% 209850

AUTISM


Alternative titles; symbols

AUTISTIC DISORDER


Other entities represented in this entry:

AUTISM, SUSCEPTIBILITY TO, 1, INCLUDED; AUTS1, INCLUDED
AUTISM SPECTRUM DISORDER, INCLUDED; ASD, INCLUDED

Cytogenetic location: 7q22     Genomic coordinates (GRCh38): 7:98,400,000-107,800,000


Gene-Phenotype Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
7q22 {Autism susceptibility 1} 209850 Mu, IC 2
Clinical Synopsis
   
Phenotypic Series

INHERITANCE
- Isolated cases
- Multifactorial
NEUROLOGIC
Central Nervous System
- Mental retardation in 75%
- Seizures in 15-30%
- EEG abnormalities in 20-50%
Behavioral Psychiatric Manifestations
- Impaired social interactions
- Impaired use of nonverbal behaviors, such as eye-to-eye gaze, facial expression, body posture, and gestures
- Impaired ability to form peer relationships
- Impaired language development
- Lack of spontaneous play
- Restrictive behavior, interests, and activities
- Stereotyped, repetitive behavior
- Inflexible adherence to routines or rituals
LABORATORY ABNORMALITIES
- Increased serum serotonin in 25%
MISCELLANEOUS
- Onset by 3 years of age
- Male to female ratio 4:1
- Occurs in 2-5 per 10,000 individuals
- Genetic heterogeneity (see, e.g., 609378, 608636, 608049, 300425, 300495, 300496)
- Associated with tuberous sclerosis (191100)
- Associated with untreated phenylketonuria (261600)
- Associated with Fragile X syndrome (309550)
Autism, susceptiblity to - PS209850 - 26 Entries
Location Phenotype Inheritance Phenotype
mapping key
Phenotype
MIM number
Gene/Locus Gene/Locus
MIM number
1q41-q42 {Autism susceptibility 11} 2 610836 AUTS11 610836
2q {Autism susceptibility 5} AD, AR 2 606053 AUTS5 606053
3q24 ?{Autism susceptibility 16} 3 613410 SLC9A9 608396
3q25-q27 {Autism susceptibility 8} Mu, IC 2 607373 AUTS8 607373
4q23 {Autism, susceptibility to, 19} 3 615091 EIF4E 133440
7q22 {Autism susceptibility 1} Mu, IC 2 209850 AUTS1 209850
7q31 {Autism, susceptibility to, 9} 2 611015 AUTS9 611015
7q35-q36 {Autism susceptibility 15} 3 612100 CNTNAP2 604569
7q36 {Autism, susceptibility to, 10} 2 611016 AUTS10 611016
11q13.3-q13.4 {Autism susceptibility 17} 3 613436 SHANK2 603290
12q14.2 {Autism susceptibility 13} 2 610908 AUTS13 610908
13q14.2-q14.1 {Autism susceptibility 3} Mu, IC 2 608049 AUTS3 608049
14q11.2 {Autism, susceptibility to, 18} AD 3 615032 CHD8 610528
15q11 {Autism susceptibility 4} AD 2 608636 AUTS4 608636
16p11.2 Chromosome 16p11.2 deletion syndrome, 593kb 4 611913 DEL16p11.2 611913
16p11.2 {Autism susceptibility 14A} 2 611913 DEL16p11.2 611913
17q11 {Autism susceptibility 6} 2 609378 AUTS6 609378
17q21 {Autism susceptibility 7} 2 610676 AUTS7 610676
21p13-q11 {Autism susceptibility 12} 2 610838 AUTS12 610838
Xp22.32-p22.31 Mental retardation, X-linked Mu, IC, XL 3 300495 NLGN4 300427
Xp22.32-p22.31 {Autism susceptibility, X-linked 2} Mu, IC, XL 3 300495 NLGN4 300427
Xp22.11 {Autism, susceptibility to, X-linked 4} XLR 3 300830 PTCHD1 300828
Xq13.1 {Autism susceptibility, X-linked 1} Mu, IC, XL 3 300425 NLGN3 300336
Xq28 {Autism susceptibility, X-linked 3} Mu, IC, XL 3 300496 MECP2 300005
Xq28 {Autism, susceptibility to, X-linked 5} 3 300847 RPL10 312173
Xq28 {Autism, susceptibility to, X-linked 6} XLR 3 300872 TMLHE 300777

TEXT

Description

Autism, the prototypic pervasive developmental disorder (PDD), is usually apparent by 3 years of age. It is characterized by a triad of limited or absent verbal communication, a lack of reciprocal social interaction or responsiveness, and restricted, stereotypic, and ritualized patterns of interests and behavior (Bailey et al., 1996; Risch et al., 1999). 'Autism spectrum disorder,' sometimes referred to as ASD, is a broader phenotype encompassing the less severe disorders Asperger syndrome (see ASPG1; 608638) and pervasive developmental disorder, not otherwise specified (PDD-NOS). 'Broad autism phenotype' includes individuals with some symptoms of autism, but who do not meet the full criteria for autism or other disorders. Mental retardation coexists in approximately two-thirds of individuals with ASD, except for Asperger syndrome, in which mental retardation is conspicuously absent (Jones et al., 2008). Genetic studies in autism often include family members with these less stringent diagnoses (Schellenberg et al., 2006).

Levy et al. (2009) provided a general review of autism and autism spectrum disorder, including epidemiology, characteristics of the disorder, diagnosis, neurobiologic hypotheses for the etiology, genetics, and treatment options.

Genetic Heterogeneity of Autism

Autism is considered to be a complex multifactorial disorder involving many genes. Accordingly, several loci have been identified, some or all of which may contribute to the phenotype. Included in this entry is AUTS1, which has been mapped to chromosome 7q22.

Other susceptibility loci include AUTS3 (608049), which maps to chromosome 13q14; AUTS4 (608636), which maps to chromosome 15q11; AUTS5 (606053), which maps to chromosome 2q; AUTS6 (609378), which maps to chromosome 17q11; AUTS7 (610676), which maps to chromosome 17q21; AUTS8 (607373), which maps to chromosome 3q25-q27; AUTS9 (611015), which maps to chromosome 7q31; AUTS10 (611016), which maps to chromosome 7q36; AUTS11 (610836), which maps to chromosome 1q41; AUTS12 (610838), which maps to chromosome 21p13-q11; AUTS13 (610908), which maps to chromosome 12q14; AUTS14A (611913), which has been found in patients with a deletion of a region of 16p11.2; AUTS14B (614671), which has been found in patients with a duplication of a region of 16p11.2; AUTS15 (612100), associated with mutation in the CNTNAP2 gene (604569) on chromosome 7q35-q36; AUTS16 (613410), associated with mutation in the SLC9A9 gene (608396) on chromosome 3q24; AUTS17 (613436), associated with mutation in the SHANK2 gene (603290) on chromosome 11q13; and AUTS18 (615032), associated with mutation in the CHD8 gene (610528). (NOTE: the symbol 'AUTS2' has been used to refer to a gene on chromosome 7q11 (KIAA0442; 607270) and therefore is not used as a part of this autism locus series.)

There are several X-linked forms of autism susceptibility: AUTSX1 (300425), associated with mutations in the NLGN3 gene (300336); AUTSX2 (300495), associated with mutations in NLGN4 (300427); AUTSX3 (300496), associated with mutations in MECP2 (300005); AUTSX4 (300830), associated with variation in the region on chromosome Xp22.11 containing the PTCHD1 gene (300828); AUTSX5 (300847), associated with mutations in the RPL10 gene (312173); and AUTSX6 (300872), associated with mutation in the TMLHE gene (300777).

Folstein and Rosen-Sheidley (2001) reviewed the genetics of autism.


Clinical Features

The DSM-IV (American Psychiatric Association, 1994) specifies several diagnostic criteria for autism. In general, patients with autism exhibit qualitative impairment in social interaction, as manifest by impairment in the use of nonverbal behaviors such as eye-to-eye gaze, facial expression, body postures, and gestures, failure to develop appropriate peer relationships, and lack of social sharing or reciprocity. Patients have impairments in communication, such as a delay in, or total lack of, the development of spoken language. In patients who do develop adequate speech, there remains a marked impairment in the ability to initiate or sustain a conversation, as well as stereotyped or idiosyncratic use of language. Patients also exhibit restricted, repetitive and stereotyped patterns of behavior, interests, and activities, including abnormal preoccupation with certain activities and inflexible adherence to routines or rituals.

In his pioneer description of infantile autism, Kanner (1943) defined the disorder as 'an innate inability to form the usual, biologically provided affective contact with people.' Kanner (1943) noted that in most cases the child's behavior was abnormal from early infancy, and he suggested the presence of an inborn, presumably genetic, defect.

In a review, Smalley (1997) stated that mental retardation is said to be present in approximately 75% of cases of autism, seizures in 15 to 30% of cases, and electroencephalographic abnormalities in 20 to 50% of cases. In addition, approximately 15 to 37% of cases of autism have a comorbid medical condition, including 5 to 14% with a known genetic disorder or chromosomal anomaly. The 4 most common associations include fragile X syndrome (300624), tuberous sclerosis (see 191100), 15q duplications (AUTS4; 608636), and untreated phenylketonuria (PKU; 261600). Significant associations at a phenotypic level may reflect disruptions in a common neurobiologic pathway, common susceptibility genes, or genes in linkage disequilibrium.

The autism spectrum disorder shows a striking sex bias, with a male:female ratio of idiopathic autism estimated at 4-10:1, and with an increase in this ratio as the intelligence of the affected individuals increases (Folstein and Rosen-Sheidley, 2001).

Lainhart et al. (2002) stated that approximately 20% of children with autism appear to have relatively normal development during the first 12 to 24 months of life. This period of relative normalcy gradually or suddenly ends and is followed by a period of regression, characterized most prominently by a significant loss of language skills, after which the full autism syndrome becomes evident.

Rarely, children with autism may exhibit hyperlexia, or precocious reading (238350). Among a group of 66 children with pervasive developmental disorder, Burd et al. (1985) identified 4 with hyperlexia.

Cohen et al. (2005) discussed several genetic disorders consistently associated with autism, including fragile X syndrome, tuberous sclerosis, Angelman syndrome (105830), Down syndrome (190685), Sanfilippo syndrome (252900), Rett syndrome (312750) and other MECP2-related disorders, phenylketonuria, Smith-Magenis syndrome (SMS; 182290), 22q13 deletion syndrome (606232), Cohen syndrome (COH1; 216550), adenylosuccinate lyase deficiency (103050), and Smith-Lemli-Opitz syndrome (SLOS; 270400).

Miles et al. (2008) presented an expert-derived consensus measure of dysmorphic features often observed in patients with autism. The goal was to enable clinicians not trained in dysmorphology to use this classification system to identify and further subphenotype patients with autism. The measure includes 12 body areas that can be scored to arrive at a determination of dysmorphic or nondysmorphic. The body areas include stature, hair growth pattern, ear structure and placement, nose size, facial structure, philtrum, mouth and lips, teeth, hands, fingers and thumbs, nails, and feet. The model performed with 81 to 82% sensitivity and 95 to 99% specificity.


Inheritance

Folstein and Rutter (1977) reported that there had been no recorded cases of an autistic child having an overtly autistic parent; however, they noted that autistic persons rarely marry and rarely give birth. Folstein and Rutter (1977) stated that about 2% of sibs are affected, and that speech delay is common in the sibships containing autistic children. In a study of 21 same-sex twin pairs, 11 monozygotic (MZ) and 10 dizygotic (DZ), in which at least 1 had infantile autism, Folstein and Rutter (1977) found 36% concordance among the MZ twins and no concordance among the DZ twins. The concordance for cognitive abnormalities was 82% for MZ pairs and 10% for DZ pairs. In 12 of the 17 pairs discordant for autism, a biologic hazard liable to cause brain damage was identified. The authors concluded that brain injury in infancy may lead to autism on its own or in combination with a genetic predisposition. An inheritance pattern was not suggested.

In 40 pairs of twins, Ritvo et al. (1985) found a concordance rate for autism of 23.5% in dizygotic twins (4 of 17 pairs) and 95.7% in monozygotic twins (22 of 23 pairs). Ritvo et al. (1985) ascertained 46 families with 2 (n = 41) or 3 (n = 5) sibs with autism. Classic segregation analysis yielded a maximum likelihood estimate of the segregation ratio of 0.19 +/- 0.07, a value significantly different from 0.50 expected of an autosomal dominant trait and not significantly different from 0.25 expected of a recessive trait. The authors rejected a polygenic threshold model and suggested autosomal recessive inheritance.

Using the Utah Genealogical Database, Jorde et al. (1990) determined kinship for all possible pairs of autistic subjects. The average kinship coefficient for autistic subjects and controls showed a strong tendency for autism to cluster in families. However, the familial aggregation was confined exclusively to sib pairs and did not extend to more distant relatives. The authors concluded that the findings excluded recessive inheritance, since the autosomal recessive hypothesis would predict several first-cousin pairs, of which none were found. The rapid fall off in risk to relatives, as well as the sib risk of 4.5%, was consistent with multifactorial causation.

By analysis of 99 autistic probands and their families, Bolton et al. (1994) found an increased familial risk for both autism and more broadly defined pervasive developmental disorders in sibs, 2.9% and 2.9%, respectively, which is about 75 times higher than the risk in the general population.

In 27 same-sex pairs of monozygotic twins and 20 dizygotic twins, Bailey et al. (1995) found that 60% of monozygotic pairs were concordant for autism compared to 0% of dizygotic pairs. When they considered a broader spectrum of related cognitive or social abnormalities, 92% of monozygotic pairs were concordant compared to 10% of dizygotic pairs. The high concordance in monozygotes indicated a high degree of genetic control, and the rapid fall off of concordance in dizygotes suggested to Bailey et al. (1995) a multilocus, epistatic model. In the nonconcordant monozygotic pairs, there was a significantly higher incidence of obstetric complications, which the authors attributed to prenatal developmental anomalies, as evidenced by the very high incidence of minor congenital anomalies in the affected twins. They also reported an association of autism with increased head circumference.

In a sample of families selected because each had exactly 2 affected sibs, Greenberg et al. (2001) observed a remarkably high proportion of affected twin pairs, both MZ and DZ. Of 166 affected sib pairs, 30 (12 MZ, 17 DZ, and 1 of unknown zygosity) were twin pairs. Deviation from expected values was statistically significant; in a similarly ascertained sample of individuals with type I diabetes (222100), there was no deviation from expected values. Greenberg et al. (2001) noted that to ascribe the excess of twins with autism solely to ascertainment bias would require very large ascertainment factors; e.g., affected twin pairs would need to be approximately 10 times more likely to be ascertained than affected nontwin sib pairs. In the extreme situation of 'complete stoppage,' a form of ascertainment bias in which parents stop having children after the birth of their first affected child, the only families to have an affected sib pair would be those with an affected twin pair, or affected triplets. The authors suggested that risk factors related to twinning or to fetal development or other factors, genetic or nongenetic, in the parents may contribute to autism. Hallmayer et al. (2002) presented information refuting the suggestion that the twinning process itself is an important risk factor in the development of autism.

Silverman et al. (2002) analyzed 3 autistic symptom domains, social interaction, communication, and repetitive behaviors, and variability in the presence and emergence of useful phrase speech in 212 multiply affected sibships with autism. They found that the variance within sibships was reduced for the repetitive behavior domain and for delays in and the presence of useful phrase speech. These features and the nonverbal communication subdomain provided evidence of familiality when only the diagnosis of autism was considered for defining multiply affected sibships.

Kolevzon et al. (2004) studied specific features of autism for decreased variance in 16 families with monozygotic twins concordant for autism. Using regression analysis, they demonstrated significant aggregation of symptoms in monozygotic twins for 2 autistic symptom domains: impairment in communication and in social interaction. Kolevzon et al. (2004) stated that selecting probands according to specific features known to show reduced variance within families may provide more homogeneous samples for genetic analysis.

Awadalla et al. (2010) hypothesized that deleterious de novo mutations may play a role in cases of ASD and schizophrenia (181500), 2 etiologically heterogeneous disorders with significantly reduced reproductive fitness. Awadalla et al. (2010) presented a direct measure of the de novo mutation rate (mu) and selective constraints from de novo mutations estimated from a deep resequencing dataset generated from a large cohort of ASD and schizophrenia cases (n = 285) and population control individuals (n = 285) with available parental DNA. A survey of approximately 430 Mb of DNA from 401 synapse-expressed genes across all cases and 25 Mb of DNA in controls found 28 candidate de novo mutations, 13 of which were cell line artifacts. Awadalla et al. (2010) calculated a direct neutral mutation rate (1.36 x 10(-8)) that was similar to previous indirect estimates, but they observed a significant excess of potentially deleterious de novo mutations in ASD and schizophrenia individuals. Awadalla et al. (2010) concluded that their results emphasized the importance of de novo mutations as genetic mechanisms in ASD and schizophrenia and the limitations of using DNA from archived cell lines to identify functional variants.

Sandin et al. (2014) examined the familial risk of autism in a population-based cohort of 2,049,973 Swedish children born from 1982 to 2006. They identified 37,570 twin pairs; 2,642,064 full-sib pairs; 432,281 maternal and 445,531 paternal half-sib pairs; and 5,799,875 cousin pairs. Diagnoses of ASD to December 31, 2009 were ascertained. Exposure refers to the presence or absence of autism in a sib. In the sample, 14,516 children were diagnosed with ASD, of whom 5,689 had autistic disorder. The relative recurrence risk (RRR) and rate per 100,000 person-years for ASD among monozygotic twins was estimated to be 153.0 (95% CI, 56.7-412.8; rate, 6,274 for exposed vs 27 for unexposed); for dizygotic twins, 8.2 (95% CI, 3.7-18.1; rate, 805 for exposed vs 55 for unexposed); for full sibs, 10.3 (95% CI, 9.4-11.3; rate, 829 for exposed vs 49 for unexposed); for maternal half sibs, 3.3 (95% CI, 2.6-4.2; rate, 492 for exposed vs 94 for unexposed); for paternal half sibs, 2.9 (95% CI, 2.2-3.7; rate, 371 for exposed vs 85 for unexposed); and for cousins, 2.0 (95% CI, 1.8-2.2; rate, 155 for exposed vs 49 for unexposed). The RRR pattern was similar for autistic disorder but of slightly higher magnitude. Sandin et al. (2014) found support for a disease etiology including only additive genetic and nonshared environmental effects. The ASD heritability was estimated to be 0.50 (95% CI, 0.45-0.56) and the autistic disorder heritability was estimated to 0.54 (95% CI, 0.44-0.64). Sandin et al. (2014) concluded that among children in Sweden, the individual risk of ASD and autistic disorder increased with increasing genetic relatedness.

Iossifov et al. (2014) applied whole-exome sequencing to more than 2,500 simplex families each having a child with an autistic spectrum disorder. By comparing affected to unaffected sibs, Iossifov et al. (2014) showed that 13% of de novo missense mutations and 43% of de novo likely gene-disrupting mutations contribute to 12% and 9% of diagnoses, respectively. Including CNVs, coding de novo mutations contribute to about 30% of all simplex and 45% of female diagnoses.


Mapping

AUTS1 Locus on Chromosome 7q22

By analyzing 125 autistic sib pairs, the International Molecular Genetic Study of Autism Consortium (2001) found a maximum multipoint lod score of 2.15 at marker D7S477 on chromosome 7q22, whereas analysis of 153 sib pairs generated a maximum multipoint lod score of 3.37. Linkage disequilibrium mapping identified 2 regions of association: one was under the peak of linkage, the other was 27 cM distal. In another study, the International Molecular Genetic Study of Autism Consortium (2001) found a multipoint maximum lod score of 3.20 at marker D7S477. They also detected a multipoint maximum lod score of 4.80 at marker D2S188 on chromosome 2q.

In 12 of 105 families with 2 or more sibs affected with autism, Yu et al. (2002) identified deletions ranging from 5 to more than 260 kb. One family had complex deletions at marker D7S630 on 7q21-q22, 3 families had different deletions at D7S517 on 7p, and another 3 families had different deletions at D8S264 on 8p. Yu et al. (2002) suggested that autism susceptibility alleles may cause the deletions by inducing errors during meiosis.

In a metaanalysis of 9 published genome scans on autism or autism spectrum disorders, Trikalinos et al. (2006) found evidence for significant linkage to 7q22-q32, confirming the findings of previous studies. The flanking region 7q32-qter reached a less stringent threshold for significance.

Genetic Heterogeneity

Using findings from a family study of autism and a similar study of twins, Pickles et al. (1995) concluded that autism has a multiple locus mode of inheritance involving 3 loci. Risch et al. (1999) performed a 2-stage genomewide screen of 2 groups of families with autism: 90 families comprising 97 affected sib pairs (ASPs) and 49 families with 50 affected sib pairs. Unaffected sibs, which provided 51 discordant sib pairs (DSPs) for the initial screen and 29 for the follow-up, were included as controls. There was a slightly increased identity by descent (IBD) in the ASPs (sharing of 51.6%) compared with the DSPs (sharing of 50.8%). The results were considered most compatible with a model specifying a large number of loci, perhaps 15 or more, and less compatible with models specifying 10 or fewer loci. The largest lod scores obtained were for a marker on 1p yielding a maximum multipoint lod score of 2.15, and on 17p, yielding a maximum lod score of 1.21.

In 51 multiplex families with autism, Philippe et al. (1999) used nonparametric linkage analysis to perform a genomewide screen with 264 microsatellite markers. By 2-point and multipoint affected sib-pair analyses, 11 regions gave nominal P values of 0.05 or lower. Philippe et al. (1999) observed overlap of 4 of these regions with regions on 2q, 7q, 6p, and 19p that had been identified by the earlier genomewide scan of autism conducted by the International Molecular Genetic Study of Autism Consortium (1998). The most significant multipoint linkage was close to marker D6S283 (maximum lod score = 2.23, p = 0.0013).

Smalley (1997) reported on the status of linkage studies in autism. Lamb et al. (2000) reviewed chromosomal aberrations, candidate gene studies, and linkage studies of autism.

Liu et al. (2001) genotyped 335 microsatellite markers in 110 multiplex families with autism. All families included at least 2 affected sibs, at least 1 of whom had autism; the remaining affected sibs carried diagnoses of either Asperger syndrome or pervasive developmental disorder. Affected sib-pair analysis yielded multipoint maximum lod scores that reached the accepted threshold for suggestive linkage on chromosomes 5, X, and 19. Further analysis yielded impressive evidence for linkage of autism and autism-spectrum disorders to markers on chromosomes 5 and 8, with suggestive linkage to a marker on chromosome 19.

Yonan et al. (2003) followed up on previously reported genomewide screens for autism performed by Liu et al. (2001) and Alarcon et al. (2002) showing suggestive evidence for linkage of autism spectrum disorders on chromosomes 5, 8, 16, 19, and X, and nominal evidence on several additional chromosomes. In their analysis, Yonan et al. (2003) increased the sample size 3-fold. Multipoint maximum lod scores obtained from affected sib-pair analysis of all 345 families yielded suggestive evidence for linkage on chromosomes 17, 5, 11, 4, and 8 (listed in order of MLS). The most significant findings were an MLS of 2.83 on 17q11 (AUTS6; 609378) and an MLS of 2.54 on 5p.

The genetic architecture of autism spectrum disorders (ASDs) is complex, requiring large samples to overcome heterogeneity. The Autism Genome Project Consortium (2007) broadened coverage and sample size relative to other studies of ASDs by using Affymetrix 10K SNP arrays and 1,181 families with at least 2 affected individuals, performing the largest linkage scan to that time while also analyzing copy number variation in these families. Linkage and copy number variation analyses implicated 11p13-p12 and neurexins, respectively, among other candidate loci. Neurexins teamed with previously implicated neuroligins for glutamatergic synaptogenesis, highlighting glutamate-related genes as promising candidates for contributing to ASDs. See neuroligin-3 (NLGN3; 300336) and neuroligin-4 (NLGN4; 300427).

Wang et al. (2009) presented the results of a genomewide association study of ASDs on a cohort of 780 families (3,101 subjects) with affected children, and a second cohort of 1,204 affected subjects and 6,491 control subjects, all of whom were of European ancestry. Six SNPs on chromosome 5p14.1 between cadherin-10 (CDH10; 604555) and cadherin-9 (CDH9; 609974), 2 genes encoding neuronal cell adhesion molecules, revealed strong association signals, with the most significant SNP being rs4307059 (p = 3.4 x 10(-8), odds ratio = 1.19). These signals were replicated in 2 independent cohorts, with combined P values ranging from 7.4 x 10(-8) to 2.1 x 10(-10). The authors concluded that their results implicated neuronal cell adhesion molecules in the pathogenesis of ASDs.

In a genomewide association study of 438 Caucasian families including 1,390 individuals with autism, Ma et al. (2009) found evidence for linkage to chromosome 5p14.1. Validation in an additional cohort of 2,390 samples from 457 families showed that 8 SNPs on chromosome 5p14.1 were significantly associated with autism (p values ranging from 3.24 x 10(-4) to 3.40 x 10(-6)). The most significant linkage was with rs10038113.

Using array CGH analysis, Roohi et al. (2009) identified a chromosome 3 copy number variation (CNV) disrupting the CNTN4 gene (607280) in 3 of 92 individuals with autism spectrum disorder. Two sibs had a deletion, and another unrelated individual had a duplication; both variations were inherited from an unaffected father. A third affected sib of the familial cases did not carry the deletion, suggesting incomplete penetrance or that he had a different disorder. The changes resulted from Alu-mediated unequal recombinations.

Glessner et al. (2009) presented the results from a whole-genome CNV study on a cohort of 859 ASD cases and 1,409 healthy children of European ancestry who were genotyped with approximately 550,000 SNP markers, in an attempt to comprehensively identify CNVs conferring susceptibility to ASDs. Positive findings were evaluated in an independent cohort of 1,336 ASD cases and 1,110 controls of European ancestry. Glessner et al. (2009) confirmed known associations, such as that with NRXN1 (600565) and CNTN4 (Roohi et al., 2009), in addition to several novel susceptibility genes encoding neuronal cell adhesion molecules, including NLGN1 (600568) and ASTN2 (612856), that were enriched with CNVs in ASD cases compared to controls (P = 9.5 x 10(-3)). Furthermore, CNVs within or surrounding genes involved in the ubiquitin pathways, including UBE3A (601623), PARK2 (602544), RFWD2 (608067), and FBXO40 (609107), were affected by CNVs not observed in controls (p = 3.3 x 10(-3)). Glessner et al. (2009) also identified duplications 55 kb upstream of complementary DNA AK123120 (p = 3.6 x 10(-6)). Glessner et al. (2009) concluded that although these variants may be individually rare, they appear to target genes involved in neuronal cell-adhesion or ubiquitin degradation, indicating that these 2 important gene networks expressed within the central nervous system (CNS) may contribute to genetic susceptibility to ASD.

Weiss et al. (2009) initiated a linkage and association mapping study using half a million genomewide SNPs in a common set of 1,031 multiplex autism families (1,553 affected offspring). They identified regions of suggestive and significant linkage on chromosomes 6q27 and 20p13, respectively. Initial analysis did not yield genomewide significant associations; however, genotyping of top hits in additional families revealed an SNP on chromosome 5p15 (rs10513025) between SEMA5A (609297) and TAS2R1 (604796) that was significantly associated with autism (p = 2.0 x 10(-7)). Weiss et al. (2009) also demonstrated that expression of SEMA5A is reduced in brains from autistic patients, further implicating SEMA5A as an autism susceptibility gene.

Kilpinen et al. (2009) carried out a genomewide microsatellite-based scan of a unique extended Finnish autism pedigree comprised of 20 families with verified genealogic links reaching back to the 17th century. Linkage analysis and fine mapping revealed significant results for SNPs (rs4806893, rs216283, and rs216276) on chromosome 19p13.3 in close proximity to TLE2 (601041) and TLE6 (612399) genes. They also obtained a significant result for a SNP rs1016732 on chromosome 1q23 near the ATP1A2 gene (182340). Kilpinen et al. (2009) noted that chromosome 1q23 had been previously reported as an autism susceptibility locus in Finnish families.

Exclusion Studies

In a multicenter study in Sweden, Blomquist et al. (1985) found the fragile X repeat (309550) in 13 of 83 boys (16%) with infantile autism, but in none of 19 girls with infantile autism.

Using the UCLA Registry for Genetic Studies of Autism, Spence et al. (1985) studied 46 families with at least 2 affected children. Linkage studies in 34 families showed no evidence of linkage with HLA (142800), and close linkage with 19 other autosomal markers was excluded. The highest lod score, 1.04, was found with haptoglobin (140100) on chromosome 16q22 at recombination fractions of 10% in males and 50% in females. There was no association of the disorder with fragile X.

Using data from 38 multiplex families with autism to perform a multipoint linkage analysis with markers on the X chromosome, Hallmayer et al. (1996) excluded a moderate to strong gene effect causing autism on the X chromosome.

