60]. These clinical findings suggest that the psychotic symptoms induced by NMDAR antagonists and anti-NMDAR autoantibodies are similar to those of schizophrenia [61].
NMDAR antagonists directly inhibit the permeability of channel pores, whereas NMDAR antibodies have no direct effect on channel-pore function [62]. However, anti-NMDAR antibodies lead to the internalization/downregulation of NMDARs [59,63,64,65]. These preclinical findings suggest that decreasing NMDAR activity by either decreasing functional NMDARs in the plasma membrane or directly inhibiting channels plays an important role in the negative symptoms and cognitive impairments caused by schizophrenia. Based on clinical and preclinical findings, encephalitis induced by anti-NMDAR antibodies has been established as an autoimmune disorder with schizophrenia-like symptoms [63]. Clinical findings showing that NMDAR antagonists and anti-NMDAR antibodies lead to schizophrenia-like symptoms contributed to the development of the glutamate hypothesis of the pathophysiology and pathomechanisms of schizophrenia.
The NMDAR antagonist esketamine is approved for treatment-resistant depression, though not for schizophrenia, and is a more rapid-acting antidepressant in comparison to monoamine-transporter inhibitors [66]. Anecdotal information about ayahuasca, which is a psychedelic plant brew originating from the Amazon rainforest, varies significantly, ranging from evangelical accounts to horror stories involving physical and psychological harm [67]. The antidepressive effects of ayahuasca were first reported in 1984, and a candidate mechanism underlying its antidepressive action was speculated to be monoamine oxidase inhibition [68]. A systematic review also highlighted the antidepressive effects of ayahuasca. However, the review was limited by small sample sizes and a lack of healthy controls [69]. Harmine, which is one of the abundant β-carboline alkaloids in ayahuasca, enhances glutamate uptake by increasing glutamate-transporter expression in animal models [70,71]. In addition, clinical evidence shows that the partial NMDAR agonist D-cycloserine induces a depressive mood [72]. Therefore, attenuating glutamatergic transmission, likely ketamine/esketamine and ayahuasca, plays an important role in the improvement of a depressive mood.

3.2. Post Mortem Brain Studies

Post mortem brain studies have shown variable changes in the expression of NMDARs (protein and mRNA) and some of the postsynaptic molecules associated with NMDAR signaling in patients with schizophrenia. Both the mRNA and protein expressions of the GluN1 and GluN2C subunits in the prefrontal cortices of patients with schizophrenia were significantly decreased in comparison to that in healthy controls, whereas changes in the other subunits, such as GluN2A, GluN2B, GluN2D, and GluN3, in the prefrontal cortex were not observed [73,74]. Decreased expression of GluN1 mRNA in the dentate gyrus of patients with schizophrenia compared to controls was also reported [75]. Furthermore, the expression of several NMDAR-related postsynaptic density molecules was also altered in the post mortem brains of patients with schizophrenia. The protein and mRNA expressions of both postsynaptic density protein-93 (PDS-93) and PDS-95 were decreased in the anterior cingulate cortex [73,76,77,78]. Other studies also revealed altered levels of glutamate, N-acetylaspartylglutamate, kynurenic acid, and the activity of glutamate carboxypeptidase II in schizophrenia [79,80,81,82] (N-acetylaspartyl-glutamate, which is degraded by glutamate carboxypeptidase II, is an endogenous NMDAR antagonist and mGlu3R agonist [83,84]). Increased levels of endogenous NMDAR antagonists, N-acetylaspartyl-glutamate and kynurenic acid [85,86,87], and/or the decreased activity of glutamate carboxypeptidase II in schizophrenia may support the hypothesis that N-acetylaspartyl-glutamate signaling is involved in NMDAR hypofunction in schizophrenia [79,80,81,82,83,84].

3.3. Genome-Wide Association Studies

Over the past 15 years, several large-scale genome-wide association studies (GWASs) of schizophrenia have been conducted; the largest-scale GWAS included over 76,000 patients with schizophrenia and 240,000 healthy controls [26,27,28]. Most of the identified single-nucleotide polymorphisms (SNPs) associated with schizophrenia were located in non-coding regions. However, notably, approximately 30% of them encoded proteins related to glutamatergic transmission [27,88]. Copy-number variants (CNVs), in which kilobases of DNA are deleted or repeated, linked to the risk of schizophrenia have been found to be enriched with gene-encoding proteins associated with glutamatergic synapses, and the postsynaptic density has also been implicated [89]. Fine-mapping/exome sequencing also identified coding variants associated with substantial risks for schizophrenia in the NMDAR (GluN2A), the AMPAR (GluA3), the KAR (GluK2 and GluK4), and PSD-95-associated genes (SHANK, NLGN, and DLGAP) and serine racemase (SR) [90,91].

3.4. Neurophysiological Findings

Proton magnetic resonance spectroscopy (MRS) and positron emission tomography (PET) enable the non-invasive visualization of the brain and the measurement of functions of numerous transmitters, including dopamine and glutamate, in living human subjects in vivo. Regional cerebral blood flow in the frontal cortices of patients with schizophrenia was predominantly increased by ketamine administration compared to healthy controls, as observed using PET [92]. An MRS study showed increased glutamatergic transmission in the limbic system and basal ganglia, glutamine levels in the thalamus, and levels of glutamate plus glutamine in the basal ganglia and medial temporal cortex [92,93,94,95,96]. Additionally, ketamine administration in healthy controls caused an acute increase in glutamine that was marginally related to cognitive function [97,98]. These findings possibly support the notion that people with schizophrenia are hypersensitive to NMDAR antagonists compared to healthy controls.

3.5. Psychophysiological Findings

Event-related potentials detected using both electroencephalography and magnetoencephalography enable the high-resolution temporal monitoring of physiological brain function. Mismatch negativity and P300 are considered to be indices for auditory sensory memory/pre-attentive processing and re-orienting/covert attention shifting, respectively [98,99,100]. Currently, reductions in the mismatch negativity and P300 amplitude are both hypothesized to reflect NMDAR hypofunction and are considered candidate biomarkers for schizophrenia [101,102]. Indeed, the administration of ketamine is known to attenuate the mismatch negativity and P300 amplitude in healthy subjects [103,104]. A reduction in the mismatch negativity amplitude has also been observed in patients with schizophrenia [105,106].

4. Preclinical Findings Supporting the Glutamate Hypothesis

4.1. Genetic Animal Model

Exploratory studies of functional abnormalities in genetic animal models with the knockout of serine racemase (SR-KO) have contributed to the acquisition of numerous findings about the pathophysiology and pathomechanisms of schizophrenia. SR, which converts L-serine to D-serine [107], is one of the risk genes for schizophrenia identified by GWASs [26,27,28]. Therefore, the findings of functional abnormalities in SR-KO can be considered as a “proof of principle” that glutamatergic transmission is a candidate therapeutic target for schizophrenia.
SR−KO exhibits a reduction in cortical D-serine (90%) and NMDAR-sensitive long-term potentiation (70%) [108,109,110]. Various behavioral abnormalities associated with schizophrenia have also been detected, such as impairments in memory components and hyperactivity [109,110,111,112]. A quantitative MRI study detected cortical atrophy (5%) and increased ventricle volume (20%) [113]. Golgi staining demonstrated decreased dendritic complexity and synaptic spine density in several cortical regions and the hippocampus [109,111,114], suggesting the loss of approximately 30% of cortical glutamatergic synapses [114,115]. Additionally, decreased parvalbumin in cortical GABAergic neurons and perineural nets surrounding the GABAergic neurons were also observed in SR-KO [116]. The normalization of D-serine levels in the brain induced by the chronic administration of D-serine improved some of the observed abnormalities in SR-KO, such as long-term potentiation, the reversal of memory deficits, the correction of cortical neurochemical abnormalities, and the partial restoration of dendritic spines [109,117]. Therefore, these findings suggest that SR-KO is a pharmacologically valid candidate genetic animal model [118] for schizophrenia [88]. Considering that the NMDAR requires three events for activation, namely the depolarization of the plasma membrane, the binding of D-serine/glycine to the glycine modulatory site in GluN1, and the binding of L-glutamate to the glutamine binding site in GluN2 [40], the glycine modulatory site in GluN1 plays an important role in the pathophysiology of schizophrenia or cognition.
Recently, several functional abnormalities, such as hypoactivity in the mPFC, increased dopaminergic signaling in the striatum, hypersensitivity to amphetamine-induced hyperlocomotion, fewer molecules associated with glutamate receptors in the glutamatergic synapse, and aberrant locomotor patterns opposite to those associated with antipsychotic-induced behavior, were observed in GluN2A knockout mice (GluN2A-KO) [119]. These findings from two genetic animal models, SR-KO and GluN2A-KO, suggest that glutamatergic transmission hypofunction plays an important role in the pathomechanisms of schizophrenia.

4.2. Pharmacological Animal Model

The systemic administration of NMDAR antagonists (phencyclidine, ketamine, and MK801) increased the release of dopamine [120,121], serotonin [121,122], and norepinephrine [123,124] in the mPFC. Local NMDAR antagonist administration into the mPFC also increased the release of dopamine, serotonin, and norepinephrine in the mPFC [121,125,126,127], but these increases in monoamine release in the mPFC were inhibited by the local administration of muscimol (a GABAA receptor agonist) into the mPFC [121,125,126,127]. These findings suggest that the presynaptic terminals of monoaminergic projections are subject to GABAergic inhibition, which is positively regulated by the NMDAR. Furthermore, the local administration of an NMDAR antagonist into the ventral tegmental area (VTA), locus coeruleus (LC), and dorsal raphe nucleus (DRN) also increased the release of dopamine, norepinephrine, and serotonin in the mPFC. These increases in monoamine release in the mPFC were also inhibited by the local administration of muscimol into the VTA, LC, and DRN [120,125,126,127,128,129,130,131]. Therefore, monoaminergic neurons in the VTA, LC, and DRN are also subject to GABAergic inhibition, which is positively regulated by the NMDAR, similar to the mPFC.
Similar to monoaminergic systems, systemic NMDAR antagonist administration also increased the release of L-glutamate in the mPFC [120,128,132]. However, the local administration of neither the NMDAR antagonist nor muscimol into the mPFC affected L-glutamate release in the mPFC [120,121,129,132]. In contrast, the local administration of an NMDAR antagonist into the mediodorsal (MDTN) and reticular thalamic nuclei (RTN) drastically increased L-glutamate release in the frontal cortex [120,122,131]. A detailed analysis of the transmission pathway among the RTN, MDTN, and mPFC indicated that the RTN projects GABAergic terminals into the MDTN (but not the mPFC), and the MDTN projects glutamatergic terminals into the mPFC [66,120,122,131,133,134]. These observations suggest that an impaired NMDAR function in the RTN leads to a GABAergic disinhibition of glutamatergic neurons in the MDTN, resulting in increased glutamate release in the mPFC.
These results indicate that the attenuation of the NMDAR leads to the increased transmission of monoamine and glutamate in the mPFC, but the mechanisms of monoaminergic and glutamatergic transmission in the mPFC are not identical. Notably, thalamocortical glutamatergic transmission is mainly regulated by intrathalamic GABAergic transmission, which is enhanced by the activation of thalamocortical glutamatergic transmission via NMDARs in the RTN [124,131,135,136]. Interestingly, monoaminergic inputs to the RTN are positively regulated via excitatory 5-HT7 receptors and α1 adrenoceptors [124,131,135,136].

5. Treatment-Resistant Schizophrenia and Limitation of Conventional Antipsychotics

5.1. Treatment-Resistant Schizophrenia (TRS)

Treatment-resistant schizophrenia (TRS) and the limitations of conventional antipsychotics are major issues in psychiatry. Since the approval of atypical antipsychotics, this topic has become more actively discussed [137,138,139]. The concept of TRS is tied to two clinical issues, namely a lack of improvement in psychotic episodes in patients with schizophrenia and criteria for initiating clozapine medication. Several studies have provided important clinical insights into TRS. Patients with TRS are more often of European descent [140,141] and affected by the paranoid subtype [141,142]. Compared with non-TRS, the onset age was younger [141,142,143,144], and premorbid social functioning was poorer in TRS [145]. The largest population-based cohort study to identify the clinical features of TRS was based on Danish national registry data, which were used to compare patients with TRS to all other patients diagnosed with schizophrenia over a ten-year period [142]. This study found that individuals with TRS were more likely to have a comorbid personality disorder, a more rural residence, more schooling, and a previous history of suicide attempts [142]. At the time of first schizophrenia diagnosis, patients with TRS were more likely to be inpatients, to have required more psychotropic medications in the previous year, and to have spent more than 30 days in the psychiatric hospital in the previous year [142].
Defining TRS, which means conventional antipsychotic-resistant schizophrenia, has been a challenge for the field, and numerous guidelines with potential criteria for TRS have been published [29,30,146,147] (Table 1). The Treatment Response and Resistance in Psychosis (TRIPP) Working Group provided the newest consensus criteria for TRS and guidelines for the treatment of TRS (Table 1) [29]. The TRIPP group reported that 95% of the previous studies adopted different criteria to define TRS (Table 1), 50% reported operationalized criteria, and 60% required that patients had not responded to at least two adequate treatment trials and defined adequate treatment as lasting for at least 6 weeks, the typical time for a medication response. However, 50% of these studies did not report the dosage used but instead stated only “adequate dose” [29]. These results emphasize the need for more rigorous and standardized criteria for TRS.

