Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Jul 8.
Published in final edited form as: Annu Rev Neurosci. 2020 Jul 8;43:509–533. doi: 10.1146/annurev-neuro-100119-024851

The Genetic Control of Stoichiometry Underlying Autism

Robert B Darnell 1
PMCID: PMC7593886  NIHMSID: NIHMS1633604  PMID: 32640929

Abstract

Autism is a common and complex neurologic disorder whose scientific underpinnings have begun to be established in the past decade. The essence of this breakthrough has been a focus on families, where genetic analyses are strongest, versus large-scale, case-control studies. Autism genetics has progressed in parallel with technology, from analyses of copy number variation to whole-exome sequencing (WES) and whole-genome sequencing (WGS). Gene mutations causing complete loss of function account for perhaps one-third of cases, largely detected through WES. This limitation has increased interest in understanding the regulatory variants of genes that contribute in more subtle ways to the disorder. Strategies combining biochemical analysis of gene regulation, WGS analysis of the noncoding genome, and machine learning have begun to succeed. The emerging picture is that careful control of the amounts of transcription, mRNA, and proteins made by key brain genes—stoichiometry—plays a critical role in defining the clinical features of autism.

Keywords: stoichiometry, RNA, translation, FMRP, synaptic plasticity

INTRODUCTION

Autism refers to a variety of clinical features observed in children in the first three years of life who develop difficulties with direct social interactions along with a variety of manifestations, including language dysfunction, repetitive behaviors, and restricted interests and activities, together termed autism spectrum disorder (ASD). It is therefore not surprising that, in parallel to the many clinical manifestations of those given this diagnosis, there may be many predisposing factors.

Similar phenomena abound in neurology. For example, epilepsy is a general diagnostic term with a multitude of predisposing conditions, and genetic understanding of epilepsy has identified mutations in a variety of channel proteins involved in the balance between excitation and inhibition, as well as in proteins involved in many different aspects of cell metabolism, and these genetic insights are driving rational therapeutics (Dhindsa & Goldstein 2015). Similarly, the set of dementias includes many disorders with many etiologies. Rapid-onset early dementia in Creutzfeldt-Jakob disease has an entirely different pathophysiology than the more gradual onset of dementia associated with APOE4 homozygosity. Common to all such general diagnostic categories are several features: They are clinical definitions with key common clinical manifestations that group them together—intellectual disability presenting in youth, seizures, or loss of cognitive function with age—even as clinicians recognize that there exist distinct and sometimes overlapping subsets of individuals within each category. It is important that basic scientists appreciate that such distinctions may underlie a plethora of different etiologies that manifest clinically as the same disease.

This complexity is worth emphasizing for several reasons. First, strategies that do not recognize the inherent diversity underlying categorical clinical terms such as autism have struggled to identify pathophysiologic causes for disease in individual cases. Efforts to take large cohorts of individuals with a general diagnosis of autism and extrapolate a common set of etiologic factors for what in reality are different physiologic disorders can lead to confusion. Genome-wide association study (GWAS)-focused autism studies struggled to identify common genetic features in what we now recognize as a multitude of different underlying genetic disorders unified by a common clinical diagnostic term.

Autism does have a genetic basis, with concordance rates of at least 90% in monozygotic twins with ASD, compared to 10% in dizygotic twins (Tick et al. 2016). Large-scale population-based epidemiologic studies, analyzing over 22,000 ASD individuals in five countries, estimated overall heritability of ASD as at least 80% (Bai et al. 2019). This review presents the current understanding of the molecular genetics of autism, framing an overarching model in which a wide variety of critical regulatory, synaptic, and other neuronal proteins underlie the common phenotypes of ASD, and explores the idea that one theme tying these disparate proteins together is tight stoichiometric control of their expression.

BREAKTHROUGHS IN AUTISM GENOMICS

Early studies to assess genetic causes of autism attempted to leverage the paradigm used in analysis of single nucleotide polymorphisms (SNPs) in case-control studies that used GWAS (Klein et al. 2005). This strategy was based on the recognition that common SNP variants reside in blocks and that linkage disequilibrium of such blocks can lead to insight into disease heritability. Such studies were more readily accessible than family studies in which genetics are much more tightly controlled but which are more difficult to assemble at scale.

GWAS searching for common but pleomorphic disorders such as schizophrenia (Pardiñas et al. 2018) have identified gene-associated loci through analysis of common SNP variants (for example, those present at ≥1% in the population), suggesting the possibility that this set of disorders results from the collection of many such variants in sporadic cases of this common disease. Nonetheless, it has been difficult to validate and make use of such genetic findings clinically, and this same problem has been true for studies of autism.

One approach to pursuing such studies has been to build larger sets of variants by increasing sample size and by looking at less common (rare) variants. Statistical power analysis argued that studying 25,000 cases of autism and comparing, via GWAS analysis, 100,000 controls would yield specific risk alleles (Pardiñas et al. 2018, Zuk et al. 2014). While entirely logical, this approach was confounded by the heterogeneity in the clinical features of autism as well as requisite a priori assumptions about the impact genetics would have on fecundity and selection coefficients. Autism is a spectrum of disorders and can range from those who are severely disabled to those who can function well, sometimes exceptionally, in the workplace and family. Consistent with this complexity, linkage studies did not unambiguously identify specific genes that caused ASD (Autism Spectr. Disord. Work. Group Psychiatr. Genom. Consort. 2017, Weiss et al. 2009), although larger GWAS (as applied to the recent collection of over 18,000 ASD cases and 28,000 controls) continue to be of interest (Grove et al. 2019).

A different approach to the study of autism came from bypassing the challenges and opportunities of GWAS and instead focusing on tightly genetically controlled sets of individual families. These studies began with analysis of copy number variants (CNVs) in families (Sebat et al. 2007) and then were refined to studies of whole-exome sequencing (WES) of trios or quads—genetic analysis of four individual family members, one affected and three not. Such analyses of an affected child (proband), a sibling, and their parents provided a tight-knit group of genetically related individuals from whom scientists could compare de novo genetic variants in probands.

It is worth noting that getting a firm grip on specific genes implicated in ASD, through studies of either small numbers of familial ASD (necessarily small because the disease itself can have an effect on fecundity) or larger numbers of de novo (simplex) ASD cases, has offered the potential to have important feed-forward effects in developing a comprehensive view of autism. By analogy, early linkage studies of the genetics of breast cancer failed to identify risk genes due to the complexity of the genetics and environmental factors that impact their manifestations. A foothold on the problem was accomplished by Mary-Claire King and colleagues (Hall et al. 1990) by focusing on families with a high incidence of breast cancer, and even there such studies required further refinement to look specifically at those with early-onset disease. This intuition, following basic tenets of a geneticist, allowed the identification of the BRCA1 gene and fed forward to the current recognition that a series of well over a dozen genetic risks all relating to mutations in DNA repair pathways underlie a great deal of breast cancer genetic risk.

Copy Number Variants Underlying ASD

The results of initial studies to identify ASD-related genes underlying ASD began with analysis of CNVs in 165 affected families. These included those defined as both simplex autism cases and multiplex cases with more than one affected sibling (Hall et al. 1990, Sebat et al. 2007).

These results were astounding and had two important consequences. First, while CNVs were not the only causative variants to be found in ASD individuals, the early focus on gene copy number introduced stoichiometry as a key variable to consider in framing disease pathogenesis. In the pioneering CNV study it immediately became evident that most of the de novo CNVs associated with ASD (12 out of 15) were deletions (Sebat et al. 2007), suggesting that a more common mechanism (although not necessarily cause, as discussed below) for CNV-associated stoichiometric variation in ASD was likely associated with loss of function. Such studies could then be tied to specific gene mutations within these regions. For example, studies of Angleman’s syndrome identified a CNV variant resulting from maternal deletion of chromosome 15q11-q13 as being associated with ASD (Cook et al. 1998). Subsequent analysis indicated that the same phenotype could be caused by duplication or triplication of the same region (Baker et al. 1994, Bundey et al. 1994, Schinzel et al. 1994) as well as point mutations of the UBE3A gene within this region (Fang et al. 1999, Kishino et al. 1997, Matsuura et al. 1997, Sutcliffe et al. 1997).

A number of additional studies have underscored the role of stoichiometric balance in ASD pathogenesis. Together, they more clearly define the remarkable observation that either too much or too little of the same gene product could lead to the same clinical ASD outcome. For example, Rett syndrome, a neurologic disorder associated with ASD, is classically defined as haploinsufficiency in the MECP2 gene in females (Amir et al. 1999). However, neurodevelopmental phenotypes are also found in males with MECP2 duplications (Van Esch et al. 2005). Another instance of this stoichiometry phenomenon is seen in Smith-Magenis syndrome, associated with either microdeletion of 17p11.2 (Van Esch et al. 2005) and ASD or de novo 17p11.2 duplication (Potocki et al. 2000). Subsequently, work identified mutations in the RAI1 gene as the dosage-sensitive gene within the 17p11.2 region (Carmona-Mora & Walz 2010).

A major consequence of the success of CNV analysis of ASD families was the robust conclusion that genetics could lead to the identification of single genes that were likely to contribute to ASD. The consequent increased impetus to study ASD genetics pushed forward attempts to undertake more sophisticated genomic studies, and technologic advances moved from copy number studies to WES studies.

Whole-Exome Sequence Analysis in ASD

Whole-exome analysis was pioneered by Hannon and colleagues (Hodges et al. 2007), who recognized that the ability to inexpensively synthesize mass-scale oligonucleotides could produce enough probes to bind to all transcribed exons. The strength of this bind and capture approach was primarily that all coding sequences in nearly all transcribed genes could be directly sequenced. Many such overlapping molecules from captured exons allowed great sequencing depth (complexity) within any individual locus, providing statistical robustness in the ability to call true variants (relative to low abundance technical sequencing errors).

The limitations of WES strategies were severalfold and in part presaged the drive to undertake whole-genome sequencing (WGS) studies (discussed below). The requirement to hybridize and capture individual fragments on an arrayed platform—literally thousands of hybridization reactions done in parallel—meant that the ensuing binding and wash conditions were an average best-case set of conditions for an average oligonucleotide capture. Variants in which GC content varied, for example, would be expected to be variable in their capture from being either too easy (AT-rich) or too difficult to release after binding and wash from capture oligos. Moreover, the length of captured overlapping RNA fragments obscured detection of small insertions and deletions, while the nonlinear nature of PCR amplification obscured analysis of CNVs.

Nonetheless, the strengths of WES combined with earlier methods were evident in the first such studies applied to ASD. These strengths still relied on the design of quad analysis, as early efforts applying WES in trios did not yield significant findings of ASD-related genes. In contrast, WES analysis of quads, published by the Wigler group (Iossifov et al. 2012, 2014) and independently by the groups of Eichler (Krumm et al. 2014, O’Roak et al. 2012), State (Sanders et al. 2012), and Daly (De Rubeis et al. 2014), made huge strides in the understanding of ASD genetics. What started as the identification of small numbers of loss-of-function de novo variants that were proband specific quickly became important and generalizable findings. The reproducibility of finding de novo variants in the same genes by these groups, and the subsequent replication of variants in these same genes in larger WES studies, confirmed with statistical robustness the emerging understanding that mutations in biologically coherent groups of genes relating to neuronal function underlay the development of autism in many cases.