Using the Autism Diagnostic Instrument-Revised (ADI-R), the Autism Diagnostic Observation Scale (ADOS), and psychometric tests, Klauck et al. (1997) identified 141 autistic patients from 105 simplex and 18 multiplex families; 131 patients met all 4 ADI-R algorithm criteria for autism and 10 patients showed a broader phenotype of autism. Using amplification of the CCG repeat at the fragile X locus, hybridization to the complete FMR1 cDNA probe, and hybridization to additional probes from the neighborhood of the FMR1 gene, the authors found no significant changes in 139 patients (99%) from 122 families. In 1 multiplex family with 3 children showing no dysmorphic features of the fragile X syndrome (1 male meeting 3 of 4 ADI-algorithm criteria, 1 normal male with slight learning disability but negative ADI-R testing, and 1 fully autistic female), the FRAXA full-mutation-specific CCG-repeat expansion in the genotype was not correlated with the autism phenotype. Further analysis revealed a mosaic pattern of methylation at the FMR1 gene locus in the 2 sons of the family, indicating at least a partly functional gene. Klauck et al. (1997) concluded that the association of autism with fragile X at Xq27.3 is nonexistent and excluded this location as a candidate gene for autism.


Cytogenetics

Lopreiato and Wulfsberg (1992) described a complex chromosomal rearrangement in a 6.5-year-old boy with autism who was otherwise normal except for minimal dysmorphism. The rearrangement seen in every cell examined involved chromosomes 1, 7 and 21: 46, XY, -1, -7, -21, t(1;7;21)(1p22.1-qter::21q22.3-qter; 7pter-q11.23::7q36.1-qter; 21pter-q22.3::7q11.23-q36.1::1pter-p22.1).

Vincent et al. (2006) reported 2 brothers with autism who both carried a paracentric inversion of chromosome 4p, inv(4)(p12-p15.3), inherited from an unaffected mother and unaffected maternal grandfather. More detailed molecular analysis showed that the proximal breakpoint on 4p12 involved a cluster of GABA receptor genes, including the GABRA4 gene (137141), which has been implicated in autism (Ma et al., 2005; Collins et al., 2006). Maestrini et al. (1999) found no association or linkage to the GABRB3 gene in 94 families comprising 174 individuals with autism.

Moessner et al. (2007) identified deletions in the SHANK3 gene (606230) on chromosome 22q13 in 3 (0.75%) of 400 unrelated patients with an autism spectrum disorder. The deletions ranged in size from 277 kb to 4.36 Mb; 1 patient also had a 1.4-Mb duplication at chromosome 20q13.33. The patients were essentially nonverbal and showed poor social interactions and repetitive behaviors. Two had global developmental delay and mild dysmorphic features. A fourth patient with a de novo missense mutation in the SHANK3 gene had autism-like features but had diagnostic scores above the cutoff for autism; she was conceived by in vitro fertilization. See also the chromosome 22q13.3 deletion syndrome (606232).

Copy Number Variation

Using high-resolution microarray analysis, Marshall et al. (2008) found 277 unbalanced copy number variations (CNV), including deletion, duplication, translocation, and inversion, in 189 (44%) of 427 families with autism spectrum disorder (ASD). These specific changes were not present in a total of about 1,600 controls, although control individuals also carried many CNV. Although most variants were inherited among the patients, 27 cases had de novo alterations, and 3 (11%) of these individuals had 2 or more changes. Marshall et al. (2008) detected 13 loci with recurrent or overlapping CNV in unrelated cases. Of note, CNV at chromosome 16p11.2 (AUTS14; see 611913) was identified in 4 (1%) of 427 families and none of 1,652 controls (p = 0.002). Some of the autism loci were also common to mental retardation loci. Marshall et al. (2008) concluded that structural gene variants were found in sufficiently high frequency influencing autism spectrum disorder to suggest that cytogenetic and microarray analyses be considered in routine clinical workup.

Cusco et al. (2009) analyzed 96 Spanish patients with idiopathic ASD by array CGH analysis. Only 13 of the 238 detected CNVs (range, 89 kb-2.4 Mb) were present specifically in 12 (12.5%) of 96 ASD patients. The CNVs consisted of 10 duplications and 3 deletions. In 5 patients with parental samples available, CNVs were inherited from a normal parent. Two CNVs mapped to regions with previously reported ASD candidates, KIAA0442 (607270) on chromosome 7q11.22 and GRM8 (601116) on chromosome 7q31.3. Of the 24 genes in the CNVs, some act in common pathways, most notably the phosphatidylinositol signaling and the glutamatergic synapse, both known to be affected in several genetic syndromes related with autism and previously associated with ASD. Cusco et al. (2009) hypothesized that functional alteration of genes in related neuronal networks is involved in the etiology of the ASD phenotype.

Sebat et al. (2007) tested the hypothesis that de novo CNV is associated with autism spectrum disorders. They performed comparative genomic hybridization (CGH) on the genomic DNA of patients and unaffected subjects to detect copy number variants not present in their respective parents. Candidate genomic regions were validated by higher-resolution CGH, paternity testing, cytogenetics, fluorescence in situ hybridization, and microsatellite genotyping. Confirmed de novo CNVs were significantly associated with autism (p = 0.0005). Such CNVs were identified in 12 of 118 (10%) of patients with sporadic autism, in 2 of 77 (3%) of patients with an affected first-degree relative, and in 2 of 196 (1%) of controls. The authors stated that most de novo CNVs were smaller than microscopic resolution. Affected genomic regions were highly heterogeneous and included mutations of single genes. Sebat et al. (2007) concluded that their findings established de novo germline mutation as a more significant risk factor for autism spectrum disorders than previously recognized.

Pinto et al. (2010) analyzed the genomewide characteristics of rare (less than 1% frequency) CNVs in ASD using dense genotyping arrays. When comparing 996 ASD individuals of European ancestry to 1,287 matched controls, cases were found to carry a higher global burden of rare, genic CNVs (1.19-fold, p = 0.012), especially so for loci previously implicated in either ASD and/or intellectual disability (1.69-fold, p = 3.4 x 10(-4)). Among the CNVs there were numerous de novo and inherited events, sometimes in combination in a given family, implicating many novel ASD genes such as SHANK2 (603290), SYNGAP1 (603384), DLGAP2 (605438), and the X-linked DDX53-PTCHD1 locus (see 300828). The authors also discovered an enrichment of CNVs disrupting functional gene sets involved in cellular proliferation, projection and motility, and GTPase/Ras signaling.

Levy et al. (2011) studied 887 families from the Simons Simplex Collection of relatively high functioning ASD families. They identified 75 de novo CNVs in 68 probands (approximately 8% of probands). Only a few were recurrent. Variation at the 16p11.2 locus was detected in more than 1% of patients (10 of 858), with deletions present in 6 and duplications in 4. In addition, the duplication at 7q11.2 of the Williams syndrome region (609757) was also seen as a recurrent CNV. Levy et al. (2011) noted that the finding of 8% of ASD probands with de novo events compared with 2% in unaffected sibs was in keeping with other reports. They speculated that the slightly lower incidence of 8% rather than the 10% reported by Sebat et al. (2007) related to a higher functioning autism group in this cohort or to smaller families in this study. Using their study and the previous literature, Levy et al. (2011) proposed a list of 'asymmetries,' or observed biases, in simplex families with autism. There is a higher incidence of de novo copy number mutation in children with ASDs from simplex families than in their sibs. There is a higher incidence of de novo copy number mutation in children with ASDs from simplex families than in children with ASDs from multiplex families. For transmitted rare events, duplications greatly outweigh deletions; deletions outweigh duplications among de novo events in children with ASDs. There is evidence of transmission distortion for ultrarare events to children with ASDs; this bias arises from families in which the sib is an unaffected male. Females are less likely to be diagnosed with ASDs than are males. A higher proportion of females with ASDs have detectable de novo copy number events than do males with ASDs, and the events are larger. Levy et al. (2011) suggested that females are protected from autism but did not propose a mechanism.

Sanders et al. (2011) examined 1,124 ASD simplex families from the Simons Simplex Collection. Each of the families was comprised of a single proband, unaffected parents, and in most kindreds an unaffected sib. Sanders et al. (2011) found significant association of ASD with de novo duplications of 7q11.23. They also identified rare recurrent de novo CNVs at 5 additional regions, including 16p13.2. Overall, large de novo CNVs conferred substantial risks (odds ratio = 5.6; CI = 2.6-12.0, p = 2.4 x 10(-7)). Sanders et al. (2011) suggested that there are 130 to 234 ASD-related CNV regions in the human genome and presented compelling evidence, based on cumulative data, for association of rare de novo events at 7q11.23, 15q11.2-q13.1 (see 608636), 16p11.2, and neurexin-1 (600565). Sanders et al. (2011) found that probands carrying a 16p11.2 or 7q11.23 de novo CNV were indistinguishable from the larger ASD group with respect to IQ, ASD severity, or categorical autism diagnosis. However, they did find a relationship between body weight and 16p11.2 deletions and duplications. When copy number was treated as an ordinal variable, BMI diminished as 16p11.2 copy number increased (p = 0.02).

Vaags et al. (2012) reported 4 families in which 1 or more members had autism spectrum disorder associated with heterozygous deletions of chromosome 14q affecting the NRXN3 gene (600567). The deletions were all different and ranged from 63 to 336 kb. One deletion affected only the NRXN3 alpha isoform, whereas 3 affected both the alpha and beta isoforms. Two families were ascertained from 1,158 Canadian individuals with ASD who were screened for copy number variations across the genome. The third family was 1 of 1,368 ASD cases screened, and the fourth was 1 of 1,796 ASD cases screened. The phenotype was variable, ranging from high-functioning Asperger syndrome to full autism with some pervasive developmental and behavioral problems. In 1 family, the deletion occurred de novo. In the other families, the deletion was inherited from a parent; 1 parent had a broader autism phenotype, 1 self-reported mild autistic-like features, and 1 was normal. In 1 family, 2 of 3 trizygotic triplets with autism carried the deletion; the third unaffected child did not carry the deletion. Small deletions affecting only the alpha isoform were found in 4 of 15,122 controls. The report suggested that deletions affecting the NRXN3 gene may predispose to the development of autism spectrum disorder, but segregation patterns within the families suggested issues of penetrance and expressivity at this locus.

Loirat et al. (2010) reported 3 unrelated boys with heterozygous de novo deletions in chromosome 17q12 (see 614527) who had cystic or hyperechogenic kidneys and autism. Their 17q12 deletions ranged from 1.5 to 1.8 Mb, and included LHX1 (601999), HNF1B (189907), and 19 other genes; sequencing of the LHX1 gene in the 3 boys and 32 control patients with autism revealed no mutations. Loirat et al. (2010) concluded that autism might be an additional manifestation associated with HNF1B deletion.

Moreno-De-Luca et al. (2010) performed cytogenomic array analysis in a discovery sample of patients with neurodevelopmental disorders and detected a recurrent 1.4-Mb deletion at chromosome 17q12 in 18 of 15,749 patients, including 6 with autism or autistic features; the deletion was not found in 4,519 controls. In a large follow-up sample, the same deletion was identified in 2 of 1,182 patients with autism spectrum disorder and/or neurocognitive impairment, and in 4 of 6,340 schizophrenia (see 181500) patients, but was not found in 47,929 controls (corrected p = 7.37 x 10 (-5)). Moreno-De-Luca et al. (2010) concluded that deletion 17q12 is a recurrent, pathogenic CNV that confers a high risk for autism spectrum disorder and schizophrenia, and that 1 or more of the 15 genes in the deleted interval is dosage-sensitive and essential for normal brain development and function.

Luo et al. (2012) interrogated gene expression in lymphoblasts from 439 individuals from 244 families with discordant sibs in the Simons Simplex Collection and found that the overall frequency of significantly misexpressed genes, which they referred to as outliers, did not differ between probands and unaffected sibs. However, in probands, but not their unaffected sibs, the group of outlier genes was significantly enriched in neural-related pathways, including neuropeptide signaling, synaptogenesis, and cell adhesion. The outlier genes clustered within large rare de novo CNVs and could be used for the prioritization of rare CNVs of potential significance. Several nonrecurrent CNVs with significant gene expression alterations were identified, including deletions in chromosome regions 3q27, 3p13, and 3p26 and duplications at 2p15, suggesting these as potential ASD loci.

See SHANK1 (604999) for discussion of a possible association between heterozygous deletions involving the SHANK1 gene on chromosome 19q13 and susceptibility to high-functioning autism.

Krumm et al. (2013) searched for disruptive, genic rare CNVs among 411 families affected by sporadic autism spectrum disorder from the Simons Simplex Collection by using available exome sequence data and CoNIFER (Copy Number Inference from Exome Reads). Compared to high density SNP microarrays, the authors' approach yielded approximately 2 times more smaller genic rare CNVs. Krumm et al. (2013) found that affected probands inherited more CNVs than did their sibs (453 vs 394, p = 0.004; odds ratio = 1.19) and that the probands' CNVs affected more genes (921 vs 726, p = 0.02; odds ratio = 1.30). These smaller CNVs (median size 18 kb) were transmitted preferentially from the mother (136 maternal vs 100 paternal, p = 0.02), although this bias occurred irrespective of affected status. The excess burden of inherited CNVs was driven primarily by sib pairs with discordant social behavior phenotypes, which contrasts with families where the phenotypes were more closely matched or less extreme. In a combined model, the inherited CNVs, de novo CNVs, and de novo single-nucleotide variants all independently contributed to the risk of autism (p less than 0.05).

Poultney et al. (2013) used the eXome Hidden Markov Model (XHMM) as well as transmission information and validation by molecular methods to confirm that small CNVs encompassing as few as 3 exons can be reliably called from whole-exome data. They applied this approach to an autism case-control sample of 811 subjects (mean per-target read depth = 161) and observed a significant increase in the burden of rare (minor allele frequency (MAF) 1% or less) 1- to 30-kb CNVs, 1- to 30-kb deletions, and 1- to 10-kb deletions in ASD. CNVs in the 1 to 30 kb range frequently hit just a single gene, allowing Poultney et al. (2013) to observe enrichment for disruption of genes in cytoskeletal and autophagy pathways in ASD. Poultney et al. (2013) concluded that rare 1- to 30-kb exonic deletions could contribute to risk in up to 7% of individuals with ASD.

Girirajan et al. (2013) exploited the repeat architecture of the genome to target segmental duplication-mediated rearrangement hotspots (n = 120, median size 1.78 Mbp, range 240 kbp to 13 Mbp) and smaller hotspots flanked by repetitive sequence (n = 1,247, median size 79 kbp, range 3-96 kbp) in 2,588 autistic individuals from simplex and multiplex families and in 580 controls. The analysis identified several recurrent large hotspot events, including association with 1q21 duplications, which are more likely to be identified in individuals with autism than in those with developmental delay (p = 0.01; odds ratio = 2.7). Within larger hotspots, Girirajan et al. (2013) also identified smaller atypical CNVs that implicated CHD1L (613039) and ACACA (200350) for the 1q21 and 17q12 deletions, respectively. The analysis, however, suggested no overall increase in the burden of smaller hotspots in autistic individuals as compared to controls. By focusing on gene-disruptive events, Girirajan et al. (2013) identified several genes that were enriched for CNVs in autism cases versus controls, including DPP10 (608209), PLCB1 (607120), TRPM1 (603576), NRXN1 (600565), FHIT (601153), and HYDIN (610812). Girirajan et al. (2013) found that as the size of deletions increases, nonverbal IQ significantly decreases, but there is no impact on autism severity; as the size of duplications increases, autism severity significantly increases but nonverbal IQ is not affected. Girirajan et al. (2013) concluded that the absence of an increased burden of smaller CNVs in individuals with autism and the failure of most large hotspots to refine to single genes is consistent with a model where imbalance of multiple genes contributes to a disease state.

Pinto et al. (2014) analyzed 2,446 ASD-affected families and confirmed an excess of genic deletions and duplications in affected versus control groups (1.41-fold, p = 1.0 x 10(-5)). They also found an increase in affected subjects carrying exonic pathogenic CNVs overlapping known loci associated with dominant or X-linked ASD and intellectual disability (odds ratio = 12.62, p = 2.7 x 10(-15), approximately 3% of ASD subjects). Pathogenic CNVs, often showing variable expressivity, included rare de novo and inherited events at 36 loci, implicating ASD-associated genes (CHD2, 602119; HDAC4, 605314; and GDI1, 300104) linked to other neurodevelopmental disorders, as well as other genes, such as SETD5 (615743), MIR137 (614304), and HDAC9 (606543). Consistent with hypothesized gender-specific modulators, females with ASD were more likely to have highly penetrant CNVs (p = 0.017) and were also overrepresented among subjects with fragile X syndrome protein targets (p = 0.02).


Molecular Genetics

Gauthier et al. (2011) identified a heterozygous 1-bp deletion (2733delT) in the NRXN2 gene (600566) on chromosome 11q13 in a boy of European ancestry with autism spectrum disorder. The mutation resulted in premature termination. In vitro functional expression studies in COS-7 cells showed that the mutant protein was unable to bind its usual partners, and in vitro studies in neuronal culture showed a loss of synaptogenic activity with lack of clustering of postsynaptic components. The findings were consistent with a loss of function. The mutation was inherited from the patient's father, who had severe language delay. A maternal aunt of the father's had schizophrenia, but DNA was not available from her. The patient was identified from a cohort of 142 patients with autism who were screened for mutations in the NRXN1 (600565), NRXN2, and NRXN3 genes.

Sanders et al. (2012) used whole-exome sequencing of 928 individuals, including 200 phenotypically discordant sib pairs, to demonstrate that highly disruptive nonsense and splice site de novo mutations in brain-expressed genes are associated with autism spectrum disorders and carry large effects. On the basis of mutation rates in unaffected individuals, they demonstrated that multiple independent de novo single-nucleotide variants in the same gene among unrelated probands reliably identifies risk alleles, providing a clear path forward for gene discovery. Among a total of 279 identified de novo coding mutations, there was a single instance in probands, and none in sibs, in which 2 independent nonsense variants disrupt the same gene, SCN2A (182390). Sanders et al. (2012) combined all de novo events in their sample with those identified in the study of O'Roak et al. (2012) and observed from a total of 414 probands 2 additional genes carrying 2 highly disruptive mutations each, KATNAL2 (614697) and CHD8 (610528).

O'Roak et al. (2012) performed whole-exome sequencing for parent-child trios exhibiting sporadic autism spectrum disorders, including 189 new trios and 20 that were previously reported (O'Roak et al., 2011). In addition, O'Roak et al. (2012) sequenced the exomes of 50 unaffected sibs corresponding to 31 of the new and 19 of the previously reported trios, for a total of 677 individual exomes from 209 families. O'Roak et al. (2012) showed that de novo point mutations are overwhelmingly paternal in origin (4:1 bias) and positively correlated with paternal age, consistent with the modest increased risk for children of older fathers to develop autism spectrum disorders. Moreover, 39% (49 of 126) of the most severe or disruptive de novo mutations mapped to a highly interconnected beta-catenin (116806)/chromatin remodeling protein network ranked significantly for autism candidate genes. In proband exomes, recurrent protein-altering mutations were observed in 2 genes: CHD8 and NTNG1. Mutation screening of 6 candidate genes in 1,703 ASD probands identified additional de novo, protein-altering mutations in GRIN2B (138252), LAMC3 (604349), and SCN1A (182389). Combined with copy number data, these data indicated extreme locus heterogeneity in ASD. O'Roak et al. (2012) concluded that their analysis predicted extreme locus heterogeneity underlying the genetic etiology of autism. Under a strict sporadic disorder-de novo mutation model, if 20 to 30% of the de novo point mutations are considered to be pathogenic, they could estimate between 384 and 821 loci. Furthermore, 1 individual inherited 3 rare gene disruptive CNVs and carried 2 de novo truncating mutations.

Neale et al. (2012) assessed the role of de novo mutations in autism spectrum disorders by sequencing the exomes of ASD cases and their parents (175 trios). Fewer than half of the cases (46.3%) carried a missense or nonsense de novo variant, and the overall rate of mutation was only modestly higher than the expected rate. In contrast, the proteins encoded by genes that harbored de novo missense or nonsense mutations showed a higher degree of connectivity among themselves and to previous ASD genes as indexed by protein-protein interaction screens. The small increase in the rate of de novo events, when taken together with the protein interaction results, are consistent with an important but limited role for de novo point mutations in ASD, similar to that documented for de novo copy number variants. Genetic models incorporating data indicated that most of the observed de novo events are unconnected to ASD; those that do confer risk are distributed across many genes and are incompletely penetrant (i.e., not necessarily sufficient for disease). Neale et al. (2012) concluded that their results supported polygenic models in which spontaneous coding mutations in any of a large number of genes increases risk by 5- to 20-fold. Despite the challenge posed by such models, results from de novo events and a large parallel case-control study provided strong evidence in favor of CHD8 and KATNAL2 as genuine autism risk factors.

O'Roak et al. (2012) developed a modified molecular inversion probe method enabling ultra-low-cost candidate gene resequencing in very large cohorts. To demonstrate the power of this approach, O'Roak et al. (2012) captured and sequenced 44 candidate genes in 2,446 ASD probands, and discovered 27 de novo events in 16 genes, 59% of which are predicted to truncate proteins or disrupt splicing. O'Roak et al. (2012) estimated that recurrent disruptive mutations in 6 genes--CHD8, DYRK1A (600855), GRIN2B, TBR1 (604616), PTEN (601728), and TBL1XR1 (608628)--may contribute to 1% of sporadic autism spectrum disorders. O'Roak et al. (2012) concluded that their data supported associations between specific genes and reciprocal subphenotypes (CHD8-macrocephaly and DYRK1A-microcephaly) and replicated the importance of a beta-catenin/chromatin-remodeling network to ASD etiology.

Jiang et al. (2013) used whole-genome sequencing to examine 32 families with ASD to detect de novo or rare inherited genetic variants predicted to be deleterious (loss-of-function and damaging missense mutations). Among ASD probands, Jiang et al. (2013) identified deleterious de novo mutations in 6 of 32 (19%) families and X-linked or autosomal inherited alterations in 10 of 32 (31%) families (some had combinations of mutations). The proportion of families identified with such putative mutations was larger than had been reported; this yield was in part due to the comprehensive and uniform coverage afforded by whole-genome sequencing. Deleterious variants were found in 4 unrecognized, 9 known, and 8 candidate ASD risk genes. Examples include CAPRIN1 (601178), AFF2 (300806), VIP (192320), SCN2A, KCNQ2 (602235), NRXN1, and CHD7 (608892).

To characterize the role of rare complete human knockouts in autism spectrum disorders, Lim et al. (2013) identified genes with homozygous or compound heterozygous loss-of-function variants (defined as nonsense and essential splice sites) from exome sequencing of 933 cases and 869 controls. Lim et al. (2013) identified a 2-fold increase in complete knockouts of autosomal genes with low rates of loss-of-function variation (less than or equal to 5% frequency) in cases, and estimated a 3% contribution to autism spectrum disorder risk by these events, confirming this observation in an independent set of 563 probands and 4,605 controls. Outside the pseudoautosomal regions on the X chromosome, Lim et al. (2013) similarly observed a significant 1.5-fold increase in rare hemizygous knockouts in males, contributing to another 2% of autism spectrum disorders in males. Lim et al. (2013) concluded that these results provided compelling evidence that rare autosomal and X chromosome complete gene knockouts are important inherited risk factors for autism spectrum disorders.

Using exome sequencing, De Rubeis et al. (2014) showed that analysis of rare coding variation in 3,871 autism cases and 9,937 ancestry-matched or parental controls implicated 22 autosomal genes at a false discovery rate (FDR) of less than 0.05, plus a set of 107 autosomal genes strongly enriched for those likely to affect risk (FDR less than 0.30). These 107 genes, which show unusual evolutionary constraint against mutations, incurred de novo loss-of-function mutations in over 5% of autistic subjects. Many of the genes implicated encode proteins for synaptic formation, transcriptional regulation, and chromatin-remodeling pathways. These included voltage-gated ion channels regulating the propagation of action potentials, pacemaking, and excitability-transcription coupling, as well as histone-modifying enzymes and chromatin remodelers, most prominently those that mediate posttranslational lysine methylation/demethylation modifications of histones.

Associations Pending Confirmation

See 605410.0001 for discussion of a possible association between variation in the KCND2 gene and infantile-onset severe refractory epilepsy (see EIEE1, 308350) and autism.

For discussion of a possible association between autism spectrum disorder and variation in the PRICKLE2 gene, see 608501.


Population Genetics

Smalley (1997) reported that autism has a population prevalence of approximately 4 to 5 in 10,000 with a male to female ratio of 4 to 1.

In a review of 20 studies on autism published between 1966 and 1997, Gillberg and Wing (1999) determined that autism is considerably more common than previously believed. The early studies yielded prevalence rates of under 0.5 per 1,000 children, whereas the later studies showed a mean rate of about 1 in 1,000. Children born after 1970 had a much higher rate than those born before 1970.

Bertrand et al. (2001) performed a prevalence study of autism spectrum disorders in Brick Township, New Jersey. There were 6.7 cases per 1,000 children, aged 3 to 10 years, in 1998. The prevalence for children whose condition met full diagnostic criteria for autistic disorder was 4.0 cases per 1,000 children, and the prevalence for PDD-not otherwise specified (NOS) and Asperger syndrome was 2.7 cases per 1,000 children.

In a review, Jones et al. (2008) noted that the significant increase in the frequency with which autism spectrum disorders is diagnosed, from 4 per 10,000 in 1950 to 40 to 60 per 10,000 as of 2008, results from greater awareness, availability of services, and changes in diagnostic criteria to include a broader spectrum of neurodevelopmental disorders, among others.


Pathogenesis

Schain and Freedman (1961) reported elevated levels of platelet serotonin (5-HT; see 182138) in patients with autism. Abramson et al. (1989) reported elevated blood serotonin in autistic probands and in their first-degree relatives. Piven et al. (1991) found that serotonin levels were significantly higher in autistic individuals with a sib with autism or PDD than in those without a sib with these disorders, and that autistic patients without an affected sib had serotonin levels that were significantly higher than controls.

A biologic basis of autism was suggested by the finding of developmental hypoplasia in lobules VI and VII of the cerebellar vermis (Courchesne et al., 1988). The ontogenetically, developmentally, and anatomically distinct vermal lobules I to V were found to be of normal size. However, Schaefer et al. (1996) disputed the relationship of cerebellar vermal atrophy to infantile autism. They found that the average relative size of lobules VI and VII of the cerebellar vermis was no different in their 13 patients with infantile autism when compared to that of 125 normal individuals. They found relative hypoplasia of lobules VI and VII in patients with Rett syndrome (312750) and Sotos cerebral gigantism (117550), 2 disorders characterized by autistic behaviors. No relative vermian atrophy was seen in other disorders associated with autistic behavior: fragile X, Angelman (AS; 105830), adult phenylketonuria (261600), and Sanfilippo (252900). Furthermore, they found a relative atrophy of lobules VI and VII in several patients with primary cerebellar hypoplasia and Usher syndrome type II (276901), syndromes not associated with autistic behavior.

Autopsy and neuroimaging studies have suggested that autism spectrum disorder is caused in part by abnormal brain development. Benayed et al. (2005) reviewed cerebellar abnormalities in autism spectrum disorder. The CNS structure most consistently affected in individuals with autism is the cerebellum, with a decrease in the number of Purkinje cells being present in a majority. Neurodegenerative signs are for the most part absent from these autopsy samples, suggesting a developmental defect. Neuroimaging studies have consistently demonstrated posterior cerebellar hypoplasia. Although the cerebellum has classically been considered a motor control center, functional imaging studies indicated that the cerebellum is also active during cognitive tasks that are defective in autism spectrum disorders, including language and attention. Thus, the identified cerebellar defects may contribute directly to some of the behavioral abnormalities associated with autism spectrum disorder. In turn, genetic alterations that perturb cerebellar development may contribute to susceptibility to autism spectrum disorder.

Regressive autism, characterized most prominently by a loss of language skills, has been attributed to environmental factors, particularly adverse reactions to vaccines; epidemiologic evidence, however, shows no association between vaccination and the rate of autism as reviewed by the Institute of Medicine Immunization Safety Reviews (2001); see also Taylor et al. (2002). Lainhart et al. (2002) noted that twin and family studies showed that the liability to autism extends beyond the full autism syndrome and includes qualitatively similar, albeit milder, deficits, referred to as the broader autism phenotype (BAP). If regressive autism is solely caused by environmental events, such as adverse reactions to vaccines, rates of the BAP in the relatives of children with regressive autism should be no greater than in the general population. If environmental events do not independently cause regressive autism, or if they act as 'second-hit' phenomena in children who already have a genetic liability to autism, rates of the BAP should be similar in relatives of autistic children with and without regression. Lainhart et al. (2002) found that the rate of the BAP was significantly higher in parents of children with regressive and nonregressive autism than in parents of nonautistic children. They concluded that environmental events are unlikely to be the sole cause of regressive autism, although environmental events may act in an additive or 'second-hit' fashion in individuals with a genetic vulnerability to autism.

In a review, Jones et al. (2008) discussed the hypothesis that dysregulation of methylation of brain-expressed genes on the X chromosome constitutes the major predisposition to the development of autism spectrum disorders. Broad evidence consistent with this epigenetic effect includes marked excess of males among individuals affected with ASD, most patients have a sporadic occurrence of the disorder, and most patients do not have syndromic features.

In studies of lymphocytes from 10 children with autism and 10 controls, Giulivi et al. (2010) found that patients with autism were more likely to have mitochondrial dysfunction, mtDNA overreplication, and mtDNA deletions compared to normally developing children. Lymphocytes from children with autism had lower mitochondrial-dependent oxygen consumption, with low complex I (6 of 10) and complex V (4 of 10) activity. Autistic children had increased plasma pyruvate levels and increased lymphocyte hydrogen peroxide production. Five autistic patients and 2 controls had mtDNA overreplication, and 2 patients and no controls had mtDNA deletion. Overall, the findings suggested that autism may be associated with mitochondrial dysfunction, which may reflect insufficient energy production. However, Giulivi et al. (2010) noted that the observations did not elucidate primary or secondary effects.

Castermans et al. (2010) described the positional cloning of SCAMP5 (613766) as a candidate gene for autism, based on finding a de novo chromosomal translocation t1;15(p36.11;q24.2) in a 40-year-old affected male. SCAMP5, which was silenced on the derivative chromosome, encodes a brain-enriched protein involved in membrane trafficking, similar to the previously identified candidate genes NBEA (604889) and AMISYN (STXBP6; 607958). Gene silencing of Nbea, Amisyn, and Scamp5 in mouse beta-TC3 cells resulted in a 2-fold increase in stimulated secretion of large dense-core vesicles (LDCVs), whereas overexpression suppressed secretion. Ultrastructural analysis of blood platelets from autism patients with haploinsufficiency of 1 of the 3 candidate genes showed morphologic abnormalities of dense-core granules, which closely resembled LDCVs. Castermans et al. (2010) suggested that in a subgroup of patients, the regulation of neuronal vesicle trafficking may be involved in the pathogenesis of autism.

Voineagu et al. (2011) analyzed postmortem brain tissue samples from 19 autism cases and 17 controls from the Autism Tissue Project and the Harvard brain bank using Illumina microarrays. For each individual, they profiled 3 regions previously implicated in autism: superior temporal gyrus, prefrontal cortex, and cerebellar vermis. Voineagu et al. (2011) demonstrated consistent differences in transcriptome organization between autistic and normal brain by gene coexpression network analysis. Remarkably, regional patterns of gene expression that typically distinguish frontal and temporal cortex are significantly attenuated in the autism spectrum disorder (ASD) brain, suggesting abnormalities in cortical patterning. Voineagu et al. (2011) further identified discrete modules of coexpressed genes associated with autism: a neuronal module enriched for known autism susceptibility genes, including the neuronal-specific splicing factor A2BP1 (also known as FOX1, 605104), and a module enriched for immune genes and glial markers. Using high-throughput RNA sequencing, they demonstrated dysregulated splicing of A2BP1-dependent alternative exons in the ASD brain. Moreover, using a published autism genomewide association study (GWAS) data set, Voineagu et al. (2011) showed that the neuronal module is enriched for genetically associated variants, providing independent support for the causal involvement of those genes in autism. The top module for differential expression between autism control groups was highly enriched for neuronal markers. The hubs of this group, called M12 in this study, which represented the genes with the highest rank of M12 membership, were A2BP1 but also APBA2 (602712), SCAMP5 (613766), CNTNAP1 (602346), KLC2 (611729), and CHRM1 (118510). In contrast, the immune-glial module showed no enrichment for autism GWAS signals, indicating a nongenetic etiology for this process. Voineagu et al. (2011) concluded that their results provided strong evidence for convergent molecular abnormalities in ASD, and implicated transcriptional and splicing dysregulation as underlying mechanisms of neuronal dysfunction in this disorder.

Gilman et al. (2011) developed a network-based analysis of genetic associations (NETBAG) and used this to identify a large biologic network of genes affected by rare de novo CNVs in autism. The genes forming the network are primarily related to synapse development, axon targeting, and neuron motility. The identified network was strongly related to genes previously implicated in autism and intellectual disability phenotypes. Gilman et al. (2011) suggested that their results were also consistent with the hypothesis that significantly stronger functional perturbations are required to trigger the autistic phenotype in females compared to males. Overall, the presented analysis of de novo variants supported the hypothesis that perturbed synaptogenesis is at the heart of autism. More generally, their study provided proof of the principle that networks underlying complex human phenotypes can be identified by a network-based functional analysis of rare genetic variants.

To study genomewide mutation rates, Kong et al. (2012) sequenced the entire genomes of 78 Icelandic parent-offspring trios at high coverage. Forty-four of the probands had autistic spectrum disorder and 21 were schizophrenic (181500). Kong et al. (2012) found that, with an average father's age of 29.7, the average de novo mutation rate is 1.20 x 10(-8) per nucleotide per generation. Most notably, the diversity in mutation rate of single-nucleotide polymorphisms was dominated by the age of the father at conception of the child. The effect is an increase of about 2 mutations per year. An exponential model estimates paternal mutations doubling every 16.5 years. After accounting for random Poisson variation, father's age is estimated to explain nearly all of the remaining variation in the de novo mutation counts. Kong et al. (2012) stated that there had been a recent transition of Icelanders from a rural agricultural to an urban industrial way of life, which engendered a rapid and sequential drop in the average age of fathers at conception from 34.9 years in 1900 to 27.9 years in 1980, followed by an equally swift climb back to 33.0 years in 2011, primarily owing to the effect of higher education and the increased use of contraception. On the basis of the fitted linear model, whereas individuals born in 1900 carried on average 73.7 de novo mutations, those born in 1980 carried on average only 59.7 such mutations (a decrease of 19.1%), and the mutational load of individuals born in 2011 had increased by 17.2% to 69.9. Kong et al. (2012) concluded that their observations shed light on the importance of the father's age on the risk of diseases such as schizophrenia and autism.

King et al. (2013) found that topotecan, a topoisomerase-1 (TOP1; 126420) inhibitor, dose-dependently reduces the expression of extremely long genes in mouse and human neurons, including nearly all genes that are longer than 200 kb. Expression of long genes is also reduced after knockdown of Top1 or Top2b (126431) in neurons, highlighting that both enzymes are required for full expression of long genes. By mapping RNA polymerase II density genomewide in neurons, King et al. (2013) found that this length-dependent effect on gene expression was due to impaired transcription elongation. Interestingly, many high-confidence autism spectrum disorder candidate genes are exceptionally long and were reduced in expression after TOP1 inhibition. King et al. (2013) concluded that chemicals and genetic mutations that impair topoisomerases could commonly contribute to autism spectrum disorders and other neurodevelopmental disorders.

Gamsiz et al. (2013) conducted a genomewide analysis of runs of homozygosity (ROH) in simplex ASD-affected families consisting of a proband diagnosed with ASD and at least 1 unaffected sib. In these families, probands with an IQ of 70 or below show more ROH than their unaffected sibs, whereas probands with an IQ greater than 70 do not show this excess. Although ASD is far more common in males than in females, the proportion of females increases with decreasing IQ. Gamsiz et al. (2013) stated that their data supported an association between ROH burden and autism diagnosis in girls; however, they were not able to show that this effect was independent of low IQ. The authors also identified several autism candidate genes on the basis of their being either a single gene that is within an ROH interval and that is recurrent in autism, or a gene that is within an ROH block and that harbors a homozygous rare deleterious variant upon analysis of exome sequencing data.

Through postmortem genomewide transcriptome analysis of the largest cohort of samples analyzed to that time, Parikshak et al. (2016) interrogated the noncoding transcriptome, alternative splicing, and upstream molecular regulators to broaden understanding of molecular convergence in ASD. The analysis revealed ASD-associated dysregulation of primate-specific long noncoding RNAs (lncRNAs, especially upregulation of LINC00693 and LINC00689), downregulation of the alternative splicing of activity-dependent neuron-specific exons, and attenuation of normal differences in gene expression between the frontal and temporal lobes. Their data suggested that SOX5 (604975), a transcription factor involved in neuron fate specification, contributes to this reduction in regional differences. Parikshak et al. (2016) further demonstrated that a genetically defined subtype of ASD, chromosome 15q11.2-13.1 duplication syndrome (dup15q; 608636), shares the core transcriptomic signature observed in idiopathic ASD. Coexpression network analysis revealed that individuals with ASD show age-related changes in the trajectory of microglial and synaptic function over the first 2 decades, and suggested that genetic risk for ASD may influence changes in regional cortical gene expression. Parikshak et al. (2016) concluded that their findings illustrated how diverse genetic perturbations can lead to phenotypic convergence at multiple biologic levels in a complex neuropsychiatric disorder.


Animal Model

Tabuchi et al. (2007) introduced the R451C (arg451 to cys; 300336.0001) substitution in neuroligin-3 into mice. R451C mutant mice showed impaired social interactions but enhanced spatial learning abilities. Unexpectedly these behavioral changes were accompanied by an increase in inhibitory synaptic transmission with no apparent effect on excitatory synapses. Deletion of neuroligin-3, in contrast, did not cause such changes, indicating that the R451C substitution represents a gain-of-function mutation. Tabuchi et al. (2007) concluded that increased inhibitory synaptic transmission may contribute to human autism spectrum disorders and that the R451C knockin mice may be a useful model for studying autism-related behaviors.

Using maternal immune activation (MIA) in a mouse model of ASD and genetic mutant mice, Choi et al. (2016) showed that Rorgt (see 602943)-dependent effector T lymphocytes, such as T-helper-17 (Th17) cells, and Il17a (603149) were required in mothers for MIA-induced behavioral abnormalities in offspring. MIA induced a maternal Il17a-dependent abnormal phenotype in cortex of fetal brain. Treating pregnant female mice with antibodies blocking Il17a ameliorated MIA-associated behavioral abnormalities. Choi et al. (2016) proposed that targeting of Th17 cells in susceptible pregnant mothers may reduce the likelihood of them bearing children with inflammation-induced ASD-like phenotypes.


History

Eisenberg (1994) provided a biographic sketch of Leo Kanner (1894-1981), the pioneer pediatric psychiatrist who first described and named infantile autism (Kanner, 1943).


REFERENCES

  1. Abramson, R. K., Wright, H. H., Carpenter, R., Brennan, W., Lumpuy, O., Cole, E., Young, S. R. Elevated blood serotonin in autistic probands and their first-degree relatives. J. Autism Dev. Disord. 19: 397-407, 1989. [PubMed: 2793785, related citations]

  2. Alarcon, M., Cantor, R. M., Liu, J., Gilliam, T. C., Autism Genetic Resource Exchange Consortium, Geschwind, D. H. Evidence for a language quantitative trait locus on chromosome 7q in multiplex autism families. Am. J. Hum. Genet. 70: 60-71, 2002. [PubMed: 11741194, images, related citations] [Full Text]

  3. American Psychiatric Association. Diagnostic and Statistical Manual of Mental Disorders. (4th ed.) Washington, D.C.: American Psychiatric Association (pub.) 1994.

  4. Autism Genome Project Consortium. Mapping autism risk loci using genetic linkage and chromosomal rearrangements. Nature Genet. 39: 319-328, 2007. Note: Erratum: Nature Genet. 39: 1285 only, 2007. [PubMed: 17322880, images, related citations] [Full Text]

  5. Awadalla, P., Gauthier, J., Myers, R. A., Casals, F., Hamdan, F. F., Griffing, A. R., Cote, M., Henrion, E., Spiegelman, D., Tarabeux, J., Piton, A., Yang, Y., and 22 others. Direct measure of the de novo mutation rate in autism and schizophrenia cohorts. Am. J. Hum. Genet. 87: 316-324, 2010. [PubMed: 20797689, related citations] [Full Text]

  6. Bailey, A., Le Couteur, A., Gottesman, I., Bolton, P., Simonoff, E., Yuzda, E., Rutter, M. Autism as a strongly genetic disorder: evidence from a British twin study. Psychol. Med. 25: 63-77, 1995. [PubMed: 7792363, related citations]

  7. Bailey, A., Phillips, W., Rutter, M. Autism: towards an integration of clinical, genetic, neuropsychological, and neurobiological perspectives. J. Child Psychol. Psychiat. 37: 89-126, 1996. [PubMed: 8655659, related citations] [Full Text]

  8. Benayed, R., Gharani, N., Rossman, I., Mancuso, V., Lazar, G., Kamdar, S., Bruse, S. E., Tischfield, S., Smith, B. J., Zimmerman, R. A., DiCicco-Bloom, E., Brzustowicz, L. M., Millonig, J. H. Support for the homeobox transcription factor gene ENGRAILED 2 as an autism spectrum disorder susceptibility locus. Am. J. Hum. Genet. 77: 851-868, 2005. [PubMed: 16252243, images, related citations] [Full Text]

  9. Bertrand, J., Mars, A., Boyle, C., Bove, F., Yeargin-Allsopp, M., Decoufle, P. Prevalence of autism in a United States population: the Brick Township, New Jersey, investigation. Pediatrics 108: 1155-1161, 2001. [PubMed: 11694696, related citations] [Full Text]

  10. Blomquist, H. K., Bohman, M., Edvinsson, S. O., Gillberg, C., Gustavson, K.-H., Holmgren, G., Wahlstrom, J. Frequency of the fragile X syndrome in infantile autism: a Swedish multicenter study. Clin. Genet. 27: 113-117, 1985. [PubMed: 3978844, related citations] [Full Text]

  11. Bolton, P., Macdonald, H., Pickles, A., Rios, P., Goode, S., Crowson, M., Bailey, A., Rutter, M. A case-control family history study of autism. J. Child Psychol. Psychiat. 35: 877-900, 1994. [PubMed: 7962246, related citations] [Full Text]

  12. Burd, L., Kerbeshian, J., Fisher, W. Inquiry into the incidence of hyperlexia in a statewide population of children with pervasive developmental disorder. Psychol. Rep. 57: 236-238, 1985. [PubMed: 4048337, related citations] [Full Text]

  13. Castermans, D., Volders, K., Crepel, A., Backx, L., De Vos, R., Freson, K., Meulemans, S., Vermeesch, J. R., Schrander-Stumpel, C. T. R. M., De Rijk, P., Del-Favero, J., Van Geet, C., Van De Ven, W. J. M., Steyaert, J. G., Devriendt, K., Creemers, J. W. M. SCAMP5, NBEA and AMISYN: three candidate genes for autism involved in secretion of large dense-core vesicles. Hum. Molec. Genet. 19: 1368-1378, 2010. [PubMed: 20071347, related citations] [Full Text]

  14. Choi, G. B., Yim, Y. S., Wong, H., Kim, S., Kim, H., Kim, S. V., Hoeffer, C. A., Littman, D. R., Huh, J. R. The maternal interleukin-17a pathway in mice promotes autism-like phenotypes in offspring. Science 351: 933-939, 2016. [PubMed: 26822608, images, related citations] [Full Text]

  15. Cohen, D., Pichard, N., Tordjman, S., Baumann, C., Burglen, L., Excoffier, E., Lazar, G., Mazet, P., Pinquier, C., Verloes, A., Heron, D. Specific genetic disorders and autism: clinical contribution towards their identification. J. Autism Dev. Disord. 35: 103-116, 2005. [PubMed: 15796126, related citations] [Full Text]

  16. Collins, A. L., Ma, D., Whitehead, P. L., Martin, E. R., Wright, H. H., Abramson, R. K., Hussman, J. P., Haines, J. L., Cuccaro, M. L., Gilbert, J. R., Pericak-Vance, M. A. Investigation of autism and GABA receptor subunit genes in multiple ethnic groups. Neurogenetics 7: 167-174, 2006. [PubMed: 16770606, related citations] [Full Text]

  17. Courchesne, E., Yeung-Courchesne, R., Press, G. A., Hesselink, J. R., Jernigan, T. L. Hypoplasia of cerebellar vermal lobules VI and VII in autism. New Eng. J. Med. 318: 1349-1354, 1988. [PubMed: 3367935, related citations] [Full Text]

  18. Cusco, I., Medrano, A., Gener, B., Vilardell, M., Gallastegui, F., Villa, O., Gonzalez, E., Rodriguez-Santiago, B., Vilella, E., Del Campo, M., Perez-Jurado, L. A. Autism-specific copy number variants further implicate the phosphatidylinositol signaling pathway and the glutamatergic synapse in the etiology of the disorder. Hum. Molec. Genet. 18: 1795-1804, 2009. [PubMed: 19246517, images, related citations] [Full Text]

  19. De Rubeis, S., He, X., Goldberg, A. P., Poultney, C. S., Samocha, K., Cicek, A. E., Kou, Y., Liu, L., Fromer, M., Walker, S., Singh, T., Klei, L., and 83 others. Synaptic, transcriptional and chromatin genes disrupted in autism. Nature 515: 209-215, 2014. [PubMed: 25363760, images, related citations] [Full Text]

  20. Eisenberg, L. Images in psychiatry: Leo Kanner (1894-1981). Am. J. Psychiat. 151: 751, 1994. [PubMed: 8166319, related citations] [Full Text]

  21. Folstein, S. E., Rosen-Sheidley, B. Genetics of autism: complex aetiology for a heterogeneous disorder. Nature Rev. Genet. 2: 943-955, 2001. [PubMed: 11733747, related citations] [Full Text]

  22. Folstein, S., Rutter, M. Genetic influences and infantile autism. Nature 265: 726-728, 1977. [PubMed: 558516, related citations]

  23. Folstein, S., Rutter, M. Infantile autism: a genetic study of 21 twin pairs. J. Child Psychol. Psychiat. 18: 297-321, 1977. [PubMed: 562353, related citations] [Full Text]

  24. Gamsiz, E. D., Viscidi, E. W., Frederick, A. M., Nagpal, S., Sanders, S. J., Murtha, M. T., Schmidt, M., Simons Simplex Collection Genetics Consortium, Triche, E. W., Geschwind, D. H., State, M. W., Istrail, S., Cook, E. H., Jr., Devlin, B., Morrow, E. M. Intellectual disability is associated with increased runs of homozygosity in simplex autism. Am. J. Hum. Genet. 93: 103-109, 2013. [PubMed: 23830515, images, related citations] [Full Text]

  25. Gauthier, J., Siddiqui, T. J., Huashan, P., Yokomaku, D., Hamdan, F. F., Champagne, N., Lapointe, M., Spiegelman, D., Noreau, A., Lafreniere, R. G., Fathalli, F., Joober, R., and 9 others. Truncating mutations in NRXN2 and NRXN1 in autism spectrum disorders and schizophrenia. Hum. Genet. 130: 563-573, 2011. [PubMed: 21424692, images, related citations] [Full Text]

  26. Gillberg, C., Wing, L. Autism: not an extremely rare disorder. Acta Psychiat. Scand. 99: 399-406, 1999. [PubMed: 10408260, related citations] [Full Text]

  27. Gilman, S. R., Iossifov, I., Levy, D., Ronemus, M., Wigler, M., Vitkup, D. Rare de novo variants associated with autism implicate a large functional network of genes involved in formation and function of synapses. Neuron 70: 898-907, 2011. [PubMed: 21658583, images, related citations] [Full Text]

  28. Girirajan, S., Dennis, M. Y., Baker, C., Malig, M., Coe, B. P., Campbell, C. D., Mark, K., Vu, T. H., Alkan, C., Cheng, Z., Biesecker, L. G., Bernier, R., Eichler, E. E. Refinement and discovery of new hotspots of copy-number variation associated with autism spectrum disorder. Am. J. Hum. Genet. 92: 221-237, 2013. [PubMed: 23375656, related citations] [Full Text]

  29. Giulivi, C., Zhang, Y-F., Omanska-Klusek, A., Ross-Inta, C., Wong, S., Hertz-Picciotto, I., Tassone, F., Pessah, I. N. Mitochondrial dysfunction in autism. J.A.M.A. 304: 2389-2396, 2010. [PubMed: 21119085, related citations] [Full Text]

  30. Glessner, J. T., Wang, K., Cai, G., Korvatska, O., Kim, C. E., Wood, S., Zhang, H., Estes, A., Brune, C. W., Bradfield, J. P., Imielinski, M., Frackelton, E. C., and 47 others. Autism genome-wide copy number variation reveals ubiquitin and neuronal genes. Nature 459: 569-573, 2009. [PubMed: 19404257, images, related citations] [Full Text]

  31. Greenberg, D. A., Hodge, S. E., Sowinski, J., Nicoll, D. Excess of twins among affected sibling pairs with autism: implications for the etiology of autism. Am. J. Hum. Genet. 69: 1062-1067, 2001. [PubMed: 11590546, related citations] [Full Text]

  32. Hallmayer, J., Glasson, E. J., Bower, C., Petterson, B., Croen, L., Grether, J., Risch, N. On the twin risk in autism. Am. J. Hum. Genet. 71: 941-946, 2002. [PubMed: 12297988, related citations] [Full Text]

  33. Hallmayer, J., Hebert, J. M., Spiker, D., Lotspeich, L., McMahon, W. M., Petersen, P. B., Nicholas, P., Pingree, C., Lin, A. A., Cavalli-Sforza, L. L., Risch, N., Ciaranello, R. D. Autism and the X chromosome: multipoint sib-pair analysis. Arch. Gen. Psychiat. 53: 985-989, 1996. [PubMed: 8911221, related citations] [Full Text]

  34. Institute of Medicine Immunization Safety Reviews. Immunization safety review: measles-mumps-rubella vaccine and autism. Washington, D.C.: National Academy Press (pub.) 2001. Pp. 13-69.

  35. International Molecular Genetic Study of Autism Consortium. A full genome screen for autism with evidence for linkage to a region on chromosome 7q. Hum. Molec. Genet. 7: 571-578, 1998. [PubMed: 9546821, related citations] [Full Text]

  36. International Molecular Genetic Study of Autism Consortium. A genomewide screen for autism: strong evidence for linkage to chromosomes 2q, 7q, and 16p. Am. J. Hum. Genet. 69: 570-581, 2001. [PubMed: 11481586, images, related citations] [Full Text]

  37. International Molecular Genetic Study of Autism Consortium. Further characterization of the autism susceptibility locus AUTS1 on chromosome 7q. Hum. Molec. Genet. 10: 973-982, 2001. [PubMed: 11392322, related citations]

  38. Iossifov, I., O'Roak, B. J., Sanders, S. J., Ronemus, M., Krumm, N., Levy, D., Stessman, H. A., Witherspoon, K. T., Vives, L., Patterson, K. E., Smith, J. D., Paeper, B., and 35 others. The contribution of de novo coding mutations to autism spectrum disorder. Nature 515: 216-221, 2014. [PubMed: 25363768, images, related citations] [Full Text]

  39. Jiang, Y., Yuen, R. K. C., Jin, X., Wang, M., Chen, N., Wu, X., Ju, J., Mei, J., Shi, Y., He, M., Wang, G., Liang, J., and 27 others. Detection of clinically relevant genetic variants in autism spectrum disorder by whole-genome sequencing. Am. J. Hum. Genet. 93: 249-263, 2013. [PubMed: 23849776, images, related citations] [Full Text]

  40. Jones, J. R., Skinner, C., Friez, M. J., Schwartz, C. E., Stevenson, R. E. Hypothesis: dysregulation of methylation of brain-expressed genes on the X chromosome and autism spectrum disorders. Am. J. Med. Genet. 146A: 2213-2220, 2008. [PubMed: 18698615, related citations] [Full Text]

  41. Jorde, L. B., Mason-Brothers, A., Waldmann, R., Ritvo, E. R., Freeman, B. J., Pingree, C., McMahon, W. M., Petersen, B., Jenson, W. R., Mo, A. The UCLA--University of Utah epidemiologic survey of autism: genealogical analysis of familial aggregation. Am. J. Med. Genet. 36: 85-88, 1990. [PubMed: 2333911, related citations] [Full Text]

  42. Kanner, L. Autistic disturbances of affective contact. Nerv. Child 2: 217-250, 1943.

  43. Kilpinen, H., Ylisaukko-oja, T., Rehnstrom, K., Gaal, E., Turunen, J. A., Kempas, E., von Wendt, L., Varilo, T., Peltonen, L. Linkage and linkage disequilibrium scan for autism loci in an extended pedigree from Finland. Hum. Molec. Genet. 18: 2912-2921, 2009. [PubMed: 19454485, images, related citations] [Full Text]

  44. King, I. F., Yandava, C. N., Mabb, A. M., Hsiao, J. S., Huang, H.-S., Pearson, B. L., Calabrese, J. M., Starmer, J., Parker, J. S., Magnuson, T., Chamberlain, S. J., Philpot, B. D., Zylka, M. J. Topoisomerases facilitate transcription of long genes linked to autism. Nature 501: 58-62, 2013. [PubMed: 23995680, images, related citations] [Full Text]

  45. Klauck, S. M., Munstermann, E., Bieber-Martig, B., Ruhl, D., Lisch, S., Schmotzer, G., Poustka, A., Poustka, F. Molecular genetic analysis of the FRM-1 gene in a large collection of autistic patients. Hum. Genet. 100: 224-229, 1997. [PubMed: 9254854, related citations] [Full Text]

  46. Kolevzon, A., Smith, C. J., Schmeidler, J., Buxbaum, J. D., Silverman, J. M. Familial symptom domains in monozygotic siblings with autism. Am. J. Med. Genet. 129B: 76-81, 2004. [PubMed: 15274045, related citations] [Full Text]

  47. Kong, A., Frigge, M. L., Masson, G., Besenbacher, S., Sulem, P., Magnusson, G., Gudjonsson, S. A., Sigurdsson, A., Jonasdottir, A., Jonasdottir, A., Wong, W. S. W., Sigurdsson, G., and 9 others. Rate of de novo mutations and the importance of father's age to disease risk. Nature 488: 471-475, 2012. [PubMed: 22914163, images, related citations] [Full Text]

  48. Krumm, N., O'Roak, B. J. Karakoc, E., Mohajeri, K., Nelson, B., Vives, L., Jacquemont, S., Munson, J., Bernier, R., Eichler, E. E. Transmission disequilibrium of small CNVs in simplex autism. Am. J. Hum. Genet. 93: 595-606, 2013. [PubMed: 24035194, images, related citations] [Full Text]

  49. Lainhart, J. E., Ozonoff, S., Coon, H., Krasny, L., Dinh, E., Nice, J., McMahon, W. Autism, regression, and the broader autism phenotype. Am. J. Med. Genet. 113: 231-237, 2002. [PubMed: 12439889, related citations] [Full Text]

  50. Lamb, J. A., Moore, J., Bailey, A., Monaco, A. P. Autism: recent molecular genetic advances. Hum. Molec. Genet. 9: 861-868, 2000. Note: Erratum: Hum. Molec. Genet. 9: 1461 only, 2000. [PubMed: 10767308, related citations] [Full Text]

  51. Levy, D., Ronemus, M., Yamrom, B., Lee, Y., Leotta, A., Kendall, J., Marks, S., Lakshmi, B., Pai, D., Ye, K., Buja, A., Krieger, A., Yoon, S., Troge, J., Rodgers, L., Iossifov, I., Wigler, M. Rare de novo and transmitted copy-number variation in autistic spectrum disorders. Neuron 70: 886-897, 2011. [PubMed: 21658582, related citations] [Full Text]

  52. Levy, S. E., Mandell, D. S., Schultz, R. T. Autism Lancet 374: 1627-1638, 2009. Note: Erratum: Lancet 378: 1546 only, 2011. [PubMed: 19819542, images, related citations] [Full Text]

  53. Lim, E. T., Raychaudhuri, S., Sanders, S. J., Stevens, C., Sabo, A., MacArthur, D. G., Neale, B. M., Kirby, A., Ruderfer, D. M., Fromer, M., Lek, M., Liu, L., and 29 others. Rare complete knockouts in humans: population distribution and significant role in autism spectrum disorders. Neuron 77: 235-242, 2013. [PubMed: 23352160, related citations] [Full Text]

  54. Liu, J., Nyholt, D. R., Magnussen, P., Parano, E., Pavone, P., Geschwind, D., Lord, C., Iversen, P., Hoh, J., Autism Genetic Resource Exchange Consortium, Ott, J., Gilliam, T. C. A genomewide screen for autism susceptibility loci. Am. J. Hum. Genet. 69: 327-340, 2001. Note: Erratum: Am. J. Hum. Genet. 69: 672 only, 2001. [PubMed: 11452361, images, related citations] [Full Text]

  55. Loirat, C., Bellanne-Chantelot, C., Husson, I., Deschenes, G., Guigonis, V., Chabane, N. Autism in three patients with cystic or hyperechogenic kidneys and chromosome 17q12 deletion. Nephrol. Dial. Transplant. 25: 3430-3433, 2010. [PubMed: 20587423, related citations] [Full Text]

  56. Lopreiato, J. O., Wulfsberg, E. A. A complex chromosome rearrangement in a boy with autism. J. Dev. Behav. Pediat. 13: 281-283, 1992. [PubMed: 1506468, related citations]

  57. Luo, R., Sanders, S. J., Tian, Y., Voineagu, I., Huang, N., Chu, S. H., Klei, L., Cai, C., Ou, J., Lowe, J. K., Hurles, M. E., Devlin, B., State, M. W., Geschwind, D. H. Genome-wide transcriptome profiling reveals the functional impact of rare de novo and recurrent CNVs in autism spectrum disorders. Am. J. Hum. Genet. 91: 38-55, 2012. [PubMed: 22726847, images, related citations] [Full Text]

  58. Ma, D. Q., Whitehead, P. L., Menold, M. M., Martin, E. R., Ashley-Koch, A. E., Mei, H., Ritchie, M. D., DeLong, G. R., Abramson, R. K., Wright, H. H., Cuccaro, M. L., Hussman, J. P., Gilbert, J. R., Pericak-Vance, M. A. Identification of significant association and gene-gene interaction of GABA receptor subunit genes in autism. Am. J. Hum. Genet. 77: 377-388, 2005. [PubMed: 16080114, related citations] [Full Text]

  59. Ma, D., Salyakina, D., Jaworski, J. M., Konidari, I., Whitehead, P. L., Andersen, A. N., Hoffman, J. D., Slifer, S. H., Hedges, D. J., Cukier, H. N., Griswold, A. J., McCauley, J. L., and 9 others. A genome-wide association study of autism reveals a common novel risk locus at 5p14.1. Ann. Hum. Genet. 73: 263-273, 2009. [PubMed: 19456320, images, related citations] [Full Text]

  60. Maestrini, E., Lai, C., Marlow, A., Matthews, N., Wallace, S., Bailey, A., Cook, E. H., Weeks, D. E., Monaco, A. P., International Molecular Genetic Study of Autism (IMGSA) Consortium. Serotonin transporter (5-HTT) and gamma-aminobutyric acid receptor subunit beta-3 (GABRB3) gene polymorphisms are not associated with autism in the IMGSA families. Am. J. Med. Genet. 88B: 492-496, 1999. [PubMed: 10490705, related citations] [Full Text]

  61. Marshall, C. R., Noor, A., Vincent, J. B., Lionel, A. C., Feuk, L., Skaug, J., Shago, M., Moessner, R., Pinto, D., Ren, Y., Thiruvahindrapduram, B., Fiebig, A., and 23 others. Structural variation of chromosomes in autism spectrum disorder. Am. J. Hum. Genet. 82: 477-488, 2008. [PubMed: 18252227, images, related citations] [Full Text]

  62. Miles, J. H., Takahashi, T. N., Hong, J., Munden, N., Flournoy, N., Braddock, S. R., Martin, R. A., Bocian, M. E., Spence, M. A., Hillman, R. E., Farmer, J. E. Development and validation of a measure of dysmorphology: useful for autism subgroup classification. Am. J. Med. Genet. 146A: 1101-1116, 2008. Note: Erratum: Am. J. Med. Genet. 146A; 2714 only, 2008. [PubMed: 18383511, related citations] [Full Text]

  63. Moessner, R., Marshall, C. R., Sutcliffe, J. S., Skaug, J., Pinto, D., Vincent, J., Zwaigenbaum, L., Fernandez, B., Roberts, W., Szatmari, P., Scherer, S. W. Contribution of SHANK3 mutations to autism spectrum disorder. Am. J. Hum. Genet. 81: 1289-1297, 2007. [PubMed: 17999366, images, related citations] [Full Text]

  64. Moreno-De-Luca, D., SGENE Consortium, Mulle, J. G., Simons Simplex Collection Genetics Consortium, Kaminsky, E. B., Sanders, S. J., GeneSTAR, Myers, S. M., Adam, M. P., Pakula, A. T., Eisenhauer, N. J., Uhas, K., and 26 others. Deletion 17q12 is a recurrent copy number variant that confers high risk of autism and schizophrenia. Am. J. Hum. Genet. 87: 618-630, 2010. Note: Erratum: Am. J. Hum. Genet. 88: 121 only, 2011. [PubMed: 21055719, images, related citations] [Full Text]

  65. Neale, B. M., Kou, Y., Liu, L., Ma'ayan, A., Samocha, K. E., Sabo, A., Lin, C.-F., Stevens, C., Wang, L.-S., Makarov, V., Polak, P., Yoon, S., and 47 others. Patterns and rates of exonic de novo mutations in autism spectrum disorders. Nature 485: 242-245, 2012. [PubMed: 22495311, images, related citations] [Full Text]

  66. O'Roak, B. J., Deriziotis, P., Lee, C., Vives, L., Schwartz, J. J., Girirajan, S., Karakoc, E., Mackenzie, A. P., Ng, S. B., Baker, C., Rieder, M. J., Nickerson, D. A., Bernier, R., Fisher, S. E., Shendure, J., Eichler, E. E. Exome sequencing in sporadic autism spectrum disorders identifies severe de novo mutations. Nature Genet. 43: 585-589, 2011. Note: Erratum: Nature Genet. 44: 471 only, 2012. [PubMed: 21572417, related citations] [Full Text]

  67. O'Roak, B. J., Vives, L., Fu, W., Egertson, J. D., Stanaway, I. B., Phelps, I. G., Carvill, G., Kumar, A., Lee, C., Ankenman, K., Munson, J., Hiatt, J. B., and 14 others. Multiplex targeted sequencing identifies recurrently mutated genes in autism spectrum disorders. Science 338: 1619-1622, 2012. [PubMed: 23160955, images, related citations] [Full Text]

  68. O'Roak, B. J., Vives, L., Girirajan, S., Karakoc, E., Krumm, N., Coe, B. P., Levy, R., Ko, A., Lee, C., Smith, J. D., Turner, E. H., Stanaway, I. B., and 11 others. Sporadic autism exomes reveal a highly interconnected protein network of de novo mutations. Nature 485: 246-250, 2012. [PubMed: 22495309, images, related citations] [Full Text]

  69. Parikshak, N. N., Swarup, V., Belgard, T. G., Irimia, M., Ramaswami, G., Gandal, M. J., Hartl, C., Leppa, V., de la Torre Ubieta, L., Huang, J., Lowe, J. K., Blencowe, B. J., Horvath, S., Geschwind, D. H. Genome-wide changes in lncRNA, splicing, and regional gene expression patterns in autism. Nature 540: 423-427, 2016. [PubMed: 27919067, related citations] [Full Text]

  70. Philippe, A., Martinez, M., Guilloud-Bataille, M., Gillberg, C., Rastam, M., Sponheim, E., Coleman, M., Zappella, M., Aschauer, H., van Maldergeme, L., Penet, C., Feingold, J., Brice, A., Leboyer, M., Paris Autism Research International Sibpair Study. Genome-wide scan for autism susceptibility genes. Hum. Molec. Genet. 8: 805-812, 1999. Note: Erratum: Hum. Molec. Genet. 8: 1353 only, 1999. [PubMed: 10196369, related citations] [Full Text]

  71. Pickles, A., Bolton, P., Macdonald, H., Bailey, A., Le Couteur, A., Sim, C.-H., Rutter, M. Latent-class analysis of recurrence risks for complex phenotypes with selection and measurement error: a twin and family history study of autism. Am. J. Hum. Genet. 57: 717-726, 1995. [PubMed: 7668301, related citations]

  72. Pinto, D., Delaby, E., Merico, D., Barbosa, M., Merikangas, A., Klei, L., Thiruvahindrapuram, B., Xu, X., Ziman, R., Wang, Z., Vorstman, J. A. S., Thompson, A., and 100 others. Convergence of genes and cellular pathways dysregulated in autism spectrum disorders. Am. J. Hum. Genet. 94: 677-694, 2014. [PubMed: 24768552, images, related citations] [Full Text]

  73. Pinto, D., Pagnamenta, A. T., Klei, L., Anney, R., Merico, D., Regan, R., Conroy, J., Magalhaes, T. R., Correia, C., Abrahams, B. S., Almeida, J., Bacchelli, E., and 165 others. Functional impact of global rare copy number variation in autism spectrum disorders. Nature 466: 368-372, 2010. [PubMed: 20531469, images, related citations] [Full Text]

  74. Piven, J., Tsai, G. C., Nehme, E., Coyle, J. T., Chase, G. A., Folstein, S. E. Platelet serotonin, a possible marker for familial autism. J. Autism Dev. Disord. 21: 51-59, 1991. [PubMed: 2037549, related citations]

  75. Poultney, C. S., Goldberg, A. P., Drapeau, E., Kou, Y., Harony-Nicolas, H., Kajiwara, Y., De Rubeis, S., Durand, S., Stevens, C., Rehnstrom, K., Palotie, A., Daly, M. J., Ma'ayan, A., Fromer, M., Buxbaum, J. D. Identification of small exonic CNV from whole-exome sequence data and application to autism spectrum disorder. Am. J. Hum. Genet. 93: 607-619, 2013. [PubMed: 24094742, images, related citations] [Full Text]

  76. Risch, N., Spiker, D., Lotspeich, L., Nouri, N., Hinds, D., Hallmayer, J., Kalaydjieva, L., McCague, P., Dimiceli, S., Pitts, T., Nguyen, L., Yang, J., and 19 others. A genomic screen of autism: evidence for a multilocus etiology. Am. J. Hum. Genet. 65: 493-507, 1999. [PubMed: 10417292, related citations] [Full Text]

  77. Ritvo, E. R., Freeman, B. J., Mason-Brothers, A. M., Mo, A., Ritvo, A. M. Concordance for the syndrome of autism in 40 pairs of affected twins. Am. J. Psychiat. 142: 74-77, 1985. [PubMed: 4038442, related citations] [Full Text]

  78. Ritvo, E. R., Spence, M. A., Freeman, B. J., Mason-Brothers, A., Mo, A., Marazita, M. L. Evidence for an autosomal recessive type of autism in 46 multiple incidence families. Am. J. Psychiat. 142: 187-192, 1985. [PubMed: 4038589, related citations] [Full Text]

  79. Roohi, J., Montagna, C., Tegay, D. H., Palmer, L. E., DeVincent, C., Pomeroy, J. C., Christian, S. L., Nowak, N., Hatchwell, E. Disruption of contactin 4 in three subjects with autism spectrum disorder. J. Med. Genet. 46: 176-182, 2009. [PubMed: 18349135, related citations] [Full Text]

  80. Sanders, S. J., Ercan-Sencicek, A. G., Hus, V., Luo, R., Murtha, M. T., Moreno-De-Luca, D., Chu, S. H., Moreau, M. P., Gupta, A. R., Thomson, S. A., Mason, C. E., Bilguvar, K., and 55 others. Multiple recurrent de novo CNVs, including duplications of the 7q11.23 Williams syndrome region, are strongly associated with autism. Neuron 70: 863-885, 2011. [PubMed: 21658581, images, related citations] [Full Text]

  81. Sanders, S. J., Murtha, M. T., Gupta, A. R., Murdoch, J. D., Raubeson, M. J., Willsey, A. J., Ercan-Sencicek, A. G., DiLullo, N. M., Parikshak, N. N., Stein, J. L., Walker, M. F., Ober, G. T., and 18 others. De novo mutations revealed by whole-exome sequencing are strongly associated with autism. Nature 485: 237-241, 2012. [PubMed: 22495306, images, related citations] [Full Text]

  82. Sandin, S., Lichtenstein, P., Kuja-Halkola, R., Larsson, H., Hultman, C. M., Reichenberg, A. The familial risk of autism. JAMA 311: 1770-1777, 2014. [PubMed: 24794370, images, related citations] [Full Text]

  83. Schaefer, G. B., Thompson, J. N., Jr., Bodensteiner, J. B., McConnell, J. M., Kimberling, W. J., Gay, C. T., Dutton, W. D., Hutchings, D. C., Gray, S. B. Hypoplasia of the cerebellar vermis in neurogenic syndromes. Ann. Neurol. 39: 382-385, 1996. [PubMed: 8602758, related citations] [Full Text]

  84. Schain, R. J., Freedman, D. X. Studies on 5-hydroxyindole metabolism in autistic and other mentally retarded children. J. Pediat. 58: 315-320, 1961. [PubMed: 13747230, related citations]

  85. Schellenberg, G. D., Dawson, G., Sung, Y. J., Estes, A., Munson, J., Rosenthal, E., Rothstein, J., Flodman, P., Smith, M., Coon, H., Leong, L., Yu, C.-E., Stodgell, C., Rodier, P. M., Spence, M. A., Minshew, N., McMahon, W. M., Wijsman, E. M. Evidence for multiple loci from a genome scan of autism kindreds. Molec. Psychiat. 11: 1049-1060, 2006. [PubMed: 16880825, related citations] [Full Text]

  86. Sebat, J., Lakshmi, B., Malhotra, D., Troge, J., Lese-Martin, C., Walsh, T., Yamrom, B., Yoon, S., Krasnitz, A., Kendall, J., Leotta, A., Pai, D., and 20 others. Strong association of de novo copy number mutations with autism. Science 316: 445-449, 2007. [PubMed: 17363630, images, related citations] [Full Text]

  87. Silverman, J. M., Smith, C. J., Schmeidler, J., Hollander, E., Lawlor, B. A., Fitzgerald, M., Buxbaum, J. D., Delaney, K., Galvin, P., Autism Genetic Research Exchange Consortium. Symptom domains in autism and related conditions: evidence for familiality. Am. J. Med. Genet. 114B: 64-73, 2002. [PubMed: 11840508, related citations] [Full Text]

  88. Smalley, S. L. Genetic influences in childhood-onset psychiatric disorders: autism and attention-deficit/hyperactivity disorder. Am. J. Hum. Genet. 60: 1276-1282, 1997. [PubMed: 9199546, related citations] [Full Text]

  89. Spence, M. A., Ritvo, E. R., Marazita, M. L., Funderburk, S. J., Sparkes, R. S., Freeman, B. J. Gene mapping studies with the syndrome of autism. Behav. Genet. 15: 1-13, 1985. [PubMed: 3985910, related citations]

  90. Tabuchi, K., Blundell, J., Etherton, M. R., Hammer, R. E., Liu, X., Powell, C. M., Sudhof, T. C. A neuroligin-3 mutation implicated in autism increases inhibitory synaptic transmission in mice. Science 318: 71-76, 2007. [PubMed: 17823315, images, related citations] [Full Text]

  91. Taylor, B., Miller, E., Lingam, R., Andrews, N., Simmons, A., Stowe, J. Measles, mumps, and rubella vaccination and bowel problems or developmental regression in children with autism: population study. Brit. Med. J. 324: 393-396, 2002. [PubMed: 11850369, related citations] [Full Text]

  92. Trikalinos, T. A., Karvouni, A., Zintzaras, E., Ylisaukko-oja, T., Peltonen, L., Jarvela, I., Ioannidis, J. P. A. A heterogeneity-based genome search meta-analysis for autism-spectrum disorders. Molec. Psychiat. 11: 29-36, 2006. [PubMed: 16189507, related citations] [Full Text]

  93. Vaags, A. K., Lionel, A. C., Sato, D., Goodenberger, M., Stein, Q. P., Curran, S., Ogilvie, C., Ahn, J. W., Drmic, I., Senman, L., Chrysler, C., Thompson, A., and 15 others. Rare deletions at the neurexin 3 locus in autism spectrum disorder. Am. J. Hum. Genet. 90: 133-141, 2012. [PubMed: 22209245, images, related citations] [Full Text]

  94. Vincent, J. B., Horike, S. I., Choufani, S., Paterson, A. D., Roberts, W., Szatmari, P., Weksberg, R., Fernandez, B., Scherer, S. W. An inversion inv(4)(p12-p15.3) in autistic siblings implicates the 4p GABA receptor gene cluster. J. Med. Genet. 43: 429-434, 2006. [PubMed: 16556609, images, related citations] [Full Text]

  95. Voineagu, I., Wang, X., Johnston, P., Lowe, J. K., Tian, Y., Horvath, S., Mill, J., Cantor, R. M., Blencowe, B. J., Geschwind, D. H. Transcriptomic analysis of autistic brain reveals convergent molecular pathology. Nature 474: 380-384, 2011. [PubMed: 21614001, images, related citations] [Full Text]

  96. Wang, K., Zhang, H., Ma, D., Bucan, M., Glessner, J. T., Abrahams, B. S., Salyakina, D., Imielinski, M., Bradfield, J. P., Sleiman, P. M. A., Kim, C. E., Hou, C., and 44 others. Common genetic variants on 5p14.1 associate with autism spectrum disorders. Nature 459: 528-533, 2009. [PubMed: 19404256, images, related citations] [Full Text]

  97. Weiss, L. A., Arking, D. E., Gene Discovery Project of Johns Hopkins & the Autism Consortium. A genome-wide linkage and association scan reveals novel loci for autism. Nature 461: 802-808, 2009. [PubMed: 19812673, images, related citations] [Full Text]

  98. Yonan, A. L., Alarcon, M., Cheng, R., Magnusson, P. K. E., Spence, S. J., Palmer, A. A., Grunn, A., Juo, S.-H. H., Terwilliger, J. D., Liu, J., Cantor, R. M., Geschwind, D. H., Gilliam, T. C. A genomewide screen of 345 families for autism-susceptibility loci. Am. J. Hum. Genet. 73: 886-897, 2003. [PubMed: 13680528, images, related citations] [Full Text]

  99. Yu, C.-E., Dawson, G., Munson, J., D'Souza, I., Osterling, J., Estes, A., Leutenegger, A.-L., Flodman, P., Smith, M., Raskind, W. H., Spence, M. A., McMahon, W., Wijsman, E. M., Schellenberg, G. D. Presence of large deletions in kindreds with autism. Am. J. Hum. Genet. 71: 100-115, 2002. [PubMed: 12058345, images, related citations] [Full Text]


Ada Hamosh - updated : 01/10/2017
Paul J. Converse - updated : 07/28/2016
Ada Hamosh - updated : 6/3/2016
Cassandra L. Kniffin - updated : 3/24/2016
Ada Hamosh - updated : 4/17/2015
Ada Hamosh - updated : 1/20/2015
Ada Hamosh - updated : 6/25/2014
Ada Hamosh - updated : 6/4/2014
Ada Hamosh - updated : 1/16/2014
Ada Hamosh - updated : 1/14/2014
Ada Hamosh - updated : 1/13/2014
Ada Hamosh - updated : 11/21/2013
Ada Hamosh - updated : 11/4/2013
Ada Hamosh - updated : 10/16/2013
Ada Hamosh - updated : 1/23/2013
Ada Hamosh - updated : 9/5/2012
Ada Hamosh - updated : 6/27/2012
Ada Hamosh - updated : 6/25/2012
Cassandra L. Kniffin - updated : 6/19/2012
Marla J. F. O'Neill - updated : 3/14/2012
Cassandra L. Kniffin - updated : 2/15/2012
George E. Tiller - updated : 11/14/2011
Cassandra L. Kniffin - updated : 11/8/2011
Ada Hamosh - updated : 10/4/2011
Ada Hamosh - updated : 7/26/2011
Cassandra L. Kniffin - updated : 4/18/2011
Ada Hamosh - updated : 11/8/2010
Ada Hamosh - updated : 8/17/2010
George E. Tiller - updated : 6/23/2010
Cassandra L. Kniffin - updated : 3/29/2010
Ada Hamosh - updated : 3/26/2010
George E. Tiller - updated : 2/22/2010
Cassandra L. Kniffin - updated : 12/10/2009
Ada Hamosh - updated : 8/24/2009
Ada Hamosh - updated : 8/17/2009
Cassandra L. Kniffin - updated : 1/5/2009
Cassandra L. Kniffin - updated : 8/20/2008
Cassandra L. Kniffin - updated : 4/18/2008
Ada Hamosh - updated : 10/26/2007
Ada Hamosh - updated : 5/30/2007
Cassandra L. Kniffin - updated : 3/12/2007
Cassandra L. Kniffin - updated : 8/29/2006
Cassandra L. Kniffin - updated : 8/18/2006
Cassandra L. Kniffin - updated : 6/13/2006
John Logan Black, III - updated : 4/6/2006
Marla J. F. O'Neill - updated : 10/6/2005
Cassandra L. Kniffin - updated : 6/28/2005
Victor A. McKusick - updated : 3/23/2005
Victor A. McKusick - updated : 3/11/2005
John Logan Black, III - updated : 3/1/2005
Victor A. McKusick - updated : 11/12/2004
Cassandra L. Kniffin - updated : 5/24/2004
Cassandra L. Kniffin - reorganized : 5/17/2004
Cassandra L. Kniffin - updated : 5/6/2004
John Logan Black, III - updated : 3/11/2004
Victor A. McKusick - updated : 12/1/2003
Victor A. McKusick - updated : 8/15/2003
Victor A. McKusick - updated : 2/28/2003
Victor A. McKusick - updated : 11/27/2002
Victor A. McKusick - updated : 11/14/2002
John Logan Black, III - updated : 8/14/2002
Victor A. McKusick - updated : 8/2/2002
Victor A. McKusick - updated : 4/12/2002
Victor A. McKusick - updated : 2/21/2002
Victor A. McKusick - updated : 2/14/2002
Victor A. McKusick - updated : 2/4/2002
Ada Hamosh - updated : 1/30/2002
Victor A. McKusick - updated : 1/22/2002
Michael B. Petersen - updated : 12/5/2001
Victor A. McKusick - updated : 11/27/2001
Victor A. McKusick - updated : 10/25/2001
George E. Tiller - updated : 10/2/2001
Victor A. McKusick - updated : 9/7/2001
Victor A. McKusick - updated : 10/3/2000
George E. Tiller - updated : 5/1/2000
Victor A. McKusick - updated : 1/11/2000
Victor A. McKusick - updated : 5/17/1999
Orest Hurko - updated : 3/24/1999
Victor A. McKusick - updated : 4/24/1998
Victor A. McKusick - updated : 9/12/1997
Victor A. McKusick - updated : 6/17/1997
Orest Hurko - updated : 5/6/1996
Orest Hurko - updated : 8/2/1995
Creation Date:
Victor A. McKusick : 6/3/1986
carol : 03/02/2017
carol : 02/28/2017
carol : 02/28/2017
carol : 01/11/2017
alopez : 01/10/2017
mgross : 07/28/2016
carol : 07/09/2016
alopez : 6/3/2016
alopez : 4/12/2016
carol : 3/24/2016
ckniffin : 3/24/2016
alopez : 4/17/2015
alopez : 1/20/2015
carol : 8/21/2014
alopez : 8/20/2014
mcolton : 8/19/2014
ckniffin : 8/14/2014
alopez : 6/25/2014
alopez : 6/4/2014
carol : 4/1/2014
alopez : 1/16/2014
alopez : 1/14/2014
alopez : 1/13/2014
alopez : 11/21/2013
mcolton : 11/13/2013
alopez : 11/4/2013
alopez : 10/16/2013
tpirozzi : 10/1/2013
tpirozzi : 10/1/2013
tpirozzi : 10/1/2013
tpirozzi : 10/1/2013
carol : 9/12/2013
ckniffin : 5/16/2013
carol : 4/19/2013
terry : 3/15/2013
alopez : 1/24/2013
terry : 1/23/2013
terry : 10/2/2012
alopez : 9/6/2012
terry : 9/5/2012
terry : 7/10/2012
terry : 7/6/2012
terry : 7/6/2012
alopez : 7/3/2012
alopez : 7/2/2012
alopez : 7/2/2012
terry : 6/27/2012
terry : 6/25/2012
carol : 6/20/2012
ckniffin : 6/19/2012
carol : 6/7/2012
ckniffin : 6/6/2012
alopez : 6/6/2012
terry : 3/16/2012
carol : 3/14/2012
carol : 2/21/2012
ckniffin : 2/15/2012
carol : 11/18/2011
terry : 11/14/2011
carol : 11/9/2011
ckniffin : 11/8/2011
alopez : 10/12/2011
terry : 10/4/2011
terry : 9/28/2011
alopez : 8/16/2011
terry : 7/26/2011
alopez : 6/16/2011
wwang : 4/21/2011
ckniffin : 4/18/2011
carol : 1/21/2011
alopez : 1/13/2011
ckniffin : 1/12/2011
alopez : 1/6/2011
alopez : 11/10/2010
terry : 11/8/2010
terry : 11/8/2010
alopez : 8/20/2010
terry : 8/17/2010
wwang : 7/6/2010
terry : 6/23/2010
alopez : 6/14/2010
alopez : 6/10/2010
ckniffin : 6/9/2010
alopez : 5/21/2010
wwang : 4/30/2010
ckniffin : 4/19/2010
wwang : 4/6/2010
ckniffin : 3/29/2010
alopez : 3/26/2010
wwang : 2/25/2010
terry : 2/22/2010
carol : 2/18/2010
wwang : 12/10/2009
wwang : 12/3/2009
ckniffin : 11/5/2009
wwang : 9/22/2009
alopez : 8/24/2009
alopez : 8/24/2009
terry : 8/17/2009
wwang : 1/7/2009
ckniffin : 1/5/2009
wwang : 8/26/2008
ckniffin : 8/20/2008
alopez : 6/6/2008
wwang : 4/24/2008
ckniffin : 4/18/2008
alopez : 3/21/2008
terry : 3/19/2008
wwang : 1/11/2008
terry : 1/3/2008
alopez : 11/1/2007
terry : 10/26/2007
carol : 10/10/2007
alopez : 6/14/2007
terry : 5/30/2007
alopez : 5/16/2007
carol : 5/14/2007
carol : 5/14/2007
ckniffin : 5/10/2007
ckniffin : 3/12/2007
carol : 3/6/2007
ckniffin : 3/5/2007
carol : 11/28/2006
carol : 11/27/2006
wwang : 9/7/2006
ckniffin : 8/29/2006
wwang : 8/25/2006
ckniffin : 8/18/2006
wwang : 6/16/2006
ckniffin : 6/13/2006
wwang : 4/10/2006
terry : 4/6/2006
carol : 1/18/2006
terry : 12/21/2005
carol : 12/12/2005
terry : 11/10/2005
wwang : 10/18/2005
alopez : 10/17/2005
terry : 10/6/2005
ckniffin : 6/28/2005
ckniffin : 5/23/2005
tkritzer : 3/25/2005
ckniffin : 3/24/2005
ckniffin : 3/24/2005
terry : 3/23/2005
wwang : 3/15/2005
terry : 3/11/2005
carol : 3/10/2005
tkritzer : 3/1/2005
alopez : 11/18/2004
terry : 11/12/2004
tkritzer : 5/28/2004
ckniffin : 5/24/2004
ckniffin : 5/19/2004
carol : 5/18/2004
ckniffin : 5/18/2004
carol : 5/17/2004
ckniffin : 5/17/2004
ckniffin : 5/6/2004
carol : 3/23/2004
joanna : 3/17/2004
terry : 3/11/2004
tkritzer : 12/8/2003
terry : 12/1/2003
alopez : 8/19/2003
terry : 8/15/2003
terry : 8/15/2003
joanna : 5/12/2003
ckniffin : 4/1/2003
ckniffin : 4/1/2003
carol : 3/31/2003
tkritzer : 3/6/2003
terry : 2/28/2003
tkritzer : 12/5/2002
tkritzer : 12/2/2002
terry : 11/27/2002
carol : 11/20/2002
carol : 11/20/2002
terry : 11/14/2002
terry : 11/14/2002
carol : 8/14/2002
tkritzer : 8/7/2002
tkritzer : 8/7/2002
tkritzer : 8/5/2002
terry : 8/2/2002
alopez : 4/26/2002
cwells : 4/19/2002
terry : 4/12/2002
mgross : 2/25/2002
mgross : 2/25/2002
terry : 2/21/2002
carol : 2/21/2002
cwells : 2/21/2002
cwells : 2/15/2002
terry : 2/14/2002
carol : 2/11/2002
terry : 2/4/2002
alopez : 2/4/2002
terry : 1/30/2002
carol : 1/30/2002
terry : 1/22/2002
cwells : 12/5/2001
alopez : 11/30/2001
terry : 11/27/2001
carol : 10/25/2001
terry : 10/25/2001
cwells : 10/10/2001
cwells : 10/2/2001
cwells : 9/20/2001
cwells : 9/13/2001
alopez : 9/7/2001
alopez : 9/7/2001
carol : 5/3/2001
terry : 10/5/2000
terry : 10/3/2000
alopez : 5/1/2000
mgross : 2/7/2000
terry : 1/11/2000
mgross : 6/4/1999
mgross : 6/4/1999
mgross : 5/25/1999
terry : 5/17/1999
carol : 3/24/1999
dholmes : 5/12/1998
dholmes : 5/11/1998
carol : 4/24/1998
terry : 4/14/1998
mark : 9/19/1997
terry : 9/12/1997
terry : 9/10/1997
terry : 9/10/1997
mark : 6/18/1997
terry : 6/17/1997
terry : 6/12/1997
terry : 6/12/1997
mark : 5/6/1996
mark : 5/6/1996
terry : 4/30/1996
mark : 9/10/1995
terry : 4/19/1995
jason : 6/22/1994
supermim : 3/16/1992
carol : 3/7/1992

% 209850

AUTISM


Alternative titles; symbols

AUTISTIC DISORDER


Other entities represented in this entry:

AUTISM, SUSCEPTIBILITY TO, 1, INCLUDED; AUTS1, INCLUDED
AUTISM SPECTRUM DISORDER, INCLUDED; ASD, INCLUDED

SNOMEDCT: 43614003, 408857007, 408856003;   ICD10CM: F84.0;   ICD9CM: 299.0;   DO: 12849;  


Cytogenetic location: 7q22     Genomic coordinates (GRCh38): 7:98,400,000-107,800,000


Gene-Phenotype Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
7q22 {Autism susceptibility 1} 209850 Multifactorial; Isolated cases 2

TEXT

Description

Autism, the prototypic pervasive developmental disorder (PDD), is usually apparent by 3 years of age. It is characterized by a triad of limited or absent verbal communication, a lack of reciprocal social interaction or responsiveness, and restricted, stereotypic, and ritualized patterns of interests and behavior (Bailey et al., 1996; Risch et al., 1999). 'Autism spectrum disorder,' sometimes referred to as ASD, is a broader phenotype encompassing the less severe disorders Asperger syndrome (see ASPG1; 608638) and pervasive developmental disorder, not otherwise specified (PDD-NOS). 'Broad autism phenotype' includes individuals with some symptoms of autism, but who do not meet the full criteria for autism or other disorders. Mental retardation coexists in approximately two-thirds of individuals with ASD, except for Asperger syndrome, in which mental retardation is conspicuously absent (Jones et al., 2008). Genetic studies in autism often include family members with these less stringent diagnoses (Schellenberg et al., 2006).

Levy et al. (2009) provided a general review of autism and autism spectrum disorder, including epidemiology, characteristics of the disorder, diagnosis, neurobiologic hypotheses for the etiology, genetics, and treatment options.

Genetic Heterogeneity of Autism

Autism is considered to be a complex multifactorial disorder involving many genes. Accordingly, several loci have been identified, some or all of which may contribute to the phenotype. Included in this entry is AUTS1, which has been mapped to chromosome 7q22.

Other susceptibility loci include AUTS3 (608049), which maps to chromosome 13q14; AUTS4 (608636), which maps to chromosome 15q11; AUTS5 (606053), which maps to chromosome 2q; AUTS6 (609378), which maps to chromosome 17q11; AUTS7 (610676), which maps to chromosome 17q21; AUTS8 (607373), which maps to chromosome 3q25-q27; AUTS9 (611015), which maps to chromosome 7q31; AUTS10 (611016), which maps to chromosome 7q36; AUTS11 (610836), which maps to chromosome 1q41; AUTS12 (610838), which maps to chromosome 21p13-q11; AUTS13 (610908), which maps to chromosome 12q14; AUTS14A (611913), which has been found in patients with a deletion of a region of 16p11.2; AUTS14B (614671), which has been found in patients with a duplication of a region of 16p11.2; AUTS15 (612100), associated with mutation in the CNTNAP2 gene (604569) on chromosome 7q35-q36; AUTS16 (613410), associated with mutation in the SLC9A9 gene (608396) on chromosome 3q24; AUTS17 (613436), associated with mutation in the SHANK2 gene (603290) on chromosome 11q13; and AUTS18 (615032), associated with mutation in the CHD8 gene (610528). (NOTE: the symbol 'AUTS2' has been used to refer to a gene on chromosome 7q11 (KIAA0442; 607270) and therefore is not used as a part of this autism locus series.)

There are several X-linked forms of autism susceptibility: AUTSX1 (300425), associated with mutations in the NLGN3 gene (300336); AUTSX2 (300495), associated with mutations in NLGN4 (300427); AUTSX3 (300496), associated with mutations in MECP2 (300005); AUTSX4 (300830), associated with variation in the region on chromosome Xp22.11 containing the PTCHD1 gene (300828); AUTSX5 (300847), associated with mutations in the RPL10 gene (312173); and AUTSX6 (300872), associated with mutation in the TMLHE gene (300777).

Folstein and Rosen-Sheidley (2001) reviewed the genetics of autism.


Clinical Features

The DSM-IV (American Psychiatric Association, 1994) specifies several diagnostic criteria for autism. In general, patients with autism exhibit qualitative impairment in social interaction, as manifest by impairment in the use of nonverbal behaviors such as eye-to-eye gaze, facial expression, body postures, and gestures, failure to develop appropriate peer relationships, and lack of social sharing or reciprocity. Patients have impairments in communication, such as a delay in, or total lack of, the development of spoken language. In patients who do develop adequate speech, there remains a marked impairment in the ability to initiate or sustain a conversation, as well as stereotyped or idiosyncratic use of language. Patients also exhibit restricted, repetitive and stereotyped patterns of behavior, interests, and activities, including abnormal preoccupation with certain activities and inflexible adherence to routines or rituals.

In his pioneer description of infantile autism, Kanner (1943) defined the disorder as 'an innate inability to form the usual, biologically provided affective contact with people.' Kanner (1943) noted that in most cases the child's behavior was abnormal from early infancy, and he suggested the presence of an inborn, presumably genetic, defect.

In a review, Smalley (1997) stated that mental retardation is said to be present in approximately 75% of cases of autism, seizures in 15 to 30% of cases, and electroencephalographic abnormalities in 20 to 50% of cases. In addition, approximately 15 to 37% of cases of autism have a comorbid medical condition, including 5 to 14% with a known genetic disorder or chromosomal anomaly. The 4 most common associations include fragile X syndrome (300624), tuberous sclerosis (see 191100), 15q duplications (AUTS4; 608636), and untreated phenylketonuria (PKU; 261600). Significant associations at a phenotypic level may reflect disruptions in a common neurobiologic pathway, common susceptibility genes, or genes in linkage disequilibrium.

The autism spectrum disorder shows a striking sex bias, with a male:female ratio of idiopathic autism estimated at 4-10:1, and with an increase in this ratio as the intelligence of the affected individuals increases (Folstein and Rosen-Sheidley, 2001).

Lainhart et al. (2002) stated that approximately 20% of children with autism appear to have relatively normal development during the first 12 to 24 months of life. This period of relative normalcy gradually or suddenly ends and is followed by a period of regression, characterized most prominently by a significant loss of language skills, after which the full autism syndrome becomes evident.

Rarely, children with autism may exhibit hyperlexia, or precocious reading (238350). Among a group of 66 children with pervasive developmental disorder, Burd et al. (1985) identified 4 with hyperlexia.

Cohen et al. (2005) discussed several genetic disorders consistently associated with autism, including fragile X syndrome, tuberous sclerosis, Angelman syndrome (105830), Down syndrome (190685), Sanfilippo syndrome (252900), Rett syndrome (312750) and other MECP2-related disorders, phenylketonuria, Smith-Magenis syndrome (SMS; 182290), 22q13 deletion syndrome (606232), Cohen syndrome (COH1; 216550), adenylosuccinate lyase deficiency (103050), and Smith-Lemli-Opitz syndrome (SLOS; 270400).

Miles et al. (2008) presented an expert-derived consensus measure of dysmorphic features often observed in patients with autism. The goal was to enable clinicians not trained in dysmorphology to use this classification system to identify and further subphenotype patients with autism. The measure includes 12 body areas that can be scored to arrive at a determination of dysmorphic or nondysmorphic. The body areas include stature, hair growth pattern, ear structure and placement, nose size, facial structure, philtrum, mouth and lips, teeth, hands, fingers and thumbs, nails, and feet. The model performed with 81 to 82% sensitivity and 95 to 99% specificity.


Inheritance

Folstein and Rutter (1977) reported that there had been no recorded cases of an autistic child having an overtly autistic parent; however, they noted that autistic persons rarely marry and rarely give birth. Folstein and Rutter (1977) stated that about 2% of sibs are affected, and that speech delay is common in the sibships containing autistic children. In a study of 21 same-sex twin pairs, 11 monozygotic (MZ) and 10 dizygotic (DZ), in which at least 1 had infantile autism, Folstein and Rutter (1977) found 36% concordance among the MZ twins and no concordance among the DZ twins. The concordance for cognitive abnormalities was 82% for MZ pairs and 10% for DZ pairs. In 12 of the 17 pairs discordant for autism, a biologic hazard liable to cause brain damage was identified. The authors concluded that brain injury in infancy may lead to autism on its own or in combination with a genetic predisposition. An inheritance pattern was not suggested.

In 40 pairs of twins, Ritvo et al. (1985) found a concordance rate for autism of 23.5% in dizygotic twins (4 of 17 pairs) and 95.7% in monozygotic twins (22 of 23 pairs). Ritvo et al. (1985) ascertained 46 families with 2 (n = 41) or 3 (n = 5) sibs with autism. Classic segregation analysis yielded a maximum likelihood estimate of the segregation ratio of 0.19 +/- 0.07, a value significantly different from 0.50 expected of an autosomal dominant trait and not significantly different from 0.25 expected of a recessive trait. The authors rejected a polygenic threshold model and suggested autosomal recessive inheritance.

Using the Utah Genealogical Database, Jorde et al. (1990) determined kinship for all possible pairs of autistic subjects. The average kinship coefficient for autistic subjects and controls showed a strong tendency for autism to cluster in families. However, the familial aggregation was confined exclusively to sib pairs and did not extend to more distant relatives. The authors concluded that the findings excluded recessive inheritance, since the autosomal recessive hypothesis would predict several first-cousin pairs, of which none were found. The rapid fall off in risk to relatives, as well as the sib risk of 4.5%, was consistent with multifactorial causation.

By analysis of 99 autistic probands and their families, Bolton et al. (1994) found an increased familial risk for both autism and more broadly defined pervasive developmental disorders in sibs, 2.9% and 2.9%, respectively, which is about 75 times higher than the risk in the general population.

In 27 same-sex pairs of monozygotic twins and 20 dizygotic twins, Bailey et al. (1995) found that 60% of monozygotic pairs were concordant for autism compared to 0% of dizygotic pairs. When they considered a broader spectrum of related cognitive or social abnormalities, 92% of monozygotic pairs were concordant compared to 10% of dizygotic pairs. The high concordance in monozygotes indicated a high degree of genetic control, and the rapid fall off of concordance in dizygotes suggested to Bailey et al. (1995) a multilocus, epistatic model. In the nonconcordant monozygotic pairs, there was a significantly higher incidence of obstetric complications, which the authors attributed to prenatal developmental anomalies, as evidenced by the very high incidence of minor congenital anomalies in the affected twins. They also reported an association of autism with increased head circumference.

In a sample of families selected because each had exactly 2 affected sibs, Greenberg et al. (2001) observed a remarkably high proportion of affected twin pairs, both MZ and DZ. Of 166 affected sib pairs, 30 (12 MZ, 17 DZ, and 1 of unknown zygosity) were twin pairs. Deviation from expected values was statistically significant; in a similarly ascertained sample of individuals with type I diabetes (222100), there was no deviation from expected values. Greenberg et al. (2001) noted that to ascribe the excess of twins with autism solely to ascertainment bias would require very large ascertainment factors; e.g., affected twin pairs would need to be approximately 10 times more likely to be ascertained than affected nontwin sib pairs. In the extreme situation of 'complete stoppage,' a form of ascertainment bias in which parents stop having children after the birth of their first affected child, the only families to have an affected sib pair would be those with an affected twin pair, or affected triplets. The authors suggested that risk factors related to twinning or to fetal development or other factors, genetic or nongenetic, in the parents may contribute to autism. Hallmayer et al. (2002) presented information refuting the suggestion that the twinning process itself is an important risk factor in the development of autism.

Silverman et al. (2002) analyzed 3 autistic symptom domains, social interaction, communication, and repetitive behaviors, and variability in the presence and emergence of useful phrase speech in 212 multiply affected sibships with autism. They found that the variance within sibships was reduced for the repetitive behavior domain and for delays in and the presence of useful phrase speech. These features and the nonverbal communication subdomain provided evidence of familiality when only the diagnosis of autism was considered for defining multiply affected sibships.

Kolevzon et al. (2004) studied specific features of autism for decreased variance in 16 families with monozygotic twins concordant for autism. Using regression analysis, they demonstrated significant aggregation of symptoms in monozygotic twins for 2 autistic symptom domains: impairment in communication and in social interaction. Kolevzon et al. (2004) stated that selecting probands according to specific features known to show reduced variance within families may provide more homogeneous samples for genetic analysis.

Awadalla et al. (2010) hypothesized that deleterious de novo mutations may play a role in cases of ASD and schizophrenia (181500), 2 etiologically heterogeneous disorders with significantly reduced reproductive fitness. Awadalla et al. (2010) presented a direct measure of the de novo mutation rate (mu) and selective constraints from de novo mutations estimated from a deep resequencing dataset generated from a large cohort of ASD and schizophrenia cases (n = 285) and population control individuals (n = 285) with available parental DNA. A survey of approximately 430 Mb of DNA from 401 synapse-expressed genes across all cases and 25 Mb of DNA in controls found 28 candidate de novo mutations, 13 of which were cell line artifacts. Awadalla et al. (2010) calculated a direct neutral mutation rate (1.36 x 10(-8)) that was similar to previous indirect estimates, but they observed a significant excess of potentially deleterious de novo mutations in ASD and schizophrenia individuals. Awadalla et al. (2010) concluded that their results emphasized the importance of de novo mutations as genetic mechanisms in ASD and schizophrenia and the limitations of using DNA from archived cell lines to identify functional variants.

Sandin et al. (2014) examined the familial risk of autism in a population-based cohort of 2,049,973 Swedish children born from 1982 to 2006. They identified 37,570 twin pairs; 2,642,064 full-sib pairs; 432,281 maternal and 445,531 paternal half-sib pairs; and 5,799,875 cousin pairs. Diagnoses of ASD to December 31, 2009 were ascertained. Exposure refers to the presence or absence of autism in a sib. In the sample, 14,516 children were diagnosed with ASD, of whom 5,689 had autistic disorder. The relative recurrence risk (RRR) and rate per 100,000 person-years for ASD among monozygotic twins was estimated to be 153.0 (95% CI, 56.7-412.8; rate, 6,274 for exposed vs 27 for unexposed); for dizygotic twins, 8.2 (95% CI, 3.7-18.1; rate, 805 for exposed vs 55 for unexposed); for full sibs, 10.3 (95% CI, 9.4-11.3; rate, 829 for exposed vs 49 for unexposed); for maternal half sibs, 3.3 (95% CI, 2.6-4.2; rate, 492 for exposed vs 94 for unexposed); for paternal half sibs, 2.9 (95% CI, 2.2-3.7; rate, 371 for exposed vs 85 for unexposed); and for cousins, 2.0 (95% CI, 1.8-2.2; rate, 155 for exposed vs 49 for unexposed). The RRR pattern was similar for autistic disorder but of slightly higher magnitude. Sandin et al. (2014) found support for a disease etiology including only additive genetic and nonshared environmental effects. The ASD heritability was estimated to be 0.50 (95% CI, 0.45-0.56) and the autistic disorder heritability was estimated to 0.54 (95% CI, 0.44-0.64). Sandin et al. (2014) concluded that among children in Sweden, the individual risk of ASD and autistic disorder increased with increasing genetic relatedness.

Iossifov et al. (2014) applied whole-exome sequencing to more than 2,500 simplex families each having a child with an autistic spectrum disorder. By comparing affected to unaffected sibs, Iossifov et al. (2014) showed that 13% of de novo missense mutations and 43% of de novo likely gene-disrupting mutations contribute to 12% and 9% of diagnoses, respectively. Including CNVs, coding de novo mutations contribute to about 30% of all simplex and 45% of female diagnoses.


Mapping

AUTS1 Locus on Chromosome 7q22

By analyzing 125 autistic sib pairs, the International Molecular Genetic Study of Autism Consortium (2001) found a maximum multipoint lod score of 2.15 at marker D7S477 on chromosome 7q22, whereas analysis of 153 sib pairs generated a maximum multipoint lod score of 3.37. Linkage disequilibrium mapping identified 2 regions of association: one was under the peak of linkage, the other was 27 cM distal. In another study, the International Molecular Genetic Study of Autism Consortium (2001) found a multipoint maximum lod score of 3.20 at marker D7S477. They also detected a multipoint maximum lod score of 4.80 at marker D2S188 on chromosome 2q.

In 12 of 105 families with 2 or more sibs affected with autism, Yu et al. (2002) identified deletions ranging from 5 to more than 260 kb. One family had complex deletions at marker D7S630 on 7q21-q22, 3 families had different deletions at D7S517 on 7p, and another 3 families had different deletions at D8S264 on 8p. Yu et al. (2002) suggested that autism susceptibility alleles may cause the deletions by inducing errors during meiosis.

In a metaanalysis of 9 published genome scans on autism or autism spectrum disorders, Trikalinos et al. (2006) found evidence for significant linkage to 7q22-q32, confirming the findings of previous studies. The flanking region 7q32-qter reached a less stringent threshold for significance.

Genetic Heterogeneity

Using findings from a family study of autism and a similar study of twins, Pickles et al. (1995) concluded that autism has a multiple locus mode of inheritance involving 3 loci. Risch et al. (1999) performed a 2-stage genomewide screen of 2 groups of families with autism: 90 families comprising 97 affected sib pairs (ASPs) and 49 families with 50 affected sib pairs. Unaffected sibs, which provided 51 discordant sib pairs (DSPs) for the initial screen and 29 for the follow-up, were included as controls. There was a slightly increased identity by descent (IBD) in the ASPs (sharing of 51.6%) compared with the DSPs (sharing of 50.8%). The results were considered most compatible with a model specifying a large number of loci, perhaps 15 or more, and less compatible with models specifying 10 or fewer loci. The largest lod scores obtained were for a marker on 1p yielding a maximum multipoint lod score of 2.15, and on 17p, yielding a maximum lod score of 1.21.

In 51 multiplex families with autism, Philippe et al. (1999) used nonparametric linkage analysis to perform a genomewide screen with 264 microsatellite markers. By 2-point and multipoint affected sib-pair analyses, 11 regions gave nominal P values of 0.05 or lower. Philippe et al. (1999) observed overlap of 4 of these regions with regions on 2q, 7q, 6p, and 19p that had been identified by the earlier genomewide scan of autism conducted by the International Molecular Genetic Study of Autism Consortium (1998). The most significant multipoint linkage was close to marker D6S283 (maximum lod score = 2.23, p = 0.0013).

Smalley (1997) reported on the status of linkage studies in autism. Lamb et al. (2000) reviewed chromosomal aberrations, candidate gene studies, and linkage studies of autism.

Liu et al. (2001) genotyped 335 microsatellite markers in 110 multiplex families with autism. All families included at least 2 affected sibs, at least 1 of whom had autism; the remaining affected sibs carried diagnoses of either Asperger syndrome or pervasive developmental disorder. Affected sib-pair analysis yielded multipoint maximum lod scores that reached the accepted threshold for suggestive linkage on chromosomes 5, X, and 19. Further analysis yielded impressive evidence for linkage of autism and autism-spectrum disorders to markers on chromosomes 5 and 8, with suggestive linkage to a marker on chromosome 19.

Yonan et al. (2003) followed up on previously reported genomewide screens for autism performed by Liu et al. (2001) and Alarcon et al. (2002) showing suggestive evidence for linkage of autism spectrum disorders on chromosomes 5, 8, 16, 19, and X, and nominal evidence on several additional chromosomes. In their analysis, Yonan et al. (2003) increased the sample size 3-fold. Multipoint maximum lod scores obtained from affected sib-pair analysis of all 345 families yielded suggestive evidence for linkage on chromosomes 17, 5, 11, 4, and 8 (listed in order of MLS). The most significant findings were an MLS of 2.83 on 17q11 (AUTS6; 609378) and an MLS of 2.54 on 5p.

The genetic architecture of autism spectrum disorders (ASDs) is complex, requiring large samples to overcome heterogeneity. The Autism Genome Project Consortium (2007) broadened coverage and sample size relative to other studies of ASDs by using Affymetrix 10K SNP arrays and 1,181 families with at least 2 affected individuals, performing the largest linkage scan to that time while also analyzing copy number variation in these families. Linkage and copy number variation analyses implicated 11p13-p12 and neurexins, respectively, among other candidate loci. Neurexins teamed with previously implicated neuroligins for glutamatergic synaptogenesis, highlighting glutamate-related genes as promising candidates for contributing to ASDs. See neuroligin-3 (NLGN3; 300336) and neuroligin-4 (NLGN4; 300427).

Wang et al. (2009) presented the results of a genomewide association study of ASDs on a cohort of 780 families (3,101 subjects) with affected children, and a second cohort of 1,204 affected subjects and 6,491 control subjects, all of whom were of European ancestry. Six SNPs on chromosome 5p14.1 between cadherin-10 (CDH10; 604555) and cadherin-9 (CDH9; 609974), 2 genes encoding neuronal cell adhesion molecules, revealed strong association signals, with the most significant SNP being rs4307059 (p = 3.4 x 10(-8), odds ratio = 1.19). These signals were replicated in 2 independent cohorts, with combined P values ranging from 7.4 x 10(-8) to 2.1 x 10(-10). The authors concluded that their results implicated neuronal cell adhesion molecules in the pathogenesis of ASDs.

In a genomewide association study of 438 Caucasian families including 1,390 individuals with autism, Ma et al. (2009) found evidence for linkage to chromosome 5p14.1. Validation in an additional cohort of 2,390 samples from 457 families showed that 8 SNPs on chromosome 5p14.1 were significantly associated with autism (p values ranging from 3.24 x 10(-4) to 3.40 x 10(-6)). The most significant linkage was with rs10038113.

Using array CGH analysis, Roohi et al. (2009) identified a chromosome 3 copy number variation (CNV) disrupting the CNTN4 gene (607280) in 3 of 92 individuals with autism spectrum disorder. Two sibs had a deletion, and another unrelated individual had a duplication; both variations were inherited from an unaffected father. A third affected sib of the familial cases did not carry the deletion, suggesting incomplete penetrance or that he had a different disorder. The changes resulted from Alu-mediated unequal recombinations.

Glessner et al. (2009) presented the results from a whole-genome CNV study on a cohort of 859 ASD cases and 1,409 healthy children of European ancestry who were genotyped with approximately 550,000 SNP markers, in an attempt to comprehensively identify CNVs conferring susceptibility to ASDs. Positive findings were evaluated in an independent cohort of 1,336 ASD cases and 1,110 controls of European ancestry. Glessner et al. (2009) confirmed known associations, such as that with NRXN1 (600565) and CNTN4 (Roohi et al., 2009), in addition to several novel susceptibility genes encoding neuronal cell adhesion molecules, including NLGN1 (600568) and ASTN2 (612856), that were enriched with CNVs in ASD cases compared to controls (P = 9.5 x 10(-3)). Furthermore, CNVs within or surrounding genes involved in the ubiquitin pathways, including UBE3A (601623), PARK2 (602544), RFWD2 (608067), and FBXO40 (609107), were affected by CNVs not observed in controls (p = 3.3 x 10(-3)). Glessner et al. (2009) also identified duplications 55 kb upstream of complementary DNA AK123120 (p = 3.6 x 10(-6)). Glessner et al. (2009) concluded that although these variants may be individually rare, they appear to target genes involved in neuronal cell-adhesion or ubiquitin degradation, indicating that these 2 important gene networks expressed within the central nervous system (CNS) may contribute to genetic susceptibility to ASD.

Weiss et al. (2009) initiated a linkage and association mapping study using half a million genomewide SNPs in a common set of 1,031 multiplex autism families (1,553 affected offspring). They identified regions of suggestive and significant linkage on chromosomes 6q27 and 20p13, respectively. Initial analysis did not yield genomewide significant associations; however, genotyping of top hits in additional families revealed an SNP on chromosome 5p15 (rs10513025) between SEMA5A (609297) and TAS2R1 (604796) that was significantly associated with autism (p = 2.0 x 10(-7)). Weiss et al. (2009) also demonstrated that expression of SEMA5A is reduced in brains from autistic patients, further implicating SEMA5A as an autism susceptibility gene.

Kilpinen et al. (2009) carried out a genomewide microsatellite-based scan of a unique extended Finnish autism pedigree comprised of 20 families with verified genealogic links reaching back to the 17th century. Linkage analysis and fine mapping revealed significant results for SNPs (rs4806893, rs216283, and rs216276) on chromosome 19p13.3 in close proximity to TLE2 (601041) and TLE6 (612399) genes. They also obtained a significant result for a SNP rs1016732 on chromosome 1q23 near the ATP1A2 gene (182340). Kilpinen et al. (2009) noted that chromosome 1q23 had been previously reported as an autism susceptibility locus in Finnish families.

Exclusion Studies

In a multicenter study in Sweden, Blomquist et al. (1985) found the fragile X repeat (309550) in 13 of 83 boys (16%) with infantile autism, but in none of 19 girls with infantile autism.

Using the UCLA Registry for Genetic Studies of Autism, Spence et al. (1985) studied 46 families with at least 2 affected children. Linkage studies in 34 families showed no evidence of linkage with HLA (142800), and close linkage with 19 other autosomal markers was excluded. The highest lod score, 1.04, was found with haptoglobin (140100) on chromosome 16q22 at recombination fractions of 10% in males and 50% in females. There was no association of the disorder with fragile X.

Using data from 38 multiplex families with autism to perform a multipoint linkage analysis with markers on the X chromosome, Hallmayer et al. (1996) excluded a moderate to strong gene effect causing autism on the X chromosome.

Using the Autism Diagnostic Instrument-Revised (ADI-R), the Autism Diagnostic Observation Scale (ADOS), and psychometric tests, Klauck et al. (1997) identified 141 autistic patients from 105 simplex and 18 multiplex families; 131 patients met all 4 ADI-R algorithm criteria for autism and 10 patients showed a broader phenotype of autism. Using amplification of the CCG repeat at the fragile X locus, hybridization to the complete FMR1 cDNA probe, and hybridization to additional probes from the neighborhood of the FMR1 gene, the authors found no significant changes in 139 patients (99%) from 122 families. In 1 multiplex family with 3 children showing no dysmorphic features of the fragile X syndrome (1 male meeting 3 of 4 ADI-algorithm criteria, 1 normal male with slight learning disability but negative ADI-R testing, and 1 fully autistic female), the FRAXA full-mutation-specific CCG-repeat expansion in the genotype was not correlated with the autism phenotype. Further analysis revealed a mosaic pattern of methylation at the FMR1 gene locus in the 2 sons of the family, indicating at least a partly functional gene. Klauck et al. (1997) concluded that the association of autism with fragile X at Xq27.3 is nonexistent and excluded this location as a candidate gene for autism.


Cytogenetics

Lopreiato and Wulfsberg (1992) described a complex chromosomal rearrangement in a 6.5-year-old boy with autism who was otherwise normal except for minimal dysmorphism. The rearrangement seen in every cell examined involved chromosomes 1, 7 and 21: 46, XY, -1, -7, -21, t(1;7;21)(1p22.1-qter::21q22.3-qter; 7pter-q11.23::7q36.1-qter; 21pter-q22.3::7q11.23-q36.1::1pter-p22.1).

Vincent et al. (2006) reported 2 brothers with autism who both carried a paracentric inversion of chromosome 4p, inv(4)(p12-p15.3), inherited from an unaffected mother and unaffected maternal grandfather. More detailed molecular analysis showed that the proximal breakpoint on 4p12 involved a cluster of GABA receptor genes, including the GABRA4 gene (137141), which has been implicated in autism (Ma et al., 2005; Collins et al., 2006). Maestrini et al. (1999) found no association or linkage to the GABRB3 gene in 94 families comprising 174 individuals with autism.

Moessner et al. (2007) identified deletions in the SHANK3 gene (606230) on chromosome 22q13 in 3 (0.75%) of 400 unrelated patients with an autism spectrum disorder. The deletions ranged in size from 277 kb to 4.36 Mb; 1 patient also had a 1.4-Mb duplication at chromosome 20q13.33. The patients were essentially nonverbal and showed poor social interactions and repetitive behaviors. Two had global developmental delay and mild dysmorphic features. A fourth patient with a de novo missense mutation in the SHANK3 gene had autism-like features but had diagnostic scores above the cutoff for autism; she was conceived by in vitro fertilization. See also the chromosome 22q13.3 deletion syndrome (606232).

Copy Number Variation

Using high-resolution microarray analysis, Marshall et al. (2008) found 277 unbalanced copy number variations (CNV), including deletion, duplication, translocation, and inversion, in 189 (44%) of 427 families with autism spectrum disorder (ASD). These specific changes were not present in a total of about 1,600 controls, although control individuals also carried many CNV. Although most variants were inherited among the patients, 27 cases had de novo alterations, and 3 (11%) of these individuals had 2 or more changes. Marshall et al. (2008) detected 13 loci with recurrent or overlapping CNV in unrelated cases. Of note, CNV at chromosome 16p11.2 (AUTS14; see 611913) was identified in 4 (1%) of 427 families and none of 1,652 controls (p = 0.002). Some of the autism loci were also common to mental retardation loci. Marshall et al. (2008) concluded that structural gene variants were found in sufficiently high frequency influencing autism spectrum disorder to suggest that cytogenetic and microarray analyses be considered in routine clinical workup.

Cusco et al. (2009) analyzed 96 Spanish patients with idiopathic ASD by array CGH analysis. Only 13 of the 238 detected CNVs (range, 89 kb-2.4 Mb) were present specifically in 12 (12.5%) of 96 ASD patients. The CNVs consisted of 10 duplications and 3 deletions. In 5 patients with parental samples available, CNVs were inherited from a normal parent. Two CNVs mapped to regions with previously reported ASD candidates, KIAA0442 (607270) on chromosome 7q11.22 and GRM8 (601116) on chromosome 7q31.3. Of the 24 genes in the CNVs, some act in common pathways, most notably the phosphatidylinositol signaling and the glutamatergic synapse, both known to be affected in several genetic syndromes related with autism and previously associated with ASD. Cusco et al. (2009) hypothesized that functional alteration of genes in related neuronal networks is involved in the etiology of the ASD phenotype.

Sebat et al. (2007) tested the hypothesis that de novo CNV is associated with autism spectrum disorders. They performed comparative genomic hybridization (CGH) on the genomic DNA of patients and unaffected subjects to detect copy number variants not present in their respective parents. Candidate genomic regions were validated by higher-resolution CGH, paternity testing, cytogenetics, fluorescence in situ hybridization, and microsatellite genotyping. Confirmed de novo CNVs were significantly associated with autism (p = 0.0005). Such CNVs were identified in 12 of 118 (10%) of patients with sporadic autism, in 2 of 77 (3%) of patients with an affected first-degree relative, and in 2 of 196 (1%) of controls. The authors stated that most de novo CNVs were smaller than microscopic resolution. Affected genomic regions were highly heterogeneous and included mutations of single genes. Sebat et al. (2007) concluded that their findings established de novo germline mutation as a more significant risk factor for autism spectrum disorders than previously recognized.

Pinto et al. (2010) analyzed the genomewide characteristics of rare (less than 1% frequency) CNVs in ASD using dense genotyping arrays. When comparing 996 ASD individuals of European ancestry to 1,287 matched controls, cases were found to carry a higher global burden of rare, genic CNVs (1.19-fold, p = 0.012), especially so for loci previously implicated in either ASD and/or intellectual disability (1.69-fold, p = 3.4 x 10(-4)). Among the CNVs there were numerous de novo and inherited events, sometimes in combination in a given family, implicating many novel ASD genes such as SHANK2 (603290), SYNGAP1 (603384), DLGAP2 (605438), and the X-linked DDX53-PTCHD1 locus (see 300828). The authors also discovered an enrichment of CNVs disrupting functional gene sets involved in cellular proliferation, projection and motility, and GTPase/Ras signaling.

Levy et al. (2011) studied 887 families from the Simons Simplex Collection of relatively high functioning ASD families. They identified 75 de novo CNVs in 68 probands (approximately 8% of probands). Only a few were recurrent. Variation at the 16p11.2 locus was detected in more than 1% of patients (10 of 858), with deletions present in 6 and duplications in 4. In addition, the duplication at 7q11.2 of the Williams syndrome region (609757) was also seen as a recurrent CNV. Levy et al. (2011) noted that the finding of 8% of ASD probands with de novo events compared with 2% in unaffected sibs was in keeping with other reports. They speculated that the slightly lower incidence of 8% rather than the 10% reported by Sebat et al. (2007) related to a higher functioning autism group in this cohort or to smaller families in this study. Using their study and the previous literature, Levy et al. (2011) proposed a list of 'asymmetries,' or observed biases, in simplex families with autism. There is a higher incidence of de novo copy number mutation in children with ASDs from simplex families than in their sibs. There is a higher incidence of de novo copy number mutation in children with ASDs from simplex families than in children with ASDs from multiplex families. For transmitted rare events, duplications greatly outweigh deletions; deletions outweigh duplications among de novo events in children with ASDs. There is evidence of transmission distortion for ultrarare events to children with ASDs; this bias arises from families in which the sib is an unaffected male. Females are less likely to be diagnosed with ASDs than are males. A higher proportion of females with ASDs have detectable de novo copy number events than do males with ASDs, and the events are larger. Levy et al. (2011) suggested that females are protected from autism but did not propose a mechanism.

Sanders et al. (2011) examined 1,124 ASD simplex families from the Simons Simplex Collection. Each of the families was comprised of a single proband, unaffected parents, and in most kindreds an unaffected sib. Sanders et al. (2011) found significant association of ASD with de novo duplications of 7q11.23. They also identified rare recurrent de novo CNVs at 5 additional regions, including 16p13.2. Overall, large de novo CNVs conferred substantial risks (odds ratio = 5.6; CI = 2.6-12.0, p = 2.4 x 10(-7)). Sanders et al. (2011) suggested that there are 130 to 234 ASD-related CNV regions in the human genome and presented compelling evidence, based on cumulative data, for association of rare de novo events at 7q11.23, 15q11.2-q13.1 (see 608636), 16p11.2, and neurexin-1 (600565). Sanders et al. (2011) found that probands carrying a 16p11.2 or 7q11.23 de novo CNV were indistinguishable from the larger ASD group with respect to IQ, ASD severity, or categorical autism diagnosis. However, they did find a relationship between body weight and 16p11.2 deletions and duplications. When copy number was treated as an ordinal variable, BMI diminished as 16p11.2 copy number increased (p = 0.02).

Vaags et al. (2012) reported 4 families in which 1 or more members had autism spectrum disorder associated with heterozygous deletions of chromosome 14q affecting the NRXN3 gene (600567). The deletions were all different and ranged from 63 to 336 kb. One deletion affected only the NRXN3 alpha isoform, whereas 3 affected both the alpha and beta isoforms. Two families were ascertained from 1,158 Canadian individuals with ASD who were screened for copy number variations across the genome. The third family was 1 of 1,368 ASD cases screened, and the fourth was 1 of 1,796 ASD cases screened. The phenotype was variable, ranging from high-functioning Asperger syndrome to full autism with some pervasive developmental and behavioral problems. In 1 family, the deletion occurred de novo. In the other families, the deletion was inherited from a parent; 1 parent had a broader autism phenotype, 1 self-reported mild autistic-like features, and 1 was normal. In 1 family, 2 of 3 trizygotic triplets with autism carried the deletion; the third unaffected child did not carry the deletion. Small deletions affecting only the alpha isoform were found in 4 of 15,122 controls. The report suggested that deletions affecting the NRXN3 gene may predispose to the development of autism spectrum disorder, but segregation patterns within the families suggested issues of penetrance and expressivity at this locus.

Loirat et al. (2010) reported 3 unrelated boys with heterozygous de novo deletions in chromosome 17q12 (see 614527) who had cystic or hyperechogenic kidneys and autism. Their 17q12 deletions ranged from 1.5 to 1.8 Mb, and included LHX1 (601999), HNF1B (189907), and 19 other genes; sequencing of the LHX1 gene in the 3 boys and 32 control patients with autism revealed no mutations. Loirat et al. (2010) concluded that autism might be an additional manifestation associated with HNF1B deletion.

Moreno-De-Luca et al. (2010) performed cytogenomic array analysis in a discovery sample of patients with neurodevelopmental disorders and detected a recurrent 1.4-Mb deletion at chromosome 17q12 in 18 of 15,749 patients, including 6 with autism or autistic features; the deletion was not found in 4,519 controls. In a large follow-up sample, the same deletion was identified in 2 of 1,182 patients with autism spectrum disorder and/or neurocognitive impairment, and in 4 of 6,340 schizophrenia (see 181500) patients, but was not found in 47,929 controls (corrected p = 7.37 x 10 (-5)). Moreno-De-Luca et al. (2010) concluded that deletion 17q12 is a recurrent, pathogenic CNV that confers a high risk for autism spectrum disorder and schizophrenia, and that 1 or more of the 15 genes in the deleted interval is dosage-sensitive and essential for normal brain development and function.

Luo et al. (2012) interrogated gene expression in lymphoblasts from 439 individuals from 244 families with discordant sibs in the Simons Simplex Collection and found that the overall frequency of significantly misexpressed genes, which they referred to as outliers, did not differ between probands and unaffected sibs. However, in probands, but not their unaffected sibs, the group of outlier genes was significantly enriched in neural-related pathways, including neuropeptide signaling, synaptogenesis, and cell adhesion. The outlier genes clustered within large rare de novo CNVs and could be used for the prioritization of rare CNVs of potential significance. Several nonrecurrent CNVs with significant gene expression alterations were identified, including deletions in chromosome regions 3q27, 3p13, and 3p26 and duplications at 2p15, suggesting these as potential ASD loci.

See SHANK1 (604999) for discussion of a possible association between heterozygous deletions involving the SHANK1 gene on chromosome 19q13 and susceptibility to high-functioning autism.

Krumm et al. (2013) searched for disruptive, genic rare CNVs among 411 families affected by sporadic autism spectrum disorder from the Simons Simplex Collection by using available exome sequence data and CoNIFER (Copy Number Inference from Exome Reads). Compared to high density SNP microarrays, the authors' approach yielded approximately 2 times more smaller genic rare CNVs. Krumm et al. (2013) found that affected probands inherited more CNVs than did their sibs (453 vs 394, p = 0.004; odds ratio = 1.19) and that the probands' CNVs affected more genes (921 vs 726, p = 0.02; odds ratio = 1.30). These smaller CNVs (median size 18 kb) were transmitted preferentially from the mother (136 maternal vs 100 paternal, p = 0.02), although this bias occurred irrespective of affected status. The excess burden of inherited CNVs was driven primarily by sib pairs with discordant social behavior phenotypes, which contrasts with families where the phenotypes were more closely matched or less extreme. In a combined model, the inherited CNVs, de novo CNVs, and de novo single-nucleotide variants all independently contributed to the risk of autism (p less than 0.05).

Poultney et al. (2013) used the eXome Hidden Markov Model (XHMM) as well as transmission information and validation by molecular methods to confirm that small CNVs encompassing as few as 3 exons can be reliably called from whole-exome data. They applied this approach to an autism case-control sample of 811 subjects (mean per-target read depth = 161) and observed a significant increase in the burden of rare (minor allele frequency (MAF) 1% or less) 1- to 30-kb CNVs, 1- to 30-kb deletions, and 1- to 10-kb deletions in ASD. CNVs in the 1 to 30 kb range frequently hit just a single gene, allowing Poultney et al. (2013) to observe enrichment for disruption of genes in cytoskeletal and autophagy pathways in ASD. Poultney et al. (2013) concluded that rare 1- to 30-kb exonic deletions could contribute to risk in up to 7% of individuals with ASD.

Girirajan et al. (2013) exploited the repeat architecture of the genome to target segmental duplication-mediated rearrangement hotspots (n = 120, median size 1.78 Mbp, range 240 kbp to 13 Mbp) and smaller hotspots flanked by repetitive sequence (n = 1,247, median size 79 kbp, range 3-96 kbp) in 2,588 autistic individuals from simplex and multiplex families and in 580 controls. The analysis identified several recurrent large hotspot events, including association with 1q21 duplications, which are more likely to be identified in individuals with autism than in those with developmental delay (p = 0.01; odds ratio = 2.7). Within larger hotspots, Girirajan et al. (2013) also identified smaller atypical CNVs that implicated CHD1L (613039) and ACACA (200350) for the 1q21 and 17q12 deletions, respectively. The analysis, however, suggested no overall increase in the burden of smaller hotspots in autistic individuals as compared to controls. By focusing on gene-disruptive events, Girirajan et al. (2013) identified several genes that were enriched for CNVs in autism cases versus controls, including DPP10 (608209), PLCB1 (607120), TRPM1 (603576), NRXN1 (600565), FHIT (601153), and HYDIN (610812). Girirajan et al. (2013) found that as the size of deletions increases, nonverbal IQ significantly decreases, but there is no impact on autism severity; as the size of duplications increases, autism severity significantly increases but nonverbal IQ is not affected. Girirajan et al. (2013) concluded that the absence of an increased burden of smaller CNVs in individuals with autism and the failure of most large hotspots to refine to single genes is consistent with a model where imbalance of multiple genes contributes to a disease state.

Pinto et al. (2014) analyzed 2,446 ASD-affected families and confirmed an excess of genic deletions and duplications in affected versus control groups (1.41-fold, p = 1.0 x 10(-5)). They also found an increase in affected subjects carrying exonic pathogenic CNVs overlapping known loci associated with dominant or X-linked ASD and intellectual disability (odds ratio = 12.62, p = 2.7 x 10(-15), approximately 3% of ASD subjects). Pathogenic CNVs, often showing variable expressivity, included rare de novo and inherited events at 36 loci, implicating ASD-associated genes (CHD2, 602119; HDAC4, 605314; and GDI1, 300104) linked to other neurodevelopmental disorders, as well as other genes, such as SETD5 (615743), MIR137 (614304), and HDAC9 (606543). Consistent with hypothesized gender-specific modulators, females with ASD were more likely to have highly penetrant CNVs (p = 0.017) and were also overrepresented among subjects with fragile X syndrome protein targets (p = 0.02).


Molecular Genetics

Gauthier et al. (2011) identified a heterozygous 1-bp deletion (2733delT) in the NRXN2 gene (600566) on chromosome 11q13 in a boy of European ancestry with autism spectrum disorder. The mutation resulted in premature termination. In vitro functional expression studies in COS-7 cells showed that the mutant protein was unable to bind its usual partners, and in vitro studies in neuronal culture showed a loss of synaptogenic activity with lack of clustering of postsynaptic components. The findings were consistent with a loss of function. The mutation was inherited from the patient's father, who had severe language delay. A maternal aunt of the father's had schizophrenia, but DNA was not available from her. The patient was identified from a cohort of 142 patients with autism who were screened for mutations in the NRXN1 (600565), NRXN2, and NRXN3 genes.

Sanders et al. (2012) used whole-exome sequencing of 928 individuals, including 200 phenotypically discordant sib pairs, to demonstrate that highly disruptive nonsense and splice site de novo mutations in brain-expressed genes are associated with autism spectrum disorders and carry large effects. On the basis of mutation rates in unaffected individuals, they demonstrated that multiple independent de novo single-nucleotide variants in the same gene among unrelated probands reliably identifies risk alleles, providing a clear path forward for gene discovery. Among a total of 279 identified de novo coding mutations, there was a single instance in probands, and none in sibs, in which 2 independent nonsense variants disrupt the same gene, SCN2A (182390). Sanders et al. (2012) combined all de novo events in their sample with those identified in the study of O'Roak et al. (2012) and observed from a total of 414 probands 2 additional genes carrying 2 highly disruptive mutations each, KATNAL2 (614697) and CHD8 (610528).

O'Roak et al. (2012) performed whole-exome sequencing for parent-child trios exhibiting sporadic autism spectrum disorders, including 189 new trios and 20 that were previously reported (O'Roak et al., 2011). In addition, O'Roak et al. (2012) sequenced the exomes of 50 unaffected sibs corresponding to 31 of the new and 19 of the previously reported trios, for a total of 677 individual exomes from 209 families. O'Roak et al. (2012) showed that de novo point mutations are overwhelmingly paternal in origin (4:1 bias) and positively correlated with paternal age, consistent with the modest increased risk for children of older fathers to develop autism spectrum disorders. Moreover, 39% (49 of 126) of the most severe or disruptive de novo mutations mapped to a highly interconnected beta-catenin (116806)/chromatin remodeling protein network ranked significantly for autism candidate genes. In proband exomes, recurrent protein-altering mutations were observed in 2 genes: CHD8 and NTNG1. Mutation screening of 6 candidate genes in 1,703 ASD probands identified additional de novo, protein-altering mutations in GRIN2B (138252), LAMC3 (604349), and SCN1A (182389). Combined with copy number data, these data indicated extreme locus heterogeneity in ASD. O'Roak et al. (2012) concluded that their analysis predicted extreme locus heterogeneity underlying the genetic etiology of autism. Under a strict sporadic disorder-de novo mutation model, if 20 to 30% of the de novo point mutations are considered to be pathogenic, they could estimate between 384 and 821 loci. Furthermore, 1 individual inherited 3 rare gene disruptive CNVs and carried 2 de novo truncating mutations.

Neale et al. (2012) assessed the role of de novo mutations in autism spectrum disorders by sequencing the exomes of ASD cases and their parents (175 trios). Fewer than half of the cases (46.3%) carried a missense or nonsense de novo variant, and the overall rate of mutation was only modestly higher than the expected rate. In contrast, the proteins encoded by genes that harbored de novo missense or nonsense mutations showed a higher degree of connectivity among themselves and to previous ASD genes as indexed by protein-protein interaction screens. The small increase in the rate of de novo events, when taken together with the protein interaction results, are consistent with an important but limited role for de novo point mutations in ASD, similar to that documented for de novo copy number variants. Genetic models incorporating data indicated that most of the observed de novo events are unconnected to ASD; those that do confer risk are distributed across many genes and are incompletely penetrant (i.e., not necessarily sufficient for disease). Neale et al. (2012) concluded that their results supported polygenic models in which spontaneous coding mutations in any of a large number of genes increases risk by 5- to 20-fold. Despite the challenge posed by such models, results from de novo events and a large parallel case-control study provided strong evidence in favor of CHD8 and KATNAL2 as genuine autism risk factors.

O'Roak et al. (2012) developed a modified molecular inversion probe method enabling ultra-low-cost candidate gene resequencing in very large cohorts. To demonstrate the power of this approach, O'Roak et al. (2012) captured and sequenced 44 candidate genes in 2,446 ASD probands, and discovered 27 de novo events in 16 genes, 59% of which are predicted to truncate proteins or disrupt splicing. O'Roak et al. (2012) estimated that recurrent disruptive mutations in 6 genes--CHD8, DYRK1A (600855), GRIN2B, TBR1 (604616), PTEN (601728), and TBL1XR1 (608628)--may contribute to 1% of sporadic autism spectrum disorders. O'Roak et al. (2012) concluded that their data supported associations between specific genes and reciprocal subphenotypes (CHD8-macrocephaly and DYRK1A-microcephaly) and replicated the importance of a beta-catenin/chromatin-remodeling network to ASD etiology.

Jiang et al. (2013) used whole-genome sequencing to examine 32 families with ASD to detect de novo or rare inherited genetic variants predicted to be deleterious (loss-of-function and damaging missense mutations). Among ASD probands, Jiang et al. (2013) identified deleterious de novo mutations in 6 of 32 (19%) families and X-linked or autosomal inherited alterations in 10 of 32 (31%) families (some had combinations of mutations). The proportion of families identified with such putative mutations was larger than had been reported; this yield was in part due to the comprehensive and uniform coverage afforded by whole-genome sequencing. Deleterious variants were found in 4 unrecognized, 9 known, and 8 candidate ASD risk genes. Examples include CAPRIN1 (601178), AFF2 (300806), VIP (192320), SCN2A, KCNQ2 (602235), NRXN1, and CHD7 (608892).

To characterize the role of rare complete human knockouts in autism spectrum disorders, Lim et al. (2013) identified genes with homozygous or compound heterozygous loss-of-function variants (defined as nonsense and essential splice sites) from exome sequencing of 933 cases and 869 controls. Lim et al. (2013) identified a 2-fold increase in complete knockouts of autosomal genes with low rates of loss-of-function variation (less than or equal to 5% frequency) in cases, and estimated a 3% contribution to autism spectrum disorder risk by these events, confirming this observation in an independent set of 563 probands and 4,605 controls. Outside the pseudoautosomal regions on the X chromosome, Lim et al. (2013) similarly observed a significant 1.5-fold increase in rare hemizygous knockouts in males, contributing to another 2% of autism spectrum disorders in males. Lim et al. (2013) concluded that these results provided compelling evidence that rare autosomal and X chromosome complete gene knockouts are important inherited risk factors for autism spectrum disorders.

Using exome sequencing, De Rubeis et al. (2014) showed that analysis of rare coding variation in 3,871 autism cases and 9,937 ancestry-matched or parental controls implicated 22 autosomal genes at a false discovery rate (FDR) of less than 0.05, plus a set of 107 autosomal genes strongly enriched for those likely to affect risk (FDR less than 0.30). These 107 genes, which show unusual evolutionary constraint against mutations, incurred de novo loss-of-function mutations in over 5% of autistic subjects. Many of the genes implicated encode proteins for synaptic formation, transcriptional regulation, and chromatin-remodeling pathways. These included voltage-gated ion channels regulating the propagation of action potentials, pacemaking, and excitability-transcription coupling, as well as histone-modifying enzymes and chromatin remodelers, most prominently those that mediate posttranslational lysine methylation/demethylation modifications of histones.

Associations Pending Confirmation

See 605410.0001 for discussion of a possible association between variation in the KCND2 gene and infantile-onset severe refractory epilepsy (see EIEE1, 308350) and autism.

For discussion of a possible association between autism spectrum disorder and variation in the PRICKLE2 gene, see 608501.


Population Genetics

Smalley (1997) reported that autism has a population prevalence of approximately 4 to 5 in 10,000 with a male to female ratio of 4 to 1.

In a review of 20 studies on autism published between 1966 and 1997, Gillberg and Wing (1999) determined that autism is considerably more common than previously believed. The early studies yielded prevalence rates of under 0.5 per 1,000 children, whereas the later studies showed a mean rate of about 1 in 1,000. Children born after 1970 had a much higher rate than those born before 1970.

Bertrand et al. (2001) performed a prevalence study of autism spectrum disorders in Brick Township, New Jersey. There were 6.7 cases per 1,000 children, aged 3 to 10 years, in 1998. The prevalence for children whose condition met full diagnostic criteria for autistic disorder was 4.0 cases per 1,000 children, and the prevalence for PDD-not otherwise specified (NOS) and Asperger syndrome was 2.7 cases per 1,000 children.

In a review, Jones et al. (2008) noted that the significant increase in the frequency with which autism spectrum disorders is diagnosed, from 4 per 10,000 in 1950 to 40 to 60 per 10,000 as of 2008, results from greater awareness, availability of services, and changes in diagnostic criteria to include a broader spectrum of neurodevelopmental disorders, among others.


Pathogenesis

Schain and Freedman (1961) reported elevated levels of platelet serotonin (5-HT; see 182138) in patients with autism. Abramson et al. (1989) reported elevated blood serotonin in autistic probands and in their first-degree relatives. Piven et al. (1991) found that serotonin levels were significantly higher in autistic individuals with a sib with autism or PDD than in those without a sib with these disorders, and that autistic patients without an affected sib had serotonin levels that were significantly higher than controls.

A biologic basis of autism was suggested by the finding of developmental hypoplasia in lobules VI and VII of the cerebellar vermis (Courchesne et al., 1988). The ontogenetically, developmentally, and anatomically distinct vermal lobules I to V were found to be of normal size. However, Schaefer et al. (1996) disputed the relationship of cerebellar vermal atrophy to infantile autism. They found that the average relative size of lobules VI and VII of the cerebellar vermis was no different in their 13 patients with infantile autism when compared to that of 125 normal individuals. They found relative hypoplasia of lobules VI and VII in patients with Rett syndrome (312750) and Sotos cerebral gigantism (117550), 2 disorders characterized by autistic behaviors. No relative vermian atrophy was seen in other disorders associated with autistic behavior: fragile X, Angelman (AS; 105830), adult phenylketonuria (261600), and Sanfilippo (252900). Furthermore, they found a relative atrophy of lobules VI and VII in several patients with primary cerebellar hypoplasia and Usher syndrome type II (276901), syndromes not associated with autistic behavior.

Autopsy and neuroimaging studies have suggested that autism spectrum disorder is caused in part by abnormal brain development. Benayed et al. (2005) reviewed cerebellar abnormalities in autism spectrum disorder. The CNS structure most consistently affected in individuals with autism is the cerebellum, with a decrease in the number of Purkinje cells being present in a majority. Neurodegenerative signs are for the most part absent from these autopsy samples, suggesting a developmental defect. Neuroimaging studies have consistently demonstrated posterior cerebellar hypoplasia. Although the cerebellum has classically been considered a motor control center, functional imaging studies indicated that the cerebellum is also active during cognitive tasks that are defective in autism spectrum disorders, including language and attention. Thus, the identified cerebellar defects may contribute directly to some of the behavioral abnormalities associated with autism spectrum disorder. In turn, genetic alterations that perturb cerebellar development may contribute to susceptibility to autism spectrum disorder.

Regressive autism, characterized most prominently by a loss of language skills, has been attributed to environmental factors, particularly adverse reactions to vaccines; epidemiologic evidence, however, shows no association between vaccination and the rate of autism as reviewed by the Institute of Medicine Immunization Safety Reviews (2001); see also Taylor et al. (2002). Lainhart et al. (2002) noted that twin and family studies showed that the liability to autism extends beyond the full autism syndrome and includes qualitatively similar, albeit milder, deficits, referred to as the broader autism phenotype (BAP). If regressive autism is solely caused by environmental events, such as adverse reactions to vaccines, rates of the BAP in the relatives of children with regressive autism should be no greater than in the general population. If environmental events do not independently cause regressive autism, or if they act as 'second-hit' phenomena in children who already have a genetic liability to autism, rates of the BAP should be similar in relatives of autistic children with and without regression. Lainhart et al. (2002) found that the rate of the BAP was significantly higher in parents of children with regressive and nonregressive autism than in parents of nonautistic children. They concluded that environmental events are unlikely to be the sole cause of regressive autism, although environmental events may act in an additive or 'second-hit' fashion in individuals with a genetic vulnerability to autism.

In a review, Jones et al. (2008) discussed the hypothesis that dysregulation of methylation of brain-expressed genes on the X chromosome constitutes the major predisposition to the development of autism spectrum disorders. Broad evidence consistent with this epigenetic effect includes marked excess of males among individuals affected with ASD, most patients have a sporadic occurrence of the disorder, and most patients do not have syndromic features.

In studies of lymphocytes from 10 children with autism and 10 controls, Giulivi et al. (2010) found that patients with autism were more likely to have mitochondrial dysfunction, mtDNA overreplication, and mtDNA deletions compared to normally developing children. Lymphocytes from children with autism had lower mitochondrial-dependent oxygen consumption, with low complex I (6 of 10) and complex V (4 of 10) activity. Autistic children had increased plasma pyruvate levels and increased lymphocyte hydrogen peroxide production. Five autistic patients and 2 controls had mtDNA overreplication, and 2 patients and no controls had mtDNA deletion. Overall, the findings suggested that autism may be associated with mitochondrial dysfunction, which may reflect insufficient energy production. However, Giulivi et al. (2010) noted that the observations did not elucidate primary or secondary effects.

Castermans et al. (2010) described the positional cloning of SCAMP5 (613766) as a candidate gene for autism, based on finding a de novo chromosomal translocation t1;15(p36.11;q24.2) in a 40-year-old affected male. SCAMP5, which was silenced on the derivative chromosome, encodes a brain-enriched protein involved in membrane trafficking, similar to the previously identified candidate genes NBEA (604889) and AMISYN (STXBP6; 607958). Gene silencing of Nbea, Amisyn, and Scamp5 in mouse beta-TC3 cells resulted in a 2-fold increase in stimulated secretion of large dense-core vesicles (LDCVs), whereas overexpression suppressed secretion. Ultrastructural analysis of blood platelets from autism patients with haploinsufficiency of 1 of the 3 candidate genes showed morphologic abnormalities of dense-core granules, which closely resembled LDCVs. Castermans et al. (2010) suggested that in a subgroup of patients, the regulation of neuronal vesicle trafficking may be involved in the pathogenesis of autism.

Voineagu et al. (2011) analyzed postmortem brain tissue samples from 19 autism cases and 17 controls from the Autism Tissue Project and the Harvard brain bank using Illumina microarrays. For each individual, they profiled 3 regions previously implicated in autism: superior temporal gyrus, prefrontal cortex, and cerebellar vermis. Voineagu et al. (2011) demonstrated consistent differences in transcriptome organization between autistic and normal brain by gene coexpression network analysis. Remarkably, regional patterns of gene expression that typically distinguish frontal and temporal cortex are significantly attenuated in the autism spectrum disorder (ASD) brain, suggesting abnormalities in cortical patterning. Voineagu et al. (2011) further identified discrete modules of coexpressed genes associated with autism: a neuronal module enriched for known autism susceptibility genes, including the neuronal-specific splicing factor A2BP1 (also known as FOX1, 605104), and a module enriched for immune genes and glial markers. Using high-throughput RNA sequencing, they demonstrated dysregulated splicing of A2BP1-dependent alternative exons in the ASD brain. Moreover, using a published autism genomewide association study (GWAS) data set, Voineagu et al. (2011) showed that the neuronal module is enriched for genetically associated variants, providing independent support for the causal involvement of those genes in autism. The top module for differential expression between autism control groups was highly enriched for neuronal markers. The hubs of this group, called M12 in this study, which represented the genes with the highest rank of M12 membership, were A2BP1 but also APBA2 (602712), SCAMP5 (613766), CNTNAP1 (602346), KLC2 (611729), and CHRM1 (118510). In contrast, the immune-glial module showed no enrichment for autism GWAS signals, indicating a nongenetic etiology for this process. Voineagu et al. (2011) concluded that their results provided strong evidence for convergent molecular abnormalities in ASD, and implicated transcriptional and splicing dysregulation as underlying mechanisms of neuronal dysfunction in this disorder.

Gilman et al. (2011) developed a network-based analysis of genetic associations (NETBAG) and used this to identify a large biologic network of genes affected by rare de novo CNVs in autism. The genes forming the network are primarily related to synapse development, axon targeting, and neuron motility. The identified network was strongly related to genes previously implicated in autism and intellectual disability phenotypes. Gilman et al. (2011) suggested that their results were also consistent with the hypothesis that significantly stronger functional perturbations are required to trigger the autistic phenotype in females compared to males. Overall, the presented analysis of de novo variants supported the hypothesis that perturbed synaptogenesis is at the heart of autism. More generally, their study provided proof of the principle that networks underlying complex human phenotypes can be identified by a network-based functional analysis of rare genetic variants.

To study genomewide mutation rates, Kong et al. (2012) sequenced the entire genomes of 78 Icelandic parent-offspring trios at high coverage. Forty-four of the probands had autistic spectrum disorder and 21 were schizophrenic (181500). Kong et al. (2012) found that, with an average father's age of 29.7, the average de novo mutation rate is 1.20 x 10(-8) per nucleotide per generation. Most notably, the diversity in mutation rate of single-nucleotide polymorphisms was dominated by the age of the father at conception of the child. The effect is an increase of about 2 mutations per year. An exponential model estimates paternal mutations doubling every 16.5 years. After accounting for random Poisson variation, father's age is estimated to explain nearly all of the remaining variation in the de novo mutation counts. Kong et al. (2012) stated that there had been a recent transition of Icelanders from a rural agricultural to an urban industrial way of life, which engendered a rapid and sequential drop in the average age of fathers at conception from 34.9 years in 1900 to 27.9 years in 1980, followed by an equally swift climb back to 33.0 years in 2011, primarily owing to the effect of higher education and the increased use of contraception. On the basis of the fitted linear model, whereas individuals born in 1900 carried on average 73.7 de novo mutations, those born in 1980 carried on average only 59.7 such mutations (a decrease of 19.1%), and the mutational load of individuals born in 2011 had increased by 17.2% to 69.9. Kong et al. (2012) concluded that their observations shed light on the importance of the father's age on the risk of diseases such as schizophrenia and autism.

King et al. (2013) found that topotecan, a topoisomerase-1 (TOP1; 126420) inhibitor, dose-dependently reduces the expression of extremely long genes in mouse and human neurons, including nearly all genes that are longer than 200 kb. Expression of long genes is also reduced after knockdown of Top1 or Top2b (126431) in neurons, highlighting that both enzymes are required for full expression of long genes. By mapping RNA polymerase II density genomewide in neurons, King et al. (2013) found that this length-dependent effect on gene expression was due to impaired transcription elongation. Interestingly, many high-confidence autism spectrum disorder candidate genes are exceptionally long and were reduced in expression after TOP1 inhibition. King et al. (2013) concluded that chemicals and genetic mutations that impair topoisomerases could commonly contribute to autism spectrum disorders and other neurodevelopmental disorders.

Gamsiz et al. (2013) conducted a genomewide analysis of runs of homozygosity (ROH) in simplex ASD-affected families consisting of a proband diagnosed with ASD and at least 1 unaffected sib. In these families, probands with an IQ of 70 or below show more ROH than their unaffected sibs, whereas probands with an IQ greater than 70 do not show this excess. Although ASD is far more common in males than in females, the proportion of females increases with decreasing IQ. Gamsiz et al. (2013) stated that their data supported an association between ROH burden and autism diagnosis in girls; however, they were not able to show that this effect was independent of low IQ. The authors also identified several autism candidate genes on the basis of their being either a single gene that is within an ROH interval and that is recurrent in autism, or a gene that is within an ROH block and that harbors a homozygous rare deleterious variant upon analysis of exome sequencing data.

Through postmortem genomewide transcriptome analysis of the largest cohort of samples analyzed to that time, Parikshak et al. (2016) interrogated the noncoding transcriptome, alternative splicing, and upstream molecular regulators to broaden understanding of molecular convergence in ASD. The analysis revealed ASD-associated dysregulation of primate-specific long noncoding RNAs (lncRNAs, especially upregulation of LINC00693 and LINC00689), downregulation of the alternative splicing of activity-dependent neuron-specific exons, and attenuation of normal differences in gene expression between the frontal and temporal lobes. Their data suggested that SOX5 (604975), a transcription factor involved in neuron fate specification, contributes to this reduction in regional differences. Parikshak et al. (2016) further demonstrated that a genetically defined subtype of ASD, chromosome 15q11.2-13.1 duplication syndrome (dup15q; 608636), shares the core transcriptomic signature observed in idiopathic ASD. Coexpression network analysis revealed that individuals with ASD show age-related changes in the trajectory of microglial and synaptic function over the first 2 decades, and suggested that genetic risk for ASD may influence changes in regional cortical gene expression. Parikshak et al. (2016) concluded that their findings illustrated how diverse genetic perturbations can lead to phenotypic convergence at multiple biologic levels in a complex neuropsychiatric disorder.


Animal Model

Tabuchi et al. (2007) introduced the R451C (arg451 to cys; 300336.0001) substitution in neuroligin-3 into mice. R451C mutant mice showed impaired social interactions but enhanced spatial learning abilities. Unexpectedly these behavioral changes were accompanied by an increase in inhibitory synaptic transmission with no apparent effect on excitatory synapses. Deletion of neuroligin-3, in contrast, did not cause such changes, indicating that the R451C substitution represents a gain-of-function mutation. Tabuchi et al. (2007) concluded that increased inhibitory synaptic transmission may contribute to human autism spectrum disorders and that the R451C knockin mice may be a useful model for studying autism-related behaviors.

Using maternal immune activation (MIA) in a mouse model of ASD and genetic mutant mice, Choi et al. (2016) showed that Rorgt (see 602943)-dependent effector T lymphocytes, such as T-helper-17 (Th17) cells, and Il17a (603149) were required in mothers for MIA-induced behavioral abnormalities in offspring. MIA induced a maternal Il17a-dependent abnormal phenotype in cortex of fetal brain. Treating pregnant female mice with antibodies blocking Il17a ameliorated MIA-associated behavioral abnormalities. Choi et al. (2016) proposed that targeting of Th17 cells in susceptible pregnant mothers may reduce the likelihood of them bearing children with inflammation-induced ASD-like phenotypes.


History

Eisenberg (1994) provided a biographic sketch of Leo Kanner (1894-1981), the pioneer pediatric psychiatrist who first described and named infantile autism (Kanner, 1943).


REFERENCES

  1. Abramson, R. K., Wright, H. H., Carpenter, R., Brennan, W., Lumpuy, O., Cole, E., Young, S. R. Elevated blood serotonin in autistic probands and their first-degree relatives. J. Autism Dev. Disord. 19: 397-407, 1989. [PubMed: 2793785]

  2. Alarcon, M., Cantor, R. M., Liu, J., Gilliam, T. C., Autism Genetic Resource Exchange Consortium, Geschwind, D. H. Evidence for a language quantitative trait locus on chromosome 7q in multiplex autism families. Am. J. Hum. Genet. 70: 60-71, 2002. [PubMed: 11741194] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(07)61283-X]

  3. American Psychiatric Association. Diagnostic and Statistical Manual of Mental Disorders. (4th ed.) Washington, D.C.: American Psychiatric Association (pub.) 1994.

  4. Autism Genome Project Consortium. Mapping autism risk loci using genetic linkage and chromosomal rearrangements. Nature Genet. 39: 319-328, 2007. Note: Erratum: Nature Genet. 39: 1285 only, 2007. [PubMed: 17322880] [Full Text: https://dx.doi.org/10.1038/ng1985]

  5. Awadalla, P., Gauthier, J., Myers, R. A., Casals, F., Hamdan, F. F., Griffing, A. R., Cote, M., Henrion, E., Spiegelman, D., Tarabeux, J., Piton, A., Yang, Y., and 22 others. Direct measure of the de novo mutation rate in autism and schizophrenia cohorts. Am. J. Hum. Genet. 87: 316-324, 2010. [PubMed: 20797689] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(10)00376-9]

  6. Bailey, A., Le Couteur, A., Gottesman, I., Bolton, P., Simonoff, E., Yuzda, E., Rutter, M. Autism as a strongly genetic disorder: evidence from a British twin study. Psychol. Med. 25: 63-77, 1995. [PubMed: 7792363]

  7. Bailey, A., Phillips, W., Rutter, M. Autism: towards an integration of clinical, genetic, neuropsychological, and neurobiological perspectives. J. Child Psychol. Psychiat. 37: 89-126, 1996. [PubMed: 8655659] [Full Text: http://onlinelibrary.wiley.com/resolve/openurl?genre=article&sid=nlm:pubmed&issn=0021-9630&date=1996&volume=37&issue=1&spage=89]

  8. Benayed, R., Gharani, N., Rossman, I., Mancuso, V., Lazar, G., Kamdar, S., Bruse, S. E., Tischfield, S., Smith, B. J., Zimmerman, R. A., DiCicco-Bloom, E., Brzustowicz, L. M., Millonig, J. H. Support for the homeobox transcription factor gene ENGRAILED 2 as an autism spectrum disorder susceptibility locus. Am. J. Hum. Genet. 77: 851-868, 2005. [PubMed: 16252243] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(07)63365-5]

  9. Bertrand, J., Mars, A., Boyle, C., Bove, F., Yeargin-Allsopp, M., Decoufle, P. Prevalence of autism in a United States population: the Brick Township, New Jersey, investigation. Pediatrics 108: 1155-1161, 2001. [PubMed: 11694696] [Full Text: http://pediatrics.aappublications.org/cgi/pmidlookup?view=long&pmid=11694696]

  10. Blomquist, H. K., Bohman, M., Edvinsson, S. O., Gillberg, C., Gustavson, K.-H., Holmgren, G., Wahlstrom, J. Frequency of the fragile X syndrome in infantile autism: a Swedish multicenter study. Clin. Genet. 27: 113-117, 1985. [PubMed: 3978844] [Full Text: http://onlinelibrary.wiley.com/resolve/openurl?genre=article&sid=nlm:pubmed&issn=0009-9163&date=1985&volume=27&issue=2&spage=113]

  11. Bolton, P., Macdonald, H., Pickles, A., Rios, P., Goode, S., Crowson, M., Bailey, A., Rutter, M. A case-control family history study of autism. J. Child Psychol. Psychiat. 35: 877-900, 1994. [PubMed: 7962246] [Full Text: http://onlinelibrary.wiley.com/resolve/openurl?genre=article&sid=nlm:pubmed&issn=0021-9630&date=1994&volume=35&issue=5&spage=877]

  12. Burd, L., Kerbeshian, J., Fisher, W. Inquiry into the incidence of hyperlexia in a statewide population of children with pervasive developmental disorder. Psychol. Rep. 57: 236-238, 1985. [PubMed: 4048337] [Full Text: https://dx.doi.org/10.2466/pr0.1985.57.1.236]

  13. Castermans, D., Volders, K., Crepel, A., Backx, L., De Vos, R., Freson, K., Meulemans, S., Vermeesch, J. R., Schrander-Stumpel, C. T. R. M., De Rijk, P., Del-Favero, J., Van Geet, C., Van De Ven, W. J. M., Steyaert, J. G., Devriendt, K., Creemers, J. W. M. SCAMP5, NBEA and AMISYN: three candidate genes for autism involved in secretion of large dense-core vesicles. Hum. Molec. Genet. 19: 1368-1378, 2010. [PubMed: 20071347] [Full Text: https://academic.oup.com/hmg/article-lookup/doi/10.1093/hmg/ddq013]

  14. Choi, G. B., Yim, Y. S., Wong, H., Kim, S., Kim, H., Kim, S. V., Hoeffer, C. A., Littman, D. R., Huh, J. R. The maternal interleukin-17a pathway in mice promotes autism-like phenotypes in offspring. Science 351: 933-939, 2016. [PubMed: 26822608] [Full Text: http://www.sciencemag.org/cgi/pmidlookup?view=long&pmid=26822608]

  15. Cohen, D., Pichard, N., Tordjman, S., Baumann, C., Burglen, L., Excoffier, E., Lazar, G., Mazet, P., Pinquier, C., Verloes, A., Heron, D. Specific genetic disorders and autism: clinical contribution towards their identification. J. Autism Dev. Disord. 35: 103-116, 2005. [PubMed: 15796126] [Full Text: https://www.springerlink.com/openurl.asp?genre=article&issn=0162-3257&volume=35&issue=1&spage=103]

  16. Collins, A. L., Ma, D., Whitehead, P. L., Martin, E. R., Wright, H. H., Abramson, R. K., Hussman, J. P., Haines, J. L., Cuccaro, M. L., Gilbert, J. R., Pericak-Vance, M. A. Investigation of autism and GABA receptor subunit genes in multiple ethnic groups. Neurogenetics 7: 167-174, 2006. [PubMed: 16770606] [Full Text: https://dx.doi.org/10.1007/s10048-006-0045-1]

  17. Courchesne, E., Yeung-Courchesne, R., Press, G. A., Hesselink, J. R., Jernigan, T. L. Hypoplasia of cerebellar vermal lobules VI and VII in autism. New Eng. J. Med. 318: 1349-1354, 1988. [PubMed: 3367935] [Full Text: http://www.nejm.org/doi/abs/10.1056/NEJM198805263182102?url_ver=Z39.88-2003&rfr_id=ori:rid:crossref.org&rfr_dat=cr_pub%3dpubmed]

  18. Cusco, I., Medrano, A., Gener, B., Vilardell, M., Gallastegui, F., Villa, O., Gonzalez, E., Rodriguez-Santiago, B., Vilella, E., Del Campo, M., Perez-Jurado, L. A. Autism-specific copy number variants further implicate the phosphatidylinositol signaling pathway and the glutamatergic synapse in the etiology of the disorder. Hum. Molec. Genet. 18: 1795-1804, 2009. [PubMed: 19246517] [Full Text: https://academic.oup.com/hmg/article-lookup/doi/10.1093/hmg/ddp092]

  19. De Rubeis, S., He, X., Goldberg, A. P., Poultney, C. S., Samocha, K., Cicek, A. E., Kou, Y., Liu, L., Fromer, M., Walker, S., Singh, T., Klei, L., and 83 others. Synaptic, transcriptional and chromatin genes disrupted in autism. Nature 515: 209-215, 2014. [PubMed: 25363760] [Full Text: https://dx.doi.org/10.1038/nature13772]

  20. Eisenberg, L. Images in psychiatry: Leo Kanner (1894-1981). Am. J. Psychiat. 151: 751, 1994. [PubMed: 8166319] [Full Text: http://psychiatryonline.org/doi/abs/10.1176/ajp.151.5.751?url_ver=Z39.88-2003&rfr_id=ori:rid:crossref.org&rfr_dat=cr_pub%3dpubmed]

  21. Folstein, S. E., Rosen-Sheidley, B. Genetics of autism: complex aetiology for a heterogeneous disorder. Nature Rev. Genet. 2: 943-955, 2001. [PubMed: 11733747] [Full Text: https://dx.doi.org/10.1038/35103559]

  22. Folstein, S., Rutter, M. Genetic influences and infantile autism. Nature 265: 726-728, 1977. [PubMed: 558516]

  23. Folstein, S., Rutter, M. Infantile autism: a genetic study of 21 twin pairs. J. Child Psychol. Psychiat. 18: 297-321, 1977. [PubMed: 562353] [Full Text: http://onlinelibrary.wiley.com/resolve/openurl?genre=article&sid=nlm:pubmed&issn=0021-9630&date=1977&volume=18&issue=4&spage=297]

  24. Gamsiz, E. D., Viscidi, E. W., Frederick, A. M., Nagpal, S., Sanders, S. J., Murtha, M. T., Schmidt, M., Simons Simplex Collection Genetics Consortium, Triche, E. W., Geschwind, D. H., State, M. W., Istrail, S., Cook, E. H., Jr., Devlin, B., Morrow, E. M. Intellectual disability is associated with increased runs of homozygosity in simplex autism. Am. J. Hum. Genet. 93: 103-109, 2013. [PubMed: 23830515] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(13)00272-3]

  25. Gauthier, J., Siddiqui, T. J., Huashan, P., Yokomaku, D., Hamdan, F. F., Champagne, N., Lapointe, M., Spiegelman, D., Noreau, A., Lafreniere, R. G., Fathalli, F., Joober, R., and 9 others. Truncating mutations in NRXN2 and NRXN1 in autism spectrum disorders and schizophrenia. Hum. Genet. 130: 563-573, 2011. [PubMed: 21424692] [Full Text: https://dx.doi.org/10.1007/s00439-011-0975-z]

  26. Gillberg, C., Wing, L. Autism: not an extremely rare disorder. Acta Psychiat. Scand. 99: 399-406, 1999. [PubMed: 10408260] [Full Text: http://onlinelibrary.wiley.com/resolve/openurl?genre=article&sid=nlm:pubmed&issn=0001-690X&date=1999&volume=99&issue=6&spage=399]

  27. Gilman, S. R., Iossifov, I., Levy, D., Ronemus, M., Wigler, M., Vitkup, D. Rare de novo variants associated with autism implicate a large functional network of genes involved in formation and function of synapses. Neuron 70: 898-907, 2011. [PubMed: 21658583] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0896-6273(11)00439-9]

  28. Girirajan, S., Dennis, M. Y., Baker, C., Malig, M., Coe, B. P., Campbell, C. D., Mark, K., Vu, T. H., Alkan, C., Cheng, Z., Biesecker, L. G., Bernier, R., Eichler, E. E. Refinement and discovery of new hotspots of copy-number variation associated with autism spectrum disorder. Am. J. Hum. Genet. 92: 221-237, 2013. [PubMed: 23375656] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(13)00033-5]

  29. Giulivi, C., Zhang, Y-F., Omanska-Klusek, A., Ross-Inta, C., Wong, S., Hertz-Picciotto, I., Tassone, F., Pessah, I. N. Mitochondrial dysfunction in autism. J.A.M.A. 304: 2389-2396, 2010. [PubMed: 21119085] [Full Text: https://jamanetwork.com/journals/jama/fullarticle/10.1001/jama.2010.1706]

  30. Glessner, J. T., Wang, K., Cai, G., Korvatska, O., Kim, C. E., Wood, S., Zhang, H., Estes, A., Brune, C. W., Bradfield, J. P., Imielinski, M., Frackelton, E. C., and 47 others. Autism genome-wide copy number variation reveals ubiquitin and neuronal genes. Nature 459: 569-573, 2009. [PubMed: 19404257] [Full Text: https://dx.doi.org/10.1038/nature07953]

  31. Greenberg, D. A., Hodge, S. E., Sowinski, J., Nicoll, D. Excess of twins among affected sibling pairs with autism: implications for the etiology of autism. Am. J. Hum. Genet. 69: 1062-1067, 2001. [PubMed: 11590546] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(07)61322-6]

  32. Hallmayer, J., Glasson, E. J., Bower, C., Petterson, B., Croen, L., Grether, J., Risch, N. On the twin risk in autism. Am. J. Hum. Genet. 71: 941-946, 2002. [PubMed: 12297988] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(07)60377-2]

  33. Hallmayer, J., Hebert, J. M., Spiker, D., Lotspeich, L., McMahon, W. M., Petersen, P. B., Nicholas, P., Pingree, C., Lin, A. A., Cavalli-Sforza, L. L., Risch, N., Ciaranello, R. D. Autism and the X chromosome: multipoint sib-pair analysis. Arch. Gen. Psychiat. 53: 985-989, 1996. [PubMed: 8911221] [Full Text: https://jamanetwork.com/journals/jamapsychiatry/fullarticle/vol/53/pg/985]

  34. Institute of Medicine Immunization Safety Reviews. Immunization safety review: measles-mumps-rubella vaccine and autism. Washington, D.C.: National Academy Press (pub.) 2001. Pp. 13-69.

  35. International Molecular Genetic Study of Autism Consortium. A full genome screen for autism with evidence for linkage to a region on chromosome 7q. Hum. Molec. Genet. 7: 571-578, 1998. [PubMed: 9546821] [Full Text: http://hmg.oxfordjournals.org/cgi/pmidlookup?view=long&pmid=9546821]

  36. International Molecular Genetic Study of Autism Consortium. A genomewide screen for autism: strong evidence for linkage to chromosomes 2q, 7q, and 16p. Am. J. Hum. Genet. 69: 570-581, 2001. [PubMed: 11481586] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(07)61167-7]

  37. International Molecular Genetic Study of Autism Consortium. Further characterization of the autism susceptibility locus AUTS1 on chromosome 7q. Hum. Molec. Genet. 10: 973-982, 2001. [PubMed: 11392322]

  38. Iossifov, I., O'Roak, B. J., Sanders, S. J., Ronemus, M., Krumm, N., Levy, D., Stessman, H. A., Witherspoon, K. T., Vives, L., Patterson, K. E., Smith, J. D., Paeper, B., and 35 others. The contribution of de novo coding mutations to autism spectrum disorder. Nature 515: 216-221, 2014. [PubMed: 25363768] [Full Text: https://dx.doi.org/10.1038/nature13908]

  39. Jiang, Y., Yuen, R. K. C., Jin, X., Wang, M., Chen, N., Wu, X., Ju, J., Mei, J., Shi, Y., He, M., Wang, G., Liang, J., and 27 others. Detection of clinically relevant genetic variants in autism spectrum disorder by whole-genome sequencing. Am. J. Hum. Genet. 93: 249-263, 2013. [PubMed: 23849776] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(13)00281-4]

  40. Jones, J. R., Skinner, C., Friez, M. J., Schwartz, C. E., Stevenson, R. E. Hypothesis: dysregulation of methylation of brain-expressed genes on the X chromosome and autism spectrum disorders. Am. J. Med. Genet. 146A: 2213-2220, 2008. [PubMed: 18698615] [Full Text: https://dx.doi.org/10.1002/ajmg.a.32396]

  41. Jorde, L. B., Mason-Brothers, A., Waldmann, R., Ritvo, E. R., Freeman, B. J., Pingree, C., McMahon, W. M., Petersen, B., Jenson, W. R., Mo, A. The UCLA--University of Utah epidemiologic survey of autism: genealogical analysis of familial aggregation. Am. J. Med. Genet. 36: 85-88, 1990. [PubMed: 2333911] [Full Text: http://onlinelibrary.wiley.com/resolve/openurl?genre=article&sid=nlm:pubmed&issn=0148-7299&date=1990&volume=36&issue=1&spage=85]

  42. Kanner, L. Autistic disturbances of affective contact. Nerv. Child 2: 217-250, 1943.

  43. Kilpinen, H., Ylisaukko-oja, T., Rehnstrom, K., Gaal, E., Turunen, J. A., Kempas, E., von Wendt, L., Varilo, T., Peltonen, L. Linkage and linkage disequilibrium scan for autism loci in an extended pedigree from Finland. Hum. Molec. Genet. 18: 2912-2921, 2009. [PubMed: 19454485] [Full Text: https://academic.oup.com/hmg/article-lookup/doi/10.1093/hmg/ddp229]

  44. King, I. F., Yandava, C. N., Mabb, A. M., Hsiao, J. S., Huang, H.-S., Pearson, B. L., Calabrese, J. M., Starmer, J., Parker, J. S., Magnuson, T., Chamberlain, S. J., Philpot, B. D., Zylka, M. J. Topoisomerases facilitate transcription of long genes linked to autism. Nature 501: 58-62, 2013. [PubMed: 23995680] [Full Text: https://dx.doi.org/10.1038/nature12504]

  45. Klauck, S. M., Munstermann, E., Bieber-Martig, B., Ruhl, D., Lisch, S., Schmotzer, G., Poustka, A., Poustka, F. Molecular genetic analysis of the FRM-1 gene in a large collection of autistic patients. Hum. Genet. 100: 224-229, 1997. [PubMed: 9254854] [Full Text: http://link.springer.de/link/service/journals/00439/bibs/7100002/71000224.htm]

  46. Kolevzon, A., Smith, C. J., Schmeidler, J., Buxbaum, J. D., Silverman, J. M. Familial symptom domains in monozygotic siblings with autism. Am. J. Med. Genet. 129B: 76-81, 2004. [PubMed: 15274045] [Full Text: https://dx.doi.org/10.1002/ajmg.b.30011]

  47. Kong, A., Frigge, M. L., Masson, G., Besenbacher, S., Sulem, P., Magnusson, G., Gudjonsson, S. A., Sigurdsson, A., Jonasdottir, A., Jonasdottir, A., Wong, W. S. W., Sigurdsson, G., and 9 others. Rate of de novo mutations and the importance of father's age to disease risk. Nature 488: 471-475, 2012. [PubMed: 22914163] [Full Text: https://dx.doi.org/10.1038/nature11396]

  48. Krumm, N., O'Roak, B. J. Karakoc, E., Mohajeri, K., Nelson, B., Vives, L., Jacquemont, S., Munson, J., Bernier, R., Eichler, E. E. Transmission disequilibrium of small CNVs in simplex autism. Am. J. Hum. Genet. 93: 595-606, 2013. [PubMed: 24035194] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(13)00344-3]

  49. Lainhart, J. E., Ozonoff, S., Coon, H., Krasny, L., Dinh, E., Nice, J., McMahon, W. Autism, regression, and the broader autism phenotype. Am. J. Med. Genet. 113: 231-237, 2002. [PubMed: 12439889] [Full Text: http://onlinelibrary.wiley.com/resolve/openurl?genre=article&sid=nlm:pubmed&issn=0148-7299&date=2002&volume=113&issue=3&spage=231]

  50. Lamb, J. A., Moore, J., Bailey, A., Monaco, A. P. Autism: recent molecular genetic advances. Hum. Molec. Genet. 9: 861-868, 2000. Note: Erratum: Hum. Molec. Genet. 9: 1461 only, 2000. [PubMed: 10767308] [Full Text: http://hmg.oxfordjournals.org/cgi/pmidlookup?view=long&pmid=10767308]

  51. Levy, D., Ronemus, M., Yamrom, B., Lee, Y., Leotta, A., Kendall, J., Marks, S., Lakshmi, B., Pai, D., Ye, K., Buja, A., Krieger, A., Yoon, S., Troge, J., Rodgers, L., Iossifov, I., Wigler, M. Rare de novo and transmitted copy-number variation in autistic spectrum disorders. Neuron 70: 886-897, 2011. [PubMed: 21658582] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0896-6273(11)00396-5]

  52. Levy, S. E., Mandell, D. S., Schultz, R. T. Autism Lancet 374: 1627-1638, 2009. Note: Erratum: Lancet 378: 1546 only, 2011. [PubMed: 19819542] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0140-6736(09)61376-3]

  53. Lim, E. T., Raychaudhuri, S., Sanders, S. J., Stevens, C., Sabo, A., MacArthur, D. G., Neale, B. M., Kirby, A., Ruderfer, D. M., Fromer, M., Lek, M., Liu, L., and 29 others. Rare complete knockouts in humans: population distribution and significant role in autism spectrum disorders. Neuron 77: 235-242, 2013. [PubMed: 23352160] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0896-6273(13)00033-0]

  54. Liu, J., Nyholt, D. R., Magnussen, P., Parano, E., Pavone, P., Geschwind, D., Lord, C., Iversen, P., Hoh, J., Autism Genetic Resource Exchange Consortium, Ott, J., Gilliam, T. C. A genomewide screen for autism susceptibility loci. Am. J. Hum. Genet. 69: 327-340, 2001. Note: Erratum: Am. J. Hum. Genet. 69: 672 only, 2001. [PubMed: 11452361] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(07)61079-9]

  55. Loirat, C., Bellanne-Chantelot, C., Husson, I., Deschenes, G., Guigonis, V., Chabane, N. Autism in three patients with cystic or hyperechogenic kidneys and chromosome 17q12 deletion. Nephrol. Dial. Transplant. 25: 3430-3433, 2010. [PubMed: 20587423] [Full Text: https://academic.oup.com/ndt/article-lookup/doi/10.1093/ndt/gfq380]

  56. Lopreiato, J. O., Wulfsberg, E. A. A complex chromosome rearrangement in a boy with autism. J. Dev. Behav. Pediat. 13: 281-283, 1992. [PubMed: 1506468]

  57. Luo, R., Sanders, S. J., Tian, Y., Voineagu, I., Huang, N., Chu, S. H., Klei, L., Cai, C., Ou, J., Lowe, J. K., Hurles, M. E., Devlin, B., State, M. W., Geschwind, D. H. Genome-wide transcriptome profiling reveals the functional impact of rare de novo and recurrent CNVs in autism spectrum disorders. Am. J. Hum. Genet. 91: 38-55, 2012. [PubMed: 22726847] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(12)00267-4]

  58. Ma, D. Q., Whitehead, P. L., Menold, M. M., Martin, E. R., Ashley-Koch, A. E., Mei, H., Ritchie, M. D., DeLong, G. R., Abramson, R. K., Wright, H. H., Cuccaro, M. L., Hussman, J. P., Gilbert, J. R., Pericak-Vance, M. A. Identification of significant association and gene-gene interaction of GABA receptor subunit genes in autism. Am. J. Hum. Genet. 77: 377-388, 2005. [PubMed: 16080114] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(07)63019-5]

  59. Ma, D., Salyakina, D., Jaworski, J. M., Konidari, I., Whitehead, P. L., Andersen, A. N., Hoffman, J. D., Slifer, S. H., Hedges, D. J., Cukier, H. N., Griswold, A. J., McCauley, J. L., and 9 others. A genome-wide association study of autism reveals a common novel risk locus at 5p14.1. Ann. Hum. Genet. 73: 263-273, 2009. [PubMed: 19456320] [Full Text: https://dx.doi.org/10.1111/j.1469-1809.2009.00523.x]

  60. Maestrini, E., Lai, C., Marlow, A., Matthews, N., Wallace, S., Bailey, A., Cook, E. H., Weeks, D. E., Monaco, A. P., International Molecular Genetic Study of Autism (IMGSA) Consortium. Serotonin transporter (5-HTT) and gamma-aminobutyric acid receptor subunit beta-3 (GABRB3) gene polymorphisms are not associated with autism in the IMGSA families. Am. J. Med. Genet. 88B: 492-496, 1999. [PubMed: 10490705] [Full Text: http://onlinelibrary.wiley.com/resolve/openurl?genre=article&sid=nlm:pubmed&issn=0148-7299&date=1999&volume=88&issue=5&spage=492]

  61. Marshall, C. R., Noor, A., Vincent, J. B., Lionel, A. C., Feuk, L., Skaug, J., Shago, M., Moessner, R., Pinto, D., Ren, Y., Thiruvahindrapduram, B., Fiebig, A., and 23 others. Structural variation of chromosomes in autism spectrum disorder. Am. J. Hum. Genet. 82: 477-488, 2008. [PubMed: 18252227] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(07)00035-3]

  62. Miles, J. H., Takahashi, T. N., Hong, J., Munden, N., Flournoy, N., Braddock, S. R., Martin, R. A., Bocian, M. E., Spence, M. A., Hillman, R. E., Farmer, J. E. Development and validation of a measure of dysmorphology: useful for autism subgroup classification. Am. J. Med. Genet. 146A: 1101-1116, 2008. Note: Erratum: Am. J. Med. Genet. 146A; 2714 only, 2008. [PubMed: 18383511] [Full Text: https://dx.doi.org/10.1002/ajmg.a.32244]

  63. Moessner, R., Marshall, C. R., Sutcliffe, J. S., Skaug, J., Pinto, D., Vincent, J., Zwaigenbaum, L., Fernandez, B., Roberts, W., Szatmari, P., Scherer, S. W. Contribution of SHANK3 mutations to autism spectrum disorder. Am. J. Hum. Genet. 81: 1289-1297, 2007. [PubMed: 17999366] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(07)63777-X]

  64. Moreno-De-Luca, D., SGENE Consortium, Mulle, J. G., Simons Simplex Collection Genetics Consortium, Kaminsky, E. B., Sanders, S. J., GeneSTAR, Myers, S. M., Adam, M. P., Pakula, A. T., Eisenhauer, N. J., Uhas, K., and 26 others. Deletion 17q12 is a recurrent copy number variant that confers high risk of autism and schizophrenia. Am. J. Hum. Genet. 87: 618-630, 2010. Note: Erratum: Am. J. Hum. Genet. 88: 121 only, 2011. [PubMed: 21055719] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(10)00520-3]

  65. Neale, B. M., Kou, Y., Liu, L., Ma'ayan, A., Samocha, K. E., Sabo, A., Lin, C.-F., Stevens, C., Wang, L.-S., Makarov, V., Polak, P., Yoon, S., and 47 others. Patterns and rates of exonic de novo mutations in autism spectrum disorders. Nature 485: 242-245, 2012. [PubMed: 22495311] [Full Text: https://dx.doi.org/10.1038/nature11011]

  66. O'Roak, B. J., Deriziotis, P., Lee, C., Vives, L., Schwartz, J. J., Girirajan, S., Karakoc, E., Mackenzie, A. P., Ng, S. B., Baker, C., Rieder, M. J., Nickerson, D. A., Bernier, R., Fisher, S. E., Shendure, J., Eichler, E. E. Exome sequencing in sporadic autism spectrum disorders identifies severe de novo mutations. Nature Genet. 43: 585-589, 2011. Note: Erratum: Nature Genet. 44: 471 only, 2012. [PubMed: 21572417] [Full Text: https://dx.doi.org/10.1038/ng.835]

  67. O'Roak, B. J., Vives, L., Fu, W., Egertson, J. D., Stanaway, I. B., Phelps, I. G., Carvill, G., Kumar, A., Lee, C., Ankenman, K., Munson, J., Hiatt, J. B., and 14 others. Multiplex targeted sequencing identifies recurrently mutated genes in autism spectrum disorders. Science 338: 1619-1622, 2012. [PubMed: 23160955] [Full Text: http://www.sciencemag.org/cgi/pmidlookup?view=long&pmid=23160955]

  68. O'Roak, B. J., Vives, L., Girirajan, S., Karakoc, E., Krumm, N., Coe, B. P., Levy, R., Ko, A., Lee, C., Smith, J. D., Turner, E. H., Stanaway, I. B., and 11 others. Sporadic autism exomes reveal a highly interconnected protein network of de novo mutations. Nature 485: 246-250, 2012. [PubMed: 22495309] [Full Text: https://dx.doi.org/10.1038/nature10989]

  69. Parikshak, N. N., Swarup, V., Belgard, T. G., Irimia, M., Ramaswami, G., Gandal, M. J., Hartl, C., Leppa, V., de la Torre Ubieta, L., Huang, J., Lowe, J. K., Blencowe, B. J., Horvath, S., Geschwind, D. H. Genome-wide changes in lncRNA, splicing, and regional gene expression patterns in autism. Nature 540: 423-427, 2016. [PubMed: 27919067] [Full Text: https://dx.doi.org/10.1038/nature20612]

  70. Philippe, A., Martinez, M., Guilloud-Bataille, M., Gillberg, C., Rastam, M., Sponheim, E., Coleman, M., Zappella, M., Aschauer, H., van Maldergeme, L., Penet, C., Feingold, J., Brice, A., Leboyer, M., Paris Autism Research International Sibpair Study. Genome-wide scan for autism susceptibility genes. Hum. Molec. Genet. 8: 805-812, 1999. Note: Erratum: Hum. Molec. Genet. 8: 1353 only, 1999. [PubMed: 10196369] [Full Text: http://hmg.oxfordjournals.org/cgi/pmidlookup?view=long&pmid=10196369]

  71. Pickles, A., Bolton, P., Macdonald, H., Bailey, A., Le Couteur, A., Sim, C.-H., Rutter, M. Latent-class analysis of recurrence risks for complex phenotypes with selection and measurement error: a twin and family history study of autism. Am. J. Hum. Genet. 57: 717-726, 1995. [PubMed: 7668301]

  72. Pinto, D., Delaby, E., Merico, D., Barbosa, M., Merikangas, A., Klei, L., Thiruvahindrapuram, B., Xu, X., Ziman, R., Wang, Z., Vorstman, J. A. S., Thompson, A., and 100 others. Convergence of genes and cellular pathways dysregulated in autism spectrum disorders. Am. J. Hum. Genet. 94: 677-694, 2014. [PubMed: 24768552] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(14)00150-5]

  73. Pinto, D., Pagnamenta, A. T., Klei, L., Anney, R., Merico, D., Regan, R., Conroy, J., Magalhaes, T. R., Correia, C., Abrahams, B. S., Almeida, J., Bacchelli, E., and 165 others. Functional impact of global rare copy number variation in autism spectrum disorders. Nature 466: 368-372, 2010. [PubMed: 20531469] [Full Text: https://dx.doi.org/10.1038/nature09146]

  74. Piven, J., Tsai, G. C., Nehme, E., Coyle, J. T., Chase, G. A., Folstein, S. E. Platelet serotonin, a possible marker for familial autism. J. Autism Dev. Disord. 21: 51-59, 1991. [PubMed: 2037549]

  75. Poultney, C. S., Goldberg, A. P., Drapeau, E., Kou, Y., Harony-Nicolas, H., Kajiwara, Y., De Rubeis, S., Durand, S., Stevens, C., Rehnstrom, K., Palotie, A., Daly, M. J., Ma'ayan, A., Fromer, M., Buxbaum, J. D. Identification of small exonic CNV from whole-exome sequence data and application to autism spectrum disorder. Am. J. Hum. Genet. 93: 607-619, 2013. [PubMed: 24094742] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(13)00414-X]

  76. Risch, N., Spiker, D., Lotspeich, L., Nouri, N., Hinds, D., Hallmayer, J., Kalaydjieva, L., McCague, P., Dimiceli, S., Pitts, T., Nguyen, L., Yang, J., and 19 others. A genomic screen of autism: evidence for a multilocus etiology. Am. J. Hum. Genet. 65: 493-507, 1999. [PubMed: 10417292] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(07)62066-7]

  77. Ritvo, E. R., Freeman, B. J., Mason-Brothers, A. M., Mo, A., Ritvo, A. M. Concordance for the syndrome of autism in 40 pairs of affected twins. Am. J. Psychiat. 142: 74-77, 1985. [PubMed: 4038442] [Full Text: http://psychiatryonline.org/doi/abs/10.1176/ajp.142.1.74?url_ver=Z39.88-2003&rfr_id=ori:rid:crossref.org&rfr_dat=cr_pub%3dpubmed]

  78. Ritvo, E. R., Spence, M. A., Freeman, B. J., Mason-Brothers, A., Mo, A., Marazita, M. L. Evidence for an autosomal recessive type of autism in 46 multiple incidence families. Am. J. Psychiat. 142: 187-192, 1985. [PubMed: 4038589] [Full Text: http://psychiatryonline.org/doi/abs/10.1176/ajp.142.2.187?url_ver=Z39.88-2003&rfr_id=ori:rid:crossref.org&rfr_dat=cr_pub%3dpubmed]

  79. Roohi, J., Montagna, C., Tegay, D. H., Palmer, L. E., DeVincent, C., Pomeroy, J. C., Christian, S. L., Nowak, N., Hatchwell, E. Disruption of contactin 4 in three subjects with autism spectrum disorder. J. Med. Genet. 46: 176-182, 2009. [PubMed: 18349135] [Full Text: http://jmg.bmj.com/cgi/pmidlookup?view=long&pmid=18349135]

  80. Sanders, S. J., Ercan-Sencicek, A. G., Hus, V., Luo, R., Murtha, M. T., Moreno-De-Luca, D., Chu, S. H., Moreau, M. P., Gupta, A. R., Thomson, S. A., Mason, C. E., Bilguvar, K., and 55 others. Multiple recurrent de novo CNVs, including duplications of the 7q11.23 Williams syndrome region, are strongly associated with autism. Neuron 70: 863-885, 2011. [PubMed: 21658581] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0896-6273(11)00374-6]

  81. Sanders, S. J., Murtha, M. T., Gupta, A. R., Murdoch, J. D., Raubeson, M. J., Willsey, A. J., Ercan-Sencicek, A. G., DiLullo, N. M., Parikshak, N. N., Stein, J. L., Walker, M. F., Ober, G. T., and 18 others. De novo mutations revealed by whole-exome sequencing are strongly associated with autism. Nature 485: 237-241, 2012. [PubMed: 22495306] [Full Text: https://dx.doi.org/10.1038/nature10945]

  82. Sandin, S., Lichtenstein, P., Kuja-Halkola, R., Larsson, H., Hultman, C. M., Reichenberg, A. The familial risk of autism. JAMA 311: 1770-1777, 2014. [PubMed: 24794370] [Full Text: https://jamanetwork.com/journals/jama/fullarticle/10.1001/jama.2014.4144]

  83. Schaefer, G. B., Thompson, J. N., Jr., Bodensteiner, J. B., McConnell, J. M., Kimberling, W. J., Gay, C. T., Dutton, W. D., Hutchings, D. C., Gray, S. B. Hypoplasia of the cerebellar vermis in neurogenic syndromes. Ann. Neurol. 39: 382-385, 1996. [PubMed: 8602758] [Full Text: http://onlinelibrary.wiley.com/resolve/openurl?genre=article&sid=nlm:pubmed&issn=0364-5134&date=1996&volume=39&issue=3&spage=382]

  84. Schain, R. J., Freedman, D. X. Studies on 5-hydroxyindole metabolism in autistic and other mentally retarded children. J. Pediat. 58: 315-320, 1961. [PubMed: 13747230]

  85. Schellenberg, G. D., Dawson, G., Sung, Y. J., Estes, A., Munson, J., Rosenthal, E., Rothstein, J., Flodman, P., Smith, M., Coon, H., Leong, L., Yu, C.-E., Stodgell, C., Rodier, P. M., Spence, M. A., Minshew, N., McMahon, W. M., Wijsman, E. M. Evidence for multiple loci from a genome scan of autism kindreds. Molec. Psychiat. 11: 1049-1060, 2006. [PubMed: 16880825] [Full Text: https://dx.doi.org/10.1038/sj.mp.4001874]

  86. Sebat, J., Lakshmi, B., Malhotra, D., Troge, J., Lese-Martin, C., Walsh, T., Yamrom, B., Yoon, S., Krasnitz, A., Kendall, J., Leotta, A., Pai, D., and 20 others. Strong association of de novo copy number mutations with autism. Science 316: 445-449, 2007. [PubMed: 17363630] [Full Text: http://www.sciencemag.org/cgi/pmidlookup?view=long&pmid=17363630]

  87. Silverman, J. M., Smith, C. J., Schmeidler, J., Hollander, E., Lawlor, B. A., Fitzgerald, M., Buxbaum, J. D., Delaney, K., Galvin, P., Autism Genetic Research Exchange Consortium. Symptom domains in autism and related conditions: evidence for familiality. Am. J. Med. Genet. 114B: 64-73, 2002. [PubMed: 11840508] [Full Text: http://onlinelibrary.wiley.com/resolve/openurl?genre=article&sid=nlm:pubmed&issn=0148-7299&date=2002&volume=114&issue=1&spage=64]

  88. Smalley, S. L. Genetic influences in childhood-onset psychiatric disorders: autism and attention-deficit/hyperactivity disorder. Am. J. Hum. Genet. 60: 1276-1282, 1997. [PubMed: 9199546] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(07)64217-7]

  89. Spence, M. A., Ritvo, E. R., Marazita, M. L., Funderburk, S. J., Sparkes, R. S., Freeman, B. J. Gene mapping studies with the syndrome of autism. Behav. Genet. 15: 1-13, 1985. [PubMed: 3985910]

  90. Tabuchi, K., Blundell, J., Etherton, M. R., Hammer, R. E., Liu, X., Powell, C. M., Sudhof, T. C. A neuroligin-3 mutation implicated in autism increases inhibitory synaptic transmission in mice. Science 318: 71-76, 2007. [PubMed: 17823315] [Full Text: http://www.sciencemag.org/cgi/pmidlookup?view=long&pmid=17823315]

  91. Taylor, B., Miller, E., Lingam, R., Andrews, N., Simmons, A., Stowe, J. Measles, mumps, and rubella vaccination and bowel problems or developmental regression in children with autism: population study. Brit. Med. J. 324: 393-396, 2002. [PubMed: 11850369] [Full Text: http://www.bmj.com/cgi/pmidlookup?view=long&pmid=11850369]

  92. Trikalinos, T. A., Karvouni, A., Zintzaras, E., Ylisaukko-oja, T., Peltonen, L., Jarvela, I., Ioannidis, J. P. A. A heterogeneity-based genome search meta-analysis for autism-spectrum disorders. Molec. Psychiat. 11: 29-36, 2006. [PubMed: 16189507] [Full Text: https://dx.doi.org/10.1038/sj.mp.4001750]

  93. Vaags, A. K., Lionel, A. C., Sato, D., Goodenberger, M., Stein, Q. P., Curran, S., Ogilvie, C., Ahn, J. W., Drmic, I., Senman, L., Chrysler, C., Thompson, A., and 15 others. Rare deletions at the neurexin 3 locus in autism spectrum disorder. Am. J. Hum. Genet. 90: 133-141, 2012. [PubMed: 22209245] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(11)00503-9]

  94. Vincent, J. B., Horike, S. I., Choufani, S., Paterson, A. D., Roberts, W., Szatmari, P., Weksberg, R., Fernandez, B., Scherer, S. W. An inversion inv(4)(p12-p15.3) in autistic siblings implicates the 4p GABA receptor gene cluster. J. Med. Genet. 43: 429-434, 2006. [PubMed: 16556609] [Full Text: http://jmg.bmj.com/cgi/pmidlookup?view=long&pmid=16556609]

  95. Voineagu, I., Wang, X., Johnston, P., Lowe, J. K., Tian, Y., Horvath, S., Mill, J., Cantor, R. M., Blencowe, B. J., Geschwind, D. H. Transcriptomic analysis of autistic brain reveals convergent molecular pathology. Nature 474: 380-384, 2011. [PubMed: 21614001] [Full Text: https://dx.doi.org/10.1038/nature10110]

  96. Wang, K., Zhang, H., Ma, D., Bucan, M., Glessner, J. T., Abrahams, B. S., Salyakina, D., Imielinski, M., Bradfield, J. P., Sleiman, P. M. A., Kim, C. E., Hou, C., and 44 others. Common genetic variants on 5p14.1 associate with autism spectrum disorders. Nature 459: 528-533, 2009. [PubMed: 19404256] [Full Text: https://dx.doi.org/10.1038/nature07999]

  97. Weiss, L. A., Arking, D. E., Gene Discovery Project of Johns Hopkins & the Autism Consortium. A genome-wide linkage and association scan reveals novel loci for autism. Nature 461: 802-808, 2009. [PubMed: 19812673] [Full Text: https://dx.doi.org/10.1038/nature08490]

  98. Yonan, A. L., Alarcon, M., Cheng, R., Magnusson, P. K. E., Spence, S. J., Palmer, A. A., Grunn, A., Juo, S.-H. H., Terwilliger, J. D., Liu, J., Cantor, R. M., Geschwind, D. H., Gilliam, T. C. A genomewide screen of 345 families for autism-susceptibility loci. Am. J. Hum. Genet. 73: 886-897, 2003. [PubMed: 13680528] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/AJHG40053]

  99. Yu, C.-E., Dawson, G., Munson, J., D'Souza, I., Osterling, J., Estes, A., Leutenegger, A.-L., Flodman, P., Smith, M., Raskind, W. H., Spence, M. A., McMahon, W., Wijsman, E. M., Schellenberg, G. D. Presence of large deletions in kindreds with autism. Am. J. Hum. Genet. 71: 100-115, 2002. [PubMed: 12058345] [Full Text: https://linkinghub.elsevier.com/retrieve/pii/S0002-9297(07)60038-X]


Contributors:
Ada Hamosh - updated : 01/10/2017
Paul J. Converse - updated : 07/28/2016
Ada Hamosh - updated : 6/3/2016
Cassandra L. Kniffin - updated : 3/24/2016
Ada Hamosh - updated : 4/17/2015
Ada Hamosh - updated : 1/20/2015
Ada Hamosh - updated : 6/25/2014
Ada Hamosh - updated : 6/4/2014
Ada Hamosh - updated : 1/16/2014
Ada Hamosh - updated : 1/14/2014
Ada Hamosh - updated : 1/13/2014
Ada Hamosh - updated : 11/21/2013
Ada Hamosh - updated : 11/4/2013
Ada Hamosh - updated : 10/16/2013
Ada Hamosh - updated : 1/23/2013
Ada Hamosh - updated : 9/5/2012
Ada Hamosh - updated : 6/27/2012
Ada Hamosh - updated : 6/25/2012
Cassandra L. Kniffin - updated : 6/19/2012
Marla J. F. O'Neill - updated : 3/14/2012
Cassandra L. Kniffin - updated : 2/15/2012
George E. Tiller - updated : 11/14/2011
Cassandra L. Kniffin - updated : 11/8/2011
Ada Hamosh - updated : 10/4/2011
Ada Hamosh - updated : 7/26/2011
Cassandra L. Kniffin - updated : 4/18/2011
Ada Hamosh - updated : 11/8/2010
Ada Hamosh - updated : 8/17/2010
George E. Tiller - updated : 6/23/2010
Cassandra L. Kniffin - updated : 3/29/2010
Ada Hamosh - updated : 3/26/2010
George E. Tiller - updated : 2/22/2010
Cassandra L. Kniffin - updated : 12/10/2009
Ada Hamosh - updated : 8/24/2009
Ada Hamosh - updated : 8/17/2009
Cassandra L. Kniffin - updated : 1/5/2009
Cassandra L. Kniffin - updated : 8/20/2008
Cassandra L. Kniffin - updated : 4/18/2008
Ada Hamosh - updated : 10/26/2007
Ada Hamosh - updated : 5/30/2007
Cassandra L. Kniffin - updated : 3/12/2007
Cassandra L. Kniffin - updated : 8/29/2006
Cassandra L. Kniffin - updated : 8/18/2006
Cassandra L. Kniffin - updated : 6/13/2006
John Logan Black, III - updated : 4/6/2006
Marla J. F. O'Neill - updated : 10/6/2005
Cassandra L. Kniffin - updated : 6/28/2005
Victor A. McKusick - updated : 3/23/2005
Victor A. McKusick - updated : 3/11/2005
John Logan Black, III - updated : 3/1/2005
Victor A. McKusick - updated : 11/12/2004
Cassandra L. Kniffin - updated : 5/24/2004
Cassandra L. Kniffin - reorganized : 5/17/2004
Cassandra L. Kniffin - updated : 5/6/2004
John Logan Black, III - updated : 3/11/2004
Victor A. McKusick - updated : 12/1/2003
Victor A. McKusick - updated : 8/15/2003
Victor A. McKusick - updated : 2/28/2003
Victor A. McKusick - updated : 11/27/2002
Victor A. McKusick - updated : 11/14/2002
John Logan Black, III - updated : 8/14/2002
Victor A. McKusick - updated : 8/2/2002
Victor A. McKusick - updated : 4/12/2002
Victor A. McKusick - updated : 2/21/2002
Victor A. McKusick - updated : 2/14/2002
Victor A. McKusick - updated : 2/4/2002
Ada Hamosh - updated : 1/30/2002
Victor A. McKusick - updated : 1/22/2002
Michael B. Petersen - updated : 12/5/2001
Victor A. McKusick - updated : 11/27/2001
Victor A. McKusick - updated : 10/25/2001
George E. Tiller - updated : 10/2/2001
Victor A. McKusick - updated : 9/7/2001
Victor A. McKusick - updated : 10/3/2000
George E. Tiller - updated : 5/1/2000
Victor A. McKusick - updated : 1/11/2000
Victor A. McKusick - updated : 5/17/1999
Orest Hurko - updated : 3/24/1999
Victor A. McKusick - updated : 4/24/1998
Victor A. McKusick - updated : 9/12/1997
Victor A. McKusick - updated : 6/17/1997
Orest Hurko - updated : 5/6/1996
Orest Hurko - updated : 8/2/1995
Creation Date:
Victor A. McKusick : 6/3/1986
Edit History:
carol : 03/02/2017
carol : 02/28/2017
carol : 02/28/2017
carol : 01/11/2017
alopez : 01/10/2017
mgross : 07/28/2016
carol : 07/09/2016
alopez : 6/3/2016
alopez : 4/12/2016
carol : 3/24/2016
ckniffin : 3/24/2016
alopez : 4/17/2015
alopez : 1/20/2015
carol : 8/21/2014
alopez : 8/20/2014
mcolton : 8/19/2014
ckniffin : 8/14/2014
alopez : 6/25/2014
alopez : 6/4/2014
carol : 4/1/2014
alopez : 1/16/2014
alopez : 1/14/2014
alopez : 1/13/2014
alopez : 11/21/2013
mcolton : 11/13/2013
alopez : 11/4/2013
alopez : 10/16/2013
tpirozzi : 10/1/2013
tpirozzi : 10/1/2013
tpirozzi : 10/1/2013
tpirozzi : 10/1/2013
carol : 9/12/2013
ckniffin : 5/16/2013
carol : 4/19/2013
terry : 3/15/2013
alopez : 1/24/2013
terry : 1/23/2013
terry : 10/2/2012
alopez : 9/6/2012
terry : 9/5/2012
terry : 7/10/2012
terry : 7/6/2012
terry : 7/6/2012
alopez : 7/3/2012
alopez : 7/2/2012
alopez : 7/2/2012
terry : 6/27/2012
terry : 6/25/2012
carol : 6/20/2012
ckniffin : 6/19/2012
carol : 6/7/2012
ckniffin : 6/6/2012
alopez : 6/6/2012
terry : 3/16/2012
carol : 3/14/2012
carol : 2/21/2012
ckniffin : 2/15/2012
carol : 11/18/2011
terry : 11/14/2011
carol : 11/9/2011
ckniffin : 11/8/2011
alopez : 10/12/2011
terry : 10/4/2011
terry : 9/28/2011
alopez : 8/16/2011
terry : 7/26/2011
alopez : 6/16/2011
wwang : 4/21/2011
ckniffin : 4/18/2011
carol : 1/21/2011
alopez : 1/13/2011
ckniffin : 1/12/2011
alopez : 1/6/2011
alopez : 11/10/2010
terry : 11/8/2010
terry : 11/8/2010
alopez : 8/20/2010
terry : 8/17/2010
wwang : 7/6/2010
terry : 6/23/2010
alopez : 6/14/2010
alopez : 6/10/2010
ckniffin : 6/9/2010
alopez : 5/21/2010
wwang : 4/30/2010
ckniffin : 4/19/2010
wwang : 4/6/2010
ckniffin : 3/29/2010
alopez : 3/26/2010
wwang : 2/25/2010
terry : 2/22/2010
carol : 2/18/2010
wwang : 12/10/2009
wwang : 12/3/2009
ckniffin : 11/5/2009
wwang : 9/22/2009
alopez : 8/24/2009
alopez : 8/24/2009
terry : 8/17/2009
wwang : 1/7/2009
ckniffin : 1/5/2009
wwang : 8/26/2008
ckniffin : 8/20/2008
alopez : 6/6/2008
wwang : 4/24/2008
ckniffin : 4/18/2008
alopez : 3/21/2008
terry : 3/19/2008
wwang : 1/11/2008
terry : 1/3/2008
alopez : 11/1/2007
terry : 10/26/2007
carol : 10/10/2007
alopez : 6/14/2007
terry : 5/30/2007
alopez : 5/16/2007
carol : 5/14/2007
carol : 5/14/2007
ckniffin : 5/10/2007
ckniffin : 3/12/2007
carol : 3/6/2007
ckniffin : 3/5/2007
carol : 11/28/2006
carol : 11/27/2006
wwang : 9/7/2006
ckniffin : 8/29/2006
wwang : 8/25/2006
ckniffin : 8/18/2006
wwang : 6/16/2006
ckniffin : 6/13/2006
wwang : 4/10/2006
terry : 4/6/2006
carol : 1/18/2006
terry : 12/21/2005
carol : 12/12/2005
terry : 11/10/2005
wwang : 10/18/2005
alopez : 10/17/2005
terry : 10/6/2005
ckniffin : 6/28/2005
ckniffin : 5/23/2005
tkritzer : 3/25/2005
ckniffin : 3/24/2005
ckniffin : 3/24/2005
terry : 3/23/2005
wwang : 3/15/2005
terry : 3/11/2005
carol : 3/10/2005
tkritzer : 3/1/2005
alopez : 11/18/2004
terry : 11/12/2004
tkritzer : 5/28/2004
ckniffin : 5/24/2004
ckniffin : 5/19/2004
carol : 5/18/2004
ckniffin : 5/18/2004
carol : 5/17/2004
ckniffin : 5/17/2004
ckniffin : 5/6/2004
carol : 3/23/2004
joanna : 3/17/2004
terry : 3/11/2004
tkritzer : 12/8/2003
terry : 12/1/2003
alopez : 8/19/2003
terry : 8/15/2003
terry : 8/15/2003
joanna : 5/12/2003
ckniffin : 4/1/2003
ckniffin : 4/1/2003
carol : 3/31/2003
tkritzer : 3/6/2003
terry : 2/28/2003
tkritzer : 12/5/2002
tkritzer : 12/2/2002
terry : 11/27/2002
carol : 11/20/2002
carol : 11/20/2002
terry : 11/14/2002
terry : 11/14/2002
carol : 8/14/2002
tkritzer : 8/7/2002
tkritzer : 8/7/2002
tkritzer : 8/5/2002
terry : 8/2/2002
alopez : 4/26/2002
cwells : 4/19/2002
terry : 4/12/2002
mgross : 2/25/2002
mgross : 2/25/2002
terry : 2/21/2002
carol : 2/21/2002
cwells : 2/21/2002
cwells : 2/15/2002
terry : 2/14/2002
carol : 2/11/2002
terry : 2/4/2002
alopez : 2/4/2002
terry : 1/30/2002
carol : 1/30/2002
terry : 1/22/2002
cwells : 12/5/2001
alopez : 11/30/2001
terry : 11/27/2001
carol : 10/25/2001
terry : 10/25/2001
cwells : 10/10/2001
cwells : 10/2/2001
cwells : 9/20/2001
cwells : 9/13/2001
alopez : 9/7/2001
alopez : 9/7/2001
carol : 5/3/2001
terry : 10/5/2000
terry : 10/3/2000
alopez : 5/1/2000
mgross : 2/7/2000
terry : 1/11/2000
mgross : 6/4/1999
mgross : 6/4/1999
mgross : 5/25/1999
terry : 5/17/1999
carol : 3/24/1999
dholmes : 5/12/1998
dholmes : 5/11/1998
carol : 4/24/1998
terry : 4/14/1998
mark : 9/19/1997
terry : 9/12/1997
terry : 9/10/1997
terry : 9/10/1997
mark : 6/18/1997
terry : 6/17/1997
terry : 6/12/1997
terry : 6/12/1997
mark : 5/6/1996
mark : 5/6/1996
terry : 4/30/1996
mark : 9/10/1995
terry : 4/19/1995
jason : 6/22/1994
supermim : 3/16/1992
carol : 3/7/1992