5.2. Ultra-Treatment-Resistant Schizophrenia (UTRS)

Traditionally, more than 30% of patients with schizophrenia are considered to have TRS [29,30]. Clozapine is the sole approved antipsychotic for TRS, owing to its higher efficacy in the treatment of TRS in comparison with typical antipsychotics (approximately 30–60% of patients with TRS can respond to clozapine [31,32,33]). Based on these figures, the criteria for defining TRS are currently recognized as useful for the indication of clozapine and are utilized as such.
The first study to demonstrate the superiority of clozapine over other antipsychotics was reported by Kane, and this report contributed to the approval of clozapine for the treatment of TRS [31]. This study revealed that 30% of haloperidol-resistant patients were improved by clozapine, but chlorpromazine improved 4% of patients [31]. Subsequently, several meta-analyses also confirmed the superiority of clozapine over other typical antipsychotics [32,33]. Two large-scale prospective effectiveness studies (CATIE and CUtLASS) also revealed the superiority of clozapine for TRS over atypical antipsychotics [148,149]. A recent network meta-analysis contradicted the evidence of the superiority of clozapine compared to other antipsychotics [150]. The discrepancy in the evaluated effectiveness of clozapine between the large-scale prospective effectiveness studies (CATIE and CUtLASS) [148,149] and the meta-analysis [150] was explained by two possibilities, namely sampling bias and the fact that the meta-analysis study did not include the data in CATIE or CUtLASS [151].
However, it is notable that evidence shows that over 30% of patients with TRS also exhibit clozapine resistance [31,152,153]. Recently, the TRIPP group proposed the term ultra-treatment-resistant schizophrenia (UTRS) due to the specific features of clozapine resistance [29,30,39]. To understand the basis of UTRS, a more detailed scientific analysis is needed.

5.3. Primary TRS and Secondary TRS

Clinically, the failure to respond to appropriate treatment with more than two antipsychotics is regarded as a fundamental criterion for TRS [29,30,146,147]. Recently, it has been argued that two factors may be involved in the pathomechanism or pathophysiology of TRS [154]. Specifically, TRS is defined as primary or secondary. In primary TRS, the antipsychotic-resistant pathology already exists at the first episode of schizophrenia, and the pathophysiology of secondary TRS is augmented by persistent adequate antipsychotic medications [155,156,157].
The hyperactivation of dopaminergic transmission, including the supersensitivity and upregulation of dopamine D2 receptors, induced by consecutive exposure to D2 receptor antagonistic antipsychotics is the basis for secondary TRS [154,158]. The overall response to all antipsychotics was reported to be 40–60% [159,160]. The response to antipsychotics in antipsychotic-naïve individuals is estimated to be approximately 75%. However, the response to a second trial of antipsychotics other than clozapine is considerably lower, ranging from 20 to 45% [161,162]. These discrepancies in response rates between the first and second trials of antipsychotic medications suggest the possibility that consecutive exposure to D2 receptor antagonists plays a fundamental role in the development of antipsychotic resistance. Indeed, the maximum response rate to a clozapine trial was 80% when treatment was initiated within the first 2–3 years after resistance was established, whereas, when resistance had been established for a longer period of time, the response rate decreased to approximately 30% [160,161,163].
The affinity of clozapine for the D2 receptor is markedly lower than that of conventional typical/atypical antipsychotics [154,164]. Additionally, the rapid dissociation of clozapine from D2 receptors has been considered a candidate mechanism for the decreased supersensitivity of the D2 receptor induced by clozapine [134,164,165,166]. However, the dissociation rate of clozapine from D2 receptors is not significantly faster than those of other antipsychotics, such as quetiapine, amisulpride, remoxipride, and sulpiride [164,167,168,169]. Rather, several antipsychotics act as pharmacological chaperones to D2 receptors independent of dopaminergic signaling, which contributes to the upregulation of D2 receptors. However, the impact of clozapine on D2 receptor trafficking to the plasma membrane is the lowest among antipsychotics [170]. Furthermore, in a recent study, clozapine did not upregulate numerous monoamine receptors via the suppression of protein phosphatase 2A, which regulates the turnover of monoamine receptors [38]. These clinical advantages of clozapine for secondary TRS over other atypical antipsychotics are likely related to the distinct pharmacological profile of this drug.

6. Novel Treatment Based on Glutamate Hypothesis

Both clinical and preclinical findings indicate the possibility that NMDAR modulation can provide novel treatments for TRS. In particular, the finding that 30% of risk genes for schizophrenia identified by GWASs affect NMDAR-associated signaling has attracted attention, as this suggests that, in a proportion of patients with schizophrenia (approximately 30%), the pathogenesis can be explained by the glutamate hypothesis [26,27,28]. Given that the prevalence of primary TRS is also approximately 30% [31,152,153], antipsychotics targeting the NMDAR are attractive novel candidates for the treatment of schizophrenia. To avoid excitotoxicity induced by the activation of NMDARs, the modulation of the glycine modulatory site in GluN1 (but not the direct activation of the glutamate binding site in GluN2) was adopted as a candidate approach [84,171].

6.1. D-Serine and D-Cycloserine

Both D-serine and glycine are endogenous agonists at the glycine modulatory site in GluN1 [40,172]. The activation of the NMDAR requires the binding of L-glutamate to the glutamate binding site in GluN2, accompanied by the binding of D-serine or glycine to the glycine modulatory site in GluN2. D-cycloserine is a non-endogenous partial agonist at the glycine modulatory site with lower efficacy (approximately 50% intrinsic affinity) than endogenous agonists [40,172].
Most double-blind placebo-controlled studies of D-serine and D-cycloserine have demonstrated their effects on negative symptoms and their modest effects on cognition [173], since adjunctive D-cycloserine was effective in addressing all three symptom clusters [174,175]. However, the results of subsequent trials on D-cycloserine efficacy were not consistent. Indeed, some trials found it to be ineffective [176,177]. The therapeutic window of D-cycloserine is speculated to be extremely narrow [178]. A moderate dose/concentration of D-cycloserine, which enhances cognition in experimental animals [179], can activate the NMDAR, but at high doses, D-cycloserine acts as an antagonist due to its low intrinsic efficacy [40,172].
Despite their efficacy, these agents have several shortcomings that render them unsuitable for long-term administration in patients with schizophrenia, such as adverse reactions (renal tubular necrosis, type 2 diabetes mellitus, and heart-valve disease) and a reduction in their efficacy induced by the desensitization of the NMDAR, resulting in the loss of their efficacy [180,181,182,183,184,185].

6.2. Glycine-Transporter Inhibitors

The glycine transporter, composed of SC6A5 and SLC6A9, is known to be a reversible transporter with Na+/Cl dependence. However, with substantial intracellular glycine accumulation, the glycine transporter conversely releases glycine into the extracellular space [24,186]. Therefore, the inhibition of the glycine transporter increases glycine levels in the synaptic cleft, resulting in the activation of the NMDAR. An endogenous glycine transporter inhibitor, sarcosine, is produced in the glycine synthesis pathway [187,188]. Sarcosine application can enhance NMDAR activity as a glycine source via either reverse uptake or heteroexchange [189]. Several randomized control studies of adjunctive sarcosine to antipsychotics demonstrated its efficacy in the treatment of positive and negative symptoms compared to placebo and D-serine intake groups [176,188,190,191]. However, sarcosine and its derivatives lead to several neurological adverse reactions, such as hypoactivity and ataxia [192].
Randomized control phase 2 studies of iclepertin (adjunct to antipsychotics) demonstrated that iclepertin improved cognitive functioning compared to placebo-treated controls [193,194]. However, the results of a series of phase 3 trials in the “CONNEX” program and long-term open-label studies remained to be published [195]. None of the studied agents have successfully completed clinical trials. However, the recent encouraging results of phase 2 studies of iclepertin have reinvigorated the exploration of NMDAR modulators as effective treatment options for schizophrenia.

7. TRS and Clozapine

7.1. Impact of Clozapine on Glutamatergic Agents

Most trials reported that adjunctive D-serine, D-cycloserine, and glycine-transporter inhibitor with clozapine were negative [196,197,198]. Notably, adjunctive D-cycloserine with clozapine and risperidone worsened and improved negative symptoms, respectively [198,199]. Preclinical studies reported that clozapine activated NMDAR via the enhancement of glutamatergic transmission. Clozapine increased the extracellular levels of L-glutamate, D-serine, and glycine through the enhancement of neuronal exocytosis, astroglial non-exocytosis, and the inhibition of the glycine transporter [34,35,36,37,38]. The other actions of clozapine have been reported, including that clozapine enhances occupancy in the glycine-modulating site [200], and interacts with metabotropic glutamate receptors [38,154,201,202,203,204]. The administration of D-cycloserine, D-serine, and glycine-transporter inhibitors in experimental animals improved the memory disruptions induced by MK801 and phencyclidine [187,205,206,207,208,209]. Notably, the co-administration of a glycine-transporter inhibitor and antipsychotics other than clozapine enhanced glutamatergic transmission but attenuated dopaminergic side effects [207].
Considering these preclinical findings, clozapine enhances cognition via the activation of both glutamate and glycine binding sites by increasing the extracellular levels of L-glutamate and D-serine [133,202]. In the absence of clozapine, D-cycloserine can enhance NMDAR function via the activation of the glycine binding site, but under the increasing extracellular D-serine level induced by clozapine, D-cycloserine conversely may suppress the glycine binding site due to its low intrinsic partial agonistic action (approximately 50%), resulting in inhibiting the stimulatory effects of clozapine on NMDAR function. In conclusion, D-cycloserine can improve negative symptoms (approximately 10–26%) and a part of cognitive dysfunction (working memory and experience-dependent neuroplasticity) in patients with schizophrenia [210]. However, adjunctive D-cycloserine to clozapine not only probably inhibits the clinical action of clozapine but also worsens them.
The accumulating pharmacodynamic findings on clozapine support that clozapine may be an antipsychotic agent and has a mechanism of action that corresponds to both revised dopamine and glutamate hypotheses. However, the serious/lethal adverse reactions to clozapine, such as metabolic, hematologic, and neurological complications; cardiotoxicities; and discontinuation syndromes, are more severe than recognized by general psychiatrists [133,202,211,212,213,214,215,216]. Therefore, in the future, it will be important to establish antipsychotic medication that overcomes the double-edged sword features of clozapine.

7.2. Novel Pharmacological Targets of Clozapine

Clinical findings for both clozapine and NMDAR enhancers (glycine-transporter inhibitor, D-serine, and D-cycloserine) suggest that the vulnerability of glutamatergic transmission probably plays an important role in the pathomechanism of a proportion of patients with TRS. The clinical finding is that the adjunctive administration of NMDAR enhancers in patients treated with clozapine had no effect or worsened the condition indicates that clozapine enhances glutamatergic transmission, but at the same time, it also suggests that it may have action targets other than enhanced glutamatergic transmission. Recently, we reported that a candidate gliotransmitter, L-BAIBA, may contribute to the mechanism underlying several adverse reactions to clozapine that could not be explained by the receptor binding profile of clozapine alone [38,154,202,203].
L-BAIBA is an isomer of GABA and is structurally similar to GABA [202]. Indeed, L-BAIBA binds to both GABA-A and GABA-B receptors [154,217]. Furthermore, L-BAIBA also binds to Mas-related G-protein-coupled receptor type D (MRGPRD), the glycine receptor, and III-GluR [154,218]. L-BAIBA increases AMP-dependent kinase (AMPK) signaling by activating MRGPRD, leading to metabolic complications [203]. The activation of AMPK signaling induced by L-BAIBA suppresses protein phosphatase 2A (PP2A) activity, resulting in the inhibition of the upregulation of dopamine D2, serotonin 5-HT1A, 5-HT2A, and 5-HT7 receptors (downregulation in animal models) [38]. The suppressive effects of clozapine on PP2A through an increase in L-BAIBA have been proposed to play an important role in serotonergic clozapine-discontinuation syndrome/withdrawal syndrome [38]. The discovery that clozapine inhibits PP2A activity by increasing L-BAIBA signaling contributed to the revision of the traditional pathophysiological hypothesis of serotonergic clozapine-discontinuation syndrome [38]. Similarly, a pathophysiological hypothesis of clozapine-discontinuation catatonia can be constructed via the effects of L-BAIBA on PP2A and GABA-B receptors [38,154,202]. These pathophysiological hypotheses of L-BAIBA can reasonably explain the time lag in the onset of serotonergic symptoms and catatonia after clozapine discontinuation (they usually appear 1–2 weeks after clozapine discontinuation) [38]. Thus, the mechanisms underlying the clinical efficacy and adverse reactions of clozapine, which cannot be fully explained by its receptor binding profile, may be mediated by novel molecules whose functions are induced by clozapine [38,133,154,202]. Considering the clinical efficacy and severe/lethal adverse reactions of clozapine, exploring and discovering novel targets of clozapine may contribute to the development of novel treatments for schizophrenia with fewer adverse reactions and a clinical efficacy comparable to that of clozapine. Although the pathomechanisms of the recently proposed UTRS (clozapine-resistant schizophrenia) [29,30,39] are speculated to involve functional abnormalities other than glutamatergic transmission, it should be noted that, unless we identify the actual mechanisms and unknown targets regarding the action of clozapine, the interpretation of hypothetical pathophysiology and/or pathomechanisms of UTRS may be off the mark.