Whole-Genome Sequencing: De Novo Variants in Noncoding RNA

While WES could readily identify many loss-of-function de novo variants, discovery of small insertions or deletions (e.g., of individual exons) or missense mutations that disrupted open reading frames leading to either nonsense-mediated decay (NMD) or loss of messenger RNA (mRNA) levels was more difficult. These limitations became apparent after WES of 10,000 quads available through the Simons Simplex Collection (SSC)—as increasing numbers of individuals were sequenced, the yield of findings began to plateau at ~25–30% of all children with de novo ASD. While the results from WES provided the breakthrough needed to begin to demarcate the functions encoded by proteins that led to susceptibility to ASD, the work also left open the question of what variants were responsible for the ~70% or more of probands that remained undiagnosed (Iossifov et al. 2014, Yuen et al. 2017).

Systematic identification of regulatory variants that could either increase or decrease the levels of gene expression of ASD-associated genes was beyond the reach of WES. Such variants were known to generally preside in noncoding regions of genes. While on occasion such variants could be picked up by WES, the capture design and depth were not able to detect them in either a consistent or comprehensive manner. This obstacle could in theory be overcome by WGS, in which coding and noncoding variants are detected with equal sensitivity. Strategies to identify even small CNVs could be enhanced by WGS relative to WES, and in fact an enrichment of such CNVs has been reported in analysis of WGS data (Fang et al. 2016, Turner et al. 2017).

However, WGS raised the issue of how to systematically interpret the impact of noncoding variants. It was anticipated that there are many variants per individual that are difficult to interpret and that even causative variants would have smaller effect sizes than missense variants in coding sequence.

Interpreting single-nucleotide variants from WGS raised big issues for investigators. Functional variants in introns, for example, could lead to aberrant splicing and switches in exon inclusion that could lead to such consequences as truncated proteins, premature stop codons and NMD, or de novo coding exons that are pathogenic. For example, NMD (Eom et al. 2013) or de novo exons (Cummings et al. 2017) are underrepresented or absent in genomic data sets. More generally, there were not established methods to systematically annotate noncoding functional variants such as, for example, those that impacted the nucleotide-binding sites of specific regulatory factors that in turn would cause RNA misregulation.

At the same time, the idea that noncoding variants might contribute to ASD gained strength from the general observation that regulatory regions are critical to rapid evolutionary change in humans (Indjeian et al. 2016, McLean et al. 2011). Moreover, it was becoming increasingly clear that regulatory sequences were likely to mediate proper stoichiometry in the context of complex disease (Liu et al. 2019). These observations were pertinent to findings that ASD probands on average harbor more noncoding de novo variants, in aggregate, than probands (Liu et al. 2019, Turner et al. 2017).

Understandably, given the large number of noncoding variants and difficulty interpreting them biologically, initial efforts to apply different analytical frameworks for such analyses of WGS data from small numbers of families present in data sets such as MSSNG, or from even larger data sets (519 families) from the SSC, were unable to demonstrate specific noncoding variants that were enriched in probands versus siblings (Werling et al. 2018, Yuen et al. 2017).

Applying Deep Learning Tools to Whole-Genome Sequencing

Machine learning tools were initially applied successfully to identify ASD-associated genetic changes with neural development, underscoring that a key feature of autism is alterations in genes expressed in and hence involved in early brain development (Krishnan et al. 2016), but these tools were not designed to analyze noncoding variants. More recently, machine learning was applied to larger WGS cohorts [up to 1,902 SSC families (Werling et al. 2018, An et al. 2018)] to compare rare and de novo noncoding variants. Using a risk score based on analysis of five large groups of transcription-associated annotations (e.g., functional annotations of histone acetylation, or GENCODE-defined promoters), researchers presented evidence for a contribution of noncoding variants in ASD probands in regions upstream (~750–2000 base pairs) of transcription start sites (An et al. 2018), suggesting that de novo mutations in promoter regions might contribute to ASD. However, after correction for multiple hypotheses (An et al. 2018), no single noncoding category was significant, and the use of a risk score could not identify specific variants with functional changes in transcription of specific genes. Similarly, while targeted and machine learning analysis of additional WGS in 493 multiplex families with autism was successful in identifying many deleterious protein-coding variants, analysis of noncoding variants identified only a few noncoding promoter deletions (in DLG2 and NR3C2), and no global enrichment in promoters was found (Ruzzo et al. 2019).

More recently, an important step forward in interpreting WGS data from the SSC came from a combination of leveraging biochemical demarcation of regulatory variants with deep learning strategies. This more integrated strategy focused on noncoding variants, pairing a deep convolutional-network-based framework with biochemical data demarcating regions of noncoding genome bound by DNA or RNA regulatory proteins. The latter included 2,000 regulatory profiles of histone marks, transcription factor binding sites, and chromatin accessibility identified by epigenetic or chromatin IP studies, as well as large sets of precise RNA regulatory sites identified by 230 crosslinking immunoprecipitation (CLIP) [a method for covalently crosslinking RNA-binding proteins (RBPs) to their RNA targets in situ in tissues (Ule et al. 2003)] biochemistry studies (Zhou et al. 2019).

This approach led to remarkable results, identifying an increased burden of de novo mutations disrupting both DNA-binding proteins and RBPs in ASD probands compared to unaffected siblings. Interestingly, the set of genes found to be under selective pressure (“loss of function intolerant” by the ExAC study; see discussion below) showed a significant concentration of noncoding RBP regulatory-disrupting mutations in ASD probands. Similar results were found for enrichment of de novo mutations disrupting DNA-binding proteins, and the loss of function–intolerant genes identified for RNA- and DNA-binding proteins were convergent, supporting the conclusion that observed noncoding regulatory events played causal roles in ASD. Importantly, both sets of results were independent of any selection for subsets of genes demarcated a priori using risk scores and were significant after multiple-hypothesis correction (Zhou et al. 2019), addressing issues raised in earlier WGS machine learning studies (An et al. 2018, Werling et al. 2018).

The biological implications of this strategy (Zhou et al. 2019) were important in finding that noncoding regulatory variants implicated gene regulation of genes and transcripts involved in synaptic transmission, neuronal development, and autism and were consistent with previous studies (Iossifov et al. 2014, Krishnan et al. 2016, Packer 2016). Moreover, they pointed the way to a means for annotating specific noncoding variants more precisely as disease associated. For example, the 57 genomic variants predicted to have high impact in transcriptional regulation and enriched in probands could be demonstrated to have differential transcriptional activity relative to wild-type (WT) sequences in cell-based assays (Zhou et al. 2019). In addition, a noncoding mutation, outside of canonical splice sites, was found to disrupt splicing of SMEK1 transcripts (encoding a protein that regulates neurogenesis through Wnt signaling; the fraction of exon inclusion [delta-I; (Ule et al. 2005)] relative to WT was highly significant [≃ 0.4 (Zhou et al. 2019)]). Moreover, reanalysis of the gene lists analyzed in Werling et al. (2018) with this new strategy revealed that noncoding regulatory mutations that were predicted to have posttranscriptional (RBP) regulatory effects in probands (i.e., large effect size between proband and sibling) were specifically enriched in ASD susceptibility genes (Zhou et al. 2019, figure 1c).

Together, these studies suggest a pathway forward for annotating noncoding variants in WGS data from ASD patients, a significant issue given the finding that only a minority of ASD individuals sequenced have been found to harbor coding mutations. While the selection coefficient for individual noncoding variants may be smaller than for coding variants, it is worthwhile noting that they also may account for more subtle differences in ASD phenotypes. For example, de novo noncoding variants were found to be associated with lower IQ in a subset of ASD individuals with RNA-binding mutations in introns flanking alternatively spliced exons (Zhou et al. 2019). At the same time, while these studies demonstrate a path forward by which WGS can identify rare regulatory variants with large effects, complex traits such as those seen in ASD are likely to be impacted by many genes contributing incrementally to any individual’s phenotype, and the actions and effects of any one trans-regulatory factor–DNA/RNA interaction may well be relatively small (see discussion in Liu et al. 2019).

ASD SUSCEPTIBILITY GENES

Genomic analysis has led to the identification of a vast number of ASD susceptibility genes. In general, their identification has relied on careful clinical descriptions combined with genomic sequencing, most often through analysis of quads. Some interesting exceptions to this successful strategy have emerged from more targeted studies. Careful clinical phenotyping of congenital syndromic disorders led to recognition that they may be associated with ASD phenotypes. Examples include Angelman syndrome (Khatri & Man 2019), the fragile X syndrome (FXS) (Niu et al. 2017), tuberous sclerosis (Krueger & Bear 2011), Rett syndrome (Moretti & Zoghbi 2006), Coffin-Siris syndrome 1 (CSS1) (Santen et al. 2012), and Asperger’s syndrome (Sebat et al. 2007). An unusual report of a syndrome involved a single sperm donor, in which an impacted recipient mother gave birth to two boys with ASD; crowdsourcing then traced a total of 12 offspring from the same donor sperm, all of whom had ASD (Cha 2019). This sperm was given to the geneticist Steve Scherer, who is undertaking genomic analysis.

Genomic studies have identified an expanding number of ASD susceptibility genes, with high confidence estimates of 53 genes in 2015 (Sanders et al. 2015), 102 genes in 2018 (Satterstrom et al. 2020), and 106 by mid-2019 (Iakoucheva et al. 2019), with some arguing that there may be up to 1,000 susceptibility genes (Ramaswami & Geschwind 2018). More may be identified in the future by analysis of whole-genome sequence, allowing an understanding of regulatory variants (see above), as well as by better clinical understanding of genes that have complex phenotypes that both contribute to and mask overt ASD. The latter refers to the fact that ASD is a neurodevelopmental disorder of variable severity presenting years postnatally, such that variants that could give rise to clinical autism may present earlier with severe clinical symptoms, precluding an ASD diagnosis.

Nearly all of the genes identified through WES as ASD susceptibility genes are associated as haploinsufficiencies (Iakoucheva et al. 2019). While there is some ascertainment bias, given the approaches used (e.g., searching for gene-disrupting/loss-of-function mutants), this nonetheless underscores the concept that stoichiometric control of key neuronal proteins is necessary for intellectual function. This point is further underscored in consideration of genes found associated with ASD susceptibility through study of CNVs (see also the above section titled Copy Number Variants Underlying ASD). Some genes in this category are linked to ASD through either overexpression or underexpression of the same gene. For example, while NRXN1 is one of the most well-studied ASD susceptibility genes associated with loss of function, by either CNV or likely gene-disrupting mutations, duplications of NRXN1 have also been associated with ASD (Wiśniowiecka-Kowalnik et al. 2010). Both deletion and duplication of 16p11.2 (Hippolyte et al. 2016, Pizzo et al. 2019, Rosenfeld et al. 2010) are associated with ASD. Similarly, duplication of 22p13, a region encoding SHANK3, and deletions of SHANK3 are clinically associated with intellectual disability and ASD (Jiang & Ehlers 2013, Wilson et al. 2003). Moreover, point mutations in SHANK3 are associated with ASD (Durand et al. 2007, Ruzzo et al. 2019), restoration of SHANK3 function in SHANK3-deleted neurons restores synaptic deficits (Shcheglovitov et al. 2013), and similar findings of reversible autism-like defects were observed in mice harboring a Shank3 mutation (Speed et al. 2015, Mei et al. 2016), although these mice studies are complex and need additional confirmation (Speed et al. 2019). Interestingly, Sebat and colleagues (Iakoucheva et al. 2019, Malhotra & Sebat 2012) have noted that such over- and underexpression of the same gene regions associated with ASD (1q21.1 and 16p11.2) can also be associated with either macrocephaly or microcephaly.

