Next Article in Journal
Identification of Novel GSK-3β Hits Using Competitive Biophysical Assays
Next Article in Special Issue
Changes in Brain Volumes Are Relevant during Natalizumab-Associated Progressive Multifocal Leukoencephalopathy: Lessons from a Case Report
Previous Article in Journal
Impact of Vitamin D in Prophylaxis and Treatment in Tuberculosis Patients
Previous Article in Special Issue
Pathogenetic Mechanisms of Hypertension–Brain-Induced Complications: Focus on Molecular Mediators
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Autism Spectrum Disorder: Focus on Glutamatergic Neurotransmission

1
Laboratory of Neurophysiology and Plasticity, IRCCS Fondazione Santa Lucia, 00179 Rome, Italy
2
Department of Systems Neuroscience, University Tor Vergata, 00133 Rome, Italy
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(7), 3861; https://doi.org/10.3390/ijms23073861
Submission received: 28 February 2022 / Revised: 24 March 2022 / Accepted: 29 March 2022 / Published: 31 March 2022
(This article belongs to the Special Issue State-of-the-Art Molecular Neurobiology in Italy)

Abstract

:
Disturbances in the glutamatergic system have been increasingly documented in several neuropsychiatric disorders, including autism spectrum disorder (ASD). Glutamate-centered theories of ASD are based on evidence from patient samples and postmortem studies, as well as from studies documenting abnormalities in glutamatergic gene expression and metabolic pathways, including changes in the gut microbiota glutamate metabolism in patients with ASD. In addition, preclinical studies on animal models have demonstrated glutamatergic neurotransmission deficits and altered expression of glutamate synaptic proteins. At present, there are no approved glutamatergic drugs for ASD, but several ongoing clinical trials are currently focusing on evaluating in autistic patients glutamatergic pharmaceuticals already approved for other conditions. In this review, we provide an overview of the literature concerning the role of glutamatergic neurotransmission in the pathophysiology of ASD and as a potential target for novel treatments.

1. Introduction

Autism Spectrum Disorder (ASD) is a mosaic of neurodevelopmental conditions, which show common deficits in two behavioral domains: (i) social interaction and communication difficulties, (ii) narrow interests, and repetitive and stereotyped behaviors (DSM-5) [1,2]. ASD can be associated with several co-occurring conditions, including seizures, attention-deficit/hyperactivity disorder (ADHD), and other cognitive impairments [3,4]. The estimated prevalence of ASD is about 1% of the human population [5], with males affected four times more frequently than females [6].
Although it has become clear that ASD has a complex and multifactorial etiopathogenesis, twin and family studies indicate a strong genetic background and high heritability, with a concordance rate of 60% to 95% between monozygotic twins versus 0% to 30% between dizygotic twins [7]. Despite the high heterogeneity between ASD cases, the occurrence of shared symptoms suggests common deficits in some neurodevelopmental pathways. Neurotransmitters play a critical role in developing both the peripheral and the central nervous systems. It is therefore conceivable that neurotransmitter dysfunctions may be involved in ASD pathophysiology. Among neurotransmitters, glutamate (Glu) is considered a good candidate [8,9,10] as it is directly involved in brain development and synaptogenesis [11,12], memory, behavior, and motor activity regulation [13,14,15], and gastrointestinal functions [16,17]. Studies on postmortem or body fluids samples provide important evidence about critical changes in Glu concentration in both pediatric and adult patients with ASD [18,19]. Furthermore, abnormalities in Glu receptors genes and deregulation of glutamatergic pathways have been reported both in ASD patients and animal models [20,21,22,23]. In general, excitotoxicity has already been connected with ASD. In addition, the increased prevalence of epilepsy in autistic patients compared to the general population [24,25,26] further strengthens the hypothesis of the dysfunction in excitatory and/or inhibitory network activity. In the past decades, two opposite theories about the role of Glu signaling in ASD have been proposed. In 1998, Carlsson suggested that ASD might result from a decrease in Glu signaling according to (i) overlapping between the ASD condition and the symptoms produced by Glu antagonist, (ii) the neuroanatomical and neuroimaging evidence of glutamatergic areas impairment in ASD patients, (iii) the similarities with ASD symptoms in hypoglutamatergic animals treated with NMDA antagonists [27]. Conversely, Fatemi proposed an opposite theory, suggesting an hyperglutamatergic state in ASD, based on the observation of increased Glu levels in blood samples [19].
In this review, starting from the signaling of Glu in typical brain development, we will present current evidence supporting Glu involvement in ASD pathophysiology, including reports on alterations in intestinal Glu metabolism and ongoing clinical trials for the treatment of the core symptoms of the spectrum.

2. The Critical Role of the Glutamatergic System in Brain Development

ASD symptoms appear early in development, within the first three years of life, during the fundamental period of rapid synapse formation and maturation [28,29]. Brain development involves a number of processes, including synaptogenesis, axonal and dendritic arborization, migration, and synaptic plasticity. These functions have the overall aim of building a functional brain. They can, however, cause cellular, biochemical, and structural alterations in the neonatal brain [30]. During brain development, the role played by neurotransmitters and their receptors is of primary importance. The distribution and molecular characteristics of Glu receptors change considerably during brain development, making the brain vulnerable to changes in Glu neurotransmission during growth. Indeed, alterations in the expression and regulation of Glu receptors are known to be implicated in some neuropathological conditions such as neurological and psychiatric disorders, including Alzheimer’s and Parkinson’s diseases, schizophrenia, mood disorders, depression, epilepsy, anxiety, stress, and ASD [31,32,33,34]. Studies in animal models of neurologic disease suggest an altered expression of Glu proteins [12,35]. Knowledge and research on the ontogenetic alterations in Glu receptor function, subunit expression, and binding properties in pathological conditions are still very incomplete. However, in vivo and in vitro studies have provided information about the differences in both regional densities and the time-course of changes in the expression levels of the different Glu receptors subunits (Figure 1) [36].
For example, the α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor (AMPAR) is constituted by tetramers of different subunits (GluR1-4) [37]. The expression pattern of the GluR2 subunit during the period of postnatal development, maturation, and aging in male mice was investigated: it was observed that the expression of the GluR2 subunit is gradually upregulated in the hippocampus from postnatal day 0 (P0) to adult age (20 weeks) and subsequently down-regulated in 70 week-old male mice [38]. GluR1 and GluR4 subunits are expressed at significant levels in the midbrain from embryonic day 15 (E15), they remain constant until delivery, and then decrease immediately after birth, while GluR2 and GluR3 levels increase after early development in the hippocampus [34]. N-methyl-D-aspartate receptors (NMDAR) are expressed during the developmental periods of intense synaptic formation. Similar to AMPAR, the expression of the various NMDAR subunits varies over the course of development and in different regions of the brain [39]. Studies in rodents have shown that the GluN1 subunit is expressed before E14 in areas of the brain related to cognitive function, such as the cortex, hippocampus, and thalamus. Three weeks after birth, GluN1 levels reach their highest levels and then decrease [40]. The GluN2 subunits are expressed at variable levels at specific stages of neonatal development in rats. In the hippocampus, low levels of GluN2A are detected at P0, while a peak is observed at P21 [41]. As the brain develops, the protein expression of the GluN2A subunit gradually increases. GluN2B is the predominant NMDA receptor subunit located at immature synapses [36]. During prenatal development, it is expressed in the cortex, thalamus, and spinal cord, and its concentrations increase at birth. Lower amounts are expressed in the colliculi, hippocampus, and hypothalamus [34]. After birth, GluN2B levels reach a peak of expression in the cortex and hypothalamus, two important brain structures involved in cognition. The GluN3A subunit is involved in developing dendritic spines and synaptogenesis [34]. This subunit starts to be expressed in the medulla, pons, and hypothalamus at E15 [42,43]; its levels reach a peak at P8 and then decrease at P20. Kainate-2-carboxy-3-carboxy-methyl-4-isopropenylpyrrolidine receptors (KAR) are expressed at low amounts from E19, during embryonic development, in the cortex, hippocampus, cerebellum, and striatum [34]. Their activation regulates the network and synaptic activity in the neonatal hippocampus [44,45]. KAR are widely expressed in the amygdala within the first postnatal week, a period that coincides with the process of synaptogenesis, suggesting its involvement in the process of synapse formation [45]. Rodent studies have demonstrated a low expression at birth of type 1, 2, and 4 Glu metabotropic receptors (mGlu1, mGlu2, mGlu4) that increase during neonatal development [43]. Conversely, at P0, the levels of type 3 and 5 Glu metabotropic receptors, mGlu3 and mGlu5, are very high and then decrease during the maturation period. In fact, both receptors are involved in synaptogenesis, and GRM5 is involved in the proliferation and survival of neural progenitor cells, as well as in the migration of cortical neurons [46]. Another important aspect of the brain development process is the maintenance of the excitation-inhibition (E/I) balance between Glu and Gamma-Aminobutyric Acid (GABA) [47]. Opposite to Glu, GABA is involved in inhibitory neurotransmission [48]. However, at the beginning of development, GABAergic neurons form excitatory synapses, which become inhibitory only later during the maturation process [49]. For normal brain development and functioning, it is essential that the E/I balance is maintained stably in neuron’s synapses and neural circuits [10,50]. Consequently, disturbances in the E/I balance have been implicated in neurodevelopmental disorders, including ASD [51,52].

3. Sex Differences in the Glutamatergic System

Accumulating evidence has revealed that gonadal steroids finely modulate neurotransmission mediated by amino acids, including Glu, in developing and adult brains [53]. Indeed, there appear to be functionally relevant baseline sex differences in neurotransmitter systems [11,54]. Clinical studies have highlighted increased Glu levels in frontal grey matter and basal ganglia (BG) of females compared to males, while parietal grey matter (PGM) Glu concentration is higher in men than in women [55]. Further studies have identified in more detail sex differences in Glu concentration in specific brain regions. Females were found to exhibit higher Glu levels than males in the sensorimotor cortex, anterior cingulate cortex (ACC), striatum (STR), and cerebellum [56,57]. In contrast, prefrontal cortex (PFC) Glu concentration is higher in males than in females [58]. Along with the Glu concentration changes observed in the CNS, gender differences were also found in plasma Glu concentrations. Clinical studies reported higher Glu levels in males than females, which appear inversely correlated to estrogen and progesterone levels [59]. Gender differences in Glu concentration appear to become more conspicuous with age. Males, but not females, display an age-dependent decline in BG and PGM Glu levels [55], while females show a more pronounced age-dependent decrease in ACC Glu levels [60]. In addition, blood Glu levels exhibit an age-related increase in plasma in females but not in males [61]. Similar to what is seen in humans, preclinical studies have revealed significant sex-related differences in the Glu system of rodents. Increased Glu levels have been found in the lateral hypothalamus and habenula of male rats [62]. Additionally, rodents show a sex bias in Glu receptors distribution. For example, female rats show an enhanced expression of mGlu2/3, mGlu5, and the GluN1 and Glu2B subunits of the NMDAR in the hippocampus [63,64,65], along with greater mGlu5 expression in the PFC [66]. In line with the reported sex-related differences in the rodent Glu system, in female rats, AMPAR-dependent synaptic events are altered: synaptic transmission is enhanced, whereas the long-term potentiation (LTP) magnitude is reduced, the latter alteration likely due to reduced phosphorylation of the GluR1 subunit [67].
Sex-specific factors have been hypothesized to increase male’s risk of ASD, or increase the protection of females from ASD, as more males than females are affected. The sex-related differences in the Glu system discussed above suggest the importance of elucidating the molecular mechanisms by which Glu dysfunction could differentially affect males and females.

4. Glutamate Signaling in ASD

Glu is the main excitatory neurotransmitter in the mammalian brain. The Glu concentration in the CNS has been estimated at around 10 mmol/L in the intracellular compartment [68], a much higher level than in the extracellular fluid (0.5–2 μmol/L) [69,70], in the cerebrospinal fluid (10 µmol/L) or plasma (150 µmol/L) [71]. As perturbations in the Glu system can have deleterious effects, brain Glu concentration is tightly regulated by a large number of mechanisms, such as the endothelial cells of the blood-brain barrier (BBB) and the Glu/glutamine (Glx) cycle between neurons and astrocytes, just to name a few. A large body of literature describes changes in both peripheral and brain levels of Glu and glutamine in ASD patients [72,73]. The first clinical reports, dating back to the 1990s, focused on plasma Glu levels. Using high-performance liquid chromatography (HPLC), Moreno et al. reported a significant increase in Glu and aspartate levels in children with ASD compared to controls [74,75]. The following study additionally reported reduced levels of plasma glutamine [74,75]. Notably, increased levels of plasma Glu and decreased levels of blood glutamine were also reported in siblings and parents of patients with ASD [18]. Since these first reports, many more described higher levels of plasma Glu and lower levels of plasma glutamine in children or adults with ASD compared to healthy controls [76,77,78,79,80,81,82,83,84,85,86]. Higher plasma levels of Glu have been associated with increased severity of ASD [82]. Similarly, it has been described as an elevated Glx ratio [87] in affected patients. It is worth noting that a few papers have shown a reduction or no change in blood or serum Glu concentration [8,73]. Altered Glu levels have been reported in plasma but also in brain regions of patients with ASD [8]. Although Glu does not enter the brain in appreciable quantities because of saturable and stereoselective facilitative transporters on the luminal membranes [70], blood Glu levels could, in principle, influence the brain concentration [73]. Despite Takado et al. demonstrating a positive correlation between plasma and brain levels only for glutamine but not for Glu [88], both these amino acids are metabolically engaged in a tightly balanced and highly dynamic cycle to maintain the proper concentrations of Glu and GABA [50,89]. In-vivo measurements of Glu levels in the human brain have been obtained by magnetic resonance spectroscopy (H-MRS). Because the molecular structures of Glu and glutamine are very similar, Glu spectral features are usually contaminated by contributions from glutamine; therefore, usually, the concentration of both amino acids is evaluated. Such work showed significantly higher concentrations of Glx in the amygdala-hippocampal region [90], primary sensory cortex [91], anterior cingulate cortex (ACC) [92,93,94], and auditory cortex [95], and a trend for lower Glu in the right medial temporal lobe [93], in people with ASD. Moreover, Siegel-Ramsay et al. reported an association of the increased Glx concentration with reduced functional connectivity between ACC and insular, limbic, and parietal cortical regions in patients with ASD [96]. However, some works reported opposite results, specifically a reduction of Glx or Glu concentration in the right ACC [97,98,99] basal ganglia [100] and cerebellum [98,101] in patients with ASD compared to healthy controls. Although with some discrepancies, overall, the data summarized above suggest an increase in Glu concentration in patients with ASD. In a healthy adult brain, both Glu and GABA maintain the excitatory/inhibitory (E/I) balance [102]. An enhanced Glu concentration causes prolonged activation of Glu receptors that may, in turn, produce a calcium overload in neurons. This phenomenon is known as excitotoxicity [103]. Several lines of evidence suggest that glutamatergic dysfunction, excitotoxicity, and neuroinflammation represent intertwined phenomena [104]. Indeed, cortical infusion of NMDA increased the expression of the pro-inflammatory cytokines tumor necrosis factor-alpha (TNF-α) and Interleukin-1beta (IL-1b) [105]. Accordingly, administration of an NMDA antagonist reduced both IL-1b and TNF-α levels [106]; on the other hand, an interleukin-1 receptor antagonist reduced excitotoxicity [107]. Glia and immune cells control pro-inflammatory activity driven by “on” and “off” signals originating from neurons. Glu release is an “on” signal that activates the release of the pro-inflammatory cytokines TNF-α and IL-1b [108]. While, on the one side, Glu can increase cytokine release, on the other side, cytokines seem able to reinforce Glu release. Indeed, the expression of astroglial high-affinity Glu transporters, contributing to Glu removal, is reduced during neuroinflammatory processes [109]. Furthermore, Glu release and accumulation are controlled by both IL-1b, by increasing Glu export via the cysteine/Glu exchanger [110] and TNF-α signaling [111]. Finally, recent evidence suggests that cytokines can induce ionotropic glutamatergic receptors mobilization. Indeed, TNF-α was shown to enhance excitatory synaptic strength through the exocytosis of AMPA receptors [112] and to decrease inhibitory synaptic strength by inducing the endocytosis of GABA(A) receptors [113], overall moving the E/I balance towards excitation. IL-1b was reported to boost NMDA receptor function by promoting GluN2A/B subunit phosphorylation [114] and inhibit GABA(A) receptor function [115], but other work showed that it down-regulates AMPA receptor surface expression [116].

5. The Genetics of the Glutamatergic System in ASD

The list of genes implicated in ASD susceptibility gets constantly richer, thanks to the large-scale genetic studies conducted on affected patients and their families. Because Glu receptors are highly enriched in brain areas associated with ASD, it is not surprising that mutations in Glu receptor genes and modifications in mRNA and protein expression of Glu receptors subunits have been found in patients. For the purpose of this review, we screened the Simons Foundation for Autism Research Initiative (SFARI) database (https://gene.sfari.org/, accessed on 31 January 2022), which includes hundreds of ASD candidate risk genes, to provide a comprehensive picture of the glutamatergic genetic risk factors for ASD (Table 1). The SFARI database is organized into four categories reflecting the overall strength of the available evidence supporting a gene’s relevance to ASD risk. The syndromic category includes mutations that are associated with a substantial degree of increased risk and consistently linked to additional characteristics not required for an ASD diagnosis. In this category, only the SLC1A2 gene, encoding for a glial high-affinity Glu transporter, is involved in the Glu neurotransmission system. Studies by Xu et al. have shown that mutations within the SLC1A2 gene occurred in ASD patients more frequently than in those without ASD [117]. Genes in category 1 are considered highly confident because they have been clearly implicated in ASD. Only two genes related to Glu neurotransmission are included in this category: GRIA2, encoding for the subunit type 2 of AMPAR [118,119,120], and GRIN2B, encoding for regulatory subunits type 2B of NMDAR [121,122,123,124]. Heterozygous de novo GRIA2 variants have been reported in patients diagnosed with intellectual disability and neurodevelopmental abnormalities, including ASD [125]. Further evidence has been provided by Ramanathan et al., who observed, in an autistic patient, a 19 megabase deletion in the chromosomal region containing the GRIA2 gene [126]. Sequencing studies of GRIN2B in ASD patients have identified a significant excess of rare missense mutations [127,128,129], severe de novo splice-site variants, and three additional de novo loss-of-function variants [130,131]. Further analysis for de novo mutations has confirmed GRIN2B as a high confidence candidate ASD gene [132,133,134]. Category 2 includes genes with two reported de novo likely-gene-disrupting mutations that have been implicated in ASD by a genome-wide association study. We have identified seven Glu genes in this category. GRIA1, encoding for the subunit type 1 of AMPAR, is a strong ASD candidate because a missense variant (p.Ala636Thr) has been identified in affected patients diagnosed with intellectual disability, developmental defects, and ASD [121,135,136]. Similarly, GRIN2A, encoding for regulatory subunits type 2A of NMDAR, has often been associated with neurodevelopmental phenotypes and ASD [121,123,124,135,136]. Interestingly, rare pathogenic deletions of the GRIN2A gene [137,138,139,140,141,142] and heterozygous de novo missense variants [143] have been implicated in childhood focal epilepsy. Moreover, next-generation sequencing of children with ASD has revealed variants in the GRIN2A gene with evidence that supports ASD pathogenicity [144,145]. Genetic studies have also found a genetic association between ASD and GRIK genes belonging to the kainate family of Glu receptors. In particular, GRIK2, encoding for kainate subunit type 2, and GRIK 5, encoding for kainate subunit type 5, have been classified as strong ASD candidates. A linkage disequilibrium has been revealed between ASD and GRIK2 in a mutation screening in a Chinese family [21,146]. Regarding GRIK5, three de novo missense mutations have been identified in children with ASD [147,148]. An association with ASD has also been found for the GRIP1 gene, encoding for a scaffold protein mediating trafficking and membrane organization of several transmembrane proteins. In particular, rare but high-incidence missense mutations have been reported in 480 cases compared to 480 controls [149]. Finally, ASD genome-wide copy number variation analysis reveals mutations in GRID1, the gene encoding for the Glu ionotropic receptor delta type subunit 1, in affected patients [150,151]. The last category comprises candidate genes with significant but unreplicated evidence of association with ASD or evidence lacking a rigorous comparison with healthy controls. GRM5 and GRM7 genes, encoding for Glu metabotropic receptors 5 and 7, respectively, are included in this category. However, there are some additional genes not included in the SFARI database. Indeed, genome-wide association studies provided evidence of a linkage between autistic disorder and variants in the chromosomal region 7q21–32, containing the Glu metabotropic receptor 8 gene (GRM8) [20]. Additionally, a microdeletion in the chromosomal region 7q31.33q32.1, containing 13 genes, including GRM8, has been found in a child with ASD and intellectual disability [152].

6. Glutamatergic Gene and Protein Expression in ASD

Post mortem studies examining the brains of individuals with ASD have pointed out the presence of alterations in Glu receptor expression. Significantly decreased AMPAR density and upregulated NMDAR subunit 1 protein levels have been observed in the cerebellum [22]. On the contrary, Blatt et al. failed to find differences in NMDAR expression in hippocampal tissue from 4 ASD patients [159]. Changes in metabotropic Glu receptors have also been described. Increased levels of mGlu5 receptor have been found in the cerebellar tissue of 11 affected patients [160]. Similarly, individuals with Fragile X syndrome (FXS), frequently co-diagnosed with ASD, showed increased mGlu5 protein expression in the prefrontal cortex compared with matched healthy controls [161]. A reduction in glutamic acid decarboxylase (GAD) has been reported in the parietal and cerebellar areas of affected patients compared to healthy controls [162], a deficiency that may account for the dysfunction in blood and brain Glu levels. GAD is a rate-limiting enzyme in the Glu/GABA cycle, converting Glu to GABA, and exists as two isoforms, GAD65 and GAD67 [163]. GAD67, whose gene is localized in a locus on chromosome 2 that showed susceptibility for ASD [164], is abundantly expressed in cerebellar Purkinje cells. Interestingly, a significant reduction of GAD67 mRNA levels in Purkinje cells of affected patients has been pointed out [165]. Notably, also the number and size of Purkinje cells were decreased in autistic patients [166]. In a further study, Yip et al. demonstrated a significant reduction of GAD65 mRNA in large-sized neurons of the dentate gyrus in children with ASD [167]. Increased mRNA levels of the excitatory amino acid transporter (EAAT) 1 and EAAT2 have also been described in the cerebellum of ASD patients [22]. The increased Glu concentration observed in plasma and spectroscopic studies are likely contributing to the enhanced EAAT expression [168]. Histological studies on brains of patients with ASD revealed small-sized neurons and increased neuronal packing in the limbic regions, including the hippocampus [169]. Aylward et al. reported a reduced volume of the amygdala and hippocampus in autistic patients with respect to the whole brain [170]. A subsequent study by Raymond et al. (1996) utilizing Golgi staining reported a reduced complexity in the dendritic branching of hippocampal pyramidal neurons, suggesting a curtailment in neuronal maturation [171].

7. The Role of Epigenetics in ASD

Epigenetics can provide insights into genetic and environmental disease risk factors and gene–environment interactions but also contribute to identifying disease-relevant genomic regions and biomarkers [172]. Epigenetic modifications appear to be strongly associated with ASD, suggesting that not only the genome but also the epigenome plays a role in ASD pathogenesis [173]. Indeed, several ASD-linked genes show epigenetic changes, such as DNA methylation or post-transcriptional histone modifications [174]. Increasing evidence suggests that early life environments may play a key role in ASD and other neurodevelopmental disorders. Epigenetics represent prime mediators of the environmental effects on genome and phenotype, and modifications caused, for example, by viral infections or the use of some drugs during pregnancy, can drive derangements of the gene regulatory processes during development [175,176,177]. There are four different mechanisms of epigenetic modifications that collectively play essential roles in gene transcription and expression: histone modifications, DNA methylation, RNA interference, and RNA modifications [178]. Epigenome-wide association studies (EWAS) on neurodevelopmental diseases focused primarily on the DNA methylation mechanism; conversely, less is known about histone modification changes [176,178,179,180]. Glutamate decarboxylase 1 (GAD1, aka GAD67), encoding an enzyme that catalyzes GABA production, is an example of epigenetic gene modifications associated with ASD [176,181]. Increased DNA methylation of the GAD1 gene promoter, accompanied by increased binding of the transcription repressor Methyl Cpg 2 (MeCP2), was described in the cerebellum of patients with ASD [182]. Accordingly, GAD1 expression was reduced in two rat models of ASD, the Poly(I:C)-induced maternal immune activation model [183] and rats prenatally exposed to the anticonvulsant valproate [184]. In addition, murine models with haploinsufficiency [185] or loss of GAD1 show deficits in social behavior and learning [186]. Interestingly, ASD is frequently associated with both an abnormal process of methylation and significantly differential DNA methylation patterns [176,187,188]. DNA methyltransferases (DNMTs) [176,188] polymorphisms seem involved in the pathogenesis of ASD [189]. The DNMT family includes three different enzymes, DNMT1, DNMT3a, and DNMT3b, playing a fundamental role in diverse cellular processes, such as neurodevelopment, synaptic plasticity, long-term memory learning, and training [187,190]. Accordingly, a reduced expression of DNMT1 or DNMT3a in excitatory neurons of the mouse forebrain causes alterations in synaptic plasticity, learning, and memory processes [191]. Excessive DNMT1 activity [192] or increased methylation of MeCP2 promoter lower the expression of MeCP2 [193] and consequently induce the GABA/Glu imbalance that might be involved in ASD pathophysiology [194,195]. Several studies have shown that epigenetic changes can alter the expression of pre- and postsynaptic proteins involved in ASD, such as Neurexins [196], Neuroligins (NLGNs; [197], PSD-95 [190], and SHANK3 [187]. There is experimental evidence describing an association between a differential methylation state of the promoter region of these genes and their downregulation, and genetic variations of these proteins have been implicated in ASD as responsible for the appearance of a variety of symptoms such as cognitive and social deficits and anxiety [172]. Despite conflicting evidence about the role in ASD pathogenesis of epigenetics, it may mediate or modify the genetic or environmental disease risk or represent the biological mechanism for gene-environment interactions [198]. On the other hand, it could be possible to identify epigenetic modifications as biomarkers of exposure to environmental risk factors [199].

8. Role of Glutamatergic Transmission in ASD Pathophysiology: Evidence from Animal Models

Alterations in Glu expression levels, reduced excitatory transmission, dysfunction of NMDAR-mediated synaptic plasticity, and mGluR-mediated signal transduction play a role in the cognitive and behavioral deficits of ASD [9]. Both ASD patients and animal models show an altered expression of synaptic proteins as well as dysfunction of excitatory or inhibitory neurons [52,200,201]. These observations gave rise to the theory of the E/I imbalance in the pathophysiology of ASD ([52,200,201]. Indeed, several ASD models, such as the Fmr1 [202] and the Cntnap2 mice [203], as well as the VPA-induced model, [204,205]) show a reduction in GABAergic interneurons numbers [206]. An in vivo optogenetic study provided for the first time a demonstration of the relationship between alterations in the E/I balance in the prefrontal cortex of WT mice and ASD-associated behaviors [207]. In addition, a highly significant correlation between a decreased Glu/GABA ratio in the prefrontal cortex and mice sociability has been recently demonstrated in the Cntnap2−/− mouse model [208]. Such alterations might be caused by reduced inhibitory postsynaptic currents (IPSC) and/or increased excitatory postsynaptic currents (EPSC), as observed in the Cntnap2−/− [209,210] and the VPA [211] model. Preclinical studies have shown a clear role of Glu in ASD-like behaviors. Indeed, glutamatergic neurotransmission dysfunctions have been demonstrated in various animal models (Table 2) [212,213] In particular, the regulatory proteins of the glutamatergic system (Figure 2), including Glu receptors (such as GluK2, mGlu8, NMDAR, GluR1, and GluR2 Glu receptors) and Glu transporters (such as glial Glu transporters, GLT1, GLAST, SLC1A1, and EAAT-1,2) are highly implicated in the pathophysiology of the disorder [10].

8.1. AMPA Receptor

Deficits at the level of AMPAR-mediated currents have been identified in diverse mouse models of ASD [201,232,233]. Shank2 (Shank2−/−) and Shank3 (Shank3αβ−/−) knock-out (KO) mice exhibit reduced levels of several cell surface Glu receptors in the striatum and thalamus: additionally, Shank3αβ−/− KO mouse models showed alterations in AMPAR expression [234]. Similarly, neuronal cultures of the embryonic brain stem from NLGN1-2-3 KO mice showed deficits in AMPAR-mediated glutamatergic transmission, namely alterations in the frequency of spontaneous and miniature excitatory synaptic currents (sEPSCs and mEPSCs) [227]. The R704C-NLGN3 mice show reduced AMPAR currents in the hippocampus [229]. It has been shown that hippocampal CA1 pyramidal neurons in the Fmr1-KO mouse model exhibit a higher AMPAR/NMDAR ratio early in development, returning to normal levels at P13. This early alteration is accompanied by increased expression of the GluR2 subunit in synaptic AMPARs leading to altered Ca2+ permeability [217,218]. Additionally, in FMR1 KO mice, a high expression of the AMPAR subunits GluR1-4 was observed in the hippocampal CA1 pyramidal neurons and Purkinje cells of the cerebellum [214,215], where anatomical alterations have also been observed [235].
Alterations in Glu receptors expression and transmission have been demonstrated both in the Cntnap2 KO model and in mice prenatally exposed to valproic acid (VPA); however, while an enhanced GluR1 and GluN2B expression was observed in VPA mice, a decrease in the subunits GluR1, GluR2, GluN2A, and GluN2B was reported to occur in Cntnap2 KO mice [52]. In line with such evidence, AMPAR modulation improved social deficits in Cntnap2 KO and VPA animal models [52]. Accordingly, two ampachins, CX1837 and CX1739, are drugs acting at AMPARs tested on BTBR mice, an inbred model line with ASD-like behaviors, which showed a selective improvement in social behaviors with no effect on self-grooming [231].

8.2. NMDA Receptor

Several animal models of ASD show impairments of NMDAR signaling [216,221,236,237], including SHANK3 [225], NLGN1-KO [238], FMR1 KO [216], and transgenic MeCP2 mice [239]. Increases in NMDAR expression were found in a prenatal VPA rat model [221], while lower levels were found in the FMR1 KO mouse [216] and the MeCP2−/− mouse model [240]. In the NLGN1 KO model, the NMDAR co-agonist d-cycloserine reduced anxiety and repetitive behaviors, suggesting that the increased NMDAR activity plays a role in some ASD-associated symptoms [228]. In addition, the increased expression of NMDAR in rats exposed to VPA is accompanied by enhanced long-term potentiation, reduced sensorimotor gating, repetitive/stereotyped movements, and abnormal social behaviors [221,222]. The NMDA/AMPA ratio was increased in the pyramidal neurons of the hippocampal CA1 region in NLGN3-R451C KI but not in NLGN3 KO mice. The R451C mutation caused an increase in the expression of the excitatory postsynaptic scaffolding proteins PSD95 and SAP-102, and of the GluN2B subunit of the NMDAR, without affecting AMPAR [230]. Emerging evidence suggests that disrupted NMDAR-mediated signaling may contribute to the pathogenesis of idiopathic ASD [9,127] and related syndromes, including Rett and Fragile X [240]. In support of this view, GluN1 hypomorphic mice, in which GluN1 expression is reduced to 5–10% of control levels, show deficits in social interaction [241]. GluN2A, which encodes one of the subunits of the NMDA receptor, has also been studied in relation to the ASD spectrum as it is located on chromosome 16p, a region identified as a possible ASD susceptibility locus [10,123].

8.3. KA Receptor

The physiological role of KAR is not characterized as for other Glu receptors, but it is clear that it is involved in the maturation of neural circuits during development [242]. Chromosome 6q21 is a region that contains the gene coding for the GluK2 subunit of KAR: this subunit is an excellent positional and functional candidate for ASD susceptibility [21]. GluK2 is known to play a role in the appropriate maturation of synaptic circuits involved in learning and memory [243]. GRIK2 KO (GRIK2−/−) mice have been shown to exhibit both functionally and morphologically delayed maturation of the hippocampal mossy fiber at CA3 (mf-CA3) pyramidal cell synapses; this mutation results in a transient reduction in the amplitude of AMPA-EPSCs at a critical time in postnatal development, affecting the proper formation of memory-related neural circuits [244]. In addition to this, it was observed that GRIK2−/− mice show ASD-like behaviors: deficits in social interaction, reduced locomotor activity, and impaired spatial reversal learning [245]. A mouse model overexpressing the KAR subunit GRIK4 showed a marked decrease in social interactions, accompanied by enhanced anxiety and depressive states. In addition, this model showed synaptic transmission alterations at hippocampal mossy fibers, as well as in the basolateral and central nuclei of the amygdala [246]. In vivo results show anxiety and social interaction deficits, as well as in vitro an alteration in synaptic transmission. These data support the hypothesis that increased expression of the GluK4 subunit of KAR may contribute to ASD-associated behaviors [247].

8.4. MGlu Receptor

Genetic studies indicate a complex and heterogeneous etiology of ASD, although many ASD-linked genes appear to converge on the pathway of synaptic transmission and synapse formation [248,249,250]. One of the common aspects of many animal models of ASD involves altered forms of synaptic plasticity mediated by glutamate metabotropic (mGlu) receptors: this was seen in animal models of Fragile X (FMRP) [219], SHANK3 [226], tuberous sclerosis (TSC1/TSC2) [251], PTEN [252], 16p11.2 16 microdeletion [253] and Rett syndrome (MECP2) [254]. Both mGlu1 and mGlu5 belong to group I metabotropic receptors. In the context of ASD, particular attention was devoted to mGlu5, in light of its role as a regulator of both natal and postnatal neurogenesis and synaptogenesis [255,256], as well as in motor and social behaviors that are specifically affected in ASD disorders [236], suggesting that dysregulation of mGlu5 signaling could be involved in the pathogenesis of neurodevelopmental disorders. At the behavioral level, mGlu5 is involved in locomotor reactivity to novel environments, sensorimotor gating, anxiety, and cognition [257]. MGlu5 KO mice show alterations in the behavioral domains affected in ASD, such as burial behavior, social interaction, locomotor activity, and anxiety [258]. Accordingly, alterations in mGlu5 have been associated to diverse neurological and neuropsychiatric disorders, including FXS [220,259], attention-deficit/hyperactivity ADHD [260], ASD [155], schizophrenia [261,262] and epilepsy [32]. An overactive mGlu5 signaling is thought to be responsible for multiple behavioral features of FXS [219,220]. Interestingly, the synaptic, biochemical, and cognitive defects observed in the TSC2+/− and FMR1−/y mouse models were corrected by treatments that modulate mGlu5 in opposite directions, suggesting that normal synaptic plasticity and cognition occur within an optimal range of mGlu5 mediated effects, and deviations in either direction can lead to shared behavioral impairments [263]. The potential role of mGlu1 has been characterized limited to cognitive aspects in mouse models of ASD. The mGlu1 antagonist, JNJ16259685, restores social deficits in Eif4ebp2 knockout mice but not in Shank2 KO rats [224,264], while CFMTI, another selective mGlu1 antagonist, ameliorates the social interaction deficits induced by the NMDAR antagonist MK-801 in rats [265]. With regard to mGlu antagonists, numerous preclinical studies have shown their key role in treating ASD. FXS pathophysiology is characterized by excessive mGlu signaling that contributes to the behavioral, electrophysiological, and molecular dysfunction associated with the disorder [266]. Preclinical studies in FMR1 KO animal models showed a restoration of aberrant AMPAR expression, behavioral deficits, electrophysiological alterations, protein expression dysregulation, and altered dendritic spine morphology [267] through the use of 2-methyl-6-(phenylethynyl) pyridine (MPEP), a selective mGlu5 antagonist [234].

9. GABA/Glutamate Balance in ASD

As the correct balance between glutamatergic and GABAergic neurotransmission is essential for the normal information processing, stability, and organization of the neuronal networks [207,268], an imbalance between excitation and inhibition, consequence of dysfunctions in terms of potentiation of glutamatergic signaling, and/or weakened GABAergic signaling might be implicated in ASD pathophysiology [200,269,270,271]. GABAergic transmission plays a pivotal role in cellular functions, most importantly during brain development. Indeed, in the embryonic and early postnatal period, GABA exerts an excitatory action on immature neurons and, in synergy with Glu, drives glutamatergic synapse development, cell migration, and differentiation [272,273]. As the developmental depolarizing-to-hyperpolarizing switch in GABA transmission is critical for the correct establishment of the E/I balance [274], alterations in GABAergic signaling can impair normal cellular processing and cause deficits such as those observed in ASD [275]. To date, increasing evidence of dysfunction in the GABAergic neurotransmission system has been reported in patients and animal models of ASD [276,277]. In vivo 1H-MRS measurements have revealed a significant increase in plasma levels of GABA in children with ASD compared to controls [81,85,278]. Conversely, analyses of brain areas reported a reduced GABA/Glu ratio in the frontal lobes of autistic children [279], in the occipital brain areas of adolescent patients with high-functioning ASD [280], and the PFC of adult patients in response to riluzole [281]. Differing results showed either increased GABA levels in the visual cortex or reduced GABA levels in the sensorimotor cortex and reduced GABA/Creatine levels in ACC in children with ASD compared to controls [282,283,284]. Alterations in GABAergic signaling have been demonstrated by brain postmortem studies: namely, an increased expression of GAD67 and reduced expression of GAD 65 in cerebellar cells, an increased density of GABAergic interneurons in the hippocampus, and a downregulation of several GABAergic receptor subunits [167,285,286,287,288]. Furthermore, genetics added further evidence of the involvement of GABAergic dysfunction in ASD. Indeed, rare mutations, including de novo and intrinsic variants, have been found in the chromosome 15q11–q13 region, a site containing coding regions of specific subunits of GABA receptors, including GABRB3, GABRA5, and GABRG3 [289,290,291,292]. Finally, direct evidence of GABAergic dysfunction has been obtained from molecular and cellular studies in several genetic and pharmacological animal models of ASD [293,294,295,296,297,298,299,300,301,302,303,304,305,306]. Overall, the dysfunctions shown in both glutamatergic and GABAergic neurotransmission suggest a pathological switch in the E/I balance underlying, at least partly, the failure of compensatory mechanisms, such as adaptation of synaptic efficacy, plasticity, membrane excitability, and/or synapse numbers, physiologically implemented to prevent over-excitation [307]. Accordingly, in a recent article, Bruining et al. demonstrated an increased functional E/I ratio (fE/I) and long-range temporal correlations (LRTC) of the network activity in children with ASD by means of a computational model of neuronal network oscillations, also providing evidence about a correlation of these findings with the heterogeneity of ASD phenotype [308]. Similarly, Spiegel et al. provided indirect evidence of altered E/I balance in the visual cortex of patients with ASD using the experimental binocular rivalry test paradigm; interestingly, the slower rivalry dynamics in the ASD brain were associated with higher autistic traits [309]. Moreover, in a recent study, a positive correlation between the severity of the autistic phenotype and the magnitude of the E/I ratio was demonstrated using transcranial magnetic stimulation to monitor cortico-spinal excitability and intracortical inhibition [310]. Sex-specific mechanisms responsible for the high male/female ratio of ASD may differentially affect the E/I balance and, consequently, the ASD phenotype, as was recently demonstrated using a combination of in-silico modeling and in-vivo chemogenetic manipulations in mice [311].

10. Alterations of Glutamate Metabolism in the Gut

Increasing evidence suggests that Glu, in addition to being the main excitatory neurotransmitter in the brain, is an important neurotransmitter in the ENS and GI tract [312]. In the gut, Glu, which is derived from food and, in minor part, from microbial activity, is strongly involved in several functions, including taste perception and digestion process. Of the total amount of dietary Glu, 75–95% is rapidly removed to maintain a low concentration in the systemic circulation and prevent dangerous accumulation in other parts of the body [16]. In healthy conditions, dietary Glu does not pass the BBB. However, in pathological situations associated with diet or microbiota alterations, the BBB permeability may change, and Glu accumulates in the brain [313,314]. Enteric Glu strongly contributes to the bidirectional communication between gut and brain, via the activation of receptors distributed on vagal, splanchnic, and pelvic afferents, and vice versa, through the receptors on efferent pathways comprising the dorsal motor nucleus of the vagus (DMV) [315]. Interestingly, the administration of probiotics in mice can stimulate microbiota to modulate metabolic pathways and increase Glu brain levels [316,317]. In addition, it has recently been demonstrated that dysregulation in protein phosphorylation found in hippocampus tissue of mice with dysbiotic gut microbiota is consistently associated with glutamatergic neurotransmitter system disturbances [318]. Indeed, ultrastructural and immunohistochemical evidence demonstrates that ENS is provided with the whole glutamatergic machinery, including vesicular glutamatergic transporters [319,320] and all the known ionotropic (iGlu) and mGlu receptors [17,321]. Both iGlu and mGlu receptors have been found in the esophagus, stomach, small and large intestine, where they may participate in specific functions, including muscle activity and local blood flow [17,315,322]. Interestingly, numerous morphological and molecular similarities exist between enteric glia and astrocytes, particularly in glutamine synthase and iGlu receptors expression [323,324,325], suggesting that also in the ENS, glial cells may contribute to preventing extracellular Glu concentrations from rising to neurotoxic levels [326]. The gut microbiome composition and relative alterations are of particular interest in ASD research in light of numerous reports regarding dysbiosis and GI problems, such as diarrhea, constipation, abdominal pain, and gastric reflux, in children and adults with ASD. Recently, the characterization of the ASD gut microbiome highlighted dysregulations of the enteric glutamatergic neurotransmitter machinery that may have a role in the development of symptoms. A study on 43 ASD children found reduced gut cortisol levels associated with changes in Glu metabolism, with a significant reduction in 2-keto-glutaramic acid, and changes in the microbiota, with particularly low levels of Bacteroides vulgatus and high levels of both Eggerthella lenta and Clostridium botulinum, in fecal samples [327]. In line with this, a ketogenic diet, rich in fats and low in carbohydrates, was shown to reduce seizure susceptibility in two different mouse models of ASD through its impact on the gut microbiota composition, causing an increase in Parabacteroides and, in parallel, in brain Glu levels [328].

11. Pharmacotherapeutics Targeting the Glutamatergic System

Even today, the development of new therapeutic strategies is hampered by the extremely heterogeneous physiopathology of ASD and the scarcity of clear diagnostic markers [329]. In the United States, pharmacological interventions approved for ASD are limited to Risperidone and Aripiprazole, both of which are indicated by the Food and Drug Administration (FDA) for the treatment of irritability associated with ASD [330,331], but not for the core symptoms identified by the DSM-5 as essential for ASD diagnosis [332]. Therefore, it is crucial to find new pharmacological targets for effective pathogenesis-based treatments aimed at reducing symptoms, improving functioning, and potentially impacting long-term outcomes. Several ongoing clinical trials (https://clinicaltrials.gov, accessed on 31 January 2022) target a number of neurobiological pathways, GABAergic and Glutamatergic transmission, neuroinflammation, neuropeptides, and the endocannabinoid system. So far, Glu receptor antagonists have shown limited clinical efficacy in treating ASD. Drugs tested in these trials include memantine, an NMDA receptor antagonist, and mGlu5 antagonists, such as basimglurant, fenobam, and mavoglurant [8]. The phenotypic rescue demonstrated in FMR1 KO mice modulating mGlu5 activity has driven extensive clinical trial work since 2008. By means of a PubMed search, we identified 22 clinical studies, of which 19 (86%) are registered on www.ClinicalTrials.gov (accessed on 31 January 2022). Most of these studies (14/22, 64%) targeted the main excitatory/inhibitory imbalance primarily through mGlu5 antagonists or GABA agonists. Trials of three different mGlu5 antagonists—phenobam, mavoglurant (AFQ056), and basimglurant (RO4917523)—in FXS were completed. Although the alteration of E/I balance seems to be strongly associated with ASD, it is still not entirely clear how it can be modulated [333,334]. Functional magnetic resonance imaging (fMRI) was used to measure the functional connectivity of the prefrontal cortex following pharmacological treatment with the drug riluzole [281]. This drug has been used because it has a wide range of actions on both GABA and Glu targets [335,336] and consequently appears to be able to modulate Glu-GABA flux in most individuals. Riluzole shows a good pharmacokinetic profile and a very low risk of side effects; in addition, a single oral dose of 50 mg (as used in the study) reaches a peak plasma concentration 1 h after administration [337,338]: riluzole blocks the presynaptic release of Glu, facilitates GABA receptor activity and appears to be involved in the altered Glx balance in bipolar disorder [339]. Treatment with riluzole increased the inhibitory index of the prefrontal cortex in patients with ASD compared to controls and rescued the reduced prefrontal functional connectivity; notably, in controls, the decrease in GABA fraction was correlated with an increase in glutamine, whereas in ASD patients, riluzole increased the inhibitory index without altering glutamine [337,339]. In summary, riluzole shifts the balance towards GABA in ASD patients and towards Glutamine in controls. Riluzole acts by inhibiting voltage-dependent sodium channels, reducing neurotransmitter release, and favoring astrocytic uptake of extracellular Glu [340]. With regard to the pharmacological effects of riluzole, it was seen that long-term treatment (10–12 weeks) had a significant effect on irritability in young children with ASD [341], whereas long-term monotherapy (excluded medications, riluzole 50–200 mg/day for 8 weeks) additionally reduced the hyperactivity [342]. In contrast, in a group of children and young adults (aged 12–25 years), no pharmacological effect of riluzole was found [343]. There are many studies using memantine, a non-selective antagonist of NMDARs. Joshi et al. used memantine in a 12-week randomized trial to treat social deficits in ASD adolescents [344]. Neuroimaging, functional magnetic resonance imaging, and resonance spectroscopy analyses were conducted pre- and post-treatment to assess functional neuronal deficits in adolescents with ASD and healthy volunteers. A significant improvement in non-verbal communication was observed in the Diagnostic Analysis of Nonverbal Accuracy Scale test and the executive function for self-report. Memantine was well-tolerated without causing serious side effects [344]. Very often, high Glu levels correspond to hyperexcitability in the brain, which can be caused by an elevated level of Glu in the synapses and/or overexpression of its receptors. In this double-blind, placebo-controlled study, 28 children with a primary diagnosis of ASD were given lamotrigine, a drug that modulates Glu release. The children who received the treatment showed improvement in the language aspect of communication and socialization [345]. NMDA receptor antagonists appear to be effective in hyperexcitability and aggressive behavior [346]. Memantine has been hypothesized to play a role in potentially modulating learning, blocking the excessive effects of Glu, which may include neuroinflammatory activity, and influencing neuroglial activity in ASD [347]. Although the results of clinical trials of pharmacological treatments for ASD so far show promising results, outcomes remain highly variable in the treatment of core symptoms, and there is still no solid demonstration of efficacy involving one or more pathways or targets for one treatment over another. When designing a drug development plan for the treatment of core symptoms of ASD, it is critical to take into account that ASD is a neurodevelopmental disorder that spans the entire human lifespan and that efficacy and response to specific treatments may differ over this timeframe [345,348,349,350]. Although there are still doubts about the hierarchy of evidence supporting the safety and efficacy of such use, it is clear that the unmet clinical needs of individuals with ASD change over the course of a lifetime, suggesting that the relevance of specific classes of new (experimental) or old drugs (and the molecular targets through which they act) may vary over the course of a lifetime [345,348,349,350]. It cannot be assumed that the biology involved in new and different targets will be consistent at all stages of development. Therefore, it is reasonable that initial approval for an ASD treatment should include children and young adults. Similarities can be drawn with attention-deficit/hyperactivity disorder, where the FDA requires initial applications for new drugs to include children up to 6 years of age. Separate clinical trials in children and adults are likely to be required for drug development to ensure efficacy and safety in both groups. After initial approval, younger and older individuals with ASD could be studied as part of the post-approval effort or through other mechanisms [348,349,350].

12. Conclusions

ASD is a multifactorial neurodevelopmental disease caused by multiple etiology involving genetic, epigenetic, and environmental risk factors. Nevertheless, the congruence of ASD behavioral symptomatology and developmental stage suggests a common neurobiological-based etiology. Here, we examined the current literature supporting the possible involvement of glutamatergic dysfunctions in ASD pathophysiology as glutamate is directly involved in brain development and function as well as the bidirectional communication axis between gut microbiota and the Central Nervous System. The hypothesis that glutamate signaling is altered in ASD is supported by post-mortem research, clinical reports, and animal studies. In addition, the large-scale genetic studies conducted on affected patients and their families provide strong support for the idea that glutamate plays a major role in ASD pathophysiology. In the last decade, the prevalence of ASD has dramatically risen, although it remains unclear if this reflects an effective increase in occurrence or if better knowledge and more awareness of the autistic status have facilitated diagnosis. Indeed, it could be a concurrence of factors: the advancement of genetic research, the capacity for early identification of autistic signs, the development of new guidelines, and diagnostic criteria. In addition, recent years have brought the understanding that in addition to genetics, other factors play a role in ASD pathogenesis, such as epigenetic modifications, prenatal exposure to drugs or infections, immune dysfunctions, and alterations in the gut microbiome. Admittedly, there are still many aspects to be clarified in ASD. A first challenge remains to elucidate the enormous phenotypic heterogeneity, including adaptive function, cognitive and language abilities, and neurological comorbidities, and characterizes different disorders within the spectrum. Secondly, we still know very little about how the autistic condition evolves in adulthood or whether or not the main symptoms persist or change [351]. The diagnosis in adults is complicated by the lack of relevant information and the camouflage behaviors that adults, more than children or adolescents, can implement [352]. Also of concern is the limited availability of targeted therapies and/or the limited success in developing new medications. In this sense, identifying new biomarkers could greatly benefit the development of targeted and precision treatments. Clarifying how glutamatergic dysfunction contributes to the symptomatology of ASD might help to identify new pharmacological targets. To date, there are no approved glutamatergic drugs for ASD, although several clinical trials are in an early stage aimed at evaluating glutamatergic pharmaceuticals already approved for other conditions in patients with ASD. The pharmacological modulation of group I mGlu receptors may represent a potential therapeutic strategy for the treatment of ASD [353], and, indeed, some antagonists of mGlu1 and mGlu5 receptors were shown to rescue some phenotypes in ASD mouse models [264,354]. Nevertheless, thus far, figuring out how Glu biology is affected in ASD has proved to be extremely challenging. On one hand, the comorbidity of ASD with other Glu-related diseases, such as epilepsy, can provide excellent insights, but, on the other hand, the ASD heterogeneity still represents an important limitation in understanding how Glu changes can impact ASD pathophysiology. Moreover, glutamatergic dysfunction may significantly differ across the various disorders that characterize the Spectrum and, further complicating matters, glutamatergic transmission is strongly influenced by gonadal hormones, and basal Glu brain concentration differs between males and females. Finally, shedding light on how glutamatergic dysfunction changes and evolves from the earliest to the late stages of disease could provide information on the efficacy of drug treatment. In this context, clinical research, along with preclinical research on genetic and neurodevelopmental models, could prove extremely useful.

Author Contributions

Conceptualization, M.M. (Martina Montanari), G.M., P.B. and M.M. (Maria Meringolo); formal analysis, P.B. and M.M. (Maria Meringolo); investigation, M.M. (Martina Montanari) and M.M. (Maria Meringolo); resources, P.B.; data curation, M.M. (Martina Montanari) and M.M. (Maria Meringolo); writing—original draft preparation, M.M. (Martina Montanari) and M.Me; writing—review and editing, M.M. (Martina Montanari), M.M. (Maria Meringolo) and P.B.; visualization, G.M.; supervision, P.B.; project administration, P.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yin, H.H.; Knowlton, B.J. The role of the basal ganglia in habit formation. Nat. Rev. Neurosci. 2006, 7, 464–476. [Google Scholar] [CrossRef] [PubMed]
  2. American Psychiatric Association. Diagnostic and Statistical Manual of Mental Disorders, 5th ed.; American Psychiatric Association: Washington, DC, USA, 2013; ISBN 0-89042-555-8. [Google Scholar]
  3. Fetit, R.; Hillary, R.F.; Price, D.J.; Lawrie, S.M. The neuropathology of autism: A systematic review of post-mortem studies of autism and related disorders. Neurosci. Biobehav. Rev. 2021, 129, 35–62. [Google Scholar] [CrossRef] [PubMed]
  4. Constantino, J.N.; Abbacchi, A.M.; Saulnier, C.; Klaiman, C.; Mandell, D.S.; Zhang, Y.; Hawks, Z.; Bates, J.; Klin, A.; Shattuck, P.; et al. Timing of the diagnosis of autism in African American children. Pediatrics 2020, 146, e20193629. [Google Scholar] [CrossRef]
  5. Marotta, R.; Risoleo, M.C.; Messina, G.; Parisi, L.; Carotenuto, M.; Vetri, L.; Roccella, M. The neurochemistry of autism. Brain Sci. 2020, 10, 163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Pickles, A.; McCauley, J.B.; Pepa, L.A.; Huerta, M.; Lord, C. The adult outcome of children referred for autism: Typology and prediction from childhood. J. Child Psychol. Psychiatry 2020, 61, 760–767. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Carroll, L.; Braeutigam, S.; Dawes, J.M.; Krsnik, Z.; Kostovic, I.; Coutinho, E.; Dewing, J.M.; Horton, C.A.; Gomez-Nicola, D.; Menassa, D.A. Autism spectrum disorders: Multiple routes to, and multiple consequences of, abnormal synaptic function and connectivity. Neuroscientist 2021, 27, 10–29. [Google Scholar] [CrossRef] [PubMed]
  8. Rojas, D.C. The role of glutamate and its receptors in autism and the use of glutamate receptor antagonists in treatment. J. Neural Transm. 2014, 121, 891–905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Carlson, G.C. Glutamate receptor dysfunction and drug targets across models of autism spectrum disorders. Pharmacol. Biochem. Behav. 2012, 100, 850–854. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Choudhury, P.R.; Lahiri, S.; Rajamma, U. Glutamate mediated signaling in the pathophysiology of autism spectrum disorders. Pharmacol. Biochem. Behav. 2012, 100, 841–849. [Google Scholar] [CrossRef] [PubMed]
  11. Wickens, M.M.; Bangasser, D.A.; Briand, L.A. Sex differences in psychiatric disease: A focus on the glutamate system. Front. Mol. Neurosci. 2018, 11, 197. [Google Scholar] [CrossRef] [PubMed]
  12. Basu, S.K.; Pradhan, S.; du Plessis, A.J.; Ben-Ari, Y.; Limperopoulos, C. GABA and glutamate in the preterm neonatal brain: In-vivo measurement by magnetic resonance spectroscopy. Neuroimage 2021, 238, 118215. [Google Scholar] [CrossRef] [PubMed]
  13. Guimaraes, I.M.; Carvalho, T.G.; Ferguson, S.S.; Pereira, G.S.; Ribeiro, F.M. The metabotropic glutamate receptor 5 role on motor behavior involves specific neural substrates. Mol. Brain 2015, 8, 24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Gécz, J. Glutamate receptors and learning and memory. Nat. Genet. 2010, 42, 925–926. [Google Scholar] [CrossRef] [PubMed]
  15. Zoicas, I.; Kornhuber, J. The role of metabotropic glutamate receptors in social behavior in rodents. Int. J. Mol. Sci. 2019, 20, 1412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Tomé, D. The roles of dietary glutamate in the intestine. Ann. Nutr. Metab. 2018, 73 (Suppl. 5), 15–20. [Google Scholar] [CrossRef] [PubMed]
  17. Kirchgessner, A.L. Glutamate in the enteric nervous system. Curr. Opin. Pharmacol. 2001, 1, 591–596. [Google Scholar] [CrossRef]
  18. Aldred, S.; Moore, K.M.; Fitzgerald, M.; Waring, R.H. Plasma amino acid levels in children with autism and their families. J. Autism Dev. Disord. 2003, 33, 93–97. [Google Scholar] [CrossRef] [PubMed]
  19. Fatemi, S.H. The hyperglutamatergic hypothesis of autism. Prog. Neuropsychopharmacol. Biol. Psychiatry 2008, 32, 912–913. [Google Scholar] [CrossRef] [PubMed]
  20. Serajee, F.J.; Zhong, H.; Nabi, R.; Huq, A.H.M.M. The metabotropic glutamate receptor 8 gene at 7q31: Partial duplication and possible association with autism. J. Med. Genet. 2003, 40, e42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Jamain, S.; Betancur, C.; Quach, H.; Philippe, A.; Fellous, M.; Giros, B.; Gillberg, C.; Leboyer, M.; Bourgeron, T. Paris Autism Research International Sibpair (PARIS) Study Linkage and association of the glutamate receptor 6 gene with autism. Mol. Psychiatry 2002, 7, 302–310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Purcell, A.E.; Jeon, O.H.; Zimmerman, A.W.; Blue, M.E.; Pevsner, J. Postmortem brain abnormalities of the glutamate neurotransmitter system in autism. Neurology 2001, 57, 1618–1628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Blandini, F. An update on the potential role of excitotoxicity in the pathogenesis of Parkinson’s disease. Funct. Neurol. 2010, 25, 65–71. [Google Scholar] [PubMed]
  24. Bozzi, Y.; Provenzano, G.; Casarosa, S. Neurobiological bases of autism-epilepsy comorbidity: A focus on excitation/inhibition imbalance. Eur. J. Neurosci. 2018, 47, 534–548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Bolton, P.F.; Carcani-Rathwell, I.; Hutton, J.; Goode, S.; Howlin, P.; Rutter, M. Epilepsy in autism: Features and correlates. Br. J. Psychiatry 2011, 198, 289–294. [Google Scholar] [CrossRef]
  26. Lee, B.H.; Smith, T.; Paciorkowski, A.R. Autism spectrum disorder and epilepsy: Disorders with a shared biology. Epilepsy Behav. 2015, 47, 191–201. [Google Scholar] [CrossRef] [Green Version]
  27. Carlsson, M.L. Hypothesis: Is infantile autism a hypoglutamatergic disorder? Relevance of glutamate–serotonin interactions for pharmacotherapy. J. Neural Transm. 1998, 105, 525–535. [Google Scholar] [CrossRef]
  28. Comery, T.A.; Harris, J.B.; Willems, P.J.; Oostra, B.A.; Irwin, S.A.; Weiler, I.J.; Greenough, W.T. Abnormal dendritic spines in fragile X knockout mice: Maturation and pruning deficits. Proc. Natl. Acad. Sci. USA 1997, 94, 5401–5404. [Google Scholar] [CrossRef] [Green Version]
  29. Südhof, T.C.; Malenka, R.C. Understanding synapses: Past, present, and future. Neuron 2008, 60, 469–476. [Google Scholar] [CrossRef] [Green Version]
  30. Fedder, K.N.; Sabo, S.L. On the role of glutamate in presynaptic development: Possible contributions of presynaptic NMDA receptors. Biomolecules 2015, 5, 3448–3466. [Google Scholar] [CrossRef] [Green Version]
  31. Nicoletti, F.; Bockaert, J.; Collingridge, G.L.; Conn, P.J.; Ferraguti, F.; Schoepp, D.D.; Wroblewski, J.T.; Pin, J.P. Metabotropic glutamate receptors: From the workbench to the bedside. Neuropharmacology 2011, 60, 1017–1041. [Google Scholar] [CrossRef] [Green Version]
  32. Catania, M.V.; D’Antoni, S.; Bonaccorso, C.M.; Aronica, E.; Bear, M.F.; Nicoletti, F. Group I metabotropic glutamate receptors: A role in neurodevelopmental disorders? Mol. Neurobiol. 2007, 35, 298–307. [Google Scholar] [CrossRef] [PubMed]
  33. Dong, X.; Wang, Y.; Qin, Z. Molecular mechanisms of excitotoxicity and their relevance to pathogenesis of neurodegenerative diseases. Acta Pharmacol. Sin. 2009, 30, 379–387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Egbenya, D.L.; Aidoo, E.; Kyei, G. Glutamate receptors in brain development. Childs Nerv. Syst. 2021, 37, 2753–2758. [Google Scholar] [CrossRef] [PubMed]
  35. Maragakis, N.J.; Rothstein, J.D. Glutamate transporters: Animal models to neurologic disease. Neurobiol. Dis. 2004, 15, 461–473. [Google Scholar] [CrossRef]
  36. Behuet, S.; Cremer, J.N.; Cremer, M.; Palomero-Gallagher, N.; Zilles, K.; Amunts, K. Developmental changes of glutamate and GABA receptor densities in wistar rats. Front. Neuroanat. 2019, 13, 100. [Google Scholar] [CrossRef]
  37. Pellegrini-Giampietro, D.E.; Gorter, J.A.; Bennett, M.V.; Zukin, R.S. The GluR2 (GluR-B) hypothesis: Ca(2+)-permeable AMPA receptors in neurological disorders. Trends Neurosci. 1997, 20, 464–470. [Google Scholar] [CrossRef]
  38. Pandey, S.P.; Rai, R.; Gaur, P.; Prasad, S. Development- and age-related alterations in the expression of AMPA receptor subunit GluR2 and its trafficking proteins in the hippocampus of male mouse brain. Biogerontology 2015, 16, 317–328. [Google Scholar] [CrossRef]
  39. Paupard, M.C.; Friedman, L.K.; Zukin, R.S. Developmental regulation and cell-specific expression of N-methyl-D-aspartate receptor splice variants in rat hippocampus. Neuroscience 1997, 79, 399–409. [Google Scholar] [CrossRef]
  40. Waters, K.A.; Machaalani, R. NMDA receptors in the developing brain and effects of noxious insults. Neurosignals 2004, 13, 162–174. [Google Scholar] [CrossRef]
  41. Szczurowska, E.; Mareš, P. NMDA and AMPA receptors: Development and status epilepticus. Physiol. Res. 2013, 62 (Suppl. 1), S21–S38. [Google Scholar] [CrossRef]
  42. Strunecka, A.; Blaylock, R.L.; Patocka, J.; Strunecky, O. Immunoexcitotoxicity as the central mechanism of etiopathology and treatment of autism spectrum disorders: A possible role of fluoride and aluminum. Surg. Neurol. Int. 2018, 9, 74. [Google Scholar] [CrossRef]
  43. Holopainen, I.E.; Laurén, H.B. Glutamate signaling in the pathophysiology and therapy of prenatal insults. Pharmacol. Biochem. Behav. 2012, 100, 825–834. [Google Scholar] [CrossRef] [PubMed]
  44. Bahn, S.; Volk, B.; Wisden, W. Kainate receptor gene expression in the developing rat brain. J. Neurosci. 1994, 14, 5525–5547. [Google Scholar] [CrossRef]
  45. Lauri, S.E.; Segerstråle, M.; Vesikansa, A.; Maingret, F.; Mulle, C.; Collingridge, G.L.; Isaac, J.T.R.; Taira, T. Endogenous activation of kainate receptors regulates glutamate release and network activity in the developing hippocampus. J. Neurosci. 2005, 25, 4473–4484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Jansson, L.C.; Åkerman, K.E. The role of glutamate and its receptors in the proliferation, migration, differentiation and survival of neural progenitor cells. J. Neural Transm. 2014, 121, 819–836. [Google Scholar] [CrossRef] [PubMed]
  47. Chaudhury, S.; Sharma, V.; Kumar, V.; Nag, T.C.; Wadhwa, S. Activity-dependent synaptic plasticity modulates the critical phase of brain development. Brain Dev. 2016, 38, 355–363. [Google Scholar] [CrossRef]
  48. Holtmaat, A.; Svoboda, K. Experience-dependent structural synaptic plasticity in the mammalian brain. Nat. Rev. Neurosci. 2009, 10, 647–658. [Google Scholar] [CrossRef] [PubMed]
  49. Gaiarsa, J.-L.; Caillard, O.; Ben-Ari, Y. Long-term plasticity at GABAergic and glycinergic synapses: Mechanisms and functional significance. Trends Neurosci. 2002, 25, 564–570. [Google Scholar] [CrossRef]
  50. Bak, L.K.; Schousboe, A.; Waagepetersen, H.S. The glutamate/GABA-glutamine cycle: Aspects of transport, neurotransmitter homeostasis and ammonia transfer. J. Neurochem. 2006, 98, 641–653. [Google Scholar] [CrossRef]
  51. Kim, S.; Kwon, S.-H.; Kam, T.-I.; Panicker, N.; Karuppagounder, S.S.; Lee, S.; Lee, J.H.; Kim, W.R.; Kook, M.; Foss, C.A.; et al. Transneuronal Propagation of Pathologic α-Synuclein from the Gut to the Brain Models Parkinson’s Disease. Neuron 2019, 103, 627–641.e7. [Google Scholar] [CrossRef]
  52. Kim, J.-W.; Park, K.; Kang, R.J.; Gonzales, E.L.T.; Kim, D.G.; Oh, H.A.; Seung, H.; Ko, M.J.; Kwon, K.J.; Kim, K.C.; et al. Pharmacological modulation of AMPA receptor rescues social impairments in animal models of autism. Neuropsychopharmacology 2019, 44, 314–323. [Google Scholar] [CrossRef] [Green Version]
  53. McCarthy, M.M.; Davis, A.M.; Mong, J.A. Excitatory neurotransmission and sexual differentiation of the brain. Brain Res. Bull. 1997, 44, 487–495. [Google Scholar] [CrossRef]
  54. Giacometti, L.L.; Barker, J.M. Sex differences in the glutamate system: Implications for addiction. Neurosci. Biobehav. Rev. 2020, 113, 157–168. [Google Scholar] [CrossRef] [PubMed]
  55. Sailasuta, N.; Ernst, T.; Chang, L. Regional variations and the effects of age and gender on glutamate concentrations in the human brain. Magn. Reson. Imaging 2008, 26, 667–675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Grachev, I.D.; Apkarian, A.V. Chemical heterogeneity of the living human brain: A proton MR spectroscopy study on the effects of sex, age, and brain region. Neuroimage 2000, 11, 554–563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Zahr, N.M.; Mayer, D.; Rohlfing, T.; Chanraud, S.; Gu, M.; Sullivan, E.V.; Pfefferbaum, A. In vivo glutamate measured with magnetic resonance spectroscopy: Behavioral correlates in aging. Neurobiol. Aging 2013, 34, 1265–1276. [Google Scholar] [CrossRef] [Green Version]
  58. O’Gorman, R.L.; Michels, L.; Edden, R.A.; Murdoch, J.B.; Martin, E. In vivo detection of GABA and glutamate with MEGA-PRESS: Reproducibility and gender effects. J. Magn. Reson. Imaging 2011, 33, 1262–1267. [Google Scholar] [CrossRef] [Green Version]
  59. Zlotnik, A.; Gruenbaum, B.F.; Mohar, B.; Kuts, R.; Gruenbaum, S.E.; Ohayon, S.; Boyko, M.; Klin, Y.; Sheiner, E.; Shaked, G.; et al. The effects of estrogen and progesterone on blood glutamate levels: Evidence from changes of blood glutamate levels during the menstrual cycle in women. Biol. Reprod. 2011, 84, 581–586. [Google Scholar] [CrossRef] [PubMed]
  60. Hädel, S.; Wirth, C.; Rapp, M.; Gallinat, J.; Schubert, F. Effects of age and sex on the concentrations of glutamate and glutamine in the human brain. J. Magn. Reson. Imaging 2013, 38, 1480–1487. [Google Scholar] [CrossRef] [PubMed]
  61. Kouchiwa, T.; Wada, K.; Uchiyama, M.; Kasezawa, N.; Niisato, M.; Murakami, H.; Fukuyama, K.; Yokogoshi, H. Age-related changes in serum amino acids concentrations in healthy individuals. Clin. Chem. Lab. Med. 2012, 50, 861–870. [Google Scholar] [CrossRef] [PubMed]
  62. Frankfurt, M.; Fuchs, E.; Wuttke, W. Sex differences in γ-aminobutyric acid and glutamate concentrations in discrete rat brain nuclei. Neurosci. Lett. 1984, 50, 245–250. [Google Scholar] [CrossRef]
  63. Zuena, A.R.; Mairesse, J.; Casolini, P.; Cinque, C.; Alemà, G.S.; Morley-Fletcher, S.; Chiodi, V.; Spagnoli, L.G.; Gradini, R.; Catalani, A.; et al. Prenatal restraint stress generates two distinct behavioral and neurochemical profiles in male and female rats. PLoS ONE 2008, 3, e2170. [Google Scholar] [CrossRef] [Green Version]
  64. Anderson, D.W.; Mettil, W.A.; Schneider, J.S. Rearing environment, sex and developmental lead exposure modify gene expression in the hippocampus of behaviorally naïve animals. Neurochem. Int. 2013, 62, 510–520. [Google Scholar] [CrossRef] [Green Version]
  65. Higuera-Matas, A.; Miguéns, M.; Coria, S.M.; Assis, M.A.; Borcel, E.; del Olmo, N.; Ambrosio, E. Sex-specific disturbances of the glutamate/GABA balance in the hippocampus of adult rats subjected to adolescent cannabinoid exposure. Neuropharmacology 2012, 62, 1975–1984. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, Y.; Ma, Y.; Hu, J.; Cheng, W.; Jiang, H.; Zhang, X.; Li, M.; Ren, J.; Li, X. Prenatal chronic mild stress induces depression-like behavior and sex-specific changes in regional glutamate receptor expression patterns in adult rats. Neuroscience 2015, 301, 363–374. [Google Scholar] [CrossRef] [PubMed]
  67. Monfort, P.; Gomez-Gimenez, B.; Llansola, M.; Felipo, V. Gender differences in spatial learning, synaptic activity, and long-term potentiation in the hippocampus in rats: Molecular mechanisms. ACS Chem. Neurosci. 2015, 6, 1420–1427. [Google Scholar] [CrossRef] [PubMed]
  68. Zhou, Y.; Danbolt, N.C. Glutamate as a neurotransmitter in the healthy brain. J. Neural Transm. 2014, 121, 799–817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Hawkins, R.A. The blood-brain barrier and glutamate. Am. J. Clin. Nutr. 2009, 90, 867S–874S. [Google Scholar] [CrossRef] [Green Version]
  70. Hawkins, R.A.; Viña, J.R. How Glutamate Is Managed by the Blood-Brain Barrier. Biology 2016, 5, 37. [Google Scholar] [CrossRef] [PubMed]
  71. Moussawi, K.; Riegel, A.; Nair, S.; Kalivas, P.W. Extracellular glutamate: Functional compartments operate in different concentration ranges. Front. Syst. Neurosci. 2011, 5, 94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Ghanizadeh, A. Increased glutamate and homocysteine and decreased glutamine levels in autism: A review and strategies for future studies of amino acids in autism. Dis. Markers 2013, 35, 281–286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Zheng, Z.; Zhu, T.; Qu, Y.; Mu, D. Blood Glutamate Levels in Autism Spectrum Disorder: A Systematic Review and Meta-Analysis. PLoS ONE 2016, 11, e0158688. [Google Scholar] [CrossRef]
  74. Moreno-Fuenmayor, H.; Borjas, L.; Arrieta, A.; Valera, V.; Socorro-Candanoza, L. Plasma excitatory amino acids in autism. Investig. Clin. 1996, 37, 113–128. [Google Scholar]
  75. Moreno, H.; Borjas, L.; Arrieta, A.; Sáez, L.; Prassad, A.; Estévez, J.; Bonilla, E. Clinical heterogeneity of the autistic syndrome: A study of 60 families. Investig. Clin. 1992, 33, 13–31. [Google Scholar]
  76. Shinohe, A.; Hashimoto, K.; Nakamura, K.; Tsujii, M.; Iwata, Y.; Tsuchiya, K.J.; Sekine, Y.; Suda, S.; Suzuki, K.; Sugihara, G.-I.; et al. Increased serum levels of glutamate in adult patients with autism. Prog. Neuropsychopharmacol. Biol. Psychiatry 2006, 30, 1472–1477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Shimmura, C.; Suda, S.; Tsuchiya, K.J.; Hashimoto, K.; Ohno, K.; Matsuzaki, H.; Iwata, K.; Matsumoto, K.; Wakuda, T.; Kameno, Y.; et al. Alteration of plasma glutamate and glutamine levels in children with high-functioning autism. PLoS ONE 2011, 6, e25340. [Google Scholar] [CrossRef] [Green Version]
  78. Tirouvanziam, R.; Obukhanych, T.V.; Laval, J.; Aronov, P.A.; Libove, R.; Banerjee, A.G.; Parker, K.J.; O’Hara, R.; Herzenberg, L.A.; Herzenberg, L.A.; et al. Distinct plasma profile of polar neutral amino acids, leucine, and glutamate in children with Autism Spectrum Disorders. J. Autism Dev. Disord. 2012, 42, 827–836. [Google Scholar] [CrossRef] [PubMed]
  79. Tu, W.-J.; Chen, H.; He, J. Application of LC-MS/MS analysis of plasma amino acids profiles in children with autism. J. Clin. Biochem. Nutr. 2012, 51, 248–249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Naushad, S.M.; Jain, J.M.N.; Prasad, C.K.; Naik, U.; Akella, R.R.D. Autistic children exhibit distinct plasma amino acid profile. Indian J. Biochem. Biophys. 2013, 50, 474–478. [Google Scholar] [PubMed]
  81. El-Ansary, A.; Al-Ayadhi, L. GABAergic/glutamatergic imbalance relative to excessive neuroinflammation in autism spectrum disorders. J. Neuroinflamm. 2014, 11, 189. [Google Scholar] [CrossRef]
  82. Cai, J.; Ding, L.; Zhang, J.-S.; Xue, J.; Wang, L.-Z. Elevated plasma levels of glutamate in children with autism spectrum disorders. Neuroreport 2016, 27, 272–276. [Google Scholar] [CrossRef]
  83. Zaki, M.M.; Abdel-Al, H.; Al-Sawi, M. Assessment of plasma amino acid profile in autism using cation-exchange chromatography with postcolumn derivatization by ninhydrin. Turk. J. Med. Sci. 2017, 47, 260–267. [Google Scholar] [CrossRef]
  84. Vargason, T.; Kruger, U.; McGuinness, D.L.; Adams, J.B.; Geis, E.; Gehn, E.; Coleman, D.; Hahn, J. Investigating Plasma Amino Acids for Differentiating Individuals with Autism Spectrum Disorder and Typically Developing Peers. Res. Autism Spectr. Disord. 2018, 50, 60–72. [Google Scholar] [CrossRef]
  85. Al-Otaish, H.; Al-Ayadhi, L.; Bjørklund, G.; Chirumbolo, S.; Urbina, M.A.; El-Ansary, A. Relationship between absolute and relative ratios of glutamate, glutamine and GABA and severity of autism spectrum disorder. Metab. Brain Dis. 2018, 33, 843–854. [Google Scholar] [CrossRef]
  86. Khalifa, D.; Shahin, O.; Salem, D.; Raafat, O. Serum glutamate was elevated in children aged 3-10 years with autism spectrum disorders when they were compared with controls. Acta Paediatr. 2019, 108, 295–299. [Google Scholar] [CrossRef] [Green Version]
  87. Yu, X.; Qian-Qian, L.; Cong, Y.; Xiao-Bing, Z.; Hong-Zhu, D. Reduction of essential amino acid levels and sex-specific alterations in serum amino acid concentration profiles in children with autism spectrum disorder. Psychiatry Res. 2021, 297, 113675. [Google Scholar] [CrossRef]
  88. Takado, Y.; Sato, N.; Kanbe, Y.; Tomiyasu, M.; Xin, L.; Near, J.; Yoshikawa, K.; Sahara, N.; Higashi, T.; Suhara, T.; et al. Association between Brain and Plasma Glutamine Levels in Healthy Young Subjects Investigated by MRS and LC/MS. Nutrients 2019, 11, 1649. [Google Scholar] [CrossRef] [Green Version]
  89. Andersen, J.V.; Markussen, K.H.; Jakobsen, E.; Schousboe, A.; Waagepetersen, H.S.; Rosenberg, P.A.; Aldana, B.I. Glutamate metabolism and recycling at the excitatory synapse in health and neurodegeneration. Neuropharmacology 2021, 196, 108719. [Google Scholar] [CrossRef]
  90. Page, L.A.; Daly, E.; Schmitz, N.; Simmons, A.; Toal, F.; Deeley, Q.; Ambery, F.; McAlonan, G.M.; Murphy, K.C.; Murphy, D.G.M. In vivo 1H-magnetic resonance spectroscopy study of amygdala-hippocampal and parietal regions in autism. Am. J. Psychiatry 2006, 163, 2189–2192. [Google Scholar] [CrossRef]
  91. He, J.L.; Oeltzschner, G.; Mikkelsen, M.; Deronda, A.; Harris, A.D.; Crocetti, D.; Wodka, E.L.; Mostofsky, S.H.; Edden, R.A.E.; Puts, N.A.J. Region-specific elevations of glutamate + glutamine correlate with the sensory symptoms of autism spectrum disorders. Transl. Psychiatry 2021, 11, 411. [Google Scholar] [CrossRef]
  92. Bejjani, A.; O’Neill, J.; Kim, J.A.; Frew, A.J.; Yee, V.W.; Ly, R.; Kitchen, C.; Salamon, N.; McCracken, J.T.; Toga, A.W.; et al. Elevated glutamatergic compounds in pregenual anterior cingulate in pediatric autism spectrum disorder demonstrated by 1H MRS and 1H MRSI. PLoS ONE 2012, 7, e38786. [Google Scholar] [CrossRef]
  93. Joshi, G.; Biederman, J.; Wozniak, J.; Goldin, R.L.; Crowley, D.; Furtak, S.; Lukas, S.E.; Gönenç, A. Magnetic resonance spectroscopy study of the glutamatergic system in adolescent males with high-functioning autistic disorder: A pilot study at 4T. Eur. Arch. Psychiatry Clin. Neurosci. 2013, 263, 379–384. [Google Scholar] [CrossRef]
  94. Jiménez-Espinoza, C.; Marcano Serrano, F.; González-Mora, J.L. N-Acetylaspartyl-Glutamate Metabolism in the Cingulated Cortices as a Biomarker of the Etiology in ASD: A 1H-MRS Model. Molecules 2021, 26, 675. [Google Scholar] [CrossRef]
  95. Brown, M.S.; Singel, D.; Hepburn, S.; Rojas, D.C. Increased glutamate concentration in the auditory cortex of persons with autism and first-degree relatives: A (1)H-MRS study. Autism Res. 2013, 6, 1–10. [Google Scholar] [CrossRef] [Green Version]
  96. Siegel-Ramsay, J.E.; Romaniuk, L.; Whalley, H.C.; Roberts, N.; Branigan, H.; Stanfield, A.C.; Lawrie, S.M.; Dauvermann, M.R. Glutamate and functional connectivity—Support for the excitatory-inhibitory imbalance hypothesis in autism spectrum disorders. Psychiatry Res. Neuroimaging 2021, 313, 111302. [Google Scholar] [CrossRef]
  97. Bernardi, S.; Anagnostou, E.; Shen, J.; Kolevzon, A.; Buxbaum, J.D.; Hollander, E.; Hof, P.R.; Fan, J. In vivo 1H-magnetic resonance spectroscopy study of the attentional networks in autism. Brain Res. 2011, 1380, 198–205. [Google Scholar] [CrossRef] [Green Version]
  98. Tebartz van Elst, L.; Maier, S.; Fangmeier, T.; Endres, D.; Mueller, G.T.; Nickel, K.; Ebert, D.; Lange, T.; Hennig, J.; Biscaldi, M.; et al. Disturbed cingulate glutamate metabolism in adults with high-functioning autism spectrum disorder: Evidence in support of the excitatory/inhibitory imbalance hypothesis. Mol. Psychiatry 2014, 19, 1314–1325. [Google Scholar] [CrossRef] [Green Version]
  99. Hollestein, V.; Buitelaar, J.K.; Brandeis, D.; Banaschewski, T.; Kaiser, A.; Hohmann, S.; Oranje, B.; Gooskens, B.; Durston, S.; Williams, S.C.R.; et al. Developmental changes in fronto-striatal glutamate and their association with functioning during inhibitory control in autism spectrum disorder and obsessive compulsive disorder. Neuroimage Clin. 2021, 30, 102622. [Google Scholar] [CrossRef]
  100. Horder, J.; Lavender, T.; Mendez, M.A.; O’Gorman, R.; Daly, E.; Craig, M.C.; Lythgoe, D.J.; Barker, G.J.; Murphy, D.G. Reduced subcortical glutamate/glutamine in adults with autism spectrum disorders: A [(1)H]MRS study. Transl. Psychiatry 2014, 4, e364. [Google Scholar] [CrossRef] [Green Version]
  101. DeVito, T.J.; Drost, D.J.; Neufeld, R.W.J.; Rajakumar, N.; Pavlosky, W.; Williamson, P.; Nicolson, R. Evidence for cortical dysfunction in autism: A proton magnetic resonance spectroscopic imaging study. Biol. Psychiatry 2007, 61, 465–473. [Google Scholar] [CrossRef]
  102. He, H.-Y.; Cline, H.T. What is excitation/inhibition and how is it regulated? A case of the elephant and the wisemen. J. Exp. Neurosci. 2019, 13, 1179069519859371. [Google Scholar] [CrossRef] [Green Version]
  103. Essa, M.M.; Braidy, N.; Vijayan, K.R.; Subash, S.; Guillemin, G.J. Excitotoxicity in the pathogenesis of autism. Neurotox. Res. 2013, 23, 393–400. [Google Scholar] [CrossRef]
  104. Viviani, B.; Boraso, M.; Marchetti, N.; Marinovich, M. Perspectives on neuroinflammation and excitotoxicity: A neurotoxic conspiracy? Neurotoxicology 2014, 43, 10–20. [Google Scholar] [CrossRef]
  105. Pearson, V.L.; Rothwell, N.J.; Toulmond, S. Excitotoxic brain damage in the rat induces interleukin-1beta protein in microglia and astrocytes: Correlation with the progression of cell death. Glia 1999, 25, 311–323. [Google Scholar] [CrossRef]
  106. Jander, S.; Schroeter, M.; Stoll, G. Role of NMDA receptor signaling in the regulation of inflammatory gene expression after focal brain ischemia. J. Neuroimmunol. 2000, 109, 181–187. [Google Scholar] [CrossRef]
  107. Relton, J.K.; Rothwell, N.J. Interleukin-1 receptor antagonist inhibits ischaemic and excitotoxic neuronal damage in the rat. Brain Res. Bull. 1992, 29, 243–246. [Google Scholar] [CrossRef]
  108. Biber, K.; Neumann, H.; Inoue, K.; Boddeke, H.W.G.M. Neuronal “On” and “Off” signals control microglia. Trends Neurosci. 2007, 30, 596–602. [Google Scholar] [CrossRef]
  109. Mandolesi, G.; Musella, A.; Gentile, A.; Grasselli, G.; Haji, N.; Sepman, H.; Fresegna, D.; Bullitta, S.; De Vito, F.; Musumeci, G.; et al. Interleukin-1β alters glutamate transmission at purkinje cell synapses in a mouse model of multiple sclerosis. J. Neurosci. 2013, 33, 12105–12121. [Google Scholar] [CrossRef]
  110. Fogal, B.; Li, J.; Lobner, D.; McCullough, L.D.; Hewett, S.J. System x(c)-activity and astrocytes are necessary for interleukin-1 beta-mediated hypoxic neuronal injury. J. Neurosci. 2007, 27, 10094–10105. [Google Scholar] [CrossRef] [Green Version]
  111. Bezzi, P.; Domercq, M.; Brambilla, L.; Galli, R.; Schols, D.; De Clercq, E.; Vescovi, A.; Bagetta, G.; Kollias, G.; Meldolesi, J.; et al. CXCR4-activated astrocyte glutamate release via TNFalpha: Amplification by microglia triggers neurotoxicity. Nat. Neurosci. 2001, 4, 702–710. [Google Scholar] [CrossRef]
  112. Stellwagen, D.; Beattie, E.C.; Seo, J.Y.; Malenka, R.C. Differential regulation of AMPA receptor and GABA receptor trafficking by tumor necrosis factor-alpha. J. Neurosci. 2005, 25, 3219–3228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Pribiag, H.; Stellwagen, D. TNF-α downregulates inhibitory neurotransmission through protein phosphatase 1-dependent trafficking of GABA(A) receptors. J. Neurosci. 2013, 33, 15879–15893. [Google Scholar] [CrossRef]
  114. Viviani, B.; Bartesaghi, S.; Gardoni, F.; Vezzani, A.; Behrens, M.M.; Bartfai, T.; Binaglia, M.; Corsini, E.; Di Luca, M.; Galli, C.L.; et al. Interleukin-1β Enhances NMDA Receptor-Mediated Intracellular Calcium Increase through Activation of the Src Family of Kinases. J. Neurosci. 2003, 23, 8692–8700. [Google Scholar] [CrossRef] [PubMed]
  115. Wang, S.; Cheng, Q.; Malik, S.; Yang, J. Interleukin-1beta inhibits gamma-aminobutyric acid type A (GABA(A)) receptor current in cultured hippocampal neurons. J. Pharmacol. Exp. Ther. 2000, 292, 497–504. [Google Scholar] [PubMed]
  116. Lai, A.Y.; Swayze, R.D.; El-Husseini, A.; Song, C. Interleukin-1 beta modulates AMPA receptor expression and phosphorylation in hippocampal neurons. J. Neuroimmunol. 2006, 175, 97–106. [Google Scholar] [CrossRef]
  117. Xu, S.; Han, J.C.; Morales, A.; Menzie, C.M.; Williams, K.; Fan, Y.S. Characterization of 11p14-p12 deletion in WAGR syndrome by array CGH for identifying genes contributing to mental retardation and autism. Cytogenet. Genome Res. 2008, 122, 181–187. [Google Scholar] [CrossRef]
  118. Ruzzo, E.K.; Pérez-Cano, L.; Jung, J.-Y.; Wang, L.-K.; Kashef-Haghighi, D.; Hartl, C.; Singh, C.; Xu, J.; Hoekstra, J.N.; Leventhal, O.; et al. Inherited and de novo genetic risk for autism impacts shared networks. Cell 2019, 178, 850–866.e26. [Google Scholar] [CrossRef] [Green Version]
  119. Yuen, E.Y.; Wei, J.; Yan, Z. Molecular and epigenetic mechanisms for the complex effects of stress on synaptic physiology and cognitive functions. Int. J. Neuropsychopharmacol. 2017, 20, 948–955. [Google Scholar] [CrossRef] [Green Version]
  120. Satterstrom, F.K.; Walters, R.K.; Singh, T.; Wigdor, E.M.; Lescai, F.; Demontis, D.; Kosmicki, J.A.; Grove, J.; Stevens, C.; Bybjerg-Grauholm, J.; et al. Autism spectrum disorder and attention deficit hyperactivity disorder have a similar burden of rare protein-truncating variants. Nat. Neurosci. 2019, 22, 1961–1965. [Google Scholar] [CrossRef]
  121. De Rubeis, S.; He, X.; Goldberg, A.P.; Poultney, C.S.; Samocha, K.; Cicek, A.E.; Kou, Y.; Liu, L.; Fromer, M.; Walker, S.; et al. Synaptic, transcriptional and chromatin genes disrupted in autism. Nature 2014, 515, 209–215. [Google Scholar] [CrossRef]
  122. Iossifov, I.; Levy, D.; Allen, J.; Ye, K.; Ronemus, M.; Lee, Y.-H.; Yamrom, B.; Wigler, M. Low load for disruptive mutations in autism genes and their biased transmission. Proc. Natl. Acad. Sci. USA 2015, 112, E5600–E5607. [Google Scholar] [CrossRef] [Green Version]
  123. Barnby, G.; Abbott, A.; Sykes, N.; Morris, A.; Weeks, D.E.; Mott, R.; Lamb, J.; Bailey, A.J.; Monaco, A.P. International Molecular Genetics Study of Autism Consortium Candidate-gene screening and association analysis at the autism-susceptibility locus on chromosome 16p: Evidence of association at GRIN2A and ABAT. Am. J. Hum. Genet. 2005, 76, 950–966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Yoo, H.J.; Cho, I.H.; Park, M.; Yang, S.Y.; Kim, S.A. Family based association of GRIN2A and GRIN2B with Korean autism spectrum disorders. Neurosci. Lett. 2012, 512, 89–93. [Google Scholar] [CrossRef] [PubMed]
  125. Salpietro, V.; Dixon, C.L.; Guo, H.; Bello, O.D.; Vandrovcova, J.; Efthymiou, S.; Maroofian, R.; Heimer, G.; Burglen, L.; Valence, S.; et al. AMPA receptor GluA2 subunit defects are a cause of neurodevelopmental disorders. Nat. Commun. 2019, 10, 3094. [Google Scholar] [CrossRef] [Green Version]
  126. Ramanathan, S.; Woodroffe, A.; Flodman, P.L.; Mays, L.Z.; Hanouni, M.; Modahl, C.B.; Steinberg-Epstein, R.; Bocian, M.E.; Spence, M.A.; Smith, M. A case of autism with an interstitial deletion on 4q leading to hemizygosity for genes encoding for glutamine and glycine neurotransmitter receptor sub-units (AMPA 2, GLRA3, GLRB) and neuropeptide receptors NPY1R, NPY5R. BMC Med. Genet. 2004, 5, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Myers, R.A.; Casals, F.; Gauthier, J.; Hamdan, F.F.; Keebler, J.; Boyko, A.R.; Bustamante, C.D.; Piton, A.M.; Spiegelman, D.; Henrion, E.; et al. A population genetic approach to mapping neurological disorder genes using deep resequencing. PLoS Genet. 2011, 7, e1001318. [Google Scholar] [CrossRef] [Green Version]
  128. Endele, S.; Rosenberger, G.; Geider, K.; Popp, B.; Tamer, C.; Stefanova, I.; Milh, M.; Kortüm, F.; Fritsch, A.; Pientka, F.K.; et al. Mutations in GRIN2A and GRIN2B encoding regulatory subunits of NMDA receptors cause variable neurodevelopmental phenotypes. Nat. Genet. 2010, 42, 1021–1026. [Google Scholar] [CrossRef]
  129. Tarabeux, J.; Kebir, O.; Gauthier, J.; Hamdan, F.F.; Xiong, L.; Piton, A.; Spiegelman, D.; Henrion, É.; Millet, B.; S2D team; et al. Rare mutations in N-methyl-D-aspartate glutamate receptors in autism spectrum disorders and schizophrenia. Transl. Psychiatry 2011, 1, e55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. O’Roak, B.J.; Deriziotis, P.; Lee, C.; Vives, L.; Schwartz, J.J.; Girirajan, S.; Karakoc, E.; Mackenzie, A.P.; Ng, S.B.; Baker, C.; et al. Exome sequencing in sporadic autism spectrum disorders identifies severe de novo mutations. Nat. Genet. 2011, 43, 585–589. [Google Scholar] [CrossRef] [PubMed]
  131. O’Roak, B.J.; Vives, L.; Fu, W.; Egertson, J.D.; Stanaway, I.B.; Phelps, I.G.; Carvill, G.; Kumar, A.; Lee, C.; Ankenman, K.; et al. Multiplex targeted sequencing identifies recurrently mutated genes in autism spectrum disorders. Science 2012, 338, 1619–1622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Satterstrom, F.K.; Kosmicki, J.A.; Wang, J.; Breen, M.S.; De Rubeis, S.; An, J.-Y.; Peng, M.; Collins, R.; Grove, J.; Klei, L.; et al. Large-Scale Exome Sequencing Study Implicates Both Developmental and Functional Changes in the Neurobiology of Autism. Cell 2020, 180, 568–584.e23. [Google Scholar] [CrossRef] [PubMed]
  133. Feliciano, P.; Zhou, X.; Astrovskaya, I.; Turner, T.N.; Wang, T.; Brueggeman, L.; Barnard, R.; Hsieh, A.; Snyder, L.G.; Muzny, D.M.; et al. Exome sequencing of 457 autism families recruited online provides evidence for autism risk genes. npj Genom. Med. 2019, 4, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Stessman, H.A.F.; Xiong, B.; Coe, B.P.; Wang, T.; Hoekzema, K.; Fenckova, M.; Kvarnung, M.; Gerdts, J.; Trinh, S.; Cosemans, N.; et al. Targeted sequencing identifies 91 neurodevelopmental-disorder risk genes with autism and developmental-disability biases. Nat. Genet. 2017, 49, 515–526. [Google Scholar] [CrossRef]
  135. Geisheker, M.R.; Heymann, G.; Wang, T.; Coe, B.P.; Turner, T.N.; Stessman, H.A.F.; Hoekzema, K.; Kvarnung, M.; Shaw, M.; Friend, K.; et al. Hotspots of missense mutation identify neurodevelopmental disorder genes and functional domains. Nat. Neurosci. 2017, 20, 1043–1051. [Google Scholar] [CrossRef] [Green Version]
  136. de Ligt, J.; Willemsen, M.H.; van Bon, B.W.M.; Kleefstra, T.; Yntema, H.G.; Kroes, T.; Vulto-van Silfhout, A.T.; Koolen, D.A.; de Vries, P.; Gilissen, C.; et al. Diagnostic exome sequencing in persons with severe intellectual disability. N. Engl. J. Med. 2012, 367, 1921–1929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Lesca, G.; Rudolf, G.; Bruneau, N.; Lozovaya, N.; Labalme, A.; Boutry-Kryza, N.; Salmi, M.; Tsintsadze, T.; Addis, L.; Motte, J.; et al. GRIN2A mutations in acquired epileptic aphasia and related childhood focal epilepsies and encephalopathies with speech and language dysfunction. Nat. Genet. 2013, 45, 1061–1066. [Google Scholar] [CrossRef] [PubMed]
  138. Lemke, J.R.; Lal, D.; Reinthaler, E.M.; Steiner, I.; Nothnagel, M.; Alber, M.; Geider, K.; Laube, B.; Schwake, M.; Finsterwalder, K.; et al. Mutations in GRIN2A cause idiopathic focal epilepsy with rolandic spikes. Nat. Genet. 2013, 45, 1067–1072. [Google Scholar] [CrossRef] [PubMed]
  139. Carvill, G.L.; Regan, B.M.; Yendle, S.C.; O’Roak, B.J.; Lozovaya, N.; Bruneau, N.; Burnashev, N.; Khan, A.; Cook, J.; Geraghty, E.; et al. GRIN2A mutations cause epilepsy-aphasia spectrum disorders. Nat. Genet. 2013, 45, 1073–1076. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. von Stülpnagel, C.; Ensslen, M.; Møller, R.S.; Pal, D.K.; Masnada, S.; Veggiotti, P.; Piazza, E.; Dreesmann, M.; Hartlieb, T.; Herberhold, T.; et al. Epilepsy in patients with GRIN2A alterations: Genetics, neurodevelopment, epileptic phenotype and response to anticonvulsive drugs. Eur. J. Paediatr. Neurol. 2017, 21, 530–541. [Google Scholar] [CrossRef] [PubMed]
  141. Addis, L.; Virdee, J.K.; Vidler, L.R.; Collier, D.A.; Pal, D.K.; Ursu, D. Epilepsy-associated GRIN2A mutations reduce NMDA receptor trafficking and agonist potency—Molecular profiling and functional rescue. Sci. Rep. 2017, 7, 66. [Google Scholar] [CrossRef] [Green Version]
  142. Gao, K.; Tankovic, A.; Zhang, Y.; Kusumoto, H.; Zhang, J.; Chen, W.; XiangWei, W.; Shaulsky, G.H.; Hu, C.; Traynelis, S.F.; et al. A de novo loss-of-function GRIN2A mutation associated with childhood focal epilepsy and acquired epileptic aphasia. PLoS ONE 2017, 12, e0170818. [Google Scholar] [CrossRef] [PubMed]
  143. Mangano, G.D.; Riva, A.; Fontana, A.; Salpietro, V.; Mangano, G.R.; Nobile, G.; Orsini, A.; Iacomino, M.; Battini, R.; Astrea, G.; et al. De novo GRIN2A variants associated with epilepsy and autism and literature review. Epilepsy Behav. 2022, 129, 108604. [Google Scholar] [CrossRef]
  144. Lee, J.; Ha, S.; Lee, S.-T.; Park, S.-G.; Shin, S.; Choi, J.R.; Cheon, K.-A. Next-Generation Sequencing in Korean Children with Autism Spectrum Disorder and Comorbid Epilepsy. Front. Pharmacol. 2020, 11, 585. [Google Scholar] [CrossRef] [PubMed]
  145. Yin, J.; Chun, C.-A.; Zavadenko, N.N.; Pechatnikova, N.L.; Naumova, O.Y.; Doddapaneni, H.V.; Hu, J.; Muzny, D.M.; Schaaf, C.P.; Grigorenko, E.L. Next Generation Sequencing of 134 Children with Autism Spectrum Disorder and Regression. Genes 2020, 11, 853. [Google Scholar] [CrossRef]
  146. Shuang, M.; Liu, J.; Jia, M.X.; Yang, J.Z.; Wu, S.P.; Gong, X.H.; Ling, Y.S.; Ruan, Y.; Yang, X.L.; Zhang, D. Family-based association study between autism and glutamate receptor 6 gene in Chinese Han trios. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2004, 131B, 48–50. [Google Scholar] [CrossRef]
  147. Krumm, N.; Turner, T.N.; Baker, C.; Vives, L.; Mohajeri, K.; Witherspoon, K.; Raja, A.; Coe, B.P.; Stessman, H.A.; He, Z.-X.; et al. Excess of rare, inherited truncating mutations in autism. Nat. Genet. 2015, 47, 582–588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Iossifov, I.; O’Roak, B.J.; Sanders, S.J.; Ronemus, M.; Krumm, N.; Levy, D.; Stessman, H.A.; Witherspoon, K.T.; Vives, L.; Patterson, K.E.; et al. The contribution of de novo coding mutations to autism spectrum disorder. Nature 2014, 515, 216–221. [Google Scholar] [CrossRef] [Green Version]
  149. Mejias, R.; Adamczyk, A.; Anggono, V.; Niranjan, T.; Thomas, G.M.; Sharma, K.; Skinner, C.; Schwartz, C.E.; Stevenson, R.E.; Fallin, M.D.; et al. Gain-of-function glutamate receptor interacting protein 1 variants alter GluA2 recycling and surface distribution in patients with autism. Proc. Natl. Acad. Sci. USA 2011, 108, 4920–4925. [Google Scholar] [CrossRef] [Green Version]
  150. Griswold, A.J.; Ma, D.; Cukier, H.N.; Nations, L.D.; Schmidt, M.A.; Chung, R.-H.; Jaworski, J.M.; Salyakina, D.; Konidari, I.; Whitehead, P.L.; et al. Evaluation of copy number variations reveals novel candidate genes in autism spectrum disorder-associated pathways. Hum. Mol. Genet. 2012, 21, 3513–3523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Glessner, J.T.; Wang, K.; Cai, G.; Korvatska, O.; Kim, C.E.; Wood, S.; Zhang, H.; Estes, A.; Brune, C.W.; Bradfield, J.P.; et al. Autism genome-wide copy number variation reveals ubiquitin and neuronal genes. Nature 2009, 459, 569–573. [Google Scholar] [CrossRef]
  152. Sangu, N.; Shimojima, K.; Takahashi, Y.; Ohashi, T.; Tohyama, J.; Yamamoto, T. A 7q31.33q32.1 microdeletion including LRRC4 and GRM8 is associated with severe intellectual disability and characteristics of autism. Hum. Gen. Var. 2017, 4, 17001. [Google Scholar] [CrossRef] [Green Version]
  153. Xu, X.; Li, C.; Gao, X.; Xia, K.; Guo, H.; Li, Y.; Hao, Z.; Zhang, L.; Gao, D.; Xu, C.; et al. Excessive UBE3A dosage impairs retinoic acid signaling and synaptic plasticity in autism spectrum disorders. Cell Res. 2018, 28, 48–68. [Google Scholar] [CrossRef]
  154. Salpietro, V.; Malintan, N.T.; Llano-Rivas, I.; Spaeth, C.G.; Efthymiou, S.; Striano, P.; Vandrovcova, J.; Cutrupi, M.C.; Chimenz, R.; David, E.; et al. Mutations in the neuronal vesicular SNARE VAMP2 affect synaptic membrane fusion and impair human neurodevelopment. Am. J. Hum. Genet. 2019, 104, 721–730. [Google Scholar] [CrossRef] [Green Version]
  155. Iossifov, I.; Ronemus, M.; Levy, D.; Wang, Z.; Hakker, I.; Rosenbaum, J.; Yamrom, B.; Lee, Y.-H.; Narzisi, G.; Leotta, A.; et al. De novo gene disruptions in children on the autistic spectrum. Neuron 2012, 74, 285–299. [Google Scholar] [CrossRef] [Green Version]
  156. Kelleher, R.J.; Geigenmüller, U.; Hovhannisyan, H.; Trautman, E.; Pinard, R.; Rathmell, B.; Carpenter, R.; Margulies, D. High-throughput sequencing of mGluR signaling pathway genes reveals enrichment of rare variants in autism. PLoS ONE 2012, 7, e35003. [Google Scholar] [CrossRef] [Green Version]
  157. Noroozi, R.; Taheri, M.; Movafagh, A.; Mirfakhraie, R.; Solgi, G.; Sayad, A.; Mazdeh, M.; Darvish, H. Glutamate receptor, metabotropic 7 (GRM7) gene variations and susceptibility to autism: A case-control study. Autism Res. 2016, 9, 1161–1168. [Google Scholar] [CrossRef]
  158. Liu, Y.; Zhang, Y.; Zhao, D.; Dong, R.; Yang, X.; Tammimies, K.; Uddin, M.; Scherer, S.W.; Gai, Z. Rare de novo deletion of metabotropic glutamate receptor 7 (GRM7) gene in a patient with autism spectrum disorder. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2015, 168B, 258–264. [Google Scholar] [CrossRef]
  159. Blatt, G.J.; Fitzgerald, C.M.; Guptill, J.T.; Booker, A.B.; Kemper, T.L.; Bauman, M.L. Density and distribution of hippocampal neurotransmitter receptors in autism: An autoradiographic study. J. Autism Dev. Disord. 2001, 31, 537–543. [Google Scholar] [CrossRef]
  160. Fatemi, S.H.; Folsom, T.D.; Kneeland, R.E.; Liesch, S.B. Metabotropic glutamate receptor 5 upregulation in children with autism is associated with underexpression of both Fragile X mental retardation protein and GABAA receptor beta 3 in adults with autism. Anat. Rec. 2011, 294, 1635–1645. [Google Scholar] [CrossRef] [Green Version]
  161. Lohith, T.G.; Osterweil, E.K.; Fujita, M.; Jenko, K.J.; Bear, M.F.; Innis, R.B. Is metabotropic glutamate receptor 5 upregulated in prefrontal cortex in fragile X syndrome? Mol. Autism 2013, 4, 15. [Google Scholar] [CrossRef] [Green Version]
  162. Fatemi, S.H.; Halt, A.R.; Stary, J.M.; Kanodia, R.; Schulz, S.C.; Realmuto, G.R. Glutamic acid decarboxylase 65 and 67 kDa proteins are reduced in autistic parietal and cerebellar cortices. Biol. Psychiatry 2002, 52, 805–810. [Google Scholar] [CrossRef]
  163. Bu, D.F.; Erlander, M.G.; Hitz, B.C.; Tillakaratne, N.J.; Kaufman, D.L.; Wagner-McPherson, C.B.; Evans, G.A.; Tobin, A.J. Two human glutamate decarboxylases, 65-kDa GAD and 67-kDa GAD, are each encoded by a single gene. Proc. Natl. Acad. Sci. USA 1992, 89, 2115–2119. [Google Scholar] [CrossRef] [Green Version]
  164. Rabionet, R.; Jaworski, J.M.; Ashley-Koch, A.E.; Martin, E.R.; Sutcliffe, J.S.; Haines, J.L.; Delong, G.R.; Abramson, R.K.; Wright, H.H.; Cuccaro, M.L.; et al. Analysis of the autism chromosome 2 linkage region: GAD1 and other candidate genes. Neurosci. Lett. 2004, 372, 209–214. [Google Scholar] [CrossRef]
  165. Yip, J.; Soghomonian, J.-J.; Blatt, G.J. Decreased GAD67 mRNA levels in cerebellar Purkinje cells in autism: Pathophysiological implications. Acta Neuropathol. 2007, 113, 559–568. [Google Scholar] [CrossRef]
  166. Kern, J.K. Purkinje cell vulnerability and autism: A possible etiological connection. Brain Dev. 2003, 25, 377–382. [Google Scholar] [CrossRef]
  167. Yip, J.; Soghomonian, J.J.; Blatt, G.J. Decreased GAD65 mRNA levels in select subpopulations of neurons in the cerebellar dentate nuclei in autism: An in situ hybridization study. Autism Res. 2009, 2, 50–59. [Google Scholar] [CrossRef] [Green Version]
  168. Levy, L.M.; Lehre, K.P.; Walaas, S.I.; Storm-Mathisen, J.; Danbolt, N.C. Down-regulation of glial glutamate transporters after glutamatergic denervation in the rat brain. Eur. J. Neurosci. 1995, 7, 2036–2041. [Google Scholar] [CrossRef]
  169. Bauman, M.; Kemper, T.L. Histoanatomic observations of the brain in early infantile autism. Neurology 1985, 35, 866–874. [Google Scholar] [CrossRef]
  170. Aylward, E.H.; Minshew, N.J.; Goldstein, G.; Honeycutt, N.A.; Augustine, A.M.; Yates, K.O.; Barta, P.E.; Pearlson, G.D. MRI volumes of amygdala and hippocampus in non-mentally retarded autistic adolescents and adults. Neurology 1999, 53, 2145–2150. [Google Scholar] [CrossRef]
  171. Raymond, G.V.; Bauman, M.L.; Kemper, T.L. Hippocampus in autism: A Golgi analysis. Acta Neuropathol. 1996, 91, 117–119. [Google Scholar] [CrossRef]
  172. Lima Caldeira, G.; Peça, J.; Carvalho, A.L. New insights on synaptic dysfunction in neuropsychiatric disorders. Curr. Opin. Neurobiol. 2019, 57, 62–70. [Google Scholar] [CrossRef]
  173. Eshraghi, A.A.; Liu, G.; Kay, S.-I.S.; Eshraghi, R.S.; Mittal, J.; Moshiree, B.; Mittal, R. Epigenetics and autism spectrum disorder: Is there a correlation? Front. Cell. Neurosci. 2018, 12, 78. [Google Scholar] [CrossRef] [Green Version]
  174. Rangasamy, S.; D’Mello, S.R.; Narayanan, V. Epigenetics, autism spectrum, and neurodevelopmental disorders. Neurotherapeutics 2013, 10, 742–756. [Google Scholar] [CrossRef] [Green Version]
  175. Kim, J.Y.; Son, M.J.; Son, C.Y.; Radua, J.; Eisenhut, M.; Gressier, F.; Koyanagi, A.; Carvalho, A.F.; Stubbs, B.; Solmi, M.; et al. Environmental risk factors and biomarkers for autism spectrum disorder: An umbrella review of the evidence. Lancet Psychiatry 2019, 6, 590–600. [Google Scholar] [CrossRef] [Green Version]
  176. Loke, Y.J.; Hannan, A.J.; Craig, J.M. The role of epigenetic change in autism spectrum disorders. Front. Neurol. 2015, 6, 107. [Google Scholar] [CrossRef] [Green Version]
  177. Vuong, C.K.; Black, D.L.; Zheng, S. The neurogenetics of alternative splicing. Nat. Rev. Neurosci. 2016, 17, 265–281. [Google Scholar] [CrossRef]
  178. Kubota, T.; Miyake, K.; Hirasawa, T. Epigenetic understanding of gene-environment interactions in psychiatric disorders: A new concept of clinical genetics. Clin. Epigenet. 2012, 4, 1. [Google Scholar] [CrossRef] [Green Version]
  179. Mill, J.; Heijmans, B.T. From promises to practical strategies in epigenetic epidemiology. Nat. Rev. Genet. 2013, 14, 585–594. [Google Scholar] [CrossRef]
  180. Montano, C.; Taub, M.A.; Jaffe, A.; Briem, E.; Feinberg, J.I.; Trygvadottir, R.; Idrizi, A.; Runarsson, A.; Berndsen, B.; Gur, R.C.; et al. Association of DNA Methylation Differences With Schizophrenia in an Epigenome-Wide Association Study. JAMA Psychiatry 2016, 73, 506–514. [Google Scholar] [CrossRef]
  181. Basu, S.N.; Kollu, R.; Banerjee-Basu, S. AutDB: A gene reference resource for autism research. Nucleic Acids Res. 2009, 37, D832–D836. [Google Scholar] [CrossRef] [Green Version]
  182. Zhubi, A.; Chen, Y.; Dong, E.; Cook, E.H.; Guidotti, A.; Grayson, D.R. Increased binding of MeCP2 to the GAD1 and RELN promoters may be mediated by an enrichment of 5-hmC in autism spectrum disorder (ASD) cerebellum. Transl. Psychiatry 2014, 4, e349. [Google Scholar] [CrossRef] [Green Version]
  183. Richetto, J.; Calabrese, F.; Meyer, U.; Riva, M.A. Prenatal versus postnatal maternal factors in the development of infection-induced working memory impairments in mice. Brain Behav. Immun. 2013, 33, 190–200. [Google Scholar] [CrossRef]
  184. Cusmano, D.M.; Mong, J.A. In utero exposure to valproic Acid changes sleep in juvenile rats: A model for sleep disturbances in autism. Sleep 2014, 37, 1489–1499. [Google Scholar] [CrossRef] [Green Version]
  185. Sandhu, K.V.; Lang, D.; Müller, B.; Nullmeier, S.; Yanagawa, Y.; Schwegler, H.; Stork, O. Glutamic acid decarboxylase 67 haplodeficiency impairs social behavior in mice. Genes Brain Behav. 2014, 13, 439–450. [Google Scholar] [CrossRef] [Green Version]
  186. Zhang, K.; Hill, K.; Labak, S.; Blatt, G.J.; Soghomonian, J.J. Loss of glutamic acid decarboxylase (Gad67) in Gpr88-expressing neurons induces learning and social behavior deficits in mice. Neuroscience 2014, 275, 238–247. [Google Scholar] [CrossRef]
  187. Tremblay, M.W.; Jiang, Y.-H. DNA methylation and susceptibility to autism spectrum disorder. Annu. Rev. Med. 2019, 70, 151–166. [Google Scholar] [CrossRef] [PubMed]
  188. Walton, E.; Pingault, J.B.; Cecil, C.A.M.; Gaunt, T.R.; Relton, C.L.; Mill, J.; Barker, E.D. Epigenetic profiling of ADHD symptoms trajectories: A prospective, methylome-wide study. Mol. Psychiatry 2017, 22, 250–256. [Google Scholar] [CrossRef]
  189. Alex, A.M.; Saradalekshmi, K.R.; Shilen, N.; Suresh, P.A.; Banerjee, M. Genetic association of DNMT variants can play a critical role in defining the methylation patterns in autism. IUBMB Life 2019, 71, 901–907. [Google Scholar] [CrossRef] [PubMed]
  190. Fan, X.-Y.; Shi, G.; Zhao, P. Methylation in Syn and Psd95 genes underlie the inhibitory effect of oxytocin on oxycodone-induced conditioned place preference. Eur. Neuropsychopharmacol. 2019, 29, 1464–1475. [Google Scholar] [CrossRef]
  191. Feng, J.; Zhou, Y.; Campbell, S.L.; Le, T.; Li, E.; Sweatt, J.D.; Silva, A.J.; Fan, G. Dnmt1 and Dnmt3a maintain DNA methylation and regulate synaptic function in adult forebrain neurons. Nat. Neurosci. 2010, 13, 423–430. [Google Scholar] [CrossRef]
  192. Zhubi, A.; Chen, Y.; Guidotti, A.; Grayson, D.R. Epigenetic regulation of RELN and GAD1 in the frontal cortex (FC) of autism spectrum disorder (ASD) subjects. Int. J. Dev. Neurosci. 2017, 62, 63–72. [Google Scholar] [CrossRef] [PubMed]
  193. Chao, H.-T.; Zoghbi, H.Y.; Rosenmund, C. MeCP2 controls excitatory synaptic strength by regulating glutamatergic synapse number. Neuron 2007, 56, 58–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Sohal, V.S.; Rubenstein, J.L.R. Excitation-inhibition balance as a framework for investigating mechanisms in neuropsychiatric disorders. Mol. Psychiatry 2019, 24, 1248–1257. [Google Scholar] [CrossRef] [PubMed]
  195. Srancikova, A.; Bacova, Z.; Bakos, J. The epigenetic regulation of synaptic genes contributes to the etiology of autism. Rev. Neurosci. 2021, 32, 791–802. [Google Scholar] [CrossRef]
  196. Hu, Z.; Xiao, X.; Zhang, Z.; Li, M. Genetic insights and neurobiological implications from NRXN1 in neuropsychiatric disorders. Mol. Psychiatry 2019, 24, 1400–1414. [Google Scholar] [CrossRef] [PubMed]
  197. Lu, X.; Yang, B.; Yu, H.; Hu, X.; Nie, J.; Wan, B.; Zhang, M.; Lü, C. Epigenetic mechanisms underlying the effects of triptolide and tripchlorolide on the expression of neuroligin-1 in the hippocampus of APP/PS1 transgenic mice. Pharm. Biol. 2019, 57, 453–459. [Google Scholar] [CrossRef] [Green Version]
  198. Ladd-Acosta, C.; Fallin, M.D. The role of epigenetics in genetic and environmental epidemiology. Epigenomics 2016, 8, 271–283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Dall’Aglio, L.; Muka, T.; Cecil, C.A.M.; Bramer, W.M.; Verbiest, M.M.P.J.; Nano, J.; Hidalgo, A.C.; Franco, O.H.; Tiemeier, H. The role of epigenetic modifications in neurodevelopmental disorders: A systematic review. Neurosci. Biobehav. Rev. 2018, 94, 17–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Lee, E.; Lee, J.; Kim, E. Excitation/inhibition imbalance in animal models of autism spectrum disorders. Biol. Psychiatry 2017, 81, 838–847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Gao, R.; Penzes, P. Common mechanisms of excitatory and inhibitory imbalance in schizophrenia and autism spectrum disorders. Curr. Mol. Med. 2015, 15, 146–167. [Google Scholar] [CrossRef]
  202. Selby, L.; Zhang, C.; Sun, Q.-Q. Major defects in neocortical GABAergic inhibitory circuits in mice lacking the fragile X mental retardation protein. Neurosci. Lett. 2007, 412, 227–232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Peñagarikano, O.; Abrahams, B.S.; Herman, E.I.; Winden, K.D.; Gdalyahu, A.; Dong, H.; Sonnenblick, L.I.; Gruver, R.; Almajano, J.; Bragin, A.; et al. Absence of CNTNAP2 leads to epilepsy, neuronal migration abnormalities, and core autism-related deficits. Cell 2011, 147, 235–246. [Google Scholar] [CrossRef] [Green Version]
  204. Qin, X.-Y.; Feng, J.-C.; Cao, C.; Wu, H.-T.; Loh, Y.P.; Cheng, Y. Association of Peripheral Blood Levels of Brain-Derived Neurotrophic Factor With Autism Spectrum Disorder in Children: A Systematic Review and Meta-analysis. JAMA Pediatr. 2016, 170, 1079–1086. [Google Scholar] [CrossRef] [PubMed]
  205. Qin, L.; Dai, X.; Yin, Y. Valproic acid exposure sequentially activates Wnt and mTOR pathways in rats. Mol. Cell. Neurosci. 2016, 75, 27–35. [Google Scholar] [CrossRef]
  206. Tang, X.; Jaenisch, R.; Sur, M. The role of GABAergic signalling in neurodevelopmental disorders. Nat. Rev. Neurosci. 2021, 22, 290–307. [Google Scholar] [CrossRef]
  207. Yizhar, O.; Fenno, L.E.; Prigge, M.; Schneider, F.; Davidson, T.J.; O’Shea, D.J.; Sohal, V.S.; Goshen, I.; Finkelstein, J.; Paz, J.T.; et al. Neocortical excitation/inhibition balance in information processing and social dysfunction. Nature 2011, 477, 171–178. [Google Scholar] [CrossRef] [PubMed]
  208. Park, G.; Jeon, S.J.; Ko, I.O.; Park, J.H.; Lee, K.C.; Kim, M.-S.; Shin, C.Y.; Kim, H.; Lee, Y.-S. Decreased in vivo glutamate/GABA ratio correlates with the social behavior deficit in a mouse model of autism spectrum disorder. Mol. Brain 2022, 15, 19. [Google Scholar] [CrossRef] [PubMed]
  209. Scott, M.; Milbourn, B.; Falkmer, M.; Black, M.; Bölte, S.; Halladay, A.; Lerner, M.; Taylor, J.L.; Girdler, S. Factors impacting employment for people with autism spectrum disorder: A scoping review. Autism 2019, 23, 869–901. [Google Scholar] [CrossRef] [PubMed]
  210. Selimbeyoglu, A.; Kim, C.K.; Inoue, M.; Lee, S.Y.; Hong, A.S.O.; Kauvar, I.; Ramakrishnan, C.; Fenno, L.E.; Davidson, T.J.; Wright, M.; et al. Modulation of prefrontal cortex excitation/inhibition balance rescues social behavior in CNTNAP2-deficient mice. Sci. Transl. Med. 2017, 9, eaah6733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  211. Banerjee, A.; García-Oscos, F.; Roychowdhury, S.; Galindo, L.C.; Hall, S.; Kilgard, M.P.; Atzori, M. Impairment of cortical GABAergic synaptic transmission in an environmental rat model of autism. Int. J. Neuropsychopharmacol. 2013, 16, 1309–1318. [Google Scholar] [CrossRef] [Green Version]
  212. Eltokhi, A.; Santuy, A.; Merchan-Perez, A.; Sprengel, R. Glutamatergic Dysfunction and Synaptic Ultrastructural Alterations in Schizophrenia and Autism Spectrum Disorder: Evidence from Human and Rodent Studies. Int. J. Mol. Sci. 2020, 22, 59. [Google Scholar] [CrossRef]
  213. Um, S.M.; Ha, S.; Lee, H.; Kim, J.; Kim, K.; Shin, W.; Cho, Y.S.; Roh, J.D.; Kang, J.; Yoo, T.; et al. NGL-2 Deletion Leads to Autistic-like Behaviors Responsive to NMDAR Modulation. Cell Rep. 2018, 23, 3839–3851. [Google Scholar] [CrossRef] [PubMed]
  214. Banke, T.G.; Barria, A. Transient enhanced glua2 expression in young hippocampal neurons of a fragile X mouse model. Front. Synaptic Neurosci. 2020, 12, 588295. [Google Scholar] [CrossRef] [PubMed]
  215. Cheng, G.-R.; Li, X.-Y.; Xiang, Y.-D.; Liu, D.; McClintock, S.M.; Zeng, Y. The implication of AMPA receptor in synaptic plasticity impairment and intellectual disability in fragile X syndrome. Physiol. Res. 2017, 66, 715–727. [Google Scholar] [CrossRef]
  216. Eadie, B.D.; Cushman, J.; Kannangara, T.S.; Fanselow, M.S.; Christie, B.R. NMDA receptor hypofunction in the dentate gyrus and impaired context discrimination in adult Fmr1 knockout mice. Hippocampus 2012, 22, 241–254. [Google Scholar] [CrossRef] [PubMed]
  217. McQuate, A.; Barria, A. Rapid exchange of synaptic and extrasynaptic NMDA receptors in hippocampal CA1 neurons. J. Neurophysiol. 2020, 123, 1004–1014. [Google Scholar] [CrossRef] [PubMed]
  218. Hanley, J.G. Subunit-specific trafficking mechanisms regulating the synaptic expression of Ca(2+)-permeable AMPA receptors. Semin. Cell Dev. Biol. 2014, 27, 14–22. [Google Scholar] [CrossRef] [Green Version]
  219. Ronesi, J.A.; Collins, K.A.; Hays, S.A.; Tsai, N.-P.; Guo, W.; Birnbaum, S.G.; Hu, J.-H.; Worley, P.F.; Gibson, J.R.; Huber, K.M. Disrupted Homer scaffolds mediate abnormal mGluR5 function in a mouse model of fragile X syndrome. Nat. Neurosci. 2012, 15, 431–440. [Google Scholar] [CrossRef] [PubMed]
  220. Dölen, G.; Bear, M.F. Role for metabotropic glutamate receptor 5 (mGluR5) in the pathogenesis of fragile X syndrome. J. Physiol. 2008, 586, 1503–1508. [Google Scholar] [CrossRef] [PubMed]
  221. Rinaldi, T.; Kulangara, K.; Antoniello, K.; Markram, H. Elevated NMDA receptor levels and enhanced postsynaptic long-term potentiation induced by prenatal exposure to valproic acid. Proc. Natl. Acad. Sci. USA 2007, 104, 13501–13506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Silva, G.T.; Le Bé, J.-V.; Riachi, I.; Rinaldi, T.; Markram, K.; Markram, H. Enhanced long-term microcircuit plasticity in the valproic Acid animal model of autism. Front. Synaptic Neurosci. 2009, 1, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Gauthier, J.; Spiegelman, D.; Piton, A.; Lafrenière, R.G.; Laurent, S.; St-Onge, J.; Lapointe, L.; Hamdan, F.F.; Cossette, P.; Mottron, L.; et al. Novel de novo SHANK3 mutation in autistic patients. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2009, 150B, 421–424. [Google Scholar] [CrossRef]
  224. Modi, M.E.; Brooks, J.M.; Guilmette, E.R.; Beyna, M.; Graf, R.; Reim, D.; Schmeisser, M.J.; Boeckers, T.M.; O’Donnell, P.; Buhl, D.L. Hyperactivity and Hypermotivation Associated With Increased Striatal mGluR1 Signaling in a Shank2 Rat Model of Autism. Front. Mol. Neurosci. 2018, 11, 107. [Google Scholar] [CrossRef] [PubMed]
  225. Duffney, L.J.; Wei, J.; Cheng, J.; Liu, W.; Smith, K.R.; Kittler, J.T.; Yan, Z. Shank3 deficiency induces NMDA receptor hypofunction via an actin-dependent mechanism. J. Neurosci. 2013, 33, 15767–15778. [Google Scholar] [CrossRef] [Green Version]
  226. Vicidomini, C.; Ponzoni, L.; Lim, D.; Schmeisser, M.J.; Reim, D.; Morello, N.; Orellana, D.; Tozzi, A.; Durante, V.; Scalmani, P.; et al. Pharmacological enhancement of mGlu5 receptors rescues behavioral deficits in SHANK3 knock-out mice. Mol. Psychiatry 2017, 22, 689–702. [Google Scholar] [CrossRef] [Green Version]
  227. Varoqueaux, F.; Aramuni, G.; Rawson, R.L.; Mohrmann, R.; Missler, M.; Gottmann, K.; Zhang, W.; Südhof, T.C.; Brose, N. Neuroligins determine synapse maturation and function. Neuron 2006, 51, 741–754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Blundell, J.; Blaiss, C.A.; Etherton, M.R.; Espinosa, F.; Tabuchi, K.; Walz, C.; Bolliger, M.F.; Südhof, T.C.; Powell, C.M. Neuroligin-1 deletion results in impaired spatial memory and increased repetitive behavior. J. Neurosci. 2010, 30, 2115–2129. [Google Scholar] [CrossRef] [Green Version]
  229. Etherton, M.R.; Tabuchi, K.; Sharma, M.; Ko, J.; Südhof, T.C. An autism-associated point mutation in the neuroligin cytoplasmic tail selectively impairs AMPA receptor-mediated synaptic transmission in hippocampus. EMBO J. 2011, 30, 2908–2919. [Google Scholar] [CrossRef]
  230. Etherton, M.; Földy, C.; Sharma, M.; Tabuchi, K.; Liu, X.; Shamloo, M.; Malenka, R.C.; Südhof, T.C. Autism-linked neuroligin-3 R451C mutation differentially alters hippocampal and cortical synaptic function. Proc. Natl. Acad. Sci. USA 2011, 108, 13764–13769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Silverman, J.L.; Oliver, C.F.; Karras, M.N.; Gastrell, P.T.; Crawley, J.N. AMPAKINE enhancement of social interaction in the BTBR mouse model of autism. Neuropharmacology 2013, 64, 268–282. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Lee, B.; Zhang, Y.; Kim, Y.; Kim, S.; Lee, Y.; Han, K. Age-dependent decrease of GAD65/67 mRNAs but normal densities of GABAergic interneurons in the brain regions of Shank3-overexpressing manic mouse model. Neurosci. Lett. 2017, 649, 48–54. [Google Scholar] [CrossRef] [PubMed]
  233. Jaramillo, T.C.; Xuan, Z.; Reimers, J.M.; Escamilla, C.O.; Liu, S.; Powell, C.M. Early Restoration of Shank3 Expression in Shank3 Knock-Out Mice Prevents Core ASD-Like Behavioral Phenotypes. eNeuro 2020, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Yan, Q.J.; Rammal, M.; Tranfaglia, M.; Bauchwitz, R.P. Suppression of two major Fragile X Syndrome mouse model phenotypes by the mGluR5 antagonist MPEP. Neuropharmacology 2005, 49, 1053–1066. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Lai, J.K.Y.; Lerch, J.P.; Doering, L.C.; Foster, J.A.; Ellegood, J. Regional brain volumes changes in adult male FMR1-KO mouse on the FVB strain. Neuroscience 2016, 318, 12–21. [Google Scholar] [CrossRef]
  236. Blundell, J.; Tabuchi, K.; Bolliger, M.F.; Blaiss, C.A.; Brose, N.; Liu, X.; Südhof, T.C.; Powell, C.M. Increased anxiety-like behavior in mice lacking the inhibitory synapse cell adhesion molecule neuroligin 2. Genes Brain Behav. 2009, 8, 114–126. [Google Scholar] [CrossRef] [Green Version]
  237. Silverman, J.L.; Tolu, S.S.; Barkan, C.L.; Crawley, J.N. Repetitive self-grooming behavior in the BTBR mouse model of autism is blocked by the mGluR5 antagonist MPEP. Neuropsychopharmacology 2010, 35, 976–989. [Google Scholar] [CrossRef]
  238. Shipman, S.L.; Nicoll, R.A. A subtype-specific function for the extracellular domain of neuroligin 1 in hippocampal LTP. Neuron 2012, 76, 309–316. [Google Scholar] [CrossRef] [Green Version]
  239. Asaka, Y.; Jugloff, D.G.M.; Zhang, L.; Eubanks, J.H.; Fitzsimonds, R.M. Hippocampal synaptic plasticity is impaired in the Mecp2-null mouse model of Rett syndrome. Neurobiol. Dis. 2006, 21, 217–227. [Google Scholar] [CrossRef]
  240. Maliszewska-Cyna, E.; Bawa, D.; Eubanks, J.H. Diminished prevalence but preserved synaptic distribution of N-methyl-D-aspartate receptor subunits in the methyl CpG binding protein 2(MeCP2)-null mouse brain. Neuroscience 2010, 168, 624–632. [Google Scholar] [CrossRef]
  241. Rompala, G.R.; Zsiros, V.; Zhang, S.; Kolata, S.M.; Nakazawa, K. Contribution of NMDA receptor hypofunction in prefrontal and cortical excitatory neurons to schizophrenia-like phenotypes. PLoS ONE 2013, 8, e61278. [Google Scholar] [CrossRef] [Green Version]
  242. Lerma, J.; Marques, J.M. Kainate receptors in health and disease. Neuron 2013, 80, 292–311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Motazacker, M.M.; Rost, B.R.; Hucho, T.; Garshasbi, M.; Kahrizi, K.; Ullmann, R.; Abedini, S.S.; Nieh, S.E.; Amini, S.H.; Goswami, C.; et al. A defect in the ionotropic glutamate receptor 6 gene (GRIK2) is associated with autosomal recessive mental retardation. Am. J. Hum. Genet. 2007, 81, 792–798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Lanore, F.; Labrousse, V.F.; Szabo, Z.; Normand, E.; Blanchet, C.; Mulle, C. Deficits in morphofunctional maturation of hippocampal mossy fiber synapses in a mouse model of intellectual disability. J. Neurosci. 2012, 32, 17882–17893. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Micheau, J.; Vimeney, A.; Normand, E.; Mulle, C.; Riedel, G. Impaired hippocampus-dependent spatial flexibility and sociability represent autism-like phenotypes in GluK2 mice. Hippocampus 2014, 24, 1059–1069. [Google Scholar] [CrossRef]
  246. Aller, M.I.; Pecoraro, V.; Paternain, A.V.; Canals, S.; Lerma, J. Increased Dosage of High-Affinity Kainate Receptor Gene grik4 Alters Synaptic Transmission and Reproduces Autism Spectrum Disorders Features. J. Neurosci. 2015, 35, 13619–13628. [Google Scholar] [CrossRef] [Green Version]
  247. Arora, V.; Pecoraro, V.; Aller, M.I.; Román, C.; Paternain, A.V.; Lerma, J. Increased Grik4 Gene Dosage Causes Imbalanced Circuit Output and Human Disease-Related Behaviors. Cell Rep. 2018, 23, 3827–3838. [Google Scholar] [CrossRef]
  248. de la Torre-Ubieta, L.; Won, H.; Stein, J.L.; Geschwind, D.H. Advancing the understanding of autism disease mechanisms through genetics. Nat. Med. 2016, 22, 345–361. [Google Scholar] [CrossRef]
  249. Trobiani, L.; Meringolo, M.; Diamanti, T.; Bourne, Y.; Marchot, P.; Martella, G.; Dini, L.; Pisani, A.; De Jaco, A.; Bonsi, P. The neuroligins and the synaptic pathway in Autism Spectrum Disorder. Neurosci. Biobehav. Rev. 2020, 119, 37–51. [Google Scholar] [CrossRef]
  250. Bonsi, P.; De Jaco, A.; Fasano, L.; Gubellini, P. Postsynaptic autism spectrum disorder genes and synaptic dysfunction. Neurobiol. Dis. 2022, 162, 105564. [Google Scholar] [CrossRef]
  251. Bateup, H.S.; Takasaki, K.T.; Saulnier, J.L.; Denefrio, C.L.; Sabatini, B.L. Loss of Tsc1 in vivo impairs hippocampal mGluR-LTD and increases excitatory synaptic function. J. Neurosci. 2011, 31, 8862–8869. [Google Scholar] [CrossRef] [Green Version]
  252. Takeuchi, K.; Gertner, M.J.; Zhou, J.; Parada, L.F.; Bennett, M.V.L.; Zukin, R.S. Dysregulation of synaptic plasticity precedes appearance of morphological defects in a Pten conditional knockout mouse model of autism. Proc. Natl. Acad. Sci. USA 2013, 110, 4738–4743. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Tian, D.; Stoppel, L.J.; Heynen, A.J.; Lindemann, L.; Jaeschke, G.; Mills, A.A.; Bear, M.F. Contribution of mGluR5 to pathophysiology in a mouse model of human chromosome 16p11.2 microdeletion. Nat. Neurosci. 2015, 18, 182–184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Moretti, P.; Levenson, J.M.; Battaglia, F.; Atkinson, R.; Teague, R.; Antalffy, B.; Armstrong, D.; Arancio, O.; Sweatt, J.D.; Zoghbi, H.Y. Learning and memory and synaptic plasticity are impaired in a mouse model of Rett syndrome. J. Neurosci. 2006, 26, 319–327. [Google Scholar] [CrossRef] [Green Version]
  255. Di Giorgi Gerevini, V.D.; Caruso, A.; Cappuccio, I.; Ricci Vitiani, L.; Romeo, S.; Della Rocca, C.; Gradini, R.; Melchiorri, D.; Nicoletti, F. The mGlu5 metabotropic glutamate receptor is expressed in zones of active neurogenesis of the embryonic and postnatal brain. Brain Res. Dev. Brain Res. 2004, 150, 17–22. [Google Scholar] [CrossRef] [PubMed]
  256. Di Giorgi-Gerevini, V.; Melchiorri, D.; Battaglia, G.; Ricci-Vitiani, L.; Ciceroni, C.; Busceti, C.L.; Biagioni, F.; Iacovelli, L.; Canudas, A.M.; Parati, E.; et al. Endogenous activation of metabotropic glutamate receptors supports the proliferation and survival of neural progenitor cells. Cell Death Differ. 2005, 12, 1124–1133. [Google Scholar] [CrossRef] [Green Version]
  257. Ballard, T.M.; Woolley, M.L.; Prinssen, E.; Huwyler, J.; Porter, R.; Spooren, W. The effect of the mGlu5 receptor antagonist MPEP in rodent tests of anxiety and cognition: A comparison. Psychopharmacology 2005, 179, 218–229. [Google Scholar] [CrossRef]
  258. Xu, J.; Marshall, J.J.; Kraniotis, S.; Nomura, T.; Zhu, Y.; Contractor, A. Genetic disruption of Grm5 causes complex alterations in motor activity, anxiety and social behaviors. Behav. Brain Res. 2021, 411, 113378. [Google Scholar] [CrossRef]
  259. Mody, M.; Petibon, Y.; Han, P.; Kuruppu, D.; Ma, C.; Yokell, D.; Neelamegam, R.; Normandin, M.D.; Fakhri, G.E.; Brownell, A.-L. In vivo imaging of mGlu5 receptor expression in humans with Fragile X Syndrome towards development of a potential biomarker. Sci. Rep. 2021, 11, 15897. [Google Scholar] [CrossRef]
  260. Huang, X.; Wang, M.; Zhang, Q.; Chen, X.; Wu, J. The role of glutamate receptors in attention-deficit/hyperactivity disorder: From physiology to disease. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2019, 180, 272–286. [Google Scholar] [CrossRef]
  261. Dogra, S.; Conn, P.J. Metabotropic glutamate receptors as emerging targets for the treatment of schizophrenia. Mol. Pharmacol. 2022. [Google Scholar] [CrossRef]
  262. Brown, J.; Iacovelli, L.; Di Cicco, G.; Grayson, B.; Rimmer, L.; Fletcher, J.; Neill, J.C.; Wall, M.J.; Ngomba, R.T.; Harte, M. The comparative effects of mGlu5 receptor positive allosteric modulators VU0409551 and VU0360172 on cognitive deficits and signalling in the sub-chronic PCP rat model for schizophrenia. Neuropharmacology 2022, 208, 108982. [Google Scholar] [CrossRef]
  263. Auerbach, B.D.; Osterweil, E.K.; Bear, M.F. Mutations causing syndromic autism define an axis of synaptic pathophysiology. Nature 2011, 480, 63–68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Aguilar-Valles, A.; Matta-Camacho, E.; Khoutorsky, A.; Gkogkas, C.; Nader, K.; Lacaille, J.-C.; Sonenberg, N. Inhibition of Group I Metabotropic Glutamate Receptors Reverses Autistic-Like Phenotypes Caused by Deficiency of the Translation Repressor eIF4E Binding Protein 2. J. Neurosci. 2015, 35, 11125–11132. [Google Scholar] [CrossRef] [PubMed]
  265. Satow, A.; Suzuki, G.; Maehara, S.; Hikichi, H.; Murai, T.; Murai, T.; Kawagoe-Takaki, H.; Hata, M.; Ito, S.; Ozaki, S.; et al. Unique antipsychotic activities of the selective metabotropic glutamate receptor 1 allosteric antagonist 2-cyclopropyl-5-[1-(2-fluoro-3-pyridinyl)-5-methyl-1H-1,2,3-triazol-4-yl]-2,3-dihydro-1H-isoindol-1-one. J. Pharmacol. Exp. Ther. 2009, 330, 179–190. [Google Scholar] [CrossRef]
  266. Bear, M.F.; Huber, K.M.; Warren, S.T. The mGluR theory of fragile X mental retardation. Trends Neurosci. 2004, 27, 370–377. [Google Scholar] [CrossRef] [PubMed]
  267. de Vrij, F.M.S.; Levenga, J.; van der Linde, H.C.; Koekkoek, S.K.; De Zeeuw, C.I.; Nelson, D.L.; Oostra, B.A.; Willemsen, R. Rescue of behavioral phenotype and neuronal protrusion morphology in Fmr1 KO mice. Neurobiol. Dis. 2008, 31, 127–132. [Google Scholar] [CrossRef] [Green Version]
  268. Sears, S.M.; Hewett, S.J. Influence of glutamate and GABA transport on brain excitatory/inhibitory balance. Exp. Biol. Med. 2021, 246, 1069–1083. [Google Scholar] [CrossRef]
  269. Culotta, L.; Penzes, P. Exploring the mechanisms underlying excitation/inhibition imbalance in human iPSC-derived models of ASD. Mol. Autism 2020, 11, 32. [Google Scholar] [CrossRef] [PubMed]
  270. Canitano, R.; Palumbi, R. Excitation/inhibition modulators in autism spectrum disorder: Current clinical research. Front. Neurosci. 2021, 15, 753274. [Google Scholar] [CrossRef] [PubMed]
  271. Port, R.G.; Oberman, L.M.; Roberts, T.P. Revisiting the excitation/inhibition imbalance hypothesis of ASD through a clinical lens. Br. J. Radiol. 2019, 92, 20180944. [Google Scholar] [CrossRef] [PubMed]
  272. Wang, D.D.; Kriegstein, A.R. GABA regulates excitatory synapse formation in the neocortex via NMDA receptor activation. J. Neurosci. 2008, 28, 5547–5558. [Google Scholar] [CrossRef] [PubMed]
  273. Manent, J.-B.; Demarque, M.; Jorquera, I.; Pellegrino, C.; Ben-Ari, Y.; Aniksztejn, L.; Represa, A. A noncanonical release of GABA and glutamate modulates neuronal migration. J. Neurosci. 2005, 25, 4755–4765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Ben-Ari, Y.; Gaiarsa, J.-L.; Tyzio, R.; Khazipov, R. GABA: A pioneer transmitter that excites immature neurons and generates primitive oscillations. Physiol. Rev. 2007, 87, 1215–1284. [Google Scholar] [CrossRef]
  275. Bonnet-Brilhault, F.; Alirol, S.; Blanc, R.; Bazaud, S.; Marouillat, S.; Thépault, R.A.; Andres, C.R.; Lemonnier, É.; Barthélémy, C.; Raynaud, M.; et al. GABA/Glutamate synaptic pathways targeted by integrative genomic and electrophysiological explorations distinguish autism from intellectual disability. Mol. Psychiatry 2016, 21, 411–418. [Google Scholar] [CrossRef] [PubMed]
  276. Zhao, H.; Mao, X.; Zhu, C.; Zou, X.; Peng, F.; Yang, W.; Li, B.; Li, G.; Ge, T.; Cui, R. Gabaergic system dysfunction in autism spectrum disorders. Front. Cell Dev. Biol. 2021, 9, 781327. [Google Scholar] [CrossRef] [PubMed]
  277. Di, J.; Li, J.; O’Hara, B.; Alberts, I.; Xiong, L.; Li, J.; Li, X. The role of GABAergic neural circuits in the pathogenesis of autism spectrum disorder. Int. J. Dev. Neurosci. 2020, 80, 73–85. [Google Scholar] [CrossRef] [PubMed]
  278. Dhossche, D.; Applegate, H.; Abraham, A.; Maertens, P.; Bland, L.; Bencsath, A.; Martinez, J. Elevated plasma gamma-aminobutyric acid (GABA) levels in autistic youngsters: Stimulus for a GABA hypothesis of autism. Med. Sci. Monit. 2002, 8, PR1-6. [Google Scholar] [PubMed]
  279. Harada, M.; Taki, M.M.; Nose, A.; Kubo, H.; Mori, K.; Nishitani, H.; Matsuda, T. Non-invasive evaluation of the GABAergic/glutamatergic system in autistic patients observed by MEGA-editing proton MR spectroscopy using a clinical 3 tesla instrument. J. Autism Dev. Disord. 2011, 41, 447–454. [Google Scholar] [CrossRef]
  280. Drenthen, G.S.; Barendse, E.M.; Aldenkamp, A.P.; van Veenendaal, T.M.; Puts, N.A.J.; Edden, R.A.E.; Zinger, S.; Thoonen, G.; Hendriks, M.P.H.; Kessels, R.P.C.; et al. Altered neurotransmitter metabolism in adolescents with high-functioning autism. Psychiatry Res. Neuroimaging 2016, 256, 44–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  281. Ajram, L.A.; Horder, J.; Mendez, M.A.; Galanopoulos, A.; Brennan, L.P.; Wichers, R.H.; Robertson, D.M.; Murphy, C.M.; Zinkstok, J.; Ivin, G.; et al. Shifting brain inhibitory balance and connectivity of the prefrontal cortex of adults with autism spectrum disorder. Transl. Psychiatry 2017, 7, e1137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Sapey-Triomphe, L.-A.; Lamberton, F.; Sonié, S.; Mattout, J.; Schmitz, C. Tactile hypersensitivity and GABA concentration in the sensorimotor cortex of adults with autism. Autism Res. 2019, 12, 562–575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Edmondson, D.A.; Xia, P.; McNally Keehn, R.; Dydak, U.; Keehn, B. A Magnetic Resonance Spectroscopy Study of Superior Visual Search Abilities in Children with Autism Spectrum Disorder. Autism Res. 2020, 13, 550–562. [Google Scholar] [CrossRef] [PubMed]
  284. Cochran, D.M.; Sikoglu, E.M.; Hodge, S.M.; Edden, R.A.E.; Foley, A.; Kennedy, D.N.; Moore, C.M.; Frazier, J.A. Relationship among Glutamine, γ-Aminobutyric Acid, and Social Cognition in Autism Spectrum Disorders. J. Child Adolesc. Psychopharmacol. 2015, 25, 314–322. [Google Scholar] [CrossRef] [Green Version]
  285. Yip, J.; Soghomonian, J.-J.; Blatt, G.J. Increased GAD67 mRNA expression in cerebellar interneurons in autism: Implications for Purkinje cell dysfunction. J. Neurosci. Res. 2008, 86, 525–530. [Google Scholar] [CrossRef] [PubMed]
  286. Lawrence, Y.A.; Kemper, T.L.; Bauman, M.L.; Blatt, G.J. Parvalbumin-, calbindin-, and calretinin-immunoreactive hippocampal interneuron density in autism. Acta Neurol. Scand. 2010, 121, 99–108. [Google Scholar] [CrossRef] [PubMed]
  287. Fatemi, S.H.; Reutiman, T.J.; Folsom, T.D.; Rooney, R.J.; Patel, D.H.; Thuras, P.D. mRNA and protein levels for GABAAalpha4, alpha5, beta1 and GABABR1 receptors are altered in brains from subjects with autism. J. Autism Dev. Disord. 2010, 40, 743–750. [Google Scholar] [CrossRef] [Green Version]
  288. Fatemi, S.H.; Reutiman, T.J.; Folsom, T.D.; Thuras, P.D. GABA(A) receptor downregulation in brains of subjects with autism. J. Autism Dev. Disord. 2009, 39, 223–230. [Google Scholar] [CrossRef] [Green Version]
  289. Buxbaum, J.D.; Silverman, J.M.; Smith, C.J.; Greenberg, D.A.; Kilifarski, M.; Reichert, J.; Cook, E.H.; Fang, Y.; Song, C.Y.; Vitale, R. Association between a GABRB3 polymorphism and autism. Mol. Psychiatry 2002, 7, 311–316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  290. Hogart, A.; Nagarajan, R.P.; Patzel, K.A.; Yasui, D.H.; Lasalle, J.M. 15q11-13 GABAA receptor genes are normally biallelically expressed in brain yet are subject to epigenetic dysregulation in autism-spectrum disorders. Hum. Mol. Genet. 2007, 16, 691–703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  291. Guo, H.; Peng, Y.; Hu, Z.; Li, Y.; Xun, G.; Ou, J.; Sun, L.; Xiong, Z.; Liu, Y.; Wang, T.; et al. Genome-wide copy number variation analysis in a Chinese autism spectrum disorder cohort. Sci. Rep. 2017, 7, 44155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  292. Delahanty, R.J.; Kang, J.Q.; Brune, C.W.; Kistner, E.O.; Courchesne, E.; Cox, N.J.; Cook, E.H.; Macdonald, R.L.; Sutcliffe, J.S. Maternal transmission of a rare GABRB3 signal peptide variant is associated with autism. Mol. Psychiatry 2011, 16, 86–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Heise, C.; Preuss, J.M.; Schroeder, J.C.; Battaglia, C.R.; Kolibius, J.; Schmid, R.; Kreutz, M.R.; Kas, M.J.H.; Burbach, J.P.H.; Boeckers, T.M. Heterogeneity of cell surface glutamate and GABA receptor expression in shank and CNTN4 autism mouse models. Front. Mol. Neurosci. 2018, 11, 212. [Google Scholar] [CrossRef] [PubMed]
  294. Sabanov, V.; Braat, S.; D’Andrea, L.; Willemsen, R.; Zeidler, S.; Rooms, L.; Bagni, C.; Kooy, R.F.; Balschun, D. Impaired GABAergic inhibition in the hippocampus of Fmr1 knockout mice. Neuropharmacology 2017, 116, 71–81. [Google Scholar] [CrossRef] [PubMed]
  295. El-Khoury, R.; Panayotis, N.; Matagne, V.; Ghata, A.; Villard, L.; Roux, J.-C. GABA and glutamate pathways are spatially and developmentally affected in the brain of Mecp2-deficient mice. PLoS ONE 2014, 9, e92169. [Google Scholar] [CrossRef]
  296. Lim, C.-S.; Kim, H.; Yu, N.-K.; Kang, S.J.; Kim, T.; Ko, H.-G.; Lee, J.; Yang, J.-E.; Ryu, H.-H.; Park, T.; et al. Enhancing inhibitory synaptic function reverses spatial memory deficits in Shank2 mutant mice. Neuropharmacology 2017, 112, 104–112. [Google Scholar] [CrossRef]
  297. Sgadò, P.; Genovesi, S.; Kalinovsky, A.; Zunino, G.; Macchi, F.; Allegra, M.; Murenu, E.; Provenzano, G.; Tripathi, P.P.; Casarosa, S.; et al. Loss of GABAergic neurons in the hippocampus and cerebral cortex of Engrailed-2 null mutant mice: Implications for autism spectrum disorders. Exp. Neurol. 2013, 247, 496–505. [Google Scholar] [CrossRef] [Green Version]
  298. Provenzano, G.; Gilardoni, A.; Maggia, M.; Pernigo, M.; Sgadò, P.; Casarosa, S.; Bozzi, Y. Altered Expression of GABAergic Markers in the Forebrain of Young and Adult Engrailed-2 Knockout Mice. Genes 2020, 11, 384. [Google Scholar] [CrossRef] [Green Version]
  299. Hammer, M.; Krueger-Burg, D.; Tuffy, L.P.; Cooper, B.H.; Taschenberger, H.; Goswami, S.P.; Ehrenreich, H.; Jonas, P.; Varoqueaux, F.; Rhee, J.-S.; et al. Perturbed Hippocampal Synaptic Inhibition and γ-Oscillations in a Neuroligin-4 Knockout Mouse Model of Autism. Cell Rep. 2015, 13, 516–523. [Google Scholar] [CrossRef] [Green Version]
  300. Jiang, D.-Y.; Wu, Z.; Forsyth, C.T.; Hu, Y.; Yee, S.-P.; Chen, G. GABAergic deficits and schizophrenia-like behaviors in a mouse model carrying patient-derived neuroligin-2 R215H mutation. Mol. Brain 2018, 11, 31. [Google Scholar] [CrossRef] [Green Version]
  301. Chen, J.; Dong, B.; Feng, X.; Jiang, D.; Chen, G.; Long, C.; Yang, L. Aberrant mPFC GABAergic synaptic transmission and fear behavior in neuroligin-2 R215H knock-in mice. Brain Res. 2020, 1730, 146671. [Google Scholar] [CrossRef]
  302. Chau, D.K.-F.; Choi, A.Y.-T.; Yang, W.; Leung, W.N.; Chan, C.W. Downregulation of glutamatergic and GABAergic proteins in valproric acid associated social impairment during adolescence in mice. Behav. Brain Res. 2017, 316, 255–260. [Google Scholar] [CrossRef] [PubMed]
  303. Hou, Q.; Wang, Y.; Li, Y.; Chen, D.; Yang, F.; Wang, S. A Developmental Study of Abnormal Behaviors and Altered GABAergic Signaling in the VPA-Treated Rat Model of Autism. Front. Behav. Neurosci. 2018, 12, 182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Varman, D.R.; Soria-Ortíz, M.B.; Martínez-Torres, A.; Reyes-Haro, D. GABAρ3 expression in lobule X of the cerebellum is reduced in the valproate model of autism. Neurosci. Lett. 2018, 687, 158–163. [Google Scholar] [CrossRef] [PubMed]
  305. Zieminska, E.; Toczylowska, B.; Diamandakis, D.; Hilgier, W.; Filipkowski, R.K.; Polowy, R.; Orzel, J.; Gorka, M.; Lazarewicz, J.W. Glutamate, glutamine and GABA levels in rat brain measured using MRS, HPLC and NMR methods in study of two models of autism. Front. Mol. Neurosci. 2018, 11, 418. [Google Scholar] [CrossRef] [Green Version]
  306. Lozovaya, N.; Nardou, R.; Tyzio, R.; Chiesa, M.; Pons-Bennaceur, A.; Eftekhari, S.; Bui, T.T.; Billon-Grand, M.; Rasero, J.; Bonifazi, P.; et al. Early alterations in a mouse model of Rett syndrome: The GABA developmental shift is abolished at birth. Sci. Rep. 2019, 9, 9276. [Google Scholar] [CrossRef] [Green Version]
  307. Maffei, A.; Fontanini, A. Network homeostasis: A matter of coordination. Curr. Opin. Neurobiol. 2009, 19, 168–173. [Google Scholar] [CrossRef] [Green Version]
  308. Bruining, H.; Hardstone, R.; Juarez-Martinez, E.L.; Sprengers, J.; Avramiea, A.-E.; Simpraga, S.; Houtman, S.J.; Poil, S.-S.; Dallares, E.; Palva, S.; et al. Measurement of excitation-inhibition ratio in autism spectrum disorder using critical brain dynamics. Sci. Rep. 2020, 10, 9195. [Google Scholar] [CrossRef]
  309. Spiegel, A.; Mentch, J.; Haskins, A.J.; Robertson, C.E. Slower binocular rivalry in the autistic brain. Curr. Biol. 2019, 29, 2948–2953.e3. [Google Scholar] [CrossRef]
  310. Oliveira, B.; Mitjans, M.; Nitsche, M.A.; Kuo, M.-F.; Ehrenreich, H. Excitation-inhibition dysbalance as predictor of autistic phenotypes. J. Psychiatr. Res. 2018, 104, 96–99. [Google Scholar] [CrossRef]
  311. Trakoshis, S.; Martínez-Cañada, P.; Rocchi, F.; Canella, C.; You, W.; Chakrabarti, B.; Ruigrok, A.N.; Bullmore, E.T.; Suckling, J.; Markicevic, M.; et al. Intrinsic excitation-inhibition imbalance affects medial prefrontal cortex differently in autistic men versus women. eLife 2020, 9, e55684. [Google Scholar] [CrossRef]
  312. Swaminathan, M.; Hill-Yardin, E.L.; Bornstein, J.C.; Foong, J.P.P. Endogenous glutamate excites myenteric calbindin neurons by activating group I metabotropic glutamate receptors in the mouse colon. Front. Neurosci. 2019, 13, 426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  313. Kelly, J.R.; Kennedy, P.J.; Cryan, J.F.; Dinan, T.G.; Clarke, G.; Hyland, N.P. Breaking down the barriers: The gut microbiome, intestinal permeability and stress-related psychiatric disorders. Front. Cell. Neurosci. 2015, 9, 392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  314. Braniste, V.; Al-Asmakh, M.; Kowal, C.; Anuar, F.; Abbaspour, A.; Tóth, M.; Korecka, A.; Bakocevic, N.; Ng, L.G.; Kundu, P.; et al. The gut microbiota influences blood-brain barrier permeability in mice. Sci. Transl. Med. 2014, 6, 263ra158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  315. Furness, J.B.; Callaghan, B.P.; Rivera, L.R.; Cho, H.-J. The enteric nervous system and gastrointestinal innervation: Integrated local and central control. Adv. Exp. Med. Biol. 2014, 817, 39–71. [Google Scholar] [CrossRef] [PubMed]
  316. Janik, R.; Thomason, L.A.M.; Stanisz, A.M.; Forsythe, P.; Bienenstock, J.; Stanisz, G.J. Magnetic resonance spectroscopy reveals oral Lactobacillus promotion of increases in brain GABA, N-acetyl aspartate and glutamate. Neuroimage 2016, 125, 988–995. [Google Scholar] [CrossRef]
  317. El-Ansary, A.; Bacha, A.B.; Bjørklund, G.; Al-Orf, N.; Bhat, R.S.; Moubayed, N.; Abed, K. Probiotic treatment reduces the autistic-like excitation/inhibition imbalance in juvenile hamsters induced by orally administered propionic acid and clindamycin. Metab. Brain Dis. 2018, 33, 1155–1164. [Google Scholar] [CrossRef]
  318. Wang, H.; Liu, L.; Rao, X.; Zeng, B.; Yu, Y.; Zhou, C.; Zeng, L.; Zheng, P.; Pu, J.; Xu, S.; et al. Integrated phosphoproteomic and metabolomic profiling reveals perturbed pathways in the hippocampus of gut microbiota dysbiosis mice. Transl. Psychiatry 2020, 10, 346. [Google Scholar] [CrossRef]
  319. Liu, M.T.; Rothstein, J.D.; Gershon, M.D.; Kirchgessner, A.L. Glutamatergic enteric neurons. J. Neurosci. 1997, 17, 4764–4784. [Google Scholar] [CrossRef] [Green Version]
  320. Iwanaga, T.; Goto, M.; Watanabe, M. Cellular distribution of glutamate transporters in the gastrointestinal tract of mice: An immunohistochemical and in situ hybridization approach. Biomed. Res. 2005, 26, 271–278. [Google Scholar] [CrossRef] [Green Version]
  321. Julio-Pieper, M.; Flor, P.J.; Dinan, T.G.; Cryan, J.F. Exciting times beyond the brain: Metabotropic glutamate receptors in peripheral and non-neural tissues. Pharmacol. Rev. 2011, 63, 35–58. [Google Scholar] [CrossRef] [Green Version]
  322. Filpa, V.; Moro, E.; Protasoni, M.; Crema, F.; Frigo, G.; Giaroni, C. Role of glutamatergic neurotransmission in the enteric nervous system and brain-gut axis in health and disease. Neuropharmacology 2016, 111, 14–33. [Google Scholar] [CrossRef] [PubMed]
  323. von Boyen, G.B.T.; Steinkamp, M.; Adler, G.; Kirsch, J. Glutamate receptor subunit expression in primary enteric glia cultures. J. Recept. Signal Transduct. Res. 2006, 26, 329–336. [Google Scholar] [CrossRef] [PubMed]
  324. Carpanese, E.; Moretto, P.; Filpa, V.; Marchet, S.; Moro, E.; Crema, F.; Frigo, G.; Giaroni, C. Antagonism of ionotropic glutamate receptors attenuates chemical ischemia-induced injury in rat primary cultured myenteric ganglia. PLoS ONE 2014, 9, e113613. [Google Scholar] [CrossRef]
  325. Jessen, K.R.; Mirsky, R. Astrocyte-like glia in the peripheral nervous system: An immunohistochemical study of enteric glia. J. Neurosci. 1983, 3, 2206–2218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  326. Schousboe, A.; Waagepetersen, H.S. Role of astrocytes in glutamate homeostasis: Implications for excitotoxicity. Neurotox. Res. 2005, 8, 221–225. [Google Scholar] [CrossRef]
  327. Wang, M.; Wan, J.; Rong, H.; He, F.; Wang, H.; Zhou, J.; Cai, C.; Wang, Y.; Xu, R.; Yin, Z.; et al. Alterations in Gut Glutamate Metabolism Associated with Changes in Gut Microbiota Composition in Children with Autism Spectrum Disorder. mSystems 2019, 4, e00321-18. [Google Scholar] [CrossRef] [Green Version]
  328. Olson, C.A.; Vuong, H.E.; Yano, J.M.; Liang, Q.Y.; Nusbaum, D.J.; Hsiao, E.Y. The Gut Microbiota Mediates the Anti-Seizure Effects of the Ketogenic Diet. Cell 2018, 173, 1728–1741.e13. [Google Scholar] [CrossRef] [Green Version]
  329. LeClerc, S.; Easley, D. Pharmacological therapies for autism spectrum disorder: A review. Pharm. Ther. 2015, 40, 389–397. [Google Scholar]
  330. Hardan, A.Y.; Hendren, R.L.; Aman, M.G.; Robb, A.; Melmed, R.D.; Andersen, K.A.; Luchini, R.; Rahman, R.; Ali, S.; Jia, X.D.; et al. Efficacy and safety of memantine in children with autism spectrum disorder: Results from three phase 2 multicenter studies. Autism 2019, 23, 2096–2111. [Google Scholar] [CrossRef]
  331. Findling, R.L.; Mankoski, R.; Timko, K.; Lears, K.; McCartney, T.; McQuade, R.D.; Eudicone, J.M.; Amatniek, J.; Marcus, R.N.; Sheehan, J.J. A Randomized Controlled Trial Investigating the Safety and Efficacy of Aripiprazole in the Long-Term Maintenance Treatment of Pediatric Patients With Irritability Associated With Autistic Disorder. J. Clin. Psychiatry 2014, 75, 22–30. [Google Scholar] [CrossRef]
  332. Ichikawa, H.; Mikami, K.; Okada, T.; Yamashita, Y.; Ishizaki, Y.; Tomoda, A.; Ono, H.; Usuki, C.; Tadori, Y. Aripiprazole in the Treatment of Irritability in Children and Adolescents with Autism Spectrum Disorder in Japan: A Randomized, Double-blind, Placebo-controlled Study. Child Psychiatry Hum. Dev. 2017, 48, 796–806. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  333. Hahamy, A.; Behrmann, M.; Malach, R. The idiosyncratic brain: Distortion of spontaneous connectivity patterns in autism spectrum disorder. Nat. Neurosci. 2015, 18, 302–309. [Google Scholar] [CrossRef] [PubMed]
  334. Just, M.A.; Keller, T.A.; Malave, V.L.; Kana, R.K.; Varma, S. Autism as a neural systems disorder: A theory of frontal-posterior underconnectivity. Neurosci. Biobehav. Rev. 2012, 36, 1292–1313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  335. Mantz, J.; Laudenbach, V.; Lecharny, J.B.; Henzel, D.; Desmonts, J.M. Riluzole, a novel antiglutamate, blocks GABA uptake by striatal synaptosomes. Eur. J. Pharmacol. 1994, 257, R7–R8. [Google Scholar] [CrossRef]
  336. Doble, A. The pharmacology and mechanism of action of riluzole. Neurology 1996, 47, S233–S241. [Google Scholar] [CrossRef]
  337. He, Y.; Benz, A.; Fu, T.; Wang, M.; Covey, D.F.; Zorumski, C.F.; Mennerick, S. Neuroprotective agent riluzole potentiates postsynaptic GABAA receptor function. Neuropharmacology 2002, 42, 199–209. [Google Scholar] [CrossRef]
  338. Chéramy, A.; Barbeito, L.; Godeheu, G.; Glowinski, J. Riluzole inhibits the release of glutamate in the caudate nucleus of the cat in vivo. Neurosci. Lett. 1992, 147, 209–212. [Google Scholar] [CrossRef]
  339. Le Liboux, A.; Lefebvre, P.; Le Roux, Y.; Truffinet, P.; Aubeneau, M.; Kirkesseli, S.; Montay, G. Single- and multiple-dose pharmacokinetics of riluzole in white subjects. J. Clin. Pharmacol. 1997, 37, 820–827. [Google Scholar] [CrossRef]
  340. Pittenger, C.; Coric, V.; Banasr, M.; Bloch, M.; Krystal, J.H.; Sanacora, G. Riluzole in the treatment of mood and anxiety disorders. CNS Drugs 2008, 22, 761–786. [Google Scholar] [CrossRef]
  341. Ghaleiha, A.; Mohammadi, E.; Mohammadi, M.-R.; Farokhnia, M.; Modabbernia, A.; Yekehtaz, H.; Ashrafi, M.; Hassanzadeh, E.; Akhondzadeh, S. Riluzole as an adjunctive therapy to risperidone for the treatment of irritability in children with autistic disorder: A double-blind, placebo-controlled, randomized trial. Paediatr. Drugs 2013, 15, 505–514. [Google Scholar] [CrossRef]
  342. de Boer, J.N.; Vingerhoets, C.; Hirdes, M.; McAlonan, G.M.; Amelsvoort, T.V.; Zinkstok, J.R. Efficacy and tolerability of riluzole in psychiatric disorders: A systematic review and preliminary meta-analysis. Psychiatry Res. 2019, 278, 294–302. [Google Scholar] [CrossRef]
  343. Wink, L.K.; Adams, R.; Horn, P.S.; Tessier, C.R.; Bantel, A.P.; Hong, M.; Shaffer, R.C.; Pedapati, E.V.; Erickson, C.A. A Randomized Placebo-Controlled Cross-Over Pilot Study of Riluzole for Drug-Refractory Irritability in Autism Spectrum Disorder. J. Autism Dev. Disord. 2018, 48, 3051–3060. [Google Scholar] [CrossRef]
  344. Joshi, G.; Wozniak, J.; Faraone, S.V.; Fried, R.; Chan, J.; Furtak, S.; Grimsley, E.; Conroy, K.; Kilcullen, J.R.; Woodworth, K.Y.; et al. A Prospective Open-Label Trial of Memantine Hydrochloride for the Treatment of Social Deficits in Intellectually Capable Adults With Autism Spectrum Disorder. J. Clin. Psychopharmacol. 2016, 36, 262–271. [Google Scholar] [CrossRef] [PubMed]
  345. Belsito, K.M.; Law, P.A.; Kirk, K.S.; Landa, R.J.; Zimmerman, A.W. Lamotrigine therapy for autistic disorder: A randomized, double-blind, placebo-controlled trial. J. Autism Dev. Disord. 2001, 31, 175–181. [Google Scholar] [CrossRef] [PubMed]
  346. Devine, D.P. Self-injurious behaviour in autistic children: A neuro-developmental theory of social and environmental isolation. Psychopharmacology 2014, 231, 979–997. [Google Scholar] [CrossRef] [PubMed]
  347. Hosenbocus, S.; Chahal, R. Memantine: A review of possible uses in child and adolescent psychiatry. J. Can. Acad. Child Adolesc. Psychiatry 2013, 22, 166–171. [Google Scholar] [PubMed]
  348. Ure, A.; Rose, V.; Bernie, C.; Williams, K. Autism: One or many spectrums? J. Paediatr. Child Health 2018, 54, 1068–1072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  349. Masi, A.; DeMayo, M.M.; Glozier, N.; Guastella, A.J. An overview of autism spectrum disorder, heterogeneity and treatment options. Neurosci. Bull. 2017, 33, 183–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  350. Pandina, G.; Ring, R.H.; Bangerter, A.; Ness, S. Current approaches to the pharmacologic treatment of core symptoms across the lifespan of autism spectrum disorder. Psychiatr. Clin. N. Am. 2020, 43, 629–645. [Google Scholar] [CrossRef] [PubMed]
  351. Huang, Y.; Arnold, S.R.; Foley, K.-R.; Trollor, J.N. Diagnosis of autism in adulthood: A scoping review. Autism 2020, 24, 1311–1327. [Google Scholar] [CrossRef] [PubMed]
  352. Cage, E.; Troxell-Whitman, Z. Understanding the reasons, contexts and costs of camouflaging for autistic adults. J. Autism Dev. Disord. 2019, 49, 1899–1911. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  353. Su, L.-D.; Wang, N.; Han, J.; Shen, Y. Group 1 metabotropic glutamate receptors in neurological and psychiatric diseases: Mechanisms and prospective. Neuroscientist 2021, 10738584211021018. [Google Scholar] [CrossRef] [PubMed]
  354. Thomas, A.M.; Bui, N.; Perkins, J.R.; Yuva-Paylor, L.A.; Paylor, R. Group I metabotropic glutamate receptor antagonists alter select behaviors in a mouse model for fragile X syndrome. Psychopharmacology 2012, 219, 47–58. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of the developmental time course of glutamatergic receptor subunits expression in different rodent brain areas. The variations in color intensity within the single bars represent changes along time in the expression levels of each receptor subunit in a particular area of the rodent brain. The midbrain, pons, and medulla are represented collectively as brain stem in the figure.
Figure 1. Schematic representation of the developmental time course of glutamatergic receptor subunits expression in different rodent brain areas. The variations in color intensity within the single bars represent changes along time in the expression levels of each receptor subunit in a particular area of the rodent brain. The midbrain, pons, and medulla are represented collectively as brain stem in the figure.
Ijms 23 03861 g001
Figure 2. Schematic drawing of the glutamatergic synaptic pathway altered in ASD. Once released, glutamate can bind to and activate AMPA, Kainate, and NMDA ionotropic receptors or G-protein coupled metabotropic receptors on postsynaptic or presynaptic neurons as well as glial cells. Glutamate, taken up into glial cells by EAAT1/2 glutamate transporters, is converted into glutamine, which can be transported into glutamatergic neurons by SNAT3/5 transporters.
Figure 2. Schematic drawing of the glutamatergic synaptic pathway altered in ASD. Once released, glutamate can bind to and activate AMPA, Kainate, and NMDA ionotropic receptors or G-protein coupled metabotropic receptors on postsynaptic or presynaptic neurons as well as glial cells. Glutamate, taken up into glial cells by EAAT1/2 glutamate transporters, is converted into glutamine, which can be transported into glutamatergic neurons by SNAT3/5 transporters.
Ijms 23 03861 g002
Table 1. Candidate genes implicated in ASD.
Table 1. Candidate genes implicated in ASD.
GeneProteinSFARI
Gene Score
EvidenceReferences
SLC1A2excitatory amino acid transporter 2 (EAAT2)Syndromic
(Score 3)
-Rare de novo deletion, Genetic Association[153]
GRIA2Glutamate ionotropic receptor AMPA type subunit 2High confidence
(Score 1)
-Heterozygous de novo variants;-19 megabase deletion in the chromosomal region containing the gene[119,120,125,126,154]
GRIN2BGlutamate Ionotropic Receptor NMDA Type Subunit 2BHigh confidence
(Score 1)
-Rare missense mutations;
-de novo splice-site variants
-de novo loss-of-function variants
[127,128,129,130,131,132,133,134]
GRIA1Glutamate Ionotropic Receptor AMPA Type Subunit 1Strong candidate
(Score 2)
-Missense variant (p.Ala636Thr)[121,135,136]
GRIN2AGlutamate Ionotropic Receptor NMDA Type Subunit 2AStrong candidate
(Score 2)
-Rare pathogenic deletions
-Heterozygous de novo missense variant
[123,137,138,139,140,141,142,143,144,145]
GRIK2Glutamate ionotropic receptor kainate type subunit 2Strong candidate
(Score 2)
-Rare Single Gene Mutation, Genetic Association[21,146]
GRIK5Glutamate ionotropic receptor kainate type subunit 5Strong candidate
(Score 2)
-Rare single gene mutations[147,148]
GRIP1Glutamate Receptor-Interacting Protein 1Strong candidate
(Score 2)
-Missense mutations[149]
GRID1Glutamate Ionotropic Receptor Delta Type Subunit 1Strong candidate
(Score 2)
-Rare Single Gene Mutation,
Genetic Association
[150,151]
GRM5Glutamate metabotropic receptor 5Suggestive evidence (Score 3)-De novo in-frame deletion variant, Genetic Association;
-Twelve rare variants
[155,156]
GRM7Glutamate metabotropic receptor 7Suggestive evidence (Score 3)gene variations and susceptibility to ASD;
Rare de novo deletion
[157,158]
GRM8Glutamate metabotropic receptor 8/-Variants in the chromosomal region 7q21–32
-Microdeletion in the chromosomal region 7q31.33q32.1
[20,152]
Table 2. Glutamatergic dysfunctions reported in pharmacological and genetic animal models of ASD.
Table 2. Glutamatergic dysfunctions reported in pharmacological and genetic animal models of ASD.
Mouse ModelAlterations in Glutamatergic SignalingRefs.
FMR1 KO
(Fragile X)
High expression of the AMPAR subunit GluR1-4 in the hippocampal CA1 pyramidal neurons and Purkinje Cells of the cerebellum.[214,215]
Lower levels of NMDAR expression.[216]
Early transient increase in AMPAR/NMDAR ratio and increased expression of the GluA2 subunit in synaptic AMPARs.[217,218]
The mGlu5 antagonist MPEP improved NMDA-mediated deficits in LTP.[8]
Altered forms of synaptic plasticity mediated by overactivation of mGlu5.[219,220]
VPA modelAlterations in E/I balance.[207]
Alterations in Glu receptor expression and synaptic transmission.[52]
Increased NMDA receptor expression and long-term potentiation.[221,222]
SHANK2 KOReduced levels of several cell surface Glu receptors in striatum and thalamus.[223]
mGlu1 antagonist ineffective in rescuing the social deficits in SHANK2 KO rats[224]
SHANK3 KOReduced levels of several cell surface Glu receptors in striatum and thalamus. Alterations in AMPAR expression.[223]
NMDAR dysfunction[225]
Altered forms of synaptic plasticity mediated by mGlu[226]
NLGN1-2-3 KOAlterations in AMPAR expression in brainstem neuronal cultures. Alterations in the frequency of spontaneous and miniature excitatory postsynaptic currents. No effect on evoked postsynaptic currents.[227]
NLGN1 KOThe NMDA co-agonist d-cycloserine reduced anxiety and repetitive behavior[228]
NLGN3-R704CAltered AMPAR-mediated currents.[229]
NLGN3-R451CIncreased NMDA/AMPA ratio in pyramidal neurons of the CA1 region of the
hippocampus.
[230]
Non-significant increase in the frequency of mEPSCs in the hippocampus.
Increased expression of the excitatory postsynaptic scaffolding proteins PSD95 and SAP-102, and the GluN2B subunit of the NMDA receptor
BTBRDecreased plasticity and excitatory postsynaptic potentials.[231]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Montanari, M.; Martella, G.; Bonsi, P.; Meringolo, M. Autism Spectrum Disorder: Focus on Glutamatergic Neurotransmission. Int. J. Mol. Sci. 2022, 23, 3861. https://doi.org/10.3390/ijms23073861

AMA Style

Montanari M, Martella G, Bonsi P, Meringolo M. Autism Spectrum Disorder: Focus on Glutamatergic Neurotransmission. International Journal of Molecular Sciences. 2022; 23(7):3861. https://doi.org/10.3390/ijms23073861

Chicago/Turabian Style

Montanari, Martina, Giuseppina Martella, Paola Bonsi, and Maria Meringolo. 2022. "Autism Spectrum Disorder: Focus on Glutamatergic Neurotransmission" International Journal of Molecular Sciences 23, no. 7: 3861. https://doi.org/10.3390/ijms23073861

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