Next Article in Journal
Analysis of DRD2 Gene Polymorphism in the Context of Personality Traits in a Group of Athletes
Next Article in Special Issue
The Genetic Diagnosis of Ultrarare DEEs: An Ongoing Challenge
Previous Article in Journal
Association between Polymorphism rs1799732 of DRD2 Dopamine Receptor Gene and Personality Traits among MMA Athletes
Previous Article in Special Issue
Study of the Interaction between Executive Function and Adaptive Behavior at School in Girls with Fragile X Syndrome
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

FMR1 and Autism, an Intriguing Connection Revisited

1
SUNY Downstate Medical Center, SUNY Downstate College of Medicine, Brooklyn, NY 11203, USA
2
Graduate Program in Neural and Behavioral Science, SUNY Downstate Medical Center, Brooklyn, NY 11203, USA
3
Rutgers Robert Wood Johnson Medical School, New Brunswick, NJ 08901, USA
4
Child Health Institute of New Jersey, New Brunswick, NJ 08901, USA
*
Author to whom correspondence should be addressed.
Genes 2021, 12(8), 1218; https://doi.org/10.3390/genes12081218
Submission received: 1 July 2021 / Revised: 3 August 2021 / Accepted: 4 August 2021 / Published: 6 August 2021

Abstract

:
Autism Spectrum Disorder (ASD) represents a distinct phenotype of behavioral dysfunction that includes deficiencies in communication and stereotypic behaviors. ASD affects about 2% of the US population. It is a highly heritable spectrum of conditions with substantial genetic heterogeneity. To date, mutations in over 100 genes have been reported in association with ASD phenotypes. Fragile X syndrome (FXS) is the most common single-gene disorder associated with ASD. The gene associated with FXS, FMR1 is located on chromosome X. Accordingly, the condition has more severe manifestations in males. FXS results from the loss of function of FMR1 due to the expansion of an unstable CGG repeat located in the 5′′ untranslated region of the gene. About 50% of the FXS males and 20% of the FXS females meet the Diagnostic Statistical Manual 5 (DSM-5) criteria for ASD. Among the individuals with ASD, about 3% test positive for FXS. FMRP, the protein product of FMR1, is a major gene regulator in the central nervous system. Multiple pathways regulated by FMRP are found to be dysfunctional in ASD patients who do not have FXS. Thus, FXS presents the opportunity to study cellular phenomena that may have wider applications in the management of ASD and to develop new strategies for ASD therapy.

1. Introduction

Neurodevelopmental disorders are a broad group of neurological and psychiatric conditions wherein the development of the central nervous system is disrupted. These disorders are typically recognized early in life and often persist into adulthood.
Of the many neurodevelopmental disorders, ASD has come to the forefront of social awareness. Recent estimates place the prevalence of Autism Spectrum Disorders (ASDs) at approximately 1 in 54 children, [1]. ASDs are characterized by persistent deficits in social communication and social interaction with restricted, repetitive patterns of behavior, interests, or activities [2]. These symptoms range from mild to severe impairments in daily functioning. It is important to note that severe impairments are more often seen in ASDs, as 78% of individuals diagnosed with ASD continue to require care into adulthood, whereas only 12% live independent lives [3]. Adding to the diversity of ASDs, a high rate of comorbidity with many other disorders appears to be the rule, rather than the exception [4]. Up to 79% of individuals with autism have motor delays, 12–70% depression, 45% intellectual disability (ID), 42–56% anxiety disorder, 9–70% gastrointestinal 0028GI) disturbances, 28–44% attention deficit hyperactivity disorder (ADHD), and 8–30% seizure disorder [5].
Twin studies have shown significant heritability of ASD [6,7], thus indicating a major contribution of genetic factors.
Nevertheless, a major obstacle in improving our understanding of the underlying pathology of ASDs has been the absence of identifiable biomarkers. With the development of advanced genomic technologies, many chromosomal copy number variants (CNV), as well as monogenetic mutations, were associated with ASD [8,9,10,11]. To date, hundreds of genes have been linked to ASD, however, most of these genes are involved in only a small percentage of cases [12,13]. A comprehensive genomic database that indexes and evaluates genetic ASD associations has been developed by the Simons Foundation Autism Research Initiative (SFARI) (https://gene.sfari.org/).
In addition to genetic links, exposures while in utero or during early development, such as toxins (e.g., lead, alcohol), medications (e.g., valproate, thalidomide), or stress, are strongly associated with neurodevelopmental disorders [14,15,16]. Risk-increasing genetic variants and early life exposures occur within unique genetic backgrounds and environmental contexts [17]. These factors combine to create an intricate etiological landscape, which presents unique challenges in comparison to other areas of medicine. This is particularly evident with regard to efforts to understand the pathophysiology and development of novel therapeutics.
A reliable biological definition of ASD would provide strong, distinct relationships between genomic variants, pathophysiological processes, and clinical phenotypes. However, many of the identified risk genes or environmental exposures are only found in a subset of patients and often share correlations with multiple disorders [18,19,20]. A similar situation is encountered with various pathophysiology at the cellular, circuit, and network levels [21].
A number of syndromic monogenetic neurodevelopmental disorders such as Fragile X Syndrome (FXS), Rett syndrome (RTT), and tuberous sclerosis (TSC), have strong associations with ASD [22,23]. On the surface, it appears logical to hypothesize that discrete genetic causes of neurodevelopmental disorders would provide relatively straightforward answers to questions about the underlying neuropathology. On the contrary, even in the case of monogenetic disorders, complex processes govern the downstream effects, as the loss of a single gene results in drastic effects on the activity of a multitude of other cellular processes [24,25].
Even though monogenetic neurodevelopmental disorders are distinct clinical entities, these disorders have provided important insights into potential common-shared pathophysiology and may provide a guide for unraveling the molecular underpinnings of non-syndromic autism [26].
One of the most important examples of syndromic ASD is Fragile X Syndrome. Approximately 30 to 50% of individuals diagnosed with FXS meet the criteria for a diagnosis of ASD, with FXS-ASD patients composing approximately 3% of all cases of ASD [27,28,29,30]. The co-occurrence of syndromic disorders, such as FXS, and the high level of heritability seen in twin studies, were among the first findings to make clear the importance of genes in the etiology of ASD [17,31]. FXS is the most common single-gene defect identified in patients with ASD. Accordingly, testing for the fragile X mental retardation 1 (FMR1) gene mutation, the mutation responsible for FXS, is recommended as a first-tier genetic test in the current expert guidelines for ASD management [22].

2. Fragile X Syndrome

Fragile X syndrome (FXS) is a neurodevelopmental disorder due to an X-linked mutation in the FMR1 gene. The overall prevalence of FXS is approximately 1 in 7000 in males and 1 in 11,000 in females [32,33]. In more than 95% of known cases, the FXS phenotype is due to an expansion of more than 200 repeats and the subsequent methylation of CGG triplets in the 5′ untranslated promoter region of the FMR1 gene [22]. The remaining 5% of FXS cases, in which triplet repeat expansions are not found, are often due to point mutations or deletions in the FMR1 gene which, as in the other 95% of cases, result in absent or markedly decreased production of the fragile X mental retardation protein (FMRP). FMRP regulates the translation of approximately 4% of fetal brain mRNA and directly regulates several classes of ion channels [34,35,36,37]. Clinically, FXS patients present with distinct physical and behavioral features. Physically, individuals have characteristic facial abnormalities (e.g., elongated face, large ears), macroorchidism, hyperlaxity of joints, and hypotonia. Behaviorally, male FXS patients typically have intellectual disability (ID), anxiety, attention deficit hyperactivity disorder (ADHD), and sensory processing deficits, while female patients have more variable manifestations [22].

2.1. FMR1

Studies from the 1980s identified nonpenetrant male carriers in families with FXS, indicating a unique pattern of inheritance for the syndrome [38]. This peculiar pattern of inheritance remained unresolved until the FMR1 gene was identified in 1991 by positional cloning [39]. The 5′-untranslated promoter region for the FMR1 gene typically contains less than fifty-five CGG trinucleotide repeats and is located at Xq27.3 (Figure 1). Individuals with the full FMR1 mutation have more than 200 repeats which result in methylation and subsequent gene silencing. Intermediate repeats of 55–200 are defined as premutation [22]. The presence or absence of FXS phenotypes from generation to generation is caused by anticipation, wherein an expansion of the premutation occurs during meiosis and produces the full mutation. Mothers with a premutation of greater than 90 to 100 repeats have up to 50% risk of passing on the full mutation to each of their children [40]. Males with full mutations are affected with FXS and typically present with intellectual disability, anxiety disorder, and attention deficit hyperactivity disorder, while full mutation females have variable and often milder manifestations (Hagerman et al., 2017). Due to the X-linked nature of this syndrome, females with the full mutation retain a functional copy of the FMR1 gene and thus have milder phenotypes than their male counterparts. As a consequence, full mutation females are often diagnosed with FXS only after confirming the condition in a male relative, while some may never be identified [41]. Individuals with the premutation do not express the severe phenotypes seen with the full mutation however several disorders are associated with premutation carriers, namely, fragile X-associated tremor and ataxia syndrome (FXTAS), fragile X associated primary ovarian insufficiency (FXPOI), and fragile X premutation associated conditions (FXPAC)) [24].

2.2. FMRP

The product of the FMR1 gene is the fragile X mental retardation protein (FMRP). This protein is highly conserved among vertebrates (~92% similarity between human and chicken homologs) and thus speaks to the importance of its biological activity [42]. FMRP is a 71 kDa protein containing a nuclear localization signal (NLS) domain, a nuclear export signal (NES) domain, three hnRNP-K-homology (KH) domains (KH0, KH1, KH2), and an arginine-glycine-glycine (RGG box) [43,44,45]. There are 12 identified isoforms of the FMRP protein that are produced as a result of alternative splicing [46]. The level of expression of each isoform varies during development and modulates binding affinities to ribosomes.
FMRP is diffusely expressed throughout the brain and varies widely among subpopulations of neurons [47]. The levels of expression of FMRP change dramatically during development, with FMRP expression reaching maximal levels early in the postnatal period, typically peaking in the early development and declining across early life [48,49]. Developmental studies with mice show a peak level of expression in the hippocampus, cerebellum, striatum, and brainstem, between postnatal days (P) 3 and P7 [50]. In the auditory and somatosensory cortex, peak FMRP levels are seen between P7 and P12. Unfortunately, differential studies of cell types and FMRP expression in the developmental brain have yet to be performed.
Within a given neuron, FMRP localizes to the soma, axon, and dendritic compartments [51]. FMRP assembles with RNAs, other binding proteins, and the homologous proteins FXR1P and FXR2P, to form large ribonucleoprotein complexes [52]. These complexes regulate the transport, translation, and metabolism of mRNAs [53]. This activity is crucial for appropriate neuronal development, synaptic connectivity and plasticity, and dendritic architecture [54,55,56,57].
Canonically, FMRP’s main function is that of an mRNA binding protein, responsible for the localization, stabilization, and translation of approximately 4% of fetal brain mRNA [37,39,58,59]. FMRP binds mRNAs through purine quartet motif regions and interacts with ribosomes via its KH domains [43,59,60,61]. Many of the mRNA transcripts that FMRP regulates are responsible for the production of synaptic proteins that are crucial for the high fidelity information processing at the synapse [57].
Non-canonical functions of FMRP, such as nuclear functions, microRNA interactions, and protein–protein interactions have been identified [34,35,36,37]. Of these functions, FMRP plays a critical role in modulating the behavior of several classes of neuronal voltage-gated ion channels [34,35,62]. Potassium channels such as Kv4.2, large conductance voltage and calcium-sensitive K+ (BKCa) channels, and Slack [34,63,64], calcium channels CaV1.2, CaV1.3 and CaV2.2, [36,65,66,67], and non-selective hyperpolarization-activated and cyclic nucleotide-gated (HCN) channels [68,69] are modulated by FMRP activity. In the case of BKCa, Slack, CaV1.2, and CaV2.2, FMRP directly complexes with these channels to regulate their function.

3. FMR1 and Autism Spectrum Disorder: Synaptic Dysfunction

There are many paradigms that can be used for developing research questions about the etiology of ASD. Currently, there is a vast array of relationships between ASD and disruptions in developmental processes due to genetic or environmental insults. Many of these perspectives provide useful frameworks for guiding research into potential causative mechanisms of ASD. One of the major areas of disruption in both FXS and ASD is found at the synapse [70,71,72,73,74] This has led to the proposal that ASD be conceptualized as a “synaptic disease” [73,75].
The loss of FMRP results in excess and dysregulated mRNA translation, delocalization of FMRP regulated proteins, and thus profound changes in the structure and physiology of the synapse [37,57]. A pivotal expansion in our understanding of the pathological effects of the FMR1 mutation comes from the “mGluR theory” of FXS [76]. Type 1 metabotropic glutamate receptors (mGluR1 or 5) are G coupled protein receptors (GCPR) that are located post-synaptically and regulate multiple cellular signaling pathways [77]. Stimulation of mGluR5 receptors induces FMRP translation at the synapse and FMRP functions as a repressor of protein synthesis [51]. In the Fmr1-KO mouse, an mGluR5 regulated form of synaptic plasticity, long-term depression (LTD) is exaggerated [78]. The pathology seen in Fmr1-KO mice is reflected in both FXS and non-syndromic ASD patients, as alterations in mGluR5 expression are seen in postmortem ASD brains [79,80]. Additionally, high-throughput sequencing of mGluR signaling pathway genes has detected enrichment of rare variants among ASD patients [81]. Since dysfunctional mGluR activity is present in FXS and some ASD patients, this has facilitated detailed investigations into downstream components of mGluR5 signaling. Targeted mutations of mGluR5 scaffolding proteins such as Homer1a, Shank3, Ngln3 produced phenotypes that approximate those seen in FXS and ASD [82,83,84]. Importantly, the mGluR hypothesis established a framework for investigations of synaptic dysfunction.
Many of the synaptic deficits which occur due to FMR1 mutations are in mechanisms of plasticity, such as α-amino-3-hydroxy-5-methyl-4-isoxazole-propionic acid (AMPA) receptor expression, N-methyl-D-aspartate (NMDA) receptor localization, and endocannabinoid (eCB) signaling [82,85,86,87,88]. Notably, mutations in many of these components, particularly FMRP, mGluR, and NMDAR, disrupt critical regulatory mechanisms such as eCB regulation of presynaptic activity [82,85,87,89]. These deficits are also found in the Fmr1-KO mouse model and importantly, genetic variants for these proteins are found in non-syndromic ASD patients and are predictive of an ASD diagnosis [90,91,92,93].
Synaptic dysfunction can be subdivided into various subcategorizations (e.g., channelopathies) [64,94]. Each of these subcategorizations provides a framework for generating hypotheses and may be useful for identifying causal mechanisms and potential treatment targets. Here we focus on a subset of presynaptic regulatory mechanisms associated with FMR1 and ASD. More specifically this review highlights the relationship between FMR1 mutations and ASD with regard to the dysfunctional regulation of presynaptic activity by the endocannabinoid system (ECS), BKCa channels, and CaV2.1 and CaV2.2 channels.

4. The Presynaptic Hypothesis of FMR1 and ASD

Of the many pathophysiological processes associated with FMR1 mutations and ASD, those which are imperative for appropriate presynaptic activity have a substantial body of clinical and preclinical evidence implicating them in the pathology of both disorders. At the synaptic level, presynaptic dysregulation results in aberrant neurotransmitter release and altered synaptic plasticity, which underlies the hyperexcitability seen at the circuit level [36,63,64,65,87]. Dysfunctional local circuits may underlie larger scale brain network dysfunction (e.g., connectopathy), and are often detected in ASD patients [95,96,97]. Here we review, several selected presynaptic regulatory components associated with FMR1 mutations and ASD: the endocannabinoid system (ECS), and BKCa channels, and, P/Q and N-type Ca2+ channels [36,65,98,99,100].
In essence, the presynaptic hypothesis posits that presynaptic dysregulation causes computational errors at the synaptic level, resulting in circuit level and systems-level brain network dysfunction that manifests as neurodevelopmental pathology.

Overview

The ECS, BKCa channels, and CaV2.1 and CaV2.2 channels are regulated by FMRP activity [35,36,65,85,101,102,103]. In the case of BKCa channels and CaV2.1 and CaV2.2 channels, FMRP interacts with these channels directly to inhibit calcium entry into the cell [35,104]. With regard to the ECS, FMRP controls the localization and translation of mRNA for DGL-α, the enzyme responsible for the production of the primary eCB, 2-AG [85]. These components make important contributions to the regulation of neurotransmitter release from the presynaptic neuron, each of which is linked to the FMRP activity, and along with FMRP itself, inhibits presynaptic CaV2.1 and CaV2.2 channels [36,65,98] (Figure 2A). When FMRP function is lost as a result of FMR1 mutations, or when there is loss of function in the aforementioned FMRP associated synaptic components, dysregulation of calcium dynamics occurs resulting inappropriate neurotransmitter release (Figure 2B) [35,36,65,84,86,101,105,106,107].
The presynaptic hypothesis is a reductionist paradigm for guiding research directed at a small subset of pathogenetic associations. It is not exclusive of other paradigms, or dysfunction in other pathways not discussed here. ASD can, and does, arise from numerous developmental insults. Each of these components has notable pre-clinical and clinical evidence linking them to lost FMRP function and ASD [36,65,85,90,100,109,110,111]. Importantly, these components can be manipulated genetically and pharmacologically to induce or rescue some ASD phenotypes, which makes them attractive targets for inquiry into underlying pathology, and potential therapies [85,99,108,112,113].

5. ECS

The endocannabinoid system (ECS) is composed of two primary cannabinoid receptors, cannabinoid type 1 receptor (CB1) and cannabinoid type 2 receptor (CB2), and two primary ligands, arachidonoyglycerol (2-AG) and N-arachidonoylethanolamine (AEA) [114,115,116]. CB1, a G-coupled protein receptor (GCPR), is expressed extensively in the central nervous system, with higher levels of expression found in the hippocampus, amygdala, striatum, and cortex [117]. CB2, also a GCPR, is expressed at low levels in the CNS and largely on microglial cells where they mediate immune responses [115,118]. Endocannabinoids are hydrophobic lipids that are biosynthesized and released on demand, unlike the majority of neurotransmitters, which are water-soluble, synthesized in advance, and stored in vesicles [119].
Of the two endocannabinoid ligands, 2-AG is the most abundant found in the mammalian CNS and is a full agonist at CB1 [120,121,122]. 2-AG synthesis follows two distinct mechanisms: First (eCBmGluR), activation of group I mGluR which activates phospholipase C β (PLC-β) to cleave phosphatidylinositol-4,5-bisphosphate (PIP2) to produce the 2-AG precursor, 1,2-diacylglycerol (DAG). This is hydrolyzed by the serine lipase, diacylglycerol lipase α (DGL-α) in central neurons and diacylglycerol lipase β (DGL-β) in immune cells (e.g., microglia, macrophages), to form 2-AG [123,124]. The second mechanism for 2-AG synthesis is dependent on rapid increases of intracellular Ca2+ via NMDA receptors (eCBNMDA) [125,126]. PLC-β is activated in a Ca2+ dependent manner and produces the precursor, DAG, needed for the production of 2-AG by DGL-α [127]. The synthesis of 2-AG in post-synaptic neurons occurs within a supramolecular complex wherein mGluR5 receptors are bound to Homer1a scaffolding proteins which also bind PLC-β and DLG-α resulting in rapid and spatially localized 2-AG synthesis [85]. Approximately 85% of 2-AG is hydrolyzed into arachidonic acid (AA) and glycerol the presynaptic enzyme monoacylglycerol lipase (MAGL), with the remaining 15% metabolized by the enzymes α/β-hydrolase-6 (ABHD6) and α/β-hydrolase-12 (ABHD12) [128,129]. The second endocannabinoid, AEA, is a partial agonist at CB1 [130,131]. AEA synthesis occurs in a Ca2+ dependent manner, in response to an influx on intracellular Ca2+ causes cleavage of phosphatidylethanolamine (PE) by N-acetyl-transferase into the AEA precursor, N-arachidonoyl-PE (NAPE), which is then cleaved by the NAPE-hydrolyzing phospholipase D (NAPE-PLD) into AEA. Metabolism of AEA is carried out by fatty acid amide hydrolase (FAAH) which hydrolyzes AEA into AA and ethanolamine (EA). Once synthesized, 2-AG and AEA diffuse retrosynaptically and interact with CB1 receptors located on the presynaptic neuron [130].
CB1 signaling by 2-AG or AEA can result in the activation of multiple signaling pathways mediated by the Gi/o protein subunits of CB1. CB1 activation inhibits adenylyl cyclase and reduces cAMP production [132]. Activation by CB1 agonists also induces mitogen-activated protein kinase (MAPK) and PI3K/AKT pathways which control gene transcription and cellular activity [133]. Crucially, CB1 inhibition of neurotransmitter release, responsible for synaptic plasticity, is mediated by Gi/o protein inhibition of presynaptic CaV2.1 and CaV2.2 channels [98].
The retrograde nature of the ECS provides a unique form of synaptic plasticity called depolarization-induced suppression of inhibition (DSI) at inhibitory GABAergic synapse and depolarization-induced suppression of excitation (DSE) at excitatory synapses [134,135]. Briefly, depolarization at the post-synaptic neuron induces the production of eCBs which act retrosynaptically to inhibit neurotransmitter release from the presynaptic neuron.
Endocannabinoids also exhibit activity at transient receptor potential cation channel subfamily V member 1 (TRPV1), G protein-coupled receptor 18, 55, and 119 (GPR 18; GPR55; GPR119) [136,137]. While the activity of these ligand–receptor interactions is not yet fully understood, it has been shown that signaling at these receptors with exogenous cannabinoids may mediate some of the anxiolytic and anti-epileptic properties of these molecules [138,139]. The ECS also has a critical developmental role, as, during gestation, DGL-α mediated 2-AG-CB1 signaling is necessary for appropriate neurogenesis, neuronal migration, and axonal targeting [140,141].

5.1. ECS, FMR1, and ASD—Clinical Correlates

A growing body of clinical evidence associates the ECS with ASD phenotypes. Post-mortem studies of brain tissue from ASD patients indicated reduced expression of CNR1, the gene for CB1R [142]. Additionally higher expression of CB2R has been found to be upregulated in some children with ASDs [143]. Analysis of multiple genomic databases found variants in CNR1 and DAGLA, the gene for DGL-α, were significantly associated with autism [144]. A series of studies investigated gaze to facial stimuli, a behavior frequently altered in FXS and ASD patients and found that polymorphisms in the CNR1 gene modulate striatal responses and gaze duration to happy faces [145,146]. Two recent studies detected lower levels of circulating endocannabinoids in ASD patients relative to controls [147,148]. Risk-increasing variants for ASD, that are also associated with FXS, have been detected in synaptic proteins important for ECS function such as GRM5, NGLN3, HOMER1A, SHANK3, and several genes coding for NMDAR subunits [81,90,93,149,150,151,152,153,154]. Given the known role that mGluR5 dysfunction plays in FXS pathology and ECS activity, it is important to note that mutations in GRM5, the gene for mGlur5, are risk variants for ASD [81]. Furthermore, alteration in both mGluR5 and FMRP expression and signaling has been detected in non FXS ASD patients [79,80,155,156].

5.2. ECS, FMRP, and ASD—Preclinical Studies

Studies with the Fmr1-KO mice consistently show evidence of ECS dysfunction [85,87,101,102]. FMRP binds the mRNA of DGL-α and controls its appropriate translation and localization at the postsynaptic density (PSD) [85]. Loss of FMRP expression resulted in delocalization of DGL-α and dysfunctional 2-AG mediated plasticity. It was demonstrated that, in the absence of FMRP, mGluR5 stimulation fails to induce 2-AG production in the prefrontal cortex (PFC), and thus the mGluR hypothesis of FXS is tied to dysfunctional eCB activity. Of note, in utero exposure to valproate, a known risk factor for ASD, is linked to decreased levels of DGL-α mRNA [157].
Molecular and physiological studies indicated that appropriate eCBmGluR production requires a scaffolding protein called Homer1a, which complexes mGluR5 and DGL-α [87,158]. In Fmr1-KO mice, interactions between mGluR5 and Homer1a are reduced and this is causal for hyperexcitability of cortical neurons and seizures [159]. Homer1a proteins also mediate mGluR5 and NMDA interactions, likely coordinating eCBmGluR and eCBNMDA forms of 2-AG synthesis [86,160]. These interactions are disrupted in Fmr1-KO mice and upregulation of Homer1a expression rescued cognitive deficits. Importantly, increasing the bioavailability of 2-AG normalized plasticity deficits and rescued the hyperactive, anxiety, and cognitive impairments phenotypes of the Fmr1-KO mouse [85,161]. Furthermore, enhancement of AEA availability rescued deficits in social approach, memory, and deficit frequently found in Fmr1-KO mice [112,162,163,164].
Pharmacological and genetic manipulations of specific ECS components strengthen the link between the ECS and ASD pathology. Mice with a targeted DGL-α deletion from direct pathway medium spiny neurons (dMSNs) of the striatum had impaired social interest and increased repetitive behaviors [165]. Mice with global DGL-α deletion showed increased anxiety, stress, and fear responses [166,167]. Additionally, optogenetic activation of basolateral amygdalar glutamatergic circuits, circuits that are inhibited by DGL-α mediated 2-AG production, caused social deficits in mice [168]. Importantly, pharmacological augmentation of 2-AG levels blocked these deficits in social behavior from occurring. Pharmacological inhibition of DGL-α activity at adulthood caused social impairments, a communication deficit, and repetitive behavior in C57BL6 mice [108].
Developmental studies show that a temporally orchestrated pattern of ECS expression and activity is imperative for appropriate brain connectivity [140,141,169,170,171]. The results of a postmortem study of brain tissue from various developmental times points revealed that CB1 and the enzymes responsible for endocannabinoid synthesis and metabolism (e.g., DGL-α, MAGL, FAAH) have distinct patterns of expression across development, particularly during neonatal and infancy age ranges [172]. This is further demonstrated by mouse studies wherein mice null for the CB1R have altered brain connectivity [173,174]. This appears to approximate a neurobiological phenotype frequently seen in patients with ASDs [97,175,176]. Genetic deletion of CB1 expression revealed deficits in social behavior, cognition, and repetitive behaviors [177,178,179]. Selective deletion of CB1 revealed that a loss of CB1 in glutamatergic, but not GABAergic, cortical neurons resulted in a reduction of social interest [180].

5.3. ECS and ASD—Potential Therapeutics

Clinically, phytocannabinoids (pCBs), plant-derived molecules with similar chemical structures as eCBs, have demonstrated success in the treatment of neurodevelopmental disorders. The pCB, cannabidiol (CBD) has FDA approval for the treatment of two forms of epilepsy: Dravet Syndrome and Lennox-Gastaut Syndrome [181,182]. In regard to FXS, a phase 1/2 study with CBD and FXS patients found that 12 weeks of treatment produced substantial reductions in hyperactivity, social avoidance, anxiety, and compulsive behavior [183]. Importantly the frequency of adverse events was low, and no serious adverse events were reported. Additionally, several prior case reports of FXS patients and CBD treatment reported improvement of symptoms [184]. Evidence indicates that CBD may be useful as a treatment for non-syndromic ASD [185,186,187,188]. Studies with the pCBs cannabidiol (CBD) showed an improvement in aggression, hyperactivity, sleep problems, speech impairment, seizures, and anxiety in ASD patients [185,189].
Cannabidivarin (CBDV), a propyl analog of CBD, has also shown promise as a treatment for ASDs [190]. A small study of 17 ASD patients and 17 matched non-ASD controls found that CBDV produced unique neurobiological effects in glutamate metabolism in the basal ganglia of ASD patients [191]. CBDV treatment in a mouse model of RTT rescued the social deficits that arise as a result of the Mecp2 mutation [192]. Additionally, CBDV treatment in the valproic acid rodent model of ASD rescued ASD-like behaviors and restored ECS activity in the hippocampus [193]. Currently, a clinical trial, funded by the United States Department of Defense, is underway for CBDV treatment in ASD patients (Clinicaltrial.gov; NCT03202303).
Importantly, these molecules largely avoid the undesired psychotropic side effects that result from CB1 activation, strengthening their appeal as potential treatments for ASDs [194]. Studies indicate that these molecules act on the ECS, however, the mechanism of action for pCBs is not well understood and requires further studies [138,195,196,197,198,199].

6. Large Conductance Voltage and Ca2+ Sensitive K+ (BKCa) Channels

Large conductance voltage and calcium-sensitive potassium (BKCa) channels are expressed ubiquitously throughout the body, however, regulatory subunits of these channels are tissue-specific [200]. In the central nervous system, the β4 regulatory subunit is referred to as the neuronal auxiliary subunit and is the most abundant of the subunits expressed with BKCa channels in central neurons [201,202]. In the CNS, BKCa channels are expressed in most brain regions at presynaptic terminals, however higher levels of expression are found in the cortex, basal ganglia, hippocampus, and cerebellum [201,203].
Functionally, the α subunit of the BKCa channel opens in response to membrane depolarization and intracellular increases in Ca2+ [204]. It has a bimodal response to these events; opening to allow a large efflux of K+ ions (thus hyperpolarizing the membrane) and complexing with P/Q and N-type Ca2+ channels to inhibit Ca2+ entry and control neurotransmitter release [205,206]. Of these two stimuli, Ca2+ entry is the rate-limiting step for BKCa activation [207]. FMRP regulates the Ca2+ sensitivity of BKCa channels through direct interactions with the α and β4 subunits [35,208]. This reduces action potential duration, controlling neurotransmitter release and repetitive neuronal activity.

6.1. BKCa, FMR1, and ASD—Clinical Correlates

Genetic studies have uncovered a relationship between genetic variants for BKCa genes and ASD. Skafidas et al., 2014 [90] examined the occurrence of specific single nucleotide polymorphisms (SNPs) and a diagnosis of ASD. A genetic diagnostic classifier of 237 SNPs in 146 genes was used with 85.6% accuracy in predicting a diagnosis of ASD in a cohort of central European individuals gathered from two different databases: SFARI and Wellcome Trust 1958 Normal Birth Cohort (WTBC) databases. Two of the SNPs determined to be most effective at determining a classification of non-syndromic ASD vs. non-ASD were found in the KCNMB4 gene, (β4 BKCa subunit), and GRM5 gene, (mGluR5) were two of the three identified genes. This is particularly important in regard to the overlap between FXS and ASD since BKCa channel activity is directly regulated by FMRP at the β4 unit and mGluR5 dysfunction in FXS has been well established [35,76,99].
Two studies that investigated chromosomal abnormalities in ASD patients discovered a mild to moderate association between mutations in KCNMA1 and a diagnosis of autism [209,210]. Additionally, mutations in the KCNMA1 gene were identified in two patients with ASD and intellectual disability [100]. Genome analysis of the first patient discovered a balanced de novo translocation (9q23/10q22) resulting in haploinsufficiency for the α subunit, while the second patient revealed a single point mutation in the KCNMA1 gene which resulted in an ALA138VAL substitution.
BKCa dysfunction is also associated with other neurodevelopmental disorders. A patient with moderate to mild intellectual disability and febrile seizures was identified as having a mutation only in the β4 BKCa regulatory domain of FMRP [211]. Analysis of the family found a maternal and paternal history of learning problems, however, this specific mutation, being X-linked, was found only in the maternal genome.

6.2. BKCa, FMR1, and ASD—Preclinical Studies

Studies with the Fmr1-KO mouse demonstrated that loss of FMRP regulation of BKCa channels increased action potential duration [35,212]. Specifically, loss of FMRP increased the after-hyperpolarization phase (AHP) of the action potential, increasing neuronal excitability, presynaptic Ca2+ influx, and neurotransmitter release. Zhang et al., 2014 [64] showed that loss of FMRP was also responsible for downregulation of BKCa channel expression in the Fmr1-KO mice. Importantly, genetic upregulation of BKCa channels expression normalized the synaptic and circuit deficits in the Fmr1-KO mouse [212]. These factors were determined to be contributory for the sensorimotor hypersensitivity phenotype in the Fmr1-KO mouse, a frequently co-morbid feature of ASD. A genetic mouse model null for the BKCa α subunit gene (Slo1) was developed to explore the role of BKCa channels in neurodevelopmental disorders [213]. This study found that mice null for BKCa α expression had impaired sensorimotor and spatial memory, with normal locomotor activity. Currently, phenotyping of the social behaviors of the BKCa−/− mouse has not been performed. Our recent pharmacological study using the BKCa channel inhibitor paxilline found that paxilline treatment induced unique social anxiety-type deficits during adulthood [113].
BKCa channels expressed outside the central nervous system respond to eCB signaling in vascular endothelial cells [214]. Additionally, in the trabecular meshwork of the eye stimulation of CB1 receptors was coupled to the activation of BKCa channels [215]. To the best of our knowledge, interactions between the ECS and CB1 in central neurons have not been investigated, and thus represent an area for future studies.

6.3. BKCa, FMR1, and ASD—Therapeutics

A BKCa channel agonist, BMS-204532 (BMS), was developed in 2002 for the treatment of ischemic stroke, however, it failed to demonstrate clinically significant therapeutic effects in phase III trials [216]. Since BMS has a favorable safety profile it is currently under investigation as a treatment of BKCa channelopathies. Detailed analyses of cells cultured from patients with ASD and BKCa mutations demonstrated that channel function could be rescued by BMS [100].
Studies with the Fmr1-KO mouse have demonstrated promise for BMS as a therapeutic for FXS. In an initial study, BMS treatment rescued social, cognitive, and anxiety phenotypes and normalized dendritic morphology in the Fmr1-KO mouse [99]. Two subsequent studies have demonstrated that BMS can rescue dendritic hyperexcitability and the increased self-grooming and sensorimotor hypersensitivity phenotypes of the Fmr1-KO mouse [64,217]. One of the challenges in using BMS clinically is the short half-life in brain tissue (t1/2 = 1.9) [216]. This would result in a difficult dosing schedule and therefore additional development is needed for molecule refinement. Despite these challenges, these preclinical studies strongly suggest that BMS or a next-generation BMS-derived molecule could provide a pharmacological intervention.

7. CaV2.2

CaV2.2 channels are key components of neurotransmission in the central nervous system and in the autonomic and sensory nervous system, and play a key role in early development [218,219]. These channels are expressed throughout the brain [220]. It is largely the α1 subunit that determines the kinetic and voltage-dependent properties of these channels [221]. FMRP interacts via its C-terminal domain to the linker region between the II and III domains of the α1 subunit [36,104]. This functions to control Ca2+ currents by reducing the expression of CaV2.2 channels via proteasomal degradation. A follow-up study found that loss of FMRP resulted in increased Cav2.2 channel currents due to increased surface expression [65]. This contributed to neuronal hyperactivity.

CaV2.2 and ASD—Clinical Correlates

There are numerous links with sequence variants in voltage-gated calcium channel genes and ASD. Iossifov et al., 2014 [20] identified a de novo variant in the CACNA2D3 associated with ASD, a gene that acts as a regulatory subunit for CaV2.2. A study of 20 ASD patients identified a duplication of the chromosomal region 9q43.3, which contains the gene CACNA1B, which produces Ca2+ currents in CaV2.2 channels [222]. This duplication was found in 12 of the 20 patients. Altered Ca2+ activity itself has been identified in ASDs. A small study of six ASD patients with six age-matched controls found significantly higher levels of Ca2+ in ASD patients [223]. These findings established a link between ASD and CaV2.2 channels and highlight their physiological importance in neuronal functions.
Due to the widespread expression of CaV2.2 channels in the body, and the effects that calcium channel-directed medications have on the cardiovascular system, the development of therapeutics which seek to target these channels in central neurons face significant challenges. However, the connection to FMRP, ECS, and BKCa channels have with alterations in CaV2.2 function and presynaptic activity represents an important pathway that may have therapeutic implications for ASD.

8. Clinical and Therapeutic Implications

In this article, we have reviewed a specific subset of deficits connected by FMR1 mutations and ASD. Specifically, a set of neural mechanisms disrupted by the FMR1 mutation that result in presynaptic dysregulation, and also share associations with ASD independently of FMR1. The link between these systems and neurodevelopmental disorders, specifically FXS and ASD, is a relatively recent area of research [190,224,225,226]. These mechanisms are amenable to both genetic and pharmacological manipulation, and thus, present an intriguing opportunity for elucidating a subset of causative mechanisms for ASD. Importantly, the ECS and BKCa channels show promising evidence that indicates potential as therapeutic targets for ASD.
Future studies are needed to explore more detailed mechanistic questions surrounding the presynaptic hypothesis of ASD with regard to the ECS and BKCa channels. Questions such as, Do the ECS and BKCa channels interact? It is well established that they each regulate the same presynaptic Ca2+ channels. It has not been explored if eCBs have activity at BKCa channels in central neurons, however, eCBs can modulate BKCa channels in cell culture [215]. Developmental studies addressing questions regarding FMRP expression, ECS, BKCa, and Cav2.2 channel activity during critical periods in development are sorely needed. Additionally, it is unknown if specific genetic variants in these FMR1 related regulatory mechanisms are able to produce a model of the spectrum of ASD phenotypes. In the case of FMR1, unique mutations that interfere with FMRPs regulatory function on BKCa are associated with unique neurodevelopmental phenotypes associated with ASD [100,211].
Due to the high level of heterogeneity seen in autism, it is imperative that systems that are mechanistically linked, and shown to modulate a spectrum of behavior, be thoroughly studied. This is particularly crucial in regard to ASD. Novel methods for modeling this disorder are needed, as are therapeutics. Therefore, the development of new models, and the identification of associated genetic variants in the population, is critical for improving our understanding of this complex and diverse disorder. The dysregulated presynaptic regulatory mechanisms discussed in this review present numerous targets which could be approached systematically to answer questions about presynaptic calcium dynamics and spectrum-like phenotypes that arise either due to FMR1 mutations, or direct insults to regulatory systems.

Author Contributions

Conceptualization, W.F. and M.V.; methodology, W.F. and M.V.; formal analysis W.F. and M.V.; investigation, W.F. and M.V.; resources, W.F. and M.V.; data curation, W.F. and M.V.; writing—original draft preparation, W.F. and M.V.; writing—review and editing, W.F. and M.V.; visualization, W.F. and M.V.; supervision, W.F. and M.V. Both authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by funds from the Rutgers Robert Wood Johnson Medical School, the Child Health Institute of New Jersey and from the School of Graduate Studies at SUNY Downstate Health Sciences University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Maenner, M.J.; Shaw, K.A.; Baio, J.; Washington, A.; Patrick, M.; DiRienzo, M.; Christensen, D.L.; Wiggins, L.D.; Pettygrove, S.; Andrews, J.G.; et al. Prevalence of Autism Spectrum Disorder Among Children Aged 8 Years—Autism and Developmental Disabilities Monitoring Network, 11 Sites, United States, 2016. MMWR Surveill. Summ. 2020, 69, 1–12. [Google Scholar] [CrossRef]
  2. American Psychiatric Association. Diagnostic and Statistical Manual of Mental Disorders (DSM-5®); American Psychiatric Publishing, Inc.: Washington, DC, USA, 2013. [Google Scholar]
  3. Billstedt, E.; Gillberg, C.; Gillberg, C. Autism after adolescence: Population-based 13-to 22-year follow-up study of 120 individuals with autism diagnosed in childhood. J. Autism Dev. Disord. 2005, 35, 351–360. [Google Scholar] [CrossRef] [PubMed]
  4. Gilger, J.W.; Kaplan, B.J. Atypical brain development: A conceptual framework for understanding developmental learning disabilities. Dev. Neuropsychol. 2001, 20, 465–481. [Google Scholar] [CrossRef] [PubMed]
  5. Lai, M.-C.; Lombardo, M.V.; Baron-Cohen, S. Autism. Lancet 2014, 383, 896–910. [Google Scholar] [CrossRef]
  6. Tick, B.; Bolton, P.; Happe, F.; Rutter, M.; Rijsdijk, F. Heritability of autism spectrum disorders: A meta-analysis of twin studies. J. Child Psychol. Psychiatry 2016, 57, 585–595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Hallmayer, J.; Cleveland, S.; Torres, A.; Phillips, J.; Cohen, B.; Torigoe, T.; Miller, J.; Fedele, A.; Collins, J.; Smith, K. Genetic heritability and shared environmental factors among twin pairs with autism. Arch. Gen. Psychiatry 2011, 68, 1095–1102. [Google Scholar] [CrossRef]
  8. Velinov, M. Genomic Copy Number Variations in the Autism Clinic—Work in Progress. Front. Cell. Neurosci. 2019, 13. [Google Scholar] [CrossRef] [PubMed]
  9. Bitar, T.; Hleihel, W.; Marouillat, S.; Vonwill, S.; Vuillaume, M.L.; Soufia, M.; Vourc’h, P.; Laumonnier, F.; Andres, C.R. Identification of rare copy number variations reveals PJA2, APCS, SYNPO, and TAC1 as novel candidate genes in Autism Spectrum Disorders. Mol. Genet. Genom. Med. 2019, 7, e786. [Google Scholar] [CrossRef] [PubMed]
  10. Sebat, J.; Lakshmi, B.; Malhotra, D.; Troge, J.; Lese-Martin, C.; Walsh, T.; Yamrom, B.; Yoon, S.; Krasnitz, A.; Kendall, J. Strong association of de novo copy number mutations with autism. Science 2007, 316, 445–449. [Google Scholar] [CrossRef] [Green Version]
  11. 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]
  12. Abrahams, B.S.; Geschwind, D.H. Advances in autism genetics: On the threshold of a new neurobiology. Nat. Rev. Genet. 2008, 9, 341–355. [Google Scholar] [CrossRef] [Green Version]
  13. State, M.W.; Levitt, P. The conundrums of understanding genetic risks for autism spectrum disorders. Nat. Neurosci. 2011, 14, 1499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Strömland, K.; Nordin, V.; Miller, M.; Akerström, B.; Gillberg, C. Autism in thalidomide embryopathy: A population study. Dev. Med. Child Neurol. 1994, 36, 351–356. [Google Scholar] [CrossRef] [PubMed]
  15. Nanson, J.L.; Bolaria, R.; Snyder, R.E.; Morse, B.A.; Weiner, L. Physician awareness of fetal alcohol syndrome: A survey of pediatricians and general practitioners. Can. Med. Assoc. J. 1995, 152, 1071. [Google Scholar]
  16. Christensen, J.; Gronborg, T.K.; Sorensen, M.J.; Schendel, D.; Parner, E.T.; Pedersen, L.H.; Vestergaard, M. Prenatal valproate exposure and risk of autism spectrum disorders and childhood autism. JAMA 2013, 309, 1696–1703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Folstein, S.; Rutter, M. Infantile autism: A genetic study of 21 twin pairs. J. Child Psychol. Psychiatry 1977, 18, 297–321. [Google Scholar] [CrossRef] [PubMed]
  18. Gilissen, C.; Hehir-Kwa, J.Y.; Thung, D.T.; van de Vorst, M.; van Bon, B.W.; Willemsen, M.H.; Kwint, M.; Janssen, I.M.; Hoischen, A.; Schenck, A.; et al. Genome sequencing identifies major causes of severe intellectual disability. Nature 2014, 511, 344–347. [Google Scholar] [CrossRef] [PubMed]
  19. De Rubeis, S.; Buxbaum, J.D. Genetics and genomics of autism spectrum disorder: Embracing complexity. Hum. Mol. Genet. 2015, 24, R24–R31. [Google Scholar] [CrossRef]
  20. 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] [PubMed] [Green Version]
  21. Sahin, M.; Sur, M. Genes, circuits, and precision therapies for autism and related neurodevelopmental disorders. Science 2015, 350. [Google Scholar] [CrossRef] [Green Version]
  22. Cassidy, S.B.; Allanson, J.E. Management of Genetic Syndromes; John Wiley & Sons: Hoboken, NJ, USA, 2010. [Google Scholar]
  23. Rylaarsdam, L.; Guemez-Gamboa, A. Genetic Causes and Modifiers of Autism Spectrum Disorder. Front. Cell Neurosci. 2019, 13, 385. [Google Scholar] [CrossRef] [PubMed]
  24. Salcedo-Arellano, M.J.; Dufour, B.; McLennan, Y.; Martinez-Cerdeno, V.; Hagerman, R. Fragile X syndrome and associated disorders: Clinical aspects and pathology. Neurobiol. Dis. 2020, 136, 104740. [Google Scholar] [CrossRef]
  25. Faundez, V.; Wynne, M.; Crocker, A.; Tarquinio, D. Molecular Systems Biology of Neurodevelopmental Disorders, Rett Syndrome as an Archetype. Front. Integr. Neurosci. 2019, 13, 30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Bernardet, M.; Crusio, W.E. Fmr1 KO mice as a possible model of autistic features. Sci. World J. 2006, 6, 1164–1176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Mendelsohn, N.J.; Schaefer, G.B. Genetic evaluation of autism. Semin. Pediatr. Neurol. 2008, 15, 27–31. [Google Scholar] [CrossRef]
  28. Schaefer, G.B.; Mendelsohn, N.J. Genetics evaluation for the etiologic diagnosis of autism spectrum disorders. Gen. Med. 2008, 10, 4–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Kaufmann, W.E.; Kidd, S.A.; Andrews, H.F.; Budimirovic, D.B.; Esler, A.; Haas-Givler, B.; Stackhouse, T.; Riley, C.; Peacock, G.; Sherman, S.L.; et al. Autism spectrum disorder in fragile X syndrome: Cooccurring conditions and current treatment. Pediatrics 2017, 139, S194–S206. [Google Scholar] [CrossRef] [Green Version]
  30. Talisa, V.B.; Boyle, L.; Crafa, D.; Kaufmann, W.E. Autism and anxiety in males with fragile X syndrome: An exploratory analysis of neurobehavioral profiles from a parent survey. Am. J. Med. Genet. A. 2014, 164A, 1198–1203. [Google Scholar] [CrossRef] [PubMed]
  31. Blomquist, H.K.S.; Bohman, M.; Edvinsson, S.O.; Gillberg, C.; Gustavson, K.H.; Holmgren, G.; Wahlström, J. Frequency of the fragile X syndrome in infantile autism. Clin. Genet. 1985, 27, 113–117. [Google Scholar] [CrossRef]
  32. Murray, A.; Youings, S.; Dennis, N.; Latsky, L.; Linehan, P.; McKechnie, N.; Macpherson, J.; Pound, M.; Jacobs, P. Population screening at the FRAXA and FRAXE loci: Molecular analyses of boys with learning difficulties and their mothers. Hum. Mol. Genet. 1996, 5, 727–735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Crawford, D.C.; Meadows, K.L.; Newman, J.L.; Taft, L.F.; Scott, E.; Leslie, M.; Shubek, L.; Holmgreen, P.; Yeargin-Allsopp, M.; Boyle, C. Prevalence of the fragile X syndrome in African-Americans. Am. J. Med. Genet. 2002, 110, 226–233. [Google Scholar] [CrossRef]
  34. Brown, M.R.; Kronengold, J.; Gazula, V.R.; Chen, Y.; Strumbos, J.G.; Sigworth, F.J.; Navaratnam, D.; Kaczmarek, L.K. Fragile X mental retardation protein controls gating of the sodium-activated potassium channel Slack. Nat. Neurosci. 2010, 13, 819–821. [Google Scholar] [CrossRef] [Green Version]
  35. Deng, P.Y.; Rotman, Z.; Blundon, J.A.; Cho, Y.; Cui, J.; Cavalli, V.; Zakharenko, S.S.; Klyachko, V.A. FMRP regulates neurotransmitter release and synaptic information transmission by modulating action potential duration via BK channels. Neuron 2013, 77, 696–711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Ferron, L.; Nieto-Rostro, M.; Cassidy, J.S.; Dolphin, A.C. Fragile X mental retardation protein controls synaptic vesicle exocytosis by modulating N-type calcium channel density. Nat. Commun. 2014, 5, 3628. [Google Scholar] [CrossRef] [PubMed]
  37. Brown, V.; Jin, P.; Ceman, S.; Darnell, J.C.; O’Donnell, W.T.; Tenenbaum, S.A.; Jin, X.; Feng, Y.; Wilkinson, K.D.; Keene, J.D.; et al. Microarray identification of FMRP-associated brain mRNAs and altered mRNA translational profiles in fragile X syndrome. Cell 2001, 107, 477–487. [Google Scholar] [CrossRef] [Green Version]
  38. Sherman, S.; Jacobs, P.; Morton, N.; Froster-Iskenius, U.; Howard-Peebles, P.; Nielsen, K.; Partington, M.; Sutherland, G.; Turner, G.; Watson, M. Further segregation analysis of the fragile X syndrome with special reference to transmitting males. Hum. Genet. 1985, 69, 289–299. [Google Scholar] [CrossRef] [PubMed]
  39. Verkerk, A.J.; Pieretti, M.; Sutcliffe, J.S.; Fu, Y.-H.; Kuhl, D.P.; Pizzuti, A.; Reiner, O.; Richards, S.; Victoria, M.F.; Zhang, F.; et al. Identification of a gene (FMR-1) containing a CGG repeat coincident with a breakpoint cluster region exhibiting length variation in fragile X syndrome. Cell 1991, 65, 905–914. [Google Scholar] [CrossRef]
  40. Nolin, S.L.; Brown, W.T.; Glicksman, A.; Houck Jr, G.E.; Gargano, A.D.; Sullivan, A.; Biancalana, V.; Bröndum-Nielsen, K.; Hjalgrim, H.; Holinski-Feder, E. Expansion of the fragile X CGG repeat in females with premutation or intermediate alleles. Am. J. Hum. Genet. 2003, 72, 454–464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Bartholomay, K.L.; Lee, C.H.; Bruno, J.L.; Lightbody, A.A.; Reiss, A.L. Closing the gender gap in fragile X syndrome: Review of females with fragile X syndrome and preliminary research findings. Brain Sci. 2019, 9, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Price, D.K.; Zhang, F.; Ashley Jr, C.T.; Warren, S.T. The ChickenFMR1Gene Is Highly Conserved with a CCT 5′-Untranslated Repeat and Encodes an RNA-Binding Protein. Genomics 1996, 31, 3–12. [Google Scholar] [CrossRef] [Green Version]
  43. Bassell, G.J.; Warren, S.T. Fragile X syndrome: Loss of local mRNA regulation alters synaptic development and function. Neuron 2008, 60, 201–214. [Google Scholar] [CrossRef] [Green Version]
  44. Myrick, L.K.; Hashimoto, H.; Cheng, X.; Warren, S.T. Human FMRP contains an integral tandem Agenet (Tudor) and KH motif in the amino terminal domain. Hum. Mol. Genet. 2015, 24, 1733–1740. [Google Scholar] [CrossRef] [PubMed]
  45. Eberhart, D.E.; Malter, H.E.; Feng, Y.; Warren, S.T. The fragile X mental retardation protein is a ribonucleoprotein containing both nuclear localization and nuclear export signals. Hum. Mol. Genet. 1996, 5, 1083–1091. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Brackett, D.M.; Qing, F.; Amieux, P.S.; Sellers, D.L.; Horner, P.J.; Morris, D.R. FMR1 transcript isoforms: Association with polyribosomes; regional and developmental expression in mouse brain. PLoS ONE 2013, 8, e58296. [Google Scholar] [CrossRef] [Green Version]
  47. Zorio, D.A.; Jackson, C.M.; Liu, Y.; Rubel, E.W.; Wang, Y. Cellular distribution of the fragile X mental retardation protein in the mouse brain. J. Comp. Neurol. 2017, 525, 818–849. [Google Scholar] [CrossRef] [Green Version]
  48. Arsenault, J.; Gholizadeh, S.; Niibori, Y.; Pacey, L.K.; Halder, S.K.; Koxhioni, E.; Konno, A.; Hirai, H.; Hampson, D.R. FMRP Expression Levels in Mouse Central Nervous System Neurons Determine Behavioral Phenotype. Hum. Gene. Ther. 2016, 27, 982–996. [Google Scholar] [CrossRef] [Green Version]
  49. Gholizadeh, S.; Halder, S.K.; Hampson, D.R. Expression of fragile X mental retardation protein in neurons and glia of the developing and adult mouse brain. Brain Res. 2015, 1596, 22–30. [Google Scholar] [CrossRef]
  50. Razak, K.A.; Dominick, K.C.; Erickson, C.A. Developmental studies in fragile X syndrome. J. Neurodevelop. Dis. 2020, 12, 1–15. [Google Scholar] [CrossRef] [PubMed]
  51. Antar, L.N.; Afroz, R.; Dictenberg, J.B.; Carroll, R.C.; Bassell, G.J. Metabotropic glutamate receptor activation regulates fragile x mental retardation protein and FMR1 mRNA localization differentially in dendrites and at synapses. J. Neurosci. 2004, 24, 2648–2655. [Google Scholar] [CrossRef]
  52. Akins, M.R.; Berk-Rauch, H.E.; Kwan, K.Y.; Mitchell, M.E.; Shepard, K.A.; Korsak, L.I.; Stackpole, E.E.; Warner-Schmidt, J.L.; Sestan, N.; Cameron, H.A. Axonal ribosomes and mRNAs associate with fragile X granules in adult rodent and human brains. Hum. Molec. Genet. 2017, 26, 192–209. [Google Scholar] [CrossRef] [Green Version]
  53. Korsak, L.I.T.; Mitchell, M.E.; Shepard, K.A.; Akins, M.R. Regulation of neuronal gene expression by local axonal translation. Curr. Genet. Med. Rep. 2016, 4, 16–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. 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]
  55. Cruz-Martin, A.; Crespo, M.; Portera-Cailliau, C. Delayed stabilization of dendritic spines in fragile X mice. J. Neurosci. 2010, 30, 7793–7803. [Google Scholar] [CrossRef] [Green Version]
  56. He, C.X.; Portera-Cailliau, C. The trouble with spines in fragile X syndrome: Density, maturity and plasticity. Neuroscience 2013, 251, 120–128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Schutt, J.; Falley, K.; Richter, D.; Kreienkamp, H.J.; Kindler, S. Fragile X mental retardation protein regulates the levels of scaffold proteins and glutamate receptors in postsynaptic densities. J. Biol. Chem. 2009, 284, 25479–25487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Ashley, C.T.; Wilkinson, K.D.; Reines, D.; Warren, S.T. FMR1 protein: Conserved RNP family domains and selective RNA binding. Science 1993, 262, 563–566. [Google Scholar] [CrossRef] [Green Version]
  59. Siomi, H.; Siomi, M.C.; Nussbaum, R.L.; Dreyfuss, G. The protein product of the fragile X gene, FMR1, has characteristics of an RNA-binding protein. Cell 1993, 74, 291–298. [Google Scholar] [CrossRef]
  60. Schaeffer, C.; Bardoni, B.; Mandel, J.L.; Ehresmann, B.; Ehresmann, C.; Moine, H. The fragile X mental retardation protein binds specifically to its mRNA via a purine quartet motif. EMBO J. 2001, 20, 4803–4813. [Google Scholar] [CrossRef]
  61. Siomi, H.; Choi, M.; Siomi, M.C.; Nussbaum, R.L.; Dreyfuss, G. Essential role for KH domains in RNA binding: Impaired RNA binding by a mutation in the KH domain of FMR1 that causes fragile X syndrome. Cell 1994, 77, 33–39. [Google Scholar] [CrossRef]
  62. Ramos, A.; Hollingworth, D.; Adinolfi, S.; Castets, M.; Kelly, G.; Frenkiel, T.A.; Bardoni, B.; Pastore, A. The structure of the N-terminal domain of the fragile X mental retardation protein: A platform for protein-protein interaction. Structure 2006, 14, 21–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Zhang, Y.; Brown, M.R.; Hyland, C.; Chen, Y.; Kronengold, J.; Fleming, M.R.; Kohn, A.B.; Moroz, L.L.; Kaczmarek, L.K. Regulation of neuronal excitability by interaction of fragile X mental retardation protein with slack potassium channels. J. Neurosci. 2012, 32, 15318–15327. [Google Scholar] [CrossRef] [Green Version]
  64. Zhang, Y.; Bonnan, A.; Bony, G.; Ferezou, I.; Pietropaolo, S.; Ginger, M.; Sans, N.; Rossier, J.; Oostra, B.; LeMasson, G.; et al. Dendritic channelopathies contribute to neocortical and sensory hyperexcitability in Fmr1(-/y) mice. Nat. Neurosci. 2014, 17, 1701–1709. [Google Scholar] [CrossRef]
  65. Ferron, L.; Novazzi, C.G.; Pilch, K.S.; Moreno, C.; Ramgoolam, K.; Dolphin, A.C. FMRP regulates presynaptic localization of neuronal voltage gated calcium channels. Neurobiol. Dis. 2020, 138, 104779. [Google Scholar] [CrossRef] [PubMed]
  66. Castagnola, S.; Delhaye, S.; Folci, A.; Paquet, A.; Brau, F.; Duprat, F.; Jarjat, M.; Grossi, M.; Beal, M.; Martin, S.; et al. New Insights Into the Role of Cav2 Protein Family in Calcium Flux Deregulation in Fmr1-KO Neurons. Front. Mol. Neurosci. 2018, 11, 342. [Google Scholar] [CrossRef]
  67. Chen, L.; Yun, S.; Seto, J.; Liu, W.; Toth, M. The fragile X mental retardation protein binds and regulates a novel class of mRNAs containing U rich target sequences. Neuroscience 2003, 120, 1005–1017. [Google Scholar] [CrossRef]
  68. Kalmbach, B.E.; Johnston, D.; Brager, D.H. Cell-Type Specific Channelopathies in the Prefrontal Cortex of the fmr1-/y Mouse Model of Fragile X Syndrome. eNeuro 2015, 2. [Google Scholar] [CrossRef] [Green Version]
  69. Brager, D.H.; Akhavan, A.R.; Johnston, D. Impaired dendritic expression and plasticity of h-channels in the fmr1−/y mouse model of fragile X syndrome. Cell Rep. 2012, 1, 225–233. [Google Scholar] [CrossRef] [Green Version]
  70. Bagni, C.; Zukin, R.S. A Synaptic Perspective of Fragile X Syndrome and Autism Spectrum Disorders. Neuron 2019, 101, 1070–1088. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Baudouin, S.J.; Gaudias, J.; Gerharz, S.; Hatstatt, L.; Zhou, K.; Punnakkal, P.; Tanaka, K.F.; Spooren, W.; Hen, R.; De Zeeuw, C.I. Shared synaptic pathophysiology in syndromic and nonsyndromic rodent models of autism. Science 2012, 338, 128–132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Bhakar, A.L.; Dolen, G.; Bear, M.F. The pathophysiology of fragile X (and what it teaches us about synapses). Annu. Rev. Neurosci. 2012, 35, 417–443. [Google Scholar] [CrossRef] [Green Version]
  73. Chen, J.; Yu, S.; Fu, Y.; Li, X. Synaptic proteins and receptors defects in autism spectrum disorders. Front. Cell. Neurosci. 2014, 8, 276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Luo, J.; Norris, R.H.; Gordon, S.L.; Nithianantharajah, J. Neurodevelopmental synaptopathies: Insights from behaviour in rodent models of synapse gene mutations. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2018, 84, 424–439. [Google Scholar] [CrossRef]
  75. 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] [Green Version]
  76. 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]
  77. Abe, T.; Sugihara, H.; Nawa, H.; Shigemoto, R.; Mizuno, N.; Nakanishi, S. Molecular characterization of a novel metabotropic glutamate receptor mGluR5 coupled to inositol phosphate/Ca2+ signal transduction. J. Biol. Chem. 1992, 267, 13361–13368. [Google Scholar] [CrossRef]
  78. Huber, K.M.; Gallagher, S.M.; Warren, S.T.; Bear, M.F. Altered synaptic plasticity in a mouse model of fragile X mental retardation. Proc. Natl. Acad. Sci. USA 2002, 99, 7746–7750. [Google Scholar] [CrossRef] [Green Version]
  79. 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. Adv. Integr. Anat. Evol. Biol. 2011, 294, 1635–1645. [Google Scholar] [CrossRef] [Green Version]
  80. Fatemi, S.H.; Folsom, T.D. GABA receptor subunit distribution and FMRP-mGluR5 signaling abnormalities in the cerebellum of subjects with schizophrenia, mood disorders, and autism. Schizophr. Res. 2015, 167, 42–56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Kelleher, R.J., 3rd; Geigenmuller, 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]
  82. Foldy, C.; Malenka, R.C.; Sudhof, T.C. Autism-associated neuroligin-3 mutations commonly disrupt tonic endocannabinoid signaling. Neuron 2013, 78, 498–509. [Google Scholar] [CrossRef] [Green Version]
  83. Sledziowska, M.; Galloway, J.; Baudouin, S.J. Evidence for a Contribution of the Nlgn3/Cyfip1/Fmr1 Pathway in the Pathophysiology of Autism Spectrum Disorders. Neuroscience 2019, 445, 31–41. [Google Scholar] [CrossRef] [PubMed]
  84. Guo, W.; Molinaro, G.; Collins, K.A.; Hays, S.A.; Paylor, R.; Worley, P.F.; Szumlinski, K.K.; Huber, K.M. Selective Disruption of Metabotropic Glutamate Receptor 5-Homer Interactions Mimics Phenotypes of Fragile X Syndrome in Mice. J. Neurosci. 2016, 36, 2131–2147. [Google Scholar] [CrossRef] [Green Version]
  85. Jung, K.M.; Sepers, M.; Henstridge, C.M.; Lassalle, O.; Neuhofer, D.; Martin, H.; Ginger, M.; Frick, A.; DiPatrizio, N.V.; Mackie, K.; et al. Uncoupling of the endocannabinoid signalling complex in a mouse model of fragile X syndrome. Nat. Commun. 2012, 3, 1080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Aloisi, E.; Le Corf, K.; Dupuis, J.; Zhang, P.; Ginger, M.; Labrousse, V.; Spatuzza, M.; Georg Haberl, M.; Costa, L.; Shigemoto, R.; et al. Altered surface mGluR5 dynamics provoke synaptic NMDAR dysfunction and cognitive defects in Fmr1 knockout mice. Nat. Commun. 2017, 8, 1103. [Google Scholar] [CrossRef] [Green Version]
  87. Tang, A.H.; Alger, B.E. Homer protein-metabotropic glutamate receptor binding regulates endocannabinoid signaling and affects hyperexcitability in a mouse model of fragile X syndrome. J. Neurosci. 2015, 35, 3938–3945. [Google Scholar] [CrossRef] [Green Version]
  88. Yang, M.; Bozdagi, O.; Scattoni, M.L.; Wohr, M.; Roullet, F.I.; Katz, A.M.; Abrams, D.N.; Kalikhman, D.; Simon, H.; Woldeyohannes, L.; et al. Reduced excitatory neurotransmission and mild autism-relevant phenotypes in adolescent Shank3 null mutant mice. J. Neurosci. 2012, 32, 6525–6541. [Google Scholar] [CrossRef] [PubMed]
  89. Krueger, D.D.; Brose, N. Evidence for a common endocannabinoid-related pathomechanism in autism spectrum disorders. Neuron 2013, 78, 408–410. [Google Scholar] [CrossRef] [Green Version]
  90. Skafidas, E.; Testa, R.; Zantomio, D.; Chana, G.; Everall, I.P.; Pantelis, C. Predicting the diagnosis of autism spectrum disorder using gene pathway analysis. Mol. Psychiatry 2014, 19, 504–510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Jamain, S.; Quach, H.; Betancur, C.; Rastam, M.; Colineaux, C.; Gillberg, I.C.; Soderstrom, H.; Giros, B.; Leboyer, M.; Gillberg, C.; et al. Mutations of the X-linked genes encoding neuroligins NLGN3 and NLGN4 are associated with autism. Nat. Genet. 2003, 34, 27–29. [Google Scholar] [CrossRef] [Green Version]
  92. Durand, C.M.; Betancur, C.; Boeckers, T.M.; Bockmann, J.; Chaste, P.; Fauchereau, F.; Nygren, G.; Rastam, M.; Gillberg, I.C.; Anckarsater, H.; et al. Mutations in the gene encoding the synaptic scaffolding protein SHANK3 are associated with autism spectrum disorders. Nat. Genet. 2007, 39, 25–27. [Google Scholar] [CrossRef] [Green Version]
  93. Pinto, D.; Delaby, E.; Merico, D.; Barbosa, M.; Merikangas, A.; Klei, L.; Thiruvahindrapuram, B.; Xu, X.; Ziman, R.; Wang, Z.; et al. Convergence of genes and cellular pathways dysregulated in autism spectrum disorders. Am. J. Hum. Genet. 2014, 94, 677–694. [Google Scholar] [CrossRef] [Green Version]
  94. Mullin, A.P.; Gokhale, A.; Moreno-De-Luca, A.; Sanyal, S.; Waddington, J.L.; Faundez, V. Neurodevelopmental disorders: Mechanisms and boundary definitions from genomes, interactomes and proteomes. Transl. Psychiatry 2013, 3, e329. [Google Scholar] [CrossRef]
  95. Assaf, M.; Jagannathan, K.; Calhoun, V.D.; Miller, L.; Stevens, M.C.; Sahl, R.; O’Boyle, J.G.; Schultz, R.T.; Pearlson, G.D. Abnormal functional connectivity of default mode sub-networks in autism spectrum disorder patients. NeuroImage 2010, 53, 247–256. [Google Scholar] [CrossRef] [Green Version]
  96. Cardon, G.J.; Hepburn, S.; Rojas, D.C. Structural Covariance of Sensory Networks, the Cerebellum, and Amygdala in Autism Spectrum Disorder. Front. Neurol. 2017, 8, 615. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. 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] [Green Version]
  98. Twitchell, W.; Brown, S.; Mackie, K. Cannabinoids inhibit N- and P/Q-type calcium channels in cultured rat hippocampal neurons. J. Neurophysiol. 1997, 78, 43–50. [Google Scholar] [CrossRef] [PubMed]
  99. Hebert, B.; Pietropaolo, S.; Meme, S.; Laudier, B.; Laugeray, A.; Doisne, N.; Quartier, A.; Lefeuvre, S.; Got, L.; Cahard, D.; et al. Rescue of fragile X syndrome phenotypes in Fmr1 KO mice by a BKCa channel opener molecule. Orphanet. J. Rare Dis. 2014, 9, 124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Laumonnier, F.; Roger, S.; Guérin, P.; Molinari, F.; M’rad, R.; Cahard, D.; Belhadj, A.; Halayem, M.; Persico, A.M.; Elia, M. Association of a functional deficit of the BK Ca channel, a synaptic regulator of neuronal excitability, with autism and mental retardation. Am. J. Psychiatry 2006, 163, 1622–1629. [Google Scholar] [CrossRef]
  101. Maccarrone, M.; Rossi, S.; Bari, M.; De Chiara, V.; Rapino, C.; Musella, A.; Bernardi, G.; Bagni, C.; Centonze, D. Abnormal mGlu 5 receptor/endocannabinoid coupling in mice lacking FMRP and BC1 RNA. Neuropsychopharmacology 2010, 35, 1500–1509. [Google Scholar] [CrossRef]
  102. Straiker, A.; Min, K.T.; Mackie, K. Fmr1 deletion enhances and ultimately desensitizes CB(1) signaling in autaptic hippocampal neurons. Neurobiol. Dis. 2013, 56, 1–5. [Google Scholar] [CrossRef] [Green Version]
  103. Darnell, J.C.; Van Driesche, S.J.; Zhang, C.; Hung, K.Y.S.; Mele, A.; Fraser, C.E.; Stone, E.F.; Chen, C.; Fak, J.J.; Chi, S.W. FMRP stalls ribosomal translocation on mRNAs linked to synaptic function and autism. Cell 2011, 146, 247–261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Ferron, L. Fragile X mental retardation protein controls ion channel expression and activity. J. Physiol. 2016, 594, 5861–5867. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Zhang, L.; Alger, B.E. Enhanced endocannabinoid signaling elevates neuronal excitability in fragile X syndrome. J. Neurosci. 2010, 30, 5724–5729. [Google Scholar] [CrossRef] [PubMed]
  106. Wang, X.; Bey, A.L.; Katz, B.M.; Badea, A.; Kim, N.; David, L.K.; Duffney, L.J.; Kumar, S.; Mague, S.D.; Hulbert, S.W.; et al. Altered mGluR5-Homer scaffolds and corticostriatal connectivity in a Shank3 complete knockout model of autism. Nat. Commun. 2016, 7, 11459. [Google Scholar] [CrossRef] [Green Version]
  107. Wang, X.; McCoy, P.A.; Rodriguiz, R.M.; Pan, Y.; Je, H.S.; Roberts, A.C.; Kim, C.J.; Berrios, J.; Colvin, J.S.; Bousquet-Moore, D.; et al. Synaptic dysfunction and abnormal behaviors in mice lacking major isoforms of Shank3. Hum. Mol. Genet. 2011, 20, 3093–3108. [Google Scholar] [CrossRef] [Green Version]
  108. Fyke, W.; Alarcon, J.M.; Velinov, M.; Chadman, K.K. Pharmacological inhibition of the primary endocannabinoid producing enzyme, DGL-α, induces autism spectrum disorder-like and co-morbid phenotypes in adult C57BL/J mice. Autism Res. 2021, 14, 1375–1389. [Google Scholar] [CrossRef]
  109. Li, J.; You, Y.; Yue, W.; Jia, M.; Yu, H.; Lu, T.; Wu, Z.; Ruan, Y.; Wang, L.; Zhang, D. Genetic Evidence for Possible Involvement of the Calcium Channel Gene CACNA1A in Autism Pathogenesis in Chinese Han Population. PLoS ONE 2015, 10, e0142887. [Google Scholar] [CrossRef]
  110. Ruzzo, E.K.; Perez-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] [PubMed] [Green Version]
  111. Yoshino, H.; Miyamae, T.; Hansen, G.; Zambrowicz, B.; Flynn, M.; Pedicord, D.; Blat, Y.; Westphal, R.S.; Zaczek, R.; Lewis, D.A.; et al. Postsynaptic diacylglycerol lipase mediates retrograde endocannabinoid suppression of inhibition in mouse prefrontal cortex. J. Physiol. 2011, 589, 4857–4884. [Google Scholar] [CrossRef]
  112. Wei, D.; Dinh, D.; Lee, D.; Li, D.; Anguren, A.; Moreno-Sanz, G.; Gall, C.M.; Piomelli, D. Enhancement of Anandamide-Mediated Endocannabinoid Signaling Corrects Autism-Related Social Impairment. Cannabis Cannabinoid Res. 2016, 1, 81–89. [Google Scholar] [CrossRef]
  113. Fyke, W.; Alarcon, J.M.; Velinov, M.; Chadman, K.K. Pharmacological inhibition of BKCa channels induces a specific social deficit in adult C57BL6/J mice. Behav. Neurosci. 2021. [Google Scholar] [CrossRef]
  114. Matsuda, L.A.; Lolait, S.J.; Brownstein, M.J.; Young, A.C.; Bonner, T.I. Structure of a cannabinoid receptor and functional expression of the cloned cDNA. Nature 1990, 346, 561–564. [Google Scholar] [CrossRef]
  115. Munro, S.; Thomas, K.L.; Abu-Shaar, M. Molecular characterization of a peripheral receptor for cannabinoids. Nature 1993, 365, 61. [Google Scholar] [CrossRef]
  116. Devane, W.A.; Hanus, L.; Breuer, A.; Pertwee, R.G.; Stevenson, L.A.; Griffin, G.; Gibson, D.; Mandelbaum, A.; Etinger, A.; Mechoulam, R. Isolation and structure of a brain constituent that binds to the cannabinoid receptor. Science 1992, 258, 1946–1949. [Google Scholar] [CrossRef]
  117. Kano, M.; Ohno-Shosaku, T.; Hashimotodani, Y.; Uchigashima, M.; Watanabe, M. Endocannabinoid-mediated control of synaptic transmission. Physiol. Rev. 2009, 89, 309–380. [Google Scholar] [CrossRef] [PubMed]
  118. Núñez, E.; Benito, C.; Pazos, M.R.; Barbachano, A.; Fajardo, O.; González, S.; Tolón, R.M.; Romero, J. Cannabinoid CB2 receptors are expressed by perivascular microglial cells in the human brain: An immunohistochemical study. Synapse 2004, 53, 208–213. [Google Scholar] [CrossRef] [PubMed]
  119. Makriyannis, A.; Tian, X.; Guo, J. How lipophilic cannabinergic ligands reach their receptor sites. Prostaglandins Other Lipid Mediat. 2005, 77, 210–218. [Google Scholar] [CrossRef] [PubMed]
  120. Sugiura, T.; Kondo, S.; Sukagawa, A.; Nakane, S.; Shinoda, A.; Itoh, K.; Yamashita, A.; Waku, K. 2-Arachidonoylgylcerol: A possible endogenous cannabinoid receptor ligand in brain. Biochem. Biophys. Res. Commun. 1995, 215, 89–97. [Google Scholar] [CrossRef]
  121. Suhara, Y.; Takayama, H.; Nakane, S.; Miyashita, T.; Waku, K.; Sugiura, T. Synthesis and biological activities of 2-arachidonoylglycerol, an endogenous cannabinoid receptor ligand, and its metabolically stable ether-linked analogues. Chem. Pharm. Bull. 2000, 48, 903–907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Stella, N.; Schweitzer, P.; Piomelli, D. A second endogenous cannabinoid that modulates long-term potentiation. Nature 1997, 388, 773–778. [Google Scholar] [CrossRef] [Green Version]
  123. Jung, K.M.; Mangieri, R.; Stapleton, C.; Kim, J.; Fegley, D.; Wallace, M.; Mackie, K.; Piomelli, D. Stimulation of endocannabinoid formation in brain slice cultures through activation of group I metabotropic glutamate receptors. Mol. Pharmacol. 2005, 68, 1196–1202. [Google Scholar] [CrossRef] [Green Version]
  124. Bisogno, T.; Howell, F.; Williams, G.; Minassi, A.; Cascio, M.G.; Ligresti, A.; Matias, I.; Schiano-Moriello, A.; Paul, P.; Williams, E.J.; et al. Cloning of the first sn1-DAG lipases points to the spatial and temporal regulation of endocannabinoid signaling in the brain. J. Cell Biol. 2003, 163, 463–468. [Google Scholar] [CrossRef] [PubMed]
  125. Ohno-Shosaku, T.; Hashimotodani, Y.; Ano, M.; Takeda, S.; Tsubokawa, H.; Kano, M. Endocannabinoid signalling triggered by NMDA receptor-mediated calcium entry into rat hippocampal neurons. J. Physiol. 2007, 584, 407–418. [Google Scholar] [CrossRef]
  126. Zhang, L.; Wang, M.; Bisogno, T.; Di Marzo, V.; Alger, B.E. Endocannabinoids generated by Ca2+ or by metabotropic glutamate receptors appear to arise from different pools of diacylglycerol lipase. PLoS ONE 2011, 6, e16305. [Google Scholar] [CrossRef] [Green Version]
  127. Brenowitz, S.D.; Regehr, W.G. Calcium dependence of retrograde inhibition by endocannabinoids at synapses onto Purkinje cells. J. Neurosci. 2003, 23, 6373–6384. [Google Scholar] [CrossRef]
  128. Dinh, T.; Carpenter, D.; Leslie, F.; Freund, T.; Katona, I.; Sensi, S.; Kathuria, S.; Piomelli, D. Brain monoglyceride lipase participating in endocannabinoid inactivation. Proc. Natl. Acad. Sci. USA 2002, 99, 10819–10824. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Gulyas, A.I.; Cravatt, B.F.; Bracey, M.H.; Dinh, T.P.; Piomelli, D.; Boscia, F.; Freund, T.F. Segregation of two endocannabinoid-hydrolyzing enzymes into pre- and postsynaptic compartments in the rat hippocampus, cerebellum and amygdala. Eur. J. Neurosci. 2004, 20, 441–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Sugiura, T.; Kobayashi, Y.; Oka, S.; Waku, K. Biosynthesis and degradation of anandamide and 2-arachidonoylglycerol and their possible physiological significance. Prostaglandins Leukot. Essent. Fat. Acids (PLEFA) 2002, 66, 173–192. [Google Scholar] [CrossRef]
  131. Felder, C.C.; Briley, E.M.; Axelrod, J.; Simpson, J.T.; Mackie, K.; Devane, W.A. Anandamide, an endogenous cannabimimetic eicosanoid, binds to the cloned human cannabinoid receptor and stimulates receptor-mediated signal transduction. Proc. Natl. Acad. Sci. USA 1993, 90, 7656–7660. [Google Scholar] [CrossRef] [Green Version]
  132. Felder, C.C.; Joyce, K.E.; Briley, E.M.; Mansouri, J.; Mackie, K.; Blond, O.; Lai, Y.; Ma, A.L.; Mitchell, R.L. Comparison of the pharmacology and signal transduction of the human cannabinoid CB1 and CB2 receptors. Mol. Pharmacol. 1995, 48, 443–450. [Google Scholar]
  133. Bouaboula, M.; Poinot-Chazel, C.; Bourrie, B.; Canat, X.; Calandra, B.; Rinaldi-Carmona, M.; Le Fur, G.; Casellas, P. Activation of mitogen-activated protein kinases by stimulation of the central cannabinoid receptor CB1. Biochem. J. 1995, 312, 637–641. [Google Scholar] [CrossRef] [Green Version]
  134. Pitler, T.; Alger, B. Postsynaptic spike firing reduces synaptic GABAA responses in hippocampal pyramidal cells. J. Neurosci. 1992, 12, 4122–4132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Wilson, R.I.; Nicoll, R.A. Endogenous cannabinoids mediate retrograde signalling at hippocampal synapses. Nature 2001, 410, 588. [Google Scholar] [CrossRef] [PubMed]
  136. Maccarrone, M.; Rossi, S.; Bari, M.; De Chiara, V.; Fezza, F.; Musella, A.; Gasperi, V.; Prosperetti, C.; Bernardi, G.; Finazzi-Agro, A.; et al. Anandamide inhibits metabolism and physiological actions of 2-arachidonoylglycerol in the striatum. Nat. Neurosci. 2008, 11, 152–159. [Google Scholar] [CrossRef]
  137. Lauckner, J.E.; Jensen, J.B.; Chen, H.-Y.; Lu, H.-C.; Hille, B.; Mackie, K. GPR55 is a cannabinoid receptor that increases intracellular calcium and inhibits M current. Proc. Natl. Acad. Sci. USA 2008, 105, 2699–2704. [Google Scholar] [CrossRef] [Green Version]
  138. Hill, T.D.; Cascio, M.G.; Romano, B.; Duncan, M.; Pertwee, R.G.; Williams, C.M.; Whalley, B.J.; Hill, A.J. Cannabidivarin-rich cannabis extracts are anticonvulsant in mouse and rat via a CB1 receptor-independent mechanism. Br. J. Pharmacol. 2013, 170, 679–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Hill, A.; Mercier, M.; Hill, T.; Glyn, S.; Jones, N.; Yamasaki, Y.; Futamura, T.; Duncan, M.; Stott, C.; Stephens, G. Cannabidivarin is anticonvulsant in mouse and rat. Br. J. Pharmacol. 2012, 167, 1629–1642. [Google Scholar] [CrossRef]
  140. Berghuis, P.; Rajnicek, A.M.; Morozov, Y.M.; Ross, R.A.; Mulder, J.; Urban, G.M.; Monory, K.; Marsicano, G.; Matteoli, M.; Canty, A.; et al. Hardwiring the brain: Endocannabinoids shape neuronal connectivity. Science 2007, 316, 1212–1216. [Google Scholar] [CrossRef] [Green Version]
  141. Keimpema, E.; Alpar, A.; Howell, F.; Malenczyk, K.; Hobbs, C.; Hurd, Y.L.; Watanabe, M.; Sakimura, K.; Kano, M.; Doherty, P.; et al. Diacylglycerol lipase alpha manipulation reveals developmental roles for intercellular endocannabinoid signaling. Sci. Rep. 2013, 3, 2093. [Google Scholar] [CrossRef]
  142. Purcell, A.; Jeon, O.; Zimmerman, A.; 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]
  143. Siniscalco, D.; Sapone, A.; Giordano, C.; Cirillo, A.; de Magistris, L.; Rossi, F.; Fasano, A.; Bradstreet, J.J.; Maione, S.; Antonucci, N. Cannabinoid receptor type 2, but not type 1, is up-regulated in peripheral blood mononuclear cells of children affected by autistic disorders. J. Autism Dev. Disord. 2013, 43, 2686–2695. [Google Scholar] [CrossRef] [PubMed]
  144. Smith, D.R.; Stanley, C.M.; Foss, T.; Boles, R.G.; McKernan, K. Rare genetic variants in the endocannabinoid system genes CNR1 and DAGLA are associated with neurological phenotypes in humans. PLoS ONE 2017, 12, e0187926. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Chakrabarti, B.; Baron-Cohen, S. Variation in the human cannabinoid receptor CNR1 gene modulates gaze duration for happy faces. Mol. Autism 2011, 2, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Chakrabarti, B.; Kent, L.; Suckling, J.; Bullmore, E.; Baron-Cohen, S. Variations in the human cannabinoid receptor (CNR1) gene modulate striatal responses to happy faces. Eur. J. Neurosci. 2006, 23, 1944–1948. [Google Scholar] [CrossRef] [PubMed]
  147. Aran, A.; Eylon, M.; Harel, M.; Polianski, L.; Nemirovski, A.; Tepper, S.; Schnapp, A.; Cassuto, H.; Wattad, N.; Tam, J. Lower circulating endocannabinoid levels in children with autism spectrum disorder. Mol. Autism 2019, 10. [Google Scholar] [CrossRef]
  148. Karhson, D.S.; Krasinska, K.M.; Dallaire, J.A.; Libove, R.A.; Phillips, J.M.; Chien, A.S.; Garner, J.P.; Hardan, A.Y.; Parker, K.J. Plasma anandamide concentrations are lower in children with autism spectrum disorder. Mol. Autism 2018, 9, 18. [Google Scholar] [CrossRef] [PubMed]
  149. Moessner, R.; Marshall, C.R.; Sutcliffe, J.S.; Skaug, J.; Pinto, D.; Vincent, J.; Zwaigenbaum, L.; Fernandez, B.; Roberts, W.; Szatmari, P.; et al. Contribution of SHANK3 mutations to autism spectrum disorder. Am. J. Hum. Genet. 2007, 81, 1289–1297. [Google Scholar] [CrossRef] [Green Version]
  150. Quartier, A.; Courraud, J.; Thi Ha, T.; McGillivray, G.; Isidor, B.; Rose, K.; Drouot, N.; Savidan, M.A.; Feger, C.; Jagline, H. Novel mutations in NLGN3 causing autism spectrum disorder and cognitive impairment. Hum. Mutat. 2019, 40, 2021–2032. [Google Scholar] [CrossRef]
  151. 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] [Green Version]
  152. Tarabeux, J.; Kebir, O.; Gauthier, J.; Hamdan, F.; Xiong, L.; Piton, A.; Spiegelman, D.; Henrion, É.; Millet, B.; Fathalli, F.; 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] [Green Version]
  153. Uzunova, G.; Hollander, E.; Shepherd, J. The role of ionotropic glutamate receptors in childhood neurodevelopmental disorders: Autism spectrum disorders and fragile x syndrome. Curr. Neuropharmacol. 2014, 12, 71–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Lee, E.-J.; Choi, S.Y.; Kim, E. NMDA receptor dysfunction in autism spectrum disorders. Curr. Opin. Pharmacol. 2015, 20, 8–13. [Google Scholar] [CrossRef]
  155. Fatemi, S.H.; Folsom, T.D. Dysregulation of fragile X mental retardation protein and metabotropic glutamate receptor 5 in superior frontal cortex of individuals with autism: A postmortem brain study. Mol. Autism 2011, 2, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Fatemi, S.H.; Folsom, T.D.; Kneeland, R.E.; Yousefi, M.K.; Liesch, S.B.; Thuras, P.D. Impairment of fragile X mental retardation protein-metabotropic glutamate receptor 5 signaling and its downstream cognates ras-related C3 botulinum toxin substrate 1, amyloid beta A4 precursor protein, striatal-enriched protein tyrosine phosphatase, and homer 1, in autism: A postmortem study in cerebellar vermis and superior frontal cortex. Mol. Autism 2013, 4, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Kerr, D.M.; Downey, L.; Conboy, M.; Finn, D.P.; Roche, M. Alterations in the endocannabinoid system in the rat valproic acid model of autism. Behav. Brain Res. 2013, 249, 124–132. [Google Scholar] [CrossRef] [PubMed]
  158. Jung, K.M.; Astarita, G.; Zhu, C.; Wallace, M.; Mackie, K.; Piomelli, D. A key role for diacylglycerol lipase-alpha in metabotropic glutamate receptor-dependent endocannabinoid mobilization. Mol. Pharmacol. 2007, 72, 612–621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Ronesi, J.A.; Huber, K.M. Homer interactions are necessary for metabotropic glutamate receptor-induced long-term depression and translational activation. J. Neurosci. 2008, 28, 543–547. [Google Scholar] [CrossRef]
  160. Tu, J.C.; Xiao, B.; Naisbitt, S.; Yuan, J.P.; Petralia, R.S.; Brakeman, P.; Doan, A.; Aakalu, V.K.; Lanahan, A.A.; Sheng, M.; et al. Coupling of mGluR/Homer and PSD-95 complexes by the Shank family of postsynaptic density proteins. Neuron 1999, 23, 583–592. [Google Scholar] [CrossRef] [Green Version]
  161. Wang, W.; Cox, B.M.; Jia, Y.; Le, A.A.; Cox, C.D.; Jung, K.M.; Hou, B.; Piomelli, D.; Gall, C.M.; Lynch, G. Treating a novel plasticity defect rescues episodic memory in Fragile X model mice. Mol. Psychiatry 2017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Qin, M.; Zeidler, Z.; Moulton, K.; Krych, L.; Xia, Z.; Smith, C.B. Endocannabinoid-mediated improvement on a test of aversive memory in a mouse model of fragile X syndrome. Behav. Brain Res. 2015, 291, 164–171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Busquets-Garcia, A.; Maldonado, R.; Ozaita, A. New insights into the molecular pathophysiology of fragile X syndrome and therapeutic perspectives from the animal model. Int. J. Biochem. Cell Biol. 2014, 53, 121–126. [Google Scholar] [CrossRef] [Green Version]
  164. Gomis-Gonzalez, M.; Busquets-Garcia, A.; Matute, C.; Maldonado, R.; Mato, S.; Ozaita, A. Possible Therapeutic Doses of Cannabinoid Type 1 Receptor Antagonist Reverses Key Alterations in Fragile X Syndrome Mouse Model. Genes 2016, 7, 56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Shonesy, B.C.; Parrish, W.P.; Haddad, H.K.; Stephenson, J.R.; Baldi, R.; Bluett, R.J.; Marks, C.R.; Centanni, S.W.; Folkes, O.M.; Spiess, K.; et al. Role of Striatal Direct Pathway 2-Arachidonoylglycerol Signaling in Sociability and Repetitive Behavior. Biol. Psychiatry 2018, 84, 304–315. [Google Scholar] [CrossRef] [Green Version]
  166. Shonesy, B.C.; Bluett, R.J.; Ramikie, T.S.; Baldi, R.; Hermanson, D.J.; Kingsley, P.J.; Marnett, L.J.; Winder, D.G.; Colbran, R.J.; Patel, S. Genetic disruption of 2-arachidonoylglycerol synthesis reveals a key role for endocannabinoid signaling in anxiety modulation. Cell Rep. 2014, 9, 1644–1653. [Google Scholar] [CrossRef] [Green Version]
  167. Jenniches, I.; Ternes, S.; Albayram, O.; Otte, D.M.; Bach, K.; Bindila, L.; Michel, K.; Lutz, B.; Bilkei-Gorzo, A.; Zimmer, A. Anxiety, Stress, and Fear Response in Mice With Reduced Endocannabinoid Levels. Biol. Psychiatry 2016, 79, 858–868. [Google Scholar] [CrossRef] [Green Version]
  168. Folkes, O.M.; Baldi, R.; Kondev, V.; Marcus, D.J.; Hartley, N.D.; Turner, B.D.; Ayers, J.K.; Baechle, J.J.; Misra, M.P.; Altemus, M.; et al. An endocannabinoid-regulated basolateral amygdala-nucleus accumbens circuit modulates sociability. J. Clin. Investig. 2020, 130, 1728–1742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Mulder, J.; Aguado, T.; Keimpema, E.; Barabas, K.; Ballester Rosado, C.J.; Nguyen, L.; Monory, K.; Marsicano, G.; Di Marzo, V.; Hurd, Y.L.; et al. Endocannabinoid signaling controls pyramidal cell specification and long-range axon patterning. Proc. Natl. Acad. Sci. USA 2008, 105, 8760–8765. [Google Scholar] [CrossRef] [Green Version]
  170. Heng, L.; Beverley, J.A.; Steiner, H.; Tseng, K.Y. Differential developmental trajectories for CB1 cannabinoid receptor expression in limbic/associative and sensorimotor cortical areas. Synapse 2011, 65, 278–286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Oudin, M.J.; Hobbs, C.; Doherty, P. DAGL-dependent endocannabinoid signalling: Roles in axonal pathfinding, synaptic plasticity and adult neurogenesis. Eur. J. Neurosci. 2011, 34, 1634–1646. [Google Scholar] [CrossRef]
  172. Long, L.E.; Lind, J.; Webster, M.; Weickert, C.S. Developmental trajectory of the endocannabinoid system in human dorsolateral prefrontal cortex. BMC Neurosci. 2012, 13, 87. [Google Scholar] [CrossRef] [Green Version]
  173. Abbas Farishta, R.; Robert, C.; Turcot, O.; Thomas, S.; Vanni, M.P.; Bouchard, J.F.; Casanova, C. Impact of CB1 Receptor Deletion on Visual Responses and Organization of Primary Visual Cortex in Adult Mice. Invest Ophthalmol. Vis. Sci. 2015, 56, 7697–7707. [Google Scholar] [CrossRef] [Green Version]
  174. Hill, M.N.; Hillard, C.J.; McEwen, B.S. Alterations in corticolimbic dendritic morphology and emotional behavior in cannabinoid CB1 receptor-deficient mice parallel the effects of chronic stress. Cereb. Cortex 2011, 21, 2056–2064. [Google Scholar] [CrossRef] [Green Version]
  175. Keown, C.L.; Datko, M.C.; Chen, C.P.; Maximo, J.O.; Jahedi, A.; Muller, R.A. Network organization is globally atypical in autism: A graph theory study of intrinsic functional connectivity. Biol. Psychiatry Cogn. Neurosci. Neuroimaging 2017, 2, 66–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Keown, C.L.; Shih, P.; Nair, A.; Peterson, N.; Mulvey, M.E.; Muller, R.A. Local functional overconnectivity in posterior brain regions is associated with symptom severity in autism spectrum disorders. Cell Rep. 2013, 5, 567–572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Haller, J.; Bakos, N.; Szirmay, M.; Ledent, C.; Freund, T.F. The effects of genetic and pharmacological blockade of the CB1 cannabinoid receptor on anxiety. Eur. J. Neurosci. 2002, 16, 1395–1398. [Google Scholar] [CrossRef]
  178. Haller, J.; Varga, B.; Ledent, C.; Barna, I.; Freund, T.F. Context-dependent effects of CB1 cannabinoid gene disruption on anxiety-like and social behaviour in mice. Eur. J. Neurosci. 2004, 19, 1906–1912. [Google Scholar] [CrossRef]
  179. Litvin, Y.; Phan, A.; Hill, M.N.; Pfaff, D.W.; McEwen, B.S. CB1 receptor signaling regulates social anxiety and memory. Genes Brain Behav. 2013, 12, 479–489. [Google Scholar] [CrossRef] [PubMed]
  180. Terzian, A.L.; Micale, V.; Wotjak, C.T. Cannabinoid receptor type 1 receptors on GABAergic vs. glutamatergic neurons differentially gate sex-dependent social interest in mice. Eur. J. Neurosci. 2014, 40, 2293–2298. [Google Scholar] [CrossRef] [PubMed]
  181. Devinsky, O.; Nabbout, R.; Miller, I.; Laux, L.; Zolnowska, M.; Wright, S.; Roberts, C. Long-term cannabidiol treatment in patients with Dravet syndrome: An open-label extension trial. Epilepsia 2018. [Google Scholar] [CrossRef] [PubMed]
  182. Devinsky, O.; Patel, A.D.; Cross, J.H.; Villanueva, V.; Wirrell, E.C.; Privitera, M.; Greenwood, S.M.; Roberts, C.; Checketts, D.; VanLandingham, K.E.; et al. Effect of Cannabidiol on Drop Seizures in the Lennox-Gastaut Syndrome. N. Engl. J. Med. 2018, 378, 1888–1897. [Google Scholar] [CrossRef] [Green Version]
  183. Heussler, H.; Cohen, J.; Silove, N.; Tich, N.; Bonn-Miller, M.O.; Du, W.; O’Neill, C.; Sebree, T. A phase 1/2, open-label assessment of the safety, tolerability, and efficacy of transdermal cannabidiol (ZYN002) for the treatment of pediatric fragile X syndrome. J. Neurodev. Disord. 2019, 11, 16. [Google Scholar] [CrossRef] [PubMed]
  184. Tartaglia, N.; Bonn-Miller, M.; Hagerman, R. Treatment of Fragile X Syndrome with Cannabidiol: A Case Series Study and Brief Review of the Literature. Cannabis Cannabinoid Res. 2019, 4, 3–9. [Google Scholar] [CrossRef] [PubMed]
  185. Barchel, D.; Stolar, O.; De-Haan, T.; Ziv-Baran, T.; Saban, N.; Fuchs, D.O.; Koren, G.; Berkovitch, M. Oral Cannabidiol Use in Children With Autism Spectrum Disorder to Treat Related Symptoms and Co-morbidities. Front. Pharmacol. 2018, 9, 1521. [Google Scholar] [CrossRef] [PubMed]
  186. Poleg, S.; Golubchik, P.; Offen, D.; Weizman, A. Cannabidiol as a suggested candidate for treatment of autism spectrum disorder. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2019, 89, 90–96. [Google Scholar] [CrossRef]
  187. Pretzsch, C.M.; Freyberg, J.; Voinescu, B.; Lythgoe, D.; Horder, J.; Mendez, M.A.; Wichers, R.; Ajram, L.; Ivin, G.; Heasman, M.; et al. Effects of cannabidiol on brain excitation and inhibition systems; a randomised placebo-controlled single dose trial during magnetic resonance spectroscopy in adults with and without autism spectrum disorder. Neuropsychopharmacology 2019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Pretzsch, C.M.; Voinescu, B.; Mendez, M.A.; Wichers, R.; Ajram, L.; Ivin, G.; Heasman, M.; Williams, S.; Murphy, D.G.; Daly, E.; et al. The effect of cannabidiol (CBD) on low-frequency activity and functional connectivity in the brain of adults with and without autism spectrum disorder (ASD). J. Psychopharmacol. 2019, 33, 1141–1148. [Google Scholar] [CrossRef]
  189. Bar-Lev Schleider, L.; Mechoulam, R.; Saban, N.; Meiri, G.; Novack, V. Real life Experience of Medical Cannabis Treatment in Autism: Analysis of Safety and Efficacy. Sci. Rep. 2019, 9. [Google Scholar] [CrossRef] [PubMed]
  190. Zamberletti, E.; Gabaglio, M.; Parolaro, D. The Endocannabinoid System and Autism Spectrum Disorders: Insights from Animal Models. Int. J. Mol. Sci. 2017, 18, 1916. [Google Scholar] [CrossRef]
  191. Pretzsch, C.M.; Voinescu, B.; Lythgoe, D.; Horder, J.; Mendez, M.A.; Wichers, R.; Ajram, L.; Ivin, G.; Heasman, M.; Edden, R.A.E.; et al. Effects of cannabidivarin (CBDV) on brain excitation and inhibition systems in adults with and without Autism Spectrum Disorder (ASD): A single dose trial during magnetic resonance spectroscopy. Transl. Psychiatry 2019, 9, 313. [Google Scholar] [CrossRef] [Green Version]
  192. Vigli, D.; Cosentino, L.; Raggi, C.; Laviola, G.; Woolley-Roberts, M.; De Filippis, B. Chronic treatment with the phytocannabinoid Cannabidivarin (CBDV) rescues behavioural alterations and brain atrophy in a mouse model of Rett syndrome. Neuropharmacol. 2018, 140, 121–129. [Google Scholar] [CrossRef]
  193. Zamberletti, E.; Gabaglio, M.; Woolley-Roberts, M.; Bingham, S.; Rubino, T.; Parolaro, D. Cannabidivarin Treatment Ameliorates Autism-Like Behaviors and Restores Hippocampal Endocannabinoid System and Glia Alterations Induced by Prenatal Valproic Acid Exposure in Rats. Front. Cell. Neurosci. 2019, 13, 367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Morales, P.; Hurst, D.P.; Reggio, P.H. Molecular Targets of the Phytocannabinoids: A Complex Picture. Prog. Chem. Org. Nat. Prod. 2017, 103, 103–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Iannotti, F.A.; Hill, C.L.; Leo, A.; Alhusaini, A.; Soubrane, C.; Mazzarella, E.; Russo, E.; Whalley, B.J.; Di Marzo, V.; Stephens, G.J. Nonpsychotropic plant cannabinoids, cannabidivarin (CBDV) and cannabidiol (CBD), activate and desensitize transient receptor potential vanilloid 1 (TRPV1) channels in vitro: Potential for the treatment of neuronal hyperexcitability. ACS Chem. Neurosci. 2014, 5, 1131–1141. [Google Scholar] [CrossRef] [Green Version]
  196. Rock, E.M.; Sticht, M.A.; Duncan, M.; Stott, C.; Parker, L.A. Evaluation of the potential of the phytocannabinoids, cannabidivarin (CBDV) and Delta(9) -tetrahydrocannabivarin (THCV), to produce CB1 receptor inverse agonism symptoms of nausea in rats. Br. J. Pharmacol. 2013, 170, 671–678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Rosenthaler, S.; Pohn, B.; Kolmanz, C.; Huu, C.N.; Krewenka, C.; Huber, A.; Kranner, B.; Rausch, W.D.; Moldzio, R. Differences in receptor binding affinity of several phytocannabinoids do not explain their effects on neural cell cultures. Neurotoxicology Teratol. 2014, 46, 49–56. [Google Scholar] [CrossRef]
  198. De Petrocellis, L.; Ligresti, A.; Moriello, A.S.; Allarà, M.; Bisogno, T.; Petrosino, S.; Stott, C.G.; Di Marzo, V. Effects of cannabinoids and cannabinoid-enriched Cannabis extracts on TRP channels and endocannabinoid metabolic enzymes. Br. J. Pharmacol. 2011, 163, 1479–1494. [Google Scholar] [CrossRef] [Green Version]
  199. Atwood, B.K.; Wager-Miller, J.; Haskins, C.; Straiker, A.; Mackie, K. Functional selectivity in CB(2) cannabinoid receptor signaling and regulation: Implications for the therapeutic potential of CB(2) ligands. Mol. Pharmacol. 2012, 81, 250–263. [Google Scholar] [CrossRef] [Green Version]
  200. Tseng-Crank, J.; Foster, C.D.; Krause, J.D.; Mertz, R.; Godinot, N.; DiChiara, T.J.; Reinhart, P.H. Cloning, expression, and distribution of functionally distinct Ca2+-activated K+ channel isoforms from human brain. Neuron 1994, 13, 1315–1330. [Google Scholar] [CrossRef]
  201. Petrik, D.; Brenner, R. Regulation of STREX exon large conductance, calcium-activated potassium channels by the beta4 accessory subunit. Neuroscience 2007, 149, 789–803. [Google Scholar] [CrossRef] [Green Version]
  202. Weiger, T.M.; Holmqvist, M.H.; Levitan, I.B.; Clark, F.T.; Sprague, S.; Huang, W.J.; Glucksmann, M.A. A novel nervous system β subunit that downregulates human large conductance calcium-dependent potassium channels. J. Neurosci. 2000, 20, 3563–3570. [Google Scholar] [CrossRef] [PubMed]
  203. Hu, H.; Shao, L.R.; Chavoshy, S.; Gu, N.; Trieb, M.; Behrens, R.; Storm, J.F. Presynaptic Ca2+-activated K+ channels in glutamatergic hippocampal terminals and their role in spike repolarization and regulation of transmitter release. J. Neurosci. 2001, 21, 9585–9597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Wallner, M.; Meera, P.; Toro, L. Determinant for β-subunit regulation in high-conductance voltage-activated and Ca2+-sensitive K+ channels: An additional transmembrane region at the N terminus. Proc. Natl. Acad. Sci. USA 1996, 93, 14922–14927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Salkoff, L.; Butler, A.; Ferreira, G.; Santi, C.; Wei, A. High-conductance potassium channels of the SLO family. Nat. Rev. Neurosci. 2006, 7, 921–931. [Google Scholar] [CrossRef] [PubMed]
  206. Berkefeld, H.; Sailer, C.A.; Bildl, W.; Rohde, V.; Thumfart, J.O.; Eble, S.; Klugbauer, N.; Reisinger, E.; Bischofberger, J.; Oliver, D.; et al. BKCa-Cav channel complexes mediate rapid and localized Ca2+-activated K+ signaling. Science 2006, 314, 615–620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Berkefeld, H.; Fakler, B. Ligand-gating by Ca2+ is rate limiting for physiological operation of BK(Ca) channels. J. Neurosci. 2013, 33, 7358–7367. [Google Scholar] [CrossRef] [Green Version]
  208. Kshatri, A.; Cerrada, A.; Gimeno, R.; Bartolomé-Martín, D.; Rojas, P.; Giraldez, T. Differential regulation of BK channels by fragile X mental retardation protein. J. Gen. Physiol. 2020, 152, e201912502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Alarcon, M.; Cantor, R.M.; Liu, J.; Gilliam, T.C.; Geschwind, D.H.; Autism Genetic Research Exchange, C. Evidence for a language quantitative trait locus on chromosome 7q in multiplex autism families. Am. J. Hum. Genet. 2002, 70, 60–71. [Google Scholar] [CrossRef] [Green Version]
  210. International Molecular Genetic Study of Autism. A genomewide screen for autism: Strong evidence for linkage to chromosomes 2q, 7q, and 16p. Am. J. Hum. Genet. 2001, 69, 570–581. [Google Scholar] [CrossRef] [Green Version]
  211. Myrick, L.K.; Deng, P.Y.; Hashimoto, H.; Oh, Y.M.; Cho, Y.; Poidevin, M.J.; Suhl, J.A.; Visootsak, J.; Cavalli, V.; Jin, P.; et al. Independent role for presynaptic FMRP revealed by an FMR1 missense mutation associated with intellectual disability and seizures. Proc. Natl. Acad. Sci. USA 2015, 112, 949–956. [Google Scholar] [CrossRef] [Green Version]
  212. Deng, P.Y.; Klyachko, V.A. Genetic upregulation of BK channel activity normalizes multiple synaptic and circuit defects in a mouse model of fragile X syndrome. J. Physiol. 2016, 594, 83–97. [Google Scholar] [CrossRef]
  213. Typlt, M.; Mirkowski, M.; Azzopardi, E.; Ruettiger, L.; Ruth, P.; Schmid, S. Mice with deficient BK channel function show impaired prepulse inhibition and spatial learning, but normal working and spatial reference memory. PLoS ONE 2013, 8, e081270. [Google Scholar] [CrossRef] [Green Version]
  214. Bondarenko, A.I.; Panasiuk, O.; Drachuk, K.; Montecucco, F.; Brandt, K.J.; Mach, F. The quest for endothelial atypical cannabinoid receptor: BKCa channels act as cellular sensors for cannabinoids in in vitro and in situ endothelial cells. Vascul. Pharmacol. 2018. [Google Scholar] [CrossRef]
  215. Sade, H.; Muraki, K.; Ohya, S.; Hatano, N.; Imaizumi, Y. Activation of large-conductance, Ca2+-activated K+ channels by cannabinoids. Am. J. Physiol. Cell Physiol. 2006, 290, C77–C86. [Google Scholar] [CrossRef]
  216. Jensen, B.S. BMS-204352: A potassium channel opener developed for the treatment of stroke. CNS Drug Rev. 2002, 8, 353–360. [Google Scholar] [CrossRef]
  217. Carreno-Munoz, M.I.; Martins, F.; Medrano, M.C.; Aloisi, E.; Pietropaolo, S.; Dechaud, C.; Subashi, E.; Bony, G.; Ginger, M.; Moujahid, A.; et al. Potential Involvement of Impaired BKCa Channel Function in Sensory Defensiveness and Some Behavioral Disturbances Induced by Unfamiliar Environment in a Mouse Model of Fragile X Syndrome. Neuropsychopharmacology 2018, 43, 492–502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Catterall, W.A.; Few, A.P. Calcium channel regulation and presynaptic plasticity. Neuron 2008, 59, 882–901. [Google Scholar] [CrossRef] [Green Version]
  219. Turner, T.J.; Adams, M.E.; Dunlap, K. Multiple Ca2+ channel types coexist to regulate synaptosomal neurotransmitter release. Proc. Natl. Acad. Sci. USA 1993, 90, 9518–9522. [Google Scholar] [CrossRef] [Green Version]
  220. Xu, J.; Long, L.; Wang, J.; Tang, Y.; Hu, H.; Soong, T.; Tang, F. Nuclear localization of Cav2. 2 and its distribution in the mouse central nervous system, and changes in the hippocampus during and after pilocarpine-induced status epilepticus. Neuropathol. Appl. Neurobiol. 2010, 36, 71–85. [Google Scholar] [CrossRef] [PubMed]
  221. Heyes, S.; Pratt, W.S.; Rees, E.; Dahimene, S.; Ferron, L.; Owen, M.J.; Dolphin, A.C. Genetic disruption of voltage-gated calcium channels in psychiatric and neurological disorders. Prog. Neurobiol. 2015, 134, 36–54. [Google Scholar] [CrossRef] [Green Version]
  222. Yatsenko, S.A.; Hixson, P.; Roney, E.K.; Scott, D.A.; Schaaf, C.P.; Ng, Y.-t.; Palmer, R.; Fisher, R.B.; Patel, A.; Cheung, S.W. Human subtelomeric copy number gains suggest a DNA replication mechanism for formation: Beyond breakage–fusion–bridge for telomere stabilization. Hum. Genet. 2012, 131, 1895–1910. [Google Scholar] [CrossRef] [Green Version]
  223. Palmieri, L.; Papaleo, V.; Porcelli, V.; Scarcia, P.; Gaita, L.; Sacco, R.; Hager, J.; Rousseau, F.; Curatolo, P.; Manzi, B. Altered calcium homeostasis in autism-spectrum disorders: Evidence from biochemical and genetic studies of the mitochondrial aspartate/glutamate carrier AGC1. Mol. Psychiatry 2010, 15, 38–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Gaffuri, A.L.; Ladarre, D.; Lenkei, Z. Type-1 cannabinoid receptor signaling in neuronal development. Pharmacology 2012, 90, 19–39. [Google Scholar] [CrossRef] [PubMed]
  225. Karhson, D.S.; Hardan, A.Y.; Parker, K.J. Endocannabinoid signaling in social functioning: An RDoC perspective. Transl. Psychiatry 2016, 6, e905. [Google Scholar] [CrossRef] [Green Version]
  226. Zou, M.; Li, D.; Li, L.; Wu, L.; Sun, C. Role of the endocannabinoid system in neurological disorders. Int. J. Dev. Neurosci. 2019, 76, 95–102. [Google Scholar] [CrossRef]
Figure 1. The FMR1 gene and Fragile X pathology. CGG repeats (yellow) in the promoter region. <55 repeats are typical. Repeat expansion resulting in the premutation (55–200) is found in 1/130–250 females and 1/260–800 males. The premutation expansion increases mRNA transcription and is associated with Fragile X primary ovarian insufficiency (FXPOI), Fragile X-associated tremor and ataxia syndrome (FXTAS), and Fragile X-associated neuropsychiatric disorder (FXAND). Repeats greater than 200 results in the methylation of the promoter region and gene silencing.
Figure 1. The FMR1 gene and Fragile X pathology. CGG repeats (yellow) in the promoter region. <55 repeats are typical. Repeat expansion resulting in the premutation (55–200) is found in 1/130–250 females and 1/260–800 males. The premutation expansion increases mRNA transcription and is associated with Fragile X primary ovarian insufficiency (FXPOI), Fragile X-associated tremor and ataxia syndrome (FXTAS), and Fragile X-associated neuropsychiatric disorder (FXAND). Repeats greater than 200 results in the methylation of the promoter region and gene silencing.
Genes 12 01218 g001
Figure 2. FMRP regulation of presynaptic activity. FMRP contributes to presynaptic regulation (A) by (a) stimulating the BKCa channel to inhibit CaV channels, (b) directly inhibiting CaV channels, or (c) controlling the translation and localization of the eCB producing enzyme diacylglycerol lipase α (DGL-α) in the post-synaptic density. DGL-α exists in a complex (synaptosome) with the scaffolding protein Homer1a and mGluR5. DGL-α production of the CB1 ligand 2-AG occurs due to mGluR5 activity (Ca2+ independent) or NMDA activity (Ca2+ dependent). CB1 responds to 2-AG stimulation by inhibiting P/Q and N-type CaV channels. The absence of FMRP due to the FMR1 mutation (B) results in a loss of appropriate presynaptic CaV channels regulation by (d,e) BKCa channels and (f) direct FMRP interactions; (g) absence of post-synaptic FMRP results in delocalized DGL-α and 2-AG production. Mutations in (h) the β4 regulatory unit of BKCa channels, (i) CB1, (j) Homer1a, (k) Shank3, and (l) NMDA channels have associations with syndromic and non-syndromic ASD. Each of these defects causes increased Ca2+ entry and neurotransmitter release (computational dysfunction). [108].
Figure 2. FMRP regulation of presynaptic activity. FMRP contributes to presynaptic regulation (A) by (a) stimulating the BKCa channel to inhibit CaV channels, (b) directly inhibiting CaV channels, or (c) controlling the translation and localization of the eCB producing enzyme diacylglycerol lipase α (DGL-α) in the post-synaptic density. DGL-α exists in a complex (synaptosome) with the scaffolding protein Homer1a and mGluR5. DGL-α production of the CB1 ligand 2-AG occurs due to mGluR5 activity (Ca2+ independent) or NMDA activity (Ca2+ dependent). CB1 responds to 2-AG stimulation by inhibiting P/Q and N-type CaV channels. The absence of FMRP due to the FMR1 mutation (B) results in a loss of appropriate presynaptic CaV channels regulation by (d,e) BKCa channels and (f) direct FMRP interactions; (g) absence of post-synaptic FMRP results in delocalized DGL-α and 2-AG production. Mutations in (h) the β4 regulatory unit of BKCa channels, (i) CB1, (j) Homer1a, (k) Shank3, and (l) NMDA channels have associations with syndromic and non-syndromic ASD. Each of these defects causes increased Ca2+ entry and neurotransmitter release (computational dysfunction). [108].
Genes 12 01218 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fyke, W.; Velinov, M. FMR1 and Autism, an Intriguing Connection Revisited. Genes 2021, 12, 1218. https://doi.org/10.3390/genes12081218

AMA Style

Fyke W, Velinov M. FMR1 and Autism, an Intriguing Connection Revisited. Genes. 2021; 12(8):1218. https://doi.org/10.3390/genes12081218

Chicago/Turabian Style

Fyke, William, and Milen Velinov. 2021. "FMR1 and Autism, an Intriguing Connection Revisited" Genes 12, no. 8: 1218. https://doi.org/10.3390/genes12081218

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