Given the early action and pleiomorphy of ASD genes, there is some evidence that the full set of susceptibility genes may be overlooked because their phenotypes with respect to ASD may be obscured in some instances by more severe phenotypes. An illustration of how ASD susceptibility genes could thus be masked is given by considering mutations associated with the chromatin remodeling protein ARID1B. Haploinsufficiency of ARID1B has been recognized as underlying the congenital disorder CSS1, and in some children this results in intellectual disability and ASD (Santen et al. 2012). However, it has subsequently been recognized that newborns presenting with ARID1B mutations may present with a wide spectrum of clinical symptoms, including life-threatening malformations of major organs [microcephaly (10%) and severe cardiac defects (ventricular septal defects, coarctation of the aorta; 8%) (Mannino et al. 2018)] that may preempt an ASD diagnosis that would otherwise manifest later in life [for example, after surgical repair of previously devastating cardiac anomalies (Bean Jaworski et al. 2017)]. Analogous gene variants can cause severe perinatal defects but may also, in milder clinical manifestations or after early aggressive medical/surgical management, remain to be identified and hence contribute to greater understanding of developmental pathways associated with ASD.

To date, two major sets of genes that encode quite distinct functions have been identified by CNV analysis and WES. These sets can be described as genes encoding proteins involved in chromatin regulation and those encoding proteins involved in synaptic function, particularly relating to the balance between excitation and inhibition. In both instances, the theme of regulation of stoichiometric levels of gene control per se is evident in ASD and consistent with the overall model that the fine-tuning of neuronal activity and gene expression may be fundamental in allowing normal neuronal development and ultimately cognitive function.

Gene Regulation and ASD Susceptibility

Proteins that regulate transcription or chromatin state are vastly overrepresented in the known set of ASD susceptibility genes, together accounting for nearly one-third of the currently identified genes (Iakoucheva et al. 2019). Transcriptional regulators and their binding partners include both general factors and those associated with specific neuronal functions, such as ADNP, which may be regulated by and associated with vasointestinal peptide signaling (Sragovich et al. 2019).

In addition to transcription factors, an even larger set of ASD susceptibility genes encode chromatin modifiers. These again include general factors, some of which are implicated in regulating neuronal function, such as BRD4 (see below). While many functions of these proteins regarding specific actions in neuronal gene regulation are not yet known, some again may be more specifically tied to regulation of neuronal genes. For example, PHF21, believed to be the gene within 11p11.2 deletions (Potocki-Shaffer syndrome) that mediates intellectual disability (Kim et al. 2012), is a histone deacetylation/demethylating protein that is also part of the LSD1-CoREST corepressor complex thought to repress expression of neuronal genes (Porter et al. 2018).

RNA regulation and ASD susceptibility.

A second aspect of gene regulation intersecting with ASD susceptibility is the identification of large numbers of RNA regulatory genes. These fall into various categories, including factors involved in splicing (see below) and other aspects of RNA regulation (3 UTR binding, RNA localization; CELF4, ELAVL3), translational regulation [EIF3G, FMRP (see below)], and microRNA regulation and mRNA turnover (TNRC6B).

Understanding the role of RBPs in ASD has benefited from the ability to identify, with nucleotide resolution, their direct RNA-binding sites using CLIP methods (Ule et al. 2018). In ASD, a number of the networks of regulated RNAs identified show remarkable overlap with ASD susceptibility genes. The foremost among this group is FMRP, discussed below. Networks regulated by 3′UTR–binding factor CELF4 have also been found to overlap ASD susceptibility gene networks, specifically those involved in neuronal excitation (Porter et al. 2018, Wagnon et al. 2012).

Splicing factor networks regulated by RBPs also show a significant relationship to ASD (Gandal et al. 2018, Voineagu et al. 2011). When the targets of RBFOX splicing factors were determined by CLIP, some regulated exons were found in ASD susceptibility genes, including NLGN3 (Zhang et al. 2008), and subsequent studies found that over 10% of RBFOX-regulated alternative exons are regulated from among 48 ASD susceptibility genes, including SHANK3 (Weyn-Vanhentenryck et al. 2014). These findings are complemented by additional analysis of splicing (Fogel et al. 2012) and cytoplasmic transcripts regulated by RBFOX1 and implicated in ASD, including cytoplasmic transcripts encoding ASD-related proteins affecting synaptic function and calcium signaling (Lee et al. 2016). Together, these studies are consistent with the strong link between RBFOX proteins in autism, observed in both studies of copy number variation in ASD (Gambin et al. 2017, Sebat et al. 2007) and the original case report of a 160-kb deletion in the RBFOX1 gene (Martin et al. 2007).

Several studies have implicated the NOVA splicing factors (Buckanovich et al. 1993, Yang et al. 1998) in both RBFOX and ASD biology. NOVA expression is extremely restricted to neurons in the central nervous system, an observation that first suggested that the mammalian brain regulates RNA in uniquely specified ways (Darnell 2013). Several NOVA-regulated precursor mRNAs harboring YCAY clusters were originally described in cellular and in vitro splicing assays and in Nova-null mice (Dredge & Darnell 2003; Dredge et al. 2001, 2005; Jensen et al. 2000; Musunuru & Darnell 2001; Saito et al. 2019; Ule et al. 2006), which established NOVA as the first tissue-specific alternative splicing factor in vertebrates. Subsequent unbiased, genome-wide approaches afforded by CLIP, including pioneering exon-junction splicing microarrays (Ule et al. 2005), HITS-CLIP (Licatalosi et al. 2008), and bioinformatic analyses (Ule et al. 2006, Zhang et al. 2010), identified a robust set of ~800 target alternative exons (Ule et al. 2018).

Together, these studies allowed bioinformatic analyses that revealed robust overlap with ASD susceptibility genes. Bioinformatic analysis of target transcripts, based largely on analysis of HITS-CLIP data, tied NOVA to both RBFOX proteins and ASD (Zhang et al. 2010). Moreover, a subset of NOVA-regulated transcripts identified by Bayesian network analysis were found to harbor orphan conserved elements not bound by NOVA proteins themselves. Further analysis revealed that these were RBFOX-binding elements, and NOVA and RBFOX proteins were found to synergize in regulating splicing cooperatively (Zhang et al. 2010). Independently, NOVA targets were examined for potential relationships to neurologic disease, including ASD. Among 358 NOVA target transcripts identified by network analysis, 88 were implicated in genetic disorders, including mental retardation, epilepsy, and autism. Interestingly, the latter were encoded by the RBFOX1 gene as well as a host of other ASD susceptibility genes (AUTS2, CADM1, GRIK2, NRXN1, SLC4A10, etc.). Together, 8.5% of transcripts coordinately regulated by both NOVA and RBFOX proteins are implicated in autism (Zhang et al. 2010).

Notably, very specific delineations of misspliced transcripts are beginning to be identified as correlated and potentially causally related to ASD (reviewed in Quesnel-Vallières et al. 2019). Conserved neuronal microexons, highly conserved exons that are believed to modulate protein-protein interactions (Scheckel & Darnell 2015), have been found to be frequently misregulated in the ASD brain, again often in transcripts associated with neuronal vesicle transport (Irimia et al. 2014), and more recently in transcripts encoding translational elongation factors (Gonatopoulos-Pournatzis et al. 2020). Several RBFOX transcripts show dysregulated splicing in the ASD brain (Voineagu et al. 2011). An assessment of allelic imbalance in proband versus siblings has implicated small nucleolar RNA–targeted splicing changes in ASD-related target genes (Lee et al. 2019), and a specific error in splicing of the CPEB4 transcript, which encodes a polyA-binding protein linked to deadenylation, has been associated with ASD (Parras et al. 2018). Finally, since RNA regulatory elements identified by CLIP have primarily been identified in noncoding regions of transcripts (with FMRP binding elements a notable exception), these studies have led the way into new approaches to understand ASD susceptibility variants in noncoding regions (see discussion above and Zhou et al. 2019).

Excitatory/inhibitory balance and ASD susceptibility.

Perhaps the earliest considerations regarding the imbalance in neuronal excitation/inhibition in ASD came from hypotheses based on early genetic studies of autism and observations from mouse models (Rubenstein & Merzenich 2003), combined with early studies on FXS. FXS is directly relevant to autism, given the extremely high prevalence of ASD in children with FXS. Early studies by Bear, Huber, and colleagues (Bear et al. 2004; Huber et al. 2000, 2002) studying the synaptic pathology in Fmr1-null neurons demonstrated that FMRP plays a critical role in the regulation of a key receptor for the excitatory neurotransmitter glutamate, specifically the metabotropic glutamate receptor mGluR5. These studies led to the mGluR hypothesis of FXS in which imbalance of glutamate-dependent excitation/inhibition underlay the development of intellectual disability in FXS (Bear et al. 2004; Huber et al. 2000, 2002; Parras et al. 2018; Zhang et al. 2009). Moreover, as mGluR5 mRNA was found to be directly bound to the translational regulatory protein FMRP (Darnell et al. 2011), these studies tied synaptic translation to long-term depression in hippocampal CA1 neurons in FXS animal models.

Subsequently, the many genes that have been identified as ASD susceptibility genes (as well as FMRP targets) have obvious direct and indirect (downstream signaling) functions in excitatory or inhibitory neurotransmission. These include both susceptibility genes themselves and direct and indirect protein-protein interaction networks, as recently reviewed in detail by Wall and colleagues (Ruzzo et al. 2019; see also Darnell et al. 2011). One unexplained observation is that large genes are disproportionately represented in ASD studies (Sebat et al. 2007) and as FMRP targets (Darnell et al. 2011). Another and perhaps more salient connection is that ASD susceptibility genes and genes encoding FMRP target RNAs are under high selective pressure (Darnell et al. 2011, Iossifov et al. 2012). This concept has since been applied more generally, with the development of loss of intolerance scores, to successfully narrow lists of candidate genes in other complex neurogenetic disorders such as epilepsy (Petrovski et al. 2013).

The protein products of genes directly identified as ASD susceptibility loci include 98 categorized by analysis of whole-genome sequences from 2,308 individuals (Ruzzo et al. 2019) and what turns out to be a largely overlapping set of 106 ASD susceptibility genes comprehensively assembled by Sebat and colleagues (Iakoucheva et al. 2019), and the reader is referred to both of those reviews for details and discussion. Many of these ASD susceptibility genes encode inhibitory (GABABR2, GABABR3, SLC6A1) or excitatory (GRIA2, GRIN2B) neurotransmitter receptors or transporters and ion channels that impact neuronal sensitivity, including potassium (KCNMA1, KCNQ3), calcium (CACNA1E, CACNA2D3), and sodium (SCN1A, SCN2A) channels and allied proteins (e.g., NRXN1, involved in calcium channel function, and ANK2, involved in ion channel localization). Together, these observations support the identification of excitatory/inhibitory imbalance as one means of leading to an ASD phenotype.

ASD and Epilepsy

The relationship between ASD and the balance between neuronal excitation and inhibition is underscored by considering the relationship between ASD and epilepsy (Lee et al. 2015). Up to 20% of children with ASD have epilepsy, and this relationship was noted early by the association of tuberous sclerosis with ASD (Walsh et al. 2008). Tuberous sclerosis is a congenital neurodevelopmental disorder associated with heterotopias of ectopic groups of neurons in the brain, and these are believed to be epileptic foci. These correlations in tuberous sclerosis complicate conclusions regarding the cause-effect relationship between epilepsy and ASD, a caveat that in general (although see discussion of BCKDK below) complicates interpretation of the coincidence of epilepsy and autism.

Specific genetic variants associated with ion channel function and the maintenance of excitatory/inhibitory balance have been associated with ASD. For example, variants in the potassium channel inward rectifier gene KCNQ3 have recently been associated with ASD. Interestingly, molecular analysis of these missense mutations indicate that they result in a gain of function (Sands et al. 2019); conversely, loss-of-function mutations in KCNQ2 or KCNQ3 can also cause neurodevelopmental problems, including epilepsy and global neurologic disability (Miceli et al. 2014, Sands et al. 2019). This observation emphasizes the delicate nature of KCNQ3 (and other potassium channels related to ASD, e.g., KCNMA1) stoichiometric control and its relationship to ASD and epilepsy.

Indeed, the frequent concurrence of epilepsy in ASD patients may directly relate to RNA regulation. Human and mouse studies have demonstrated that haploinsufficiency of RNA regulatory proteins is frequently associated with epilepsy, as seen with FMRP (Lee et al. 2015), NOVA1 and NOVA2 (Eom et al. 2013), ELAVL4 (Ince-Dunn et al. 2012), RBFOX proteins (Bhalla et al. 2004; Lal et al. 2013a,b), and CELF4 (Wagnon et al. 2012). The relationship between RBP stoichiometry and the nature of excitatory/inhibitory balance in ASD is particularly relevant given the findings of Zhou et al. (2019), discussed above, demonstrating that noncoding variants that affect RBP binding are enriched in ASD probands and ASD susceptibility genes.

Another relationship exists between dysregulation of metabolic pathways, ASD, and epilepsy. In one particularly interesting example, recessive gene-disrupting mutations in BCKDK, the gene encoding branched-chain ketoacid dehydrogenase kinase, were identified in families with autism, epilepsy, and intellectual disability (Novarino et al. 2012). Interestingly, mouse models (Bckdk-null mice) recapitulated neurobehavioral defects that were ameliorated with dietary supplements, suggesting both a potential treatment for these patients and a dissociation of epilepsy from the developmental aspects of intellectual disability in this instance of ASD. The gene encoding the HDL-binding protein HDLBP is downregulated in ASD (Satterstrom et al. 2020), and there is a relationship between epilepsy, metabolic defects, and verbal learning defects, specifically including lower HDL levels (Hermann et al. 2017).

Finally, a smaller but significant number of ASD susceptibility genes encode proteins involved in actin or microtubule biology (see, for example, Ruzzo et al. 2019), and there is emerging interest in the relationship between epilepsy and microtubule dynamics (Xu et al. 2016). This intersection is not well articulated in experimental science other than through observations of actions of pharmacologic agents that act on microtubules, which may exacerbate or ameliorate human epilepsy. In ASD, such a relationship may converge on the need for neurons to organize processes in development and to do so in rapid time scales, for example, to reorganize microtubular arrays in response to synaptic plasticity stress. This question may be developed further, given the finding that ASD susceptibility genes include proteins associated with microtubule dynamics such as MAP1A (Myers et al. 2011), NCKAP1 (Zweier et al. 2019), and DYNC1H1 (Ding et al. 2016), which is essential for retrograde axonal transport, as well as several that encode proteins [KATNAL2, SPAST (Matthews et al. 2017)] that are believed to act to sever microtubules and promote rapid reorganization of the neuronal cytoskeleton.

AUTISM AND THE STOICHIOMETRY OF BRAIN REGULATION

A general concept of gene balance underlying variations in human phenotype and disease was put forth over a decade ago based on considerations of the consequences of gene duplication (Conrad & Antonarakis 2007). More specifically in ASD studies, gene dosage effects leading to cognitive development tie the control of neuronal plasticity to the control of protein levels mediated by transcriptional and translational controls. In part, this stoichiometric control seems likely to include those proteins encoding the balance between neuronal excitation and inhibition, a point emphasized above by the observation that 20% of ASD patients have epilepsy. More generally, the idea of gene balance relates to the concept of homeostatic plasticity, discussed below and considered in great detail by Turrigiano and colleagues (Turrigiano & Nelson 2004, Turrigiano et al. 1998).

Autism and the Fragile X Syndrome

An aspect of stoichiometric misregulation predisposing to ASD—related to but distinct from variants in chromatin or synaptic function–associated genes—came from the study of FXS. Approximately 50% of children with FXS are diagnosed with ASD. Nearly all children with FXS harbor a triplet repeat in the 5′ UTR of the fragile X gene, FMR1 (Santoro et al. 2012). This leads to hypermethylation of the DNA locus and a loss of function through transcriptional silencing of the FMR1 gene. The FMR1-encoded protein, FMRP, was recognized to harbor two KH-type RNA-binding domains, a spacer region, and a third RGG-type RNA-binding domain.

Precisely what mechanism might underlie the connection between FXS and ASD first came to light by identifying FMRP’s RNA targets in the brain. FMRP was recognized as similar in overall architecture to the KH domain containing NOVA neuron-specific RBPs, leading to the application of CLIP—first developed for analysis of Nova function (Ule et al. 2003)—to FMRP. This identified 842 robust mRNAs bound by FMRP in the mouse brain (Darnell et al. 2011), a list now expanded to more than 1,000 transcripts (Van Driesche et al. 2019).

Remarkably, these studies revealed that the ranked list of FMRP-bound brain transcripts fell into several major gene ontology categories that closely overlapped the nascent set of ASD targets (Darnell et al. 2011). This observation was replicated independently as more ASD targets were identified by WES analysis of quads (Iossifov et al. 2012). At the top of the FMRP mRNA-bound target list were synaptic proteins and chromatin modifiers. These studies also led to the discovery that FMRP target and ASD risk genes were under a high degree of selective pressure, implying that they encoded key cellular functions (Iossifov et al. 2012). This concept, subsequently termed loss-of-function intolerance, has been applied more widely in genomics (Petrovski et al. 2013), including, for example, to amyotrophic lateral sclerosis (Cirulli et al. 2015). The variant intolerance evident in FMRP and ASD joint targets underscores how, when mutated, such genes may lead to severe neurologic phenotypes, including not only intellectual disability and ASD but potentially other neurologic disorders such as schizophrenia (Fromer et al. 2014).

FMRP CLIP not only delivered a biochemically robust list of FMRP targets but revealed what is likely the major function of the FMRP protein. FMRP CLIP results differed from those previously seen with most other RBPs such as NOVA, which bound to discrete RNA elements through sequence-specific interactions mediated by the KH-type RNA-binding domains. In vitro, FMRP was found to bind complex RNA elements (G-quadruplex and RNA pseudoknots), revealed by RNA selection, biochemistry (Darnell et al. 2001, 2005), and structural studies (Phan et al. 2011, Vasilyev et al. 2015). In vivo, while FMRP CLIP demonstrated binding directly to a specific subset of brain transcripts, within those transcripts its primary binding appeared indiscriminate, covering entire coding sequences (Darnell et al. 2011). This observation, together with the finding that the vast majority of FMRP protein within the brain is present on polyribosome fractions of sucrose gradients (Darnell et al. 2011, Stefani et al. 2004), led to biochemical studies of the role of FMRP in translational regulation, an issue that had previously been raised by in vitro (nonstoichiometric) FMRP studies (Laggerbauer et al. 2001).

Assessment of a role for FMRP in translation done in WT and Fmr1-null brains revealed several points. FMRP RNA binding was reliant on its KH domains, as mutations in the protein’s KH1 (Myrick et al. 2014) or KH2 (Zang et al. 2009) domains (mutations that reproduced those seen in two clinical rare cases of FXS in which the FMR1 gene did not harbor triplet repeats) abrogated polysome association and RNA binding. Moreover, it was found that FMRP binding to polyribosomes on its target mRNAs was associated with both binding to ribosomes (observed by electron microscopy) and stalled ribosomes (observed in biochemical assays) (Darnell et al. 2011). These findings were supported by additional in vitro studies showing an increased translation rate in the absence of FMRP (Udagawa et al. 2013). The findings suggested that FMRP binding to specific mRNAs (via still undiscovered specificity determinants) leads to reversible ribosomal elongation arrest, thereby inhibiting translation of those transcripts.

Additional direct actions of FMRP have since been described, including association with CAPRIN1, in vitro assembly of protein-protein (and likely RNA) complexes into insoluble cell fractions (El Fatimy et al. 2012, Kim et al. 2019), and undefined (direct and/or indirect) actions of FMRP on elongation factors and translational initiation that may act to balance FMRP actions on ribosomal elongation (Das Sharma et al. 2019). To date, these studies together point to a model in which FMRP loss in FXS (and by proxy ASD) results from a primary defect in translational control of specific target mRNAs.

Since some of the FMRP targets themselves mediate translational controls, one net takeaway from studies of FMRP function is that FMRP may act at multiple levels to regulate the stoichiometry of proteins involved in synaptic function and chromatin modification. This is consistent with the observation made above that FMRP acts to inhibit the synthesis of dosage-sensitive gene products in neurons. Several such targets, including some whose loss of function has been associated with ASD, are NRXN1, NLGN3, SHANK3, TSC2, and CENTG2, consistent with the idea that stoichiometric control of key neuronal transcripts is important in autism.

Interestingly, the number of splice site variants (as well as other nonsense mutations) present in FMRP targets was also found to be significantly decreased (Iossifov et al. 2012). Thus, there appears to be a strong purifying selection in humans against splicing errors or nonsense mutations in FMRP, suggesting the importance of RNA-protein stoichiometry for key autism-related transcripts, and the potential for overlap between splicing and translational regulatory networks. A potential relationship between FMRP-dependent translational control and RBFOX/NOVA-dependent splicing control in autism remains largely unexplored.

Stoichiometric Control of Synaptic Excitatory/Inhibitory Balance and Homeostatic Plasticity

A summary of all genes identified to date in ASD suggests subgroups of disorders—for example, those involved in excitatory/inhibitory balance, and those involved in regulation of transcription. A unifying feature of these two sets is that they themselves are regulated by the translational control protein FMRP.

The genes involved in these biological balances, between neuronal excitability and molecular output of the genome, may be considered as a group to be involved in maintaining normal stoichiometric output of these proteins. Each of these domains in turn relates to ASD in more specific ways. Fifteen percent of males with FXS have epilepsy, and 25% of FXS and ASD children have concurrent epilepsy (Niu et al. 2017). While this remains a poorly understood general phenomenon, the observations fit well with the critical need to control excitation/inhibition balance, a key underlying feature of both epilepsy and ASD.

The need to control the translation of excitatory/inhibitory proteins for cognitive function fits with our understanding of Hebbian synaptic control and synaptic learning, in which coincidence detection requires translation during a brief (~20–60 min) sensitive window. There is great interest in the molecular neuroscience community in whether this window of translational control may in turn relate to the local regulation of translation of synaptic mRNAs (Hafner et al. 2019). Indeed, it will be of interest to understand whether such regulation by local translational regulators such as FMRP may play such a role.

A local synaptic mechanism for tight regulation of the stoichiometry of synaptic proteins does not preclude the need to maintain general homeostasis in the neuron. This concept, termed homeostatic plasticity, has been considered in great detail in studies of monocular deprivation and plasticity in the visual system by Turrigiano and colleagues (Miska et al. 2018). Interestingly, preliminary studies suggest that the synaptic scaffolding protein SHANK3, which itself is a stoichiometric target in ASD (see above), is also necessary for homeostatic plasticity in the brain (Hafner et al. 2019, Tatavarty et al. 2018).

Moreover, SHANK3 variants can lead to either excitatory or inhibitory imbalance in different neuronal populations (Monteiro & Feng 2017, Van Driesche et al. 2019). While this argument could reflect a general requirement for SHANK3 (and other ASD targets) in homeostatic plasticity across whole neurons or neuronal circuits (Hafner et al. 2019, Tatavarty et al. 2018), they clearly reflect the critical nature of proper stoichiometric control of SHANK3 in ASD.