8. Conclusions

This review discusses the attractive viewpoints of the glutamate hypothesis as a background for the development of novel antipsychotics independent of the dopamine hypothesis; part of the discussion includes the possibility that patients with glutamatergic dysfunction may be one of the major subgroups of TRS. However, considering the results of GWASs, patients with schizophrenia with glutamatergic dysfunction or acquired risks related to glutamatergic transmission are speculated to account for approximately 30% of all individuals with schizophrenia. Therefore, the population of patients with schizophrenia is probably composed of individuals resistant to not only dopaminergic agents but also glutamatergic agents or both (resistant to dopaminergic and glutamatergic agents). In other words, when evaluating the efficacy of novel glutamatergic agents through randomized controlled trials in the entire population of patients with schizophrenia, the results may be skewed due to individuals who are sensitive to dopaminergic antipsychotics but insensitive to glutamatergic agents. Furthermore, because of severe adverse reactions to clozapine, which is currently the sole approved antipsychotic for the treatment of TRS, it is important to strive to develop novel antipsychotics that are more effective in the treatment of TRS than conventional antipsychotics, are safer than clozapine, and can prevent the development of secondary TRS. Considering the pharmacological features of clozapine (increasing extracellular levels of endogenous NMDAR agonists), second-generation NMDAR PAMs may be candidate agents that are clinically efficacious in addition to clozapine. Finally, the development of treatment for UTRS will be the most important challenge in psychiatry in the future. Thus, pharmacodynamic analyses to clarify the unknown pathophysiology of clozapine are becoming increasingly important.

Author Contributions

Conceptualization, M.O.; data curation, R.O., E.M. and M.O.; funding acquisition, M.O.; methodology, R.O. and M.O.; project administration; M.O.; validation, R.O., E.M. and M.O.; writing—original draft, R.O. and M.O.; writing—review and editing, E.M. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Japan Society for the Promotion of Science (23K06986 and 24K10730).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. GBD 2019 Mental Disorders Collaborators. Global, regional, and national burden of 12 mental disorders in 204 countries and territories, 1990–2019: A systematic analysis for the global burden of disease study 2019. Lancet Psychiatry 2022, 9, 137–150. [Google Scholar]
  2. Diseases, G.B.D.; Injuries, C. Global burden of 369 diseases and injuries in 204 countries and territories, 1990–2019: A systematic analysis for the global burden of disease study 2019. Lancet 2020, 396, 1204–1222. [Google Scholar]
  3. Meltzer, H.Y.; Li, Z.; Kaneda, Y.; Ichikawa, J. Serotonin receptors: Their key role in drugs to treat schizophrenia. Prog. Neuropsychopharmacol. Biol. Psychiatry 2003, 27, 1159–1172. [Google Scholar] [CrossRef]
  4. Tarazi, F.I.; Neill, J.C. The preclinical profile of asenapine: Clinical relevance for the treatment of schizophrenia and bipolar mania. Expert. Opin. Drug Discov. 2013, 8, 93–103. [Google Scholar] [CrossRef]
  5. Goff, D.C. Future perspectives on the treatment of cognitive deficits and negative symptoms in schizophrenia. World Psychiatry 2013, 12, 99–107. [Google Scholar] [CrossRef] [PubMed]
  6. Greenberg, W.M.; Citrome, L. Pharmacokinetics and pharmacodynamics of lurasidone hydrochloride, a second-generation antipsychotic: A systematic review of the published literature. Clin. Pharmacokinet. 2017, 56, 493–503. [Google Scholar] [CrossRef]
  7. Gessa, G.L.; Devoto, P.; Diana, M.; Flore, G.; Melis, M.; Pistis, M. Dissociation of haloperidol, clozapine, and olanzapine effects on electrical activity of mesocortical dopamine neurons and dopamine release in the prefrontal cortex. Neuropsychopharmacology 2000, 22, 642–649. [Google Scholar] [CrossRef]
  8. Rusheen, A.E.; Gee, T.A.; Jang, D.P.; Blaha, C.D.; Bennet, K.E.; Lee, K.H.; Heien, M.L.; Oh, Y. Evaluation of electrochemical methods for tonic dopamine detection in vivo. Trends Anal. Chem. 2020, 132, 116049. [Google Scholar] [CrossRef]
  9. Pow, J.L.; Baumeister, A.A.; Hawkins, M.F.; Cohen, A.S.; Garand, J.C. Deinstitutionalization of american public hospitals for the mentally ill before and after the introduction of antipsychotic medications. Harv. Rev. Psychiatry 2015, 23, 176–187. [Google Scholar] [CrossRef]
  10. Liddle, P.F. The symptoms of chronic schizophrenia. A re-examination of the positive-negative dichotomy. Br. J. Psychiatry 1987, 151, 145–151. [Google Scholar] [CrossRef]
  11. Crow, T.J. Positive and negative schizophrenic symptoms and the role of dopamine. Br. J. Psychiatry 1980, 137, 383–386. [Google Scholar] [CrossRef] [PubMed]
  12. Johnstone, E.C.; Owens, D.G.; Frith, C.D.; Crow, T.J. The relative stability of positive and negative features in chronic schizophrenia. Br. J. Psychiatry 1987, 150, 60–64. [Google Scholar] [CrossRef]
  13. Huhn, M.; Nikolakopoulou, A.; Schneider-Thoma, J.; Krause, M.; Samara, M.; Peter, N.; Arndt, T.; Backers, L.; Rothe, P.; Cipriani, A.; et al. Comparative efficacy and tolerability of 32 oral antipsychotics for the acute treatment of adults with multi-episode schizophrenia: A systematic review and network meta-analysis. Lancet 2019, 394, 939–951. [Google Scholar] [CrossRef] [PubMed]
  14. Kaar, S.J.; Natesan, S.; McCutcheon, R.; Howes, O.D. Antipsychotics: Mechanisms underlying clinical response and side-effects and novel treatment approaches based on pathophysiology. Neuropharmacology 2020, 172, 107704. [Google Scholar] [CrossRef] [PubMed]
  15. Kane, J.M. Newer antipsychotic drugs. A review of their pharmacology and therapeutic potential. Drugs 1993, 46, 585–593. [Google Scholar] [CrossRef] [PubMed]
  16. Worrel, J.A.; Marken, P.A.; Beckman, S.E.; Ruehter, V.L. Atypical antipsychotic agents: A critical review. Am. J. Health Syst. Pharm. 2000, 57, 238–255. [Google Scholar] [CrossRef]
  17. Lim, J.; Lee, S.A.; Lam, M.; Rapisarda, A.; Kraus, M.; Keefe, R.S.; Lee, J. The relationship between negative symptom subdomains and cognition. Psychol. Med. 2016, 46, 2169–2177. [Google Scholar] [CrossRef]
  18. Brannan, S.K.; Sawchak, S.; Miller, A.C.; Lieberman, J.A.; Paul, S.M.; Breier, A. Muscarinic cholinergic receptor agonist and peripheral antagonist for schizophrenia. N. Engl. J. Med. 2021, 384, 717–726. [Google Scholar] [CrossRef]
  19. Correll, C.U.; Angelov, A.S.; Miller, A.C.; Weiden, P.J.; Brannan, S.K. Safety and tolerability of karxt (xanomeline-trospium) in a phase 2, randomized, double-blind, placebo-controlled study in patients with schizophrenia. Schizophrenia 2022, 8, 109. [Google Scholar] [CrossRef]
  20. Sauder, C.; Allen, L.A.; Baker, E.; Miller, A.C.; Paul, S.M.; Brannan, S.K. Effectiveness of karxt (xanomeline-trospium) for cognitive impairment in schizophrenia: Post hoc analyses from a randomised, double-blind, placebo-controlled phase 2 study. Transl. Psychiatry 2022, 12, 491. [Google Scholar] [CrossRef]
  21. Weiden, P.J.; Breier, A.; Kavanagh, S.; Miller, A.C.; Brannan, S.K.; Paul, S.M. Antipsychotic efficacy of karxt (xanomeline-trospium): Post hoc analysis of positive and negative syndrome scale categorical response rates, time course of response, and symptom domains of response in a phase 2 study. J. Clin. Psychiatry 2022, 83, 40913. [Google Scholar] [CrossRef] [PubMed]
  22. Koblan, K.S.; Kent, J.; Hopkins, S.C.; Krystal, J.H.; Cheng, H.; Goldman, R.; Loebel, A. A non-d2-receptor-binding drug for the treatment of schizophrenia. N. Engl. J. Med. 2020, 382, 1497–1506. [Google Scholar] [CrossRef] [PubMed]
  23. Correll, C.U.; Koblan, K.S.; Hopkins, S.C.; Li, Y.; Heather, D.; Goldman, R.; Loebel, A. Safety and effectiveness of ulotaront (sep-363856) in schizophrenia: Results of a 6-month, open-label extension study. NPJ Schizophr. 2021, 7, 63. [Google Scholar] [CrossRef] [PubMed]
  24. Rosenbrock, H.; Desch, M.; Wunderlich, G. Development of the novel glyt1 inhibitor, iclepertin (bi 425809), for the treatment of cognitive impairment associated with schizophrenia. Eur. Arch. Psychiatry Clin. Neurosci. 2023, 273, 1557–1566. [Google Scholar] [CrossRef]
  25. Tsapakis, E.M.; Diakaki, K.; Miliaras, A.; Fountoulakis, K.N. Novel compounds in the treatment of schizophrenia-a selective review. Brain Sci. 2023, 13, 1193. [Google Scholar] [CrossRef]
  26. Trubetskoy, V.; Pardinas, A.F.; Qi, T.; Panagiotaropoulou, G.; Awasthi, S.; Bigdeli, T.B.; Bryois, J.; Chen, C.Y.; Dennison, C.A.; Hall, L.S.; et al. Mapping genomic loci implicates genes and synaptic biology in schizophrenia. Nature 2022, 604, 502–508. [Google Scholar] [CrossRef]
  27. Owen, M.J.; Legge, S.E.; Rees, E.; Walters, J.T.R.; O’Donovan, M.C. Genomic findings in schizophrenia and their implications. Mol. Psychiatry 2023, 28, 3638–3647. [Google Scholar] [CrossRef]
  28. Schizophrenia Working Group of the Psychiatric Genomics Consortium. Biological insights from 108 schizophrenia-associated genetic loci. Nature 2014, 511, 421–427. [Google Scholar] [CrossRef]
  29. Howes, O.D.; McCutcheon, R.; Agid, O.; de Bartolomeis, A.; van Beveren, N.J.; Birnbaum, M.L.; Bloomfield, M.A.; Bressan, R.A.; Buchanan, R.W.; Carpenter, W.T.; et al. Treatment-resistant schizophrenia: Treatment response and resistance in psychosis (trrip) working group consensus guidelines on diagnosis and terminology. Am. J. Psychiatry 2017, 174, 216–229. [Google Scholar] [CrossRef]
  30. Lehman, A.F.; Lieberman, J.A.; Dixon, L.B.; McGlashan, T.H.; Miller, A.L.; Perkins, D.O.; Kreyenbuhl, J.; American Psychiatric, A.; Steering Committee on Practice, G. Practice guideline for the treatment of patients with schizophrenia, second edition. Am. J. Psychiatry 2004, 161, 1–56. [Google Scholar]
  31. Kane, J.; Honigfeld, G.; Singer, J.; Meltzer, H. Clozapine for the treatment-resistant schizophrenic. A double-blind comparison with chlorpromazine. Arch. Gen. Psychiatry 1988, 45, 789–796. [Google Scholar] [CrossRef] [PubMed]
  32. Chakos, M.; Lieberman, J.; Hoffman, E.; Bradford, D.; Sheitman, B. Effectiveness of second-generation antipsychotics in patients with treatment-resistant schizophrenia: A review and meta-analysis of randomized trials. Am. J. Psychiatry 2001, 158, 518–526. [Google Scholar] [CrossRef]
  33. Siskind, D.; McCartney, L.; Goldschlager, R.; Kisely, S. Clozapine v. First- and second-generation antipsychotics in treatment-refractory schizophrenia: Systematic review and meta-analysis. Br. J. Psychiatry 2016, 209, 385–392. [Google Scholar] [CrossRef] [PubMed]
  34. Javitt, D.C.; Duncan, L.; Balla, A.; Sershen, H. Inhibition of system a-mediated glycine transport in cortical synaptosomes by therapeutic concentrations of clozapine: Implications for mechanisms of action. Mol. Psychiatry 2005, 10, 275–287. [Google Scholar] [CrossRef] [PubMed]
  35. Tanahashi, S.; Yamamura, S.; Nakagawa, M.; Motomura, E.; Okada, M. Clozapine, but not haloperidol, enhances glial d-serine and l-glutamate release in rat frontal cortex and primary cultured astrocytes. Br. J. Pharmacol. 2012, 165, 1543–1555. [Google Scholar] [CrossRef]
  36. Fukuyama, K.; Okubo, R.; Murata, M.; Shiroyama, T.; Okada, M. Activation of astroglial connexin is involved in concentration-dependent double-edged sword clinical action of clozapine. Cells 2020, 9, 414. [Google Scholar] [CrossRef]
  37. Fukuyama, K.; Okada, M. Effects of atypical antipsychotics, clozapine, quetiapine and brexpiprazole on astroglial transmission associated with connexin43. Int. J. Mol. Sci. 2021, 22, 5623. [Google Scholar] [CrossRef]
  38. Okada, M.; Fukuyama, K.; Motomura, E. Impacts of exposure to and subsequent discontinuation of clozapine on tripartite synaptic transmission. Br. J. Pharmacol. 2024, 181, 1–22. [Google Scholar] [CrossRef]
  39. Luykx, J.J.; Gonzalez-Diaz, J.M.; Guu, T.W.; van der Horst, M.Z.; van Dellen, E.; Boks, M.P.; Guloksuz, S.; DeLisi, L.E.; Sommer, I.E.; Cummins, R.; et al. An international research agenda for clozapine-resistant schizophrenia. Lancet Psychiatry 2023, 10, 644–652. [Google Scholar] [CrossRef]
  40. Blanke, M.L.; VanDongen, A.M.J. Activation mechanisms of the NMDA receptor. In Biology of the NMDA Receptorf; Van Dongen, A.M., Ed.; CRC Press: Boca Raton, FL, USA, 2009. [Google Scholar]
  41. Vyklicky, V.; Korinek, M.; Smejkalova, T.; Balik, A.; Krausova, B.; Kaniakova, M.; Lichnerova, K.; Cerny, J.; Krusek, J.; Dittert, I.; et al. Structure, function, and pharmacology of NMDA receptor channels. Physiol. Res. 2014, 63, S191–S203. [Google Scholar] [CrossRef]
  42. 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]
  43. Yao, Y.; Belcher, J.; Berger, A.J.; Mayer, M.L.; Lau, A.Y. Conformational analysis of NMDA receptor glun1, glun2, and glun3 ligand-binding domains reveals subtype-specific characteristics. Structure 2013, 21, 1788–1799. [Google Scholar] [CrossRef] [PubMed]
  44. Lu, W.; Du, J.; Goehring, A.; Gouaux, E. Cryo-em structures of the triheteromeric NMDA receptor and its allosteric modulation. Science 2017, 355, eaal3729. [Google Scholar] [CrossRef]
  45. Sullivan, J.M.; Traynelis, S.F.; Chen, H.S.; Escobar, W.; Heinemann, S.F.; Lipton, S.A. Identification of two cysteine residues that are required for redox modulation of the NMDA subtype of glutamate receptor. Neuron 1994, 13, 929–936. [Google Scholar] [CrossRef]
  46. Jiang, L.; Liu, N.; Zhao, F.; Huang, B.; Kang, D.; Zhan, P.; Liu, X. Discovery of glun2a subtype-selective n-methyl-d-aspartate (NMDA) receptor ligands. Acta Pharm. Sin. B 2024, 14, 1987–2005. [Google Scholar] [CrossRef] [PubMed]
  47. Chou, T.H.; Epstein, M.; Fritzemeier, R.G.; Akins, N.S.; Paladugu, S.; Ullman, E.Z.; Liotta, D.C.; Traynelis, S.F.; Furukawa, H. Molecular mechanism of ligand gating and opening of NMDA receptor. Nature 2024, 632, 209–217. [Google Scholar] [CrossRef]
  48. Niu, M.; Yang, X.; Li, Y.; Sun, Y.; Wang, L.; Ha, J.; Xie, Y.; Gao, Z.; Tian, C.; Wang, L.; et al. Progresses in glun2a-containing NMDA receptors and their selective regulators. Cell. Mol. Neurobiol. 2023, 43, 139–153. [Google Scholar] [CrossRef]
  49. Li, D.C.; Pitts, E.G.; Dighe, N.M.; Gourley, S.L. Glun2b inhibition confers resilience against long-term cocaine-induced neurocognitive sequelae. Neuropsychopharmacology 2023, 48, 1108–1117. [Google Scholar] [CrossRef]
  50. Tarres-Gatius, M.; Miquel-Rio, L.; Campa, L.; Artigas, F.; Castane, A. Involvement of NMDA receptors containing the glun2c subunit in the psychotomimetic and antidepressant-like effects of ketamine. Transl. Psychiatry 2020, 10, 427. [Google Scholar] [CrossRef]
  51. Wang, J.X.; Irvine, M.W.; Burnell, E.S.; Sapkota, K.; Thatcher, R.J.; Li, M.; Simorowski, N.; Volianskis, A.; Collingridge, G.L.; Monaghan, D.T.; et al. Structural basis of subtype-selective competitive antagonism for glun2c/2d-containing NMDA receptors. Nat. Commun. 2020, 11, 423. [Google Scholar] [CrossRef]
  52. Lai, Y.; Chu, X.; Di, L.; Gao, W.; Guo, Y.; Liu, X.; Lu, C.; Mao, J.; Shen, H.; Tang, H.; et al. Recent advances in the translation of drug metabolism and pharmacokinetics science for drug discovery and development. Acta Pharm. Sin. B 2022, 12, 2751–2777. [Google Scholar] [CrossRef]
  53. Luisada, P.V. The phencyclidine psychosis: Phenomenology and treatment. NIDA Res. Monogr. 1978, 12, 241–253. [Google Scholar]
  54. Snyder, S.H. Phencyclidine. Nature 1980, 285, 355–356. [Google Scholar] [CrossRef] [PubMed]
  55. Lahti, A.C.; Holcomb, H.H.; Medoff, D.R.; Tamminga, C.A. Ketamine activates psychosis and alters limbic blood flow in schizophrenia. Neuroreport 1995, 6, 869–872. [Google Scholar] [CrossRef]
  56. 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]
  57. Zandi, M.S.; Irani, S.R.; Lang, B.; Waters, P.; Jones, P.B.; McKenna, P.; Coles, A.J.; Vincent, A.; Lennox, B.R. Disease-relevant autoantibodies in first episode schizophrenia. J. Neurol. 2011, 258, 686–688. [Google Scholar] [CrossRef] [PubMed]
  58. Pollak, T.A.; McCormack, R.; Peakman, M.; Nicholson, T.R.; David, A.S. Prevalence of anti-n-methyl-d-aspartate (NMDA) receptor [corrected] antibodies in patients with schizophrenia and related psychoses: A systematic review and meta-analysis. Psychol. Med. 2014, 44, 2475–2487. [Google Scholar] [CrossRef]
  59. Jezequel, J.; Johansson, E.M.; Leboyer, M.; Groc, L. Pathogenicity of antibodies against NMDA receptor: Molecular insights into autoimmune psychosis. Trends Neurosci. 2018, 41, 502–511. [Google Scholar] [CrossRef]
  60. Tong, J.; Huang, J.; Luo, X.; Chen, S.; Cui, Y.; An, H.; Xiu, M.; Tan, S.; Wang, Z.; Yuan, Y.; et al. Elevated serum anti-NMDA receptor antibody levels in first-episode patients with schizophrenia. Brain Behav. Immun. 2019, 81, 213–219. [Google Scholar] [CrossRef]
  61. Nakazawa, K.; Sapkota, K. The origin of NMDA receptor hypofunction in schizophrenia. Pharmacol. Ther. 2020, 205, 107426. [Google Scholar] [CrossRef]
  62. Moscato, E.H.; Peng, X.; Jain, A.; Parsons, T.D.; Dalmau, J.; Balice-Gordon, R.J. Acute mechanisms underlying antibody effects in anti-n-methyl-d-aspartate receptor encephalitis. Ann. Neurol. 2014, 76, 108–119. [Google Scholar] [CrossRef] [PubMed]
  63. Dalmau, J.; Tuzun, E.; Wu, H.Y.; Masjuan, J.; Rossi, J.E.; Voloschin, A.; Baehring, J.M.; Shimazaki, H.; Koide, R.; King, D.; et al. Paraneoplastic anti-n-methyl-d-aspartate receptor encephalitis associated with ovarian teratoma. Ann. Neurol. 2007, 61, 25–36. [Google Scholar] [CrossRef] [PubMed]
  64. Masdeu, J.C.; Dalmau, J.; Berman, K.F. NMDA receptor internalization by autoantibodies: A reversible mechanism underlying psychosis? Trends Neurosci. 2016, 39, 300–310. [Google Scholar] [CrossRef]
  65. Gibson, L.L.; McKeever, A.; Coutinho, E.; Finke, C.; Pollak, T.A. Cognitive impact of neuronal antibodies: Encephalitis and beyond. Transl. Psychiatry 2020, 10, 304. [Google Scholar] [CrossRef]
  66. Okada, M.; Kawano, Y.; Fukuyama, K.; Motomura, E.; Shiroyama, T. Candidate strategies for development of a rapid-acting antidepressant class that does not result in neuropsychiatric adverse effects: Prevention of ketamine-induced neuropsychiatric adverse reactions. Int. J. Mol. Sci. 2020, 21, 7951. [Google Scholar] [CrossRef] [PubMed]
  67. Ruffell, S.G.D.; Crosland-Wood, M.; Palmer, R.; Netzband, N.; Tsang, W.; Weiss, B.; Gandy, S.; Cowley-Court, T.; Halman, A.; McHerron, D.; et al. Ayahuasca: A review of historical, pharmacological, and therapeutic aspects. PCN Rep. 2023, 2, e1462023. [Google Scholar] [CrossRef]
  68. McKenna, D.J.; Towers, G.H.; Abbott, F. Monoamine oxidase inhibitors in south american hallucinogenic plants: Tryptamine and beta-carboline constituents of ayahuasca. J. Ethnopharmacol. 1984, 10, 195–223. [Google Scholar] [CrossRef]
  69. Dos Santos, R.G.; Balthazar, F.M.; Bouso, J.C.; Hallak, J.E. The current state of research on ayahuasca: A systematic review of human studies assessing psychiatric symptoms, neuropsychological functioning, and neuroimaging. J. Psychopharmacol. 2016, 30, 1230–1247. [Google Scholar] [CrossRef]
  70. Li, Y.; Sattler, R.; Yang, E.J.; Nunes, A.; Ayukawa, Y.; Akhtar, S.; Ji, G.; Zhang, P.W.; Rothstein, J.D. Harmine, a natural beta-carboline alkaloid, upregulates astroglial glutamate transporter expression. Neuropharmacology 2011, 60, 1168–1175. [Google Scholar] [CrossRef]
  71. Liu, F.; Wu, J.; Gong, Y.; Wang, P.; Zhu, L.; Tong, L.; Chen, X.; Ling, Y.; Huang, C. Harmine produces antidepressant-like effects via restoration of astrocytic functions. Prog. Neuropsychopharmacol. Biol. Psychiatry 2017, 79, 258–267. [Google Scholar] [CrossRef]
  72. Gao, W.; Yang, N.; Mei, X.; Zhu, X.; Hu, W.; Zeng, Y. Influence of anti-tuberculosis drugs plus cycloserine on sputum negative conversion rate, adverse reactions and inflammatory factors in multi-drug resistant tuberculosis. Am. J. Transl. Res. 2021, 13, 9332–9339. [Google Scholar] [PubMed]
  73. Kristiansen, L.V.; Huerta, I.; Beneyto, M.; Meador-Woodruff, J.H. NMDA receptors and schizophrenia. Curr. Opin. Pharmacol. 2007, 7, 48–55. [Google Scholar] [CrossRef] [PubMed]
  74. 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]
  75. Gao, X.M.; Sakai, K.; Roberts, R.C.; Conley, R.R.; Dean, B.; Tamminga, C.A. Ionotropic glutamate receptors and expression of n-methyl-d-aspartate receptor subunits in subregions of human hippocampus: Effects of schizophrenia. Am. J. Psychiatry 2000, 157, 1141–1149. [Google Scholar] [CrossRef]
  76. Kristiansen, L.V.; Beneyto, M.; Haroutunian, V.; Meador-Woodruff, J.H. 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]
  77. Dracheva, S.; Marras, S.A.; Elhakem, S.L.; Kramer, F.R.; Davis, K.L.; Haroutunian, V. N-methyl-d-aspartic acid receptor expression in the dorsolateral prefrontal cortex of elderly patients with schizophrenia. Am. J. Psychiatry 2001, 158, 1400–1410. [Google Scholar] [CrossRef]
  78. Ohnuma, T.; Kato, H.; Arai, H.; Faull, R.L.; McKenna, P.J.; Emson, P.C. Gene expression of psd95 in prefrontal cortex and hippocampus in schizophrenia. Neuroreport 2000, 11, 3133–3137. [Google Scholar] [CrossRef]
  79. Tsai, G.; Passani, L.A.; Slusher, B.S.; Carter, R.; Baer, L.; Kleinman, J.E.; Coyle, J.T. Abnormal excitatory neurotransmitter metabolism in schizophrenic brains. Arch. Gen. Psychiatry 1995, 52, 829–836. [Google Scholar] [CrossRef]
  80. Ghose, S.; Chin, R.; Gallegos, A.; Roberts, R.; Coyle, J.; Tamminga, C. Localization of naag-related gene expression deficits to the anterior hippocampus in schizophrenia. Schizophr. Res. 2009, 111, 131–137. [Google Scholar] [CrossRef]
  81. Bergeron, R.; Coyle, J.T. Naag, NMDA receptor and psychosis. Curr. Med. Chem. 2012, 19, 1360–1364. [Google Scholar] [CrossRef]
  82. Wonodi, I.; Schwarcz, R. Cortical kynurenine pathway metabolism: A novel target for cognitive enhancement in schizophrenia. Schizophr. Bull. 2010, 36, 211–218. [Google Scholar] [CrossRef] [PubMed]
  83. Coyle, J.T.; Balu, D.T. The role of serine racemase in the pathophysiology of brain disorders. Adv. Pharmacol. 2018, 82, 35–56. [Google Scholar] [PubMed]
  84. Balu, D.T. The NMDA receptor and schizophrenia: From pathophysiology to treatment. Adv. Pharmacol. 2016, 76, 351–382. [Google Scholar] [PubMed]
  85. Fukuyama, K.; Tanahashi, S.; Hoshikawa, M.; Shinagawa, R.; Okada, M. Zonisamide regulates basal ganglia transmission via astroglial kynurenine pathway. Neuropharmacology 2014, 76, 137–145. [Google Scholar] [CrossRef] [PubMed]
  86. Yamamura, S.; Hoshikawa, M.; Dai, K.; Saito, H.; Suzuki, N.; Niwa, O.; Okada, M. Ono-2506 inhibits spike-wave discharges in a genetic animal model without affecting traditional convulsive tests via gliotransmission regulation. Br. J. Pharmacol. 2013, 168, 1088–1100. [Google Scholar] [CrossRef]
  87. Fukuyama, K.; Okada, M. Effects of levetiracetam on astroglial release of kynurenine-pathway metabolites. Br. J. Pharmacol. 2018, 175, 4253–4265. [Google Scholar] [CrossRef]
  88. Coyle, J.T.; Ruzicka, W.B.; Balu, D.T. Fifty years of research on schizophrenia: The ascendance of the glutamatergic synapse. Am. J. Psychiatry 2020, 177, 1119–1128. [Google Scholar] [CrossRef]
  89. Kirov, G.; Pocklington, A.J.; Holmans, P.; Ivanov, D.; Ikeda, M.; Ruderfer, D.; Moran, J.; Chambert, K.; Toncheva, D.; Georgieva, L.; et al. De novo cnv analysis implicates specific abnormalities of postsynaptic signalling complexes in the pathogenesis of schizophrenia. Mol. Psychiatry 2012, 17, 142–153. [Google Scholar] [CrossRef]
  90. Singh, T.; Poterba, T.; Curtis, D.; Akil, H.; Al Eissa, M.; Barchas, J.D.; Bass, N.; Bigdeli, T.B.; Breen, G.; Bromet, E.J.; et al. Rare coding variants in ten genes confer substantial risk for schizophrenia. Nature 2022, 604, 509–516. [Google Scholar] [CrossRef]
  91. Hu, T.M.; Wang, Y.C.; Wu, C.L.; Hsu, S.H.; Tsai, H.Y.; Cheng, M.C. Multiple rare risk coding variants in postsynaptic density-related genes associated with schizophrenia susceptibility. Front. Genet. 2020, 11, 524258. [Google Scholar] [CrossRef]
  92. Holcomb, H.H.; Lahti, A.C.; Medoff, D.R.; Cullen, T.; Tamminga, C.A. Effects of noncompetitive NMDA receptor blockade on anterior cingulate cerebral blood flow in volunteers with schizophrenia. Neuropsychopharmacology 2005, 30, 2275–2282. [Google Scholar] [CrossRef] [PubMed]
  93. Merritt, K.; Egerton, A.; Kempton, M.J.; Taylor, M.J.; McGuire, P.K. Nature of glutamate alterations in schizophrenia: A meta-analysis of proton magnetic resonance spectroscopy studies. JAMA Psychiatry 2016, 73, 665–674. [Google Scholar] [CrossRef] [PubMed]
  94. Demjaha, A.; Egerton, A.; Murray, R.M.; Kapur, S.; Howes, O.D.; Stone, J.M.; McGuire, P.K. Antipsychotic treatment resistance in schizophrenia associated with elevated glutamate levels but normal dopamine function. Biol. Psychiatry 2014, 75, e11–e13. [Google Scholar] [CrossRef] [PubMed]
  95. Theberge, J.; Bartha, R.; Drost, D.J.; Menon, R.S.; Malla, A.; Takhar, J.; Neufeld, R.W.; Rogers, J.; Pavlosky, W.; Schaefer, B.; et al. Glutamate and glutamine measured with 4.0 t proton mrs in never-treated patients with schizophrenia and healthy volunteers. Am. J. Psychiatry 2002, 159, 1944–1946. [Google Scholar] [CrossRef]
  96. de la Fuente-Sandoval, C.; Leon-Ortiz, P.; Azcarraga, M.; Stephano, S.; Favila, R.; Diaz-Galvis, L.; Alvarado-Alanis, P.; Ramirez-Bermudez, 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]
  97. Rowland, L.M.; Bustillo, J.R.; Mullins, P.G.; Jung, R.E.; Lenroot, R.; Landgraf, E.; Barrow, R.; Yeo, R.; Lauriello, J.; Brooks, W.M. Effects of ketamine on anterior cingulate glutamate metabolism in healthy humans: A 4-t proton mrs study. Am. J. Psychiatry 2005, 162, 394–396. [Google Scholar] [CrossRef]
  98. Stone, J.M.; Dietrich, C.; Edden, R.; Mehta, M.A.; De Simoni, S.; Reed, L.J.; Krystal, J.H.; Nutt, D.; Barker, G.J. Ketamine effects on brain gaba and glutamate levels with 1h-mrs: Relationship to ketamine-induced psychopathology. Mol. Psychiatry 2012, 17, 664–665. [Google Scholar] [CrossRef]
  99. Nagai, T.; Kirihara, K.; Tada, M.; Koshiyama, D.; Koike, S.; Suga, M.; Araki, T.; Hashimoto, K.; Kasai, K. Reduced mismatch negativity is associated with increased plasma level of glutamate in first-episode psychosis. Sci. Rep. 2017, 7, 2258. [Google Scholar] [CrossRef]
  100. Picton, T.W. The p300 wave of the human event-related potential. J. Clin. Neurophysiol. 1992, 9, 456–479. [Google Scholar] [CrossRef]
  101. Ross, J.M.; Hamm, J.P. Cortical microcircuit mechanisms of mismatch negativity and its underlying subcomponents. Front. Neural Circuits 2020, 14, 13. [Google Scholar] [CrossRef]
  102. Haaf, M.; Leicht, G.; Curic, S.; Mulert, C. Glutamatergic deficits in schizophrenia—Biomarkers and pharmacological interventions within the ketamine model. Curr. Pharm. Biotechnol. 2018, 19, 293–307. [Google Scholar] [CrossRef]
  103. Umbricht, D.; Schmid, L.; Koller, R.; Vollenweider, F.X.; Hell, D.; Javitt, D.C. Ketamine-induced deficits in auditory and visual context-dependent processing in healthy volunteers: Implications for models of cognitive deficits in schizophrenia. Arch. Gen. Psychiatry 2000, 57, 1139–1147. [Google Scholar] [CrossRef] [PubMed]
  104. Oranje, B.; van Berckel, B.N.; Kemner, C.; van Ree, J.M.; Kahn, R.S.; Verbaten, M.N. The effects of a sub-anaesthetic dose of ketamine on human selective attention. Neuropsychopharmacology 2000, 22, 293–302. [Google Scholar] [CrossRef]
  105. Nagai, T.; Tada, M.; Kirihara, K.; Yahata, N.; Hashimoto, R.; Araki, T.; Kasai, K. Auditory mismatch negativity and p3a in response to duration and frequency changes in the early stages of psychosis. Schizophr. Res. 2013, 150, 547–554. [Google Scholar] [CrossRef] [PubMed]
  106. Naatanen, R.; Todd, J.; Schall, U. Mismatch negativity (mmn) as biomarker predicting psychosis in clinically at-risk individuals. Biol. Psychol. 2016, 116, 36–40. [Google Scholar] [CrossRef]
  107. Billard, J.M. Changes in serine racemase-dependent modulation of NMDA receptor: Impact on physiological and pathological brain aging. Front. Mol. Biosci. 2018, 5, 106. [Google Scholar] [CrossRef] [PubMed]
  108. Wolosker, H.; Balu, D.T.; Coyle, J.T. The rise and fall of the d-serine-mediated gliotransmission hypothesis. Trends Neurosci. 2016, 39, 712–721. [Google Scholar] [CrossRef]
  109. Balu, D.T.; Li, Y.; Puhl, M.D.; Benneyworth, M.A.; Basu, A.C.; Takagi, S.; Bolshakov, V.Y.; Coyle, J.T. Multiple risk pathways for schizophrenia converge in serine racemase knockout mice, a mouse model of NMDA receptor hypofunction. Proc. Natl. Acad. Sci. USA 2013, 110, E2400–E2409. [Google Scholar] [CrossRef]
  110. Basu, A.C.; Tsai, G.E.; Ma, C.L.; Ehmsen, J.T.; Mustafa, A.K.; Han, L.; Jiang, Z.I.; Benneyworth, M.A.; Froimowitz, M.P.; Lange, N.; et al. Targeted disruption of serine racemase affects glutamatergic neurotransmission and behavior. Mol. Psychiatry 2009, 14, 719–727. [Google Scholar] [CrossRef]
  111. DeVito, L.M.; Balu, D.T.; Kanter, B.R.; Lykken, C.; Basu, A.C.; Coyle, J.T.; Eichenbaum, H. Serine racemase deletion disrupts memory for order and alters cortical dendritic morphology. Genes. Brain Behav. 2011, 10, 210–222. [Google Scholar] [CrossRef]
  112. Balu, D.T.; Presti, K.T.; Huang, C.C.Y.; Muszynski, K.; Radzishevsky, I.; Wolosker, H.; Guffanti, G.; Ressler, K.J.; Coyle, J.T. Serine racemase and d-serine in the amygdala are dynamically involved in fear learning. Biol. Psychiatry 2018, 83, 273–283. [Google Scholar] [CrossRef]
  113. Puhl, M.D.; Mintzopoulos, D.; Jensen, J.E.; Gillis, T.E.; Konopaske, G.T.; Kaufman, M.J.; Coyle, J.T. In vivo magnetic resonance studies reveal neuroanatomical and neurochemical abnormalities in the serine racemase knockout mouse model of schizophrenia. Neurobiol. Dis. 2015, 73, 269–274. [Google Scholar] [CrossRef]
  114. Balu, D.T.; Basu, A.C.; Corradi, J.P.; Cacace, A.M.; Coyle, J.T. The NMDA receptor co-agonists, d-serine and glycine, regulate neuronal dendritic architecture in the somatosensory cortex. Neurobiol. Dis. 2012, 45, 671–682. [Google Scholar] [CrossRef] [PubMed]
  115. Balu, D.T.; Coyle, J.T. Neuronal d-serine regulates dendritic architecture in the somatosensory cortex. Neurosci. Lett. 2012, 517, 77–81. [Google Scholar] [CrossRef] [PubMed]
  116. Steullet, P.; Cabungcal, J.H.; Coyle, J.; Didriksen, M.; Gill, K.; Grace, A.A.; Hensch, T.K.; LaMantia, A.S.; Lindemann, L.; Maynard, T.M.; et al. Oxidative stress-driven parvalbumin interneuron impairment as a common mechanism in models of schizophrenia. Mol. Psychiatry 2017, 22, 936–943. [Google Scholar] [CrossRef] [PubMed]
  117. Balu, D.T.; Coyle, J.T. Chronic d-serine reverses arc expression and partially rescues dendritic abnormalities in a mouse model of NMDA receptor hypofunction. Neurochem. Int. 2014, 75, 76–78. [Google Scholar] [CrossRef]
  118. Okada, M. Can rodent models elucidate the pathomechanisms of genetic epilepsy? Br. J. Pharmacol. 2022, 179, 1620–1639. [Google Scholar] [CrossRef]
  119. Farsi, Z.; Nicolella, A.; Simmons, S.K.; Aryal, S.; Shepard, N.; Brenner, K.; Lin, S.; Herzog, L.; Moran, S.P.; Stalnaker, K.J.; et al. Brain-region-specific changes in neurons and glia and dysregulation of dopamine signaling in grin2a mutant mice. Neuron 2023, 111, 3378–3396.e9. [Google Scholar] [CrossRef]
  120. Okada, M.; Fukuyama, K.; Ueda, Y. Lurasidone inhibits NMDA receptor antagonist-induced functional abnormality of thalamocortical glutamatergic transmission via 5-ht(7) receptor blockade. Br. J. Pharmacol. 2019, 176, 4002–4018. [Google Scholar] [CrossRef]
  121. Yamamura, S.; Ohoyama, K.; Hamaguchi, T.; Kashimoto, K.; Nakagawa, M.; Kanehara, S.; Suzuki, D.; Matsumoto, T.; Motomura, E.; Shiroyama, T.; et al. Effects of quetiapine on monoamine, gaba, and glutamate release in rat prefrontal cortex. Psychopharmacology 2009, 206, 243–258. [Google Scholar] [CrossRef]
  122. Amargos-Bosch, M.; Lopez-Gil, X.; Artigas, F.; Adell, A. Clozapine and olanzapine, but not haloperidol, suppress serotonin efflux in the medial prefrontal cortex elicited by phencyclidine and ketamine. Int. J. Neuropsychopharmacol. 2006, 9, 565–573. [Google Scholar] [CrossRef]
  123. Li, Z.; Boules, M.; Williams, K.; Peris, J.; Richelson, E. The novel neurotensin analog nt69l blocks phencyclidine (pcp)-induced increases in locomotor activity and pcp-induced increases in monoamine and amino acids levels in the medial prefrontal cortex. Brain Res. 2010, 1311, 28–36. [Google Scholar] [CrossRef]
  124. Okada, M.; Fukuyama, K. Interaction between mesocortical and mesothalamic catecholaminergic transmissions associated with NMDA receptor in the locus coeruleus. Biomolecules 2020, 10, 990. [Google Scholar] [CrossRef] [PubMed]
  125. Tanahashi, S.; Yamamura, S.; Nakagawa, M.; Motomura, E.; Okada, M. Dopamine d2 and serotonin 5-ht1a receptors mediate the actions of aripiprazole in mesocortical and mesoaccumbens transmission. Neuropharmacology 2012, 62, 765–774. [Google Scholar] [CrossRef] [PubMed]
  126. Tanahashi, S.; Ueda, Y.; Nakajima, A.; Yamamura, S.; Nagase, H.; Okada, M. Novel delta1-receptor agonist knt-127 increases the release of dopamine and l-glutamate in the striatum, nucleus accumbens and median pre-frontal cortex. Neuropharmacology 2012, 62, 2057–2067. [Google Scholar] [CrossRef]
  127. Yamamura, S.; Ohoyama, K.; Hamaguchi, T.; Nakagawa, M.; Suzuki, D.; Matsumoto, T.; Motomura, E.; Tanii, H.; Shiroyama, T.; Okada, M. Effects of zotepine on extracellular levels of monoamine, gaba and glutamate in rat prefrontal cortex. Br. J. Pharmacol. 2009, 157, 656–665. [Google Scholar] [CrossRef]
  128. Okada, M.; Fukuyama, K.; Kawano, Y.; Shiroyama, T.; Ueda, Y. Memantine protects thalamocortical hyper-glutamatergic transmission induced by NMDA receptor antagonism via activation of system xc. Pharmacol. Res. Perspect. 2019, 7, e004572019. [Google Scholar] [CrossRef] [PubMed]
  129. Okada, M.; Fukuyama, K.; Okubo, R.; Shiroyama, T.; Ueda, Y. Lurasidone sub-chronically activates serotonergic transmission via desensitization of 5-ht1a and 5-ht7 receptors in dorsal raphe nucleus. Pharmaceuticals 2019, 12, 149. [Google Scholar] [CrossRef]
  130. Okada, M.; Okubo, R.; Fukuyama, K. Vortioxetine subchronically activates serotonergic transmission via desensitization of serotonin 5-ht(1a) receptor with 5-ht(3) receptor inhibition in rats. Int. J. Mol. Sci. 2019, 20, 6235. [Google Scholar] [CrossRef]
  131. Okada, M.; Fukuyama, K.; Nakano, T.; Ueda, Y. Pharmacological discrimination of effects of mk801 on thalamocortical, mesothalamic, and mesocortical transmissions. Biomolecules 2019, 9, 746. [Google Scholar] [CrossRef]
  132. Fukuyama, K.; Hasegawa, T.; Okada, M. Cystine/glutamate antiporter and aripiprazole compensate NMDA antagonist-induced dysfunction of thalamocortical l-glutamatergic transmission. Int. J. Mol. Sci. 2018, 19, 3645. [Google Scholar] [CrossRef]
  133. Okada, M.; Fukuyama, K.; Shiroyama, T.; Murata, M. A working hypothesis regarding identical pathomechanisms between clinical efficacy and adverse reaction of clozapine via the activation of connexin43. Int. J. Mol. Sci. 2020, 21, 7019. [Google Scholar] [CrossRef] [PubMed]
  134. Okubo, R.; Hasegawa, T.; Fukuyama, K.; Shiroyama, T.; Okada, M. Current limitations and candidate potential of 5-ht7 receptor antagonism in psychiatric pharmacotherapy. Front. Psychiatry 2021, 12, 623684. [Google Scholar] [CrossRef]
  135. Fukuyama, K.; Nakano, T.; Shiroyama, T.; Okada, M. Chronic administrations of guanfacine on mesocortical catecholaminergic and thalamocortical glutamatergic transmissions. Int. J. Mol. Sci. 2021, 22, 4122. [Google Scholar] [CrossRef] [PubMed]
  136. Okada, M.; Fukuyama, K.; Kawano, Y.; Shiroyama, T.; Suzuki, D.; Ueda, Y. Effects of acute and sub-chronic administrations of guanfacine on catecholaminergic transmissions in the orbitofrontal cortex. Neuropharmacology 2019, 156, 107547. [Google Scholar] [CrossRef]
  137. Caspi, A.; Davidson, M.; Tamminga, C.A. Treatment-refractory schizophrenia. Dialogues Clin. Neurosci. 2004, 6, 61–70. [Google Scholar] [CrossRef]
  138. Conley, R.R.; Kelly, D.L. Management of treatment resistance in schizophrenia. Biol. Psychiatry 2001, 50, 898–911. [Google Scholar] [CrossRef] [PubMed]
  139. Suzuki, T.; Remington, G.; Mulsant, B.H.; Rajji, T.K.; Uchida, H.; Graff-Guerrero, A.; Mamo, D.C. Treatment resistant schizophrenia and response to antipsychotics: A review. Schizophr. Res. 2011, 133, 54–62. [Google Scholar] [CrossRef]
  140. Meltzer, H.Y.; Rabinowitz, J.; Lee, M.A.; Cola, P.A.; Ranjan, R.; Findling, R.L.; Thompson, P.A. Age at onset and gender of schizophrenic patients in relation to neuroleptic resistance. Am. J. Psychiatry 1997, 154, 475–482. [Google Scholar]
  141. Teo, C.; Borlido, C.; Kennedy, J.L.; De Luca, V. The role of ethnicity in treatment refractory schizophrenia. Compr. Psychiatry 2013, 54, 167–172. [Google Scholar] [CrossRef]
  142. Wimberley, T.; Stovring, H.; Sorensen, H.J.; Horsdal, H.T.; MacCabe, J.H.; Gasse, C. Predictors of treatment resistance in patients with schizophrenia: A population-based cohort study. Lancet Psychiatry 2016, 3, 358–366. [Google Scholar] [CrossRef] [PubMed]
  143. Reichert, A.; Kreiker, S.; Mehler-Wex, C.; Warnke, A. The psychopathological and psychosocial outcome of early-onset schizophrenia: Preliminary data of a 13-year follow-up. Child. Adolesc. Psychiatry Ment. Health 2008, 2, 6. [Google Scholar] [CrossRef] [PubMed]
  144. Hollis, C. Adult outcomes of child- and adolescent-onset schizophrenia: Diagnostic stability and predictive validity. Am. J. Psychiatry 2000, 157, 1652–1659. [Google Scholar] [CrossRef] [PubMed]
  145. Schennach, R.; Riedel, M.; Musil, R.; Moller, H.J. Treatment response in first-episode schizophrenia. Clin. Psychopharmacol. Neurosci. 2012, 10, 78–87. [Google Scholar] [CrossRef]
  146. National Institute for Health and Care Excellence. Psychosis and Schizophrenia in Adults: Prevention and Management. Available online: https://www.ncbi.nlm.nih.gov/pubmed/32207892 (accessed on 1 September 2024).
  147. Hasan, A.; Falkai, P.; Wobrock, T.; Lieberman, J.; Glenthoj, B.; Gattaz, W.F.; Thibaut, F.; Moller, H.J.; World Federation of Societies of Biological Psychiatry Task Force on Treatment Guidelines for Schizophrenia. World federation of societies of biological psychiatry (wfsbp) guidelines for biological treatment of schizophrenia, part 1: Update 2012 on the acute treatment of schizophrenia and the management of treatment resistance. World J. Biol. Psychiatry 2012, 13, 318–378. [Google Scholar] [CrossRef]
  148. McEvoy, J.P.; Lieberman, J.A.; Stroup, T.S.; Davis, S.M.; Meltzer, H.Y.; Rosenheck, R.A.; Swartz, M.S.; Perkins, D.O.; Keefe, R.S.; Davis, C.E.; et al. Effectiveness of clozapine versus olanzapine, quetiapine, and risperidone in patients with chronic schizophrenia who did not respond to prior atypical antipsychotic treatment. Am. J. Psychiatry 2006, 163, 600–610. [Google Scholar] [CrossRef]
  149. Lewis, S.W.; Barnes, T.R.; Davies, L.; Murray, R.M.; Dunn, G.; Hayhurst, K.P.; Markwick, A.; Lloyd, H.; Jones, P.B. Randomized controlled trial of effect of prescription of clozapine versus other second-generation antipsychotic drugs in resistant schizophrenia. Schizophr. Bull. 2006, 32, 715–723. [Google Scholar] [CrossRef]
  150. Samara, M.T.; Dold, M.; Gianatsi, M.; Nikolakopoulou, A.; Helfer, B.; Salanti, G.; Leucht, S. Efficacy, acceptability, and tolerability of antipsychotics in treatment-resistant schizophrenia: A network meta-analysis. JAMA Psychiatry 2016, 73, 199–210. [Google Scholar] [CrossRef]
  151. Kane, J.M.; Correll, C.U. The role of clozapine in treatment-resistant schizophrenia. JAMA Psychiatry 2016, 73, 187–188. [Google Scholar] [CrossRef]
  152. Meltzer, H.Y. Treatment of the neuroleptic-nonresponsive schizophrenic patient. Schizophr. Bull. 1992, 18, 515–542. [Google Scholar] [CrossRef]
  153. Lieberman, J.A.; Safferman, A.Z.; Pollack, S.; Szymanski, S.; Johns, C.; Howard, A.; Kronig, M.; Bookstein, P.; Kane, J.M. Clinical effects of clozapine in chronic schizophrenia: Response to treatment and predictors of outcome. Am. J. Psychiatry 1994, 151, 1744–1752. [Google Scholar] [PubMed]
  154. Fukuyama, K.; Motomura, E.; Okada, M. A novel gliotransmitter, l-beta-aminoisobutyric acid, contributes to pathophysiology of clinical efficacies and adverse reactions of clozapine. Biomolecules 2023, 13, 1288. [Google Scholar] [CrossRef]
  155. Correll, C.U.; Howes, O.D. Treatment-resistant schizophrenia: Definition, predictors, and therapy options. J. Clin. Psychiatry 2021, 82, 36608. [Google Scholar] [CrossRef]
  156. Mouchlianitis, E.; McCutcheon, R.; Howes, O.D. Brain-imaging studies of treatment-resistant schizophrenia: A systematic review. Lancet Psychiatry 2016, 3, 451–463. [Google Scholar] [CrossRef]
  157. Potkin, S.G.; Kane, J.M.; Correll, C.U.; Lindenmayer, J.P.; Agid, O.; Marder, S.R.; Olfson, M.; Howes, O.D. The neurobiology of treatment-resistant schizophrenia: Paths to antipsychotic resistance and a roadmap for future research. NPJ Schizophr. 2020, 6, 1. [Google Scholar] [CrossRef]
  158. Chouinard, G.; Samaha, A.N.; Chouinard, V.A.; Peretti, C.S.; Kanahara, N.; Takase, M.; Iyo, M. Antipsychotic-induced dopamine supersensitivity psychosis: Pharmacology, criteria, and therapy. Psychother. Psychosom. 2017, 86, 189–219. [Google Scholar] [CrossRef]
  159. Siskind, D.; Siskind, V.; Kisely, S. Clozapine response rates among people with treatment-resistant schizophrenia: Data from a systematic review and meta-analysis. Can. J. Psychiatry 2017, 62, 772–777. [Google Scholar] [CrossRef]
  160. Yoshimura, B.; Yada, Y.; So, R.; Takaki, M.; Yamada, N. The critical treatment window of clozapine in treatment-resistant schizophrenia: Secondary analysis of an observational study. Psychiatry Res. 2017, 250, 65–70. [Google Scholar] [CrossRef] [PubMed]
  161. Agid, O.; Arenovich, T.; Sajeev, G.; Zipursky, R.B.; Kapur, S.; Foussias, G.; Remington, G. An algorithm-based approach to first-episode schizophrenia: Response rates over 3 prospective antipsychotic trials with a retrospective data analysis. J. Clin. Psychiatry 2011, 72, 1439–1444. [Google Scholar] [CrossRef] [PubMed]
  162. Kahn, R.S.; Winter van Rossum, I.; Leucht, S.; McGuire, P.; Lewis, S.W.; Leboyer, M.; Arango, C.; Dazzan, P.; Drake, R.; Heres, S.; et al. Amisulpride and olanzapine followed by open-label treatment with clozapine in first-episode schizophrenia and schizophreniform disorder (optimise): A three-phase switching study. Lancet Psychiatry 2018, 5, 797–807. [Google Scholar] [CrossRef]
  163. John, A.P.; Ko, E.K.F.; Dominic, A. Delayed initiation of clozapine continues to be a substantial clinical concern. Can. J. Psychiatry 2018, 63, 526–531. [Google Scholar] [CrossRef] [PubMed]
  164. Seeman, P. Clozapine, a fast-off-d2 antipsychotic. ACS Chem. Neurosci. 2014, 5, 24–29. [Google Scholar] [CrossRef] [PubMed]
  165. Kapur, S.; Seeman, P. Does fast dissociation from the dopamine d2 receptor explain the action of atypical antipsychotics?: A new hypothesis. Am. J. Psychiatry 2001, 158, 360–369. [Google Scholar] [CrossRef]
  166. Stepnicki, P.; Kondej, M.; Kaczor, A.A. Current concepts and treatments of schizophrenia. Molecules 2018, 23, 2087. [Google Scholar] [CrossRef]
  167. Sykes, D.A.; Moore, H.; Stott, L.; Holliday, N.; Javitch, J.A.; Lane, J.R.; Charlton, S.J. Extrapyramidal side effects of antipsychotics are linked to their association kinetics at dopamine d(2) receptors. Nat. Commun. 2017, 8, 763. [Google Scholar] [CrossRef] [PubMed]
  168. Sahlholm, K.; Zeberg, H.; Nilsson, J.; Ogren, S.O.; Fuxe, K.; Arhem, P. The fast-off hypothesis revisited: A functional kinetic study of antipsychotic antagonism of the dopamine d2 receptor. Eur. Neuropsychopharmacol. 2016, 26, 467–476. [Google Scholar] [CrossRef] [PubMed]
  169. Kapur, S.; Seeman, P. Antipsychotic agents differ in how fast they come off the dopamine d2 receptors. Implications for atypical antipsychotic action. J. Psychiatry Neurosci. 2000, 25, 161–166. [Google Scholar]
  170. Schrader, J.M.; Irving, C.M.; Octeau, J.C.; Christian, J.A.; Aballo, T.J.; Kareemo, D.J.; Conti, J.; Camberg, J.L.; Lane, J.R.; Javitch, J.A.; et al. The differential actions of clozapine and other antipsychotic drugs on the translocation of dopamine d2 receptors to the cell surface. J. Biol. Chem. 2019, 294, 5604–5615. [Google Scholar] [CrossRef]
  171. Lawlor, B.A.; Davis, K.L. Does modulation of glutamatergic function represent a viable therapeutic strategy in alzheimer’s disease? Biol. Psychiatry 1992, 31, 337–350. [Google Scholar] [CrossRef]
  172. Hashimoto, K.; Malchow, B.; Falkai, P.; Schmitt, A. Glutamate modulators as potential therapeutic drugs in schizophrenia and affective disorders. Eur. Arch. Psychiatry Clin. Neurosci. 2013, 263, 367–377. [Google Scholar] [CrossRef]
  173. Tsai, G.E.; Lin, P.Y. Strategies to enhance n-methyl-d-aspartate receptor-mediated neurotransmission in schizophrenia, a critical review and meta-analysis. Curr. Pharm. Des. 2010, 16, 522–537. [Google Scholar] [CrossRef]
  174. Heresco-Levy, U.; Javitt, D.C.; Ebstein, R.; Vass, A.; Lichtenberg, P.; Bar, G.; Catinari, S.; Ermilov, M. D-serine efficacy as add-on pharmacotherapy to risperidone and olanzapine for treatment-refractory schizophrenia. Biol. Psychiatry 2005, 57, 577–585. [Google Scholar] [CrossRef]
  175. Kantrowitz, J.T.; Malhotra, A.K.; Cornblatt, B.; Silipo, G.; Balla, A.; Suckow, R.F.; D’Souza, C.; Saksa, J.; Woods, S.W.; Javitt, D.C. High dose d-serine in the treatment of schizophrenia. Schizophr. Res. 2010, 121, 125–130. [Google Scholar] [CrossRef]
  176. Lane, H.Y.; Chang, Y.C.; Liu, Y.C.; Chiu, C.C.; Tsai, G.E. Sarcosine or d-serine add-on treatment for acute exacerbation of schizophrenia: A randomized, double-blind, placebo-controlled study. Arch. Gen. Psychiatry 2005, 62, 1196–1204. [Google Scholar] [CrossRef]
  177. Weiser, M.; Heresco-Levy, U.; Davidson, M.; Javitt, D.C.; Werbeloff, N.; Gershon, A.A.; Abramovich, Y.; Amital, D.; Doron, A.; Konas, S.; et al. A multicenter, add-on randomized controlled trial of low-dose d-serine for negative and cognitive symptoms of schizophrenia. J. Clin. Psychiatry 2012, 73, e728–e734. [Google Scholar] [CrossRef] [PubMed]
  178. Cascella, N.G.; Macciardi, F.; Cavallini, C.; Smeraldi, E. D-cycloserine adjuvant therapy to conventional neuroleptic treatment in schizophrenia: An open-label study. J. Neural Transm. Gen. Sect. 1994, 95, 105–111. [Google Scholar] [CrossRef] [PubMed]
  179. Baxter, M.G.; Lanthorn, T.H.; Frick, K.M.; Golski, S.; Wan, R.Q.; Olton, D.S. D-cycloserine, a novel cognitive enhancer, improves spatial memory in aged rats. Neurobiol. Aging 1994, 15, 207–213. [Google Scholar] [CrossRef] [PubMed]
  180. Meftah, A.; Hasegawa, H.; Kantrowitz, J.T. D-serine: A cross species review of safety. Front. Psychiatry 2021, 12, 726365. [Google Scholar] [CrossRef]
  181. Boje, K.M.; Wong, G.; Skolnick, P. Desensitization of the NMDA receptor complex by glycinergic ligands in cerebellar granule cell cultures. Brain Res. 1993, 603, 207–214. [Google Scholar] [CrossRef]
  182. Parnas, A.S.; Weber, M.; Richardson, R. Effects of multiple exposures to d-cycloserine on extinction of conditioned fear in rats. Neurobiol. Learn. Mem. 2005, 83, 224–231. [Google Scholar] [CrossRef]
  183. Goff, D.C.; Herz, L.; Posever, T.; Shih, V.; Tsai, G.; Henderson, D.C.; Freudenreich, O.; Evins, A.E.; Yovel, I.; Zhang, H.; et al. A six-month, placebo-controlled trial of d-cycloserine co-administered with conventional antipsychotics in schizophrenia patients. Psychopharmacology 2005, 179, 144–150. [Google Scholar] [CrossRef] [PubMed]
  184. Gorelik, E.; Gorelik, B.; Masarwa, R.; Perlman, A.; Hirsh-Raccah, B.; Matok, I. The cardiovascular safety of antiobesity drugs-analysis of signals in the fda adverse event report system database. Int. J. Obes. 2020, 44, 1021–1027. [Google Scholar] [CrossRef] [PubMed]
  185. Suwandhi, L.; Hausmann, S.; Braun, A.; Gruber, T.; Heinzmann, S.S.; Galvez, E.J.C.; Buck, A.; Legutko, B.; Israel, A.; Feuchtinger, A.; et al. Chronic d-serine supplementation impairs insulin secretion. Mol. Metab. 2018, 16, 191–202. [Google Scholar] [CrossRef]
  186. Marques, B.L.; Oliveira-Lima, O.C.; Carvalho, G.A.; de Almeida Chiarelli, R.; Ribeiro, R.I.; Parreira, R.C.; da Madeira Freitas, E.M.; Resende, R.R.; Klempin, F.; Ulrich, H.; et al. Neurobiology of glycine transporters: From molecules to behavior. Neurosci. Biobehav. Rev. 2020, 118, 97–110. [Google Scholar] [CrossRef] [PubMed]
  187. Peyrovian, B.; Rosenblat, J.D.; Pan, Z.; Iacobucci, M.; Brietzke, E.; McIntyre, R.S. The glycine site of NMDA receptors: A target for cognitive enhancement in psychiatric disorders. Prog. Neuropsychopharmacol. Biol. Psychiatry 2019, 92, 387–404. [Google Scholar] [CrossRef]
  188. Lane, H.Y.; Liu, Y.C.; Huang, C.L.; Chang, Y.C.; Liau, C.H.; Perng, C.H.; Tsai, G.E. Sarcosine (N-methylglycine) treatment for acute schizophrenia: A randomized, double-blind study. Biol. Psychiatry 2008, 63, 9–12. [Google Scholar] [CrossRef]
  189. Supplisson, S. Dynamic role of glyt1 as glycine sink or source: Pharmacological implications for the gain control of NMDA receptors. Neuroscience 2024, in press. [Google Scholar]
  190. Lane, H.Y.; Lin, C.H.; Green, M.F.; Hellemann, G.; Huang, C.C.; Chen, P.W.; Tun, R.; Chang, Y.C.; Tsai, G.E. Add-on treatment of benzoate for schizophrenia: A randomized, double-blind, placebo-controlled trial of d-amino acid oxidase inhibitor. JAMA Psychiatry 2013, 70, 1267–1275. [Google Scholar] [CrossRef]
  191. Amiaz, R.; Kent, I.; Rubinstein, K.; Sela, B.A.; Javitt, D.; Weiser, M. Safety, tolerability and pharmacokinetics of open label sarcosine added on to anti-psychotic treatment in schizophrenia—Preliminary study. Isr. J. Psychiatry Relat. Sci. 2015, 52, 12–15. [Google Scholar]
  192. Uno, Y.; Coyle, J.T. Glutamate hypothesis in schizophrenia. Psychiatry Clin. Neurosci. 2019, 73, 204–215. [Google Scholar] [CrossRef]
  193. Fleischhacker, W.W.; Podhorna, J.; Groschl, M.; Hake, S.; Zhao, Y.; Huang, S.; Keefe, R.S.E.; Desch, M.; Brenner, R.; Walling, D.P.; et al. Efficacy and safety of the novel glycine transporter inhibitor bi 425809 once daily in patients with schizophrenia: A double-blind, randomised, placebo-controlled phase 2 study. Lancet Psychiatry 2021, 8, 191–201. [Google Scholar] [CrossRef]
  194. Harvey, P.D.; Bowie, C.R.; McDonald, S.; Podhorna, J. Evaluation of the efficacy of bi 425809 pharmacotherapy in patients with schizophrenia receiving computerized cognitive training: Methodology for a double-blind, randomized, parallel-group trial. Clin. Drug Investig. 2020, 40, 377–385. [Google Scholar] [CrossRef] [PubMed]
  195. Dudzik, P.; Lustyk, K.; Pytka, K. Beyond dopamine: Novel strategies for schizophrenia treatment. Med. Res. Rev. 2024, 44, 2307–2330. [Google Scholar] [CrossRef] [PubMed]
  196. Evins, A.E.; Fitzgerald, S.M.; Wine, L.; Rosselli, R.; Goff, D.C. Placebo-controlled trial of glycine added to clozapine in schizophrenia. Am. J. Psychiatry 2000, 157, 826–828. [Google Scholar] [CrossRef] [PubMed]
  197. Tsai, G.E.; Yang, P.; Chung, L.C.; Tsai, I.C.; Tsai, C.W.; Coyle, J.T. D-serine added to clozapine for the treatment of schizophrenia. Am. J. Psychiatry 1999, 156, 1822–1825. [Google Scholar] [CrossRef]
  198. Goff, D.C.; Tsai, G.; Levitt, J.; Amico, E.; Manoach, D.; Schoenfeld, D.A.; Hayden, D.L.; McCarley, R.; Coyle, J.T. A placebo-controlled trial of d-cycloserine added to conventional neuroleptics in patients with schizophrenia. Arch. Gen. Psychiatry 1999, 56, 21–27. [Google Scholar] [CrossRef]
  199. Evins, A.E.; Amico, E.; Posever, T.A.; Toker, R.; Goff, D.C. D-cycloserine added to risperidone in patients with primary negative symptoms of schizophrenia. Schizophr. Res. 2002, 56, 19–23. [Google Scholar] [CrossRef]
  200. Schwieler, L.; Linderholm, K.R.; Nilsson-Todd, L.K.; Erhardt, S.; Engberg, G. Clozapine interacts with the glycine site of the NMDA receptor: Electrophysiological studies of dopamine neurons in the rat ventral tegmental area. Life Sci. 2008, 83, 170–175. [Google Scholar] [CrossRef]
  201. Fukuyama, K.; Kato, R.; Murata, M.; Shiroyama, T.; Okada, M. Clozapine normalizes a glutamatergic transmission abnormality induced by an impaired NMDA receptor in the thalamocortical pathway via the activation of a group iii metabotropic glutamate receptor. Biomolecules 2019, 9, 234. [Google Scholar] [CrossRef]
  202. Fukuyama, K.; Motomura, E.; Okada, M. Enhanced l-beta-aminoisobutyric acid is involved in the pathophysiology of effectiveness for treatment-resistant schizophrenia and adverse reactions of clozapine. Biomolecules 2023, 13, 862. [Google Scholar] [CrossRef]
  203. Fukuyama, K.; Motomura, E.; Okada, M. Opposing effects of clozapine and brexpiprazole on beta-aminoisobutyric acid: Pathophysiology of antipsychotics-induced weight gain. Schizophrenia 2023, 9, 8. [Google Scholar] [CrossRef]
  204. Gray, L.; van den Buuse, M.; Scarr, E.; Dean, B.; Hannan, A.J. Clozapine reverses schizophrenia-related behaviours in the metabotropic glutamate receptor 5 knockout mouse: Association with n-methyl-d-aspartic acid receptor up-regulation. Int. J. Neuropsychopharmacol. 2009, 12, 45–60. [Google Scholar] [CrossRef]
  205. Manahan-Vaughan, D.; Wildforster, V.; Thomsen, C. Rescue of hippocampal ltp and learning deficits in a rat model of psychosis by inhibition of glycine transporter-1 (glyt1). Eur. J. Neurosci. 2008, 28, 1342–1350. [Google Scholar] [CrossRef] [PubMed]
  206. Shimazaki, T.; Kaku, A.; Chaki, S. D-serine and a glycine transporter-1 inhibitor enhance social memory in rats. Psychopharmacology 2010, 209, 263–270. [Google Scholar] [CrossRef] [PubMed]
  207. Chaki, S.; Shimazaki, T.; Karasawa, J.; Aoki, T.; Kaku, A.; Iijima, M.; Kambe, D.; Yamamoto, S.; Kawakita, Y.; Shibata, T.; et al. Efficacy of a glycine transporter 1 inhibitor tasp0315003 in animal models of cognitive dysfunction and negative symptoms of schizophrenia. Psychopharmacology 2015, 232, 2849–2861. [Google Scholar] [CrossRef]
  208. Deiana, S.; Hauber, W.; Munster, A.; Sommer, S.; Ferger, B.; Marti, A.; Schmid, B.; Dorner-Ciossek, C.; Rosenbrock, H. Pro-cognitive effects of the glyt1 inhibitor bitopertin in rodents. Eur. J. Pharmacol. 2022, 935, 175306. [Google Scholar] [CrossRef] [PubMed]
  209. Hashimoto, K.; Fujita, Y.; Ishima, T.; Chaki, S.; Iyo, M. Phencyclidine-induced cognitive deficits in mice are improved by subsequent subchronic administration of the glycine transporter-1 inhibitor nfps and d-serine. Eur. Neuropsychopharmacol. 2008, 18, 414–421. [Google Scholar] [CrossRef]
  210. Goff, D.C. D-cycloserine: An evolving role in learning and neuroplasticity in schizophrenia. Schizophr. Bull. 2012, 38, 936–941. [Google Scholar] [CrossRef]
  211. Nielsen, J.; Young, C.; Ifteni, P.; Kishimoto, T.; Xiang, Y.T.; Schulte, P.F.; Correll, C.U.; Taylor, D. Worldwide differences in regulations of clozapine use. CNS Drugs 2016, 30, 149–161. [Google Scholar] [CrossRef]
  212. Kikuchi, Y.S.; Sato, W.; Ataka, K.; Yagisawa, K.; Omori, Y.; Kanbayashi, T.; Shimizu, T. Clozapine-induced seizures, electroencephalography abnormalities, and clinical responses in japanese patients with schizophrenia. Neuropsychiatr. Dis. Treat. 2014, 10, 1973–1978. [Google Scholar] [CrossRef] [PubMed]
  213. De Leon, J.; Sanz, E.J.; De Las Cuevas, C. Data from the world health organization’s pharmacovigilance database supports the prominent role of pneumonia in mortality associated with clozapine adverse drug reactions. Schizophr. Bull. 2020, 46, 1–3. [Google Scholar] [CrossRef]
  214. Rohde, C.; Siskind, D.; de Leon, J.; Nielsen, J. Antipsychotic medication exposure, clozapine, and pneumonia: Results from a self-controlled study. Acta Psychiatr. Scand. 2020, 142, 78–86. [Google Scholar] [CrossRef]
  215. Blackman, G.; Oloyede, E. Clozapine discontinuation withdrawal symptoms in schizophrenia. Ther. Adv. Psychopharmacol. 2021, 11, 20451253211032053. [Google Scholar] [CrossRef] [PubMed]
  216. Lander, M.; Bastiampillai, T.; Sareen, J. Review of withdrawal catatonia: What does this reveal about clozapine? Transl. Psychiatry 2018, 8, 139. [Google Scholar] [CrossRef] [PubMed]
  217. Schmieden, V.; Betz, H. Pharmacology of the inhibitory glycine receptor: Agonist and antagonist actions of amino acids and piperidine carboxylic acid compounds. Mol. Pharmacol. 1995, 48, 919–927. [Google Scholar] [PubMed]
  218. Minato, T.; Nakamura, N.; Saiki, T.; Miyabe, M.; Ito, M.; Matsubara, T.; Naruse, K. Beta-aminoisobutyric acid, l-baiba, protects pc12 cells from hydrogen peroxide-induced oxidative stress and apoptosis via activation of the ampk and pi3k/akt pathway. IBRO Neurosci. Rep. 2022, 12, 65–72. [Google Scholar] [CrossRef]
Figure 1. Schematic structure of the N-methyl-D-aspartate glutamate receptor (NMDAR). NMDAR is a heterotetrameric complex with GluN1, GluN2, and GluN3 subunits. S1: glutathione binding site (in GluN1 and GluN2A). S2: Zn2+ binding site (Zn2+: in GluN2A and ifenprodil/Ro25-6981 in GluN2B). S3: glycine binding site in GluN1 (glycine, D-serine, and kynurenic acid). S4: L-glutamate binding site in GluN2. S5: phencyclidine binding site in transmembrane domain (MK801, ketamine, and phencyclidine).
Figure 1. Schematic structure of the N-methyl-D-aspartate glutamate receptor (NMDAR). NMDAR is a heterotetrameric complex with GluN1, GluN2, and GluN3 subunits. S1: glutathione binding site (in GluN1 and GluN2A). S2: Zn2+ binding site (Zn2+: in GluN2A and ifenprodil/Ro25-6981 in GluN2B). S3: glycine binding site in GluN1 (glycine, D-serine, and kynurenic acid). S4: L-glutamate binding site in GluN2. S5: phencyclidine binding site in transmembrane domain (MK801, ketamine, and phencyclidine).
Biomolecules 14 01128 g001
Table 1. Proposed Defining Treatment-Resistant Schizophrenia.
Table 1. Proposed Defining Treatment-Resistant Schizophrenia.
Treatment FailureDurationFailure Criteria
APA
(2004)
>2
>1 (atypical)
>6 weeksLittle/no response to adequate duration/dose (therapeutic range)
NICE
(2014)
>2 (Sequential)
>1 (no clozapine atypical)
4–6 weeksLittle/no response to adequate dose and correct duration
(despite established adherence to medication)
WFSBP (2012)>2 unidentical classes
>1 (atypical)
2–8 weeksNo significant improvement in the psychopathology and/or target symptoms (ensured treatment adherence)
TRRIP (2017)>2 (atypical)
>1 (LAI > 4 Month)
>6 weeks
(therapeutic dose)
More than moderate severity
<20% symptom reduction during prospective or observation of >6 weeks
At least moderate functional impairment based on a validated scale
adherence (>80% of prescribed doses) confirmed by serum levels
Abbreviations: APA: American Psychiatric Association [30], NICE: National Institute for Health and Care Excellence [146], TRRIP: Treatment Response and Resistance in Psychosis [29], WFSBP: World Federation of Societies of Biological Psychiatry [147]. LAI: long-acting injectable antipsychotics.
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

Okubo, R.; Okada, M.; Motomura, E. Dysfunction of the NMDA Receptor in the Pathophysiology of Schizophrenia and/or the Pathomechanisms of Treatment-Resistant Schizophrenia. Biomolecules 2024, 14, 1128. https://doi.org/10.3390/biom14091128

AMA Style

Okubo R, Okada M, Motomura E. Dysfunction of the NMDA Receptor in the Pathophysiology of Schizophrenia and/or the Pathomechanisms of Treatment-Resistant Schizophrenia. Biomolecules. 2024; 14(9):1128. https://doi.org/10.3390/biom14091128

Chicago/Turabian Style

Okubo, Ruri, Motohiro Okada, and Eishi Motomura. 2024. "Dysfunction of the NMDA Receptor in the Pathophysiology of Schizophrenia and/or the Pathomechanisms of Treatment-Resistant Schizophrenia" Biomolecules 14, no. 9: 1128. https://doi.org/10.3390/biom14091128

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop