Next Article in Journal
Structural and Functional Basis of JAMM Deubiquitinating Enzymes in Disease
Next Article in Special Issue
ASC Transporters Mediate D-Serine Transport into Astrocytes Adjacent to Synapses in the Mouse Brain
Previous Article in Journal
Data Size and Quality Matter: Generating Physically-Realistic Distance Maps of Protein Tertiary Structures
Previous Article in Special Issue
Biochemical Properties and Physiological Functions of pLG72: Twenty Years of Investigations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Systematic Review

Rational and Translational Implications of D-Amino Acids for Treatment-Resistant Schizophrenia: From Neurobiology to the Clinics

1
Laboratory of Translational and Molecular Psychiatry and Unit of Treatment-Resistant Psychosis, Section of Psychiatry, Department of Neuroscience, Reproductive Sciences and Dentistry, University of Naples Federico II, 80131 Naples, Italy
2
Clinical Psychopharmacology Program, College of Pharmacy, Idaho State University, Pocatello, ID 83209, USA
*
Author to whom correspondence should be addressed.
Biomolecules 2022, 12(7), 909; https://doi.org/10.3390/biom12070909
Submission received: 15 June 2022 / Revised: 25 June 2022 / Accepted: 26 June 2022 / Published: 29 June 2022

Abstract

:
Schizophrenia has been conceptualized as a neurodevelopmental disorder with synaptic alterations and aberrant cortical–subcortical connections. Antipsychotics are the mainstay of schizophrenia treatment and nearly all share the common feature of dopamine D2 receptor occupancy, whereas glutamatergic abnormalities are not targeted by the presently available therapies. D-amino acids, acting as N-methyl-D-aspartate receptor (NMDAR) modulators, have emerged in the last few years as a potential augmentation strategy in those cases of schizophrenia that do not respond well to antipsychotics, a condition defined as treatment-resistant schizophrenia (TRS), affecting almost 30–40% of patients, and characterized by serious cognitive deficits and functional impairment. In the present systematic review, we address with a direct and reverse translational perspective the efficacy of D-amino acids, including D-serine, D-aspartate, and D-alanine, in poor responders. The impact of these molecules on the synaptic architecture is also considered in the light of dendritic spine changes reported in schizophrenia and antipsychotics’ effect on postsynaptic density proteins. Moreover, we describe compounds targeting D-amino acid oxidase and D-aspartate oxidase enzymes. Finally, other drugs acting at NMDAR and proxy of D-amino acids function, such as D-cycloserine, sarcosine, and glycine, are considered in the light of the clinical burden of TRS, together with other emerging molecules.

1. Introduction

Schizophrenia is a severe psychiatric disorder that affects approximately 0.7% of the world population, striking young adults or adolescents at the onset, and causing significant impairment in social and occupational functioning, seriously impacting the quality of life of patients and their families [1]. Clinically, three main psychopathological domains can be present: positive (hallucinations, delusions, disorganization), negative (flat affect, social withdrawal, anhedonia, avolition), and cognitive (altered executive functions, impaired working memory) symptoms.
Antipsychotics, acting mainly albeit not exclusively through the occupancy of dopamine D2 receptor (D2R), represent the cornerstone of schizophrenia pharmacological treatment and have proven highly effective in managing positive symptoms while being limitedly efficacious on negative and cognitive ones. However, approximately 30% of schizophrenia patients do not respond to two consecutive antipsychotics, each of them used at the highest dose and at least for six weeks of treatment; these patients are defined as treatment-resistant [2,3]. Treatment-resistant schizophrenia (TRS) patients have more severe abnormal functioning [4], cognitive deficits [5,6], and neurological soft signs [7]. The only antipsychotic available for TRS therapy is clozapine [2,8,9,10], a highly efficacious drug whose utilization in clinical practice is undermined by relatively rare but severe adverse events such as agranulocytosis, myocarditis, intestinal hypomotility, and convulsions [11]. Therefore, clozapine is largely underused and frequently introduced very late, when the trajectory of the disease is already advanced [12]. Moreover, it is unrealistic that a complex disorder such as schizophrenia could be treated with just one pharmacological agent tackling all the psychopathological domains of the disease. The complexity of the clinical presentation is mirrored by the complexity of molecular aberrations underpinning the disease neurobiology, driving the concept of schizophrenia as a polygenic and multifactorial disorder characterized by abnormal synaptic plasticity and altered cortical–subcortical connectivity [13,14]. In this framework, there is a significant need for novel pharmacological strategies that take into account preclinical and clinical findings in a direct and reverse translational fashion [15,16]. One of the proposed strategies to counteract poor response to antipsychotics in schizophrenia patients is based on D-amino acids’ utilization or augmentation, and thus several clinical trials have been performed with D-serine, D-alanine, and D-amino oxidase inhibitors to assess the efficacy and tolerability of these novel therapeutic molecules as well as their possible future applicability [17,18,19,20,21,22,23,24,25,26,27,28,29,30]. Finally, a further therapeutic approach in a similar direction is represented by the investigation of a few compounds equipped with D-amino acid-related mechanism of action, such as D-cycloserine, sarcosine [31,32], and glycine, whose preclinical investigation [33] as well as utilization in experimental design have indirectly contributed to expanding our knowledge on D-amino acids potential role in TRS. This review aims at addressing the preclinical and clinical evidence in support of the role of D-amino acids in treatment resistance or poor responder schizophrenia patients. We will tackle the following questions:
(1)
How robust is the evidence supporting an alteration of the glutamate system and dopamine-glutamate interaction in TRS?
(2)
How and to what extent can D-amino acids dysregulation intercept the glutamatergic abnormalities in TRS?
(3)
Do preclinical and clinical findings substantiate the D-amino acid strategy in TRS?
(4)
What are the putative innovative future trials and targets?
With these questions in mind, we describe the complex and promising landscape of D-amino acid-based potential interventions in TRS.

2. Materials and Methods

We conducted a systematic review of the literature according to the Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) guidelines [34], based on a PubMed/MEDLINE database search from inception up until 5 May 2022, using the following search string: schizophrenia[MeSH Terms] AND ((D-amino acid oxidase[MeSH Terms]) OR (D-aspartic acid[MeSH Terms]) OR (D-serine) OR (D-alanine) OR (D-amino acid)). Searches were also conducted using ClinicalTrials.gov, using the keywords “D-amino acids”, “D-serine”, “D-aspartate”, “DAO inhibitors” or their combinations. Additional records were retrieved by hand-searching the reference lists of included articles. We included in the qualitative synthesis English-written clinical and preclinical studies relevant to the subject, (i) exploring the implications of D-amino acids in the pathophysiology of schizophrenia and TRS; (ii) reporting original data on D-amino acid-centered strategies for the treatment of schizophrenia and refractory psychotic disorders. We did not apply design methodology constraints. After removing duplicates, 430 records were identified. After pilot-testing the study selection, screening for title/abstract and full-text assessment were independently performed by two investigators (AB and LV). Discrepancies were resolved by reaching consensus. The details of the selection process are reported in the PRISMA flow diagram (please see Figure 1). After the screening process, 329 studies were included in the qualitative synthesis.

3. The neurobiology of Treatment-Resistant Schizophrenia and Amino Acid-Related Metabolism

Even though the molecular and pharmacological mechanisms of TRS are still elusive, consistent in vivo evidence mainly from multimodal imaging studies has emerged in recent years, shedding light on the putative neurobiological underpinning of this severe medical condition, especially when compared to treatment-responsive schizophrenia.
Among the most replicated imaging findings addressing in vivo neurotransmission in schizophrenia patients, two stand over all: a) the increased release of striatal dopamine after intravenous stimulants (i.e., amphetamines) in schizophrenia patients compared to normal controls as measured by Positron emission tomography (PET) adopting 11C-raclopride [35,36,37,38]; and (2) the increased dopamine capacity detected by 3,4-dihydroxy-6-[18F]fluoro-l-phenylalanine (18F-DOPA) PET [39]. Both methods and related results strongly support the dopamine hypothesis of schizophrenia [40]. More recently, reduced radioligand binding at D2R, indirectly indicating a high level of dopamine release, has been detected in the thalamus, a region considered a key hub in the cortical–subcortical dopamine–glutamate interaction [41]. However, in patients not responding adequately to antipsychotics and fulfilling the criteria of TRS, a significant increase in striatal dopamine capacity is not detectable, suggesting that in this group of patients the potential neurobiological underpinning of antipsychotics poor response does not lie exclusively on the dopamine cortico-striatal dysregulation [42,43,44].
Among other neurotransmitter systems involved in TRS pathophysiology, the glutamatergic pathway has been the strongest related to disease neurobiology. Glutamate levels have been reported significantly different in multiple brain regions of TRS patients compared to responders and healthy controls, as measured by 1H-magnetic resonance spectroscopy (MRS) [44,45,46,47]. In this regard, the anterior cingulate cortex (ACC) has strongly been implicated in TRS pathophysiology, showing an increase in glutamate concentrations not detectable in antipsychotics responder patients [44,45,46,47]. In a multicenter study including 48 responder and 44 non-responder schizophrenia patients, beyond the significant increase in ACC glutamate levels exhibited by the TRS group, the authors measured 18F-DOPA striatal uptake and no differences were detected between the two groups [44], supporting the glutamatergic and not dopaminergic involvement in TRS pathophysiology. Moreover, the ACC (glutamate + glutamine)/N-acetyl aspartate (NAA) ratio not only has been found higher in TRS patients in comparison with schizophrenia responsive and healthy control groups but has also been negatively associated with working memory scores measured by the working memory index (WMI) of the Wechsler Adult Intelligence Scales—Fourth Edition (WAIS-IV) [47]. In line with this trend, a decrease in ACC cortical thickness has been related to an increase in ACC glutamate + glutamine levels within TRS and clozapine-resistant patients [48], suggesting that a possible correlation between central glutamate variations and changes in brain structures might provide a fertile background for antipsychotics unresponsiveness. Abnormal ACC glutamate metabolism, poor cognitive performance, and the changes in white matter tracts revealed by MRS have also been reported in early-onset TRS patients, highlighting the possible existence of a predisposing pathological environment present since the beginning of the disease [49]. Further to the clear involvement of ACC, a post hoc analysis conducted in first-episode psychosis (FEP) patients, allowing clinicians to retrospectively define antipsychotic responsiveness or resistance, showed higher glutamate levels in non-responders, positively associated with striatal volume [50]. Again, in the putamen, the total glutamate + glutamine levels have been found increased in TRS patients when compared to first-line responders [51].
To summarize, increasing evidence has been provided in supporting the glutamatergic hypothesis of schizophrenia and even more TRS, with higher glutamate and glutamine levels running in parallel with morphological changes in brain architecture and cognitive decline. The increase in glutamate concentrations has been attributed to N-methyl-D-aspartate receptor (NMDAR) hypofunction, a glutamate ionotropic receptor expressed on the surface of γ-aminobutyric acid (GABA) ergic parvalbumin-positive (PV+) neurons. In fact, poor NMDAR functioning may lead to the disruption of the inhibitory control of the excitatory pyramidal neurons, and thus to the so-called glutamatergic storm [52].
It is noteworthy that glutamate can be affected at multiple molecular levels in refractory schizophrenia [53]: (1) synthesis, which directly involves astrocytes [54]; (2) release [55]; (3) action at postsynaptic [56] and perisynaptic receptors [57]; (4) modulation of postsynaptic density proteins (PSD) [58] and intracellular signaling cascades [59]; (5) reuptake by glial cells through excitatory amino acid transporters (EAATs) [60]. In this framework, D-amino acids and related enzymes may directly or indirectly influence these processes.
Considering the discussed role of glutamate in general schizophrenia pathophysiology and, even more, in TRS, it is not surprising that D-amino acids have been envisioned as a potential pharmacological intervention for strengthening the response to antipsychotics in those patients that do not show improvement with the available treatments [61]. Of interest, one of the earliest reports assessing D-amino acids’ involvement in TRS pathophysiology demonstrated a difference in D-serine plasma levels between TRS and healthy controls, as well as an increase in glycine and D-serine concentrations after clozapine treatment [18], despite the limitation of the “peripheral” detection, which may partially mirror central measures.
It has been proposed that clozapine, the cornerstone of treatment in resistant schizophrenia, may exert its unique antipsychotic effect by directly tuning glutamatergic signaling rather than merely blocking dopaminergic receptors. In fact, neither typical antipsychotics nor D1R agonists can mimic its molecular effects in restoring the glutamatergic pathways [62]. Several lines of evidence emphasized the glutamatergic mode of action of clozapine [63,64], paving the way for the use of modulators of glutamatergic transmission, namely D-amino acids, in the treatment of TRS [62].
The D-amino acid-centered pharmacological strategies may intercept multiple steps of glutamate signaling, directly targeting NMDAR function or increasing D-serine and D-aspartate synaptic concentration via inhibition of D-amino acid oxidase (DAO) and D-aspartate oxidase (DASPO), respectively, both involved in the regulation of NMDAR function [65,66].

4. Dopamine–Glutamate Interaction and the Role of D-Amino Acids

The dopamine hypothesis of schizophrenia has been one of the most enduring assumptions on the origin of the disease, based on the evidence of antipsychotics acting as D2R blockers, and considering the striatal dopamine hyperactivity as responsible for the occurrence of positive symptoms [35,67,68]. Further evidence has been derived from neuroimaging studies showing an increase in striatal dopamine levels after acute amphetamine administration in schizophrenia patients compared to healthy controls [35,37]. On the other hand, clinical observations that acute and repetitive exposure to phencyclidine (PCP), ketamine, and other NMDAR non-competitive antagonists could similarly trigger psychotic-like symptoms in healthy subjects or exacerbate psychotic symptoms in schizophrenia patients have fueled the alternative but not mutually exclusive glutamate hypothesis of schizophrenia [69,70]. This hypothesis has been corroborated by multiple in vitro and in vivo experimental findings [71,72,73]. For instance, in vivo PET imaging studies have shown that NMDAR antagonists, similarly to stimulants, are able to reduce striatal 11C-raclopride D2R binding, suggesting an increase in striatal dopamine release in humans [74,75,76]. Moreover, NMDAR antagonists have been found to induce impairment in executive functions, mimicking cognitive symptoms of schizophrenia, thus supporting the involvement of glutamatergic transmission in the pathophysiology of the disorders [77,78,79,80,81]. The rise of the glutamate hypothesis of schizophrenia has not eclipsed the well-established dopamine theory of schizophrenia, therefore originating a more complex interpretation of the disease based on dopamine-glutamate interaction in which glutamate and dopamine pathways in schizophrenia pathophysiology are converging at multiple levels and with reciprocal influence [82,83].
Furthermore, several lines of evidence demonstrated that NMDAR hypofunction affects corticolimbic GABAergic PV+ interneurons, reducing their excitability with a subsequent loss of their inhibitory action on glutamatergic intracortical neurons controlling downstream midbrain dopaminergic neurons [84]. In turn, dopaminergic firing increases [85,86,87] and contributes to specific schizophrenia-like phenotypes, as shown in transgenic mice, supporting the GABAergic hypothesis of schizophrenia [86,88]. Nikolaus and colleagues presented a series of experiments exploring the effects of agonists and antagonists at both GABAA receptors and NMDARs on dopamine controls in mesolimbic and mesostriatal pathways. They observed that the GABAA receptor agonist muscimol and the NMDAR antagonist amantadine exert similar actions on regional dopamine release, supporting the hypothesis of triple glutamate—GABA—dopamine interaction underlying the dysfunctions observed in schizophrenia [89,90].
Beyond the well-known effect on the dopaminergic pathway, stimulants such as cocaine have been seen to variously modulate the glutamatergic system, participating in the cross-talk between the two neurotransmitters. Specifically, after cocaine withdrawal, a reduction in glutamate concentration has been detected in the nucleus accumbens of rodents, associated with an increase in dendritic spine head diameter and α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor (AMPAR) surface expression, hence raising the AMPAR/NMDAR ratio [91,92]. Furthermore, the dopamine D1 receptor (D1R) stimulation has been shown to enhance glutamatergic neurotransmission, increasing NMDAR phosphorylation and activity via the induction of the downstream protein kinase A (PKA)/dopamine and cAMP-regulated phosphoprotein-32 (DARPP-32)/protein phosphatase 1 (PP1) pathway [93].
The integration of dopaminergic and glutamatergic signals occurs in the PSD, a functional interface acting as a bridge between postsynaptic receptor structures and intracellular downstream transduction pathways [58,94,95,96]. The PSD has been described as a thickening of the glutamatergic postsynaptic terminals, composed of glutamate receptors, ion channels, scaffolding proteins, adaptors, cytoskeletal proteins, and modulators of signaling processes [97,98,99]. Several lines of evidence point to PSD proteins as key molecules implicated in schizophrenia pathophysiology and possibly in its pharmacological treatment [100,101,102,103,104].
Multiple reports have demonstrated that the striatal induction of immediate early genes (IEG), coding for PSD proteins, depends on D1R and its intracellular signaling cascade [105,106,107,108,109]. Nonetheless, the effects of pro-dopaminergic agents such as amphetamines and cocaine on IEG induction may be inhibited by NMDAR antagonists, suggesting the presence of a dopamine–glutamate interaction at the level of striatal circuits or at least at the intracellular level [110]. Moreover, the striatal IEG expression may be differentially modulated according to the specific NMDAR subunit expressed [111]. In fact, the blockade of GluN2A-containing-NMDAR has been associated with a reduction in c-Fos and Zif-268 expression, which otherwise increased after D1R agonist administration [111]. Contrarily, the antagonism at GluN2B-containing-NMDAR has an opposite effect on gene expression, suggesting different regulation of striatal function depending on the specific subunit composition [111].
Through the expression of several PSD proteins, dopaminergic perturbations have been accounted for regulating the shape and growth of dendritic spines and thus the synaptic NMDAR functioning [112]. In particular, the activation of D1Rs but not D2Rs or NMDARs has been found to induce the transcription of Homer1a, a component and regulator of PSD, in the prefrontal cortex (PFC), nucleus accumbens, and ventral tegmental area (VTA) [113]. The elevated intracellular levels of Homer1a protein may lead to: (i) changes in dendritic spines and axons via altered PSD proteins recruitment (such as PSD-95); (ii) disruption of protein complexes and subsequent translocation of metabotropic glutamate receptors (mGluRs) to the membrane surface; (iii) increased calcium entry via transient receptor potential (TRP) calcium channels; (iv) decoupling of mGluR5 and extracellular signal-regulated kinase (ERK) signaling, exerting neuroprotective properties [113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135]. Moreover, Homer1a overexpression has been found to blunt the glutamate response associated with acute cocaine administration [136,137].
With regard to D1R-NMDAR cross-talk converging on PKA/DARPP-32/PP1 pathway, PSD-95 overexpression has been found to disrupt physical interactions by interposing between dopamine and glutamate receptors [138]. On the other hand, lack of PSD-95, as shown in mutant mice, is responsible for the overactivation of both D1R and NMDAR, increasing the susceptibility to NMDAR-mediated excitotoxicity, with subsequent neuronal damage [139].
In this framework, D-amino acids including D-serine may exert a relevant role in dopamine and glutamate interplay, enhancing NMDAR-dependent excitatory postsynaptic currents (EPSCs), and modulating the downstream dopamine neurons’ excitability, indicating a putative involvement in the neurobiological core of schizophrenia [140]. Of interest, D-serine has shown the capability to modulate NMDAR firing, mainly acting as a receptor co-agonist at synaptic sites, in contrast to glycine, exerting its role at the extrasynaptic level [141]. Moreover, it has been observed that the genetic deletion of the DASPO encoding gene has a relevant impact on the expression of PSD proteins [142], highlighting potential implications of D-amino acids for the architecture of the dendritic spine that relies on PSD composition. Indeed, considering PSD, a molecular hub integrating dopamine and glutamate interaction, the D-amino acid impact on this structure is even more appealing as a therapeutic strategy for schizophrenia.
Even a dopaminergic influence on D-serine availability has been suggested, with different roles exerted by D1R and D3R on D-amino acid levels. In particular, tonic spontaneous dopamine release not only enhances NMDAR firing through D1-type receptors but also increases D-serine concentration in PFC, ameliorating cognitive performances in animal models [143]. On the contrary, high phasic dopamine release activates D2-type receptors, which reduce both D-serine levels and NMDAR activation [143].
D-serine has been involved also in GABA–glutamate interplay, being synthesized in striatal GABAergic neurons, serving as a modulator of NMDAR excitability at postsynaptic GABAergic dendrites, which receive glutamatergic afferents, thus limiting the diffusion of glutamatergic inputs from the brain cortex to other structures [144].
Therefore, the relevance of dopamine–glutamate interplay for understanding the neurobiology of schizophrenia, as well as the ability of D-amino acids to interfere with these complex interactions, have raised keen interest in developing novel D-amino acid-based therapeutic strategies.

5. D-Amino Acids as an Innovative Potential Therapeutic Approach to Mitigate NMDAR Hypofunction in Schizophrenia

In the last two decades, the theories about neurotransmission dysfunction in schizophrenia shifted from a classical “dopamine-centric” hypothesis toward a more complex neurotransmitter imbalance, including mainly glutamatergic, serotonergic, GABAergic, and peptidergic transmission [75,77,78,79,80,81,86,88].
Glutamate is the major excitatory neurotransmitter in the central nervous system (CNS), and it is involved in various processes, such as memory, learning, and brain development, through the action on two fundamental types of postsynaptic receptors, namely ionotropic (i.e., NMDAR, AMPAR, and kainate) and metabotropic ones (divided into Groups I, II, and III) [145]. The activation of NMDAR requires its endogenous agonist glutamate and the co-agonist glycine or D-serine to be bound to specific binding sites on the receptor complex [145,146]. The hypofunction of NMDAR has been proposed in schizophrenia upon observing that PCP and its analogues, such as ketamine and MK-801, non-competitive antagonists at NMDAR, can induce a wide range of psychotic symptoms in healthy individuals, not only positive but also negative and cognitive ones, recapitulating the most relevant features of the disease [77]. Particularly, acute and chronic PCP abuse can replicate a schizophrenia-like clinical picture more closely than amphetamines, which are unlikely to mimic negative symptoms and cognitive deficits [79,147,148].
In this framework, NMDAR dysfunction in schizophrenia represents a convergence point of dopaminergic, glutamatergic, and GABAergic alterations, as well as the final common pathway leading from the pathophysiology to the symptom progression [149].
Of interest, the exacerbation of symptoms in schizophrenia patients induced by NMDAR antagonists is only partially relieved by canonical antipsychotics, reacting better at the second-generation neuroleptics having a broader receptor profile, which allows interaction with different neurotransmitter systems. In this context, the possibility of minimizing the hypofunction of NMDAR through the use of D-amino acids acting as co-agonists at this site opens new attractive scenarios.
Psychotropic modulation of glutamate neurotransmission may represent potential monotherapy or add-on treatment strategy for TRS [150]. In particular, D-amino acids augmenting the NMDAR-related transmission may be used in schizophrenia to ameliorate the clinical response to conventional medications.
In the following sections, we will extensively review the preclinical and clinical evidence supporting research attention to D-amino acids and their therapeutic potential in refractory psychiatric disorders.

5.1. NMDAR Co-Agonists

NMDARs in CNS are heteromeric protein complexes composed of at least one GluN1 subunit together with different combinations of GluN2 and/or GluN3 [151]. Alternative splicing of the GRIN1 gene leads to eight GluN1 isoforms, which are required for channel function. The association of GluN1 subunits with different variants of GluN2 (GluN2A-D) and GluN3 (GluN3A and GluN3B) produces a variety of NMDARs with peculiar biophysical and pharmacological properties and specific expression patterns in development and adulthood [152,153]. The dynamic regulation of NMDAR’s activity is also achieved through post-translational modifications, such as glycosylation, phosphorylation, and ubiquitination [154].
At resting membrane potential, the pore of the NMDAR channel is blocked by Mg2+, a block that can be removed upon depolarization by neighboring AMPARs. Aside from the glutamate/NMDA recognition site, GluN1 exhibits a second site for glycine/D-serine (named glycine modulatory site or glycine B site) (Figure 2) [155], and both must be occupied for channel activation. Once finally opened, the NMDAR channel allows for an influx of calcium ions, which activates an intracellular signaling cascade that can ultimately involve gene transcription and synaptic plasticity. NMDAR recruitment during high presynaptic glutamatergic activity results in a permanent increase in synaptic efficacy known as long-term potentiation (LTP) [156]. The persistent activation of NMDAR can cause the sprouting of postsynaptic spines and also have trophic effects on postsynaptic neurons [157], but, on the other hand, an excessive activation can produce oxidative stress and subsequent neuronal death [158]. Indeed, direct agonists at NMDAR are known to be neurotoxic by providing excitotoxicity [159]. Thus, in order to mitigate the NMDAR hypofunction and avoid the excitotoxic consequences of direct receptor activation, current strategies have been carried out on agonists at the glycine B site of the NMDAR complex, using full agonists including D-serine, D-alanine, and glycine, as well as partial agonists such as D-cycloserine.

5.2. D-Serine

Serine is one of the naturally occurring proteinogenic amino acids synthesized in the human body from other metabolites, including glycine. D-serine is synthesized in the brain by serine racemase from L-serine (Figure 3), whose biosynthesis controls its levels [160]. It acts as an endogenous ligand at the glycine B site of NMDAR, playing a central role in mediating NMDAR signaling and plasticity [161]. Recently, abnormalities in the D-serine pathway have been found to suggest a significant contribution to glutamatergic dysfunctions.
As mentioned above, NMDAR activation requires the concomitant binding of glutamate and at least one of glycine or D-serine. However, D-serine was found to be more effective than glycine in increasing glutamatergic neurotransmission [162,163,164,165,166,167,168].
In particular, previous studies have shown that the effective dose required to activate NMDAR was lower for D-serine as compared to glycine, probably due to aromatic residues affecting binding kinetics at the glycine binding site [166,169]. Moreover, immunohistochemical studies have shown that in the telencephalon and developing cerebellum D-serine is expressed in close proximity to NMDARs, whereas the distribution of glycine overlaps the expression of NMDAR in the brainstem, olfactory bulb, and adult cerebellum [170], pointing to D-serine as the major endogenous ligand at the glycine B site at least in the forebrain. In addition, in vivo microdialysis revealed that the extracellular content of free endogenous D-serine was approximately 2.5 times higher than that of glycine in the striatum while being markedly lower in the cerebellum [171].
Basu and colleagues demonstrated that a lack of D-serine may be crucial in the pathophysiology of schizophrenia as observed in a murine model of constitutive D-serine deficiency [172]. They found that mutant mice lacking the capacity to endogenously produce D-serine presented significant alterations in glutamatergic transmission with a subsequent serious deficit in spatial memory and synaptic plasticity, thus reproducing cognitive impairments associated with the schizophrenic endophenotype [172]. Moreover, mutant mice carrying mutations in serine racemase, resulting in a complete loss of enzyme activity, exhibit dramatically reduced D-serine levels and a number of psychotic traits, a significant reduction in the density of inhibitory synapses in the hippocampus [173], and impairments in sensorimotor gating, sociability and spatial discrimination of stimuli [174,175]. Persistent latent inhibition observed in mutant mice carrying a point mutation in NMDAR co-agonist site was effectively reversed by D-serine, which enhances NMDAR glycine B site function, as well as the atypical antipsychotic clozapine [176], whose mechanism of action could involve glial and neuronal exocytosis of D-serine in medial PFC [177].
The relevance of the D-serine pathway in regulating glutamatergic tone is also suggested by in vitro tests indicating the ability of D-serine to enhance NMDAR currents in the hippocampus and striatum [164,178,179]. According to this line of research, increased hippocampal activation after D-serine administration was also corroborated by functional magnetic resonance imaging (fMRI) findings [180].
Several abnormalities in the levels of D-serine and enzymes that modulate D-serine availability were checked in patients suffering from schizophrenia [181]. For instance, a significant reduction in D-serine levels has been found in cerebrospinal fluid (CSF) [182,183] and the serum of patients affected by schizophrenia [184,185], despite discrepancies [186,187]. Furthermore, Ono and colleagues [188] reported a significant elevation in DAO expression in choroid plexus epithelial cells, hypothesizing a subsequent reduction in D-serine cerebrospinal levels in schizophrenia. In a cross-sectional study, Ozeki and colleagues observed an increase in the activity of phosphoserine phosphatase, a rate-limiting enzyme in L-serine synthesis, in peripheral blood mononuclear cells from schizophrenia patients, and a subsequent rise in plasma L-serine accompanied by a reduction in D-serine levels [189].
With regard to the brain concentration of D-serine and related enzymes, as revealed by a machine learning analysis conducted in a post-mortem schizophrenia brain, D-serine resulted to be among the molecules mostly contributing to discriminating schizophrenia from controls [190]. Several studies have also identified abnormalities in DAO expression and activity in the cerebellum [188,190,191], hippocampus [192], and cortex [193]. Verrall et al. detected an elevation in DAO mRNA in the cerebellum of patients, consistent with increased D-serine degradation, whereas serine racemase levels increased in dorsolateral PFC [194], probably reflecting an attempted compensatory mechanism for adapting to the reduction in cortical D-serine. However, given the physiological absence of DAO in the forebrain, which is enriched with D-serine, serine racemase in these areas is known to perform also α,β-elimination reaction with both L-serine and D-serine, contributing to the control of D-serine levels [195,196]. Genetic studies evaluated the association of schizophrenia phenotype with polymorphisms in genes involved in D-serine metabolism. For instance, protein interacting with C kinase (PICK1), a protein required for serine racemase functioning [197], has been found to be altered in “disorganized” schizophrenia, as reported in a case–control study enrolling Japanese subjects [198]. Several studies indicate an association between DAO gene, its potential regulator G72, and schizophrenia [199,200,201,202,203,204]. Of interest, Disrupted-In-Schizophrenia-1 (DISC-1), which has been repeatedly associated with schizophrenia [205], has been found to bind to and stabilize serine racemase. In a preclinical study, it has been demonstrated that mutant DISC-1 fails to bind serine racemase, resulting in a reduction in D-serine levels with subsequent behavioral abnormalities consistent with hypofunction of NMDAR [206]. Conversely, the administration of D-serine in rats seems to have a sex-specific effect, inducing an increase in DISC-1 levels in dorsolateral PFC of female rats and a decrease in males [207]. In the same study, the authors showed that D-serine reduced the expression, in male rats, of nitric oxide synthase 1 adaptor protein (NOS1AP), which has been found overexpressed in the cortex of patients with schizophrenia [207]. Based on the evidence of NOS1AP gene involvement in schizophrenia [208], the relevance for dendrite branching [209], and its role in mediating NMDAR-PSD-95 signaling, further studies suggested that its overexpression in dorsolateral PFC of patients may result in disruption of NMDAR functions [210,211,212]. Similar behavioral abnormalities in adulthood are elicited also by neonatal injection of a selective inhibitor of the enzyme [213].
Although neither conventional nor atypical antipsychotics directly affect brain levels of D-serine [19,194,207,214], it has been argued that amelioration of schizophrenia symptoms is related to elevation in plasma D-serine levels [215,216]. Moreover, several preclinical and clinical studies have shown an association between D-serine availability and cognitive and negative symptoms in schizophrenia [217,218,219]. Indeed, in an animal model of anhedonia with a deficit in sucrose consumption induced by MK-801, acute D-serine administration at 1280 mg/kg is responsible for attenuating MK-801 effects similarly to clozapine, but not haloperidol administration [218]. Furthermore, the systemic administration of D-serine (50 mg/kg/day) is responsible for improved performance in tasks linked to recognition learning and working memory [220].
In a recent preclinical study, the involvement of D-serine in the fine-tuning modulation of dopamine-glutamate cross-talk through D1R and D3R was demonstrated. Of interest, the activation of D1R and D3R has been found to regulate the extracellular levels of D-serine by exerting, respectively, a facilitatory and inhibitory influence on D-serine availability, and thus on NMDAR activation. Thus, dopamine may exert a dual effect on glutamatergic transmission due to, among other mechanisms, the mediation of D-serine, which could represent a nodal role in the dopamine–glutamate interface [143]. Furthermore, D-serine seems to potentiate PFC-dependent cognitive processes by D3R blockade [143].
In the light of this evidence, D-serine has been considered a promising strategy for the rational design of novel pharmacological approaches for treating schizophrenia. Clinical trials testing D-serine administration in schizophrenia subjects are discussed below.

Clinical Reports of D-Serine Efficacy in Treating Schizophrenia Patients Not Responding to Antipsychotics

D-serine has been tested in several randomized double-blind placebo-controlled trials [221,222,223,224,225,226,227,228] as an adjunctive medication in patients receiving risperidone [224], olanzapine [222], clozapine [227], other antipsychotics, or placebo (Table 1). Negative results were obtained with the clozapine combination [227], probably due to the clozapine mechanism of action already involving the modulation of NMDAR functions. A lack of improvement in the Positive and Negative Syndrome Scale (PANSS) total score, positive and negative subscales, and the scale for the assessment of negative symptoms (SANS) in other trials have been ascribed to the dose used (30 mg/kg or 2g/day instead of 60 mg/kg) [221,228] or the confounder of psychotic exacerbation in trials enrolling acutely ill patients instead of chronic ones [224]. However, a recent systematic review and meta-analysis including the above-cited studies demonstrated the effectiveness of D-serine in reducing the total SANS score and the negative subscale of PANSS [229].
Underlining the relevance of the dose administered, also in the light of the extensive and rapid metabolism via DAO, responsible for a reduced bioavailability of the drug [61], it should be noted that Kantrowitz et al. performed a dose-escalation study demonstrating large effect-size improvements in psychotic symptoms both at 60 mg/kg and 120 mg/kg [230]. The same group reported a decrease in the PANSS total score and improvements in auditory mismatch negativity in patients suffering from schizophrenia and schizoaffective disorder after administration of 60 mg/kg/day of D-serine as an add-on treatment to antipsychotics [223].
Taken together, these findings may suggest that 60 mg/kg may represent an adequate dosage to exert a full antipsychotic response, but the optimal dose required remains to be elucidated. Further research exploring the effects of higher doses of D-serine is underway [231]. However, since the catabolism of high doses of D-serine has been shown to produce putative cytotoxic metabolites for glial cells in vitro [232,233] and cause acute tubular necrosis in rats [234], inhibitors of the DAO function have recently been proposed as a safer alternative to D-serine [235,236,237,238].
Although so far considered a promising strategy to treat refractory schizophrenia patients, a study testing D-serine effects in recent-onset schizophrenia is ongoing [239].

5.3. D-Alanine

D-alanine is one of the D-amino acids naturally occurring in the mammalian tissues, present in the human CNS at a concentration estimated to be approximately 12.3 nmol/g in white matter and 9.5 nmol/g in gray matter [240,241,242].
Studies conducted in rats showed that D-alanine concentration reaches a peak at 6 weeks and decreases significantly with age [243]. Moreover, the D-alanine amount slowly declines during the night while increasing during the daytime [243]. Mammals are not capable of synthesizing D-alanine [244], which is primarily derived from the diet, as opposed to D-serine that is biosynthesized by serine racemase [245]. In the mammalian brain, D-alanine plays as a selective and potent co-agonist at NMDARs glycine B site [246] and, similarly to D-serine, it is metabolized by DAO [247]. However, it has been observed that D-alanine may also be directly excreted by the kidneys without being metabolized by DAO [248]. In fact, the co-administration of the DAO inhibitor sodium benzoate and D-alanine did not change D-alanine concentration in CSF in comparison to D-alanine alone, suggesting that this enzyme is not crucial in the regulation of the D-alanine availability [249].
Consistent with glutamatergic dysfunction in schizophrenia, D-alanine administration in animal models is responsible for inhibiting PCP-induced locomotor activity [250], and methamphetamine-induced hyperactivity [251]. Moreover, intracortical infusion of D-alanine was able to attenuate PCP-induced accelerating effects on prefrontal dopamine metabolism [252].
The role of D-alanine in the neurobiology of schizophrenia requires further investigations but increasing evidence has highlighted differences in plasma levels of this amino acid in patients compared to healthy controls. In this regard, Hatano and colleagues demonstrated that D-alanine plasma levels in patients diagnosed with schizophrenia increased significantly from hospital admission, due to higher severity of symptoms, to discharge, positively correlating with clinical improvements [253].
Based on these clinical and preclinical findings, D-alanine was also evaluated in clinical studies (Table 1) testing its efficacy in the treatment of schizophrenia as an add-on to other antipsychotics.
A 6-week double-blind placebo-controlled study reported an improvement in schizophrenia symptoms (based on PANSS positive and Cognitive, and SANS score) after D-alanine administration at 100 mg/kg/day in conjunction with antipsychotics [27].

5.4. D-Aspartate

Along with D-serine, D-aspartate is probably the only other D-amino acid synthesized endogenously by a specific aspartate racemase [254]. In the animal brain, D-aspartate has been detected in hippocampal and frontal cortex neurons, mainly concentrated at the distal tip of axons as well as in neuroendocrine cells and CSF [255,256,257,258]. As an excitatory neurotransmitter, D-aspartate is stored in synaptic vesicles and secretory granules at the axon terminal and is released via a calcium-dependent exocytotic mechanism [259,260]. The transporter system responsible for D-aspartate traffic toward vesicles and its reuptake is yet to be identified but may involve glutamate transporters [261].
D-aspartate exhibits a high time-dependent variability in brain expression, switching from a peak in the gestation period to a dramatic reduction in postnatal life [256,262], running in parallel to the progressive increase in DASPO levels, its catabolizing enzyme [259,263,264,265].
Unlike D-serine, which potentiates NMDAR transmission by acting at the modulatory glycine B site on the GluN1 subunit, D-aspartate stimulates postsynaptic NMDARs by directly binding to the glutamate site on the GluN2 subunit [266,267]. Whereas the spatiotemporal distribution of D-serine seems to closely mimic that of NMDARs, the rise and decline in D-aspartate levels, independently of NMDAR expression, suggests its early contribution to the critical period of CNS maturation [262] and a limited function in the following stages of development.
DASPO is a peroxisomal flavoenzyme primarily expressed in neuronal cells, encoded by the Ddo gene, and involved in the oxidative deamination of acidic D-amino acids, including D-aspartate and D-glutamate, with the following α-keto acid, ammonium, and hydrogen peroxide (H2O2) production [24,268,269]. Even though DAO and DASPO may share a similar molecular structure and a common ancestral origin, the two enzymes diverge for several properties, showing different kinetic binding dynamics, flavin adenine dinucleotide (FAD) affinity, and substrate specificity, with a higher selectivity for neutral and basic or acidic D-amino acids, respectively [24,270]. In this regard, DASPO has shown a 10-fold higher specificity for D-aspartate compared with DAO specificity for D-serine [271,272].
Given the selective D-aspartate time-dependent expression pattern, the availability of Ddo knock-out murine models (Ddo−/−) has generated great interest in the scientific community, allowing clinicians to investigate the effect of DASPO and related metabolites during the whole neurodevelopmental process, thus providing a reliable translational model for mammalian. Moreover, the persistent accumulation of D-aspartate observed in Ddo−/− mice has suggested that DASPO is the unique or at least the major enzyme capable of degrading D-aspartate. Therefore, high enzymatic efficiency might account for enhanced D-aspartate catabolism, responsible for impairments in early NMDAR-related critical processes [22].
Both oral administration of D-aspartate and genetic deletion of Ddo have been found to enhance LTP in mice hippocampus and induce several neuroplasticity modifications, including increased basal brain metabolic activity, dendritic arborization, and spine density [273,274]. Moreover, oral D-aspartate supplementation but not Ddo inactivation has been associated with a significant reduction in cognitive flexibility, interpreted by the authors as the result of LTP saturation [273]. Another plausible explanation for cognitive deterioration may rely on the putative burst in H2O2 production resulting from DASPO enhancement in the D-aspartate treated group, which promotes the establishment of a neuroinflammatory environment.
Since NMDAR expression has not been found altered in several Ddo knock-out models, it has been proposed that the D-aspartate facilitation of LTP induction and maintenance may be exerted by enhancing the sensitivity of NMDAR to endogenous glutamate. Of interest, the effects of D-aspartate administration on glutamatergic transmission are not completely reversed by the NMDAR antagonist MK-801, suggesting that D-aspartate action may be mediated, at least in part, by other receptor complexes, including metabotropic mGluR5 and presynaptic AMPARs [275,276,277].
In light of the glutamatergic hypothesis of schizophrenia centered on NMDAR hypofunction, D-aspartate’s ability to influence NMDAR-dependent postsynaptic currents has gained much attention, and thus, several studies have focused on a possible link with psychotic disorders [142,278,279]. Moreover, the possibility to modulate NMDAR and mGluR5 transmission without producing severe excitotoxic effects has turned the therapeutic focus to D-aspartate and DASPO inhibitors. Accordingly, post-mortem studies revealed that D-aspartate content was reduced by 30-40% in PFC and striatum of schizophrenia patients compared with healthy individuals, due to a concomitant 25% increase in DASPO cortical activity [278,280]. A recent machine learning approach confirmed that, among the molecules of glutamatergic pathway highly predictive of schizophrenia in post-mortem dorsolateral PFC, there were D-serine, D-aspartate/total aspartate ratio, and other postsynaptic proteins such as PSD-95, exhibiting a complex non-linear relationship with the probability of developing the disorder [190]. Errico and colleagues demonstrated also that D-aspartate elevation attenuated the disruptive effects of stimulants and MK-801 on pre-pulse inhibition, a measure of sensorimotor gating generally found abnormal in schizophrenia [281].
Furthermore, several lines of evidence point to the involvement of the D-aspartate pathway in the mechanism of action of well-known antipsychotic agents. For instance, the atypical antipsychotic olanzapine has been reported to exert its pharmacological effects through the inhibition of the DASPO enzyme [282]. Furthermore, both clozapine and olanzapine increase the L-glutamate efflux in PFC as observed after D-aspartate administration [282]. Therefore, the increase in DASPO activity resulting from post-mortem findings should not be ascribed to the action of antipsychotics—which, on the contrary, can even inhibit enzyme activity—but should be considered a primitive feature of the disease.
Intriguingly, whereas D-aspartate increase in Ddo−/− mutant mice has been found to improve memory and LTP in young rodents, it accelerated the age-related decline in cognitive functioning [283]. This biphasic effect on cognition and synaptic plasticity may be explained by striking age-dependent abnormalities induced in the phosphorylation status of ERK1/2 [283]. In fact, the chronic dysregulation of D-aspartate signaling might account for both the enhancement and the worsening of cognitive abilities, mirrored at the molecular level by the initial pro-survival effects of synaptic NMDAR activity and the subsequent harmful effects of excessive NMDAR stimulation [284]. This dual action may account for conflicting results of D-aspartate-centered strategies in ameliorating cognitive symptoms [257,273]. In this context, the physiological postnatal arise in DASPO expression may be necessary to terminate D-aspartate signaling no longer required in adulthood, thus preventing the detrimental neurodegeneration triggered by excessive NMDAR stimulation.
However, it should be noted that D-aspartate is currently approved and marketed as a dietary supplement without producing toxic effects in humans [285,286].

6. D-Amino Acid Oxidase

DAO is an enzyme involved in D-serine and other D-amino acids catabolism via a process of oxidative deamination leading to the production of α-keto acids, ammonium, and H2O2 [287]. As discussed above, DAO and DASPO share a common structure and probably the same origin but exhibit a different substrate specificity and a dissimilar brain expression pattern.
Whilst DAO activity has been detected in human and other mammalian brains, kidneys, and livers [288], the enzyme has not been found expressed in mouse liver [289]. Therefore, given the limitations of species-to-species translatability, caution should be paid in interpreting the results of preclinical studies assessing the efficacy and tolerability of DAO inhibitors in murine models.
Several studies have investigated the oxidase localization in the CNS aiming at identifying the exact enzyme site of action and multiple regional patterns of DAO expression. At the subcellular level, the oxidase was detected in microperoxisomes of both kidney and brain tissues [290]. Moreover, DAO has been traced almost exclusively within type I and type II astrocytes and other neuron supporting cells such as Bergmann glial cells in the cerebellum [290,291,292], while pyramidal neurons have been identified as the primary enzyme localization in other brain regions, including the hippocampus, PFC, and substantia nigra [194]. Although ubiquitously present, higher amounts of DAO mRNA levels have been detected in the cerebellum than in other areas [291,293,294], as opposed to enzyme protein levels, which are mostly expressed in the amygdala, striatum, and PFC and less in the cerebellum [293]. A plausible explanation for different brain patterns exhibited by DAO mRNA and protein expression may include post-transcriptional modifications via microRNA regulation [293].
Considering the evidence of D-serine-mediated NMDAR modulation and the relevant role in D-amino acid metabolism exerted by the oxidase, it is not surprising the growing interest in DAO as a possible neurobiological target for treating psychotic disorders. The first evidence linking DAO with schizophrenia came from a genetic association study seeking disease-related single nucleotide polymorphism (SNPs) in a 5-Mb segment of the chromosome 13q34 [203]. The authors were able to identify G72 as a potential gene involved in schizophrenia pathophysiology and discovered that the gene encoded for a DAO binding protein, known as DAO activator (DAOA), capable of enhancing the enzyme-induced D-serine oxidation [203]. Thus, an analysis of the DAO gene was performed, and four intronic SNPs were found to be associated with schizophrenia in a French-Canadian sample [203]. Noteworthy, as demonstrated by post-mortem studies, the oxidase activity and expression increased bilaterally in the hippocampus and cerebellum of patients suffering from schizophrenia than in healthy controls [183,191,192]. Further, the ratio between serine racemase and DAO protein levels was shown to be significantly lower in the disease group, possibly explaining the decrease in D-serine concentrations observed in the CSF of the affected population [183,191,192]. Afterwards, several experiments have been conducted in murine models to elucidate the putative impact of DAO impairment on cognitive functions. In this regard, DAO knockout mice exhibited better spatial and non-spatial short-term memory performance probably via an increase in D-serine and D-alanine levels followed by an enhancement in NMDAR-mediated glutamatergic neurotransmission [295]. Moreover, in transgenic Grin1D481N mice displaying psychotic phenotypes, the concurrent homozygous point mutation G181R in the DAO1 gene and the subsequent oxidase hypofunction reverted the schizophrenia-like negative and cognitive symptoms, improving spatial memory, cognitive flexibility, promoting social behaviors, and interestingly running in parallel with an increase in D-serine levels [296]. An attempt to explore in vivo DAO influence on CNS signaling pathways has been conducted using the fMRI technique, unveiling differences in brain connectivity patterns within the left putamen, the right posterior cingulate, and the left middle frontal gyri in patients with schizophrenia when split into two subgroups based on different DAOA genotypes [297]. In addition, differences in DAO genotype have been associated with alterations in the functional connectivity of the left precuneus and right posterior cingulate gyrus, both comprised in the default mode network, in patients affected by schizophrenia but not with bipolar disorder nor in healthy controls, suggesting that DAO dysfunctions may be detrimental when acting in a complex set of multiple neurobiological defects [298]. First- and second-generation antipsychotics, beyond the well-defined action on dopaminergic and serotonergic receptors, have been seen to exert an in vitro inhibitory effect on DAO activity, which should be considered as a part of the drug mechanism of action, providing further evidence to support the putative involvement of the oxidase in the pathogenesis of schizophrenia [299,300,301].
Beyond the established hypothesis of D-serine reduction and the following NMDAR hypofunction as the putative mechanism linking oxidase to psychiatric disorders, alterations in other DAO-regulated pathways might be relevant to the emergence of a molecular milieu predisposing to the development of schizophrenia. Specifically, DAO has been found to modulate the dopaminergic system by converting D-DOPA into L-DOPA and hence constituting an alternative metabolic pathway in the dopamine synthesis [302,303,304]. On the other hand, the increase in both dopaminergic VTA neurons firing rate and frontal cortex dopamine levels under DAO genetic or pharmacological inhibitory conditions has probably been due to D-serine-mediated hyperactivation of VTA neurons expressing NMDAR [305,306]. Thus, the controversial role in dopaminergic pathway regulation exerted by the oxidase needs to be better clarified in future studies focusing on DAO net effect apart from the downstream D-serine modulation.
Furthermore, an increase in peripheral DAO levels has been observed in several conditions characterized by cognitive impairment, including Alzheimer’s disease, mild cognitive impairment, and post-stroke dementia [307,308]. Even though the putative mechanism responsible for the association between DAO levels and poor performances in cognitive tasks has not been clearly established, the production of H2O2, resulting from the enzyme catalytic activity and leading to oxidative stress, has been indicated as accountable for the cognitive decline [308]. In this regard, a burst of DAO activity and the following excess in H2O2 levels may disrupt the dynamic pro- and anti-inflammatory balance in favor of the latter, promoting the establishment of a neuroinflammatory environment [309]. Considering DAO-induced reactive oxygen species production, the administration of D-amino acids alone as an augmentation strategy for patients with schizophrenia may be counterproductive, especially on long-term cognitive endpoints. Of interest, the administration of 5-chloro-benzo[d]isoxazol-3-ol (CBIO), a DAO inhibitor, showed to enhance D-serine efficacy in mice, whereas the D-amino acid alone was not effective in attenuating the pre-pulse inhibition deficit induced by MK-801, probably via reduced catabolism, higher D-serine availability, and lower oxidative stress [236,310]. Thus, only future studies addressing the effect of DAO inhibition on neuroinflammatory markers will clarify potential clinical advantages, especially in cognitive domains, coming from the combination between D-amino acids and DAO inhibitors.
In the light of multiple scientific reports assessing and supporting the hypothesis of DAO alteration as a neurobiological hallmark of schizophrenia, novel chemical compounds targeting and suppressing the oxidase activity have been developed to ameliorate psychotic symptoms (Table 2). In this regard, sodium benzoate has been the cornerstone of oxidase inhibitory drugs, showing efficacy in improving positive but not negative or cognitive symptoms in patients with schizophrenia, as assessed by a recent meta-analysis [25]. When sodium benzoate was administered to TRS patients, a significant improvement was detected in overall symptomatology and quality of life in comparison to the placebo group [311]. Furthermore, the combination between sodium benzoate and sarcosine was revealed to be efficient at ameliorating cognitive performances and global functioning in patients with chronic schizophrenia [312]. Moreover, preclinical studies have shown controversial evidence on sodium benzoate pharmacodynamic mechanism, testifying as the drug antipsychotic effect could be achieved in a murine PCP model of schizophrenia even without increasing D-serine levels [313]. Other more selective DAO inhibitory drugs are currently tested in patients with schizophrenia and among these, luvadaxistat has shown a promising improvement in cognitive functions [314,315,316]. Thus, future studies exploring the tolerability and efficacy profile in specific disease subgroups, in addition or not to other drugs enhancing the D-amino acid pathway, will shed light on the real potential of DAO inhibitors in schizophrenia and other psychotic disorders.
The relative specificity of DAO or DAOA polymorphisms in combination with brain functional connectivity alterations or other risk genetic variants for schizophrenia and the ongoing development of more selective and potent oxidase inhibitory agents may open novel possible future directions in both diagnosis and treatment of psychotic disorders, driving to a theranostic innovation in the psychiatric field too.

7. Other Modulators of the Glycine B Site at NMDAR

7.1. Glycine

Glycine is a small nonessential amino acid not containing a chiral center functioning as both an inhibitory and excitatory neurotransmitter depending on the binding site in CNS. The inhibitory action is mediated by a ligand-gated ion channel known as GlyR, pharmacologically characterized by strychnine sensitivity, and mainly localized in the spinal cord where it plays an important role in nociceptive transmission [319,320,321,322,323,324,325]. On the other hand, the excitatory effect depends on the glycine B site, defined as a modulatory strychnine-insensitive site of the GluN1 subunit at the NMDAR ion channel. In the synaptic space, glycine is released by neighboring glial cells and in small amounts by glycinergic neurons. Serine, its endogenous precursor, is converted into glycine by serine-hydroxymethyltranferase (SHMT). A driblet of glycine derives also from N-methyl-glycine (sarcosine) demethylation. A normal diet contains about 2 g of glycine and normally has no significant effect on brain levels. The glycine permeability through the blood–brain barrier (BBB) is very low after peripheral administration; therefore, high oral doses are necessary to obtain an effective increase in CNS levels [326]. The potential involvement of glycine in the modulation of other neurotransmitters systems (such as dopamine-glutamate interplay) and in the neurobiological substrates of mental disorders is supported by the association between clinical phenotypes of schizophrenia with genetic variants in glycinergic pathways and abnormal circulating levels of this molecule [327,328,329,330]. In fact, circulating levels of glycine have been found to correlate with schizophrenia symptoms [331,332]. Moreover, a negative association has been demonstrated between glycine plasma levels and pre-pulse inhibition, a measure of sensorimotor gating commonly found reduced in schizophrenia [333].
An MRS study, evaluating glutamate and glycine levels in both the anterior and posterior cingulate cortex in patients with first-episode psychosis, detected an increase in these two amino acids compared to controls, consistent with the glutamatergic dysfunctions since the acute early phase of psychotic illnesses [328].
Based on these findings, and given the NMDAR hypofunction hypothesis of schizophrenia, the ability of glycine to enhance NMDAR-mediated neurotransmission and its good tolerability profile in both acute and chronic treatment strongly suggests that this molecule might be useful as a therapeutic approach in the pharmacotherapy of schizophrenia, and possibly in TRS [150]. For this purpose, several studies were conducted testing glycine as an add-on therapy to standard antipsychotic treatment. In the literature, fourteen studies have been conducted exploring the efficacy of glycine as an augmentation strategy in schizophrenia treatment, including 11 placebo-controlled randomized clinical trials (RCT) [326,334,335,336,337,338,339,340,341,342,343] and three open-label studies [344,345,346] (Table 3). Recently, data from RCTs have been meta-analysed, showing interesting results. In particular, among NMDAR co-agonists administered in addition to canonical antipsychotics, glycine has proved to be effective in reducing SANS and PANSS total scores, as well as in treatment-refractory patients [229]. On the contrary, NMDAR modulator augmentation was not effective in patients receiving clozapine, probably due to the clozapine “ceiling” effect on the enhancement of NMDAR transmission [229,347,348,349]. These observations may account for divergent and somewhat inconsistent results of glycine augmentation efficacy in those clinical trials enrolling only patients receiving clozapine or which did not distinguish patients on clozapine from the ones on other antipsychotics. With regard to safety outcomes, patients treated with glycine experienced more dry mouth and nausea in comparison with the placebo group [229].

7.1.1. Glycine Transport Inhibitors for the Treatment of Schizophrenia

Another possible approach to improve glutamatergic neurotransmission via NMDAR would be to pharmacologically increase glycine synaptic levels by Glycine Transporter-1 (GlyT-1) inhibitors. GlyT-1 is a sodium/chloride-dependent transporter mainly expressed on glial cells and neurons, particularly on both pre- and postsynaptic terminals (Figure 4) [368,369], allowing for glycine reuptake from the synaptic cleft [370], thus maintaining subsaturating concentrations of glycine at NMDARs [371]. Multiple interactions between GlyT-1 and PSD proteins have been proposed. In particular, PSD-95 stabilizes and anchors GlyT1 at the plasma membranes [369], whereas several kinases, including the Ca2+/calmodulin-dependent protein kinase, phosphorylate the cytosolic regions of GlyT-1, regulating its activity [372].
The importance of GlyT-1 has been proven in preclinical studies conducted in GlyT-1 knock-out mice. In fact, the absence of the GlyT-1 gene has been found to lead to severe respiratory and motor deficits causing premature death [373]. Moreover, GlyT-1 deletion has been associated with neonatal encephalopathy in humans, characterized by impaired consciousness, unresponsiveness, absence of reflexes, hypotonia, and respiratory failure [374]. However, heterozygous GlyT1+/− mice, associated with a reduced but not null activity of GlyT-1, displayed increased memory retention [330] and a distinctive resistance to psychotogenic effects of amphetamines [375].
Given that glycine transport maintains local synaptic glycine concentration at very low levels [376], we can assume that GlyT-1 activity might be targeted in schizophrenia, in order to mitigate the NMDAR hypofunction, reaching a fine-tuning of the excitation-inhibition balance.
Among GlyT-1 inhibitors, sarcosine and bitopertin were tested as adjunctive therapy for schizophrenia (Table 2), as discussed in the sections below.

7.1.2. Sarcosine

Sarcosine, also known as N-methylglycine, is a degradation product of the amino acid glycine. Preclinical studies supported by in vivo brain neuroimaging demonstrated that sarcosine acts as a potent and selective GlyT-1 inhibitor, thus potentiating NMDAR functions and enhancing hippocampal LTP [377], suggesting that this compound may exert antipsychotic and precognitive effects [378,379]. Sarcosine mechanism of action might involve the specific synaptic elevation of glycine but not D-serine, as suggested by microdialysis findings in the PFC [380]. It has been proposed that, given the structural similarity to glycine, sarcosine may directly act also as an NMDAR co-agonist [381] and inhibitory glycine receptor ligand, even less potent than glycine [382]. In fact, sarcosine has been found to elicit the same effects as glycine on NMDAR activation, even in the absence of glycine itself, probably directly binding to the glycine B site [381].
As observed with other D-amino acids, the combination with clozapine does not appear to be effective, probably due to the glycinergic action intrinsically exerted by clozapine, currently the only antipsychotic indicated and evidence-based treatment for TRS [3]. Moreover, sarcosine administration induced a pattern of c-Fos expression resembling that induced by clozapine [377]. Therefore, sarcosine and other D-amino acid-centered strategies may represent alternatives to clozapine in refractory patients.
Several studies investigated sarcosine augmentation efficacy in schizophrenia patients [224,225,312,347,350,352,353,354]. The addition of sarcosine (2 g/day) to antipsychotic medications, including risperidone [350] and other atypicals [225] but not clozapine [347], was found to benefit schizophrenia patients. A randomized double-blind placebo-controlled study demonstrated that sarcosine (2 g/day) is more effective in reducing psychotic symptoms than placebo or D-serine (2 g/day) [224,347]. Moreover, a randomized double-blind study reported that sarcosine alone at the dose of 2 g/day was useful in the treatment of acutely symptomatic drug-free patients affected by schizophrenia [351], suggesting that sarcosine can benefit not only chronic patients but also acutely ill subjects. Furthermore, the addition of sarcosine (2 g/day) has been reported to positively affect the glutamatergic transmission by reducing Glx (a complex of glutamate, glutamine, and GABA)/creatine ratio in white matter of the left frontal lobe as well as in the hippocampus [353,354].
However, these studies have been based on a small sample size of patients, so further large-size placebo-controlled dose finding trials are needed to fully understand the role of sarcosine in schizophrenia. A recent meta-analysis showed that sarcosine surpassed non-sarcosine GlyT-1 inhibitors (i.e., bitopertin) as an augmentation strategy in head-to-head comparisons, mainly targeting negative symptoms [229].
According to this evidence, novel selective GlyT-1 inhibitors have been developed and evaluated for the treatment of schizophrenia. Gaining a more complete understanding of GlyT-1 role in the treatment of schizophrenia patients is expected to provide new therapeutic perspectives for managing this disabling disorder.

7.1.3. Non-Sarcosine Derivatives GlyT-1 Inhibitors

Preclinical studies investigating the effects of sarcosine derivates in rodents have shown the occurrence of numerous side effects such as ataxia, motor and respiratory dysfunctions [327], probably related to the slow dissociation kinetics that allow sarcosine-derivatives to act as pseudo-irreversible inhibitors of GlyT-1 [383]. In this framework, pharmaceutical companies decided to develop non-sarcosine derivatives GlyT-1 inhibitors, such as bitopertin, also known as RG1678 or RO4917838.
Regarding bitopertin, preclinical studies showed its ability to modulate both glutamatergic and dopaminergic signaling, enhance hippocampal LTP [377], and improve performance in PFC-dependent tasks [384], thus mitigating schizophrenia-like behaviors. Early clinical studies pointed to the efficacy of adjunctive bitopertin in treating negative symptoms [355,384], although an inverted U-shaped dose–response profile emerged, leading to inconsistent results when bitopertin was administered at high doses (=60 mg). These results were paralleled, at the preclinical level, by the observations that, at higher concentrations, bitopertin did not affect LTP induction processes, probably as a result of internalization of the NMDAR following an excessive release of glycine [327]. However, in a phase III 24-week double-blind placebo-controlled study, bitopertin failed to reach the primary and secondary endpoints in patients with persistent predominant negative symptoms [356], and further clinical trials testing this molecule were discontinued by Roche. Kantrowitz et al. reported that bitopertin, even at the dose that was most effective in the clinical development programs (10 mg), did not affect NMDAR-based biomarkers including auditory mismatch negativity and visual P1, neurophysiological measures indexing local circuit dysfunctions involved in altered sensory processing and cognitive functioning [357].
Though the bitopertin mechanism of action appeared to be promising, the inverted U-shaped dose–response curve, associated with the wide range of inter-individual variability, may have complicated the identification of the dose needed for efficacy in humans.
Another non-sarcosine GlyT-1 inhibitor, BI 425809, has been recently tested in a double-blind randomized controlled phase II study in patients affected by schizophrenia, resulting in improvements in cognitive symptoms, as assessed by Measurement and Treatment Research to Improve Cognition in Schizophrenia (MATRICS) cognitive battery [358]. The study investigated the effects of 2 mg, 5 mg, 10 mg, or 25 mg BI 425809 administered daily for 12 weeks, as an add-on strategy to antipsychotic monotherapy and cotreatment with a second antipsychotic [358]. BI 425809 was well tolerated in healthy volunteers as well as in patients [358,385]. A randomized trial combining BI 425809 with computerized cognitive training has been recently planned in order to consolidate the results obtained on cognitive functions [384].

7.2. D-Cycloserine

D-cycloserine is a cyclic analogue of serine, which, due to its ability to inhibit the bacterial D-amino acid transaminase, has been marketed as a second-line antibiotic in tuberculosis treatment [386]. Given its action as a partial agonist at the glycine B site, it has been tested in clinical trials as an adjuvant treatment for anxiety disorders, depression, trauma-related, obsessive compulsive disorder [387,388,389,390,391,392], and schizophrenia [388,393]. Regarding D-cycloserine effects in psychiatric disorders, a preclinical study has investigated its role in the modulation of Homer 1a expression after typical and atypical acute antipsychotics administration, showing that the haloperidol and clozapine effect of increasing Homer1a levels in the caudate-putamen and nucleus accumbens was attenuated by D-cycloserine co-administration [394]. D-cycloserine, reducing Homer1a expression and thus its “negative dominant” effect, may enhance mGluR clustering, resulting in a modulation of neuronal and postsynaptic plasticity that might be implicated in the antipsychotic mechanism of action [112,394].
In line with these preclinical observations, several RCTs have been conducted to explore the potential of D-cycloserine as an add-on to antipsychotic treatment for schizophrenia [343,359,360,361,362,363,364,365,366,367,395] but two recent meta-analyses agreed that D-cycloserine failed to exhibit significant efficacy in treating negative, cognitive, or positive symptoms of schizophrenia [229,396].

7.3. D-Peptides

Several D-peptides have recently emerged as potential therapeutic candidates for Alzheimer’s disease and other neurological conditions [397,398]. D-peptides, synthetized in D-enantiomer form by mirror image phage display methodology, are designed to extend the half-life of the peptides and increase resistance to enzyme degradation, prolonging in vivo activity [399]. In particular, D-enantiomeric forms of L-peptides interact with the target proteins of the natural L-amino acid configuration. This technique was employed to develop Tau aggregation inhibitors [398,400] as well as promising D-Aβ-peptides to reduce plaque formation and inflammatory reactions [401,402,403] in Alzheimer’s disease.
Despite limited clinical evidence, D-peptide therapeutics is a rapidly expanding field [404] offering multiple advantages, including low-cost synthesis, low immunogenicity, and serum stability compared to L-peptides analogues that are faster degraded in serum by active endopeptidase. Thus, D-peptides are currently designed and tested in preclinical research setting for a variety of medical conditions, including infections and cancer [405,406,407].

8. Discussion

In the last twenty years, the difficulty of explaining the complexity of schizophrenia clinical presentation by a single brain neurochemical abnormality or one synaptic molecular alteration, as well as the evidence of multiple patterns of connectivity derangements, have raised a keen interest in innovative pharmacological treatments targeting different hallmarks of the disorder, in the attempt to ameliorate the multidomain psychopathological manifestations and provide novel alternatives for TRS patients [1,13,406,407,408,409,410,411,412,413].
In this framework, among multiple strategies proposed to bypass the difficulties of TRS treatment, especially in patients not eligible or non-responder to clozapine, D-amino acids (i.e., D-serine, D-alanine, and D-aspartate) and related proxy molecules (e.g., D-cycloserine, sarcosine) have attracted interest as potential innovative therapies [414]. All D-amino acids and proxy molecules clinically tested have been explored usually as add-on/augmentation therapy or in protocols including patients with poor response to antipsychotics [150], showing a significant capability in decreasing the severity of the overall or subsets of symptoms in schizophrenia and TRS groups [229], even though several limitations of the trials should be acknowledged.
First, the number of patients included in each single study is low, and even if coherent with a pilot study, there is a need for trials with larger sample sizes in order to draw firmer conclusions.
Second, single trials are not easily comparable even among studies investigating the same compound due to different study designs adopted by clinicians, specifically exploring disparate populations and using different antipsychotics when the D-amino acid is tested as add-on therapy [222,224,311,312,318].
Third, all the available D-amino acids tested in schizophrenia and TRS patients are given orally at high doses to ensure significant adsorption and even more important reliable brain levels of the molecules. This is a crucial issue: at present, due to the lack of imaging strategies as the ones that can mirror the quantification of dopamine receptors occupancy by antipsychotics, it is impossible to ascertain with precision, which is the in vivo pharmacodynamics of D-amino acids, hindering the search for the optimal dose accordingly.
Fourth, the cluster of symptoms targeted by the treatment with D-amino acids appears to be different in different studies and complicates the overall comparative interpretation of the studies’ outcomes.
Despite these limitations, the search for a “D-aminoacidergic” compound is strongly pursued both by academia and industry.
In this regard, beyond the molecules extensively discussed above, an emerging field is represented by the investigation of D-cysteine function in the brain and its possible involvement in the pathophysiology and treatment of behavioral disorders. Specifically, D-cysteine binds to Myristoylated Alanine Rich C Kinase Substrate (MARCKS), which is relevant for neuronal survival and migration via Akt pathway regulation [415]. Akt dysfunction has been associated with abnormal dopamine–glutamate interaction in schizophrenia [416,417,418,419] and is reported to be involved in the mechanism of action of antipsychotics [420,421]. The augmentation of the antipsychotic therapy with D-cysteine in TRS could represent an innovative strategy deserving attention and worth being explored in pilot clinical studies, also considering the available information regarding the role of L-cysteine and acetylcysteine in the treatment of behavioral disorders [422].
Moreover, D-aspartate has been well-characterized in mammalian brains [24], and multiple findings from in vivo and in vitro preclinical studies as well as post-mortem investigations in patients have highlighted the possible association with schizophrenia pathophysiology and the putative translational value of this D-amino acid in schizophrenia treatment. In this regard, two findings are worth considering: (1) D-aspartate levels have been found increased after olanzapine chronic administration in mice possibly via the DASPO inhibition, which may represent an unveiled antipsychotic mechanism of action [282] (2) DASPO is responsible for a stable control on D-aspartate concentrations in the brain, differently from DAO, whose low activity affects D-serine concentrations only to a lesser extent [271,272]. Based on the NMDAR dysfunction hypothesis, it has been proposed that inhibitors of the human D-aspartate oxidase (hDASPO) [24] could represent a potentially innovative strategy to counteract the loss of response to antipsychotics in schizophrenia.
The search for “D-aminoacidergic” compounds is an emerging field with significant room for further expansion and few initial results are at the present under evaluation: sodium benzoate (NaBen) for TRS and TAK-831 (luvadaxistat) potentially for negative symptoms of schizophrenia are two examples of a pharmacological strategy based on D-amino acid indirect modulation [309,310,314]. The discovery of a novel class of DAO inhibitors with the Schrödinger computational platform needs to be followed closely in the next few years to establish the real possibility of a clinical application [314].
Regarding possible future directions in clinical and preclinical research, in the light of previous findings extensively discussed above, D-amino-acid-centered strategies might represent an alternative rather than an augmentation approach to clozapine in TRS patients due to a putative “ceiling” effect on NMDAR neurotransmission. In addition, to counteract the possible detrimental effects of high doses of D-amino acids, co-administration of DAO inhibitors might represent a novel therapeutic strategy as testified by preclinical models [236], exploiting the concomitant benefits of complementary molecular pathways, and minimizing possible adverse effects. Furthermore, the development of specific DAO ligands for PET imaging studies, overcoming the problem of BBB low permeability [423], might provide a novel in vivo perspective and serve as a new line of research, clarifying the impact of enzyme changes under multiple pathophysiological conditions, including TRS. The ongoing fine-tuning of genetic and brain imaging techniques, combined with computational approaches, will implement current knowledge on D-amino acids and related enzymes, shedding light on their relevance for TRS neurobiology and providing novel hallmarks of the disease. In this regard, the “D-aminoacidergic” pathway might drive a theranostic innovation, offering both diagnostic and therapeutic targets in TRS.
In conclusion, TRS represents a relevant epidemiological burden affecting almost 30% of schizophrenia patients and significantly impacting the life of the subjects in terms of reduced functioning, cognitive impairment, and overall quality. Only one pharmacological treatment almost fifty years after its first introduction, clozapine, is still available at the present for TRS. Therefore, the search for better and innovative pharmacological strategies is needed, and direct and reverse translational implications as well as preliminary clinical evidence support the D-amino acids strategy for the development of novel and safer therapeutic agents worth exploration in clinical trials.

Author Contributions

Conceptualization: A.d.B.; methodology: A.B., A.d.B., G.D.S. and L.V.; writing—original draft preparation: A.d.B., A.B., L.V. and G.D.S.; writing—review and editing: A.d.B. and M.C.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding. No funders had role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are available upon request.

Conflicts of Interest

A.d.B. has received unrestricted research support from Janssen, Lundbeck, and Otsuka and lecture honoraria for educational meeting from Chiesi, Lundbeck, Roche, Sunovion, Vitria, Recordati, Angelini and Takeda; he has served on advisory boards for Eli Lilly, Jansen, Lundbeck, Otsuka, Roche, Vitria, Chiesi, Recordati, Angelini, and Takeda. No activity is related directly or indirectly to the present manuscript content. All the other authors declare no conflict of interest.

References

  1. McCutcheon, R.A.; Marques, T.R.; Howes, O.D. Schizophrenia-An Overview. JAMA Psychiatry 2020, 77, 201–210. [Google Scholar] [CrossRef] [PubMed]
  2. Correll, C.U.; Howes, O.D. Treatment-Resistant Schizophrenia: Definition, Predictors, and Therapy Options. J. Clin. Psychiatry 2021, 82, 36608. [Google Scholar] [CrossRef] [PubMed]
  3. Howes, O.D.; McCutcheon, R.; Agid, O.; De Bartolomeis, A.; Van Beveren, N.J.; Birnbaum, M.L.; Bloomfield, M.; 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] [PubMed]
  4. Harvey, P.D.; Heaton, R.K.; Carpenter, W.T.; Green, M.F.; Gold, J.M.; Schoenbaum, M. Functional impairment in people with schizophrenia: Focus on employability and eligibility for disability compensation. Schizophr. Res. 2012, 140, 1–8. [Google Scholar] [CrossRef] [Green Version]
  5. De Bartolomeis, A.; Balletta, R.; Giordano, S.; Buonaguro, E.F.; Latte, G.; Iasevoli, F. Differential cognitive performances between schizophrenic responders and non-responders to antipsychotics: Correlation with course of the illness, psychopathology, attitude to the treatment and antipsychotics doses. Psychiatry Res. 2013, 210, 387–395. [Google Scholar] [CrossRef] [PubMed]
  6. Iasevoli, F.; Avagliano, C.; Altavilla, B.; Barone, A.; D’Ambrosio, L.; Matrone, M.; Francesco, D.N.; Razzino, E.; De Bartolomeis, A. Disease Severity in Treatment Resistant Schizophrenia Patients Is Mainly Affected by Negative Symptoms, Which Mediate the Effects of Cognitive Dysfunctions and Neurological Soft Signs. Front. Psychiatry 2018, 9, 553. [Google Scholar] [CrossRef]
  7. De Bartolomeis, A.; Prinzivalli, E.; Callovini, G.; D’Ambrosio, L.; Altavilla, B.; Avagliano, C.; Iasevoli, F. Treatment resistant schizophrenia and neurological soft signs may converge on the same pathology: Evidence from explanatory analysis on clinical, psychopathological, and cognitive variables. Prog. Neuropsychopharmacol. Biol. Psychiatry 2018, 81, 356–366. [Google Scholar] [CrossRef]
  8. Mizuno, K.; Mizuno, E.; Suekane, A.; Shiratsuchi, T. Efficacy of clozapine for long-stay patients with treatment-resistant schizophrenia: 4-year observational study. Neuropsychopharmacol. Rep. 2022, 42, 183–190. [Google Scholar] [CrossRef]
  9. Correll, C.U.; Martin, A.; Patel, C.; Benson, C.; Goulding, R.; Kern-Sliwa, J.; Joshi, K.; Schiller, E.; Kim, E. Systematic literature review of schizophrenia clinical practice guidelines on acute and maintenance management with antipsychotics. Schizophrenia 2022, 8, 5. [Google Scholar] [CrossRef]
  10. Gammon, D.; Cheng, C.; Volkovinskaia, A.; Baker, G.; Dursun, S. Clozapine: Why Is It So Uniquely Effective in the Treatment of a Range of Neuropsychiatric Disorders? Biomolecules 2021, 11, 1030. [Google Scholar] [CrossRef]
  11. Segev, A.; Iqbal, E.; McDonagh, T.A.; Casetta, C.; Oloyede, E.; Piper, S.; Plymen, C.M.; MacCabe, J.H. Clozapine-induced myocarditis: Electronic health register analysis of incidence, timing, clinical markers and diagnostic accuracy. Br. J. Psychiatry 2021, 219, 644–651. [Google Scholar] [CrossRef] [PubMed]
  12. Kane, J.M. Clozapine is underutilized. Shanghai Arch. Psychiatry 2012, 24, 114–115. [Google Scholar] [CrossRef] [PubMed]
  13. Zamanpoor, M. Schizophrenia in a genomic era: A review from the pathogenesis, genetic and environmental etiology to diagnosis and treatment insights. Psychiatr. Genet. 2020, 30, 1–9. [Google Scholar] [CrossRef] [PubMed]
  14. Mäki-Marttunen, T.; Kaufmann, T.; Elvsåshagen, T.; Devor, A.; Djurovic, S.; Westlye, L.T.; Linne, M.-L.; Rietschel, M.; Schubert, D.; Borgwardt, S.; et al. Biophysical Psychiatry—How Computational Neuroscience Can Help to Understand the Complex Mechanisms of Mental Disorders. Front. Psychiatry 2019, 10, 534. [Google Scholar] [CrossRef] [Green Version]
  15. Wada, M.; Noda, Y.; Iwata, Y.; Tsugawa, S.; Yoshida, K.; Tani, H.; Hirano, Y.; Koike, S.; Sasabayashi, D.; Katayama, H.; et al. Dopaminergic dysfunction and excitatory/inhibitory imbalance in treatment-resistant schizophrenia and novel neuromodulatory treatment. Mol. Psychiatry 2022, 27, 2950–2967. [Google Scholar] [CrossRef]
  16. 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. Schizophrenia 2020, 6, 1. [Google Scholar] [CrossRef]
  17. Billard, J.-M. D-Amino acids in brain neurotransmission and synaptic plasticity. Amino Acids 2012, 43, 1851–1860. [Google Scholar] [CrossRef] [PubMed]
  18. Yamamori, H.; Hashimoto, R.; Fujita, Y.; Numata, S.; Yasuda, Y.; Fujimoto, M.; Ohi, K.; Umeda-Yano, S.; Ito, A.; Ohmori, T.; et al. Changes in plasma d-serine, l-serine, and glycine levels in treatment-resistant schizophrenia before and after clozapine treatment. Neurosci. Lett. 2014, 582, 93–98. [Google Scholar] [CrossRef]
  19. Hons, J.; Vasatova, M.; Čermáková, E.; Doubek, P.; Libiger, J. Different serine and glycine metabolism in patients with schizophrenia receiving clozapine. J. Psychiatr. Res. 2012, 46, 811–818. [Google Scholar] [CrossRef]
  20. Harrison, P.J. D-Amino Acid Oxidase Inhibition: A New Glutamate Twist for Clozapine Augmentation in Schizophrenia? Biol. Psychiatry 2018, 84, 396–398. [Google Scholar] [CrossRef]
  21. Leppik, L.; Kriisa, K.; Koido, K.; Koch, K.; Kajalaid, K.; Haring, L.; Vasar, E.; Zilmer, M. Profiling of Amino Acids and Their Derivatives Biogenic Amines Before and After Antipsychotic Treatment in First-Episode Psychosis. Front. Psychiatry 2018, 9, 155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Errico, F.; Nuzzo, T.; Carella, M.; Bertolino, A.; Usiello, A. The Emerging Role of Altered d-Aspartate Metabolism in Schizophrenia: New Insights from Preclinical Models and Human Studies. Front. Psychiatry 2018, 9, 559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Errico, F.; Napolitano, F.; Nisticò, R.; Usiello, A. New insights on the role of free d-aspartate in the mammalian brain. Amino Acids 2012, 43, 1861–1871. [Google Scholar] [CrossRef] [PubMed]
  24. Pollegioni, L.; Molla, G.; Sacchi, S.; Murtas, G. Human D-aspartate Oxidase: A Key Player in D-aspartate Metabolism. Front. Mol. Biosci. 2021, 8, 689719. [Google Scholar] [CrossRef]
  25. Seetharam, J.C.; Maiti, R.; Mishra, A.; Mishra, B.R. Efficacy and safety of add-on sodium benzoate, a D-amino acid oxidase inhibitor, in treatment of schizophrenia: A systematic review and meta-analysis. Asian J. Psychiatry 2021, 68, 102947. [Google Scholar] [CrossRef]
  26. Kantrowitz, J.T.; Woods, S.W.; Petkova, E.; Cornblatt, B.; Corcoran, C.; Chen, H.; Silipo, G.; Javitt, D.C. D-serine for the treatment of negative symptoms in individuals at clinical high risk of schizophrenia: A pilot, double-blind, placebo-controlled, randomised parallel group mechanistic proof-of-concept trial. Lancet Psychiatry 2015, 2, 403–412. [Google Scholar] [CrossRef]
  27. Tsai, G.E.; Yang, P.; Chang, Y.-C.; Chong, M.-Y. D-Alanine Added to Antipsychotics for the Treatment of Schizophrenia. Biol. Psychiatry 2006, 59, 230–234. [Google Scholar] [CrossRef]
  28. Keller, S.; Punzo, D.; Cuomo, M.; Affinito, O.; Coretti, L.; Sacchi, S.; Florio, E.; Lembo, F.; Carella, M.; Copetti, M.; et al. DNA methylation landscape of the genes regulating D-serine and D-aspartate metabolism in post-mortem brain from controls and subjects with schizophrenia. Sci. Rep. 2018, 8, 10163. [Google Scholar] [CrossRef] [Green Version]
  29. Abdulbagi, M.; Wang, L.; Siddig, O.; Di, B.; Li, B. D-Amino Acids and D-Amino Acid-Containing Peptides: Potential Disease Biomarkers and Therapeutic Targets? Biomolecules 2021, 11, 1716. [Google Scholar] [CrossRef]
  30. Taniguchi, K.; Sawamura, H.; Ikeda, Y.; Tsuji, A.; Kitagishi, Y.; Matsuda, S. D-Amino Acids as a Biomarker in Schizophrenia. Diseases 2022, 10, 9. [Google Scholar] [CrossRef]
  31. Marchi, M.; Galli, G.; Magarini, F.M.; Mattei, G.; Galeazzi, G.M. Sarcosine as an add-on treatment to antipsychotic medication for people with schizophrenia: A systematic review and meta-analysis of randomized controlled trials. Expert Opin. Drug Metab. Toxicol. 2021, 17, 483–493. [Google Scholar] [CrossRef] [PubMed]
  32. Strzelecki, D.; Podgórski, M.; Kałużyńska, O.; Stefańczyk, L.; Kotlicka-Antczak, M.; Gmitrowicz, A.; Grzelak, P. Adding Sarcosine to Antipsychotic Treatment in Patients with Stable Schizophrenia Changes the Concentrations of Neuronal and Glial Metabolites in the Left Dorsolateral Prefrontal Cortex. Int. J. Mol. Sci. 2015, 16, 24475–24489. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Yao, L.; Wang, Z.; Deng, D.; Yan, R.; Ju, J.; Zhou, Q. The impact of D-cycloserine and sarcosine on in vivo frontal neural activity in a schizophrenia-like model. BMC Psychiatry 2019, 19, 314. [Google Scholar] [CrossRef] [Green Version]
  34. Page, M.J.; McKenzie, J.E.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E.; et al. The PRISMA 2020 statement: An updated guideline for reporting systematic reviews. Syst. Rev. 2021, 10, 89. [Google Scholar] [CrossRef] [PubMed]
  35. Breier, A.; Su, T.-P.; Saunders, R.; Carson, R.E.; Kolachana, B.S.; de Bartolomeis, A.; Weinberger, D.R.; Weisenfeld, N.; Malhotra, A.K.; Eckelman, W.C.; et al. Schizophrenia is associated with elevated amphetamine-induced synaptic dopamine concentrations: Evidence from a novel positron emission tomography method. Proc. Natl. Acad. Sci. USA 1997, 94, 2569–2574. [Google Scholar] [CrossRef] [Green Version]
  36. Slifstein, M.; Kegeles, L.S.; Xu, X.; Thompson, J.L.; Urban, N.; Castrillon, J.; Hackett, E.; Bae, S.-A.; Laruelle, M.; Abi-Dargham, A. Striatal and extrastriatal dopamine release measured with PET and [18F] fallypride. Synapse 2010, 64, 350–362. [Google Scholar] [CrossRef] [Green Version]
  37. Laruelle, M.; Abi-Dargham, A. Dopamine as the wind of the psychotic fire: New evidence from brain imaging studies. J. Psychopharmacol. 1999, 13, 358–371. [Google Scholar] [CrossRef]
  38. Caravaggio, F.; Porco, N.; Kim, J.; Carmona, E.T.; Brown, E.; Iwata, Y.; Nakajima, S.; Gerretsen, P.; Remington, G.; Graff-Guerrero, A. Measuring Amphetamine-Induced Dopamine Release in Humans: A Comparative Meta-Analysis of [11C]-Raclopride and [11C]-(+)-PHNO Studies. Synapse 2021, 75, e22195. [Google Scholar] [CrossRef]
  39. Howes, O.D.; Bose, S.K.; Turkheimer, F.; Valli, I.; Egerton, A.; Valmaggia, L.R.; Murray, R.M.; McGuire, P. Dopamine Synthesis Capacity Before Onset of Psychosis: A Prospective [18F]-DOPA PET Imaging Study. Am. J. Psychiatry 2011, 168, 1311–1317. [Google Scholar] [CrossRef] [Green Version]
  40. De Rosa, A.; Di Maio, A.; Torretta, S.; Garofalo, M.; Giorgelli, V.; Masellis, R.; Nuzzo, T.; Errico, F.; Bertolino, A.; Subramaniam, S.; et al. Abnormal RasGRP1 Expression in the Post-Mortem Brain and Blood Serum of Schizophrenia Patients. Biomolecules 2022, 12, 328. [Google Scholar] [CrossRef]
  41. Plavén-Sigray, P.; Victorsson, P.I.; Santillo, A.; Matheson, G.J.; Lee, M.; Collste, K.; Fatouros-Bergman, H.; Sellgren, C.M.; Erhardt, S.; Agartz, I.; et al. Thalamic dopamine D2-receptor availability in schizophrenia: A study on antipsychotic-naive patients with first-episode psychosis and a meta-analysis. Mol. Psychiatry 2021, 27, 1233–1240. [Google Scholar] [CrossRef] [PubMed]
  42. Demjaha, A.; Murray, R.; McGuire, P.; Kapur, S.; Howes, O. Dopamine Synthesis Capacity in Patients with Treatment-Resistant Schizophrenia. Am. J. Psychiatry 2012, 169, 1203–1210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kim, E.; Howes, O.D.; Veronese, M.; Beck, K.; Seo, S.; Park, J.W.; Lee, J.S.; Lee, Y.-S.; Kwon, J.S. Presynaptic Dopamine Capacity in Patients with Treatment-Resistant Schizophrenia Taking Clozapine: An [18F]DOPA PET Study. Neuropsychopharmacology 2016, 42, 941–950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Egerton, A.; Murphy, A.; Donocik, J.; Anton, A.; Barker, G.J.; Collier, T.; Deakin, B.; Drake, R.; Eliasson, E.; Emsley, R.; et al. Dopamine and Glutamate in Antipsychotic-Responsive Compared with Antipsychotic-Nonresponsive Psychosis: A Multicenter Positron Emission Tomography and Magnetic Resonance Spectroscopy Study (STRATA). Schizophr. Bull. 2020, 47, 505–516. [Google Scholar] [CrossRef]
  45. Mouchlianitis, E.; Bloomfield, M.; Law, V.; Beck, K.; Selvaraj, S.; Rasquinha, N.; Waldman, A.D.; Turkheimer, F.; Egerton, A.; Stone, J.; et al. Treatment-Resistant Schizophrenia Patients Show Elevated Anterior Cingulate Cortex Glutamate Compared to Treatment-Responsive. Schizophr. Bull. 2016, 42, 744–752. [Google Scholar] [CrossRef]
  46. Tarumi, R.; Tsugawa, S.; Noda, Y.; Plitman, E.; Honda, S.; Matsushita, K.; Chavez, S.; Sawada, K.; Wada, M.; Matsui, M.; et al. Levels of glutamatergic neurometabolites in patients with severe treatment-resistant schizophrenia: A proton magnetic resonance spectroscopy study. Neuropsychopharmacology 2019, 45, 632–640. [Google Scholar] [CrossRef]
  47. Huang, L.-C.; Lin, S.-H.; Tseng, H.-H.; Chen, K.C.; Abdullah, M.; Yang, Y.K. Altered glutamate level and its association with working memory among patients with treatment-resistant schizophrenia (TRS): A proton magnetic resonance spectroscopy study. Psychol. Med. 2022, 1–8. [Google Scholar] [CrossRef]
  48. Ochi, R.; Plitman, E.; Patel, R.; Tarumi, R.; Iwata, Y.; Tsugawa, S.; Kim, J.; Honda, S.; Noda, Y.; Uchida, H.; et al. Investigating structural subdivisions of the anterior cingulate cortex in schizophrenia, with implications for treatment resistance and glutamatergic levels. J. Psychiatry Neurosci. 2022, 47, E1–E10. [Google Scholar] [CrossRef]
  49. Matrone, M.; Kotzalidis, G.D.; Romano, A.; Bozzao, A.; Cuomo, I.; Valente, F.; Gabaglio, C.; Lombardozzi, G.; Trovini, G.; Amici, E.; et al. Treatment-resistant schizophrenia: Addressing white matter integrity, intracortical glutamate levels, clinical and cognitive profiles between early- and adult-onset patients. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2021, 114, 110493. [Google Scholar] [CrossRef]
  50. Reyes-Madrigal, F.; Guma, E.; León-Ortiz, P.; Gómez-Cruz, G.; Mora-Durán, R.; Graff-Guerrero, A.; Kegeles, L.S.; Chakravarty, M.M.; de la Fuente-Sandoval, C. Striatal glutamate, subcortical structure and clinical response to first-line treatment in first-episode psychosis patients. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2021, 113, 110473. [Google Scholar] [CrossRef]
  51. Goldstein, M.E.; Anderson, V.M.; Pillai, A.; Kydd, R.R.; Russell, B.R. Glutamatergic Neurometabolites in Clozapine-Responsive and-Resistant Schizophrenia. Int. J. Neuropsychopharmacol. 2015, 18, pyu117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Duarte, J.M.N.; Xin, L. Magnetic Resonance Spectroscopy in Schizophrenia: Evidence for Glutamatergic Dysfunction and Impaired Energy Metabolism. Neurochem. Res. 2018, 44, 102–116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Hyun, J.S.; Inoue, T.; Hayashi-Takagi, A. Multi-Scale Understanding of NMDA Receptor Function in Schizophrenia. Biomolecules 2020, 10, 1172. [Google Scholar] [CrossRef]
  54. Mei, Y.-Y.; Wu, D.C.; Zhou, N. Astrocytic Regulation of Glutamate Transmission in Schizophrenia. Front. Psychiatry 2018, 9, 544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Egerton, A.; Grace, A.A.; Stone, J.; Bossong, M.G.; Sand, M.; McGuire, P. Glutamate in schizophrenia: Neurodevelopmental perspectives and drug development. Schizophr. Res. 2020, 223, 59–70. [Google Scholar] [CrossRef]
  56. Geyer, M.A.; Krebs-Thomson, K.; Braff, D.L.; Swerdlow, N.R. Pharmacological studies of prepulse inhibition models of sensorimotor gating deficits in schizophrenia: A decade in review. Psychopharmacology 2001, 156, 117–154. [Google Scholar] [CrossRef]
  57. Smith, K.R.; Kopeikina, K.J.; Fawcett-Patel, J.M.; Leaderbrand, K.; Gao, R.; Schürmann, B.; Myczek, K.; Radulovic, J.; Swanson, G.T.; Penzes, P. Psychiatric risk factor ANK3/ankyrin-G nanodomains regulate the structure and function of glutamatergic synapses. Neuron 2014, 84, 399–415. [Google Scholar] [CrossRef] [Green Version]
  58. De Bartolomeis, A.; Avagliano, C.; Vellucci, L.; D’Ambrosio, L.; Manchia, M.; D’Urso, G.; Buonaguro, E.F.; Iasevoli, F. Translating preclinical findings in clinically relevant new antipsychotic targets: Focus on the glutamatergic postsynaptic density. Implications for treatment resistant schizophrenia. Neurosci. Biobehav. Rev. 2019, 107, 795–827. [Google Scholar] [CrossRef]
  59. Dienel, S.J.; Schoonover, K.E.; Lewis, D.A. Cognitive Dysfunction and Prefrontal Cortical Circuit Alterations in Schizophrenia: Developmental Trajectories. Biol. Psychiatry 2022. [Google Scholar] [CrossRef]
  60. O’Donovan, S.M.; Sullivan, C.R.; McCullumsmith, R.E. The role of glutamate transporters in the pathophysiology of neuropsychiatric disorders. Schizophrenia 2017, 3, 32. [Google Scholar] [CrossRef]
  61. MacKay, M.B.; Kravtsenyuk, M.; Thomas, R.; Mitchell, N.D.; Dursun, S.M.; Baker, G.B. D-Serine: Potential Therapeutic Agent and/or Biomarker in Schizophrenia and Depression? Front. Psychiatry 2019, 10, 25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Abekawa, T.; Ito, K.; Koyama, T. Role of the simultaneous enhancement of NMDA and dopamine D1 receptor-mediated neurotransmission in the effects of clozapine on phencyclidine-induced acute increases in glutamate levels in the rat medial prefrontal cortex. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2006, 374, 177–193. [Google Scholar] [CrossRef]
  63. 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] [PubMed] [Green Version]
  64. Lipina, T.; Labrie, V.; Weiner, I.; Roder, J. Modulators of the glycine site on NMDA receptors, d-serine and ALX 5407, display similar beneficial effects to clozapine in mouse models of schizophrenia. Psychopharmacology 2005, 179, 54–67. [Google Scholar] [CrossRef] [PubMed]
  65. Verrall, L.; Burnet, P.; Betts, J.F.; Harrison, P.J. The neurobiology of D-amino acid oxidase and its involvement in schizophrenia. Mol. Psychiatry 2009, 15, 122–137. [Google Scholar] [CrossRef] [PubMed]
  66. Ferraris, D.; Tsukamoto, T. Recent Advances in the Discovery of D-Amino Acid Oxidase Inhibitors and Their Therapeutic Utility in Schizophrenia. Curr. Pharm. Des. 2011, 17, 103–111. [Google Scholar] [CrossRef]
  67. Seeman, P.; Lee, T. Antipsychotic drugs: Direct correlation between clinical potency and presynaptic action on dopamine neurons. Science 1975, 188, 1217–1219. [Google Scholar] [CrossRef]
  68. Creese, I.; Burt, D.R.; Snyder, S.H. Dopamine Receptor Binding Predicts Clinical and Pharmacological Potencies of Antischizophrenic Drugs. Science 1976, 192, 481–483. [Google Scholar] [CrossRef]
  69. Olney, J.W.; Newcomer, J.W.; Farber, N. NMDA receptor hypofunction model of schizophrenia. J. Psychiatr. Res. 1999, 33, 523–533. [Google Scholar] [CrossRef]
  70. Newcomer, J.W.; Farber, N.; Jevtovic-Todorovic, V.; Selke, G.; Melson, A.K.; Hershey, T.; Craft, S.; Olney, J.W. Ketamine-Induced NMDA Receptor Hypofunction as a Model of Memory Impairment and Psychosis. Neuropsychopharmacology 1999, 20, 106–118. [Google Scholar] [CrossRef]
  71. Farber, N.B.; Wozniak, D.F.; Price, M.T.; Labruyere, J.; Huss, J.; St Peter, H.; Olney, J.W. Age-specific neurotoxicity in the rat associated with NMDA receptor blockade: Potential relevance to schizophrenia? Biol. Psychiatry 1995, 38, 788–796. [Google Scholar] [CrossRef]
  72. Olney, J.W.; Farber, N.B. NMDA antagonists as neurotherapeutic drugs, psychotogens, neurotoxins, and research tools for studying schizophrenia. Neuropsychopharmacology 1995, 13, 335–345. [Google Scholar] [CrossRef]
  73. Martinez, Z.A.; Halim, N.D.; Oostwegel, J.L.; Geyer, M.A.; Swerdlow, N.R. Ontogeny of Phencyclidine and Apomorphine-Induced Startle Gating Deficits in Rats. Pharmacol. Biochem. Behav. 2000, 65, 449–457. [Google Scholar] [CrossRef]
  74. Vollenweider, F.X.; Vontobel, P.; Øye, I.; Hell, D.; Leenders, K.L. Effects of (S)-ketamine on striatal dopamine: A [11C]raclopride PET study of a model psychosis in humans. J. Psychiatr. Res. 2000, 34, 35–43. [Google Scholar] [CrossRef]
  75. Breier, A.; Adler, C.M.; Weisenfeld, N.; Su, T.-P.; Elman, I.; Picken, L.; Malhotra, A.K.; Pickar, D. Effects of NMDA antagonism on striatal dopamine release in healthy subjects: Application of a novel PET approach. Synapse 1998, 29, 142–147. [Google Scholar] [CrossRef]
  76. Smith, G.; Schloesser, R.; Brodie, J.; Dewey, S.L.; Logan, J.; Vitkun, S.A.; Simkowitz, P.; Hurley, A.; Cooper, T.; Volkow, N.D.; et al. Glutamate Modulation of Dopamine Measured in Vivo with Positron Emission Tomography (PET) and 11C-Raclopride in Normal Human Subjects. Neuropsychopharmacology 1998, 18, 18–25. [Google Scholar] [CrossRef] [Green Version]
  77. Adler, C.M.; Malhotra, A.K.; Elman, I.; Goldberg, T.; Egan, M.; Pickar, D.; Breier, A. Comparison of Ketamine-Induced Thought Disorder in Healthy Volunteers and Thought Disorder in Schizophrenia. Am. J. Psychiatry 1999, 156, 1646–1649. [Google Scholar] [CrossRef]
  78. Olney, J.W.; Farber, N.B. Glutamate receptor dysfunction and schizophrenia. Arch. Gen. Psychiatry 1995, 52, 998–1007. [Google Scholar] [CrossRef]
  79. Javitt, D.C.; Zukin, S. Recent advances in the phencyclidine model of schizophrenia. Am. J. Psychiatry 1991, 148, 1301–1308. [Google Scholar] [CrossRef]
  80. Javitt, D.C. Negative schizophrenic symptomatology and the PCP (phencyclidine) model of schizophrenia. Hillside J. Clin. Psychiatry 1987, 9, 12–35. [Google Scholar]
  81. Coyle, J.T. The Glutamatergic Dysfunction Hypothesis for Schizophrenia. Harv. Rev. Psychiatry 1996, 3, 241–253. [Google Scholar] [CrossRef] [PubMed]
  82. Fiore, G.; Iasevoli, F. Dopamine-Glutamate Interaction and Antipsychotics Mechanism of Action: Implication for New Pharmacological Strategies in Psychosis. Curr. Pharm. Des. 2005, 11, 3561–3594. [Google Scholar] [CrossRef]
  83. Buck, S.A.; Erickson-Oberg, M.Q.; Logan, R.W.; Freyberg, Z. Relevance of interactions between dopamine and glutamate neurotransmission in schizophrenia. Mol. Psychiatry 2022. [Google Scholar] [CrossRef] [PubMed]
  84. Parellada, E.; Gassó, P. Glutamate and microglia activation as a driver of dendritic apoptosis: A core pathophysiological mechanism to understand schizophrenia. Transl. Psychiatry 2021, 11, 271. [Google Scholar] [CrossRef] [PubMed]
  85. Carlsson, M.; Carlsson, A. Interactions between glutamatergic and monoaminergic systems within the basal ganglia-implications for schizophrenia and Parkinson’s disease. Trends Neurosci. 1990, 13, 272–276. [Google Scholar] [CrossRef]
  86. Nakazawa, K.; Zsiros, V.; Jiang, Z.; Nakao, K.; Kolata, S.; Zhang, S.; Belforte, J.E. GABAergic interneuron origin of schizophrenia pathophysiology. Neuropharmacology 2012, 62, 1574–1583. [Google Scholar] [CrossRef] [Green Version]
  87. Kegeles, L.S.; Abi-Dargham, A.; Zea-Ponce, Y.; Rodenhiser-Hill, J.; Mann, J.J.; Van Heertum, R.L.; Cooper, T.B.; Carlsson, A.; Laruelle, M. Modulation of amphetamine-induced striatal dopamine release by ketamine in humans: Implications for schizophrenia. Biol. Psychiatry 2000, 48, 627–640. [Google Scholar] [CrossRef]
  88. Belforte, J.E.; Zsiros, V.; Sklar, E.R.; Jiang, Z.; Yu, G.; Li, Y.; Quinlan, E.M.; Nakazawa, K. Postnatal NMDA receptor ablation in corticolimbic interneurons confers schizophrenia-like phenotypes. Nat. Neurosci. 2010, 13, 76–83. [Google Scholar] [CrossRef] [Green Version]
  89. Nikolaus, S.; Beu, M.; Wittsack, H.-J.; Müller-Lutz, A.; Antke, C.; Hautzel, H.; Mori, Y.; Mamlins, E.; Antoch, G.; Müller, H.-W. GABAergic and glutamatergic effects on nigrostriatal and mesolimbic dopamine release in the rat. Rev. Neurosci. 2020, 31, 569–588. [Google Scholar] [CrossRef]
  90. Kim, J.; Horti, A.G.; Mathews, W.B.; Pogorelov, V.; Valentine, H.; Brašić, J.R.; Holt, D.P.; Ravert, H.T.; Dannals, R.F.; Zhou, L.; et al. Quantitative Multi-modal Brain Autoradiography of Glutamatergic, Dopaminergic, Cannabinoid, and Nicotinic Receptors in Mutant Disrupted-In-Schizophrenia-1 (DISC1) Mice. Mol. Imaging Biol. 2014, 17, 355–363. [Google Scholar] [CrossRef]
  91. Kalivas, P.W. The glutamate homeostasis hypothesis of addiction. Nat. Rev. Neurosci. 2009, 10, 561–572. [Google Scholar] [CrossRef] [PubMed]
  92. Schmidt, H.D.; Pierce, R.C. Cocaine-induced neuroadaptations in glutamate transmission: Potential therapeutic targets for craving and addiction. Ann. N. Y. Acad. Sci. 2010, 1187, 35–75. [Google Scholar] [CrossRef] [PubMed]
  93. Greengard, P.; Allen, P.B.; Nairn, A.C. Beyond the Dopamine Receptor: The DARPP-32/Protein Phosphatase-1 Cascade. Neuron 1999, 23, 435–447. [Google Scholar] [CrossRef] [Green Version]
  94. Tomasetti, C.; Iasevoli, F.; Buonaguro, E.F.; De Berardis, D.; Fornaro, M.; Fiengo, A.L.C.; Martinotti, G.; Orsolini, L.; Valchera, A.; Di Giannantonio, M.; et al. Treating the Synapse in Major Psychiatric Disorders: The Role of Postsynaptic Density Network in Dopamine-Glutamate Interplay and Psychopharmacologic Drugs Molecular Actions. Int. J. Mol. Sci. 2017, 18, 135. [Google Scholar] [CrossRef]
  95. Iasevoli, F.; Tomasetti, C.; Buonaguro, E.F.; de Bartolomeis, A. The Glutamatergic Aspects of Schizophrenia Molecular Pathophysiology: Role of the Postsynaptic Density, and Implications for Treatment. Curr. Neuropharmacol. 2014, 12, 219–238. [Google Scholar] [CrossRef] [Green Version]
  96. 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]
  97. Yamauchi, T. Molecular constituents and phosphorylation-dependent regulation of the post-synaptic density. Mass Spectrom. Rev. 2002, 21, 266–286. [Google Scholar] [CrossRef]
  98. Boeckers, T.M. The postsynaptic density. Cell Tissue Res. 2006, 326, 409–422. [Google Scholar] [CrossRef]
  99. Gold, M.G. A frontier in the understanding of synaptic plasticity: Solving the structure of the postsynaptic density. BioEssays 2012, 34, 599–608. [Google Scholar] [CrossRef] [Green Version]
  100. De Bartolomeis, A.; Sarappa, C.; Buonaguro, E.F.; Marmo, F.; Eramo, A.; Tomasetti, C.; Iasevoli, F. Different effects of the NMDA receptor antagonists ketamine, MK-801, and memantine on postsynaptic density transcripts and their topography: Role of Homer signaling, and implications for novel antipsychotic and pro-cognitive targets in psychosis. Prog. Neuropsychopharmacol. Biol. Psychiatry 2013, 46, 1–12. [Google Scholar] [CrossRef]
  101. De Bartolomeis, A.; Latte, G.; Tomasetti, C.; Iasevoli, F. Glutamatergic postsynaptic density protein dysfunctions in synaptic plasticity and dendritic spines morphology: Relevance to schizophrenia and other behavioral disorders pathophysiology, and implications for novel therapeutic approaches. Mol. Neurobiol. 2014, 49, 484–511. [Google Scholar] [CrossRef] [PubMed]
  102. Iasevoli, F.; Tomasetti, C.; Ambesi-Impiombato, A.; Muscettola, G.; de Bartolomeis, A. Dopamine receptor subtypes contribution to Homer1a induction: Insights into antipsychotic molecular action. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2009, 33, 813–821. [Google Scholar] [CrossRef] [PubMed]
  103. Tomasetti, C.; Dell’Aversano, C.; Iasevoli, F.; de Bartolomeis, A. Homer splice variants modulation within cortico-subcortical regions by dopamine D2 antagonists, a partial agonist, and an indirect agonist: Implication for glutamatergic postsynaptic density in antipsychotics action. Neuroscience 2007, 150, 144–158. [Google Scholar] [CrossRef]
  104. Barone, A.; Signoriello, S.; Latte, G.; Vellucci, L.; Giordano, G.; Avagliano, C.; Buonaguro, E.F.; Marmo, F.; Tomasetti, C.; Iasevoli, F.; et al. Modulation of glutamatergic functional connectivity by a prototypical antipsychotic: Translational inference from a postsynaptic density immediate-early gene-based network analysis. Behav. Brain Res. 2021, 404, 113160. [Google Scholar] [CrossRef] [PubMed]
  105. Graybiel, A.M.; Moratalla, R.; Robertson, H.A. Amphetamine and cocaine induce drug-specific activation of the c-fos gene in striosome-matrix compartments and limbic subdivisions of the striatum. Proc. Natl. Acad. Sci. USA 1990, 87, 6912–6916. [Google Scholar] [CrossRef] [Green Version]
  106. Moratalla, R.; Robertson, H.; Graybiel, A. Dynamic regulation of NGFI-A (zif268, egr1) gene expression in the striatum. J. Neurosci. 1992, 12, 2609–2622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Nguyen, T.V.; Kosofsky, B.E.; Birnbaum, R.; Cohen, B.M.; Hyman, S.E. Differential expression of c-fos and zif268 in rat striatum after haloperidol, clozapine, and amphetamine. Proc. Natl. Acad. Sci. USA 1992, 89, 4270–4274. [Google Scholar] [CrossRef] [Green Version]
  108. Konradi, C.; Cole, R.L.; Heckers, S.; Hyman, S.E. Amphetamine regulates gene expression in rat striatum via transcription factor CREB. J. Neurosci. 1994, 14, 5623–5634. [Google Scholar] [CrossRef] [Green Version]
  109. Cole, R.L.; Konradi, C.; Douglass, J.; Hyman, S.E. Neuronal adaptation to amphetamine and dopamine: Molecular mechanisms of prodynorphin gene regulation in rat striatum. Neuron 1995, 14, 813–823. [Google Scholar] [CrossRef] [Green Version]
  110. Konradi, C.; Leveque, J.C.; Hyman, S.E. Amphetamine and dopamine-induced immediate early gene expression in striatal neurons depends on postsynaptic NMDA receptors and calcium. J. Neurosci. 1996, 16, 4231–4239. [Google Scholar] [CrossRef] [Green Version]
  111. Keefe, K.A.; Ganguly, A. Effects of NMDA receptor antagonists on D1 dopamine receptor-mediated changes in striatal immediate early gene expression: Evidence for involvement of pharmacologically distinct NMDA receptors? Dev. Neurosci. 1998, 20, 216–228. [Google Scholar] [CrossRef] [PubMed]
  112. De Bartolomeis, A.; Barone, A.; Buonaguro, E.F.; Tomasetti, C.; Vellucci, L.; Iasevoli, F. The Homer1 family of proteins at the crossroad of dopamine-glutamate signaling: An emerging molecular “Lego” in the pathophysiology of psychiatric disorders. A systematic review and translational insight. Neurosci. Biobehav. Rev. 2022, 136, 104596. [Google Scholar] [CrossRef] [PubMed]
  113. Ghasemzadeh, M.B.; Windham, L.K.; Lake, R.W.; Acker, C.J.; Kalivas, P.W. Cocaine activates Homer1 immediate early gene transcription in the mesocorticolimbic circuit: Differential regulation by dopamine and glutamate signaling. Synapse 2008, 63, 42–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Kammermeier, P.J.; Worley, P.F. Homer 1a uncouples metabotropic glutamate receptor 5 from postsynaptic effectors. Proc. Natl. Acad. Sci. USA 2007, 104, 6055–6060. [Google Scholar] [CrossRef] [Green Version]
  115. Kim, J.Y.; Zeng, W.; Kiselyov, K.; Yuan, J.P.; Dehoff, M.H.; Mikoshiba, K.; Worley, P.F.; Muallem, S. Homer 1 Mediates Store- and Inositol 1,4,5-Trisphosphate Receptor-dependent Translocation and Retrieval of TRPC3 to the Plasma Membrane. J. Biol. Chem. 2006, 281, 32540–32549. [Google Scholar] [CrossRef] [Green Version]
  116. Mao, L.; Yang, L.; Tang, Q.; Samdani, S.; Zhang, G.; Wang, J.Q. The Scaffold Protein Homer1b/c Links Metabotropic Glutamate Receptor 5 to Extracellular Signal-Regulated Protein Kinase Cascades in Neurons. J. Neurosci. 2005, 25, 2741–2752. [Google Scholar] [CrossRef]
  117. Klugmann, M.; Symes, C.W.; Leichtlein, C.B.; Klaussner, B.K.; Dunning, J.; Fong, D.; Young, D.; During, M.J. AAV-mediated hippocampal expression of short and long Homer 1 proteins differentially affect cognition and seizure activity in adult rats. Mol. Cell. Neurosci. 2005, 28, 347–360. [Google Scholar] [CrossRef]
  118. Shiraishi, Y.; Mizutani, A.; Yuasa, S.; Mikoshiba, K.; Furuichi, T. Glutamate-induced declustering of post-synaptic adaptor protein Cupidin (Homer 2/vesl-2) in cultured cerebellar granule cells. J. Neurochem. 2003, 87, 364–376. [Google Scholar] [CrossRef] [Green Version]
  119. Yuan, J.P.; Kiselyov, K.; Shin, D.M.; Chen, J.; Shcheynikov, N.; Kang, S.H.; Dehoff, M.H.; Schwarz, M.K.; Seeburg, P.H.; Muallem, S.; et al. Homer Binds TRPC Family Channels and Is Required for Gating of TRPC1 by IP3 Receptors. Cell 2003, 114, 777–789. [Google Scholar] [CrossRef] [Green Version]
  120. Hennou, S.; Kato, A.; Schneider, E.M.; Lundstrom, K.; Gahwiler, B.H.; Inokuchi, K.; Gerber, U.; Ehrengruber, M.U. Homer-1a/Vesl-1S enhances hippocampal synaptic transmission. Eur. J. Neurosci. 2003, 18, 811–819. [Google Scholar] [CrossRef]
  121. Sala, C.; Futai, K.; Yamamoto, K.; Worley, P.F.; Hayashi, Y.; Sheng, M. Inhibition of Dendritic Spine Morphogenesis and Synaptic Transmission by Activity-Inducible Protein Homer1a. J. Neurosci. 2003, 23, 6327–6337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Ango, F.; Robbe, D.; Tu, J.C.; Xiaob, B.; Worley, P.F.; Pin, J.-P.; Bockaert, J.; Fagnia, L. Homer-Dependent Cell Surface Expression of Metabotropic Glutamate Receptor Type 5 in Neurons. Mol. Cell. Neurosci. 2002, 20, 323–329. [Google Scholar] [CrossRef] [PubMed]
  123. Sergé, A.; Fourgeaud, L.; Hémar, A.; Choquet, D. Receptor Activation and Homer Differentially Control the Lateral Mobility of Metabotropic Glutamate Receptor 5 in the Neuronal Membrane. J. Neurosci. 2002, 22, 3910–3920. [Google Scholar] [CrossRef] [Green Version]
  124. Sala, C.; Piëch, V.; Wilson, N.R.; Passafaro, M.; Liu, G.; Sheng, M. Regulation of Dendritic Spine Morphology and Synaptic Function by Shank and Homer. Neuron 2001, 31, 115–130. [Google Scholar] [CrossRef] [Green Version]
  125. Ango, F.; Prézeau, L.; Muller, T.; Tu, J.C.; Xiao, B.; Worley, P.F.; Pin, J.-P.; Bockaert, J.; Fagni, L. Agonist-independent activation of metabotropic glutamate receptors by the intracellular protein Homer. Nature 2001, 411, 962–965. [Google Scholar] [CrossRef]
  126. Foa, L.; Rajan, I.; Haas, K.; Wu, G.-Y.; Brakeman, P.; Worley, P.; Cline, H. The scaffold protein, Homer1b/c, regulates axon pathfinding in the central nervous system in vivo. Nat. Neurosci. 2001, 4, 499–506. [Google Scholar] [CrossRef]
  127. Kammermeier, P.J.; Xiao, B.; Tu, J.C.; Worley, P.F.; Ikeda, S. Homer Proteins Regulate Coupling of Group I Metabotropic Glutamate Receptors to N-Type Calcium and M-Type Potassium Channels. J. Neurosci. 2000, 20, 7238–7245. [Google Scholar] [CrossRef]
  128. Ango, F.; Pin, J.-P.; Tu, J.C.; Xiao, B.; Worley, P.F.; Bockaert, J.; Fagni, L. Dendritic and Axonal Targeting of Type 5 Metabotropic Glutamate Receptor Is Regulated by Homer1 Proteins and Neuronal Excitation. J. Neurosci. 2000, 20, 8710–8716. [Google Scholar] [CrossRef] [Green Version]
  129. Ciruela, F.; Soloviev, M.M.; Chan, W.-Y.; McIlhinney, R. Homer-1c/Vesl-1L Modulates the Cell Surface Targeting of Metabotropic Glutamate Receptor Type 1α: Evidence for an Anchoring Function. Mol. Cell. Neurosci. 2000, 15, 36–50. [Google Scholar] [CrossRef] [Green Version]
  130. Tadokoro, S.; Tachibana, T.; Imanaka, T.; Nishida, W.; Sobue, K. Involvement of unique leucine-zipper motif of PSD-Zip45 (Homer 1c/vesl-1L) in group 1 metabotropic glutamate receptor clustering. Proc. Natl. Acad. Sci. USA 1999, 96, 13801–13806. [Google Scholar] [CrossRef] [Green Version]
  131. Shiraishi, Y.; Mizutani, A.; Bito, H.; Fujisawa, K.; Narumiya, S.; Mikoshiba, K.; Furuichi, T. Cupidin, an Isoform of Homer/Vesl, Interacts with the Actin Cytoskeleton and Activated Rho Family Small GTPases and Is Expressed in Developing Mouse Cerebellar Granule Cells. J. Neurosci. 1999, 19, 8389–8400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Roche, K.W.; Tu, J.C.; Petralia, R.S.; Xiao, B.; Wenthold, R.J.; Worley, P.F. Homer 1b Regulates the Trafficking of Group I Metabotropic Glutamate Receptors. J. Biol. Chem. 1999, 274, 25953–25957. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Ciruela, F.; Soloviev, M.M.; McIlhinney, R.A. Co-expression of metabotropic glutamate receptor type 1alpha with homer-1a/Vesl-1S increases the cell surface expression of the receptor. Biochem. J. 1999, 341 Pt 3, 795–803. [Google Scholar] [CrossRef] [PubMed]
  134. Tu, J.C.; Xiao, B.; Yuan, J.P.; Lanahan, A.A.; Leoffert, K.; Li, M.; Linden, D.J.; Worley, P.F. Homer Binds a Novel Proline-Rich Motif and Links Group 1 Metabotropic Glutamate Receptors with IP3 Receptors. Neuron 1998, 21, 717–726. [Google Scholar] [CrossRef] [Green Version]
  135. Xiao, B.; Tu, J.C.; Petralia, R.S.; Yuan, J.P.; Doan, A.; Breder, C.D.; Ruggiero, A.; Lanahan, A.A.; Wenthold, R.J.; Worley, P.F. Homer Regulates the Association of Group 1 Metabotropic Glutamate Receptors with Multivalent Complexes of Homer-Related, Synaptic Proteins. Neuron 1998, 21, 707–716. [Google Scholar] [CrossRef] [Green Version]
  136. Lominac, K.; Oleson, E.B.; Pava, M.; Klugmann, M.; Schwarz, M.K.; Seeburg, P.H.; During, M.J.; Worley, P.F.; Kalivas, P.W.; Szumlinski, K.K. Distinct Roles for Different Homer1 Isoforms in Behaviors and Associated Prefrontal Cortex Function. J. Neurosci. 2005, 25, 11586–11594. [Google Scholar] [CrossRef]
  137. Tappe, A.; Kuner, R. Regulation of motor performance and striatal function by synaptic scaffolding proteins of the Homer1 family. Proc. Natl. Acad. Sci. USA 2006, 103, 774–779. [Google Scholar] [CrossRef] [Green Version]
  138. Zhang, J.; Xu, T.X.; Hallett, P.J.; Watanabe, M.; Grant, S.G.; Isacson, O.; Yao, W.D. PSD-95 uncouples dopamine-glutamate interaction in the D1/PSD-95/NMDA receptor complex. J. Neurosci. 2009, 29, 2948–2960. [Google Scholar] [CrossRef]
  139. Zhang, J.; Saur, T.; Duke, A.N.; Grant, S.G.N.; Platt, D.M.; Rowlett, J.K.; Isacson, O.; Yao, W.-D. Motor Impairments, Striatal Degeneration, and Altered Dopamine-Glutamate Interplay in Mice Lacking PSD-95. J. Neurogenetics 2014, 28, 98–111. [Google Scholar] [CrossRef] [Green Version]
  140. Destreel, G.; Seutin, V.; Engel, D. Subsaturation of the N-methyl-D-aspartate receptor glycine site allows the regulation of bursting activity in juvenile rat nigral dopamine neurons. Eur. J. Neurosci. 2019, 50, 3454–3471. [Google Scholar] [CrossRef]
  141. Papouin, T.; Ladépêche, L.; Ruel, J.; Sacchi, S.; Labasque, M.; Hanini, M.; Groc, L.; Pollegioni, L.; Mothet, J.-P.; Oliet, S.H. Synaptic and Extrasynaptic NMDA Receptors Are Gated by Different Endogenous Coagonists. Cell 2012, 150, 633–646. [Google Scholar] [CrossRef] [Green Version]
  142. De Bartolomeis, A.; Errico, F.; Aceto, G.; Tomasetti, C.; Usiello, A.; Iasevoli, F. D-aspartate dysregulation in Ddo(−/−) mice modulates phencyclidine-induced gene expression changes of postsynaptic density molecules in cortex and striatum. Prog. Neuropsychopharmacol. Biol. Psychiatry 2015, 62, 35–43. [Google Scholar] [CrossRef] [PubMed]
  143. Dallérac, G.; Li, X.; Lecouflet, P.; Morisot, N.; Sacchi, S.; Asselot, R.; Pham, T.H.; Potier, B.; Watson, D.J.G.; Schmidt, S.; et al. Dopaminergic neuromodulation of prefrontal cortex activity requires the NMDA receptor coagonist d-serine. Proc. Natl. Acad. Sci. USA 2021, 118, e2023750118. [Google Scholar] [CrossRef] [PubMed]
  144. Takagi, S.; Puhl, M.D.; Anderson, T.; Balu, D.T.; Coyle, J.T. Serine Racemase Expression by Striatal Neurons. Cell. Mol. Neurobiol. 2020, 42, 279–289. [Google Scholar] [CrossRef]
  145. Collingridge, G.; Abraham, W. Glutamate receptors and synaptic plasticity: The impact of Evans and Watkins. Neuropharmacology 2021, 206, 108922. [Google Scholar] [CrossRef] [PubMed]
  146. Kleckner, N.W.; Dingledine, R. Requirement for Glycine in Activation of NMDA-Receptors Expressed in Xenopus Oocytes. Science 1988, 241, 835–837. [Google Scholar] [CrossRef]
  147. Zhang, B.; Xiong, F.; Ma, Y.; Li, B.; Mao, Y.; Zhou, Z.; Yu, H.; Li, J.; Li, C.; Fu, J.; et al. Chronic phencyclidine treatment impairs spatial working memory in rhesus monkeys. Psychopharmacology 2019, 236, 2223–2232. [Google Scholar] [CrossRef]
  148. Aniline, O.; Pitts, F.N. Phencyclidine (PCP): A Review and Perspectives. CRC Crit. Rev. Toxicol. 1982, 10, 145–177. [Google Scholar] [CrossRef]
  149. Kantrowitz, J.T.; Javitt, D.C. N-methyl-d-aspartate (NMDA) receptor dysfunction or dysregulation: The final common pathway on the road to schizophrenia? Brain Res. Bull. 2010, 83, 108–121. [Google Scholar] [CrossRef] [Green Version]
  150. De Bartolomeis, A.; Manchia, M.; Marmo, F.; Vellucci, L.; Iasevoli, F.; Barone, A. Glycine Signaling in the Framework of Dopamine-Glutamate Interaction and Postsynaptic Density. Implications for Treatment-Resistant Schizophrenia. Front. Psychiatry 2020, 11, 369. [Google Scholar] [CrossRef]
  151. Cull-Candy, S.; Brickley, S.; Farrant, M. NMDA receptor subunits: Diversity, development and disease. Curr. Opin. Neurobiol. 2001, 11, 327–335. [Google Scholar] [CrossRef]
  152. Monyer, H.; Burnashev, N.; Laurie, D.J.; Sakmann, B.; Seeburg, P.H. Developmental and regional expression in the rat brain and functional properties of four NMDA receptors. Neuron 1994, 12, 529–540. [Google Scholar] [CrossRef]
  153. Dumas, T.C. Developmental regulation of cognitive abilities: Modified composition of a molecular switch turns on associative learning. Prog. Neurobiol. 2005, 76, 189–211. [Google Scholar] [CrossRef] [PubMed]
  154. Lussier, M.; Sanz-Clemente, A.; Roche, K.W. Dynamic Regulation of N-Methyl-d-aspartate (NMDA) and α-Amino-3-hydroxy-5-methyl-4-isoxazolepropionic Acid (AMPA) Receptors by Posttranslational Modifications. J. Biol. Chem. 2015, 290, 28596–28603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Zafra, F.; Ibáñez, I.; Bartolomé-Martín, D.; Piniella, D.; Arribas-Blázquez, M.; Giménez, C. Glycine Transporters and Its Coupling with NMDA Receptors. Adv. Neurobiol. 2017, 16, 55–83. [Google Scholar] [CrossRef] [PubMed]
  156. Lüscher, C.; Malenka, R.C. NMDA receptor-dependent long-term potentiation and long-term depression (LTP/LTD). Cold Spring Harb. Perspect. Biol. 2012, 4, a005710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Lládo, J.; Calderó, J.; Ribera, J.; Tarabal, O.; Oppenheim, R.W.; Esquerda, J.E. Opposing Effects of Excitatory Amino Acids on Chick Embryo Spinal Cord Motoneurons: Excitotoxic Degeneration or Prevention of Programmed Cell Death. J. Neurosci. 1999, 19, 10803–10812. [Google Scholar] [CrossRef] [Green Version]
  158. Benarroch, E.E. NMDA receptors: Recent insights and clinical correlations. Neurology 2011, 76, 1750–1757. [Google Scholar] [CrossRef]
  159. Zhou, X.; Hollern, D.; Liao, J.; Andrechek, E.; Wang, H. NMDA receptor-mediated excitotoxicity depends on the coactivation of synaptic and extrasynaptic receptors. Cell Death Dis. 2013, 4, e560. [Google Scholar] [CrossRef] [Green Version]
  160. Yang, J.H.; Wada, A.; Yoshida, K.; Miyoshi, Y.; Sayano, T.; Esaki, K.; Kinoshita, M.O.; Tomonaga, S.; Azuma, N.; Watanabe, M.; et al. Brain-specific Phgdh Deletion Reveals a Pivotal Role for l-Serine Biosynthesis in Controlling the Level of d-Serine, an N-methyl-d-aspartate Receptor Co-agonist, in Adult Brain. J. Biol. Chem. 2010, 285, 41380–41390. [Google Scholar] [CrossRef] [Green Version]
  161. Henneberger, C.; Bard, L.; Rusakov, D.A. D-Serine: A key to synaptic plasticity? Int. J. Biochem. Cell Biol. 2012, 44, 587–590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Stevens, E.R.; Gustafson, E.C.; Miller, R.F. Glycine transport accounts for the differential role of glycine vs. d-serine at NMDA receptor coagonist sites in the salamander retina. Eur. J. Neurosci. 2010, 31, 808–816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Gustafson, E.C.; Stevens, E.R.; Wolosker, H.; Miller, R.F. Endogenous d-Serine Contributes to NMDA-Receptor–Mediated Light-Evoked Responses in the Vertebrate Retina. J. Neurophysiol. 2007, 98, 122–130. [Google Scholar] [CrossRef] [Green Version]
  164. Yang, Y.; Ge, W.; Chen, Y.; Zhang, Z.; Shen, W.; Wu, C.; Poo, M.; Duan, S. Contribution of astrocytes to hippocampal long-term potentiation through release of d-serine. Proc. Natl. Acad. Sci. USA 2003, 100, 15194–15199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Mothet, J.-P.; Parent, A.T.; Wolosker, H.; Brady, R.O., Jr.; Linden, D.J.; Ferris, C.D.; Rogawski, M.A.; Snyder, S.H. D-Serine is an endogenous ligand for the glycine site of the N-methyl-D-aspartate receptor. Proc. Natl. Acad. Sci. USA 2000, 97, 4926–4931. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Matsui, T.-A.; Sekiguchi, M.; Hashimoto, A.; Tomita, U.; Nishikawa, T.; Wada, K. Functional Comparison of d-Serine and Glycine in Rodents: The Effect on Cloned NMDA Receptors and the Extracellular Concentration. J. Neurochem. 2002, 65, 454–458. [Google Scholar] [CrossRef]
  167. Mothet, J.P.; Rouaud, E.; Sinet, P.-M.; Potier, B.; Jouvenceau, A.; Dutar, P.; Videau, C.; Epelbaum, J.; Billard, J.-M. A critical role for the glial-derived neuromodulator d-serine in the age-related deficits of cellular mechanisms of learning and memory. Aging Cell 2006, 5, 267–274. [Google Scholar] [CrossRef]
  168. Panatier, A.; Theodosis, D.T.; Mothet, J.-P.; Touquet, B.; Pollegioni, L.; Poulain, D.A.; Oliet, S.H. Glia-Derived d-Serine Controls NMDA Receptor Activity and Synaptic Memory. Cell 2006, 125, 775–784. [Google Scholar] [CrossRef]
  169. Berger, A.J.; Dieudonné, S.; Ascher, P. Glycine Uptake Governs Glycine Site Occupancy at NMDA Receptors of Excitatory Synapses. J. Neurophysiol. 1998, 80, 3336–3340. [Google Scholar] [CrossRef] [Green Version]
  170. Schell, M.J.; Brady, R.O., Jr.; Molliver, M.E.; Snyder, S.H. D-Serine as a Neuromodulator: Regional and Developmental Localizations in Rat Brain Glia Resemble NMDA Receptors. J. Neurosci. 1997, 17, 1604–1615. [Google Scholar] [CrossRef] [Green Version]
  171. Hashimoto, A.; Oka, T.; Nishikawa, T. Extracellular concentration of endogenous free d-serine in the rat brain as revealed by in vivo microdialysis. Neuroscience 1995, 66, 635–643. [Google Scholar] [CrossRef]
  172. 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 2008, 14, 719–727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Jami, S.A.; Cameron, S.; Wong, J.M.; Daly, E.R.; McAllister, A.K.; Gray, J.A. Increased excitation-inhibition balance and loss of GABAergic synapses in the serine racemase knockout model of NMDA receptor hypofunction. J. Neurophysiol. 2021, 126, 11–27. [Google Scholar] [CrossRef] [PubMed]
  174. Labrie, V.; Fukumura, R.; Rastogi, A.; Fick, L.J.; Wang, W.; Boutros, P.C.; Kennedy, J.L.; Semeralul, M.O.; Lee, F.H.; Baker, G.B.; et al. Serine racemase is associated with schizophrenia susceptibility in humans and in a mouse model. Hum. Mol. Genet. 2009, 18, 3227–3243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Matveeva, T.M.; Pisansky, M.T.; Young, A.; Miller, R.F.; Gewirtz, J.C. Sociality deficits in serine racemase knockout mice. Brain Behav. 2019, 9, e01383. [Google Scholar] [CrossRef] [Green Version]
  176. Labrie, V.; Lipina, T.; Roder, J.C. Mice with reduced NMDA receptor glycine affinity model some of the negative and cognitive symptoms of schizophrenia. Psychopharmacology 2008, 200, 217–230. [Google Scholar] [CrossRef]
  177. 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. J. Cereb. Blood Flow Metab. 2011, 165, 1543–1555. [Google Scholar] [CrossRef] [Green Version]
  178. Martina, M.; Krasteniakov, N.V.; Bergeron, R. D-Serine differently modulates NMDA receptor function in rat CA1 hippocampal pyramidal cells and interneurons. J. Physiol. 2003, 548 Pt 2, 411–423. [Google Scholar] [CrossRef]
  179. Chapman, D.E.; Keefe, K.A.; Wilcox, K.S. Evidence for Functionally Distinct Synaptic NMDA Receptors in Ventromedial Versus Dorsolateral Striatum. J. Neurophysiol. 2003, 89, 69–80. [Google Scholar] [CrossRef]
  180. Panizzutti, R.; Rausch, M.; Zurbrügg, S.; Baumann, D.; Beckmann, N.; Rudin, M. The pharmacological stimulation of NMDA receptors via co-agonist site: An fMRI study in the rat brain. Neurosci. Lett. 2005, 380, 111–115. [Google Scholar] [CrossRef]
  181. Burnet, P.; Hutchinson, L.; Von Hesling, M.; Gilbert, E.-J.; Brandon, N.; Rutter, A.; Hutson, P.; Harrison, P. Expression of D-serine and glycine transporters in the prefrontal cortex and cerebellum in schizophrenia. Schizophr. Res. 2008, 102, 283–294. [Google Scholar] [CrossRef]
  182. Hashimoto, A.; Yoshikawa, M. Effect of aminooxyacetic acid on extracellular level of d-serine in rat striatum: An in vivo microdialysis study. Eur. J. Pharmacol. 2005, 525, 91–93. [Google Scholar] [CrossRef]
  183. Bendikov, I.; Nadri, C.; Amar, S.; Panizzutti, R.; De Miranda, J.; Wolosker, H.; Agam, G. A CSF and postmortem brain study of d-serine metabolic parameters in schizophrenia. Schizophr. Res. 2007, 90, 41–51. [Google Scholar] [CrossRef] [PubMed]
  184. Hashimoto, K.; Fukushima, T.; Shimizu, E.; Komatsu, N.; Watanabe, H.; Shinoda, N.; Nakazato, M.; Kumakiri, C.; Okada, S.; Hasegawa, H.; et al. Decreased serum levels of D-serine in patients with schizophrenia: Evidence in support of the N-methyl-D-aspartate receptor hypofunction hypothesis of schizophrenia. Arch. Gen. Psychiatry 2003, 60, 572–576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Yamada, K.; Ohnishi, T.; Hashimoto, K.; Ohba, H.; Iwayama-Shigeno, Y.; Toyoshima, M.; Okuno, A.; Takao, H.; Toyota, T.; Minabe, Y.; et al. Identification of Multiple Serine Racemase (SRR) mRNA Isoforms and Genetic Analyses of SRR and DAO in Schizophrenia and d-Serine Levels. Biol. Psychiatry 2005, 57, 1493–1503. [Google Scholar] [CrossRef] [PubMed]
  186. Hons, J.; Zirko, R.; Ulrychova, M.; Cermakova, E.; Libiger, J. D-serine serum levels in patients with schizophrenia: Relation to psychopathology and comparison to healthy subjects. Neuro Endocrinol. Lett. 2008, 29, 485–492. [Google Scholar]
  187. Fuchs, S.A.; De Barse, M.M.; Scheepers, F.E.; Cahn, W.; Dorland, L.; Velden, M.G.D.S.-V.D.; Klomp, L.W.; Berger, R.; Kahn, R.S.; de Koning, T.J. Cerebrospinal fluid d-serine and glycine concentrations are unaltered and unaffected by olanzapine therapy in male schizophrenic patients. Eur. Neuropsychopharmacol. 2008, 18, 333–338. [Google Scholar] [CrossRef]
  188. Ono, K.; Shishido, Y.; Park, H.K.; Kawazoe, T.; Iwana, S.; Chung, S.P.; El-Magd, R.M.A.; Yorita, K.; Okano, M.; Watanabe, T.; et al. Potential pathophysiological role of d-amino acid oxidase in schizophrenia: Immunohistochemical and in situ hybridization study of the expression in human and rat brain. J. Neural Transm. 2009, 116, 1335–1347. [Google Scholar] [CrossRef]
  189. Ozeki, Y.; Sekine, M.; Fujii, K.; Watanabe, T.; Okayasu, H.; Takano, Y.; Shinozaki, T.; Aoki, A.; Akiyama, K.; Homma, H.; et al. Phosphoserine phosphatase activity is elevated and correlates negatively with plasma d-serine concentration in patients with schizophrenia. Psychiatry Res. 2016, 237, 344–350. [Google Scholar] [CrossRef]
  190. De Rosa, A.; Fontana, A.; Nuzzo, T.; Garofalo, M.; Di Maio, A.; Punzo, D.; Copetti, M.; Bertolino, A.; Errico, F.; Rampino, A.; et al. Machine Learning algorithm unveils glutamatergic alterations in the post-mortem schizophrenia brain. Schizophrenia 2022, 8, 8. [Google Scholar] [CrossRef]
  191. Burnet, P.W.J.; Eastwood, S.L.; Bristow, G.C.; Godlewska, B.R.; Sikka, P.; Walker, M.; Harrison, P.J. D-Amino acid oxidase activity and expression are increased in schizophrenia. Mol. Psychiatry 2008, 13, 658–660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Habl, G.; Zink, M.; Petroianu, G.; Bauer, M.; Schneider-Axmann, T.; von Wilmsdorff, M.; Falkai, P.; Henn, F.A.; Schmitt, A. Increased d-amino acid oxidase expression in the bilateral hippocampal CA4 of schizophrenic patients: A post-mortem study. J. Neural Transm. 2009, 116, 1657–1665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Madeira, C.; Freitas, M.E.; Vargas-Lopes, C.; Wolosker, H.; Panizzutti, R. Increased brain d-amino acid oxidase (DAAO) activity in schizophrenia. Schizophr. Res. 2008, 101, 76–83. [Google Scholar] [CrossRef] [PubMed]
  194. Verrall, L.; Walker, M.; Rawlings, N.; Benzel, I.; Kew, J.N.C.; Harrison, P.J.; Burnet, P.W.J. D-Amino acid oxidase and serine racemase in human brain: Normal distribution and altered expression in schizophrenia. Eur. J. Neurosci. 2007, 26, 1657–1669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Schell, M.J.; Molliver, M.E.; Snyder, S.H. D-serine, an endogenous synaptic modulator: Localization to astrocytes and glutamate-stimulated release. Proc. Natl. Acad. Sci. USA 1995, 92, 3948–3952. [Google Scholar] [CrossRef] [Green Version]
  196. Foltyn, V.N.; Bendikov, I.; De Miranda, J.; Panizzutti, R.; Dumin, E.; Shleper, M.; Li, P.; Toney, M.D.; Kartvelishvily, E.; Wolosker, H. Serine racemase modulates intracellular D-serine levels through an α,β-elimination activity. J. Biol. Chem. 2005, 280, 1754–1763. [Google Scholar] [CrossRef] [Green Version]
  197. Hikida, T.; Mustafa, A.K.; Maeda, K.; Fujii, K.; Barrow, R.K.; Saleh, M.; Huganir, R.L.; Snyder, S.H.; Hashimoto, K.; Sawa, A. Modulation of d-Serine Levels in Brains of Mice Lacking PICK. Biol. Psychiatry 2008, 63, 997–1000. [Google Scholar] [CrossRef] [Green Version]
  198. Fujii, K.; Maeda, K.; Hikida, T.; Mustafa, A.K.; Balkissoon, R.; Xia, J.; Yamada, T.; Ozeki, Y.; Kawahara, R.; Okawa, M.; et al. Serine racemase binds to PICK1: Potential relevance to schizophrenia. Mol. Psychiatry 2005, 11, 150–157. [Google Scholar] [CrossRef]
  199. Wang, X.; He, G.; Gu, N.; Yang, J.; Tang, J.; Chen, Q.; Liu, X.; Shen, Y.; Qian, X.; Lin, W.; et al. Association of G72/G30 with schizophrenia in the Chinese population. Biochem. Biophys Res. Commun. 2004, 319, 1281–1286. [Google Scholar] [CrossRef]
  200. Schumacher, J.; Jamra, R.A.; Freudenberg, J.; Becker, T.; Ohlraun, S.; Otte, A.C.J.; Tullius, M.; Kovalenko, S.; Bogaert, A.V.D.; Maier, W.; et al. Examination of G72 and D-amino-acid oxidase as genetic risk factors for schizophrenia and bipolar affective disorder. Mol. Psychiatry 2004, 9, 203–207. [Google Scholar] [CrossRef]
  201. Addington, A.M.; Gornick, M.; Sporn, A.L.; Gogtay, N.; Greenstein, D.; Lenane, M.; Gochman, P.; Baker, N.; Balkissoon, R.; Vakkalanka, R.K.; et al. Polymorphisms in the 13q33.2 gene G72/G30 are associated with childhood-onset schizophrenia and psychosis not otherwise specified. Biol. Psychiatry 2004, 55, 976–980. [Google Scholar] [CrossRef]
  202. Hattori, E.; Liu, C.; Badner, J.A.; Bonner, T.I.; Christian, S.L.; Maheshwari, M.; Detera-Wadleigh, S.D.; Gibbs, R.A.; Gershon, E.S. Polymorphisms at the G72/G30 Gene Locus, on 13q33, Are Associated with Bipolar Disorder in Two Independent Pedigree Series*. Am. J. Hum. Genet. 2003, 72, 1131–1140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Chumakov, I.; Blumenfeld, M.; Guerassimenko, O.; Cavarec, L.; Palicio, M.; Abderrahim, H.; Bougueleret, L.; Barry, C.; Tanaka, H.; La Rosa, P.; et al. Genetic and physiological data implicating the new human gene G72 and the gene for d-amino acid oxidase in schizophrenia. Proc. Natl. Acad. Sci. USA 2002, 99, 13675–13680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Chen, Y.S.; Akula, N.; Detera-Wadleigh, S.D.; Schulze, T.G.; Thomas, J.; Potash, J.B.; DePaulo, J.R.; McInnis, M.G.; Cox, N.J.; McMahon, F.J. Findings in an independent sample support an association between bipolar affective disorder and the G72/G30 locus on chromosome 13q. Mol. Psychiatry 2004, 9, 87–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Balu, D.T.; Coyle, J.T. Neuroplasticity signaling pathways linked to the pathophysiology of schizophrenia. Neurosci. Biobehav. Rev. 2011, 35, 848–870. [Google Scholar] [CrossRef] [Green Version]
  206. Ma, T.M.; Abazyan, S.; Abazyan, B.; Nomura, J.; Yang, C.; Seshadri, S.J.; Sawa, A.; Snyder, S.H.; Pletnikov, M.V. Pathogenic disruption of DISC1-serine racemase binding elicits schizophrenia-like behavior via D-serine depletion. Mol. Psychiatry 2012, 18, 557–567. [Google Scholar] [CrossRef]
  207. Svane, K.C.; Asis, E.-K.; Omelchenko, A.; Kunnath, A.J.; Brzustowicz, L.M.; Silverstein, S.M.; Firestein, B.L. D-Serine administration affects nitric oxide synthase 1 adaptor protein and DISC1 expression in sex-specific manner. Mol. Cell. Neurosci. 2018, 89, 20–32. [Google Scholar] [CrossRef]
  208. Brzustowicz, L.M.; Simone, J.; Mohseni, P.; Hayter, J.E.; Hodgkinson, K.A.; Chow, E.W.; Bassett, A.S. Linkage Disequilibrium Mapping of Schizophrenia Susceptibility to the CAPON Region of Chromosome 1q. Am. J. Hum. Genet. 2004, 74, 1057–1063. [Google Scholar] [CrossRef] [Green Version]
  209. Carrel, D.; Du, Y.; Komlos, D.; Hadzimichalis, N.M.; Kwon, M.; Wang, B.; Brzustowicz, L.M.; Firestein, B.L. NOS1AP Regulates Dendrite Patterning of Hippocampal Neurons through a Carboxypeptidase E-Mediated Pathway. J. Neurosci. 2009, 29, 8248–8258. [Google Scholar] [CrossRef]
  210. Li, L.-L.; De Mera, R.M.M.-F.; Chen, J.; Ba, W.; Kasri, N.N.; Zhang, M.; Courtney, M.J.; De Mera, R.M.M.F. Unexpected Heterodivalent Recruitment of NOS1AP to nNOS Reveals Multiple Sites for Pharmacological Intervention in Neuronal Disease Models. J. Neurosci. 2015, 35, 7349–7364. [Google Scholar] [CrossRef] [Green Version]
  211. Hadzimichalis, N.M.; Previtera, M.L.; Moreau, M.P.; Li, B.; Lee, G.H.; Dulencin, A.M.; Matteson, P.G.; Buyske, S.; Millonig, J.H.; Brzustowicz, L.M.; et al. NOS1AP protein levels are altered in BA46 and cerebellum of patients with schizophrenia. Schizophr. Res. 2010, 124, 248–250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Eastwood, S.L. Does the CAPON gene confer susceptibility to schizophrenia? PLoS Med. 2005, 2, e348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Hagiwara, H.; Iyo, M.; Hashimoto, K. Neonatal Disruption of Serine Racemase Causes Schizophrenia-Like Behavioral Abnormalities in Adulthood: Clinical Rescue by D-Serine. PLoS ONE 2013, 8, e62438. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Sakurai, S.-I.; Ishii, S.; Umino, A.; Shimazu, D.; Yamamoto, N.; Nishikawa, T. Effects of psychotomimetic and antipsychotic agents on neocortical and striatal concentrations of various amino acids in the rat. J. Neurochem. 2004, 90, 1378–1388. [Google Scholar] [CrossRef]
  215. Panizzutti, R.; Fisher, M.; Garrett, C.; Man, W.H.; Sena, W.; Madeira, C.; Vinogradov, S. Association between increased serum d-serine and cognitive gains induced by intensive cognitive training in schizophrenia. Schizophr. Res. 2018, 207, 63–69. [Google Scholar] [CrossRef]
  216. Ohnuma, T.; Sakai, Y.; Maeshima, H.; Hatano, T.; Hanzawa, R.; Abe, S.; Kida, S.; Shibata, N.; Suzuki, T.; Arai, H. Changes in plasma glycine, l-serine, and d-serine levels in patients with schizophrenia as their clinical symptoms improve: Results from the Juntendo University Schizophrenia Projects (JUSP). Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2008, 32, 1905–1912. [Google Scholar] [CrossRef]
  217. Hons, J.; Zirko, R.; Vasatova, M.; Doubek, P.; Klimova, B.; Masopust, J.; Valis, M.; Kuca, K. Impairment of Executive Functions Associated With Lower D-Serine Serum Levels in Patients With Schizophrenia. Front. Psychiatry 2021, 12, 514579. [Google Scholar] [CrossRef]
  218. Vardigan, J.D.; Huszar, S.L.; McNaughton, C.H.; Hutson, P.H.; Uslaner, J.M. MK-801 produces a deficit in sucrose preference that is reversed by clozapine, D-serine, and the metabotropic glutamate 5 receptor positive allosteric modulator CDPPB: Relevance to negative symptoms associated with schizophrenia? Pharmacol. Biochem. Behav. 2010, 95, 223–229. [Google Scholar] [CrossRef]
  219. Karasawa, J.-I.; Hashimoto, K.; Chaki, S. D-Serine and a glycine transporter inhibitor improve MK-801-induced cognitive deficits in a novel object recognition test in rats. Behav. Brain Res. 2008, 186, 78–83. [Google Scholar] [CrossRef]
  220. Bado, P.; Madeira, C.; Vargas-Lopes, C.; Moulin, T.C.; Wasilewska-Sampaio, A.P.; Maretti, L.; de Oliveira, R.V.; Amaral, O.B.; Panizzutti, R. Effects of low-dose d-serine on recognition and working memory in mice. Psychopharmacology 2011, 218, 461–470. [Google Scholar] [CrossRef]
  221. Dsouza, D.C.; Radhakrishnan, R.; Perry, E.; Bhakta, S.G.; Singh, N.M.; Yadav, R.; Abi-Saab, D.; Pittman, B.; Chaturvedi, S.K.; Sharma, M.P.; et al. Feasibility, Safety, and Efficacy of the Combination of D-Serine and Computerized Cognitive Retraining in Schizophrenia: An International Collaborative Pilot Study. Neuropsychopharmacology 2012, 38, 492–503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. 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] [PubMed]
  223. Kantrowitz, J.T.; Epstein, M.; Lee, M.; Lehrfeld, N.; Nolan, K.A.; Shope, C.; Petkova, E.; Silipo, G.; Javitt, D.C. Improvement in mismatch negativity generation during d-serine treatment in schizophrenia: Correlation with symptoms. Schizophr. Res. 2018, 191, 70–79. [Google Scholar] [CrossRef] [PubMed]
  224. 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] [PubMed] [Green Version]
  225. Lane, H.Y.; Lin, C.H.; Huang, Y.J.; Liao, C.H.; Chang, Y.C.; Tsai, G.E. A randomized, double-blind, placebo-controlled comparison study of sarcosine (N-methylglycine) and D-serine add-on treatment for schizophrenia. Int. J. Neuropsychopharmacol. 2010, 13, 451–460. [Google Scholar] [CrossRef] [Green Version]
  226. Tsai, G.; Yang, P.; Chung, L.-C.; Lange, N.; Coyle, J.T. D-serine added to antipsychotics for the treatment of schizophrenia. Biol. Psychiatry 1998, 44, 1081–1089. [Google Scholar] [CrossRef]
  227. 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]
  228. 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]
  229. Goh, K.K.; Wu, T.H.; Chen, C.H.; Lu, M.L. Efficacy of N-methyl-D-aspartate receptor modulator augmentation in schizophrenia: A meta-analysis of randomised, placebo-controlled trials. J. Psychopharmacol. 2021, 35, 236–252. [Google Scholar] [CrossRef]
  230. 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] [Green Version]
  231. D-Serine AudRem: R33 Phase. Available online: https://clinicaltrials.gov/ct2/show/NCT05046353?term=NCT05046353&draw=2&rank=1 (accessed on 14 June 2022).
  232. Chung, S.P.; Sogabe, K.; Park, H.K.; Song, Y.; Ono, K.; El-Magd, R.A.; Shishido, Y.; Yorita, K.; Sakai, T.; Fukui, K. Potential cytotoxic effect of hydroxypyruvate produced from D-serine by astroglial D-amino acid oxidase. J. Biochem. 2010, 148, 743–753. [Google Scholar] [CrossRef] [PubMed]
  233. Park, H.K.; Shishido, Y.; Ichise-Shishido, S.; Kawazoe, T.; Ono, K.; Iwana, S.; Tomita, Y.; Yorita, K.; Sakai, T.; Fukui, K. Potential Role for Astroglial d-Amino Acid Oxidase in Extracellular d-Serine Metabolism and Cytotoxicity. J. Biochem. 2006, 139, 295–304. [Google Scholar] [CrossRef] [PubMed]
  234. Carone, F.A.; Nakamura, S.; Goldman, B. Urinary loss of glucose, phosphate, and protein by diffusion into proximal straight tubules injured by D-serine and maleic acid. Lab. Investig. 1985, 52, 605–610. [Google Scholar]
  235. Adage, T.; Trillat, A.-C.; Quattropani, A.; Perrin, D.; Cavarec, L.; Shaw, J.; Guerassimenko, O.; Giachetti, C.; Gréco, B.; Chumakov, I.; et al. In vitro and in vivo pharmacological profile of AS057278, a selective d-amino acid oxidase inhibitor with potential anti-psychotic properties. Eur. Neuropsychopharmacol. 2008, 18, 200–214. [Google Scholar] [CrossRef]
  236. Hashimoto, K.; Fujita, Y.; Horio, M.; Kunitachi, S.; Iyo, M.; Ferraris, D.; Tsukamoto, T. Co-Administration of a D-Amino Acid Oxidase Inhibitor Potentiates the Efficacy of D-Serine in Attenuating Prepulse Inhibition Deficits After Administration of Dizocilpine. Biol. Psychiatry 2009, 65, 1103–1106. [Google Scholar] [CrossRef] [PubMed]
  237. Smith, S.M.; Uslaner, J.M.; Yao, L.; Mullins, C.M.; Surles, N.O.; Huszar, S.L.; McNaughton, C.H.; Pascarella, D.M.; Kandebo, M.; Hinchliffe, R.M.; et al. The behavioral and neurochemical effects of a novel D-amino acid oxidase inhibitor compound 8 [4H-thieno [3,2-b]pyrrole-5-carboxylic acid] and D-serine. J. Pharmacol. Exp. Ther. 2009, 328, 921–930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Duplantier, A.J.; Becker, S.L.; Bohanon, M.J.; Borzilleri, K.A.; Chrunyk, B.A.; Downs, J.T.; Hu, L.-Y.; El-Kattan, A.; James, L.C.; Liu, S.; et al. Discovery, SAR, and Pharmacokinetics of a Novel 3-Hydroxyquinolin-2(1H)-one Series of Potent d-Amino Acid Oxidase (DAAO) Inhibitors. J. Med. Chem. 2009, 52, 3576–3585. [Google Scholar] [CrossRef]
  239. The Effect of D-Serine as Add-on Therapy in Recent-Onset Psychosis. Available online: https://clinicaltrials.gov/ct2/show/NCT04140773?term=NCT04140773&draw=2&rank=1 (accessed on 14 June 2022).
  240. Fisher, G.H.; D’Aniello, A.; Vetere, A.; Padula, L.; Cusano, G.P.; Man, E.H. Free D-aspartate and D-alanine in normal and Alzheimer brain. Brain Res. Bull. 1991, 26, 983–985. [Google Scholar] [CrossRef]
  241. Nagata, Y.; Kubota, K. A trial to determine D-amino acids in tissue proteins of mice. Amino Acids 1993, 4, 121–125. [Google Scholar] [CrossRef]
  242. Nagata, Y.; Konno, R.; Niwa, A. Amino acid levels in d-alanine-administered mutant mice lacking d-amino acid oxidase. Metabolism 1994, 43, 1153–1157. [Google Scholar] [CrossRef]
  243. Morikawa, A.; Hamase, K.; Zaitsu, K. Determination of d-alanine in the rat central nervous system and periphery using column-switching high-performance liquid chromatography. Anal. Biochem. 2003, 312, 66–72. [Google Scholar] [CrossRef]
  244. Sasabe, J.; Suzuki, M. Emerging Role of D-Amino Acid Metabolism in the Innate Defense. Front. Microbiol. 2018, 9, 933. [Google Scholar] [CrossRef] [Green Version]
  245. Saitoh, Y.; Katane, M.; Miyamoto, T.; Sekine, M.; Sakai-Kato, K.; Homma, H. d-Serine and d-Alanine Regulate Adaptive Foraging Behavior in Caenorhabditis elegans via the NMDA Receptor. J. Neurosci. 2020, 40, 7531–7544. [Google Scholar] [CrossRef] [PubMed]
  246. McBain, C.J.; Kleckner, N.W.; Wyrick, S.; Dingledine, R. Structural requirements for activation of the glycine coagonist site of N-methyl-D-aspartate receptors expressed in Xenopus oocytes. Mol. Pharmacol. 1989, 36, 556–565. [Google Scholar] [PubMed]
  247. Chairoungdua, A.; Kanai, Y.; Matsuo, H.; Inatomi, J.; Kim, D.K.; Endou, H. Identification and Characterization of a Novel Member of the Heterodimeric Amino Acid Transporter Family Presumed to be Associated with an Unknown Heavy Chain. J. Biol. Chem. 2001, 276, 49390–49399. [Google Scholar] [CrossRef] [Green Version]
  248. Rojas, C.; Alt, J.; Ator, N.A.; Wilmoth, H.; Rais, R.; Hin, N.; DeVivo, M.; Popiolek, M.; Tsukamoto, T.; Slusher, B.S. Oral administration of D-alanine in monkeys robustly increases plasma and cerebrospinal fluid levels but experimental D-amino acid oxidase inhibitors had minimal effect. J. Psychopharmacol. 2016, 30, 887–895. [Google Scholar] [CrossRef] [PubMed]
  249. Popiolek, M.; Tierney, B.; Steyn, S.J.; De Vivo, M. Lack of Effect of Sodium Benzoate at Reported Clinical Therapeutic Concentration on d-Alanine Metabolism in Dogs. ACS Chem. Neurosci. 2018, 9, 2832–2837. [Google Scholar] [CrossRef]
  250. Tanii, Y.; Nishikawa, T.; Hashimoto, A.; Takahashi, K. Stereoselective antagonism by enantiomers of alanine and serine of phencyclidine-induced hyperactivity, stereotypy and ataxia in the rat. J. Pharmacol. Exp. Ther. 1994, 269, 1040–1048. [Google Scholar]
  251. Hashimoto, A.; Nishikawa, T.; Oka, T.; Takahashi, K. D-alanine inhibits methamphetamine-induced hyperactivity in rats. Eur. J. Pharmacol. 1991, 202, 105–107. [Google Scholar]
  252. Umino, A.; Takahashi, K.; Nishikawa, T. Characterization of the phencyclidine-induced increase in prefrontal cortical dopamine metabolism in the rat. J. Cereb. Blood Flow Metab. 1998, 124, 377–385. [Google Scholar] [CrossRef] [Green Version]
  253. Hatano, T.; Ohnuma, T.; Sakai, Y.; Shibata, N.; Maeshima, H.; Hanzawa, R.; Suzuki, T.; Arai, H. Plasma alanine levels increase in patients with schizophrenia as their clinical symptoms improve—Results from the Juntendo University Schizophrenia Projects (JUSP). Psychiatry Res. 2010, 177, 27–31. [Google Scholar] [CrossRef] [PubMed]
  254. Wolosker, H.; Blackshaw, S.; Snyder, S.H. Serine racemase: A glial enzyme synthesizing d-serine to regulate glutamate-N-methyl-d-aspartate neurotransmission. Proc. Natl. Acad. Sci. USA 1999, 96, 13409–13414. [Google Scholar] [CrossRef] [Green Version]
  255. D’aniello, A.; Di Fiore, M.M.; Fisher, G.H.; Milone, A.; Seleni, A.; D’Aniello, S.; Perna, A.; Ingrosso, D. Occurrence of D-aspartic acid and N-methyl-D-aspartic acid in rat neuroendocrine tissues and their role in the modulation of luteinizing hormone and growth hormone release. FASEB J. 2000, 14, 699–714. [Google Scholar] [CrossRef] [PubMed]
  256. Hashimoto, A.; Kumashiro, S.; Nishikawa, T.; Oka, T.; Takahashi, K.; Mito, T.; Takashima, S.; Doi, N.; Mizutani, Y.; Yamazaki, T.; et al. Embryonic Development and Postnatal Changes in Free d-Aspartate and d-Serine in the Human Prefrontal Cortex. J. Neurochem. 1993, 61, 348–351. [Google Scholar] [CrossRef] [PubMed]
  257. Topo, E.; Soricelli, A.; Di Maio, A.; D’Aniello, E.; Di Fiore, M.M.; D’Aniello, A. Evidence for the involvement of d-aspartic acid in learning and memory of rat. Amino Acids 2009, 38, 1561–1569. [Google Scholar] [CrossRef]
  258. Spinelli, P.; Brown, E.R.; Ferrandino, G.; Branno, M.; Montarolo, P.G.; D’Aniello, E.; Rastogi, R.K.; D’Aniello, B.; Baccari, G.C.; Fisher, G.; et al. D-aspartic acid in the nervous system ofAplysia limacina: Possible role in neurotransmission. J. Cell. Physiol. 2005, 206, 672–681. [Google Scholar] [CrossRef]
  259. D’Aniello, S.; Somorjai, I.; Garcia-Fernàndez, J.; Topo, E.; D’Aniello, A. D-Aspartic acid is a novel endogenous neurotransmitter. FASEB J. 2010, 25, 1014–1027. [Google Scholar] [CrossRef] [Green Version]
  260. Nakatsuka, S.; Hayashi, M.; Muroyama, A.; Otsuka, M.; Kozaki, S.; Yamada, H.; Moriyama, Y. d-Aspartate Is Stored in Secretory Granules and Released through a Ca2+-dependent Pathway in a Subset of Rat Pheochromocytoma PC12 Cells. J. Biol. Chem. 2001, 276, 26589–26596. [Google Scholar] [CrossRef] [Green Version]
  261. Palacín, M.; Estevez, R.; Bertran, J.; Zorzano, A. Molecular Biology of Mammalian Plasma Membrane Amino Acid Transporters. Physiol. Rev. 1998, 78, 969–1054. [Google Scholar] [CrossRef]
  262. Hashimoto, A.; Oka, T.; Nishikawa, T. Anatomical Distribution and Postnatal Changes in Endogenous Free D-Aspartate and D-Serine in Rat Brain and Periphery. Eur. J. Neurosci. 1995, 7, 1657–1663. [Google Scholar] [CrossRef]
  263. Schell, M.J.; Cooper, O.B.; Snyder, S.H. D-aspartate localizations imply neuronal and neuroendocrine roles. Proc. Natl. Acad. Sci. USA 1997, 94, 2013–2018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Errico, F.; Cuomo, M.; Canu, N.; Caputo, V.; Usiello, A. New insights on the influence of free d-aspartate metabolism in the mammalian brain during prenatal and postnatal life. Biochim. Biophys. Acta (BBA) Proteins Proteom. 2020, 1868, 140471. [Google Scholar] [CrossRef] [PubMed]
  265. De Rosa, A.; Mastrostefano, F.; Di Maio, A.; Nuzzo, T.; Saitoh, Y.; Katane, M.; Isidori, A.M.; Caputo, V.; Marotta, P.; Falco, G.; et al. Prenatal expression of d-aspartate oxidase causes early cerebral d-aspartate depletion and influences brain morphology and cognitive functions at adulthood. Amino Acids 2020, 52, 597–617. [Google Scholar] [CrossRef]
  266. Olverman, H.; Jones, A.; Mewett, K.; Watkins, J. Structure/activity relations of N-methyl-d-aspartate receptor ligands as studied by their inhibition of [3H]d2-amino-5-phosphonopentanoic acid binding in rat brain membranes. Neuroscience 1988, 26, 17–31. [Google Scholar] [CrossRef]
  267. Krashia, P.; Ledonne, A.; Nobili, A.; Cordella, A.; Errico, F.; Usiello, A.; D’Amelio, M.; Mercuri, N.B.; Guatteo, E.; Carunchio, I. Persistent elevation of D-Aspartate enhances NMDA receptor-mediated responses in mouse substantia nigra pars compacta dopamine neurons. Neuropharmacology 2016, 103, 69–78. [Google Scholar] [CrossRef]
  268. Van Veldhoven, P.P.; Brees, C.; Mannaerts, G.P. D-Aspartate oxidase, a peroxisomal enzyme in liver of rat and man. Biochim. Biophys. Acta (BBA) Gen. Subj. 1991, 1073, 203–208. [Google Scholar] [CrossRef]
  269. Katane, M.; Kawata, T.; Nakayama, K.; Saitoh, Y.; Kaneko, Y.; Matsuda, S.; Saitoh, Y.; Miyamoto, T.; Sekine, M.; Homma, H. Characterization of the Enzymatic and Structural Properties of Human D-Aspartate Oxidase and Comparison with Those of the Rat and Mouse Enzymes. Biol. Pharm. Bull. 2015, 38, 298–305. [Google Scholar] [CrossRef] [Green Version]
  270. Negri, A.; Ceciliani, F.; Tedeschi, G.; Simonic, T.; Ronchi, S. The primary structure of the flavoprotein D-aspartate oxidase from beef kidney. J. Biol. Chem. 1992, 267, 11865–11871. [Google Scholar] [CrossRef]
  271. Molla, G.; Sacchi, S.; Bernasconi, M.; Pilone, M.S.; Fukui, K.; Polegioni, L. Characterization of human D-amino acid oxidase. FEBS Lett. 2006, 580, 2358–2364. [Google Scholar] [CrossRef] [Green Version]
  272. Molla, G.; Chaves-Sanjuan, A.; Savinelli, A.; Nardini, M.; Pollegioni, L. Structure and kinetic properties of human d-aspartate oxidase, the enzyme-controlling d-aspartate levels in brain. FASEB J. 2020, 34, 1182–1197. [Google Scholar] [CrossRef] [Green Version]
  273. Errico, F.; Nisticò, R.; Palma, G.; Federici, M.; Affuso, A.; Brilli, E.; Topo, E.; Centonze, D.; Bernardi, G.; Bozzi, Y.; et al. Increased levels of d-aspartate in the hippocampus enhance LTP but do not facilitate cognitive flexibility. Mol. Cell. Neurosci. 2008, 37, 236–246. [Google Scholar] [CrossRef] [PubMed]
  274. Errico, F.; Nisticò, R.; Di Giorgio, A.; Squillace, M.; Vitucci, D.; Galbusera, A.; Piccinin, S.; Mango, D.; Fazio, L.; Middei, S.; et al. Free D-aspartate regulates neuronal dendritic morphology, synaptic plasticity, gray matter volume and brain activity in mammals. Transl. Psychiatry 2014, 4, e417. [Google Scholar] [CrossRef] [PubMed]
  275. Molinaro, G.; Pietracupa, S.; Di Menna, L.; Pescatori, L.; Usiello, A.; Battaglia, G.; Nicoletti, F.; Bruno, V. D-Aspartate activates mGlu receptors coupled to polyphosphoinositide hydrolysis in neonate rat brain slices. Neurosci. Lett. 2010, 478, 128–130. [Google Scholar] [CrossRef] [PubMed]
  276. Cristino, L.; Luongo, L.; Squillace, M.; Paolone, G.; Mango, D.; Piccinin, S.; Zianni, E.; Imperatore, R.; Iannotta, M.; Longo, F.; et al. d-Aspartate oxidase influences glutamatergic system homeostasis in mammalian brain. Neurobiol. Aging 2015, 36, 1890–1902. [Google Scholar] [CrossRef]
  277. Gong, X.-Q.; Frandsen, A.; Lu, W.-Y.; Wan, Y.; Zabek, R.L.; Pickering, D.S.; Bai, D. D-Aspartate and NMDA, but not L-aspartate, block AMPA receptors in rat hippocampal neurons. J. Cereb. Blood Flow Metab. 2005, 145, 449–459. [Google Scholar] [CrossRef] [Green Version]
  278. Errico, F.; Napolitano, F.; Squillace, M.; Vitucci, D.; Blasi, G.; de Bartolomeis, A.; Bertolino, A.; D’Aniello, A.; Usiello, A. Decreased levels of d-aspartate and NMDA in the prefrontal cortex and striatum of patients with schizophrenia. J. Psychiatr. Res. 2013, 47, 1432–1437. [Google Scholar] [CrossRef]
  279. Errico, F.; Mothet, J.-P.; Usiello, A. D-Aspartate: An endogenous NMDA receptor agonist enriched in the developing brain with potential involvement in schizophrenia. J. Pharm. Biomed. Anal. 2015, 116, 7–17. [Google Scholar] [CrossRef]
  280. Nuzzo, T.; Sacchi, S.; Errico, F.; Keller, S.; Palumbo, O.; Florio, E.; Punzo, D.; Napolitano, F.; Copetti, M.; Carella, M.; et al. Decreased free d-aspartate levels are linked to enhanced d-aspartate oxidase activity in the dorsolateral prefrontal cortex of schizophrenia patients. Schizophrenia 2017, 3, 16. [Google Scholar] [CrossRef]
  281. Errico, F.; Rossi, S.; Napolitano, F.; Catuogno, V.; Topo, E.; Fisone, G.; D’Aniello, A.; Centonze, D.; Usiello, A. D-Aspartate Prevents Corticostriatal Long-Term Depression and Attenuates Schizophrenia-Like Symptoms Induced by Amphetamine and MK. J. Neurosci. 2008, 28, 10404–10414. [Google Scholar] [CrossRef] [Green Version]
  282. Sacchi, S.; De Novellis, V.; Paolone, G.; Nuzzo, T.; Iannotta, M.; Belardo, C.; Squillace, M.; Bolognesi, P.; Rosini, E.; Motta, Z.; et al. Olanzapine, but not clozapine, increases glutamate release in the prefrontal cortex of freely moving mice by inhibiting D-aspartate oxidase activity. Sci. Rep. 2017, 7, 46288. [Google Scholar] [CrossRef] [Green Version]
  283. Errico, F.; Nisticò, R.; Napolitano, F.; Oliva, A.B.; Romano, R.; Barbieri, F.; Florio, T.; Russo, C.; Mercuri, N.B.; Usiello, A. Persistent increase of d-aspartate in d-aspartate oxidase mutant mice induces a precocious hippocampal age-dependent synaptic plasticity and spatial memory decay. Neurobiol. Aging 2009, 32, 2061–2074. [Google Scholar] [CrossRef] [PubMed]
  284. Hardingham, G.E.; Bading, H. The Yin and Yang of NMDA receptor signalling. Trends Neurosci. 2003, 26, 81–89. [Google Scholar] [CrossRef]
  285. Melville, G.W.; Siegler, J.C.; Marshall, P.W.M. The effects of d-aspartic acid supplementation in resistance-trained men over a three month training period: A randomised controlled trial. PLoS ONE 2017, 12, e0182630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Melville, G.W.; Siegler, J.C.; Marshall, P.W. Three and six grams supplementation of d-aspartic acid in resistance trained men. J. Int. Soc. Sports Nutr. 2015, 12, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Pollegioni, L.; Diederichs, K.; Molla, G.; Umhau, S.; Welte, W.; Ghisla, S.; Pilone, M.S. Yeast d-Amino Acid Oxidase: Structural Basis of its Catalytic Properties. J. Mol. Biol. 2002, 324, 535–546. [Google Scholar] [CrossRef] [Green Version]
  288. Pollegioni, L.; Piubelli, L.; Sacchi, S.; Pilone, M.S.; Molla, G. Physiological functions of D-amino acid oxidases: From yeast to humans. Cell. Mol. Life Sci. 2007, 64, 1373–1394. [Google Scholar] [CrossRef] [PubMed]
  289. Konno, R.; Sasaki, M.; Asakura, S.; Fukui, K.; Enami, J.; Niwa, A. D-Amino-acid oxidase is not present in the mouse liver. Biochim. Biophys. Acta (BBA) Gen. Subj. 1997, 1335, 173–181. [Google Scholar] [CrossRef]
  290. Arnold, G.; Liscum, L.; Holtzman, E. Ultrastructural localization of D-amino acid oxidase in microperoxisomes of the rat nervous system. J. Histochem. Cytochem. 1979, 27, 735–745. [Google Scholar] [CrossRef] [Green Version]
  291. Urai, Y.; Jinnouchi, O.; Kwak, K.T.; Suzue, A.; Nagahiro, S.; Fukui, K. Gene expression of D-amino acid oxidase in cultured rat astrocytes: Regional and cell type specific expression. Neurosci. Lett. 2002, 324, 101–104. [Google Scholar] [CrossRef]
  292. Horiike, K.; Tojo, H.; Arai, R.; Yamano, T.; Nozaki, M.; Maeda, T. Localization of D-amino acid oxidase in Bergmann glial cells and astrocytes of rat cerebellum. Brain Res. Bull. 1987, 19, 587–596. [Google Scholar] [CrossRef]
  293. Jagannath, V.; Marinova, Z.; Monoranu, C.-M.; Walitza, S.; Grünblatt, E. Expression of D-Amino Acid Oxidase (DAO/DAAO) and D-Amino Acid Oxidase Activator (DAOA/G72) during Development and Aging in the Human Post-mortem Brain. Front. Neuroanat. 2017, 11, 31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Kapoor, R.; Lim, K.S.; Cheng, A.; Garrick, T.; Kapoor, V. Preliminary evidence for a link between schizophrenia and NMDA-glycine site receptor ligand metabolic enzymes, d-amino acid oxidase (DAAO) and kynurenine aminotransferase-1 (KAT-1). Brain Res 2006, 1106, 205–210. [Google Scholar] [CrossRef] [PubMed]
  295. Pritchett, D.; Hasan, S.; Tam, S.K.; Engle, S.J.; Brandon, N.J.; Sharp, T.; Foster, R.G.; Harrison, P.J.; Bannerman, D.M.; Peirson, S.N. d-amino acid oxidase knockout (Dao(−/−)) mice show enhanced short-term memory performance and heightened anxiety, but no sleep or circadian rhythm disruption. Eur. J. Neurosci. 2015, 41, 1167–1179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Labrie, V.; Wang, W.; Barger, S.W.; Baker, G.B.; Roder, J.C. Genetic loss of D-amino acid oxidase activity reverses schizophrenia-like phenotypes in mice. Genes Brain Behav. 2010, 9, 11–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  297. Liu, A.; Chen, X.; Wang, Z.J.; Xu, Q.; Appel-Cresswell, S.; McKeown, M.J. A Genetically Informed, Group fMRI Connectivity Modeling Approach: Application to Schizophrenia. IEEE Trans. Biomed. Eng. 2013, 61, 946–956. [Google Scholar] [CrossRef]
  298. Papagni, S.A.; Mechelli, A.; Prata, D.P.; Kambeitz, J.; Fu, C.H.; Picchioni, M.; Walshe, M.; Toulopoulou, T.; Bramon, E.; Murray, R.M.; et al. Differential effects of DAAO on regional activation and functional connectivity in schizophrenia, bipolar disorder and controls. NeuroImage 2011, 56, 2283–2291. [Google Scholar] [CrossRef]
  299. Shishikura, M.; Hakariya, H.; Iwasa, S.; Yoshio, T.; Ichiba, H.; Yorita, K.; Fukui, K.; Fukushima, T. Evaluation of human D-amino acid oxidase inhibition by anti-psychotic drugs in vitro. Biosci. Trends 2014, 8, 149–154. [Google Scholar] [CrossRef] [Green Version]
  300. Iwasa, S.; Tabara, H.; Song, Z.; Nakabayashi, M.; Yokoyama, Y.; Fukushima, T. Inhibition of D-Amino Acid Oxidase Activity by Antipsychotic Drugs Evaluated by a Fluorometric Assay Using D-Kynurenine as Substrate. Yakugaku Zasshi 2011, 131, 1111–1116. [Google Scholar] [CrossRef] [Green Version]
  301. Abou El-Magd, R.M.; Park, H.K.; Kawazoe, T.; Iwana, S.; Ono, K.; Chung, S.P.; Miyano, M.; Yorita, K.; Sakai, T.; Fukui, K. The effect of risperidone on D-amino acid oxidase activity as a hypothesis for a novel mechanism of action in the treatment of schizophrenia. J. Psychopharmacol. 2010, 24, 1055–1067. [Google Scholar] [CrossRef]
  302. Karoum, F.; Freed, W.J.; Chuang, L.-W.; Cannon-Spoor, E.; Wyatt, R.J.; Costa, E. D-DOPA and l-DOPA similarly elevate brain dopamine and produce turning behavior in rats. Brain Res. 1988, 440, 190–194. [Google Scholar] [CrossRef]
  303. Shindo, H.; Maeda, T. Studies on the Metabolism of D-and L-Isomers of 3, 4-Dihydroxyphenylalanine (DOPA). VI. Metabolism of D-DOPA in Rat Kidney. Chem. Pharm. Bull. 1974, 22, 1721–1731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Wu, M.; Zhou, X.-J.; Konno, R.; Wang, Y.-X. D-dopa is unidirectionally converted to l-dopa by d-amino acid oxidase, followed by dopa transaminase. Clin. Exp. Pharmacol. Physiol. 2006, 33, 1042–1046. [Google Scholar] [CrossRef] [PubMed]
  305. Betts, J.F.; Schweimer, J.; Burnham, K.E.; Burnet, P.; Sharp, T.; Harrison, P.J. D-amino acid oxidase is expressed in the ventral tegmental area and modulates cortical dopamine. Front. Synaptic Neurosci. 2014, 6, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Schweimer, J.V.; Coullon, G.S.; Betts, J.F.; Burnet, P.W.; Engle, S.J.; Brandon, N.J.; Harrison, P.J.; Sharp, T. Increased burst-firing of ventral tegmental area dopaminergic neurons in D-amino acid oxidase knockout mice in vivo. Eur. J. Neurosci. 2014, 40, 2999–3009. [Google Scholar] [CrossRef] [PubMed]
  307. Lin, C.-H.; Yang, H.-T.; Chiu, C.-C.; Lane, H.-Y. Blood levels of D-amino acid oxidase vs. D-amino acids in reflecting cognitive aging. Sci. Rep. 2017, 7, 14849. [Google Scholar] [CrossRef] [PubMed]
  308. Chen, Y.-C.; Chou, W.-H.; Tsou, H.-H.; Fang, C.-P.; Liu, T.-H.; Tsao, H.-H.; Hsu, W.-C.; Weng, Y.-C.; Wang, Y.; Liu, Y.-L. A Post-hoc Study of D-Amino Acid Oxidase in Blood as an Indicator of Post-stroke Dementia. Front. Neurol. 2019, 10, 402. [Google Scholar] [CrossRef]
  309. Yang, J.; Yang, J.; Liang, S.H.; Xu, Y.; Moore, A.; Ran, C. Imaging hydrogen peroxide in Alzheimer’s disease via cascade signal amplification. Sci. Rep. 2016, 6, 35613. [Google Scholar] [CrossRef] [Green Version]
  310. Ferraris, D.; Duvall, B.; Ko, Y.-S.; Thomas, A.G.; Rojas, C.; Majer, P.; Hashimoto, K.; Tsukamoto, T. Synthesis and Biological Evaluation of d-Amino Acid Oxidase Inhibitors. J. Med. Chem. 2008, 51, 3357–3359. [Google Scholar] [CrossRef]
  311. Lin, C.H.; Lin, C.H.; Chang, Y.C.; Huang, Y.J.; Chen, P.W.; Yang, H.T.; Lane, H.Y. Sodium Benzoate, a D-Amino Acid Oxidase Inhibitor, Added to Clozapine for the Treatment of Schizophrenia: A Randomized, Double-Blind, Placebo-Controlled Trial. Biol. Psychiatry 2018, 84, 422–432. [Google Scholar] [CrossRef]
  312. Lin, C.Y.; Liang, S.Y.; Chang, Y.C.; Ting, S.Y.; Kao, C.L.; Wu, Y.H.; Tsai, G.E.; Lane, H.Y. Adjunctive sarcosine plus benzoate improved cognitive function in chronic schizophrenia patients with constant clinical symptoms: A randomised, double-blind, placebo-controlled trial. World J. Biol. Psychiatry 2017, 18, 357–368. [Google Scholar] [CrossRef]
  313. Matsuura, A.; Fujita, Y.; Iyo, M.; Hashimoto, K. Effects of sodium benzoate on pre-pulse inhibition deficits and hyperlocomotion in mice after administration of phencyclidine. Acta Neuropsychiatr. 2015, 27, 159–167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  314. Tang, H.; Jensen, K.; Houang, E.; McRobb, F.M.; Bhat, S.; Svensson, M.; Bochevarov, A.; Day, T.; Dahlgren, M.K.; Bell, J.A.; et al. Discovery of a Novel Class of d-Amino Acid Oxidase Inhibitors Using the Schrödinger Computational Platform. J. Med. Chem. 2022, 65, 6775–6802. [Google Scholar] [CrossRef] [PubMed]
  315. A Study to Evaluate Efficacy, Safety, Tolerability, and Pharmacokinetics of 3 Dose Levels of TAK-831 in Adjunctive Treatment of Adult Participants with Negative Symptoms of Schizophrenia. Available online: https://clinicaltrials.gov/ct2/show/NCT03382639?id=NCT03382639&draw=2&rank=1&load=cart (accessed on 14 June 2022).
  316. CLINICAL TRIALS SND12. Available online: http://syneurxtrials.com/Trials/Refractory-Schizophrenia (accessed on 14 June 2022).
  317. 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] [PubMed]
  318. Scott, J.G.; Baker, A.; Lim, C.C.W.; Foley, S.; Dark, F.; Gordon, A.; Ward, D.; Richardson, D.; Bruxner, G.; Beckmann, K.M.; et al. Effect of Sodium Benzoate vs. Placebo among Individuals with Early Psychosis: A Randomized Clinical Trial. JAMA Netw. Open. 2020, 3, e2024335. [Google Scholar] [CrossRef]
  319. Naas, E.; Zilles, K.; Gnahn, H.; Betz, H.; Becker, C.-M.; Schröder, H. Glycine receptor immunoreactivity in rat and human cerebral cortex. Brain Res. 1991, 561, 139–146. [Google Scholar] [CrossRef]
  320. Gielen, M.C.; Thomas, P.; Smart, T.G. The desensitization gate of inhibitory Cys-loop receptors. Nat. Commun. 2015, 6, 6829. [Google Scholar] [CrossRef] [Green Version]
  321. Harvey, R.J.; Vandenberg, R.J. Glycine Transporters and Receptors as Targets for Analgesics. Biomolecules 2021, 11, 1676. [Google Scholar] [CrossRef] [PubMed]
  322. Kuo, A.; Corradini, L.; Nicholson, J.; Smith, M. Assessment of the Anti-Allodynic and Anti-Hyperalgesic Efficacy of a Glycine Transporter 2 Inhibitor Relative to Pregabalin, Duloxetine and Indomethacin in a Rat Model of Cisplatin-Induced Peripheral Neuropathy. Biomolecules 2021, 11, 940. [Google Scholar] [CrossRef]
  323. Cioffi, C. Inhibition of Glycine Re-Uptake: A Potential Approach for Treating Pain by Augmenting Glycine-Mediated Spinal Neurotransmission and Blunting Central Nociceptive Signaling. Biomolecules 2021, 11, 864. [Google Scholar] [CrossRef]
  324. Zeilhofer, H.U.; Werynska, K.; Gingras, J.; Yévenes, G.E. Glycine Receptors in Spinal Nociceptive Control—An Update. Biomolecules 2021, 11, 846. [Google Scholar] [CrossRef]
  325. Barsch, L.; Werdehausen, R.; Leffler, A.; Eulenburg, V. Modulation of Glycinergic Neurotransmission may Contribute to the Analgesic Effects of Propacetamol. Biomolecules 2021, 11, 493. [Google Scholar] [CrossRef] [PubMed]
  326. Heresco-Levy, U.; Ermilov, M.; Lichtenberg, P.; Bar, G.; Javitt, D.C. High-dose glycine added to olanzapine and risperidone for the treatment of schizophrenia. Biol. Psychiatry 2004, 55, 165–171. [Google Scholar] [CrossRef]
  327. Pinard, E.; Borroni, E.; Koerner, A.; Umbricht, D.; Alberati, D. Glycine Transporter Type I (GlyT1) Inhibitor, Bitopertin: A Journey from Lab to Patient. CHIMIA 2018, 72, 477–484. [Google Scholar] [CrossRef]
  328. Kim, S.-Y.; Kaufman, M.J.; Cohen, B.M.; Jensen, J.E.; Coyle, J.T.; Du, F.; Öngür, D. In Vivo Brain Glycine and Glutamate Concentrations in Patients with First-Episode Psychosis Measured by Echo Time–Averaged Proton Magnetic Resonance Spectroscopy at 4T. Biol. Psychiatry 2018, 83, 484–491. [Google Scholar] [CrossRef]
  329. Cioffi, C.L.; Guzzo, P.R. Inhibitors of Glycine Transporter-1: Potential Therapeutics for the Treatment of CNS Disorders. Curr. Top. Med. Chem. 2016, 16, 3404–3437. [Google Scholar] [CrossRef]
  330. Tsai, G.; Ralph-Williams, R.J.; Martina, M.; Bergeron, R.; Berger-Sweeney, J.; Dunham, K.S.; Jiang, Z.; Caine, S.B.; Coyle, J.T. Gene knockout of glycine transporter 1: Characterization of the behavioral phenotype. Proc. Natl. Acad. Sci. USA 2004, 101, 8485–8490. [Google Scholar] [CrossRef] [Green Version]
  331. Hons, J.; Zirko, R.; Ulrychova, M.; Cermakova, E.; Doubek, P.; Libiger, J. Glycine serum level in schizophrenia: Relation to negative symptoms. Psychiatry Res. 2010, 176, 103–108. [Google Scholar] [CrossRef]
  332. Neeman, G.; Blanaru, M.; Bloch, B.; Kremer, I.; Ermilov, M.; Javitt, D.C.; Heresco-Levy, U. Relation of Plasma Glycine, Serine, and Homocysteine Levels to Schizophrenia Symptoms and Medication Type. Am. J. Psychiatry 2005, 162, 1738–1740. [Google Scholar] [CrossRef]
  333. Heresco-Levy, U.; Bar, G.; Levin, R.; Ermilov, M.; Ebstein, R.P.; Javitt, D.C. High glycine levels are associated with prepulse inhibition deficits in chronic schizophrenia patients. Schizophr. Res. 2007, 91, 14–21. [Google Scholar] [CrossRef] [PubMed]
  334. Serrita, J.; Ralevski, E.; Yoon, G.; Petrakis, I. A Pilot Randomized, Placebo-Controlled Trial of Glycine for Treatment of Schizophrenia and Alcohol Dependence. J. Dual Diagn. 2019, 15, 46–55. [Google Scholar] [CrossRef] [PubMed]
  335. Potkin, S.G.; Jin, Y.; Bunney, B.G.; Costa, J.; Gulasekaram, B. Effect of Clozapine and Adjunctive High-Dose Glycine in Treatment-Resistant Schizophrenia. Am. J. Psychiatry 1999, 156, 145–147. [Google Scholar] [CrossRef] [PubMed]
  336. Javitt, D.C.; Zylberman, I.; Zukin, S.R.; Heresco-Levy, U.; Lindenmayer, J.P. Amelioration of negative symptoms in schizophrenia by glycine. Am. J. Psychiatry 1994, 151, 1234–1236. [Google Scholar] [CrossRef]
  337. Javitt, D.C.; Silipo, G.; Cienfuegos, A.; Shelley, A.-M.; Bark, N.; Park, M.; Lindenmayer, J.-P.; Suckow, R.; Zukin, S.R. Adjunctive high-dose glycine in the treatment of schizophrenia. Int. J. Neuropsychopharmacol. 2001, 4, 385–391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  338. Heresco-Levy, U.; Javitt, D.C.; Ermilov, M.; Mordel, C.; Horowitz, A.; Kelly, D. Double-blind, placebo-controlled, crossover trial of glycine adjuvant therapy for treatment-resistant schizophrenia. Br. J. Psychiatry 1996, 169, 610–617. [Google Scholar] [CrossRef] [PubMed]
  339. Heresco-Levy, U.; Javitt, D.C.; Ermilov, M.; Mordel, C.; Silipo, G.; Lichtenstein, M. Efficacy of High-Dose Glycine in the Treatment of Enduring Negative Symptoms of Schizophrenia. Arch. Gen. Psychiatry 1999, 56, 29–36. [Google Scholar] [CrossRef] [PubMed]
  340. Greenwood, L.-M.; Leung, S.; Michie, P.T.; Green, A.; Nathan, P.J.; Fitzgerald, P.; Johnston, P.; Solowij, N.; Kulkarni, J.; Croft, R.J. The effects of glycine on auditory mismatch negativity in schizophrenia. Schizophr. Res. 2017, 191, 61–69. [Google Scholar] [CrossRef] [PubMed]
  341. Evins, A.E. Placebo-Controlled Trial of Glycine Added to Clozapine in Schizophrenia. Am. J. Psychiatry 2000, 157, 826–828. [Google Scholar] [CrossRef]
  342. Diaz, P.; Bhaskara, S.; Dursun, S.M.; Deakin, B. Double-blind, Placebo-Controlled, Crossover Trial of Clozapine Plus Glycine in Refractory Schizophrenia Negative Results. J. Clin. Psychopharmacol. 2005, 25, 277–278. [Google Scholar] [CrossRef]
  343. Buchanan, R.W.; Javitt, D.C.; Marder, S.R.; Schooler, N.R.; Gold, J.M.; McMahon, R.P.; Heresco-Levy, U.; Carpenter, W.T. The Cognitive and Negative Symptoms in Schizophrenia Trial (CONSIST): The Efficacy of Glutamatergic Agents for Negative Symptoms and Cognitive Impairments. Am. J. Psychiatry 2007, 164, 1593–1602. [Google Scholar] [CrossRef] [Green Version]
  344. Costa, J.; Khaled, E.; Sramek, J.; Bunney, W.; Potkin, S.G. An Open Trial of Glycine as an Adjunct to Neuroleptics in Chronic Treatment-Refractory Schizophrenics. J. Clin. Psychopharmacol. 1990, 10, 71–72. [Google Scholar] [CrossRef]
  345. Rosse, R.B.; Theut, S.K.; Banay-Schwartz, M.; Leighton, M.; Scarcella, E.; Cohen, C.G.; Deutsch, S.I. Glycine adjuvant therapy to conventional neuroleptic treatment in schizophrenia: An open-label, pilot study. Clin. Neuropharmacol. 1989, 12, 416–424. [Google Scholar] [CrossRef] [PubMed]
  346. Leiderman, E.; Zylberman, I.; Zukin, S.; Cooper, T.B.; Javitt, D.C. Preliminary investigation of high-dose oral glycine on serum levels and negative symptoms in schizophrenia: An open-label trial. Biol. Psychiatry 1996, 39, 213–215. [Google Scholar] [CrossRef]
  347. Lane, H.-Y.; Huang, C.-L.; Wu, P.-L.; Liu, Y.-C.; Chang, Y.-C.; Lin, P.-Y.; Chen, P.-W.; Tsai, G. Glycine Transporter I Inhibitor, N-methylglycine (Sarcosine), Added to Clozapine for the Treatment of Schizophrenia. Biol. Psychiatry 2006, 60, 645–649. [Google Scholar] [CrossRef] [PubMed]
  348. Dravid, S.M.; Burger, P.B.; Prakash, A.; Geballe, M.T.; Yadav, R.; Le, P.; Vellano, K.; Snyder, J.P.; Traynelis, S.F. Structural determinants of D-cycloserine efficacy at the NR1/NR2C NMDA receptors. J. Neurosci. 2010, 30, 2741–2754. [Google Scholar] [CrossRef] [PubMed]
  349. Heresco-Levy, U. N-Methyl-D-aspartate (NMDA) receptor-based treatment approaches in schizophrenia: The first decade. Int. J. Neuropsychopharmacol. 2000, 3, 243–258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  350. Tsai, G.; Lane, H.-Y.; Yang, P.; Chong, M.-Y.; Lange, N. Glycine transporter I inhibitor, N-Methylglycine (sarcosine), added to antipsychotics for the treatment of schizophrenia. Biol. Psychiatry 2004, 55, 452–456. [Google Scholar] [CrossRef]
  351. 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] [PubMed]
  352. Strzelecki, D.; Urban-Kowalczyk, M.; Wysokiński, A. Serum levels of interleukin 6 in schizophrenic patients during treatment augmentation with sarcosine (results of the PULSAR study). Hum. Psychopharmacol. Clin. Exp. 2018, 33, e2652. [Google Scholar] [CrossRef]
  353. Strzelecki, D.; Podgórski, M.; Kałużyńska, O.; Gawlik-Kotelnicka, O.; Stefańczyk, L.; Kotlicka-Antczak, M.; Gmitrowicz, A.; Grzelak, P. Supplementation of Antipsychotic Treatment with the Amino Acid Sarcosine Influences Proton Magnetic Resonance Spectroscopy Parameters in Left Frontal White Matter in Patients with Schizophrenia. Nutrients 2015, 7, 8767–8782. [Google Scholar] [CrossRef] [Green Version]
  354. Strzelecki, D.; Podgórski, M.; Kałużyńska, O.; Gawlik-Kotelnicka, O.; Stefańczyk, L.; Kotlicka-Antczak, M.; Gmitrowicz, A.; Grzelak, P. Supplementation of antipsychotic treatment with sarcosine—GlyT1 inhibitor—Causes changes of glutamatergic 1NMR spectroscopy parameters in the left hippocampus in patients with stable schizophrenia. Neurosci. Lett. 2015, 606, 7–12. [Google Scholar] [CrossRef]
  355. Umbricht, D.; Alberati, D.; Martin-Facklam, M.; Borroni, E.; Youssef, E.A.; Ostland, M.; Wallace, T.L.; Knoflach, F.; Dorflinger, E.; Wettstein, J.G.; et al. Effect of bitopertin, a glycine reuptake inhibitor, on negative symptoms of schizophrenia: A randomized, double-blind, proof-of-concept study. JAMA Psychiatry 2014, 71, 637–646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  356. Bugarski-Kirola, D.; Iwata, N.; Sameljak, S.; Reid, C.; Blaettler, T.; Millar, L.; Marques, T.R.; Garibaldi, G.; Kapur, S. Efficacy and safety of adjunctive bitopertin versus placebo in patients with suboptimally controlled symptoms of schizophrenia treated with antipsychotics: Results from three phase 3, randomised, double-blind, parallel-group, placebo-controlled, multicentre studies in the SearchLyte clinical trial programme. Lancet Psychiatry 2016, 3, 1115–1128. [Google Scholar] [CrossRef]
  357. Kantrowitz, J.T.; Nolan, K.A.; Epstein, M.; Lehrfeld, N.; Shope, C.; Petkova, E.; Javitt, D.C. Neurophysiological Effects of Bitopertin in Schizophrenia. J. Clin. Psychopharmacol. 2017, 37, 447–451. [Google Scholar] [CrossRef]
  358. Fleischhacker, W.W.; Podhorna, J.; Gröschl, 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]
  359. Cain, C.K.; McCue, M.; Bello, I.; Creedon, T.; Tang, D.-I.; Laska, E.; Goff, D.C. D-Cycloserine augmentation of cognitive remediation in schizophrenia. Schizophr. Res. 2014, 153, 177–183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  360. Duncan, E.J.; Szilagyi, S.; Schwartz, M.P.; Bugarski-Kirola, D.; Kunzova, A.; Negi, S.; Stephanides, M.; Efferen, T.R.; Angrist, B.; Peselow, E.; et al. Effects of d-cycloserine on negative symptoms in schizophrenia. Schizophr. Res. 2004, 71, 239–248. [Google Scholar] [CrossRef] [PubMed]
  361. Goff, D.C.; Henderson, D.C.; Evins, A.; Amico, E. A placebo-controlled crossover trial of d-cycloserine added to clozapine in patients with schizophrenia. Biol. Psychiatry 1999, 45, 512–514. [Google Scholar] [CrossRef]
  362. 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] [Green Version]
  363. 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. Psychopharmacologia 2004, 179, 144–150. [Google Scholar] [CrossRef]
  364. Goff, D.C.; Cather, C.; Gottlieb, J.D.; Evins, A.E.; Walsh, J.; Raeke, L.; Otto, M.W.; Schoenfeld, D.; Green, M.F. Once-weekly d-cycloserine effects on negative symptoms and cognition in schizophrenia: An exploratory study. Schizophr. Res. 2008, 106, 320–327. [Google Scholar] [CrossRef] [Green Version]
  365. Heresco-Levy, U.; Ermilov, M.; Shimoni, J.; Shapira, B.; Silipo, G.; Javitt, D.C. Placebo-controlled trial of D-cycloserine added to conventional neuroleptics, olanzapine, or risperidone in schizophrenia. Am. J. Psychiatry 2002, 159, 480–482. [Google Scholar] [CrossRef]
  366. Rosse, R.B.; Fay-Mccarthy, M.; Kendrick, K.; Davis, E.R.; Deutsch, S.I. D-Cycloserine Adjuvant Therapy to Molindone in the Treatment of Schizophrenia. Clin. Neuropharmacol. 1996, 19, 444–450. [Google Scholar] [CrossRef] [PubMed]
  367. Takiguchi, K.; Uezato, A.; Itasaka, M.; Atsuta, H.; Narushima, K.; Yamamoto, N.; Kurumaji, A.; Tomita, M.; Oshima, K.; Shoda, K.; et al. Association of schizophrenia onset age and white matter integrity with treatment effect of D-cycloserine: A randomized placebo-controlled double-blind crossover study. BMC Psychiatry 2017, 17, 249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  368. Zafra, F.; Aragon, C.; Olivares, L.; Danbolt, N.; Gimenez, C.; Storm-Mathisen, J. Glycine transporters are differentially expressed among CNS cells. J. Neurosci. 1995, 15, 3952–3969. [Google Scholar] [CrossRef] [PubMed]
  369. Cubelos, B.; Gonzalez-Gonzalez, I.M.; Gimenez, C.; Zafra, F. The scaffolding protein PSD-95 interacts with the glycine transporter GLYT1 and impairs its internalization. J. Neurochem. 2005, 95, 1047–1058. [Google Scholar] [CrossRef] [PubMed]
  370. Betz, H.; Gomeza, J.; Armsen, W.; Scholze, P.; Eulenburg, V. Glycine transporters: Essential regulators of synaptic transmission. Biochem. Soc. Trans. 2006, 34 Pt 1, 55–58. [Google Scholar] [CrossRef] [PubMed]
  371. Martina, M.; Gorfinkel, Y.; Halman, S.; Lowe, J.A.; Periyalwar, P.; Schmidt, C.J.; Bergeron, R. Glycine transporter type 1 blockade changes NMDA receptor-mediated responses and LTP in hippocampal CA1 pyramidal cells by altering extracellular glycine levels. J. Physiol. 2004, 557, 489–500. [Google Scholar] [CrossRef]
  372. Vargas-Medrano, J.; Castrejon-Tellez, V.; Plenge, F.; Ramirez, I.; Miranda, M. PKCβ-dependent phosphorylation of the glycine transporter. Neurochem. Int. 2011, 59, 1123–1132. [Google Scholar] [CrossRef] [Green Version]
  373. Gomeza, J.; Hülsmann, S.; Ohno, K.; Eulenburg, V.; Szöke, K.; Richter, D.; Betz, H. Inactivation of the Glycine Transporter 1 Gene Discloses Vital Role of Glial Glycine Uptake in Glycinergic Inhibition. Neuron 2003, 40, 785–796. [Google Scholar] [CrossRef] [Green Version]
  374. Kurolap, A.; Hershkovitz, T.; Baris, H.N.; Adam, M.P.; Mirzaa, G.M.; Pagon, R.A.; Wallace, S.E.; Bean, L.J.H.; Gripp, K.W.; Amemiya, A. (Eds.) GLYT1 Encephalopathy, in GeneReviews(®®); University of Washington: Seattle, WA, USA, 1993. [Google Scholar]
  375. Gabernet, L.; Pauly-Evers, M.; Schwerdel, C.; Lentz, M.; Bluethmann, H.; Vogt, K.; Alberati, D.; Mohler, H.; Boison, D. Enhancement of the NMDA receptor function by reduction of glycine transporter-1 expression. Neurosci. Lett. 2004, 373, 79–84. [Google Scholar] [CrossRef]
  376. Bergeron, R.; Meyer, T.M.; Coyle, J.T.; Greene, R.W. Modulation of N-methyl-D-aspartate receptor function by glycine transport. Proc. Natl. Acad. Sci. USA 1998, 95, 15730–15734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  377. Alberati, D.; Moreau, J.-L.; Lengyel, J.; Hauser, N.; Mory, R.; Borroni, E.; Pinard, E.; Knoflach, F.; Schlotterbeck, G.; Hainzl, D.; et al. Glycine reuptake inhibitor RG1678: A pharmacologic characterization of an investigational agent for the treatment of schizophrenia. Neuropharmacology 2012, 62, 1152–1161. [Google Scholar] [CrossRef] [PubMed]
  378. Kinney, G.G.; Sur, C.; Burno, M.; Mallorga, P.J.; Williams, J.B.; Figueroa, D.J.; Wittmann, M.; Lemaire, W.; Conn, P.J. The glycine transporter type 1 inhibitor N-[3-(4′-fluorophenyl)-3-(4′-phenylphenoxy)propyl]sarcosine potentiates NMDA receptor-mediated responses in vivo and produces an antipsychotic profile in rodent behavior. J. Neurosci. 2003, 23, 7586–7591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  379. Depoortère, R.; Dargazanli, G.; Estenne-Bouhtou, G.; Coste, A.; Lanneau, C.; Desvignes, C.; Poncelet, M.; Héaulme, M.; Santucci, V.; Decobert, M.; et al. Neurochemical, Electrophysiological and Pharmacological Profiles of the Selective Inhibitor of the Glycine Transporter-1 SSR504734, a Potential New Type of Antipsychotic. Neuropsychopharmacology 2005, 30, 1963–1985. [Google Scholar] [CrossRef]
  380. Fone, K.C.; Watson, D.J.; Billiras, R.I.; Sicard, D.I.; Dekeyne, A.; Rivet, J.-M.; Gobert, A.; Millan, M.J. Comparative Pro-cognitive and Neurochemical Profiles of Glycine Modulatory Site Agonists and Glycine Reuptake Inhibitors in the Rat: Potential Relevance to Cognitive Dysfunction and Its Management. Mol. Neurobiol. 2020, 57, 2144–2166. [Google Scholar] [CrossRef] [Green Version]
  381. Zhang, H.X.; Hyrc, K.; Thio, L.L. The glycine transport inhibitor sarcosine is an NMDA receptor co-agonist that differs from glycine. J. Physiol. 2009, 587, 3207–3220. [Google Scholar] [CrossRef]
  382. Zhang, H.X.; Lyons-Warren, A.; Thio, L.L. The glycine transport inhibitor sarcosine is an inhibitory glycine receptor agonist. Neuropharmacology 2009, 57, 551–555. [Google Scholar] [CrossRef] [Green Version]
  383. Kopec, K.; Flood, D.G.; Gasior, M.; McKenna, B.A.W.; Zuvich, E.; Schreiber, J.; Salvino, J.M.; Durkin, J.T.; Ator, M.A.; Marino, M.J. Glycine transporter (GlyT1) inhibitors with reduced residence time increase prepulse inhibition without inducing hyperlocomotion in DBA/2 mice. Biochem. Pharmacol. 2010, 80, 1407–1417. [Google Scholar] [CrossRef]
  384. 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] [Green Version]
  385. Desch, M.; Wunderlich, G.; Goettel, M.; Goetz, S.; Liesenfeld, K.-H.; Chan, T.S.; Rosenbrock, H.; Sennewald, R.; Link, J.; Keller, S.; et al. Effects of Cytochrome P450 3A4 Induction and Inhibition on the Pharmacokinetics of BI 425809, a Novel Glycine Transporter 1 Inhibitor. Eur. J. Drug Metab. Pharmacokinet. 2021, 47, 91–103. [Google Scholar] [CrossRef]
  386. Wolinsky, E. Statement of the Tuberculosis Committee of the Infectious Diseases Society of America. Clin. Infect. Dis. 1993, 16, 627–628. [Google Scholar] [CrossRef] [PubMed]
  387. Smits, J.A.; Rosenfield, D.; Otto, M.W.; Marques, L.; Davis, M.L.; Meuret, A.E.; Simon, N.M.; Pollack, M.H.; Hofmann, S.G. D-cycloserine enhancement of exposure therapy for social anxiety disorder depends on the success of exposure sessions. J. Psychiatr. Res. 2013, 47, 1455–1461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  388. Farrell, L.J.; Waters, A.M.; Tiralongo, E.; Mathieu, S.; McKenzie, M.; Garbharran, V.; Ware, R.S.; Zimmer-Gembeck, M.J.; McConnell, H.; Lavell, C.; et al. Efficacy of D-cycloserine augmented brief intensive cognitive-behavioural therapy for paediatric obsessive-compulsive disorder: A randomised clinical trial. Depress. Anxiety 2022. [Google Scholar] [CrossRef] [PubMed]
  389. Kvale, G.; Hansen, B.; Hagen, K.; Abramowitz, J.S.; Børtveit, T.; Craske, M.G.; Franklin, M.E.; Haseth, S.; Himle, J.A.; Hystad, S.; et al. Effect of D-Cycloserine on the Effect of Concentrated Exposure and Response Prevention in Difficult-to-Treat Obsessive-Compulsive Disorder: A Randomized Clinical Trial. JAMA Netw. Open. 2020, 3, e2013249. [Google Scholar] [CrossRef] [PubMed]
  390. Olden, M.; Wyka, K.; Cukor, J.; Peskin, M.; Altemus, M.; Lee, F.S.; Finkelstein-Fox, L.; Rabinowitz, T.; Difede, J. Pilot Study of a Telehealth-Delivered Medication-Augmented Exposure Therapy Protocol for PTSD. J. Nerv. Ment. Dis. 2017, 205, 154–160. [Google Scholar] [CrossRef] [Green Version]
  391. Cole, J.; Selby, B.; Ismail, Z.; McGirr, A. D-cycloserine normalizes long-term motor plasticity after transcranial magnetic intermittent theta-burst stimulation in major depressive disorder. Clin. Neurophysiol. 2021, 132, 1770–1776. [Google Scholar] [CrossRef]
  392. Rothbaum, B.O.; Price, M.; Jovanovic, T.; Norrholm, S.D.; Gerardi, M.; Dunlop, B.; Davis, M.; Bradley, B.; Duncan, E.J.; Rizzo, A.; et al. A randomized, double-blind evaluation of D-cycloserine or alprazolam combined with virtual reality exposure therapy for posttraumatic stress disorder in Iraq and Afghanistan War veterans. Am. J. Psychiatry 2014, 171, 640–648. [Google Scholar] [CrossRef]
  393. Mataix-Cols, D.; Fernández de la Cruz, L.; Monzani, B.; Rosenfield, D.; Andersson, E.; Pérez-Vigil, A.; Frumento, P.; de Kleine, R.A.; Difede, J.; Dunlop, B.W.; et al. D-Cycloserine Augmentation of Exposure-Based Cognitive Behavior Therapy for Anxiety, Obsessive-Compulsive, and Posttraumatic Stress Disorders: A Systematic Review and Meta-analysis of Individual Participant Data. JAMA Psychiatry 2017, 74, 501–510. [Google Scholar] [CrossRef]
  394. Polese, D.; De Serpis, A.A.; Ambesi-Impiombato, A.; Muscettola, G.; De Bartolomeis, A. Homer 1a Gene Expression Modulation by Antipsychotic Drugs Involvement of the Glutamate Metabotropic System and Effects of D-Cycloserine. Neuropsychopharmacology 2002, 27, 906–913. [Google Scholar] [CrossRef]
  395. Van Berckel, B.N.; Evenblij, C.N.; van Loon, B.J.; Maas, M.F.; van der Geld, M.A.; Wynne, H.J.; van Ree, J.M.; Kahn, R.S. D-cycloserine increases positive symptoms in chronic schizophrenic patients when administered in addition to antipsychotics: A double-blind, parallel, placebo-controlled study. Neuropsychopharmacology 1999, 21, 203–210. [Google Scholar] [CrossRef] [Green Version]
  396. Kuppili, P.P.; Menon, V.; Sathyanarayanan, G.; Sarkar, S.; Andrade, C. Efficacy of adjunctive d-Cycloserine for the treatment of schizophrenia: A systematic review and meta-analysis of randomized controlled trials. J. Neural Transm. 2021, 128, 253–262. [Google Scholar] [CrossRef] [PubMed]
  397. Takayama, K.; Hitachi, K.; Okamoto, H.; Saitoh, M.; Odagiri, M.; Ohfusa, R.; Shimada, T.; Taguchi, A.; Taniguchi, A.; Tsuchida, K.; et al. Development of Myostatin Inhibitory d-Peptides to Enhance the Potency, Increasing Skeletal Muscle Mass in Mice. ACS Med. Chem. Lett. 2022, 13, 492–498. [Google Scholar] [CrossRef] [PubMed]
  398. Aillaud, I.; Kaniyappan, S.; Chandupatla, R.R.; Ramirez, L.M.; Alkhashrom, S.; Eichler, J.; Horn, A.H.C.; Zweckstetter, M.; Mandelkow, E.; Sticht, H.; et al. A novel D-amino acid peptide with therapeutic potential (ISAD1) inhibits aggregation of neurotoxic disease-relevant mutant Tau and prevents Tau toxicity in vitro. Alzheimer’s Res. Ther. 2022, 14, 15. [Google Scholar] [CrossRef] [PubMed]
  399. Schumacher, T.N.M.; Mayr, L.M.; Minor, D.L.; Milhollen, M.A.; Burgess, M.W.; Kim, P.S. Identification of d-Peptide Ligands Through Mirror-Image Phage Display. Science 1996, 271, 1854–1857. [Google Scholar] [CrossRef] [Green Version]
  400. Malhis, M.; Kaniyappan, S.; Aillaud, I.; Chandupatla, R.R.; Ramirez, L.M.; Zweckstetter, M.; Horn, A.H.C.; Mandelkow, E.; Sticht, H.; Funke, S.A. Potent Tau Aggregation Inhibitor D-Peptides Selected against Tau-Repeat 2 Using Mirror Image Phage Display. ChemBioChem 2021, 22, 3049–3059. [Google Scholar] [CrossRef]
  401. Sadowski, M.; Pankiewicz, J.; Scholtzova, H.; Ripellino, J.A.; Li, Y.; Schmidt, S.D.; Mathews, P.M.; Fryer, J.D.; Holtzman, D.M.; Sigurdsson, E.M.; et al. A Synthetic Peptide Blocking the Apolipoprotein E/β-Amyloid Binding Mitigates β-Amyloid Toxicity and Fibril Formation in Vitro and Reduces β-Amyloid Plaques in Transgenic Mice. Am. J. Pathol. 2004, 165, 937–948. [Google Scholar] [CrossRef]
  402. Chalifour, R.J.; McLaughlin, R.W.; Lavoie, L.; Morissette, C.; Tremblay, N.; Boulé, M.; Sarazin, P.; Stéa, D.; Lacombe, D.; Tremblay, P.; et al. Stereoselective Interactions of Peptide Inhibitors with the β-Amyloid Peptide. J. Biol. Chem. 2003, 278, 34874–34881. [Google Scholar] [CrossRef] [Green Version]
  403. Funke, S.A.; Van Groen, T.; Kadish, I.; Bartnik, D.; Nagel-Steger, L.; Brener, O.; Sehl, T.; Batra-Safferling, R.; Moriscot, C.; Schoehn, G.; et al. Oral Treatment with the d-Enantiomeric Peptide D3 Improves the Pathology and Behavior of Alzheimer’s Disease Transgenic Mice. ACS Chem. Neurosci. 2010, 1, 639–648. [Google Scholar] [CrossRef]
  404. Engel, H.; Guischard, F.; Krause, F.; Nandy, J.; Kaas, P.; Höfflin, N.; Köhn, M.; Kilb, N.; Voigt, K.; Wolf, S.; et al. finDr: A web server for in silico D-peptide ligand identification. Synth. Syst. Biotechnol. 2021, 6, 402–413. [Google Scholar] [CrossRef]
  405. Valiente, P.A.; Wen, H.; Nim, S.; Lee, J.; Kim, H.J.; Kim, J.; Perez-Riba, A.; Paudel, Y.P.; Hwang, I.; Kim, K.-D.; et al. Computational Design of Potent D-Peptide Inhibitors of SARS-CoV. J. Med. Chem. 2021, 64, 14955–14967. [Google Scholar] [CrossRef]
  406. He, W.; Zhang, Z.; Yang, W.; Zheng, X.; You, W.; Yao, Y.; Yan, J.; Liu, W. Turing milk into pro-apoptotic oral nanotherapeutic: De novo bionic chiral-peptide supramolecule for cancer targeted and immunological therapy. Theranostics 2022, 12, 2322–2334. [Google Scholar] [CrossRef] [PubMed]
  407. Zhou, Y.; Chen, Y.; Tan, Y.; Hu, R.; Niu, M.M. An NRP1/MDM2-Targeted D-Peptide Supramolecular Nanomedicine for High-Efficacy and Low-Toxic Liver Cancer Therapy. Adv. Healthc. Mater. 2021, 10, e2002197. [Google Scholar] [CrossRef]
  408. Porcelli, S.; Van Der Wee, N.; van der Werff, S.; Aghajani, M.; Glennon, J.C.; van Heukelum, S.; Mogavero, F.; Lobo, A.; Olivera, F.J.; Lobo, E.; et al. Social brain, social dysfunction and social withdrawal. Neurosci. Biobehav. Rev. 2018, 97, 10–33. [Google Scholar] [CrossRef] [PubMed]
  409. Azmanova, M.; Pitto-Barry, A.; Barry, N.P.E. Schizophrenia: Synthetic strategies and recent advances in drug design. MedChemComm 2018, 9, 759–782. [Google Scholar] [CrossRef] [PubMed]
  410. Dogra, S.; Conn, P.J. Metabotropic glutamate receptors as emerging targets for the treatment of schizophrenia. Mol. Pharmacol. 2022, 101, 275–285. [Google Scholar] [CrossRef]
  411. Jankowska, A.; Satała, G.; Partyka, A.; Wesołowska, A.; Bojarski, A.J.; Pawłowski, M.; Chłoń-Rzepa, G. Discovery and Development of Non-Dopaminergic Agents for the Treatment of Schizophrenia: Overview of the Preclinical and Early Clinical Studies. Curr. Med. Chem. 2019, 26, 4885–4913. [Google Scholar] [CrossRef]
  412. Gouvêa-Junqueira, D.; Falvella, A.C.B.; Antunes, A.; Seabra, G.; Brandão-Teles, C.; Martins-De-Souza, D.; Crunfli, F. Novel Treatment Strategies Targeting Myelin and Oligodendrocyte Dysfunction in Schizophrenia. Front. Psychiatry 2020, 11, 379. [Google Scholar] [CrossRef]
  413. Hashimoto, K. Recent Advances in the Early Intervention in Schizophrenia: Future Direction from Preclinical Findings. Curr. Psychiatry Rep. 2019, 21, 75. [Google Scholar] [CrossRef]
  414. Sonnenschein, S.F.; Grace, A.A. Emerging therapeutic targets for schizophrenia: A framework for novel treatment strategies for psychosis. Expert Opin. Ther. Targets 2020, 25, 15–26. [Google Scholar] [CrossRef]
  415. Roychaudhuri, R.; Snyder, S.H. Mammalian D-cysteine: A novel regulator of neural progenitor cell proliferation: Endogenous D-cysteine, the stereoisomer with rapid spontaneous in vitro racemization rate, has major neural roles. Bioessays 2022, 44, e2200002. [Google Scholar] [CrossRef]
  416. Wang, L.; Zhou, K.; Fu, Z.; Yu, D.; Huang, H.; Zang, X.; Mo, X. Brain Development and Akt Signaling: The Crossroads of Signaling Pathway and Neurodevelopmental Diseases. J. Mol. Neurosci. 2016, 61, 379–384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  417. Li, Y.-C.; Gao, W.-J. GSK-3β activity and hyperdopamine-dependent behaviors. Neurosci. Biobehav. Rev. 2011, 35, 645–654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  418. Kunii, Y.; Matsumoto, J.; Izumi, R.; Nagaoka, A.; Hino, M.; Shishido, R.; Sainouchi, M.; Akatsu, H.; Hashizume, Y.; Kakita, A.; et al. Evidence for Altered Phosphoinositide Signaling-Associated Molecules in the Postmortem Prefrontal Cortex of Patients with Schizophrenia. Int. J. Mol. Sci. 2021, 22, 8280. [Google Scholar] [CrossRef] [PubMed]
  419. Luo, D.-Z.; Chang, C.-Y.; Huang, T.-R.; Studer, V.; Wang, T.-W.; Lai, W.-S. Lithium for schizophrenia: Supporting evidence from a 12-year, nationwide health insurance database and from Akt1-deficient mouse and cellular models. Sci. Rep. 2020, 10, 647. [Google Scholar] [CrossRef] [Green Version]
  420. Moretti, P.N.; Ota, V.K.; Gouvea, E.S.; Pedrini, M.; Santoro, M.L.; Talarico, F.; Spindola, L.M.; Carvalho, C.M.; Noto, C.; Xavier, G.; et al. Accessing Gene Expression in Treatment-Resistant Schizophrenia. Mol. Neurobiol. 2018, 55, 7000–7008. [Google Scholar] [CrossRef]
  421. Rampino, A.; Marakhovskaia, A.; Soares-Silva, T.; Torretta, S.; Veneziani, F.; Beaulieu, J.M. Antipsychotic Drug Responsiveness and Dopamine Receptor Signaling; Old Players and New Prospects. Front. Psychiatry 2019, 9, 702. [Google Scholar] [CrossRef] [Green Version]
  422. Martínez-Banaclocha, M. N-acetyl-cysteine in Schizophrenia: Potential Role on the Sensitive Cysteine Proteome. Curr. Med. Chem. 2020, 27, 6424–6439. [Google Scholar] [CrossRef]
  423. Deng, X.; Zhang, Y.; Chen, Z.; Kumata, K.; Van, R.; Rong, J.; Shao, T.; Hatori, A.; Mori, W.; Yu, Q.; et al. Synthesis and preliminary evaluation of 4-hydroxy-6-(3-[11C]methoxyphenethyl)pyridazin-3(2H)-one, a 11C-labeled-amino acid oxidase (DAAO) inhibitor for PET imaging. Bioorganic Med. Chem. Lett. 2020, 30, 127326. [Google Scholar] [CrossRef]
Figure 1. Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) flow chart. The diagram details the database searches, the number of abstracts screened, the full-text documents retrieved, and the number of included and excluded studies.
Figure 1. Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) flow chart. The diagram details the database searches, the number of abstracts screened, the full-text documents retrieved, and the number of included and excluded studies.
Biomolecules 12 00909 g001
Figure 2. NMDAR structures and binding sites. Whereas glutamate binds to a dedicated glutamate-binding site, the co-agonists D-serine and glycine bind to the so-called “glycine B site”. D-cycloserine acts as a partial agonist at this site. NMDAR activation requires the concomitant binding of glutamate and co-agonists. NMDAR: N-Methyl-D-aspartate receptors; Mg2+: magnesium. Created with BioRender.com (accessed on 14 June 2022).
Figure 2. NMDAR structures and binding sites. Whereas glutamate binds to a dedicated glutamate-binding site, the co-agonists D-serine and glycine bind to the so-called “glycine B site”. D-cycloserine acts as a partial agonist at this site. NMDAR activation requires the concomitant binding of glutamate and co-agonists. NMDAR: N-Methyl-D-aspartate receptors; Mg2+: magnesium. Created with BioRender.com (accessed on 14 June 2022).
Biomolecules 12 00909 g002
Figure 3. D-serine is synthesized in astrocytes by serine racemase, an enzyme that converts L-serine into D-serine. D-serine acts as a co-agonist at synaptic NMDARs, whereas its reuptake is performed by the neutral amino acid transporters Alanine-serine-cysteine-threonine (ASCT) 1 and 2. D-amino acid oxidase is responsible for D-serine degradation in glial cells. ASCT: Alanine-serine-cysteine-threonine transporter; EEAT1: Excitatory amino acid transporter 1; EEAT2: Excitatory amino acid transporter 2; SNAT: Sodium-coupled neutral amino acid transporter; NMDAR: N-Methyl-D-aspartate receptors; DAO: D-amino acid oxidase. Created with BioRender.com (accessed on 14 June 2022).
Figure 3. D-serine is synthesized in astrocytes by serine racemase, an enzyme that converts L-serine into D-serine. D-serine acts as a co-agonist at synaptic NMDARs, whereas its reuptake is performed by the neutral amino acid transporters Alanine-serine-cysteine-threonine (ASCT) 1 and 2. D-amino acid oxidase is responsible for D-serine degradation in glial cells. ASCT: Alanine-serine-cysteine-threonine transporter; EEAT1: Excitatory amino acid transporter 1; EEAT2: Excitatory amino acid transporter 2; SNAT: Sodium-coupled neutral amino acid transporter; NMDAR: N-Methyl-D-aspartate receptors; DAO: D-amino acid oxidase. Created with BioRender.com (accessed on 14 June 2022).
Biomolecules 12 00909 g003
Figure 4. Glycine transporter 1 (GlyT1) controls glycine concentration in the synaptic cleft. GlyT1 is localized at the postsynaptic density of glutamatergic synapses and may physically interact with NMDAR complex. GlyT1 activity is regulated by indirect phosphorylation by CaMKII. It has been proposed that clozapine may exert its antipsychotic activity, among other mechanisms, by inhibiting GlyT1. GlyT1: Glycine Receptor Transporter 1; NMDAR: N-Methyl-D-aspartate receptors; PSD-95: postsynaptic density protein 95; CaMKII: Ca2+/calmodulin-dependent protein kinase; AMPAR: α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor; PSD-95: postsynaptic density protein 95. Mg2+: magnesium. Created with BioRender.com (accessed on 14 June 2022).
Figure 4. Glycine transporter 1 (GlyT1) controls glycine concentration in the synaptic cleft. GlyT1 is localized at the postsynaptic density of glutamatergic synapses and may physically interact with NMDAR complex. GlyT1 activity is regulated by indirect phosphorylation by CaMKII. It has been proposed that clozapine may exert its antipsychotic activity, among other mechanisms, by inhibiting GlyT1. GlyT1: Glycine Receptor Transporter 1; NMDAR: N-Methyl-D-aspartate receptors; PSD-95: postsynaptic density protein 95; CaMKII: Ca2+/calmodulin-dependent protein kinase; AMPAR: α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor; PSD-95: postsynaptic density protein 95. Mg2+: magnesium. Created with BioRender.com (accessed on 14 June 2022).
Biomolecules 12 00909 g004
Table 1. Randomized clinical trials investigating the efficacy of D-amino acids in treating schizophrenia.
Table 1. Randomized clinical trials investigating the efficacy of D-amino acids in treating schizophrenia.
Agent AddedAuthorStudy DesignGroups of TreatmentMean AgenDoseDurationStable Antipsychotic RegimensOutcomeSide Effects
D-serineKantrowitz et al., 2018 [223]Double-blind crossover trialD-serine40 ± 111460 mg/kg/day6 weeksCPZ equivalents (mg): 965 ± 760Improvement in total PANSS symptoms
Reduction in evoked power for the α band
-
Placebo-
Weiser et al., 2012 [228]RCTD-serineNot retrieved1952 g/day-Not retrievedNo significant differences in SANS and MATRICS battery score-
Placebo
Tsai et al., 1998 [226]Double-blind, placebo-controlled trialD-serine31.7 ± 7.51530 mg/kg/day6 weeksSulpiride (n = 6),
Haloperidol (n = 5);
Risperidone (n = 4);
Pipotiazine (n = 2)
Fluphenthixol (n = 2)
Thiothixene (n = 1)
Pipotiazine (n = 2)
Haloperidol and pipotiazine (n = 1)
Pipotiazine and chlorpromazine (n = 1)
Fluphenazine and chlorpromazine (n = 1)
Risperidone and chlorpromazine (n = 1)
Trifluoperazine and chlorpromazine (n = 1)
Trifluoperazine and haloperidol (n = 1)
One patient was antipsychotic free.
Improvement in PANSS positive
SANS
PANSS-cognitive
WCST
CGI
No significant side effects
Placebo33.9 ± 6.614-
Heresco-Levy et al., 2005 [222]double-blind, placebo-controlled trialRisperidone + D-serine
Olanzapine + D-serine
42.7 ± 13.12130 mg/kg/day6 weeks-Improvement in BPRS
SANS
PANSS (negative, positive, cognitive)
No significant clinical or laboratory side effects
Risperidone + Placebo
Olanzapine + Placebo
47.4 ± 14.218
Lane et al., 2005 [224]RCTD-serine31.8 ± 10.2212 g/day6 weeksRisperidoneNo significant differencesWeight gain (67%)
Palpitation (42%)
Placebo34.1 ± 8.723
Lane et al., 2010 [225]RCTD-serine30.7 ± 9.6202 g/day6 weeksRisperidone
Olanzapine
Quetiapine
No significant differencesNo significant side effects
Placebo31.5 ± 7.920
Tsai et al., 2010 [227]RCTD-serine42.6 ± 3.61030 mg/kg/day6 weeksClozapine mean dose: 363 ± 128No significant differencesNo differences in side effects
Placebo39.5 ± 5.510Clozapine mean dose: 315 ± 146
Kantrowitz et al., 2010 [230]Open label trialD-serine low dose41.7 ± 11.41230 mg/kg CPZ Equivalents/day468.8 ± 252Improvement in PANSS and CGI-
D-serine mild dose43.5 ± 9.41960 mg/kg602.6 ± 295-
D-serine high dose43.2 ± 9.616120 mg/kg493.7 ± 319Nephrotoxic-like pattern at 120 mg/kg which resolved upon D-serine discontinuation (6.25%)
Asymptomatic transaminitis (12.5%)
Insomnia after one dose and GI distress (12.5%)
Kantrowitz et al., 2015 [26]Double-blind, placebo-controlled trialD-serine13–351560 mg/kg/day16 weeks-Scale of Prodromal Symptoms negative score-
Placebo20-
D-alanineTsai et al., 2006 [27]RCTD-alanine30.9 ± 6.514100 mg/kg daily6 weeksCPZ equivalents (mg): 468 ± 478Sulpiride; Haloperidol; Risperidone; Pipotiazine; Pipotiazine and chlorpromazine; Pipotiazine and sulpiride; Thioridazine; Trifluoperazine and chlorpromazine; Fluphenazine decanoate and chlorpromazineImprovement in PANSS-positive
PANSS-cognitive and SANS score
No significant differences
Placebo31.8 ± 7.418CPZ equivalents (mg): 364 ± 220
PANSS, Positive and Negative Syndrome Scale; SANS, Scale for the Assessment of Negative Symptoms; MATRICS, Measurement and Treatment Research to Improve Cognition in Schizophrenia; WCST, Wisconsin Card Sorting Test; CGI, Clinical Global Impression; BPRS, Brief Psychiatric Rating Scale; RCT, Randomized Controlled Trial; TRS, Treatment-Resistant Schizophrenia; CPZ, chlorpromazine.
Table 2. Randomized clinical trials investigating the efficacy of sodium benzoate as a DAO-inhibitory agent in treating chronic schizophrenia and early psychosis.
Table 2. Randomized clinical trials investigating the efficacy of sodium benzoate as a DAO-inhibitory agent in treating chronic schizophrenia and early psychosis.
AuthorsGroups of TreatmentMean AgenDoseDurationStable Antipsychotic RegimensOutcomeSide Effects
Lane et al., 2013
[317]
Sodium benzoate38.4 ± 9.7251 g/day6 weeksAmisulpride
Chlorpromazine
Flupenthixol
Haloperidol
Quetiapine fumarate
Risperidone
Sulpiride
Ziprasidone
Zotepine
The sodium benzoate group exhibited better performances in SANS, GAF, QOLS, CGI, HDRS, speed of processing, visual
learning and memory
No significant differences
were found between the two groups in Simpson-Angus Rating
Scale, AIMS, and Barnes Akathisia Scale
scores
Placebo36.3 ± 7.927-
Scott et al., 2020 [318]Sodium benzoate21.7 ± 4.7491 g/day12 weeksAripiprazole
Amisulpride
Brexpiprazole
Clozapine
Haloperidol
Olanzapine
Lurasidone
Paliperidone
Quetiapine
Risperidone
No improvements were found in the sodium benzoate groupNo significant differences
were found between groups
Placebo21.2 ± 3.450-
Lin et al., 2018 [311]Sodium benzoate44.3 ± 7.2201 g/day6 weeksClozapineImprovements in negative and overall symptomatology and quality of life were detected in sodium benzoate groups compared to placeboNo significant differences
were found between the two groups in Simpson-Angus Rating
Scale, AIMS, and Barnes Akathisia Scale
scores
Sodium benzoate44.8 ± 8.1202 g/day
Placebo47 ± 11.920-
Lin et al., 2017 [312]Sarcosine + sodium benzoate37.8 ± 9.6212 g/day + 1 g/day12 weeksAmisulpride
Aripiprazole
Olanzapine
Paliperidone
Quetiapine
Risperidone
Zotepine
Improvement in GAF and cognitive battery compared to sarcosine group-
Sarcosine38.2 ± 9.3212 g/dayImprovement in reasoning and problem solving compared to placebo
Placebo39.1 ± 9.521--
AIMS, Abnormal Involuntary Movement Scale; CGI, Clinical Global Impression; GAF, Global Assessment of Function; HDRS, Hamilton Depression Rating Scale–17 items; QOLS, Quality of Life Scale; SANS, Scales for the Assessment of Negative Symptoms–20 item; RCT, randomized controlled trial.
Table 3. Randomized clinical trials investigating the efficacy of glycine-centered agents in treating schizophrenia.
Table 3. Randomized clinical trials investigating the efficacy of glycine-centered agents in treating schizophrenia.
Agent AddedAuthorStudy DesignGroups of TreatmentMean AgenDosesDurationStable Antipsychotic RegimensOutcomeSide Effects
GlycineSerrita et al., 2019 [334]RCTGlycine48.60 ± 5.01100.8 g/kg12 weeksAntipsychoticsNo significant differences in PANSS scoreNo significant differences
Placebo49.20 ± 4.8410
Rosse et al., 1989 [345]Open-label Pilot studyGlycine38 ± 15618.8 g/day8 weeksHaloperidol;
Benztropine;
Thiothixene;
Vitamin E;
Loxapine
Beneficial effects in two patients at SANS scale.
Two others are worsened
Reduction in neuroleptic-induced muscle stiffness and extrapyramidal dysfunction in three patients.
Potkin et al., 1999 [335]RCTGlycine12.4 ± 7.2930 mg/day12 weeksClozapine (400–1200 mg/day)No significant treatment effectsNo significant differences
Placebo10Improvement in BPRS positive symptoms
Javitt et al., 1994 [336]RCTGlycine-1430 mg/day8 weeksUnknownImprovement in PANSS negative symptoms domain-
Placebo
Heresco-Levy et al., 1999 [339]RCTGlycine38.8 ± 11.0961.2 ± 13.46 weeksClozapine mean dose: 471.0 ± 207.8 mg
Chlorpromazine equivalents: 240.0 ± 146.3
Significant increase in glycine and serine levels
Significant improvement in PANSS—negative symptoms and BPRS total score
No significant differences
Placebo10
Buchanan et al., 2007 [343]RCTGlycine42.6 ± 10.84260 g/day16 weeksUnknownnon-significant differences in SANS score
Significant improvement in CGI score in D-cycloserine group compared to placebo
Worsened nausea and dry mouth in glycine group compared to placebo
D-cycloserine44.4 ± 10.44650 g/day
Placebo43.4 ± 11.445-
Costa et al., 1990 [344]Open-label Pilot studyGlycine34.81115 g/day5 weeksIncreasing amount in second generation antipsychoticsDecrease in BPRS score in 33% of patientsUpper gastrointestinal discomfort in 17% of patients
Liederman et al., 1996 [346]Open-label Pilot studyGlycine45.0 ± 7.6560.0 g/day8 weeksClozapine (2)
Risperidone (2)
Haloperidol (1)
Significant improvement in negative symptoms was found using the SANSNo adverse effects; Reduction in EPS
Heresco-Levy et al., 1996 [338]RCTGlycine41.36 ± 12.931160.63 ± 12.986 weeksClozapine (4)
Thioridazine (3)
Haloperidol (2)
Chlorpromazine (1)
Perphenazine (1)
Improvement in PANSS-negative symptoms
PANSS-general psychopathology and PANSS total score
No adverse effects
Placebo
Javitt et al., 2001 [337]RCTGlycine39.6 ± 5.5120.8 g/kg/day6 weeksOlanzapine (6)
Clozapine (4)
Mesoridiazine (1)
Haloperidol (1)
Improvement in PANSS- negative symptoms No significant adverse effects
Unknown
Placebo
Evins et al., 2000 [341]RCTGlycine39.0 ± 7.02760 g/day8 weeksClozapineNo significant treatment effectsNo significant adverse effects
Placebo
Diaz et al., 2005 [342]RCTGlycine39.5 ± 12.44660 g/day8 weeksClozapine mean dose: 575 mg/dayNo significant treatment effectsNausea and vomiting (16%)
Placebo6-
Greenwood et al., 2018 [340]RCTGlycine36.0 ± 7.81224.8 g/day6 weeksUnspecified antipsychotics medicationImprovement in PANSS-Total, PANSS-Negative and
PANSS-General symptoms
Improvement in MMN
-
Placebo40.2 ± 8.910
Heresco-Levy et al., 2004
[326]
RCTGlycine44.7 ± 10.8755.2 ± 10.1 g/day6 weeksOlanzapine (10) mean dose: 14.3 ± 6.2 mg/day
Risperidone mean dose (4): 6.20 ± 3.08 mg/day
Improvement in five-factor PANSSDecrease in EPS,
upper gastrointestinal tract discomfort with nausea (12%)
Placebo7
SarcosineLane et al., 2005b [224]RCTSarcosine36.1 ± 10.2212 g/day6 weeksRisperidoneImprovement in PANSS total, positive, and negative score; SANS-20 and SANS-17Treatment-emergent adverse events other than extrapyramidal symptoms are similar in the three groups
D-serine31.8 ± 10.4212 g/day-
Placebo34.1 ± 8.723--
Lane et al., 2006 [347]RCTSarcosine36.7 ± 10.1102 g/day6 weeksClozapine mean dose: 306 ± 159 mgNon-significant differences Non-significant differences
Placebo35.5 ± 6.610Clozapine mean dose: 305 ± 55 mg
Lane et al., 2010 [225]RCTSarcosine30.4 ± 10.620Not retrieved6 weeksRisperidone
Olanzapine
Quetiapine
Not retrieved-
Placebo31.5 ± 7.920
D-serine30.7 ± 9.620
Lin et al., 2017 [312]RCTSarcosine + sodium benzoate 37.8 ± 9.6212 g/day + 1 g/day12 weeksAmisulpride
Aripiprazole
Olanzapine
Paliperidone
Quetiapine
Risperidone
Zotepine
Improvement in GAF and cognitive battery compared to sarcosine group-
Sarcosine38.2 ± 9.3212 g/dayImprovement in reasoning and problem solving compared to placebo
Placebo39.1 ± 9.521--
Tsai et al., 2004 [350]RCTSarcosine29.8 ± 7.2172 g/day6 weeksChlorpromazine equivalence: 409 ± 320Improvement in PANSS scoreNo significant side effects
Placebo33.4 ± 8.321Chlorpromazine equivalence: 433 ± 243
Lane et al., 2008 [351]RCTSarcosine31.3 ± 10.491 g/day6 weeksDrug-freeNo significant differences between groupsInsomnia, weight gain (1-2 kg), sedation, constipation, and fatigability. These effects were all mild and brief.
Sarcosine34.3 ± 11.2112 g/day
Strzelecki et al., 2018 [352]RCTSarcosine37.3 ± 11.3292 g/day6 monthsFirst- and second-generation antipsychoticsImprovement in total, general, and negative PANSS scoresNo significant side effects
Placebo40.2 ± 10.130
Strzelecki et al., 2015 [353,354]RCTSarcosine36.5252 g/day6 monthsAntipsychotics except clozapineImprovement in total PANSS scoresNo significant side effects
Placebo4025
BitopertinUmbricht et al., 2014 [355]RCTBitopertin high dose38.9 ± 9.57960 mg/day8 weeksAripiprazole
Olanzapine
Quetiapine
Paliperidone
Risperidone
Risperidone long-acting injection
-Serious adverse effects in 3.8% of patients
Bitopertin mild dose40.7 ± 9.48130 mg/dayImprovement in PANSS and NSFS scoreSerious adverse effects in 2.4% of patients
Bitopertin low dose41.1 ± 10.48210 mg/dayImprovement in PANSS, NSFS, CGI-I-N, and PSP scoreSerious adverse effects in 1.21% of patients
Placebo30.0 ± 10.881 --
Bugarski-Kirola et al., 2016 [356]RCTBitopertin high dose39.1 ± 12.220220 mg/day12 weeksUnspecified-1%
Bitopertin low dose40.2 ± 12.420010 mg/dayImprovement in PANSS and PSFS score compared to placebo2%
three deaths due to upper gastrointestinal hemorrhage caused by duodenal ulcer, alcohol poisoning and related head injury, and suicide
Placebo39.7 ± 12.7200--3.5%
Kantrowitz et al., 2017 [357]RCTBitopertin 40 ± 111710 mg/day6 weeksCPZ equivalents > 600 mgNo differences between groups No side effect
Placebo43 ± 1212-
BI425809Fleischhacker et al., 2021 [358]RCTBI42580936.5 ± 8.5852 mg/day12 weeksUnspecified antipsychotics treatmentImprovement in cognitive functions in BI425809 groups compared to placebo. The largest enhancement from baseline vs. placebo was observed in the 10 mg and 25 mg dose group Headache
Somnolence
Gastrointestinal disorders
BI42580937.5 ± 7.9845 mg/day
BI42580937.9 ± 6.88510 mg/day
BI42580936.2 ± 7.88525 mg/day
Placebo37.2 ± 7.7170-
D-cycloserineCain et al., 2014 [359]RCTD-cycloserine48.8 ± 11.51850 mg/day8 weeksClozapine and others (unspecified)Improvement in performance on the auditory discrimination task, working memory, and negative symptoms-
Placebo46.2 ± 13.318
Duncan et al., 2004 [360]RCTD-cycloserine48.7 ± 12.110250 mg/day4 weeksUnspecifiedImprovement in SANS and BPRS scoreNo side effects
Placebo54.4 ± 11.812
Goff et al., 1999 [361]RCTD-cycloserine36.6 ± 9.61750 mg/day6 weeksUnspecifiedWorsening in negative symptoms-
Placebo
Goff et al., 1999 [362]RCTD-cycloserine46.8 ± 12.32350 mg/day8 weeksUnspecifiedImprovement in SANS score-
Placebo41.2 ± 8.124
Goff et al., 2005 [363]RCTD-cycloserine45.9 ± 7.42250 mg/day6 monthsUnspecifiedNo significant differences9%
Placebo47.0 ± 8.62821%
Goff et al., 2008 [364]RCTD-cycloserine50.1 ± 9.151950 mg/day8 weeksAll antipsychotic except for clozapineImprovement in SANS scoreNo side effects
Placebo48.0 ± 6.6619
Heresco-Levy et al., 2002 [365]RCTD-cycloserine40.0 ± 12.1850 mg/day6 weeksConventional antipsychotics and olanzapine, or risperidoneImprovement in SANS and PANSS negative scoreWorsening of psychotic symptoms in two patients on treatment and two during placebo
Placebo8
Rosse et al., 1996 [366]Open-label studyD-cycloserine high dose38.1 ± 6.8630 mg/day4 weeksMolindone 150 mg/dayNo significant differencesNo side effects
D-cycloserine low dose310 mg/day
Placebo4-
Takiguchi et al., 2017 [367]RCT crossover studyD-cycloserine44.0 ± 14.03650 mg/day6 weeksUnspecified: 827.0 ± 609.5 CPZ equivalentsNo significant differencesLiver enzyme elevation in two patients of the placebo group
Placebo
PANSS, Positive and Negative Syndrome Scale; SANS, Scale for the Assessment of Negative Symptoms; CGI, Clinical Global Impression; CGI-I-N, CGI Improvement of Negative Symptoms; GAF, Global Assessment of Function; BPRS, Brief Psychiatric Rating Scale; MMN, Mismatch negativity; NFSF, Need Satisfaction and Frustration Scale; RCT, Randomized Controlled Trial; PSP, Personal Social Performance; PSFS, Patient Specific Functional Scale; CPZ, chlorpromazine; EPS, extrapyramidal symptoms.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

de Bartolomeis, A.; Vellucci, L.; Austin, M.C.; De Simone, G.; Barone, A. Rational and Translational Implications of D-Amino Acids for Treatment-Resistant Schizophrenia: From Neurobiology to the Clinics. Biomolecules 2022, 12, 909. https://doi.org/10.3390/biom12070909

AMA Style

de Bartolomeis A, Vellucci L, Austin MC, De Simone G, Barone A. Rational and Translational Implications of D-Amino Acids for Treatment-Resistant Schizophrenia: From Neurobiology to the Clinics. Biomolecules. 2022; 12(7):909. https://doi.org/10.3390/biom12070909

Chicago/Turabian Style

de Bartolomeis, Andrea, Licia Vellucci, Mark C. Austin, Giuseppe De Simone, and Annarita Barone. 2022. "Rational and Translational Implications of D-Amino Acids for Treatment-Resistant Schizophrenia: From Neurobiology to the Clinics" Biomolecules 12, no. 7: 909. https://doi.org/10.3390/biom12070909

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

Article Metrics

Back to TopTop