Given the role of SHANK3 and other synaptic proteins in local synaptic control and their relationship to Hebbian function and local synaptic plasticity, their stoichiometric control is again likely to reflect excitatory/inhibitory imbalance at the level of individual synapses. Indeed, SHANK3 and many other synaptic proteins are direct targets of FMRP (Darnell et al. 2011), linking FMRP’s role in local synaptic translational controls to its role in the neuronal cell body to circuit wide homeostatic control.

Stoichiometric Control of Transcriptional Regulation in ASD and Homeostatic Plasticity

Proper stoichiometric control of the level of chromatin regulators is critical in ASD, given their role both as genetic susceptibility genes and as targets of translational overexpression in the absence of FMRP. Indeed, FMRP loss leads to overexpression of one such regulator, the bromo-domain protein BRD4 (Korb et al. 2017, Monteiro & Feng 2017, Van Driesche et al. 2019), and missense mutations in BRD4 [for example, P113Q (Iossifov et al. 2012)] have been seen in ASD probands. Mutations leading to decrease of BRD4 levels lead to downstream dysregulation of synaptic proteins, memory defects, and impacts on seizures. Interestingly, correcting stoichiometric imbalance of BRD4 overexpression in Fmr1-null mice, using the BRD4-specific inhibitor JQ1, corrects corresponding synaptic and physiologic abnormalities (Korb et al. 2015, 2017).

FMRP as General Molecular Organizer

Three arguments may be made regarding FMRP as a central regulator of stoichiometric control in autism. First, there is an undisputed relationship between the two disorders; ~60% of males with FXS meet diagnostic criteria for ASD (Niu et al. 2017). Second, FMRP plays a role in the control of overall gene expression—most clearly through translational regulation of transcriptional controls. Such a relationship between factors impacting translational state, through regulation of FMRP stoichiometry itself and/or its impact on gene expression, for example, through modulation of chromatin state and transcription factors, is suggestive of a global role of FMRP on transcriptional regulation (Korb et al. 2017). This suggests a potential relationship between FMRP and neuronal homeostasis.

The third activity of FMRP relates to its role in the synapse. Most of this literature (but not all) relates to the presence, regulation, and activity of FMRP locally in the cell dendrite. Neuronal dendrites harbor ribosomes at the bases of spines and beneath the synapse (Ostroff et al. 2002), harbor a discrete but sizable number of transcripts (Glock et al. 2017, Tushev et al. 2018), and harbor the FMRP protein itself (Bassell & Warren 2008), as do axon termini (Akins et al. 2017, Christie et al. 2009), and FMRP may act in both compartments. Functionally, FMRP acts to stall ribosomal elongation (Darnell et al. 2011). Taken together, these observations suggest that synaptic activity could have unique (and localized) actions on translation and the stoichiometry of key synaptic proteins.

For example, FMRP binds to the mRNA encoding several dendritic neurotransmitter receptors, including key receptors such as the metabotropic glutamate receptor and the NMDA receptor, which are present at low molar stoichiometry within individual dendrites [i.e., ~20 molecules (Sheng & Hoogenraad 2007)]. Regulated translation of even two extra receptor molecules would yield a 10% increase in receptor stoichiometry, which in turn could have a long-term impact on the function of that synapse. Such a model would be consistent with a number of proposed means by which FMRP could be regulated, including those involving Hebbian mechanisms that lead to calcium influx, potentially triggering FMRP dephosphorylation (Lee et al. 2011) and acute and transient changes in its ability to regulate translation (Ceman et al. 2003, Muddashetty et al. 2011). Since FMRP’s action to bind mRNA and inhibit ribosomal elongation is reversible in experimental settings (Darnell et al. 2011), these observations underscore a model for FMRP action in normal and ASD synapses in which local Hebbian input leads to stoichiometric variation in protein molarity and synaptic function.

An overarching model (Figure 1) would place FMRP as a sensor of neuronal activity (for example, through calcium-dependent coincidence detection) in both the synapse and cell soma, impacting its ability to drive translation-dependent synaptic activity and transcriptional state. The readout of its activity would be a direct action on protein stoichiometry, both within the synapse (as one example, via modulation of NMDA receptors) and globally (on transcription state, as one example, acting on BRD4 levels); Figure 1 relates additional aspects of this model. An interesting aspect of such a model is that it provides a means to contribute simultaneously to regulation of both localized protein synthesis within specific synaptic compartments and global gene regulation, contributing toward a potential mechanism governing both synaptic learning and homeostatic plasticity within the neuron. Such a model integrating mechanisms that may jointly impact the neurobiology of learning and homeostatic plasticity in the neuron reflects the growing understanding of how genomics contributes to our understanding of neurologic disorders of intellectual function of the complex sort observed in ASD.

Figure 1.

Figure 1

A model for stoichiometric integration of neuronal activities in autism. Autism spectrum disorder (ASD) susceptibility genes fall into three major classes under different levels of stoichiometric control. Each has a connection to ASD and FMRP: synaptic proteins and chromatin regulators are both major ASD gene susceptibility targets and major FMRP targets, while the loss of FMRP in fragile X syndrome leads to ASD in about half of affected children. This model connecting FMRP, intellectual function, and ASD proposes that FMRP integrates protein expression and stoichiometry in two ways. First (➊), within the synapse, FMRP integrates local signals (such as Ca2+-mediated signals) to modulate translation (Darnell et al. 2011) and local stoichiometry of proteins involved in synaptic plasticity and protein degradation. Second (➋), within the cell soma, FMRP integrates the sum of dendritic signaling globally, modulating RNA translation to regulate the stoichiometry of nuclear proteins—chromatin modifiers and transcription factors (➌)—to effect global homeostasis in the neuron. This model is supported by the recognition that local dendritic translation plays an important role in protein synthesis–dependent synaptic plasticity (Kang & Schuman 1996). Such local control of synaptic protein stoichiometry can thereby be differentially regulated in independent synapses within a single neuron (via translational control, by FMRP as a local regulator of ribosomal elongation, and by other factors). Notably, joint synaptic-related, FMRP-regulated target messenger RNAs (mRNAs) and ASD risk alleles also include proteins involved in microtubule/actin dynamics and protein turnover (not pictured), including those involved in protein degradation pathways [e.g., the ubiquitin ligase Ube3A; reviewed in Kasherman et al. (2020)]. A connection with Hebbian plasticity (involving NMDA receptor–mediated Ca2+ signaling) and FMRP regulation is consistent with data suggesting that the ability of FMRP to suppress translation may be regulated by Ca2+-dependent phosphatases (see main text). Globally, there is a need for homeostasis within the whole neuron, as it simultaneously balances thousands of synaptic excitatory/inhibitory signals without triggering neuronal electrical instability. This concept reflects the need for homeostatic plasticity, as articulated by Turrigiano and colleagues (Turrigiano & Nelson 2004; Miska et al. 2018), with particular connection to ASD exemplified through studies of SHANK3 (Tatavarty et al. 2018). A molecular mechanism in such neuronal homeostatic control, acting over different timescales ranging from milliseconds to days (Abraham & Williams 2003), is proposed here as the net feedback between translational control of synaptic activity locally, cell soma regulation of mRNAs encoding chromatin regulators and transcription factors (➌), and additional joint FMRP-regulated targets and ASD risk alleles, including proteins involved in RNA binding/regulation (not pictured; ELAVL3, EIF3G, TNRC6B; see text and Iakoucheva et al. 2019). Again, these transcripts encoding gene regulatory functions are enriched as FMRP targets (Darnell et al. 2011, Iossifov et al. 2012) and are implicated in ASD (Satterstrom et al. 2020; see text). An untested hypothesis is that mRNAs encoding different functions—synaptic vs. nuclear factors—are differentially localized in the dendrite and soma. In sum, this model suggests that net signaling from synapses may be integrated in the soma in a fast timescale by electrical integration at the axon hillock and at a slower timescale by translational regulation (e.g., as mediated by FMRP, via a balance between phosphorylation/dephosphorylation status), leading to stoichiometric homeostasis of synaptic proteins and nuclear transcriptional state.

ACKNOWLEDGMENTS

The work presented here incorporates the efforts of many wonderful colleagues, to whom the author is deeply indebted, and particularly thoughtful discussions with Olga Troyanskaya, Chris Park, Jian Zhou, Chandra Theesfeld, Julia Moore-Vogel, Evan Eichler, Michael Wigler, Rick Lifton, and Jim Simons. This work was supported by the National Institutes of Health (NS034389, NS081706, NS097404, and 1UM1HG008901), the Simons Foundation (SFARI 240432), and the Rockefeller University Hospital NCATS and CTSA program (UL1 TR000043). R.B.D. is an Investigator of the Howard Hughes Medical Institute.

Footnotes

DISCLOSURE STATEMENT

The author is a scientific consultant and receives fees or stock options for services to Atreca Inc. and OPNA.

LITERATURE CITED

  1. Abraham WC, Williams JM. 2003. Properties and mechanisms of LTP maintenance. Neuroscientist 9(6):463–74 [DOI] [PubMed] [Google Scholar]
  2. Akins MR, Berk-Rauch HE, Kwan KY, Mitchell ME, Shepard KA, et al. 2017. Axonal ribosomes and mRNAs associate with fragile X granules in adult rodent and human brains. Hum. Mol. Genet 26(1):192–209 [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Amir RE, Van den Veyver IB, Wan M, Tran CQ, Francke U, Zoghbi HY. 1999. Rett syndrome is caused by mutations in X-linked MECP2, encoding methyl-CpG-binding protein 2. Nat. Genet 23(2):185–88 [DOI] [PubMed] [Google Scholar]
  4. An J-Y, Lin K, Zhu L, Werling DM, Dong S, et al. 2018. Genome-wide de novo risk score implicates promoter variation in autism spectrum disorder. Science 362(6420):eaat6576. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Autism Spectr. Disord. Work. Group Psychiatr. Genom. Consort. 2017. Meta-analysis of GWAS of over 16,000 individuals with autism spectrum disorder highlights a novel locus at 10q24.32 and a significant overlap with schizophrenia. Mol. Autism 8:21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Bai D, Yip BHK, Windham GC, Sourander A, Francis R, et al. 2019. Association of genetic and environmental factors with autism in a 5-country cohort. JAMA Psychiatry 76(10):1035–43 [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Baker P, Piven J, Schwartz S, Patil S. 1994. Brief report: duplication of chromosome 15q11–13 in two individuals with autistic disorder. J. Autism Dev. Disord 24(4):529–35 [DOI] [PubMed] [Google Scholar]
  8. Bassell GJ, Warren ST. 2008. Fragile X syndrome: loss of local mRNA regulation alters synaptic development and function. Neuron 60(2):201–14 [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Bean Jaworski JL, Flynn T, Burnham N, Chittams JL et al. 2017. Rates of autism and potential risk factors in children with congenital heart defects. Congenit. Heart Dis 12(4):421–29 [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Bear MF, Huber KM, Warren ST. 2004. The mGluR theory of fragile X mental retardation. Trends Neurosci. 27(7):370–77 [DOI] [PubMed] [Google Scholar]
  11. Bhalla K, Phillips HA, Crawford J, McKenzie OK et al. 2004. The de novo chromosome 16 translocations of two patients with abnormal phenotypes (mental retardation and epilepsy) disrupt the A2BP1 gene. J. Hum. Genet 49:308–311 [DOI] [PubMed] [Google Scholar]
  12. Buckanovich RJ, Posner JB, Darnell RB. 1993. Nova, the paraneoplastic Ri antigen, is homologous to an RNA-binding protein and is specifically expressed in the developing motor system. Neuron 11(4):657–72 [DOI] [PubMed] [Google Scholar]
  13. Bundey S, Hardy C, Vickers S, Kilpatrick MW, Corbett JA. 1994. Duplication of the 15q11–13 region in a patient with autism, epilepsy and ataxia. Dev. Med. Child Neurol 36(8):736–42 [DOI] [PubMed] [Google Scholar]
  14. Carmona-Mora P, Walz K. 2010. Retinoic acid induced 1, RAI1: a dosage sensitive gene related to neurobehavioral alterations including autistic behavior. Curr. Genom 11(8):607–17 [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Ceman S, O’Donnell WT, Reed M, Patton S, Pohl J, Warren ST. 2003. Phosphorylation influences the translation state of FMRP-associated polyribosomes. Hum. Mol. Genet 12(24):3295–305 [DOI] [PubMed] [Google Scholar]
  16. Cha AE. 2019. The children of Donor H898. The Washington Post, Sept. 14. https://www.washingtonpost.com/health/the-children-of-donor-h898/2019/09/14/dcc191d8-86da-11e9-a491-25df61c78dc4_story.html [Google Scholar]
  17. Christie SB, Akins MR, Schwob JE, Fallon JR. 2009. The FXG: a presynaptic fragile X granule expressed in a subset of developing brain circuits. J. Neurosci 29(5):1514–24 [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Cirulli ET, Lasseigne BN, Petrovski S, Sapp PC, Dio PA, et al. 2015. Exome sequencing in amyotrophic lateral sclerosis identifies risk genes and pathways. Science 347(6229):1436–41 [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Conrad B, Antonarakis SE. 2007. Gene duplication: a drive for phenotypic diversity and cause of human disease. Annu. Rev. Genom. Hum. Genet 8:17–35 [DOI] [PubMed] [Google Scholar]
  20. Cook EH Jr., Courchesne RY, Cox NJ, Lord C, Gonen D, et al. 1998. Linkage-disequilibrium mapping of autistic disorder, with 15q11–13 markers. Am. J. Hum. Genet 62(5):1077–83 [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Cummings BB, Marshall JL, Tukiainen T, et al. 2017. Improving genetic diagnosis in Mendelian disease with transcriptome sequencing. Sci. Transl. Med 9(386):eaaI5209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Darnell JC, Fraser CE, Mostovetsky O, Stefani G, Jones TA, et al. 2005. Kissing complex RNAs mediate interaction between the fragile-X mental retardation protein KH2 domain and brain polyribosomes. Genes Dev. 19(8):903–18 [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Darnell JC, Jensen KB, Jin P, Brown V, Warren ST, Darnell RB. 2001. Fragile X mental retardation protein targets G quartet mRNAs important for neuronal function. Cell 107(4):489–99 [DOI] [PubMed] [Google Scholar]
  24. Darnell JC, Van Driesche SJ, Zhang C, Hung KYS, Mele A, et al. 2011. FMRP stalls ribosomal translocation on mRNAs linked to synaptic function and autism. Cell 146(2):247–61 [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Darnell RB. 2013. RNA protein interactions in neurons. Annu. Rev. Neurosci 36:243–70 [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Das Sharma S, Metz JB, Li H, Hobson BD, Hornstein N, et al. 2019. Widespread alterations in translation elongation in the brain of juvenile Fmr1 knockout mice. Cell Rep. 26(12):3313–22.e5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. De Rubeis S, He X, Goldberg AP, Poultney CS, Samocha K, et al. 2014. Synaptic, transcriptional and chromatin genes disrupted in autism. Nature 515(7526):209–15 [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Dhindsa RS, Goldstein DB. 2015. Genetic discoveries drive molecular analyses and targeted therapeutic options in the epilepsies. Curr. Neurol. Neurosci. Rep 15(10):70. [DOI] [PubMed] [Google Scholar]
  29. Ding D, Chen Z, Li K, Long Z, Ye W, et al. 2016. Identification of a de novo DYNC1H1 mutation via WES according to published guidelines. Sci. Rep 6:20423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Dredge BK, Darnell RB. 2003. Nova regulates GABAA receptor γ2 alternative splicing via a distal downstream UCAU-rich intronic splicing enhancer. Mol. Cell. Biol 23(13):4687–700 [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Dredge BK, Polydorides AD, Darnell RB. 2001. The splice of life: alternative splicing and neurological disease. Nat. Rev. Neurosci 2(1):43–50 [DOI] [PubMed] [Google Scholar]
  32. Dredge BK, Stefani G, Engelhard CC, Darnell RB. 2005. Nova autoregulation reveals dual functions in neuronal splicing. EMBO J. 24(8):1608–20 [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Durand CM, Betancur C, Boeckers TM, Bockmann J, Chaste P, et al. 2007. Mutations in the gene encoding the synaptic scaffolding protein SHANK3 are associated with autism spectrum disorders. Nat. Genet 39(1):25–27 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. El Fatimy R, Tremblay S, Dury AY, Solomon S, De Koninck P, et al. 2012. Fragile X mental retardation protein interacts with the RNA-binding protein Caprin1 in neuronal RiboNucleoProtein complexes. PLOS ONE 7(6):e39338 Correction. 2012. PLOS ONE 7(9). 10.1371/annotation/05374d07-34cf-483f-80f4-ec87374cbeb6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Eom T, Zhang C, Wang H, Lay K, Fak J, et al. 2013. NOVA-dependent regulation of cryptic NMD exons controls synaptic protein levels after seizure. eLife 2:e00178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Fang H, Bergmann EA, Arora K, Vacic V, Zody MC, et al. 2016. Indel variant analysis of short-read sequencing data with Scalpel. Nat. Protoc 11(12):2529–48 [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Fang P, Lev-Lehman E, Tsai TF, Matsuura T, Benton CS, et al. 1999. The spectrum of mutations in UBE3A causing Angelman syndrome. Hum. Mol. Genet 8(1):129–35 [DOI] [PubMed] [Google Scholar]
  38. Fogel BL, Wexler E, Wahnich A, Friedrich T, Vijayendran C, et al. 2012. RBFOX1 regulates both splicing and transcriptional networks in human neuronal development. Hum. Mol. Genet 21(19):4171–86 [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Fromer M, Pocklington AJ, Kavanagh DH, Williams HJ, Dwyer S, et al. 2014. De novo mutations in schizophrenia implicate synaptic networks. Nature 506(7487):179–84 [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Gambin T, Yuan B, Bi W, Liu P, Rosenfeld JA, et al. 2017. Identification of novel candidate disease genes from de novo exonic copy number variants. Genome Med. 9(1):83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Gandal MJ, Zhang P, Hadjimichael E, Walker RL, Chen C, et al. 2018. Transcriptome-wide isoform-level dysregulation in ASD, schizophrenia, and bipolar disorder. Science 362(6420):eaat8127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Glock C, Heumüller M, Schuman EM. 2017. mRNA transport & local translation in neurons. Curr. Opin. Neurobiol 45:169–77 [DOI] [PubMed] [Google Scholar]
  43. Gonatopoulos-Pournatzis T, Niibori R, Salter EW, Weatheritt RJ, Tsang B, et al. 2020. Autism-misregulated eIF4G microexons control synaptic translation and higher order cognitive functions. Mol. Cell 77:1176–92 [DOI] [PubMed] [Google Scholar]
  44. Grove J, Ripke S, Als TD, Mattheisen M, Walters RK, et al. 2019. Identification of common genetic risk variants for autism spectrum disorder. Nat. Genet 51(3):431–44 [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Hafner A-S, Donlin-Asp PG, Leitch B, Herzog E, Schuman EM. 2019. Local protein synthesis is a ubiquitous feature of neuronal pre- and postsynaptic compartments. Science 364(6441):eaau3644. [DOI] [PubMed] [Google Scholar]
  46. Hall JM, Lee MK, Newman B, Morrow JE, Anderson LA, et al. 1990. Linkage of early-onset familial breast cancer to chromosome 17q21. Science 250(4988):1684–89 [DOI] [PubMed] [Google Scholar]
  47. Hermann BP, Sager MA, Koscik RL, Young K, Nakamura K. 2017. Vascular, inflammatory, and metabolic factors associated with cognition in aging persons with chronic epilepsy. Epilepsia 58(11):e152–56 [DOI] [PubMed] [Google Scholar]
  48. Hippolyte L, Maillard AM, Rodriguez-Herreros B, Pain A, Martin-Brevet S, et al. 2016. The number of genomic copies at the 16p11.2 locus modulates language, verbal memory, and inhibition. Biol. Psychiatry 80(2):129–39 [DOI] [PubMed] [Google Scholar]
  49. Hodges E, Xuan Z, Balija V, Kramer M, Molla MN, et al. 2007. Genome-wide in situ exon capture for selective resequencing. Nat. Genet 39(12):1522–27 [DOI] [PubMed] [Google Scholar]
  50. Huber KM, Gallagher SM, Warren ST, Bear MF. 2002. Altered synaptic plasticity in a mouse model of fragile X mental retardation. PNAS 99(11):7746–50 [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Huber KM, Kayser MS, Bear MF. 2000. Role for rapid dendritic protein synthesis in hippocampal mGluR-dependent long-term depression. Science 288(5469):1254–57 [DOI] [PubMed] [Google Scholar]
  52. Iakoucheva LM, Muotri AR, Sebat J. 2019. Getting to the cores of autism. Cell 178(6):1287–98 [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Ince-Dunn G, Okano HJ, Jensen KB, Park W-Y, Zhong R, et al. 2012. Neuronal Elav-like (Hu) proteins regulate RNA splicing and abundance to control glutamate levels and neuronal excitability. Neuron 75(6):1067–80 [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Indjeian VB, Kingman GA, Jones FC, Guenther CA, Grimwood J, et al. 2016. Evolving new skeletal traits by cis-regulatory changes in bone morphogenetic proteins. Cell 164(1–2):45–56 [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Iossifov I, O’Roak BJ, Sanders SJ, Ronemus M, Krumm N, et al. 2014. The contribution of de novo coding mutations to autism spectrum disorder. Nature 515(7526):216–21 [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Iossifov I, Ronemus M, Levy D, Wang Z, Hakker I, et al. 2012. De novo gene disruptions in children on the autistic spectrum. Neuron 74(2):285–99 [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Irimia M, Weatheritt RJ, Ellis JD, Parikshak NN, Gonatopoulos-Pournatzis T, et al. 2014. A highly conserved program of neuronal microexons is misregulated in autistic brains. Cell 159(7):1511–23 [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Jensen KB, Dredge BK, Stefani G, Zhong R, Buckanovich RJ, et al. 2000. Nova-1 regulates neuron-specific alternative splicing and is essential for neuronal viability. Neuron 25(2):359–71 [DOI] [PubMed] [Google Scholar]
  59. Jiang YH, Ehlers MD. 2013. Modeling autism by SHANK gene mutations in mice. Neuron 78(10):8–27 [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Kang H, Schuman EM. 1996. A requirement for local protein synthesis in neurotrophin-induced hippocampal synaptic plasticity. Science 273(5280):1402–6 [DOI] [PubMed] [Google Scholar]
  61. Kasherman MA, Premarathne S, Burne TH, Wood SA, Piper M. 2020. The ubiquitin system: a regulatory hub for intellectual disability and autism spectrum disorder. Mol. Neurobiol 57:2179–93 [DOI] [PubMed] [Google Scholar]
  62. Khatri N, Man H-Y. 2019. The autism and Angelman syndrome protein Ube3A/E6AP: the gene, E3 ligase ubiquitination targets and neurobiological functions. Front. Mol. Neurosci 12:109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Kim H-G, Kim H-T, Leach NT, Lan F, Ullmann R, et al. 2012. Translocations disrupting PHF21A in the Potocki-Shaffer-syndrome region are associated with intellectual disability and craniofacial anomalies. Am. J. Hum. Genet 91(1):56–72 [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Kim TH, Tsang B, Vernon RM, Sonenberg N, Kay LE, Forman-Kay JD. 2019. Phospho-dependent phase separation of FMRP and CAPRIN1 recapitulates regulation of translation and deadenylation. Science 365:825–29 [DOI] [PubMed] [Google Scholar]
  65. Kishino T, Lalande M, Wagstaff J. 1997. UBE3A/E6-AP mutations cause Angelman syndrome. Nat. Genet 15(1):70–73 [DOI] [PubMed] [Google Scholar]
  66. Klein RJ, Zeiss C, Chew EY, Tsai J-Y, Sackler RS, et al. 2005. Complement factor H polymorphism in age-related macular degeneration. Science 308(5720):385–89 [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Korb E, Herre M, Zucker-Scharff I, Darnell RB, Allis CD. 2015. BET protein Brd4 activates transcription in neurons and BET inhibitor Jq1 blocks memory in mice. Nat. Neurosci 18(10):1464–73 [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Korb E, Herre M, Zucker-Scharff I, Gresack J, Allis CD, Darnell RB. 2017. Excess translation of epigenetic regulators contributes to fragile X syndrome and is alleviated by Brd4 inhibition. Cell 170(6):1209–23.e20 [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Krishnan A, Zhang R, Yao V, Theesfeld CL, Wong AK, et al. 2016. Genome-wide prediction and functional characterization of the genetic basis of autism spectrum disorder. Nat. Neurosci 19(11):1454–62 [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Krueger DD, Bear MF. 2011. Toward fulfilling the promise of molecular medicine in fragile X syndrome. Annu. Rev. Med 62:411–29 [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Krumm N, O’Roak BJ, Shendure J, Eichler EE. 2014. A de novo convergence of autism genetics and molecular neuroscience. Trends Neurosci. 37(2):95–105 [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Laggerbauer B, Ostareck D, Keidel EM, Ostareck-Lederer A, Fischer U. 2001. Evidence that fragile X mental retardation protein is a negative regulator of translation. Hum. Mol. Genet 10(4):329–38 [DOI] [PubMed] [Google Scholar]
  73. Lal D, Reinthaler EM, Altmüller J, Toliat MR et al. 2013a. RBFOX1 and RBFOX3 mutations in Rolandic epilepsy. PLOS ONE 8:e73323. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Lal D, Trucks H, Møller RS, Hjalgrim H, et al. 2013b. Rare exonic deletions of the RBFOX1 gene increase risk of idiopathic generalized epilepsy. Epilepsia 54:265–71 [DOI] [PubMed] [Google Scholar]
  75. Lee BH, Smith T, Paciorkowski AR. 2015. Autism spectrum disorder and epilepsy: disorders with a shared biology. Epilepsy Behav. 47:191–201 [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Lee C, Kang EY, Gandal MJ, Eskin E, Geschwind DH. 2019. Profiling allele-specific gene expression in brains from individuals with autism spectrum disorder reveals preferential minor allele usage. Nat. Neurosci 22(9):1521–32 [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Lee HY, Ge W-P, Huang W, He Y, Wang GX, et al. 2011. Bidirectional regulation of dendritic voltage-gated potassium channels by the fragile X mental retardation protein. Neuron 72(4):630–42 [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Lee J-A, Damianov A, Lin C-H, Fontes M, Parikshak NN, et al. 2016. Cytoplasmic Rbfox1 regulates the expression of synaptic and autism-related genes. Neuron 89(1):113–28 [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Licatalosi DD, Mele A, Fak JJ, Ule J, Kayikci M, et al. 2008. HITS-CLIP yields genome-wide insights into brain alternative RNA processing. Nature 456(7221):464–69 [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Liu X, Li YI, Pritchard JK. 2019. Trans effects on gene expression can drive omnigenic inheritance. Cell 177(4):1022–34.e6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Malhotra D, Sebat J. 2012. CNVs: harbingers of a rare variant revolution in psychiatric genetics. Cell 148(6):1223–41 [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Mannino EA, Miyawaki H, Santen G, Schrier Vergano SA. 2018. First data from a parent-reported registry of 81 individuals with Coffin-Siris syndrome: natural history and management recommendations. Am. J. Med. Genet . A 176(11):2250–58 [DOI] [PubMed] [Google Scholar]
  83. Martin CL, Duvall JA, Ilkin Y, Simon JS, Arreaza MG, et al. 2007. Cytogenetic and molecular characterization of A2BP1/FOX1 as a candidate gene for autism. Am. J. Med. Genet. B Neuropsychiatr. Genet 144B(7):869–76 [DOI] [PubMed] [Google Scholar]
  84. Matsuura T, Sutcliffe JS, Fang P, Galjaard RJ, Jiang YH, et al. 1997. De novo truncating mutations in E6-AP ubiquitin-protein ligase gene (UBE3A) in Angelman syndrome. Nat. Genet 15(1):74–77 [DOI] [PubMed] [Google Scholar]
  85. Matthews AM, Tarailo-Graovac M, Price EM, Blydt-Hansen I, Ghani A, et al. 2017. A de novo mosaic mutation in SPAST with two novel alternative alleles and chromosomal copy number variant in a boy with spastic paraplegia and autism spectrum disorder. Eur. J. Med. Genet 60(10):548–52 [DOI] [PubMed] [Google Scholar]
  86. McLean CY, Reno PL, Pollen AA, Bassan AI, Capellini TD, et al. 2011. Human-specific loss of regulatory DNA and the evolution of human-specific traits. Nature 471(7337):216–19 [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Mei Y, Monteiro P, Zhou Y, Kim J-A, Gao X, et al. 2016. Adult restoration of Shank3 expression rescues selective autistic-like phenotypes. Nature 530(7591):481–84 [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Miceli F, Soldovieri MV, Joshi N, Weckhuysen S, Cooper EC, Taglialatela M. 2014. KCNQ3-related disorders In GeneReviews, ed. Adam MP, Ardinger HH, Pagon RA, Wallace SE, Bean LJH, et al. Seattle, WA: Univ. Washington [Google Scholar]
  89. Miska NJ, Richter LM, Cary BA, Gjorgjieva J, Turrigiano GG. 2018. Sensory experience inversely regulates feedforward and feedback excitation-inhibition ratio in rodent visual cortex. eLife 7:e38846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Monteiro P, Feng G. 2017. SHANK proteins: roles at the synapse and in autism spectrum disorder. Nat. Rev. Neurosci 18(3):147–57 [DOI] [PubMed] [Google Scholar]
  91. Moretti P, Zoghbi HY. 2006. MeCP2 dysfunction in Rett syndrome and related disorders. Curr. Opin. Genet. Dev 16(3):276–81 [DOI] [PubMed] [Google Scholar]
  92. Muddashetty RS, Nalavadi VC, Gross C, Yao X, Xing L, et al. 2011. Reversible inhibition of PSD-95 mRNA translation by miR-125a, FMRP phosphorylation, and mGluR signaling. Mol. Cell 42(5):673–88 [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Musunuru K, Darnell RB. 2001. Paraneoplastic neurologic disease antigens: RNA-binding proteins and signaling proteins in neuronal degeneration. Annu. Rev. Neurosci 24:239–62 [DOI] [PubMed] [Google Scholar]
  94. Myers RA, Casals F, Gauthier J, Hamdan FF, Keebler J, et al. 2011. A population genetic approach to mapping neurological disorder genes using deep resequencing. PLOS Genet. 7(2):e1001318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Myrick LK, Nakamoto-Kinoshita M, Lindor NM, Kirmani S, Cheng X, Warren ST. 2014. Fragile X syndrome due to a missense mutation. Eur. J. Hum. Genet 22(10):1185–89 [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Niu M, Han Y, Dy ABC, Du J, Jin H, et al. 2017. Autism symptoms in fragile X syndrome. J. Child Neurol 32(10):903–9 [DOI] [PubMed] [Google Scholar]
  97. Novarino G, El-Fishawy P, Kayserili H, Meguid NA, Scott EM, et al. 2012. Mutations in BCKD-kinase lead to a potentially treatable form of autism with epilepsy. Science 338(6105):394–97 [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. O’Roak BJ, Vives L, Girirajan S, Karakoc E, Krumm N, et al. 2012. Sporadic autism exomes reveal a highly interconnected protein network of de novo mutations. Nature 485(7397):246–50 [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Ostroff LE, Fiala JC, Allwardt B, Harris KM. 2002. Polyribosomes redistribute from dendritic shafts into spines with enlarged synapses during LTP in developing rat hippocampal slices. Neuron 35(3):535–45 [DOI] [PubMed] [Google Scholar]
  100. Packer A 2016. Neocortical neurogenesis and the etiology of autism spectrum disorder. Neurosci. Biobehav. Rev 64:185–95 [DOI] [PubMed] [Google Scholar]
  101. Pardiñas AF, Holmans P, Pocklington AJ, Escott-Price V, Ripke S, et al. 2018. Common schizophrenia alleles are enriched in mutation-intolerant genes and in regions under strong background selection. Nat. Genet 50(3):381–89 [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Parras A, Anta H, Santos-Galindo M, Swarup V, Elorza A, et al. 2018. Autism-like phenotype and risk gene mRNA deadenylation by CPEB4 mis-splicing. Nature 560(7719):441–46 [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Petrovski S, Wang Q, Heinzen EL, Allen AS, Goldstein DB. 2013. Genic intolerance to functional variation and the interpretation of personal genomes. PLOS Genet. 9(8):e1003709. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Phan AT, Kuryavyi V, Darnell JC, Serganov A, Majumdar A, et al. 2011. Structure-function studies of FMRP RGG peptide recognition of an RNA duplex-quadruplex junction. Nat. Struct. Mol. Biol 18(7):796–804 [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Pizzo L, Jensen M, Polyak A, Rosenfeld JA, Mannik K, et al. 2019. Rare variants in the genetic background modulate cognitive and developmental phenotypes in individuals carrying disease-associated variants. Genet. Med 21(4):816–25 [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Porter RS, Murata-Nakamura Y, Nagasu H, Kim H-G, Iwase S. 2018. Transcriptome analysis revealed impaired cAMP responsiveness in PHF21A-deficient human cells. Neuroscience 370:170–80 [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Potocki L, Chen KS, Park SS, Osterholm DE, Withers MA, et al. 2000. Molecular mechanism for duplication 17p11.2—the homologous recombination reciprocal of the Smith-Magenis microdeletion. Nat. Genet 24(1):84–87 [DOI] [PubMed] [Google Scholar]
  108. Quesnel-Vallières T, Weatheritt RJ, Cordes SP, Blencowe BJ. 2019. Autism spectrum disorder: insights into convergent mechanisms from transcriptomics. Nat. Rev. Genet 20(1):51–63 [DOI] [PubMed] [Google Scholar]
  109. Ramaswami G, Geschwind DH. 2018. Genetics of autism spectrum disorder. Handb. Clin. Neurol 147:321–29 [DOI] [PubMed] [Google Scholar]
  110. Rosenfeld JA, Coppinger J, Bejjani BA, Girirajan S, Eichler EE, et al. 2010. Speech delays and behavioral problems are the predominant features in individuals with developmental delays and 16p11.2 microdeletions and microduplications. J. Neurodev. Disord 2(1):26–38 [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Rubenstein JL, Merzenich MM. 2003. Model of autism: increased ratio of excitation/inhibition in key neural systems. Genes Brain Behav. 2(5):255–67 [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Ruzzo EK, Pérez-Cano L, Jung J-Y, Wang L-K, Kashef-Haghighi D, et al. 2019. Inherited and de novo genetic risk for autism impacts shared networks. Cell 178(4):850–66.e26 [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Saito Y, Yuan Y, Zucker-Scharff I, Fak JJ, Jereb S, et al. 2019. Differential NOVA2-mediated splicing in excitatory and inhibitory neurons regulates cortical development and cerebellar function. Neuron 101(4):707–20.e5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Sanders SJ, He X, Willsey AJ, Ercan-Sencicek AG, Samocha KE, et al. 2015. Insights into autism spectrum disorder genomic architecture and biology from 71 risk loci. Neuron 87(6):1215–33 [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Sanders SJ, Murtha MT, Gupta AR, Murdoch JD, Raubeson MJ, et al. 2012. De novo mutations revealed by whole-exome sequencing are strongly associated with autism. Nature 485(7397):237–41 [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Sands TT, Miceli F, Lesca G, Beck AE, Sadleir LG, et al. 2019. Autism and developmental disability caused by KCNQ3 gain-of-function variants. Ann. Neurol 86(2):181–92 [DOI] [PubMed] [Google Scholar]
  117. Santen GWE, Aten E, Sun Y, Almomani R, Gilissen C, et al. 2012. Mutations in SWI/SNF chromatin remodeling complex gene ARID1B cause Coffin-Siris syndrome. Nat. Genet 44(4):379–80 [DOI] [PubMed] [Google Scholar]
  118. Santoro MR, Bray SM, Warren ST. 2012. Molecular mechanisms of fragile X syndrome: a twenty-year perspective. Annu. Rev. Pathol 7:219–45 [DOI] [PubMed] [Google Scholar]
  119. Satterstrom FK, Kosmicki JA, Wang J, Breen MS, De Rubeis S, et al. 2020. Large-scale exome sequencing study implicates both developmental and functional changes in the neurobiology of autism. Cell 180:568–84.e23 [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Scheckel C, Darnell RB. 2015. Microexons–tiny but mighty. EMBO J. 34(3):273–74 [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Schinzel AA, Brecevic L, Bernasconi F, Binkert F, Berthet F, et al. 1994. Intrachromosomal triplication of 15q11-q13. J. Med. Genet 31(10):798–803 [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Sebat J, Lakshmi B, Malhotra D, Troge J, Lese-Martin C, et al. 2007. Strong association of de novo copy number mutations with autism. Science 316(5823):445–49 [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Shcheglovitov A, Shcheglovitova O, Yazawa M, Portmann T, Shu R, et al. 2013. SHANK3 and IGF1 restore synaptic deficits in neurons from 22q13 deletion syndrome patients. Nature 503(7475):267–71 [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Sheng M, Hoogenraad CC. 2007. The postsynaptic architecture of excitatory synapses: a more quantitative view. Annu. Rev. Biochem 76:823–47 [DOI] [PubMed] [Google Scholar]
  125. Speed HE, Kouser M, Xuan Z, Liu S, Duong A, Powell CM. 2019. Apparent genetic rescue of adult Shank3 exon 21 insertion mutation mice tempered by appropriate control experiments. eNeuro 6(5):ENEURO.0317–19.2019 [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Speed HE, Kouser M, Xuan Z, Reimers JM, Ochoa CF, et al. 2015. Autism-associated insertion mutation (InsG) of Shank3 exon 21 causes impaired synaptic transmission and behavioral deficits. J. Neurosci 35(26):9648–65 [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Sragovich S, Ziv Y, Vaisvaser S, Shomron N, Hendler T, Gozes I. 2019. The autism-mutated ADNP plays a key role in stress response. Transl. Psychiatry 9(1):235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Stefani G, Fraser CE, Darnell JC, Darnell RB. 2004. Fragile X mental retardation protein is associated with translating polyribosomes in neuronal cells. J. Neurosci 24(33):7272–76 [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Sutcliffe JS, Jiang YH, Galijaard RJ, Matsuura T, Fang P, et al. 1997. The E6-Ap ubiquitin-protein ligase (UBE3A) gene is localized within a narrowed Angelman syndrome critical region. Genome Res. 7(4):368–77 [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Tatavarty V, Torrado Pacheco A, Lin H, Miska NJ, Hengen KB, et al. 2018. Autism-associated Shank3 is essential for homeostatic plasticity and neuronal circuit stability. bioRxiv 365445 10.1101/365445 [DOI] [Google Scholar]
  131. Tick B, Bolton P, Happé F, Rutter M, Rijsdijk F. 2016. Heritability of autism spectrum disorders: a meta-analysis of twin studies. J. Child Psychol. Psychiatry 57(5):585–95 [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Turner TN, Coe BP, Dickel DE, Hoekzema K, Nelson BJ, et al. 2017. Genomic patterns of de novo mutation in simplex autism. Cell 171(3):710–22.e12 [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Turrigiano GG, Leslie KR, Desai NS, Rutherford LC, Nelson SB. 1998. Activity-dependent scaling of quantal amplitude in neocortical neurons. Nature 391(6670):892–96 [DOI] [PubMed] [Google Scholar]
  134. Turrigiano GG, Nelson SB. 2004. Homeostatic plasticity in the developing nervous system. Nat. Rev. Neurosci 5(2):97–107 [DOI] [PubMed] [Google Scholar]
  135. Tushev G, Glock C, Heumüller M, Biever A, Jovanovic M, Schuman EM. 2018. Alternative 3′ UTRs modify the localization, regulatory potential, stability, and plasticity of mRNAs in neuronal compartments. Neuron 98(3):495–511.e6 [DOI] [PubMed] [Google Scholar]
  136. Udagawa T, Farny NG, Jakovcevski M, Kaphzan H, Alarcon JM, et al. 2013. Genetic and acute CPEB1 depletion ameliorate fragile X pathophysiology. Nat. Med 19(11):1473–77 [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Ule J, Hwang H-W, Darnell RB. 2018. The future of Cross-Linking and Immunoprecipitation (CLIP). Cold Spring Harb. Perspect. Biol 10(8):a032243. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Ule J, Jensen KB, Ruggiu M, Mele A, Ule A, Darnell RB. 2003. CLIP identifies Nova-regulated RNA networks in the brain. Science 302(5648):1212–15 [DOI] [PubMed] [Google Scholar]
  139. Ule J, Stefani G, Mele A, Ruggiu M, Wang X, et al. 2006. An RNA map predicting Nova-dependent splicing regulation. Nature 444(7119):580–86 [DOI] [PubMed] [Google Scholar]
  140. Ule J, Ule A, Spencer J, Williams A, Hu J-S, et al. 2005. Nova regulates brain-specific splicing to shape the synapse. Nat. Genet 37(8):844–52 [DOI] [PubMed] [Google Scholar]
  141. Van Driesche SJ, Sawicka K, Zhang C, Hung SKY, Park CY, et al. 2019. FMRP binding to a ranked subset of long genes is revealed by coupled CLIP and TRAP in specific neuronal cell types. bioRxiv 762500 10.1101/762500 [DOI] [Google Scholar]
  142. Van Esch H, Bauters M, Ignatius J, Jansen M, Raynaud M, et al. 2005. Duplication of the MECP2 region is a frequent cause of severe mental retardation and progressive neurological symptoms in males. Am. J. Hum. Genet 77(3):442–53 [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Vasilyev N, Polonskaia A, Darnell JC, Darnell RB, Patel DJ, Serganov A. 2015. Crystal structure reveals specific recognition of a G-quadruplex RNA by a β-turn in the RGG motif of FMRP. PNAS 112(39):E5391–400 [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Voineagu I, Wang X, Johnston P, Lowe JK, Tian Y, et al. 2011. Transcriptomic analysis of autistic brain reveals convergent molecular pathology. Nature 474(7351):380–84 [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Wagnon JL, Briese M, Sun W, Mahaffey CL, Curk T, et al. 2012. CELF4 regulates translation and local abundance of a vast set of mRNAs, including genes associated with regulation of synaptic function. PLOS Genet. 8(11):e1003067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Walsh CA, Morrow EM, Rubenstein JL. 2008. Autism and brain development. Cell 135:396–400 [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Weiss LA, Arking DE, Gene Discov. Proj. Johns Hopkins Autism Consort., Daly MJ, Chakravarti A. 2009. A genome-wide linkage and association scan reveals novel loci for autism. Nature 461(7265):802–8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Werling DM, Brand H, An J-Y, Stone MR, Zhu L, et al. 2018. An analytical framework for whole-genome sequence association studies and its implications for autism spectrum disorder. Nat. Genet 50(5):727–36 [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Weyn-Vanhentenryck SM, Mele A, Yan Q, Sun S, Farny N, et al. 2014. HITS-CLIP and integrative modeling define the Rbfox splicing-regulatory network linked to brain development and autism. Cell Rep. 6(6):1139–52 [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Wilson HL, Wong ACC, Shaw SR, Tse W-Y, Stapleton GA, et al. 2003. Molecular characterisation of the 22q13 deletion syndrome supports the role of haploinsufficiency of SHANK3/PROSAP2 in the major neurological symptoms. J. Med. Genet 40(8):575–84 [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Wiśniowiecka-Kowaednik B, Nesteruk M, Peters SU, Xia Z, Cooper ML, et al. 2010. Intragenic rearrangements in NRXN1 in three families with autism spectrum disorder, developmental delay, and speech delay. Am. J. Med. Genet. B Neuropsychiatr. Genet 153B(5):983–93 [DOI] [PubMed] [Google Scholar]
  152. Xu X, Hu Y, Xiong Y, Li Z, Wang W, et al. 2016. Association of microtubule dynamics with chronic epilepsy. Mol. Neurobiol 53(7):5013–24 [DOI] [PubMed] [Google Scholar]
  153. Yang YY, Yin GL, Darnell RB. 1998. The neuronal RNA-binding protein Nova-2 is implicated as the autoantigen targeted in POMA patients with dementia. PNAS 95(22):13254–59 [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Yuen RKC, Merico D, Bookman M, Howe JL, Thiruvahindrapuram B, et al. 2017. Whole genome sequencing resource identifies 18 new candidate genes for autism spectrum disorder. Nat. Neurosci 20(4):602–11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Zang JB, Nosyreva ED, Spencer CM, Volk LJ, Musunuru K, et al. 2009. A mouse model of the human fragile X syndrome I304N mutation. PLOS Genet. 5(12):e1000758. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Zhang C, Frias MA, Mele A, Ruggiu M, Eom T, et al. 2010. Integrative modeling defines the Nova splicing-regulatory network and its combinatorial controls. Science 329(5990):439–43 [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Zhang C, Zhang Z, Castle J, Sun S, Johnson J, et al. 2008. Defining the regulatory network of the tissue-specific splicing factors Fox-1 and Fox-2. Genes Dev. 22(18):2550–63 [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Zhang J, Hou L, Klann E, Nelson DL. 2009. Altered hippocampal synaptic plasticity in the FMR1 gene family knockout mouse models. J. Neurophysiol 101(5):2572–80 [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Zhou J, Park CY, Theesfeld CL, Wong AK, Yuan Y, et al. 2019. Whole-genome deep-learning analysis identifies contribution of noncoding mutations to autism risk. Nat. Genet 51(6):973–80 [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Zuk O, Schaffner SF, Samocha K, Do R, Hechter E, et al. 2014. Searching for missing heritability: designing rare variant association studies. PNAS 111(4):E455–64 [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Zweier M, Begemann A, McWalter K, Cho MT, Abela L, et al. 2019. Spatially clustering de novo variants in CYFIP2, encoding the cytoplasmic FMRP interacting protein 2, cause intellectual disability and seizures. Eur. J. Hum. Genet 27(5):747–59 [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES