Skip to main content
Erschienen in: Molecular Autism 1/2017

Open Access 01.12.2017 | Review

Bio-collections in autism research

verfasst von: Jamie Reilly, Louise Gallagher, June L. Chen, Geraldine Leader, Sanbing Shen

Erschienen in: Molecular Autism | Ausgabe 1/2017

Abstract

Autism spectrum disorder (ASD) is a group of complex neurodevelopmental disorders with diverse clinical manifestations and symptoms. In the last 10 years, there have been significant advances in understanding the genetic basis for ASD, critically supported through the establishment of ASD bio-collections and application in research. Here, we summarise a selection of major ASD bio-collections and their associated findings. Collectively, these include mapping ASD candidate genes, assessing the nature and frequency of gene mutations and their association with ASD clinical subgroups, insights into related molecular pathways such as the synapses, chromatin remodelling, transcription and ASD-related brain regions. We also briefly review emerging studies on the use of induced pluripotent stem cells (iPSCs) to potentially model ASD in culture. These provide deeper insight into ASD progression during development and could generate human cell models for drug screening. Finally, we provide perspectives concerning the utilities of ASD bio-collections and limitations, and highlight considerations in setting up a new bio-collection for ASD research.
Abkürzungen
ADI-R
Autism Diagnostic Interview–Revised
ADOS
Autistic Diagnostic Observation Schedule
AGP
Autism Genome Project
AGRE
Autism Genetic Resource Exchange
AIC
Autism Inpatient Collection
ASD
Autism spectrum disorders
ATP
Autism tissue program
CAN
Cure Autism Now Foundation
CCGSMD
Centre for Collaborative Genetic Studies on Mental Disorders
CNV
Copy number variation
DBC
Danish Birth Cohort
DBSS
Dried blood spot samples
DNSB
Danish Newborn Screening Biobank
DSM-5
Diagnostic and Statistical Manual of Mental Disorders
GSH
Glutathione
GWAS
Genome-wide association study
HAWAS
Histone acetylome-wide association study
HBDI
Human Biological Data Interchange
ICD
International Statistical Classification of Diseases and Related Health Problems
iPSC
Induced pluripotent stem cells
MMP
Matrix metalloproteinase
MPLA
Multiplex ligation-dependent probe amplification
PDD-NOS
Pervasive developmental disorder not otherwise specified
RUCDR
Rutgers University Cell and DNA Repository
SCD
Social (pragmatic) communication disorder
SNP
Single nucleotide polymorphism
SNV
Single nucleotide variation
SSC
Simons Simplex Collection
TASC
Autism Simplex Collection
WES
Whole exome sequencing study

Background

Autism spectrum disorder (ASD) is a group of early onset and heterogeneous neurodevelopmental disorders affecting males (1/42) more often than females (1/189) [1]. The prevalence of ASD has risen rapidly; from 0.5/1000 people in early epidemiological studies of 1960–1970 [2, 3] to 1/68 children of school age according to recent data from the Centre for Disease Control [1].
ASD is characterised by atypical development of social behaviour, communication deficits and the presence of repetitive and stereotyped behaviours [4]. It is highly clinically heterogeneous and accompanied by commonly occurring comorbidities that are not core to the disorder but frequently disabling. Communication deficit also persists in social communication disorder (SCD), and the new diagnosis of SCD (DSM-5) makes it possible to distinguish ASD from SCD individuals. The severity may vary across a range of parameters including ASD symptoms, IQ and comorbid behaviours [4]. For example, 70% ASD patients will have at least 1 comorbid psychiatric disorder [5], such as social anxiety, depression and bipolar disorder [6]. In addition, ASD is frequently associated with epilepsy, gastrointestinal and immune disorders [7].
ASD is a highly heritable complex polygenic condition. Estimated heritability based on family and twin studies are 50–80% [8, 9]. It is strongly linked to genetic factors involving the development and function of the nervous system [10], mitochondrial function [11], the immune system [12] and epigenetic regulations [13]. Genetic risk is attributed to rare copy number variants (CNV) and single nucleotide variants (SNV) acting on the background of common genetic variation (reviewed by [14]). High throughput genome sequencing technologies have facilitated genomic discovery, and advanced bioinformatics methodologies have enabled investigation of protein-protein interactions [15, 16] and functionally related pathways [17, 18]. The pathway to gene discovery has required large-scale international collaborative efforts based on the assembly of large bio-collections that are now publicly available and the subject of this review. In parallel to bio-collections, large-scale patient registries have provided epidemiological data that illustrate the course and prognosis of ASD and are helping to identify environmental factors influencing the aetiology [1922]. Despite the advances, significant gaps in our knowledge of the aetiology remain and effective treatments for core ASD symptoms are elusive. The genetic and clinical heterogeneity of ASD means that further advancement will require larger bio-collections coupled with rich clinical data, ideally longitudinally to obtain a clear picture of the disorder both on the molecular and physiological levels.

Autism bio-collections

A bio-collection is a large set of biologically characterised samples, such as blood or tissue collected from a group of individuals who typically have a specific medical condition. Bio-collections are useful as a dedicated resource to generate clinical and scientific data for the analysis of medical conditions on a large scale [23], as well as to create functional disease models to explore the biology of clinical conditions. Large-scale bio-collections and associated comprehensive data that can aid the interrogation of the relationship between the genotype and phenotype effects at the individual and group levels can address the issue of heterogeneity. The purpose of this review is to provide a summary of the publicly available ASD bio-collections, to highlight the impact of these on ASD research and to identify new directions for ASD bio-collection for future research purposes.

Methods and search criteria

A literature search was conducted amongst published studies from Jan 2001 to Nov 2016 on electronic databases of Web of Science, EBSCO, PubMed, Science Direct, MEDLINE, Wiley Online Library. The search terms included “biobank”, “registry”, “collection”, “autism” and citation of bio-collections. A total of 263 studies from ASD bio-collections have been included in the tables and references of this review (Tables 2, 3, 4, 5 and 6).

Inclusion criteria

This review included (a) studies using original samples of human tissues in ASD bio-collections; (b) studies using bio-samples extracted from systematically collected bio-resources (i.e. DNA, RNA, protein) for investigating the risk or influence of ASD; (c) the population studies involving participants of autism, Asperger and pervasive developmental disorder not otherwise specified (PDD-NOS); (d) studies published in peer-reviewed journals and (e) in English.

Exclusion criteria

Studies were excluded (a) if they did not mention the collection(s) in the research data, references, acknowledgements or supplementary materials; (b) if the bio-samples were not derived from a systematic sample collection; and (d) if studies only concerned animal models of ASD without using ASD bio-collections or data.
We focus largely on studies from five bio-collections, four providing DNA, cell lines and metabolites, the Autism Genetic Resource Exchange (AGRE), Simons Simplex Collection (SSC), The Danish Newborn Screening Biobank (DNSB) and The Autism Simplex Collection (TASC) one providing brain tissue, Autism BrainNet (formerly the Autism Tissue Program (ATP)).We also included two emerging bio-collections that have fewer or no publications released yet, but could be of significant impact in the future. They are the Autism Inpatient Collection (AIC) [24] and the Autism Spectrum Stem Cell Resource [25]. An overview of the bio-collections and their website links can be found in Table 1.
Table 1
Information on the biobanks covered in the review
Name
Founded
Location
Type of sample collected and stored
Website/source paper
Autism Genetic Resource Exchange (AGRE)
1997
USA
Blood and immortalised cell lines
Simons Simplex Collection (SSC)
2010
USA
Blood and immortalised cell lines
Danish Newborn Screening Bio-collection
1980’s
Denmark
Dried blood spot samples
Autism Tissue Program
1998
USA
Post-mortem brain tissue
Autism Spectrum Stem cell Resource
2014
USA
Skin fibroblasts, blood, induced pluripotent stem cells, neural stem cells, neuronal cells, glial cells
[25]
The Autism Simplex Collection
2010
USA and Europe
Blood
[97]
Autism Inpatient Collection
2014
USA
Blood and lymphoblasts
[24]

Results

Autism Genetic Resource Exchange (AGRE)

AGRE was established in 1997 by the Cure Autism Now (CAN) Foundation and the Human Biological Data Interchange (HBDI). Samples are provided by families with children affected by ASD and are coupled with anonymously coded clinical diagnostic data, such as Autism Diagnostic Interview–Revised (ADI–R) and Autism Diagnostic Observational Schedule (ADOS). Additional clinical data include photographic dysmorphology, neurological and physical examination, and family and medical history. AGRE is currently managed by Autism Speaks. It contains over 2500 families and the resource has contributed to high profile genetic discoveries relating to ASD (Table 2). Samples are housed at the National Institute of Mental Health repository at Rutgers’ University in the form of immortalised cell lines, DNA and serum samples which can be accessed by researchers through applications [20].
Table 2
Overview of studies using the AGRE collection
Reference
Bio-collection
Samples
Number
Study
Findings
[35]
AGRE
Genomic data (AGRE), brain tissue (mouse)
4327 samples (AGRE)
8789 samples (total)
Genotype-phenotype study
HMGN1 found to be a negative regulator of MECP2 expression. Dysregulation alters behaviour in mice, and AGRE collection contains genotypes linked to altered expression
[180]
AGRE
Blood
152 subjects
Quantitative trait analysis
Chromosome region 7q found to be a risk region for Autism Symptoms
[181]
AGRE
Lymphoblasts
1438 subjects
Association study
CNTNAP2 identified as an ASD susceptibility gene
[182]
AGRE
Blood
1794 subjects
Linkage analysis
Chromosome 7q35 may harbour a gene that could contribute to variability in spoken language
[183]
AGRE
Genomic data
455 subjects
Pedigree study
Association found with chromosome X region Xp22.11-P21.2, where gene IL1RAPL1 is located and also implicated in ASD
[184]
AGRE
Blood and lymphoblasts
252 families
Gene expression analysis and association
ROBO1-4 found to be associated with ASD. Low expression levels of ROBO1-2 found in ASD patients
[185]
AGRE
Blood and lymphoblasts
3211 subjects
Gene association study
Analysis of SNP polymorphisms in PCDHA suggest it as a potential candidate gene for ASD
[186]
AGRE and ATP
Lymphoblasts and brain tissue
3211 subjects (AGRE)
21 subjects (ATP)
Gene association study
ZNF804A found to be associated with ASD and verbal deficits, where knockdown of this gene reduced expression of SNAP25, and both are reduced in the anterior cingulate gyrus in ASD brains.
[187]
AGRE
Blood and lymphoblasts
72 families
Association study
Common variant found in CNTNAP2 that is linked to ASD susceptibility
[43]
AGRE
Blood
470 families (total) 224(AGRE)
Association study
GABRB3 and GABRG3 found to be associated with ASD
[188]
AGRE
Blood and lymphoblasts
975 subjects
CNV analysis
Analysis of 15q13.1-3 region revealed APBA2 as an ASD candidate gene
[189]
AGRE
Blood and lymphoblasts
1577 subjects (total)
1526 subjects (AGRE)
CNV analysis
CNTNAP2 detected in ASD patients suggested to have a contribution to the disorder
[74]
AGRE
Lymphoblasts
6 subjects
Proof of principle
48 genes showed differential expression between patients and controls. Many genes involved in signalling, focal adhesion and metabolism
[190]
AGRE
Lymphoblasts
18 subjects* (controls provided by AGRE)
Profiling study
Altered levels of UBE3A (1.5–2 fold increase) expression found in ASD patients with 15q11-14 duplications. APP and SUMO found to be decreased, and are involved in apoptosis
[40]
AGRE
Blood and lymphoblasts
334 families
Reanalysis of data set using different analysis method
Association found in chromosome 1, which was previously overlooked. Further evidence that 17q11 is associated with ASD
[191]
AGRE
Genomic data
12 families
Method paper
Description of parent of origin method to detect mosaic chromosomal abnormalities.
[192]
AGRE
Blood and lymphoblasts
518 families
Replication study and functional study
The gene EN2 suggested to act as ASD susceptibility locus, and mutations could alter brain development
[41]
AGRE
Blood and lymphoblasts
389 families (AGRE) 518 families (total)
Association study
Haplotypes found in ASD families found to affect regulation of EN2 gene expression
[75]
AGRE
Blood and lymphoblasts
954 subjects
Gene-gene interaction study
Glutathione pathway is implicated in autism
[28]
AGRE
Blood and lymphoblasts
6056 subjects (TOTAL)
4444 subjects (AGRE)
GWAS
UBE3A, NRXN1, BZRAP, and MDGA2 found to have disruptive CNVs amongst ASD patients, some only occurring once amongst patients
[83]
AGRE
Genomic data
830 subjects
Methods paper
Use of disease symptoms improves detection of linkage in genetic data. Useful when heterogeneity is involved
[38]
AGRE
Blood
18 subjects
Genotype-phenotype study
3 out of 18 patients with ASD and macrocephaly had mutations in PTEN gene. Considered as ASD gene to be explored
[39]
AGRE
Blood and lymphoblasts
88 subjects (total)
39 subjects (AGRE)
Mutation screening
De novo missense mutation found in one patient with ASD and macrocephaly.
[193]
AGRE
Blood and lymphoblasts
95 families
Gene linkage study
Chromosome region 2q suggested to contain an autism susceptibility gene
[53]
AGRE
Blood
88 families(total)
62 families (AGRE)
Linkage analysis
GABRB3 polymorphism found to be associated with ASD
[194]
AGRE
Blood
115 families
Linkage analysis
Analysis carried out for a ASD family subset with obsessive compulsive behaviours (n = 35) found evidence of linkage to chromosome 1 and further evidence on chromosome 6 and 19
[82]
AGRE
Blood and lymphoblasts
279 subjects
Method paper
Multiplex ligation-dependent probe amplification shown to be effective at detecting microduplications and deletions
[50]
AGRE
Genomic data
748 subjects
Association study
MET variants associated with social and communication phenotypes amongst people ASD
[49]
AGRE
Blood and lymphoblasts
2712 subjects (total)
631 subjects (AGRE)
Association study
Multiple genes implicated in the MET pathway with ASD, such as PLAUR and SERPINE1
[48]
AGRE
Blood
743 families (total)
283(AGRE)
Association study
MET promoter variant that decreases expression found to be associated with ASD
[195]
AGRE
Blood and lymphoblasts
109 subjects
Replication study
Independent sample from the same cohort showed same linkage association to chromosome region 17q21
[196]
AGRE
Blood
480 families
Genetic score study
3 risk SNPs (ATP2B2, PITX1, HOXA1) had high reproducibility in males, 2 in females (MARK1, ITGB3), and 3 across both genders (CTNAP2, JARID2, EN2).
[197]
AGRE
Blood
381 subjects
Association study
Association between ASD in males and ATP2B2
[198]
AGRE
Blood
2569 subjects
Functional genomics study
Combining functional genomics and statistical analysis helped identify common variants in ASD
[199]
AGRE
Blood
2837 subjects
Association study
Rare haplotype affecting promoter of DLX1 found to be associated with ASD. No common variants found for DLX genes and GAD1
[200]
AGRE
Blood
2261 subjects
GWAS
The chromosome regions Xp22.33/Yp11.31 suggested to harbour male specific variants for ASD
[201]
AGRE
Blood
1132 subjects
QTL analysis
Chromosome regions 16p12-13 and 8q23-24 linked to harbour genes contributing to deficits in non-verbal communication in autistic patients
[202]
AGRE
Blood
993 subjects
Association study
Glu27 allele of ADRB2 gene suggested to confer increased risk of autism, with pregnancy related stressors having an increased effect
[203]
AGRE
Blood and brain tissue
90 subjects
Gene identification
Identification of the gene CORTBP2 from autism candidate region 7q31
[54]
AGRE
Blood
611 families
Association study
Reinforced evidence that GABRA4 and GABRB1 are implicated in ASD. Other ethnic groups found to have SNPs in these genes
[204]
AGRE
Blood
228 families (total)
38 (AGRE)
Association study
HOXG1 polymorphism A218G found to be associated in increased head circumference amongst ASD patients
[205]
AGRE
Genomic data
2165 subjects + 1165 families (total)
2165 subjects (AGRE)
GWAS
Associations found in the following genes with ASD and linked co-morbidities; KCND2, NOS2A and NELL1
[206]
AGRE
DNA
37 twin sets (total)
15 twin sets (AGRE)
Association study
Terbutaline exposure for two or more weeks associated with increased concordance for ASD. 2 polymorphisms for ADRB2 associated with ASD
[207]
AGRE
Blood
284 families (total)
38 families (AGRE)
Linkage/association study
Variants of PON1 found to be associated with ASD families in North America, but not in Italian families
[208]
AGRE
Blood
38 subjects
CNV study
Microdeletions and duplications on chromosome regions 3p26.3, 6q24, 22q11.2, 4q34.2 and 1q24 linked to ASD with physical anomalies. Genes STXBP5 and LRRN1 identified as candidate genes
[209]
AGRE
SSC
Genomic data
2294 subjects (SSC)
579 subjects (AGRE)
35663
CNV analysis
Exploration of evolution of human specific SRGAP2 genes. Rare duplications observed in SSC cohort for SRGAP2C.
[210]
AGRE
Genomic data from [211]
121 families
QTL-analysis
2 loci were identified in chromosomes 11 and 17 associated with social responsiveness in ASD families
[81]
AGRE
Blood
411 families (total)
371 families (AGRE)
Method paper
Detection of amplicons using mismatch repair. More amplicon variants were found in patients compared to controls
[212]
AGRE
Blood
66 subjects
Metabolite analysis
ASD families have lower levels of unprocessed Reelin protein in blood than controls
[213]
AGRE
Blood
90 subjects
Gene characterisation
CADPS and CADPS2 characterised and cloned. Found to be activators of protein secretion. No disease specific variants found amongst ASD patients
[214]
AGRE
Genomic data
1146 subjects
Linkage analysis
Linkage peaks found for language—speech phenotypes consistent with potential motor speed disorder in following chromosome regions; 1q24.2, 3q25.31, 4q22.3, 5p12, 5q33.1, 17p12, 17q11.2, 17q22, 4p15.2 and 21q22.2. multiple candidate genes were also identified
[215]
AGRE
Blood
2140 subjects
Linkage analysis
Parental origin effect significantly linked to chromosomes 4, 15 and 20
[42]
AGRE
Blood
167 families
Association study
EN2 found to be associated with ASD susceptibly
[216]
AGRE
Blood and lymphoblasts
537 subjects (total)
34 subjects (AGRE)
CNV analysis
Proposal that increased CNV load, particularly duplication of base pairs, predisposes to ASD. Negative correlation found with CNV load and social and communication skills. Applied to both common and rare CNVs
[73]
AGRE
Blood and lymphoblasts
4714 subjects (total)
1336 subjects (AGRE)
CNV analysis
Genes involved in Neuronal adhesion (NLGN1, ASTN2) and ubiquitin pathways (UBE3, PARK2, RFWD2, FBXO40) were found in ASD patients. Further evidence of NRXN1 and CNTN4 involved with ASD.
[217]
AGRE
Blood
147 subjects
Genotype phenotype
Suggested relationship between polymorphism MTFR 677C → T and autism-related behaviours
[218]
AGRE
Blood and lymphoblasts
693 subjects (AGRE)
5878 subjects (total)
CNV analysis
Microduplications and microdeletions in chromosome 16p11.2 associated with psychiatric disorders; duplications associated with schizophrenia, bipolar disorder and ASD, and deletions with ASD and other neurodevelopmental disorders
[219]
AGRE
Blood
219 subjects
Variant analysis
DLX1/2 and DLX5/6 gene analysis may not contribute to ASD but functional analysis of variants still worth investigation
[36]
AGRE
Blood
1410 (total)
401 (AGRE)
Association study
No association found for a sequence variant in mental retardation found in exon 1 of MECP gene in autism cohort
[220]
AGRE
Blood and lymphoblasts
112 families(total)
79 families (AGRE)
Association study
A haplotype for DRD1 is found to be associated with ASD risk amongst males
[221]
AGRE
Data from [222]
551 subjects (AGRE)
SNP analysis
Analysis of SNPs revealed variants of CD38 associated with ASD. Variants of CD38 linked to control of OXT secretion.
[223]
AGRE
Lymphoblastoid cell
14 subjects
Gene expression analysis
First study to show differential expression between lymphoblastoid cell lines. Genes affected implicated in cell death and development, nervous system development and immune development and function
[224]
AGRE
Lymphoblasts
116 subjects
Gene expression analysis
Patients with severe ASD showed altered expression of genes involved in Circadian rhythm. 20 novel genes found putative non-coding regions associated with androgen sensitivity
[225]
AGRE
Genomic data
1295 families (total)
696 families (AGRE)
GWAS
Noise reduction filter for GWAS leads to list of 830 candidate genes, where they impact dendrite and axon outgrowth and guidance
[29]
AGRE ATP
Blood and brain
133 sib pairs (total) 77 Sib pairs (AGRE) 17 brain tissue (ATP)
Ogliogenic hypothesis study
Evidence of epigenetic and genetic factors possibly contributing to ASD and UBE3 having a possible role in ASD
[179]
AGRE
Blood and lymphoblasts
192 subjects (AGRE)
483 subjects (total)
Association study
Disruptions in NRXN1 gene found to be associated with ASD
[226]
AGRE
Genomic data
476 subjects (total)
290 subjects (AGRE)
Association study
Suggestive association of parent and maternal origin effect on SLC6A4 promoter variant and ASD. Further testing required on biological model or larger cohort
[26]
AGRE
Blood and lymphoblasts
1549 subjects
410 subjects (AGRE)
Mutation screening
Recurrent microdeletions in chromosome region 16p11.2 were observed in ASD patients and not in controls
[227]
AGRE
Blood and lymphoblasts
974 subjects (total)
512 subjects (AGRE)
Mutation screening
RIMS3 identified as a possible ASD susceptibility gene
[228]
AGRE
Blood
33 families (AGRE)
49 families (total)
Association study
Association found for HLA-DR4 gene in higher frequency in geographically defined subtype, but not in controls or AGRE sample
[229]
AGRE
Blood
508 families (total)
139 families (AGRE)
Association study
Analysis of 2p15-16.1 microdeletions region identified two candidate genes; XPO1 and OXT1
[230]
AGRE
Blood and lymphoblasts
407 families (total)
138 families
Association analysis
Polymorphisms found in or near DLX1 and DLX2 found to be associated with ASD
[231]
AGRE
Blood and lymphoblasts
512 families (total) 138 families (AGRE)
Association study
Association found between ASD and MTHFR gene in simplex families but not in multiplex families
[37]
AGRE
Blood and lymphoblasts
219 families (total)
98 families (AGRE)
Association study
Polymorphisms in MECP2 found to be associated with ASD
[232]
AGRE
Genomic data
990 families
Association study
2 genes found to be associated with ASD; RYR2 and UPP2
[233]
AGRE
Genomic data
2194 families (total)
543 families (AGRE)
Association study
Association found between the calcium channel genes (CACNA1L, CACNA1C and CACNA1) with ASD
[32]
AGRE
Blood
470 families (total)
224 families (AGRE)
Gene association studies
GABRA4 and GABRB1 found to be associated with ASD
[234]
AGRE
Genomic data
680 families (AGRE) 1167 families (total)
GWAS
Identification of a common novel risk locus as chromosome region 5p14.1. Common and rare variants identified. AGRE used as validation dataset
[75]
AGRE
Lymphoblasts
12 subjects
Cell necrosis
Cells from autistic patients were more susceptible to oxidative and nitrosative stress
[235]
AGRE
Blood and lymphoblasts
1142 subjects (total)
139 families (AGRE)
Association study
GTF2i found to be associated with ASD
[236]
AGRE
Blood and lymphoblasts
207 families
CNV analysis
Translocation between short arms of chromosome 16 and 15 reported in 1 female patient. Nominal association of A2BP1/FOX1 observed in ASD cohort.
[237]
AGRE
Serum
21 human subjects
13 rhesus monkeys
Exposure study
Monkeys exposed to antibodies from human mothers of autistic children displayed stereotypies and hyperactive behaviour. Autoimmune component suggested to contribute to ASD
[238]
AGRE
ATP
Blood, lymphoblasts and brain tissue
276 families (AGRE)
17 subjects (ATP)
Association study
MARK1 gene found to be associated with ASD. Overexpression of gene also found in prefrontal cortex (BA46) but not cerebellum in human post-mortem tissue. Mouse model showed abnormalities in dendrites.
[55]
AGRE
Blood
123 families (total)
75 Families (AGRE)
Linkage disequilibrium study
Nominal evidence found for ASD risk alleles in GABAa Receptor subunits
[52]
AGRE
Blood
137 families (total)
80 families (AGRE)
Linkage and association study
SLC6A4 found not to be associated to rigid-compulsive subset of ASD patients.
[239]
AGRE
Blood
158 families (total)
Linkage analysis
Increased support that chromosome regions 19p13 and 17q11.2 harbour ASD susceptibility loci
[240]
AGRE
Blood and lymphoblasts
1336 subjects (AGRE)
1509 subjects (total)
CNV analysis
Large-scale survey of 15q24 microdeletion syndrome identifies atypical deletion that narrows critical region and (776 kb versus 1.75mb) and number of genes (15 versus 38) sequencing of genes recommended
[241]
AGRE
Genomic data
4278 subjects (total)
1518 subjects (AGRE)
Transmission disequilibrium testing
AGRE dataset found to have a genome wide signals at chromosome region 10q26.13 in both sexes and paternal signals in 6p21.1
[30]
AGRE
Blood and lymphoblasts
2886 subjects (total)
1441 subjects (AGRE)
CNV analysis
Microdeletions and duplications at chromosome region 15q13.2q13.3 found to be associated with ASD symptoms and other psychiatric disorders
[242]
AGRE
Blood and Lymphoblasts
34 subjects
Linkage analysis study
Chromosomes 7q and 21q are associated with a subset of ASD patients with developmental regression
[222]
AGRE
Blood and brain tissue
1221 subjects (total)
263 subjects (AGRE)
Association study
Two genetic variants of CD38 found to be associated with ASD
[243]
 
Blood
233 subjects
Association study
HOXA1 A218G alleles found to significantly influence head growth rates.
[244]
AGRE
Blood
196 families
Association study
Association not found between SNPs in DLX6 and PLCO on chromosome 7q21-22 and ASD
[245]
AGRE
Blood
196 families
Association study
Presence of a susceptibility mutation found in TDO2 or nearby gene
[246]
AGRE
Blood and lymphoblasts
249 families
Association study
Elevated levels of STX1A found to be associated with ASD
[47]
AGRE and ATP
Lymphoblasts
14 subjects (AGRE)
84 subjects (ATP)
Methylation study
Different methylation patterns found for genes involved in cell death/survival, neurodevelopment and gene transcription. Decreased expression of RORA and BCL2 was found in brain samples of ASD patients
[247]
AGRE
Blood and lymphoblasts
110 subjects
Genetic association study
Association found between PER1 and NPAS2 and ASD
[248]
AGRE
Blood and lymphoblasts
104 families
Genetic association study
BDNF associated with ASD; significantly higher expression in ASD subjects
[249]
AGRE
Blood and lymphoblasts
13,205 subjects (total)
80 subjects (AGRE)
CNV analysis
Disruption of the PTCHD1 locus on Xp22.11 identified in families with ASD and in families with Intellectual disability. Novel CNVs identified in DPYD and DPP6.
[80]
AGRE and ATP
Lymphoblasts and brain tissue
13 subjects (AGRE)
3 subjects (ATP)
Genotype-phenotype study
Increased dosage of the gene CYFIP1 results in altered cellular and dendritic morphology and dysregulates mTOR pathway in ASD patients with duplications in 15q11-13
[250]
AGRE
Blood and lymphoblasts
95 subjects (AGRE)
134 subjects (total)
Genomic and molecular study
No coding mutations or parental-specific expression found in ASD and Gilles de la tourettes syndrome (GTS) in the gene IMMP2L. Gene should not be written out as factor for both conditions
[251]
AGRE
Blood and lymphoblasts
283 families
Linkage mapping study
PRKCB1 shown to be associated with ASD
[252]
AGRE
Blood and lymphoblasts
1086 subjects
Candidate gene study
PITX1 shown to be associated with ASD
[253]
AGRE
Blood
406 families (total)
99 Families (AGRE)
Association and linkage disequilibrium study
GAD1 SNPs found not to be associated with ASD
[254]
AGRE
Blood
322 families (total)
86 families (AGRE)
Association study
No association found with APOE gene and ASD.
[255]
AGRE
Genomic data
4530 subjects
Association study
Immune function genes CD99L2, JARID2 and TPO show association with ASD
[256]
AGRE
Blood and lymphoblasts
334 families
Association study
Analysis of 2q24-q33 region found following genes associated with ASD; SLC25A12, STK39 and ITGA4
[257]
AGRE
Blood and lymphoblasts
411 families (total)
371 families (AGRE)
Linkage analysis
Linkage analysis of SNPs suggests SLC25A12 to be associated with ASD
[258]
AGRE
Blood and lymphoblasts
352 families
Association study
No association found between polymorphisms in TPH1 and TPH2 and ASD susceptibility or endophenotypes
[259]
AGRE
Blood and lymphoblasts
352 families (total)
295 families (AGRE)
Association study
No association found between SLC6A4 variants and susceptibility to ASD
[260]
AGRE
Blood and lymphoblasts
1011 subjects
Association study
AHI1, a gene associated with Joubert Syndrome, is also implicated in ASD
[261]
AGRE
Genomic data
2883 individuals
Methods paper
Tool that provides visualisation of SNP data
[262]
AGRE
Serum
34 subjects
Metabolite study
ASD patients had lower levels of the enzyme AAT in serum compared to controls. Difference is much more significant in ASD patients with regressive onset
[263]
AGRE
Blood and lymphoblasts
486 subjects (total)
252 subjects (AGRE)
Genotype-phenotype study
Mice with CADPS2 knockout display autistic-like behaviour and cellular features. Analysis of human Cadps2 mRNA revealed aberrant splicing that resulted in some patients lacking exon 3 of the transcribed gene
[264]
AGRE
Blood and genomic data
860 subjects (total)
468 subjects (AGRE)
GWAS
Regions in 5q21.1 and 15q22.1-q22.2 found to have most significant association in combined data for Asperger. 8 regions overlap with ASD linkage areas, and 3 overlapped with a Finnish cohort
[79]
AGRE
Lymphoblasts
14 subjects
MicroRNA analysis
Dysregulation of MicroRNA expression contributes to gene expression in ASD. Gene targets ID3 and PLK2 were validated by knockdown and overexpression assays
[265]
AGRE
Genomic data
289 families
Method paper
SNPs involved in three-way epistatic interactions found and all located in gene GLRX3
[58]
AGRE
Blood and lymphoblasts
264 families
CNV analysis
De novo CNVs were found to be strongly associated with Autism
[266]
AGRE
Blood and lymphoblasts
248 subjects (total)
146 subjects (AGRE)
Association study
Results suggestive that a y-chromosome haplotype effect is associated with ASD
[267]
AGRE
Blood and lymphoblasts
196 families
Transmission analysis
Polymorphisms in INPP1, PIK3G and TSC2 found to have linkage disequilibrium in ASD subjects
[268]
AGRE
Blood and lymphoblasts
196 families
Transmission analysis
Suggestive evidence that GRM8 is a susceptibility gene in ASD
[269]
AGRE
Blood and lymphoblasts
196 families
Association study
Suggestive but tentative evidence for MTF1 and SLC11A3 as ASD suspectability genes
[270]
AGRE
Blood and lymphoblasts
10 subjects
Whole genome sequencing
59 candidate genes suggested to be associated with ASD susceptibility, with ANK3 being the top result. 33 non-coding variants were also identified.
[271]
AGRE
Genomic data [73]
1336 subjects
Method paper
CNV analysis method that uses both B-allele frequency and log R ratio to find CNVs. Found all 21 validated short duplications in AGRE dataset. Analysis is much faster.
[272]
AGRE
Blood and lymphoblasts
Data taken from Ramoz, 2004
Association study
Suggestive association found for ASD-related routines and rituals with a polymorphism in SLC25A12
[273]
AGRE
Blood and lymphoblasts
144 subjects
Sequencing study
7 rare variants found in NLGN3 and NLGN4X. UTR found not to be significant. 2 intronic variants suggested to influence regulation of genes. Limited by throughput and cost
[274]
AGRE
Blood
351 families
Association study
Nominal significance found for 15 genes, top 3 being MYO1D, ACCN1 and LASP1 suggested for further study
[275]
AGRE
Genomic data
148 families
Linkage analysis
Male-specific linkage mapped to chromosome 17q11. Evidence of sex specific risk alleles in ASD
[56]
AGRE
Lymphoblasts
284 subjects
Association study
CACNAG identified as a candidate gene for ASD
[276]
AGRE
Lymphoblasts
267 subjects (AGRE)
Linkage and association study
SLC6A4 shown to contribute to ASD susceptibility
[78]
AGRE
Lymphoblasts
12 subjects
MicroRNA study
Lymphoblastoid cell lines from ASD patients can be used to assess microRNAs in ASD. Dysregulated MicroRNAs found to target genes linked to ASD
[277]
AGRE
Blood samples
100 subjects
Cholesterol metabolism
20% of the samples have shown hypercholestolemia, indicating that cholesterol metabolism could be perturbed in ASD
[278]
AGRE
Genomic DNA
756 subjects
Association study
EGF found to have significant association with ASD
[279]
AGRE and ATP
Data mining (AGRE) brain tissue (ATP) and blood
83 subjects
Linkage study
3p26.1, 3p26.3, 3q25-27 and 5p15 enriched for differentially expressed genes in blood and brain tissue. CNTN4, CADPS2, SUMF1, SLC9A9 and NTRK3 implicated in ASD and even more genes involved in neurological disorders that are co-morbid with ASD
[280]
AGRE
Blood
97 families
Expression profile analysis
RAY1/ST7 locus found to contain a multi-transcript system. Screening of ASD patients found rare variants not present in controls
[281]
AGRE
Blood and lymphoblasts
196 families (total)
95 families (AGRE)
Mutation screening
No mutations found in coding regions of X-chromosomal NLGN genes.
[282]
AGRE
Blood and lymphoblasts
136 families (total)
96 families (AGRE)
Association study
High association of FMR1 gene variant found amongst east Asian individuals, but not when whole sample was analysed, stratification confounded result
[283]
AGRE
Lymphoblasts
11 subjects
Neurotoxicity
Both ASD patients and controls showed upregulation of heat shock proteins when expressed to thimerosal
[33]
AGRE
Blood and lymphoblasts
3101 subjects (AGRE)
10796 subjects (total)
GWAS
Genome-wide SNPs found in CDH10 and CDH9 found to be associated with ASD
[284]
AGRE
ATP
Blood, lymphoblasts and brain tissue
1031 families (AGRE)
3104 families (total)
30 subjects (ATP)
GWAS
Analysis found association in chromosome region 5p15, where genes SEMA5A and TASR2 are located. Analysis of brain tissue showed reduced expression of SEMA5A in ASD subjects
[27]
AGRE
Lymphoblasts
5675 subjects (AGRE)
Association study
Micro deletion found in chromosome 16p11.2. amongst AGRE, Boston Children’s Hospital and Icelandic population data sets
[57]
AGRE
Blood
229 families
Association study
Sodium channel genes SCN1A1-3 contained SNPs of interest amongst ASD families for future studies
[285]
AGRE
Blood
564 families (total)
327 families (AGRE)
genetic analysis only
261 subjects (serotonin analysis)
Association study
ITGB3 genetic variation found to be associated with serotonin blood levels and ASD susceptibility
[286]
AGRE
Genomic data
5328 subjects
Recurrence rate study
Significant difference in recurrence rates between male only families and female carriers in regard to ASD. Female protective effect suggested to be at work in high genetic-risk families involving female carriers. Shorter interbirth intervals correlated to ASD risk.
[287]
AGRE
Blood lymphoblasts
1587 subjects
Linkage analysis
Replication of linkage on 20p13. Linkage found for chromosomes 6q27, 8q13.2, 1p31.3, 8p21.2 and v8p12
[288]
AGRE
Lymphoblasts
75 subjects (total)
50 subjects (AGRE)
Gene characterisation
Gene characterised and assessed for mutation amongst ASD patients. No concrete association found
[289]
AGRE
Genomic data
487 families
Method paper
Pathways of interest analysed using GWAS SNP data. 5 pathways shown to be of significance in regards to ASD
[290]
AGRE
Blood and lymphoblasts
383 subjects
Loci analysis
AGRE and Finnish ASD dataset both showed strong association with 3p24-26 locus containing the gene OXTR
[211]
AGRE
Blood and lymphoblasts
833 families
Genome-wide screen
Evidence of linkage to ASD found on chromosomes 17, 5, 11, 4 and 8, of which 17 having the highest association score in the group
[291]
AGRE
Blood and lymphoblasts
110 families
Genome-wide linkage analysis
Nominal evidence for linkage found in chromosomes 2–4,8, 10–12,15-16,18 and 20. significant linkage found for chromosomes 5 and 8 after reanalysis
[292]
AGRE
Blood and lymphoblasts
389 families
Association study
No evidence found that RH -ABO foetal-maternal incompatibility is associated with ASD
[46]
AGRE
Blood
126 families (total) 81 families (AGRE)
Association study
RELN alleles with large CGG repeats may play a role in aetiology of certain ASD cases
[293]
AGRE
Blood and lymphoblasts
165 subjects
Population genetics
Study suggested two groups: low risk families caused by spontaneous mutations, and high risk caused by female offspring that carry ASD-causing mutation that is passed onto their own offspring
[294]
AGRE
Blood and lymphoblasts
205 families
Gene association study
No association found between ASD and variant of the gene EN2
[295]
AGRE
Lymphoblasts
20 subjects
Intracellular redox study
Inbalance of glutathione redox in cell lines derived from patients with ASD
[76]
AGRE
Lymphoblasts
86 subjects
Transmethylation/transsulfuration study
Cell lines derived from parents of ASD children showed abnormal transmethylation/transsulfuration metabolism and DNA hypomethylation
Study numbers listed as families or subjects wherever applicable
The AGRE resource has been used extensively in genomics studies in ASD. Approaches have included gene-mapping such as genome-wide linkage and association studies in addition to studies of chromosomal structure, particularly the identification of copy number variants. Important ASD chromosomal regions identified include microdeletions and microduplications of 16p11.2 [26, 27], rearrangements and microdeletion/duplication of 15q13.2q13.3 [2831], common variants in the 5p14.1 region [32, 33], Neurexins and 11p12–p13 [34].
It has also helped in identification of recurrent candidate genes, such as MECP2 [3537], PTEN [38, 39], EN2 [4042], RELN [40, 4346], RORA [47], MET [4850], NGLN3-4 [51], BZRAP1 [28], SLC6A4 [40, 52] GABA receptors [32, 43, 5355], CACNA1G [56] and the sodium channel genes SCN1A, SCN2A and SCN3A [57].
These studies particularly highlighted an important role of de novo and large inherited copy number variations (CNVs), which are detected in 10% of sporadic ASD [58], which has been widely replicated in other bio-collections [5971]. The use of AGRE combined with other AGP resources have uncovered SHANK2, SYNGAP1, DLGAP2 and the X-linked DDX53-PTCHD1 locus as novel ASD genes, as well as pathways of cellular proliferation, signalling, neuronal projection and motility [72]. AGRE samples formed a replication set in a separate analysis highlighting CNVs of neuronal cell adhesion and ubiquitin pathway in ASD [73].
AGRE lymphoblastoid cells enabled studies into shared ubiquitin and neuronal gene expression in lymphoblastoid cells and brain [73, 74], gluthathione metabolism, oxidative stress [75, 76] and stress response [77], microRNAs and their use in ASD profiling [78, 79], CYFIP1 dosage effect on mTOR regulation [80], and changes in methylation patterns of RORA and BLC2 and their effects on apoptosis, cellular differentiation, inflammation and neural development [47].
The AGRE collection was also used to establish genetic methodologies and bioinformatic tools. This included using mismatch repair to detect amplicons in ASD [81], using multiplex ligation-dependent probe amplification (MPLA) to improve detection of microduplications and microdeletions [82], and incorporating disease symptoms to improve linkage detection in genetic data [83] and analysis of genetic loci to search for candidate genes [84].

Simons Simplex Collection (SSC)

The SSC is a genetic and clinical repository, which contains material derived from 2600 families. Whereas the AGRE contains multiplex families and trios, The SSC ascertained “simplex” ASD families defined as families where only one child has ASD and at least one other typically developing sibling. DNA is available for both parents, the affected child and an unaffected sibling. Thus the SSC samples are particularly valuable in evaluating parental inheritance. Samples were collected at multiple sites and were stored as immortalised cell lines at Rutgers University Cell and DNA Repository (RUCDR). Each sample was verified for parentage, gender and Fragile X mutation. In-depth clinical phenotypes were characterised for all participants to support genotype-phenotype analyses. These included data on diagnostic status, medical and psychiatric comorbidity, family history and medication use for the affected person. Broader ASD phenotype measures were collected for unaffected family members.
The SSC has become a vast resource of ASD and contributed significantly to numerous Whole exome sequencing studies of ASD in the past ~7 years (Table 3). The main findings showed that de novo mutations were frequently enriched in ASD patients [60]. Whole-genome sequencing results showed a significant enrichment of de novo and private disruptive mutations in putative regulatory regions of previously identified ASD risk genes. It also identified novel risk factors of CANX, SAE1 and PIK3CA with small CNVs and exon-specific SNPs, which were overlooked in previous CNV studies or exon sequencing [85]. It has also been observed that many de novo mutations were of paternal origin (4:1) and positively correlated with paternal age, [65]. The disruptive mutations were located in genes involve in transcription regulation, chromatin remodelling and synapse formation [86, 87].
Table 3
Overview of studies using the SSC collection
Reference
Bio-collection
Samples
Number
Study
Findings
[296]
SSC
Genomic data
2760 subjects
CNV analysis
No association found between conception-assisted reproduction and risk of ASD
[297]
SSC
Lymphoblasts
900 subjects
Sequencing study
Rare functional variants of TSC1/TSC2 did not show association with ASD
[298]
SSC
Genomic data
965 subjects (SSC)
Integrative analysis
Integrative analysis of data from 4 exome sequencing studies revealed enrichment of genes involved in chromatin remodelling and transcription in ASD patients
[88]
SSC
Blood
3730 subjects
Genotype-phenotype
Subtype of autism was caused by mutations to CHD8, of which 15 were found.
[299]
SSC
Blood
259 subjects
CNV analysis
Paired duplications mark cryptic inversions and other complex structural variations in CNV data.
[300]
SSC
Blood
552 subjects (total)
412 subjects (SSC)
Transcriptome analysis
Neuron development, nitric oxide signalling, neurogenesis and skeletal development were found outliers amongst ASD patients in TGEN cohort, whereas outliers were found in neurogenesis in ASD patients from SSC cohort
[301]
SSC
Blood and lymphoblasts
99 families
CNV analysis
55 potential pathogenic CNVs were identified and validated. 20% were considered rare when compared to the database of genomic variants. CNVs found in lymphoblast DNA but not in blood, suggesting pre-existing mutations may have been present in initial lymphoblast cells
[302]
SSC
AGRE
Blood urine
12600 subjects (total)
1887 subjects (SSC)
752 subjects (AGRE)
Association study
TMLHE found to have high levels of deletion in male-male multiplex families (1 in 190) and deficiency of this gene could be a susceptibility factor for ASD.
[303]
SSC
Genomic and exomic data
Taken from earlier studies [60, 61, 70]
Genotype-phenotype study
Mutations in ASD candidate genes have greatest impact on pyramidal neurons, cortical neurons and medium spiny neurons. Truncating de novo mutations play a small role in high-functioning cases. The greater the functional disruption of genes, the more severe the phenotypes are.
[304]
SSC
Blood
2575 subjects
GWAS
Reducing phenotypic heterogeneity within the cohort did not have a significant effect on increasing genetic homogeneity.
[305]
SSC
AGRE
Blood
14989 subjects (total)
5981 subjects (AGRE)
1815 subjects (SSC)
GWAS
CNVs found in SEMA5 regulated gene network found to be associated with ASD
[86]
SSC
Blood
13,804 subjects
WES
104 genes were implicated in 5% of ASD cases, where they are involved in transcription, chromatin remodelling and synapse formation.
[59]
SSC
Blood
2963 subjects
WES
De novo INDELS primarily originate from father, frameshift INDELS associated with ASD, Frameshift INDELS more frequent in females. RIMS1 and KMT2E found to be associated with ASD
[306]
SSC
Blood
8 subjects
Methods Paper
WGS data more effective than WES for detection of INDELS. x60 sequencing required to recover 95% of detected Indels
[307]
SSC
Genomic data
2066 subjects
Homozygosity study
In ASD simplex families, increased runs of homozygosity is associated with Intellectual disability
[308]
SSC
Blood
1227 subjects (total)
350 subjects (SSC)
CNV analysis
CNV burden correlates to certain disorders; high CNV burden to Intellectual disability and low CNV burden to dyslexia
[178]
SSC
AGRE
Blood
3168 subjects (total)
2478 subjects (SSC)
719 subjects (AGRE)
Rearrangement hotspot study
1q21 duplications found to be associated with Autism. CNVs identified in CHD1L, ACACA, DPP10, PLCB1, TRPM1, NRXN1, FHIT and HYDIN enriched in ASD. Duplications linked to decreased non-verbal IQ and duplications linked to severity of ASD.
[149]
SSC
IPSCs and lymphoblasts
1041 subjects
Disease modelling study
Disruption of TRPC6 causes disruption in human neurons and linked to a non-syndromic form of ASD. First Study to use Patent-derived IPSCs to model non-syndromic form of ASD
[309]
SSC
AGRE
Blood
2975 subjects (total)
1429 subjects (SSC)
14 subjects (AGRE
GWAS sequencing
Rare variants in synaptic genes associated with ASD. Loss of function in candidate genes a major risk factor for ASD.
[310]
SSC
TASC
Blood
932 families (total)
Method paper
Transmission and de novo association(TADA) is a method that incorporates WES data, as well as inherited variants, and variants identified between cases and controls
[311]
SSC
Exome data
597 subjects
Method description
Association was found between ASD and rare variants of the gene ABCA7 in exome data
[312]
SSC
Blood
15479 subjects (total)
9479 subjects (SSC)
Transmission analysis
Demonstration that high and low IQs could be distinguished by LGD load in respective gene targets. Transmission of rare variants with low LGD load occurs more often to affected offspring. Biased transmission towards children with low IQ
[61]
SSC
Blood
1478 subjects
WES
Gene disrupting mutations were twice as frequent in ASD subjects compared to controls. Genes disrupted were associated with Fragile X Protein FMRP.
[121]
SSC
Blood
762 subjects
CNV study
Female subjects showed a higher mutational burden before developing ASD.
[313]
SSC
Blood
720 subjects
Association study
Association was found between gene SLC25A12 and restricted and repetitive behaviour.
[314]
SSC
Blood
2106 families (TOTAL)
965 families(SSC)
Common variation study
Multiple common variants of genes additively contribute to ASD risk. Simplex families found to closely follow additive model compared to multiplex families
[315]
SSC
Blood
285 subjects
Transcriptomic study
Enriched genes found in long term potentiation/depression, Notch signalling and neurogenesis amongst ASD Patients. 55 gene prediction model performed well on male subjects, but not female subjects
[316]
SSC
Blood
58 subjects
Transcriptomic study
Upregulation of spliceosome, mitochondrial and ribosomal pathways and downregulation of neuroreceptor-ligand, immune response and calcium signalling pathways in ASD patients compared to controls
[317]
SSC
Genomic data
78349 subjects (total)
3080 subjects (SSC)
SNP study
17–29% of variance in liability explained by SNPS. Genetic correlation found between disorders;
High: Schizophrenia and bipolar disorder
Moderate: Schizophrenia and major depressive disorder, major depressive disorder and ADHD, major depressive disorder and bipolar disorder
Low: Schizophrenia and ASD
[60]
SSC
Genomic Data
1784 subjects
CNV study
De novo duplications and deletions are major contributors to ASD. Females shown to have a greater genetic resistance to autism.
[318]
SSC
AGRE
TASC
Blood
6970 subjects (total)
806 subjects (AGRE)
996 subjects (TASC)
563 subjects (AGRE)
WES
2-fold enrichment of complete knockout of autosomal genes with low LoF variation, and 1.5-fold enrichment for rare hemizygous knockout in males. Both contribute 3 and 2% to ASD risk, respectively.
[63]
SSC
Lymphoblasts
386 subjects
CNV study
Recurrent and rare de novo CNVs were discovered to alter gene expression in chromosome regions 3q27, 3p13, 3p26, 2p15, 16p11.2 and 7q11.23.
[129]
SSC
IPSCs
12 subjects
Disease modelling
Overexpression of FOXG1 was linked to increased head circumference and ASD severity in idiopathic autism subjects. An overabundance of inhibitory neurons in ASD cell lines was also found.
[319]
SSC
Genomic and clinical data
2478 subjects
Gene-environment study
Individuals with ASD-associated CNVs were more susceptible to effects of febrile episodes and maternal infection during pregnancy and have impact on behavioural outcomes
[320]
SSC
Blood
10118 (TOTAL)
1974 (SSC)
Genetic association
Higher prevalence of SLC12A5 variants containing altered CpG sites amongst ASD patients.
[321]
SSC
DNSB
Blood
2418 subjects (SSC)
1353 subjects (DSNB)
CNV analysis
17q12 deletion identified as a CNV variant that confers high risk of ASD and Schizophrenia
[322]
SSC
AGRE
Genomic data
49167 subjects (total)
1124 subjects (SSC)
1835 subjects (AGRE)
CNV analysis
More significant CNVs that could infer ASD risk were identified using combined large clinical datasets of neurodevelopmental disorders than with ASD cohorts alone
[323]
SSC
Lymphoblasts
5451 subjects
Association study
No association was found for heterozygous mutations in CNTNAP2 and contribution to ASD risk
[324]
SSC
Blood and lymphoblastoid cell lines
593 families
Method description
A novel method was used to detect de novo and transmitted insert-deletions(Intel’s) in exomic data
[325]
SSC
Blood
1315 subjects (total)
145 subjects (SSC)
CNV analysis
Duplication CNVs enriched in negative regulation categories, deletion CNVs enriched in positive regulation categories. Highly connected genes in network enriched in patients with a single gene CNV change
[65]
SSC
Blood
677 subjects (SSC)
WES
De novo mutations paternal in origin (4:1) and positive correlation with age. Recurrent mutations in genes CDH8 and NTNG1.
[64]
SSC
Blood
20 families (total)
19 families(SSC)
WES
21 de novo mutations identified. 11 of which found to be protein altering. Mutations identified in FOXP1, GRIN2B, SCN1A, LAMC3 and CNTNAP2.
[87]
SSC
Blood
2246 subjects (SSC)
WES
27 de novo events found in 16 genes, 59% predicted to truncate proteins. further support for genotype-phenotype relationship in CDH8 and DYRKA1
[326]
SSC
Blood
19 subjects (total)
4 subjects (SSC)
Genotype-phenotype
Overexpression/increased dosage of MECP2 related with core features of ASD
[133]
AGRE
SSC
Data taken from [327]
8816 subjects (total)
737 subjects (SSC)
4449 (AGRE)
Replication study
Findings could not be replicated from Skafidas paper
[328]
SSC
DNBS
Genomic data
38000 subjects (total)
4358 subjects (SSC)
19142 subjects (DNBS)
General population study
Genetic influences on ASD risk found to influence typical variation in social and communication ability in the general population
[67]
SSC
Blood
2256 subjects
De novo and familial influences
Familial influences were more significant in cases of high-functioning ASD conditions.
[327]
SSC
Lymphoblasts
1 subject
Clinical report
De novo microdeletion in chromosome 3q29 associated in person with ASD, childhood psychosis and intellectual disability
[68]
SSC
Genomic data
1024 families
De novo mutation analysis
Significant role for loss of function mutations in ASD cases.
[329]
AGRE
SSC
Blood
8816 subjects (total)
737 subjects (SSC)
4449 subjects (AGRE)
Predictive testing
Diagnostic classifier containing 237 SNPs and 146 genes
[330]
SSC
AGRE
Blood
975 subjects (total)
392 subjects (SSC)
585 subjects (AGRE)
Genotype-phenotype study
NPAS1 found to repress generation of specific subtypes of cortical interneurons
[85]
SSC
Blood
53 families
Whole genome sequencing
Enrichment of disruptive mutations in putative regulatory regions in ASD patients
[71]
SSC
Blood
9231 subjects
Genotype-phenotype study
Disrupting mutations in DYRK1A were linked to a subset of 15 patients with a syndromic form of ASD/ID.
[331]
SSC
Blood
903 families
WES
Enrichment of non-synonymous and potentially pathogenic mutations in mitochondrial DNA in ASD patients compared to controls. Transmission of potential pathogenic mutations differed between mother-ASD pairs and mother-sibling pairs
[332]
SSC
Lymphoblasts
1 family
Mutation analysis
PKA found to be an upstream regulator of UBE3A, where mutation in phosphorylation site results in hyperactivity of UBE3A
[333]
SSC
Blood
686 subjects +
612 families (SSC)
WES
Bi-allelic mutations found in genes enriched in inherited ASD cases (AMT, PEX7, SYNE1, VPS13B, PAH, POMGNT1)
[333]
SSC
Blood
928 subjects
WES
Strong evidence that de novo mutations are associated with ASD
[69]
SSC
Blood, lymphoblasts and saliva
1174 families
CNV analysis
Significant associations found between ASD and de novo duplications of chromosome 7q11.23. de novo CNVs identified in 5 other regions, including 16p13.2
[334]
SSC
Blood and lymphoblasts
2591 families
CNV analysis
De novo CNVs associated with ASD. 6 loci and 65 genes identified, many targeting the chromatin or synapse
[335]
SSC
Genomic data
2337 families
Transmission disequilibrium
Excess of truncating inherited mutations associated with ASD. RIMS1, CUL7, LTZR1 identified as candidate genes
[336]
SSC
Genomic data
411 families
Transmission disequilibrium
Affected ASD patients inherited more CNVs than their unaffected siblings, and these CNVs of ASD patients affected more genes. Enrichment of brain-specific genes in inherited CNVs amongst ASD patients
[312]
SSC
Genomic data
10,942 subjects (total)
4942 subjects (SSC)
Biased transmission study
Frequent biased transmission of disruptive mutations to Low IQ ASD patients. Low and high IQ subjects can be distinguished by mutational load.
Study numbers listed as families or subjects wherever applicable
The SSC has enabled detection of the ultra-rare “recurrent” CNVs. This included duplications of 7q11.23, 15q11.2 (NIPA) and 16p13.11, and deletion/duplication of 16p11.2, 16p13.2 (USP7), 1q21.1, 2p16.3, 7q31.1, 15q13.2–q13.3, 16p13.3, 20q13.33 and 22q11.21 [60]. The SSC also helped identify recurrent gene mutations in ASD include CHD8, NTNG1, GRIN2B, SCN1A and LAMC3, which are important for transcriptional regulation, neuronal differentiation and function [87].
CHD8 was further evaluated as an ASD candidate gene in children with developmental delay or ASD, and 15 independent mutations were identified and enriched in a subset of ASD with altered brain size, distinct facial features and gastrointestinal complaints. Disruption of CDH8 in zebra fish recapitulated some of the patient phenotypes including increased head size and impaired gastrointestinal motility [88]. CHD8 is shown to control expression of other high-confidence de novo ASD risk genes such as DYRK1A, GRIN2B and POGZ [89]. Mutation of DYRK1A was strongly linked to a subset of ASD patients with seizures at infancy, hypertonia, intellectual disability, microencephaly, dysmorphic facial features and impaired speech [71, 89]. POGZ gene which plays a role in cell cycle progression is also found to contribute to a subset of ASD with varying developmental delay, vision problems, motor coordination impairment, tendency of obesity, microcephaly, hyperactivity and feeding problems [90].

Danish Newborn Screening (NBS) Biobank

The NBS Biobank has a large collection of dried blood spot samples (DBSS), which are taken from new-borns 5–7 days after birth. They are sent to the New-born Screening lab at the Statens Serum Institute for analysis, and stored at −20 °C in a separate freezing facility at the NBS Biobank. Prior to collection, parents are informed via leaflets about the biobank, with focus on what the samples will be used for (documentation, testing and retesting, research, etc.). Participants can opt out of storage at any time via a letter to the department. For security, both the clinical data and biological samples are linked via a unique number, kept in separate buildings, and are accessible by authorised personnel only [91]. The advantage of the NBS resources is that it provides a large amount of non-ASD controls as well as Danish ASD samples.
In the past 30 years the NBS Bio-collection has accumulated samples from 2.2 million individuals, around 65,000–70,000 samples per year from Denmark, Greenland and the Faroe Islands. Most recently this resource has been included under the Danish iPsych consortium with the Psychiatric Genomics Consortium (PGC), added 8–12 k samples to the PGC analysis and significantly increased its power to detect common genetic effects for ASD, which have been recently published [92].
DBSS were also used to examine metabolites. A group led by Abdallah carried out a series of studies on Danish collections (Table 4) to examine the potential role of cytokines and chemokines involved in signalling and immune response of ASD. Initially using amniotic fluid from the Danish Birth Cohort (DBC) collection [93, 94], they followed up with DBSS from new-borns crossed referenced from that cohort [95, 96]; they detected an imbalance of cytokines amongst ASD subjects compared to the controls. Most of the chemicals were lower than normal, such as Th-1 and Th-2 like cytokines involved in proliferation, priming and activation of these cell types, whereas a small number of cytokines displayed increased expression in ASD. The abnormal levels of these chemicals could lead to a hypoactive or “inactive” immune system in the brain, making it more susceptible to infection-related ASD. However, when chemokine levels were examined in amniotic fluid, no concrete relationship could be established.

The Autism Simplex Collection (TASC)

TASC is a trio-based international bio-collection that was assembled in collaboration with the Autism Genome Project and funded by Autism Speaks [97]. Trios, comprised of both parents and a child affected with ASD with no known medical or genetic cause. Collection of samples took place between 2008 and 2010 across 13 sites; 9 in North America and 4 in Europe. Management, storage and distribution of TASC data are handled by the Centre for Collaborative Genetic Studies on Mental Disorders (CCGSMD) [97]. Samples are housed at the NIMH and AGRE repositories both of which are located at Rutgers University.
So far, TASC has been used for GWAS studies [66] and CNV studies [72, 98, 99] and WES Studies [16, 100, 101]. In addition, TASC has also been used in WGS as part of the MSSNG project, which is discussed below

Autism Inpatient Collection (AIC)

The AIC is a bio-collection for ASD research based on those on the serve end of the spectrum with severe language impairment, intellectual disability and self-injurious behaviour. This collection was founded on the basis that this segment of ASD patients are largely unrepresented in current studies. Bio-samples are initially recruited from 147 patients, and ongoing recruitment is estimated at 400 per year. Psychiatric, clinical and phenotypic data are collected in addition to blood samples for the creation of lymphoblastoid cell lines by RUCDR. Amongst this collection, over half are non-verbal, over 40% have intellectual disability and a quarter exhibit self-injurious behaviour [24]. This collection has yet to be used in any genetics-based studies. The fact that many patients are on the severe end of the spectrum makes it a welcome addition, and it opens opportunities to explore this under-represented group.

Autism Tissue Program (ATP)/Autism BrainNet

The Autism Tissue Program, now the Autism Brain Network, is a post-mortem ASD brain collection coordinated by a network of parents, caregivers, physicians and pathologists. Brain samples are preserved in formalin and/or in −80 °C freezers to maximise the potential studies. In some cases, both hemispheres are fixed in formalin when there is freezing capacity or if the post-mortem interval exceeds 24 h. Corresponding clinical data include age, sex, ethnicity, diagnosis, brain size, cause of death, post-mortem interval and preservation method for the left and right hemisphere of the brain. Due to the rarity of the sample, a thorough application procedure assesses scope, scale and feasibility of proposed projects prior to access of tissue, with the expectation that data, images and presentations generated by research on the samples are provided back to the Autism Brain Network 3 months after formal release of publications [102].
Brain pathology and molecular mechanisms have been the focus of studies using the ATP resource (Table 5) although many studies looking at brain anatomy and cell morphology employed samples from this collection, molecular and genetic studies are the primary focus of this review. Such studies included transcriptomics [103105], epigenetics [29, 106115] and alternative splicing [116, 117]. A key discovery was the identification of convergent molecular pathology linking to neuronal, glial and immune genes [105] in a transcriptomics study that investigated the gene co-expression network between autistic and control brains. This led to the proposal of abnormal cortical patterning as an underlying mechanism due to attenuated differential expression in frontal and temporal cortices in ASD brains.
A recent study showed reduced Vitamin B12 in ASD brains [118] where the ATP made a very large contribution. Post-mortem examination of brain tissue ranging from foetal to the elderly subjects also showed a marked decline of the brain vitamin B12 with age, together with lower activity of methionine synthase in the elderly, but the differences were more pronounced in ASD and schizophrenia subjects when compared to controls. Acetylation is an important post-translational modification in the field of epigenetics. ATP also made a significant contribution to a large-scale histone acetylome-wide association study (HAWAS) using the prefrontal cortex, cerebellum and temporal cortex in ASD patients and controls. Despite their heterogeneity, 68% of syndromic and idiopathic ASD cases shared a common acetylome signature at >5000 cis-regulatory elements in the prefrontal cortex and temporal cortex. Aberrant acetylome was found to be associated with synaptic transmission, ion transport, epilepsy, behavioural abnormality, chemokinesis, histone deacetylation and immunity [113].
The ATP sample was used in a methylation study that investigated differential methylation in CpG loci in three brain regions: temporal cortex, dorsolateral prefrontal cortex and cerebellum. Differential methylation of four genes (PRRT1, C11orf21/TSPAN32, ZFP57 and SDHAP3) was detected. PRRT1, C11orf21/TSPAN32 were hypomethylated while the latter two were hypermethylated [109]. A further investigation in Brodmann’s area also found a pattern of hypomethylation of a number of genes including C11orf21/TSPAN32 that are implicated in immune function and synaptic pruning [111]. These hypomethylated genes correlated with those showing overexpression by Voineagu.
The methylation studies have further uncovered dysregulation of OXTR and SHANK3 genes in ASD. OXTR gene encoding oxytoxcin receptor was significantly hypermethylated in the peripheral blood cells and temporal cortex of ASD, highlighting a reduced oxytocin signalling in the aetiology of ASD [108] and a therapeutic target of ASD. Differential methylation of the SHANK3 gene was detected between ASD and control brains. They found that when three 5′ CpG islands of the gene were examined, they observed altered methylation also changed SHANK3 splicing, with specific SHANK3 isoforms expressed in ASD [114].
This is echoed by a recent study, which reveals a dynamic microexon regulation associated with the remodelling of protein-interaction networks during neurogenesis. The neural microexons are frequently dysregulated in the brains of ASD, which is associated with reduced expression of SRRM4 [116]. The neuronal-specific splicing factor A2BP1/FOX1 and A2BP1-dependent splicing of alternative exons are also dysregulated in ASD brain [105].

Replication studies and pooling resources

Research data from one bio-collection is not always replicable in another sample set. Therefore, cross-validation between different bio-collections will not only minimise false positive, but also identify the common risk factors and subset-specific factors. For example, a genome-wide survey was carried out to test trans-generational effects of mother-child interactions, and the AGRE and SSC samples were used to replicate the original findings of 16 ASD risk genes (PCDH9, FOXP1, GABRB3, NRXN1, RELN, MACROD2, FHIT, RORA, CNTN4, CNTNAP2, FAM135B, LAMA1, NFIA, NLGN4X, RAPGEF4 and SDK1) involving urea transport and neural development. The results from the AGRE and SSC cohorts did not match the original study and showed fewer associations. When post-correction of the statistics was applied, the results lost their significance [119]. This could partially be due to the differences in the array design with different coverage of SNPs and/or different methodologies.
The meta-analysis of five data sets including the AGRE and SSC demonstrates that females have a greater tolerance to CNV burden. This leads to a speculation that the maternal tolerance of the CNVs can result in decreased foetal loss amongst females compared to males, and that ASD-specific CNV burden contributes to high sibling occurrence. What is interesting about this study is that the results for high CNV burden in females are consistent throughout each data set. This is an example showing how multiple bio-collections can give a clearer picture in a combined study where individual studies may be ambiguous [120, 121].
Many major studies on the genetics of ASD have also been accomplished as a result of the collaborations amongst the institutions (Tables 2, 3, 4, 5 and 6). An effort was made to evaluate the association of Fragile X Mental Retardation 2 locus (AFF2) with ASD using joint resources from AGRE (127 males) and SSC (75 males). AFF2 encodes an RNA-binding protein, which is silenced in Fragile X. The study found that 2.5% of ASD males carry highly conserved missense mutations on AFF2 gene which was significantly enriched in ASD patients, when compared to >5000 unaffected controls [122]. A WES was published recently, which sequenced the exomes of over 20,000 individuals, including those from the SSC and Swedish registries. The study identified 107 candidate genes, and reinforced ASD pathways of synaptic formation, chromatin remodelling and gene transcription. This study detected mutations in genes involved in calcium- (CACNA2D3, CACNA1D) and sodium-gated channels (SCN2A) which were related to neuronal function, and in genes involved in post-translational methylation (SUV420H1, KMT2C, ASH1L, SETD5, WHSC1) and demethylation (KDM4B, KDM3A, KDM5B, KDM6B) of lysine residues on histones which provided molecular basis linking to neuronal excitation and epigenetic changes in ASD [86].
Table 4
Overview of studies using the DNSB collection
Reference
Bio-collection
Samples
Number
Study
Findings
[136]
DNSB
DBSS
1100 subjects
Chemokine analysis
Analysis of crude estimates showed decreased levels of RANTES. Adjusted estimates showed no significance amongst 3 chemokines studied (RANTES, MCP-1, MIP-1A). Cautious suggestion of altered immunity in neonatal period amongst ASD patients
[96]
DNSB
DBSS
1200 subjects
Cytokine analysis
Suggestive evidence of decreased levels of certain th-1 and th-2 like cytokines in newborns later diagnosed with ASD.
[136]
DNSB
DBSS
1029 subjects
Neurotropic factor analysis
Decreased level of neurotropic factors found in ASD patients during Neonatal period
Study numbers listed as families or subjects wherever applicable
Table 5
Overview of studies using the ATP/Autism BrainNet collection
Reference
Bio-collection
Samples
Number
Study
Findings
[337]
ATP and AGRE
Brain tissue, blood and lymphoblasts
18 subjects (ATP)
841 families (AGRE)
1029 families (total)
Gene expression and association analysis
Altered expression of mitochondrial genes in anterior cingulate gyrus, motor cortex and thalamus of ASD patients. Polymorphisms in MTX2, NEFL and SLC25A27 found to be associated with ASD.
[338]
ATP
Brain tissue
18 subjects
Gene expression analysis
Reduced expression of several genes related to electron transport in anterior cingulate gyrus, motor cortex and thalamus of ASD patients
[339]
ATP
Brain tissue
57 subjects
Functional genomic study
Analysis of CNVs showed differences of what pathways are altered between children and adults; cell number, cortical patterning and differentiation in the former, and signalling and repair pathways in the latter. Prefrontal cortex samples were used
[106]
ATP
Brain tissue
33 subjects
GWAS
Patients with ASD had more genes that were up- or down-regulated in an- individual specific manner when prefrontal cortex tissue was examined
[340]
ATP
Brain tissue
126 subjects (total)
42 subjects (ATP)
Sequencing study
Recurrent deleterious mutations found in ARID1B, SCN1A, SCN2A and SETD2. Higher proportion of mutations that are deleterious, protein-altering or cause loss-of-function in ASD patients compared to controls. Cortical and cerebellar tissue was used.
[107]
ATP
Brain tissue
25 subjects
Deep sequencing study
Altered adenosine to inosine editing found in cerebella tissue from ASD patients. Dysfunctional for of editing enzyme ADARB1 more frequently in ASD Cerebella
[341]
ATP
Brain tissue
28 subjects (ATP)
43 subjects (total)
Gene expression analysis
Signalling partners of FMRP and GRM5 (HOMER1, APP, RAC1, STEP) shown to have altered expression in the cerebellar vermis and superior frontal cortex in ASD patients compared to controls.
[342]
ATP
Brain tissue
19 subjects
mRNA analysis
Reduction of multiple GABA receptor subtypes (A6, B2, D, E, G2, T and P2) detected in cerebella vermis and superior frontal cortex ASD patients
[343]
ATP
Brain tissue
25 subjects
Assay study
Imbalance in isoforms of precursor BDNF protein found in fusiform gryrus of ASD patients
[103]
ATP
Brain tissue
18 subjects
Transcriptional and epigenetic association analysis
Downregulation of genes related to oxidative phosphorylation and protein translation. Associations were found between specific behaviour domains of ASD and gene expression modules related to myelination, immune response and purinergic signalling. Cerebral and Brodmann area 19 tissue was used
[108]
ATP
Brain tissue
16 subjects
Methylation study
Increased methylation was found for the gene OTXR in ASD patients in blood and DNA from the temporal cortex
[104]
ATP
Brain tissue
107 subjects
Transcriptome analysis
Dysregulated gene expression associated with glial cells shown to have negative correlation with gene expression relating to synaptic transmission in ASD patients when Brodmann areas 10, 19 and 44 were analysed
[344]
ATP
Brain tissue
32 subjects
Transcription analysis
RORA may have dimorphic effects on gene expression in certain areas of cortical tissue between genders, and deficiency appears to cause greater gene dysregulation amongst males in both mice and humans
[345]
ATP
Brain tissue
30 subjects
Transcription analysis
RPP25 expression is decreased in the prefrontal cortex of ASD patients
[116]
ATP
Brain tissue
23 subjects
Alternate splicing analysis and discovery
A conserved group of microexons involved in modulation of interaction domains of proteins and neurogenesis is disrupted in patients with ASD
[29]
ATP
Brain tissue
17 subjects
Methylation study
UBE3 implicated as a contributing gene to autism and Angelman syndrome
[346]
ATP
Brain tissue
20 subjects
Anti-sense RNA study
Discovery of anti-sense non-coding RNA that binds to moesin at 5p14.1 in ASD cerebral cortex tissue
[109]
ATP
Brain tissue
40 subjects
Methylation study
4 differentially methylated regions; 3 in temporal cortex and 1 in cerebellum. 3/4 regions were again found in different samples and brain regions.
[117]
ATP
Brain tissue and lymphoblasts
36 subjects (total)
Transcription and alternative splicing study
Accelerated decrease of MS gene transcription across ageing found in ASD patient cerebral cortex samples
[347]
ATP
Brain tissue
73 subjects
Methylation study
Correlation found between reduced expression of MECP2 and increased methylation on the promoter region
[110]
ATP
Brain tissue
24 subjects
Methylation study
Hypomethylation of mir142 and upregulation of mi-RNAs targeting OXTR gene in prefrontal cortex of ASD brains
[348]
ATP
Brain tissue
24 subjects
Signal transduction study
Downregulation of PI3K-Akt genes observed in fusiform gyrus tissue of ASD patients. Similar effects noted in rat brain tissue exposed to valproic acid
[349]
ATP
Brain tissue (ATP)
neuronal cells
6 subjects
CHiP study
RORA found to regulate 2BP1, CYP19A1, HSD17B10, ITPR1,
NLGN1 and NTRK2 via transcription. Low levels of RORA causes dysregulation of these genes and associated pathways. Prefrontal cortex and cerebellum tissue was used.
[350]
ATP
Brain tissue (ATP)
153 families (other)
54 subjects (ATP)
Functional characterisation study
Variant of the HTR2A gene rs6311 in ASD patients has lower level of expression and contains extended 5′untranslated region. Speculation that this variant could be a risk factor in ASD. Frontopolar cortex tissue was used.
[351]
ATP
Brain tissue
28 subjects
Micro-RNA study
Difference in pattern of micro-RNA expression between ASD superior temporal gyrus samples and controls. Further evidence that Mir-320, Mir-132 and Mir-302 are involved in ASD.
[113]
ATP
Brain tissue
94 subjects (total)
51 subjects (ATP)
Acetylome study
Common acetylome signatures found amongst 68% of ASD cases in 5000 regulatory regions in the prefrontal and temporal cortex. Acetylome profiles were not affected by SNPs at these regulatory regions.
[352]
ATP and AGRE
Brain tissue, blood and lymphoblasts
21 subjects (ATP)
252 families (AGRE)
Association study
Variants of LMX1B show modest association with ASD. Analysis of mRNA from anterior cingulate gyrus is much lower in ASD patients compared to controls.
[105]
ATP
Brain tissue
36 subjects
Gene co-expression network analysis
Transcriptional and splicing dysfunction implicated in disorder. Enrichment for genes in glial, immune and neuronal modules. Gene A2BP1 linked to alterations in splicing. Studies based on using temporal cortex, frontal cortex and cerebellum
[353]
ATP
Brain tissue
28 subjects (total)
8 subjects (ATP)
Gene expression analysis
Genes expressed at higher levels in males enriched in upregulated genes in post-mortem neocortical tissue in ASD patients, including astrocyte and microglia markers
[118]
ATP
Brain tissue (ATP)
and placenta
12 subjects (ATP)
64 subjects (total)
Vitamin B12 study
Reduced levels of B12 found in ASD, aged and Schizophrenic patients compared to controls. oxidative stress found in ASD and Schizophrenia patients. Frontal cortex tissue was used
[114]
ATP
Brain tissue
98 subjects
Methylation study
Altered methylation patterns discovered in SHANK3 gene in cerebella tissue of ASD patients
[115]
ATP
Brain tissue
20 subjects
Epigenetic study
Enrichment of 5-hmc in cerebella tissue may be associated with increased binding by MECP2 to RELN and GAD1 promotors
Study numbers listed as families or subjects wherever applicable
Table 6
Overview of studies using Multipe collections
Reference
Biobank
Sample type
Number
Study
Findings
[119]
SSC and AGRE
Blood and genomic data
8044 subjects (AGRE)
4348 subjects (SSC)
Genome-wide survey on translational effects
Investigation of maternal genetic effects in ASD. Validation using other data sets (SSC and AGRE) did not reproduce similar results).
[354]
SSC and AGRE
Genomic and clinical data
Subjects (AGRE) 941
1048 subjects (SSC)
Gene association study
ATP2C2 and MRPL19 found to be associated with language impairment and dyslexia, respectively
[122]
SSC and AGRE
Blood
AGRE–127 subjects
SSC–75 subjects
Parallel sequencing study
Rare variants of the AFF2 gene found to associated with ASD susceptibility in males
[355]
SSC and AGRE
Genomic data
359 subjects (AGRE)
SS–885 subjects
GWAS
Female protective effect in ASD is not mediated by a single genetic locus.
[356]
AGRE and SSC
Genomic data
13 subjects (AGRE)
3 subjects (SSC)
WES
Loss of CTNND2 function linked to severe ASD
[86]
SSC and TASC
Blood and lymphoblasts
15480 (total)
2475(SSC)
601(TASC)
WES
107 genes implicated in ASD. These genes are responsible for synaptic formation, chromatin remodelling and transcriptional regulation
[123]
SSC and AGRE
Blood and lymphoblasts, genomic and clinical data
5657 subjects (total)
1555 subjects (AGRE)
872 subjects (SSC)
WES
Mutations in SHANK1, 2 and 3 accounts for 1 in 50 ASD cases. SHANK1 mutations linked to mild effects, SHANK2 for moderate and SHANK2 for severe.
Study numbers listed as families or subjects wherever applicable
Multiple bio-collections were employed to investigate SHANK1, 2 and 3, which are scaffolding proteins implicated in ASD. They devised a genetic screen and meta-analysis on patients and controls including cohorts from the AGRE, SSC and Swedish twin registry. In total, ~1% of all patients in the study had a mutation in this group of genes. The mutations in SHANK3 had the highest frequency (0.69%) in patients with ASD and profound intellectual disability. SHANK1 (0.04%) and SHANK2 (0.17%) mutations occurred less frequently and were present in individuals with ASD and normal IQ, and ASD with moderate intellectual disability [123].
Recently Autism, Speaks, in coordination with Google and Genome Canada, have launched another initiative; MSSNG (https://​www.​mss.​ng/​). The objective of the MSSNG project is whole genome sequencing of 10000 genomes of families affected by ASD. This incorporates AGRE along with other bio-collections to sequence the entire genomes of families with autistic children, and as of the summer of 2016, it has reached the halfway goal of 5000 genomes out of 10000, with the contribution of the AGRE (1746) and TASC (458). Two studies have been published from this initiative. In the first study, genomes from 200 families were sequenced [124]. The findings revealed many of the de novo mutations (75%) from fathers, which increased dramatically with paternal age. Clustered de novo mutations however were mostly maternal origin, and located near CNV regions subject to high mutation. The ASD genomes were enriched with damaging de novo mutations, of which 15.6% were non-coding and 22.5% genic non-coding, respectively. Many of the mutations affected regulatory regions that are targeted by DNase 1 or involved in exon skipping [124]. The second study [125] featured 5205 sequenced genomes with clinical data, where an average of 73.8 de novo single nucleotide variants and 12.6 insertions/deletions/CNVs were detected per ASD patient. Eighteen new genes were also discovered (CIC, CNOT3, DIP2C, MED13, PAX5, PHF3, SMARCC2, SRSF11, UBN2, DYNC1H1, AGAP2, ADCY3, CLASP1, MYO5A, TAF6, PCDH11X, KIAA2022 and FAM47A) that were not reported in ASD previously. These data clearly demonstrate that ASD is associated with multiple risk factors, and within an ASD individual, and multiple genetic alterations may be present. The Whole genome sequencing is therefore a powerful tool to detect genetic changes at all levels. Resources like MSSNG are valuable, and pooling of ASD bio-collections are essential for identification of the common and subgroup-specific pathways and drug targets of such a multi-factorial disease of ASD which involves hundreds of risk factors.

Stem cell research and autism spectrum stem cell resource

A major impediment to recent drug discovery particularly in the field of neuroscience is the lack of human cell models. The iPSC technology developed by Nobel Laureate Shinya Yamanaka has provided an excellent opportunity [126]. Fibroblasts from patients’ biopsy can be converted into iPSCs with defined transcription factors, which resemble embryonic stem cells and can become most cell types in our body. Therefore, patient-derived iPSCs may be used to investigate disease pathology, progression and mechanisms to create human disease models for drug screening and testing [127, 128].
The SSC has also commenced efforts to create iPSC lines from idiopathic ASD patients who have large head circumference but unknown gene association [129]. The iPSCs were grown into organoids to mimic cortical development, and ASD organoids were shown to display a disproportionate ratio of inhibitory: excitatory neurons. The cortical gene FOXG1 was overexpressed in ASD organoids, and this overexpression correlated with the severity of ASD and their head size [129]. This study has demonstrated a proof-of-concept to model ASD in culture stem cells.
The Children’s Hospital in Orange County California has set up a bio-collection dedicated to this task, the ASD Stem Cell Resource. ASD patients were screened and accepted based on the following criteria: ASD patients if they have no other conditions (i.e. trauma, stroke, seizure disorders) affecting the central nervous system other than ASD; if they have no features of other known genetic conditions (e.g. tuberous sclerosis); Fragile X patients if they are genotypically confirmed for the CGG repeat number of the FMR1 mutation; idiopathic autism patients who are negative for FMR1 mutation and chromosomal abnormality; if they possess an IQ of 40 or greater, and if they are 8-year-old or above. Skin punches and blood were collected in one location (MIND Institute), and fibroblasts were cultured and stored at the Hospital. The collection has been organised into seven groups; unaffected controls, Fragile X without ASD, Fragile X with ASD, permutations without ASD, permutations with ASD, ASD (not meeting full criteria for idiopathic status) and idiopathic ASD.
As of 2014, this resource was composed of iPSCs from 200 unaffected donors and patients. The collection includes fibroblasts, blood, iPSCs, iPSC-derived neuronal and glial cells. The first study published using this bio-collection was the iPSC models of Fragile X syndrome [130]. The Fragile X patient fibroblasts were used to derive iPSCs and differentiate into neurons for transcriptomic analysis. The neuronal differentiation genes (WNT1, BMP4, POU3F4, TFAP2C, PAX3) were shown to be upregulated, whereas potassium channel genes (KCNA1, KCNC3, KCNG2, KCNIP4, KCNJ3, KCNK9, KCNT1) were downregulated in Fragile X iPSC-derived neurons. The temporal regulation of SHANK1 and NNAT genes were also altered, with reduced SHANK1 mRNA and increased NNAT mRNA in patient cells. While the stem cell collection is relatively new, it has great potential to facilitate brain cell culture in vitro, which would otherwise not be feasible by using post mortem brain tissue.

Discussion

It is clear from the studies reviewed here that large ASD bio-collections have had an undisputable impact on progressing genomic discovery in ASD, leading to enhanced understanding of ASD neurobiology. While many studies used private collections as sources for tissue and data, large and well characterised samples from the collections reviewed have supported the discovery of small genetic effects, e.g. in GWAS and rare genetic mutations such as pathogenic CNV and SNV but it is clear, as highlighted for other neurodevelopmental disorders such as Schizophrenia that larger samples are required. Both genetic and phenotypic heterogeneity are impediments to gene discovery. Large bio-collections aim to reduce these effects but challenges remain. Each of the bio-collections reviewed has its own strengths and limitations.

Phenotypic and genotypic heterogeneity

Some of the bio-collections, e.g. SSC, AGRE, TASC, reduced phenotypic heterogeneity through the use of research gold standards for ASD diagnosis, ADI-R and ADOS. Different versions of these instruments based on the timeline when these data have been collected have been used. IQ measurement is more complex to calculate due to the broad range of IQ commonly included within bio-collections. Differences also exist in the clinical profile of subjects included in the different collections with some samples, e.g. SSC, comprised of more individuals with higher cognitive functioning relative to AGRE, TASC or AIC. Medical and psychiatric comorbidities [7] have greater recognition but are not as systematically evaluated in each of the collections. Differences in ascertainment are also relevant. The SSC focused on simplex autism, i.e. families where only one child was affected to maximise the detection of rare variants. Consequently, the relative contribution of common genetic risk within the SSC sample appears reduced. In contrast to autism specific bio-collections, the DNSB, provides a large population-based sample with clinical diagnosis that can maximise power within GWAS studies to detect common genetic variation but does not provide in-depth clinical data for phenotype-genotype analyses. This was evident in the studies on amniotic fluid and DBSS where different diagnostic criteria would have been applied at the time of the subjects’ diagnoses, meaning one criteria would have excluded subjects(ICD8) whereas another would not (ICD10) [95, 96] [93, 94].
Throughout the studies listed here, there is an imbalance of ethnicities of bio-collections, as many of the studies rely heavily on Caucasian/European descent, which has been pointed out in some journals [131] and should consider diverse family structures [132], which can otherwise lead to population stratification [133]. Fortunately, efforts are underway to explore genetics of ASD in other countries such as China [134] and Brazil [135], which will reinforce many of the earlier findings covered in this review.

Samples

Large collections providing DNA for genomics studies have been advantageous; however, as studies move beyond the scope of genetics into transcriptomics, epigenomics and proteomics, a wider variety of sample types will be required. Serum will be valuable for investigating circulating metabolites and proteins that are expressed peripherally, including chemokines [93, 95], cytokines [94, 96], neurotropins [136], MMPs [137] and hormones [138]; however, this may not be the most appropriate tissue to investigate brain relevant ASD genes and proteins. DBSS, which can be useful for WES [139, 140], methylation [141] and gene expression [142], would not be as useful as fresh drawn blood for WGS, as DBSS-derived DNA would need to be amplified prior to use for analysis, potentially causing bias.
However, human brain tissue is a rare resource; brain tissue is very difficult to access due to its scarcity, and the preservation methods used may limit studies being carried out. Also, the types of brain cells are dependent on brain tissue being used; neuronal tissue in grey matter or glial tissue in white matter. Many of the studies listed in the Autism BrainNet, for example, utilised certain sections of the brain; and the most commonly used sections are the prefrontal cortex, temporal cortex, Brodmann’s area, cerebellum and cingulate gyrus. While findings from these sections have been of crucial importance, a capacity to model the entire brain and to observe progression of ASD development would be ideal, and patient’s somatic cells can now be converted to iPSCs and then into disease cell types.
IPSCs have been used as disease models for Fragile X syndrome [143145] and Rett syndrome [146], and iPSCs have been generated from patents with deletions in SHANK3 [147] which are implicated in a number of neurodevelopmental disorders. The three-dimensional culture is developed and iPSCs can also be used to create mini-organoids, which can come very close to mimicking aspects of brain development [129, 148]. In addition to the brain cell types discussed earlier [129, 149], the iPSCs could be used to generate other cell types implicated in ASD co-morbidities, such as the gut [88, 150] and the blood brain barrier [151, 152].
Fibroblasts are the first cell type used to make iPSCs from mice [126] and humans [153] and remain as the most popular cell type for generating neural stem cells, neurons or iPSCs. Fibroblasts are easier to reprogram than many other somatic cells, and the reprogramming efficiency is between 0.1–1% depending on the reprogramming method [154]. They require basic culture media and proliferate rapidly, so large numbers of fibroblasts can be generated in a short period. Unlike keratinocytes they require trained medical personnel to obtain skin biopsies, which could be distressing to some ASD patients. Low passages of fibroblasts are required for reprogramming as higher passages dramatically reduce reprogramming efficiency and increase genomic instability [155].In addition to their use for IPSCs, fibroblasts can be used to investigate amino acid transport, and ASD fibroblasts were found to have greater affinity for transporting alanine, but less affinity for tyrosine—a key component for the synthesis of the neurotransmitter dopamine [156]. Fibroblasts can be used as a proxy to investigate transport across the blood-brain barrier [156, 157] and to investigate calcium signalling [158, 159].
Keratinocytes can also be used for generating IPSCs [160]. Collection is less invasive than skin biopsy and can be carried out by non-medical personnel. The hair samples are easy to transport and culture and transformed cells are easier to identify and isolate. Similar to fibroblasts, keratinocytes are reprogrammed at low passages and fewer methods have been employed to reprogram keratinocytes than fibroblasts. The lentiviral, retroviral and episomal reprogramming were tried successfully [155, 161, 162], and keratinocytes were shown to have high reprogramming efficiency of 1–2%. The major challenge is the reproducibility of keratinocyte growth, and it often requires repeated rounds of hair plucking from a same donor.

Organization

There are many generic articles and white papers for biobanks available, including consensus best-practice recommendations. For those who may wish to start their own bio-collections, we have listed a few articles in Table 7 for further reading on topics pertaining to collection, management, sustainability and quality control. In addition, links to international guidelines can be found here (http://​www.​oecd.​org/​sti/​biotech/​guidelinesforhum​anbiobanksandgen​eticresearchdata​baseshbgrds.​htm; http://​www.​isber.​org/​?​page=​BPR; https://​biospecimens.​cancer.​gov/​practices/​). However, even when using best practice guidelines, the storage and use of bio samples will be subject to the laws where the facilities are located, and will vary from country to country [163].
Table 7
Description of papers relating to aspects of biobanking
Reference
Subject of paper
[357]
Introduces concept of adding value to stakeholders (patient donors/funders/research customers) and to find balance between aspects of sustainability (acceptability/efficiency/accomplishment)
[358]
Feasibility of simplified consent form for biobanking. Result indicates simplified forms combined with supplemental information for further reading effective in minimising form length and complexity
[359]
Review paper detailing best practice guidelines for sample collection and storage, management of data and infrastructure. In addition, ethical, legal and social issues are explored
[360]
Paper discussing aspects of embryonic stem cell banking that can be applied to iPSCS
[361]
Key issues relating to delivery and safety testing of iPSC stocks for use in research and therapy. Importance of international and national coordinated banking systems are also discussed
[362]
Description of enclosed culture system for iPSCS and neural precursors for use in preclinical and basic research

Participation and ethics

Stakeholders can have a considerable influence on how a bio-collection operates and how a bio-collection can be set up, managed and monitored [164]. In addition to researchers, clinicians and parents in bio-collections of ASD research, autistic stakeholders should be included as part of the stakeholder group, which could help guide and inform how research is carried out. A recent survey [165] was carried out amongst researcher-community engagement on ASD research in the UK. A high dissatisfaction and level of disengagement was expressed by parents and patients, who felt that research outcomes made little or no difference to their day-to-day lives and that they were not communicated, not involved or valued. Patients also felt that they did not receive follow-up and researchers were unapproachable and driven by data collection. Establishment and sustainability of a good stakeholder engagement are essential in ASD research and in biobanking. This will not only help guide research to subjects that matter to the community, but also the future of the biobank. One initiative, such as SPARK (Simons Foundation Powering Autism Research for Knowledge) is underway to encourage ASD communities in the USA to participate in ASD research. While such a goal is laudable, it is crucial that participants are engaged in the entire process. They are not just the suppliers of bio-collections for research and data collection, but also make an input into research areas, which directly impinge on the quality of their life. Meanwhile, regular public events to update research progress and challenges to the stakeholder community may help win their understanding, appreciation and continuous support.
The ethics and obtainment of consent are significant factors for bio-collection research. The main considerations include what information shall be given to potential donors regarding the protocol and its implications of the research, how consent should be obtained [166] or what shall be done if consent was not clearly given [167]. It is also a matter of debate whether the consent should be “broad” and if the patient shall consent to a framework of research; if ethical review of each project shall be carried out by independent committees, and what are the strategies to inform and renew consent if there is significant deviation of framework; where shall the consent be revisited and renewed for every new study [168]; how the data will be protected and accessed [169, 170]; and how the findings will be communicated [171]. The latter is especially important if findings are of clinical significance to certain donors or it may affect their health or well-being [167]. These are the issues that each ethical application faces in making the application.
For people with ASD, it can be very complicated. Parents will give consent for their children if they want to donate samples for the bio-collection, but there is a question of adults who may not have the ability to give consent or to fully understand the implications. It is also important to clearly communicate what this research will mean for the patient and the family, including findings that may be of pathological as well as clinical significance. Liu and Scott have commented on how the discoveries made in ASD research can be distorted by media. If parents/patients are misled to believe that a cure will come out a few years down the road, this may lead to disappointment and make them reluctant to participate in further research. Liu and Scott pointed out that the Neurodiversity Movement group (high-functioning autists) would have issues with certain research. They will not participate in research if they feel it may threaten or undermine people with ASD [128]. They prefer investment on services and therapies, rather than on genetic studies which may result in prevention of autistics being born [172174], and the idea of curing autism is a complicated topic of debate [175].
For iPSC research, it was suggested to educate participants on the current state of research, to clearly explain the benefits and risks of biopsy donation and to consult the ASD community on research focus of an ASD bio-collection and on distribution of the cell lines [128]. For clinical trials of stem cells, stem cell counsellors shall inform participants the benefits and risks of enrolling in stem cell trials and to safeguard them from the dangers of stem cell tourism. Such an approach should also be considered for ASD-related studies [176].

Conclusions

In conclusion, bio-collections have been shown as valuable resources and enabled large-scale studies on ASD. The recent genetic studies have begun to reveal de novo mutations on major cellular pathways [17, 177]. There is also emerging evidence that ASD continuum contains subgroups with discrete mutations in specific genes such as CDH8 [88], DYRK1A [71] and POGZ [90] and gene mutations like NRXN1 [28, 60, 73, 178, 179] and SHANKs [72, 98, 114, 123] recurring in broad populations. There is a vast amount of clinical and biological information available in these bio-collections, and the data are in the need for concrete guidelines on ethics and governance. The communication and trust shall be maintained between the researchers and families who have given biological and personal information. Finally, the availability of iPSC resources dedicated to idiopathic and syndromic forms of ASD could be a tremendous boon to the research community and such models are anticipated to be complementary with animal models and to speed up the development of therapeutic interventions for ASD. They could open up the possibilities of functional studies of ASD on a large scale and could become a future model for other iPSC bio-collections to be set up worldwide.

Acknowledgements

I would like to thank all authors who contributed to this manuscript

Funding

Not applicable.

Availability of data and materials

Not applicable.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Authors’ information

Not applicable.
Not applicable.
Not applicable.

Competing interests

The authors declare that they have no competing interests.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated.
Literatur
1.
Zurück zum Zitat Christensen DL, Baio J, Van Naarden Braun K, Bilder D, Charles J, Constantino JN, et al. Prevalence and Characteristics of Autism Spectrum Disorder Among Children Aged 8 Years--Autism and Developmental Disabilities Monitoring Network, 11 Sites, United States, 2012. MMWR Surveill Summ. 2016;65:1–23. doi:10.15585/mmwr.ss6503a1.PubMedCrossRef Christensen DL, Baio J, Van Naarden Braun K, Bilder D, Charles J, Constantino JN, et al. Prevalence and Characteristics of Autism Spectrum Disorder Among Children Aged 8 Years--Autism and Developmental Disabilities Monitoring Network, 11 Sites, United States, 2012. MMWR Surveill Summ. 2016;65:1–23. doi:10.​15585/​mmwr.​ss6503a1.PubMedCrossRef
5.
Zurück zum Zitat Simonoff E, Pickles A, Charman T, Chandler S, Loucas T, Baird G. Psychiatric disorders in children with autism spectrum disorders: Prevalence, comorbidity, and associated factors in a population-derived sample. J Am Acad Child Adolesc Psychiatry. 2008;47:921–9.PubMedCrossRef Simonoff E, Pickles A, Charman T, Chandler S, Loucas T, Baird G. Psychiatric disorders in children with autism spectrum disorders: Prevalence, comorbidity, and associated factors in a population-derived sample. J Am Acad Child Adolesc Psychiatry. 2008;47:921–9.PubMedCrossRef
7.
Zurück zum Zitat Chen M-H, Su T-P, Chen Y-S, Hsu J-W, Huang K-L, Chang W-H, et al. Comorbidity of allergic and autoimmune diseases in patients with autism spectrum disorder: A nationwide population-based study. Res Autism Spectr Disord. 2013;7:205–12. doi:10.1016/j.rasd.2012.08.008.CrossRef Chen M-H, Su T-P, Chen Y-S, Hsu J-W, Huang K-L, Chang W-H, et al. Comorbidity of allergic and autoimmune diseases in patients with autism spectrum disorder: A nationwide population-based study. Res Autism Spectr Disord. 2013;7:205–12. doi:10.​1016/​j.​rasd.​2012.​08.​008.CrossRef
9.
14.
Zurück zum Zitat Bourgeron T. From the genetic architecture to synaptic plasticity in autism spectrum disorder. Nat Rev Neurosci. 2015;16:551–63.PubMedCrossRef Bourgeron T. From the genetic architecture to synaptic plasticity in autism spectrum disorder. Nat Rev Neurosci. 2015;16:551–63.PubMedCrossRef
17.
Zurück zum Zitat Oron O, Elliott E. Delineating the Common Biological Pathways Perturbed by ASD’s Genetic Etiology: Lessons from Network-Based Studies. Int J Mol Sci. 2017;18:828.PubMedCentralCrossRef Oron O, Elliott E. Delineating the Common Biological Pathways Perturbed by ASD’s Genetic Etiology: Lessons from Network-Based Studies. Int J Mol Sci. 2017;18:828.PubMedCentralCrossRef
20.
Zurück zum Zitat Geschwind DH, Sowinski J, Lord C, Iversen P, Shestack J, Jones P, et al. The Autism Genetic Resource Exchange: A Resource for the Study of Autism and Related Neuropsychiatric Conditions. Am J Hum Genet. 2001;69:463–6. doi:10.1086/321292.PubMedPubMedCentralCrossRef Geschwind DH, Sowinski J, Lord C, Iversen P, Shestack J, Jones P, et al. The Autism Genetic Resource Exchange: A Resource for the Study of Autism and Related Neuropsychiatric Conditions. Am J Hum Genet. 2001;69:463–6. doi:10.​1086/​321292.PubMedPubMedCentralCrossRef
25.
Zurück zum Zitat Brick DJ, Nethercott HE, Montesano S, Banuelos MG, Stover AE, Schutte SS, et al. The autism spectrum disorders stem cell resource at children’s hospital of orange county: implications for disease modeling and drug discovery. Stem Cells Transl Med. 2014;3:1275–86. doi:10.5966/sctm.2014-0073.PubMedPubMedCentralCrossRef Brick DJ, Nethercott HE, Montesano S, Banuelos MG, Stover AE, Schutte SS, et al. The autism spectrum disorders stem cell resource at children’s hospital of orange county: implications for disease modeling and drug discovery. Stem Cells Transl Med. 2014;3:1275–86. doi:10.​5966/​sctm.​2014-0073.PubMedPubMedCentralCrossRef
27.
29.
Zurück zum Zitat Jiang Y-H, Sahoo T, Michaelis RC, Bercovich D, Bressler J, Kashork CD, et al. A mixed epigenetic/genetic model for oligogenic inheritance of autism with a limited role for UBE3A. Am J Med Genet A. 2004;131:1–10. doi:10.1002/ajmg.a.30297.PubMedCrossRef Jiang Y-H, Sahoo T, Michaelis RC, Bercovich D, Bressler J, Kashork CD, et al. A mixed epigenetic/genetic model for oligogenic inheritance of autism with a limited role for UBE3A. Am J Med Genet A. 2004;131:1–10. doi:10.​1002/​ajmg.​a.​30297.PubMedCrossRef
30.
Zurück zum Zitat Miller DT, Shen Y, Weiss LA, Korn J, Anselm I, Bridgemohan C, et al. Microdeletion/duplication at 15q13.2q13.3 among individuals with features of autism and other neuropsychiatric disorders. J Med Genet. 2009;46:242–8. doi:10.1136/jmg.2008.059907.PubMedCrossRef Miller DT, Shen Y, Weiss LA, Korn J, Anselm I, Bridgemohan C, et al. Microdeletion/duplication at 15q13.2q13.3 among individuals with features of autism and other neuropsychiatric disorders. J Med Genet. 2009;46:242–8. doi:10.​1136/​jmg.​2008.​059907.PubMedCrossRef
31.
Zurück zum Zitat Sahoo T, Shaw CA, Young AS, Whitehouse NL, Schroer RJ, Stevenson RE, et al. Array-based comparative genomic hybridization analysis of recurrent chromosome 15q rearrangements. Am J Med Genet A. 2005;139A:106–13. doi:10.1002/ajmg.a.31000.PubMedCrossRef Sahoo T, Shaw CA, Young AS, Whitehouse NL, Schroer RJ, Stevenson RE, et al. Array-based comparative genomic hybridization analysis of recurrent chromosome 15q rearrangements. Am J Med Genet A. 2005;139A:106–13. doi:10.​1002/​ajmg.​a.​31000.PubMedCrossRef
32.
Zurück zum Zitat Ma DQ, Whitehead PL, Menold MM, Martin ER, Ashley-Koch AE, Mei H, et al. Identification of significant association and gene-gene interaction of GABA receptor subunit genes in autism. Am J Hum Genet. 2005;77:377–88. doi:10.1086/433195.PubMedPubMedCentralCrossRef Ma DQ, Whitehead PL, Menold MM, Martin ER, Ashley-Koch AE, Mei H, et al. Identification of significant association and gene-gene interaction of GABA receptor subunit genes in autism. Am J Hum Genet. 2005;77:377–88. doi:10.​1086/​433195.PubMedPubMedCentralCrossRef
34.
Zurück zum Zitat Autism Genome Project Consortium, Szatmari P, Paterson AD, Zwaigenbaum L, Roberts W, Brian J, et al. Mapping autism risk loci using genetic linkage and chromosomal rearrangements. Nat Genet. 2007;39:319–28. doi:10.1038/ng1985.CrossRef Autism Genome Project Consortium, Szatmari P, Paterson AD, Zwaigenbaum L, Roberts W, Brian J, et al. Mapping autism risk loci using genetic linkage and chromosomal rearrangements. Nat Genet. 2007;39:319–28. doi:10.​1038/​ng1985.CrossRef
36.
Zurück zum Zitat Harvey CG, Menon SD, Stachowiak B, Noor A, Proctor A, Mensah AK, et al. Sequence variants within exon 1 of MECP2 occur in females with mental retardation. Am J Med Genet B Neuropsychiatr Genet. 2007;144B:355–60. doi:10.1002/ajmg.b.30425.PubMedCrossRef Harvey CG, Menon SD, Stachowiak B, Noor A, Proctor A, Mensah AK, et al. Sequence variants within exon 1 of MECP2 occur in females with mental retardation. Am J Med Genet B Neuropsychiatr Genet. 2007;144B:355–60. doi:10.​1002/​ajmg.​b.​30425.PubMedCrossRef
38.
42.
43.
Zurück zum Zitat Ashley-Koch AE, Mei H, Jaworski J, Ma DQ, Ritchie MD, Menold MM, et al. An analysis paradigm for investigating multi-locus effects in complex disease: examination of three GABA receptor subunit genes on 15q11-q13 as risk factors for autistic disorder. Ann Hum Genet. 2006;70(Pt 3):281–92. doi:10.1111/j.1469-1809.2006.00253.x.PubMedCrossRef Ashley-Koch AE, Mei H, Jaworski J, Ma DQ, Ritchie MD, Menold MM, et al. An analysis paradigm for investigating multi-locus effects in complex disease: examination of three GABA receptor subunit genes on 15q11-q13 as risk factors for autistic disorder. Ann Hum Genet. 2006;70(Pt 3):281–92. doi:10.​1111/​j.​1469-1809.​2006.​00253.​x.PubMedCrossRef
47.
Zurück zum Zitat Nguyen A, Rauch TA, Pfeifer GP, Hu VW. Global methylation profiling of lymphoblastoid cell lines reveals epigenetic contributions to autism spectrum disorders and a novel autism candidate gene, RORA, whose protein product is reduced in autistic brain. FASEB J. 2010;24:3036–51. doi:10.1096/fj.10-154484.PubMedPubMedCentralCrossRef Nguyen A, Rauch TA, Pfeifer GP, Hu VW. Global methylation profiling of lymphoblastoid cell lines reveals epigenetic contributions to autism spectrum disorders and a novel autism candidate gene, RORA, whose protein product is reduced in autistic brain. FASEB J. 2010;24:3036–51. doi:10.​1096/​fj.​10-154484.PubMedPubMedCentralCrossRef
50.
Zurück zum Zitat Campbell DB, Warren D, Sutcliffe JS, Lee EB, Levitt P. Association of MET with social and communication phenotypes in individuals with autism spectrum disorder. Am J Med Genet B Neuropsychiatr Genet. 2010;153B:438–46. doi:10.1002/ajmg.b.30998.PubMedCrossRef Campbell DB, Warren D, Sutcliffe JS, Lee EB, Levitt P. Association of MET with social and communication phenotypes in individuals with autism spectrum disorder. Am J Med Genet B Neuropsychiatr Genet. 2010;153B:438–46. doi:10.​1002/​ajmg.​b.​30998.PubMedCrossRef
52.
Zurück zum Zitat McCauley JL, Olson LM, Dowd M, Amin T, Steele A, Blakely RD, et al. Linkage and association analysis at the serotonin transporter (SLC6A4) locus in a rigid-compulsive subset of autism. Am J Med Genet B Neuropsychiatr Genet. 2004;127B:104–12. doi:10.1002/ajmg.b.20151.PubMedCrossRef McCauley JL, Olson LM, Dowd M, Amin T, Steele A, Blakely RD, et al. Linkage and association analysis at the serotonin transporter (SLC6A4) locus in a rigid-compulsive subset of autism. Am J Med Genet B Neuropsychiatr Genet. 2004;127B:104–12. doi:10.​1002/​ajmg.​b.​20151.PubMedCrossRef
55.
Zurück zum Zitat McCauley JL, Olson LM, Delahanty R, Amin T, Nurmi EL, Organ EL, et al. A linkage disequilibrium map of the 1-Mb 15q12 GABA(A) receptor subunit cluster and association to autism. Am J Med Genet B Neuropsychiatr Genet. 2004;131B:51–9. doi:10.1002/ajmg.b.30038.PubMedCrossRef McCauley JL, Olson LM, Delahanty R, Amin T, Nurmi EL, Organ EL, et al. A linkage disequilibrium map of the 1-Mb 15q12 GABA(A) receptor subunit cluster and association to autism. Am J Med Genet B Neuropsychiatr Genet. 2004;131B:51–9. doi:10.​1002/​ajmg.​b.​30038.PubMedCrossRef
56.
Zurück zum Zitat Strom SP, Stone JL, Ten Bosch JR, Merriman B, Cantor RM, Geschwind DH, et al. High-density SNP association study of the 17q21 chromosomal region linked to autism identifies CACNA1G as a novel candidate gene. Mol Psychiatry. 2010;15:996–1005. doi:10.1038/mp.2009.41.PubMedCrossRef Strom SP, Stone JL, Ten Bosch JR, Merriman B, Cantor RM, Geschwind DH, et al. High-density SNP association study of the 17q21 chromosomal region linked to autism identifies CACNA1G as a novel candidate gene. Mol Psychiatry. 2010;15:996–1005. doi:10.​1038/​mp.​2009.​41.PubMedCrossRef
71.
Zurück zum Zitat Van Bon BWM, Coe BP, Bernier R, Green C, Gerdts J, Witherspoon K, et al. Disruptive de novo mutations of DYRK1A lead to a syndromic form of autism and ID. Mol Psychiatry. 2016;21:126–32. doi:10.1038/mp.2015.5.PubMedCrossRef Van Bon BWM, Coe BP, Bernier R, Green C, Gerdts J, Witherspoon K, et al. Disruptive de novo mutations of DYRK1A lead to a syndromic form of autism and ID. Mol Psychiatry. 2016;21:126–32. doi:10.​1038/​mp.​2015.​5.PubMedCrossRef
77.
Zurück zum Zitat Main PAE, Thomas P, Esterman A, Fenech MF. Necrosis is increased in lymphoblastoid cell lines from children with autism compared with their non-autistic siblings under conditions of oxidative and nitrosative stress. Mutagenesis. 2013;28:475–84. doi:10.1093/mutage/get025.PubMedPubMedCentralCrossRef Main PAE, Thomas P, Esterman A, Fenech MF. Necrosis is increased in lymphoblastoid cell lines from children with autism compared with their non-autistic siblings under conditions of oxidative and nitrosative stress. Mutagenesis. 2013;28:475–84. doi:10.​1093/​mutage/​get025.PubMedPubMedCentralCrossRef
79.
Zurück zum Zitat Sarachana T, Zhou R, Chen G, Manji HK, Hu VW. Investigation of post-transcriptional gene regulatory networks associated with autism spectrum disorders by microRNA expression profiling of lymphoblastoid cell lines. Genome Med. 2010;2:23. doi:10.1186/gm144.PubMedPubMedCentralCrossRef Sarachana T, Zhou R, Chen G, Manji HK, Hu VW. Investigation of post-transcriptional gene regulatory networks associated with autism spectrum disorders by microRNA expression profiling of lymphoblastoid cell lines. Genome Med. 2010;2:23. doi:10.​1186/​gm144.PubMedPubMedCentralCrossRef
80.
Zurück zum Zitat Oguro-Ando A, Rosensweig C, Herman E, Nishimura Y, Werling D, Bill BR, et al. Increased CYFIP1 dosage alters cellular and dendritic morphology and dysregulates mTOR. Mol Psychiatry. 2015;20:1069–78. doi:10.1038/mp.2014.124.PubMedCrossRef Oguro-Ando A, Rosensweig C, Herman E, Nishimura Y, Werling D, Bill BR, et al. Increased CYFIP1 dosage alters cellular and dendritic morphology and dysregulates mTOR. Mol Psychiatry. 2015;20:1069–78. doi:10.​1038/​mp.​2014.​124.PubMedCrossRef
81.
82.
Zurück zum Zitat Cai G, Edelmann L, Goldsmith JE, Cohen N, Nakamine A, Reichert JG, et al. Multiplex ligation-dependent probe amplification for genetic screening in autism spectrum disorders: efficient identification of known microduplications and identification of a novel microduplication in ASMT. BMC Med Genomics. 2008;1:50. doi:10.1186/1755-8794-1-50.PubMedPubMedCentralCrossRef Cai G, Edelmann L, Goldsmith JE, Cohen N, Nakamine A, Reichert JG, et al. Multiplex ligation-dependent probe amplification for genetic screening in autism spectrum disorders: efficient identification of known microduplications and identification of a novel microduplication in ASMT. BMC Med Genomics. 2008;1:50. doi:10.​1186/​1755-8794-1-50.PubMedPubMedCentralCrossRef
84.
Zurück zum Zitat Yonan AL, Palmer AA, Smith KC, Feldman I, Lee HK, Yonan JM, et al. Bioinformatic analysis of autism positional candidate genes using biological databases and computational gene network prediction. Genes Brain Behav. 2003;2:303–20.PubMedCrossRef Yonan AL, Palmer AA, Smith KC, Feldman I, Lee HK, Yonan JM, et al. Bioinformatic analysis of autism positional candidate genes using biological databases and computational gene network prediction. Genes Brain Behav. 2003;2:303–20.PubMedCrossRef
85.
92.
Zurück zum Zitat The Autism Spectrum Disorders Working Group of The Psychiatric Genomics Consortium. Meta-analysis of GWAS of over 16,000 individuals with autism spectrum disorder highlights a novel locus at 10q24.32 and a significant overlap with schizophrenia. Mol Autism. 2017;8:21. doi:10.1186/s13229-017-0137-9.CrossRef The Autism Spectrum Disorders Working Group of The Psychiatric Genomics Consortium. Meta-analysis of GWAS of over 16,000 individuals with autism spectrum disorder highlights a novel locus at 10q24.32 and a significant overlap with schizophrenia. Mol Autism. 2017;8:21. doi:10.​1186/​s13229-017-0137-9.CrossRef
93.
Zurück zum Zitat Abdallah MW, Larsen N, Grove J, Nørgaard-Pedersen B, Thorsen P, Mortensen EL, et al. Amniotic fluid chemokines and autism spectrum disorders: an exploratory study utilizing a Danish Historic Birth Cohort. Brain Behav Immun. 2012;26:170–6. doi:10.1016/j.bbi.2011.09.003.PubMedCrossRef Abdallah MW, Larsen N, Grove J, Nørgaard-Pedersen B, Thorsen P, Mortensen EL, et al. Amniotic fluid chemokines and autism spectrum disorders: an exploratory study utilizing a Danish Historic Birth Cohort. Brain Behav Immun. 2012;26:170–6. doi:10.​1016/​j.​bbi.​2011.​09.​003.PubMedCrossRef
94.
Zurück zum Zitat Abdallah MW, Larsen N, Grove J, Nørgaard-Pedersen B, Thorsen P, Mortensen EL, et al. Amniotic fluid inflammatory cytokines: potential markers of immunologic dysfunction in autism spectrum disorders. World J Biol Psychiatry. 2013;14:528–38. doi:10.3109/15622975.2011.639803.PubMedCrossRef Abdallah MW, Larsen N, Grove J, Nørgaard-Pedersen B, Thorsen P, Mortensen EL, et al. Amniotic fluid inflammatory cytokines: potential markers of immunologic dysfunction in autism spectrum disorders. World J Biol Psychiatry. 2013;14:528–38. doi:10.​3109/​15622975.​2011.​639803.PubMedCrossRef
95.
Zurück zum Zitat Abdallah MW, Larsen N, Grove J, Bonefeld-Jørgensen EC, Nørgaard-Pedersen B, Hougaard DM, et al. Neonatal chemokine levels and risk of autism spectrum disorders: findings from a Danish historic birth cohort follow-up study. Cytokine. 2013;61:370–6. doi:10.1016/j.cyto.2012.11.015.PubMedCrossRef Abdallah MW, Larsen N, Grove J, Bonefeld-Jørgensen EC, Nørgaard-Pedersen B, Hougaard DM, et al. Neonatal chemokine levels and risk of autism spectrum disorders: findings from a Danish historic birth cohort follow-up study. Cytokine. 2013;61:370–6. doi:10.​1016/​j.​cyto.​2012.​11.​015.PubMedCrossRef
96.
Zurück zum Zitat Abdallah MW, Larsen N, Mortensen EL, Atladóttir HÓ, Nørgaard-Pedersen B, Bonefeld-Jørgensen EC, et al. Neonatal levels of cytokines and risk of autism spectrum disorders: an exploratory register-based historic birth cohort study utilizing the Danish Newborn Screening Biobank. J Neuroimmunol. 2012;252:75–82. doi:10.1016/j.jneuroim.2012.07.013.PubMedCrossRef Abdallah MW, Larsen N, Mortensen EL, Atladóttir HÓ, Nørgaard-Pedersen B, Bonefeld-Jørgensen EC, et al. Neonatal levels of cytokines and risk of autism spectrum disorders: an exploratory register-based historic birth cohort study utilizing the Danish Newborn Screening Biobank. J Neuroimmunol. 2012;252:75–82. doi:10.​1016/​j.​jneuroim.​2012.​07.​013.PubMedCrossRef
104.
106.
Zurück zum Zitat Ben-David E, Shohat S, Shifman S. Allelic expression analysis in the brain suggests a role for heterogeneous insults affecting epigenetic processes in autism spectrum disorders. Hum Mol Genet. 2014;23:4111–24. doi:10.1093/hmg/ddu128.PubMedCrossRef Ben-David E, Shohat S, Shifman S. Allelic expression analysis in the brain suggests a role for heterogeneous insults affecting epigenetic processes in autism spectrum disorders. Hum Mol Genet. 2014;23:4111–24. doi:10.​1093/​hmg/​ddu128.PubMedCrossRef
109.
114.
Zurück zum Zitat Zhu L, Wang X, Li X-L, Towers A, Cao X, Wang P, et al. Epigenetic dysregulation of SHANK3 in brain tissues from individuals with autism spectrum disorders. Hum Mol Genet. 2014;23:1563–78. doi:10.1093/hmg/ddt547.PubMedCrossRef Zhu L, Wang X, Li X-L, Towers A, Cao X, Wang P, et al. Epigenetic dysregulation of SHANK3 in brain tissues from individuals with autism spectrum disorders. Hum Mol Genet. 2014;23:1563–78. doi:10.​1093/​hmg/​ddt547.PubMedCrossRef
115.
120.
Zurück zum Zitat Desachy G, Croen LA, Torres AR, Kharrazi M, Delorenze GN, Windham GC, et al. Increased female autosomal burden of rare copy number variants in human populations and in autism families. Mol Psychiatry. 2015;20:170–5. doi:10.1038/mp.2014.179.PubMedCrossRef Desachy G, Croen LA, Torres AR, Kharrazi M, Delorenze GN, Windham GC, et al. Increased female autosomal burden of rare copy number variants in human populations and in autism families. Mol Psychiatry. 2015;20:170–5. doi:10.​1038/​mp.​2014.​179.PubMedCrossRef
124.
Zurück zum Zitat Yuen RKC, Merico D, Cao H, Pellecchia G, Alipanahi B, Thiruvahindrapuram B, et al. Genome-wide characteristics of de novo mutations in autism. npj. Genomic Med. 2016;1:160271–1602710. doi:10.1038/npjgenmed.2016.27. Yuen RKC, Merico D, Cao H, Pellecchia G, Alipanahi B, Thiruvahindrapuram B, et al. Genome-wide characteristics of de novo mutations in autism. npj. Genomic Med. 2016;1:160271–1602710. doi:10.​1038/​npjgenmed.​2016.​27.
125.
Zurück zum Zitat C Yuen RK, Merico D, Bookman M, L Howe J, Thiruvahindrapuram B, Patel RV, et al. Whole genome sequencing resource identifies 18 new candidate genes for autism spectrum disorder. Nat Neurosci. 2017;20:602–11. doi:10.1038/nn.4524.PubMedCrossRef C Yuen RK, Merico D, Bookman M, L Howe J, Thiruvahindrapuram B, Patel RV, et al. Whole genome sequencing resource identifies 18 new candidate genes for autism spectrum disorder. Nat Neurosci. 2017;20:602–11. doi:10.​1038/​nn.​4524.PubMedCrossRef
130.
Zurück zum Zitat Lu P, Chen X, Feng Y, Zeng Q, Jiang C, Zhu X, et al. Integrated transcriptome analysis of human iPS cells derived from a fragile X syndrome patient during neuronal differentiation. Sci China Life Sci. 2016;59:1093–105. doi:10.1007/s11427-016-0194-6.PubMedCrossRef Lu P, Chen X, Feng Y, Zeng Q, Jiang C, Zhu X, et al. Integrated transcriptome analysis of human iPS cells derived from a fragile X syndrome patient during neuronal differentiation. Sci China Life Sci. 2016;59:1093–105. doi:10.​1007/​s11427-016-0194-6.PubMedCrossRef
131.
Zurück zum Zitat Pierce NP, O’Reilly MF, Sorrells AM, Fragale CL, White PJ, Aguilar JM, et al. Ethnicity reporting practices for empirical research in three autism-related journals. J Autism Dev Disord. 2014;44:1507–19. doi:10.1007/s10803-014-2041-x.PubMedCrossRef Pierce NP, O’Reilly MF, Sorrells AM, Fragale CL, White PJ, Aguilar JM, et al. Ethnicity reporting practices for empirical research in three autism-related journals. J Autism Dev Disord. 2014;44:1507–19. doi:10.​1007/​s10803-014-2041-x.PubMedCrossRef
132.
Zurück zum Zitat Hilton CL, Fitzgerald RT, Jackson KM, Maxim RA, Bosworth CC, Shattuck PT, et al. Brief report: Under-representation of African americans in autism genetic research: a rationale for inclusion of subjects representing diverse family structures. J Autism Dev Disord. 2010;40:633–9. doi:10.1007/s10803-009-0905-2.PubMedPubMedCentralCrossRef Hilton CL, Fitzgerald RT, Jackson KM, Maxim RA, Bosworth CC, Shattuck PT, et al. Brief report: Under-representation of African americans in autism genetic research: a rationale for inclusion of subjects representing diverse family structures. J Autism Dev Disord. 2010;40:633–9. doi:10.​1007/​s10803-009-0905-2.PubMedPubMedCentralCrossRef
133.
Zurück zum Zitat Robinson EB, Howrigan D, Yang J, Ripke S, Anttila V, Duncan LE, et al. Response to “Predicting the diagnosis of autism spectrum disorder using gene pathway analysis”. Mol Psychiatry. 2014;19:859–61. doi:10.1038/mp.2013.125.PubMedCrossRef Robinson EB, Howrigan D, Yang J, Ripke S, Anttila V, Duncan LE, et al. Response to “Predicting the diagnosis of autism spectrum disorder using gene pathway analysis”. Mol Psychiatry. 2014;19:859–61. doi:10.​1038/​mp.​2013.​125.PubMedCrossRef
135.
Zurück zum Zitat Nascimento PP, Bossolani-Martins AL, Rosan DBA, Mattos LC, Brandão-Mattos C, Fett-Conte AC. Single nucleotide polymorphisms in the CNTNAP2 gene in Brazilian patients with autistic spectrum disorder. Genet Mol Res. 2016;15. doi:10.4238/gmr.15017422. Nascimento PP, Bossolani-Martins AL, Rosan DBA, Mattos LC, Brandão-Mattos C, Fett-Conte AC. Single nucleotide polymorphisms in the CNTNAP2 gene in Brazilian patients with autistic spectrum disorder. Genet Mol Res. 2016;15. doi:10.​4238/​gmr.​15017422.
136.
Zurück zum Zitat Abdallah MW, Mortensen EL, Greaves-Lord K, Larsen N, Bonefeld-Jørgensen EC, Nørgaard-Pedersen B, et al. Neonatal levels of neurotrophic factors and risk of autism spectrum disorders. Acta Psychiatr Scand. 2013;128:61–9. doi:10.1111/acps.12020.PubMedCrossRef Abdallah MW, Mortensen EL, Greaves-Lord K, Larsen N, Bonefeld-Jørgensen EC, Nørgaard-Pedersen B, et al. Neonatal levels of neurotrophic factors and risk of autism spectrum disorders. Acta Psychiatr Scand. 2013;128:61–9. doi:10.​1111/​acps.​12020.PubMedCrossRef
137.
Zurück zum Zitat Abdallah MW, Pearce BD, Larsen N, Greaves-Lord K, Nørgaard-Pedersen B, Hougaard DM, et al. Amniotic fluid MMP-9 and neurotrophins in autism spectrum disorders: an exploratory study. Autism Res. 2012;5:428–33. doi:10.1002/aur.1254.PubMedCrossRef Abdallah MW, Pearce BD, Larsen N, Greaves-Lord K, Nørgaard-Pedersen B, Hougaard DM, et al. Amniotic fluid MMP-9 and neurotrophins in autism spectrum disorders: an exploratory study. Autism Res. 2012;5:428–33. doi:10.​1002/​aur.​1254.PubMedCrossRef
138.
139.
Zurück zum Zitat Hollegaard MV, Grauholm J, Nielsen R, Grove J, Mandrup S, Hougaard DM. Archived neonatal dried blood spot samples can be used for accurate whole genome and exome-targeted next-generation sequencing. Mol Genet Metab. 2013;110:65–72. doi:10.1016/j.ymgme.2013.06.004.PubMedCrossRef Hollegaard MV, Grauholm J, Nielsen R, Grove J, Mandrup S, Hougaard DM. Archived neonatal dried blood spot samples can be used for accurate whole genome and exome-targeted next-generation sequencing. Mol Genet Metab. 2013;110:65–72. doi:10.​1016/​j.​ymgme.​2013.​06.​004.PubMedCrossRef
142.
144.
Zurück zum Zitat Bar-Nur O, Caspi I, Benvenisty N. Molecular analysis of FMR1 reactivation in fragile-X induced pluripotent stem cells and their neuronal derivatives. J Mol Cell Biol. 2012;4:180–3. doi:10.1093/jmcb/mjs007.PubMedCrossRef Bar-Nur O, Caspi I, Benvenisty N. Molecular analysis of FMR1 reactivation in fragile-X induced pluripotent stem cells and their neuronal derivatives. J Mol Cell Biol. 2012;4:180–3. doi:10.​1093/​jmcb/​mjs007.PubMedCrossRef
147.
Zurück zum Zitat Cocks G, Curran S, Gami P, Uwanogho D, Jeffries AR, Kathuria A, et al. The utility of patient specific induced pluripotent stem cells for the modelling of Autistic Spectrum Disorders. Psychopharmacology (Berl). 2014;231:1079–88. doi:10.1007/s00213-013-3196-4.CrossRef Cocks G, Curran S, Gami P, Uwanogho D, Jeffries AR, Kathuria A, et al. The utility of patient specific induced pluripotent stem cells for the modelling of Autistic Spectrum Disorders. Psychopharmacology (Berl). 2014;231:1079–88. doi:10.​1007/​s00213-013-3196-4.CrossRef
149.
Zurück zum Zitat Griesi-Oliveira K, Acab A, Gupta AR, Sunaga DY, Chailangkarn T, Nicol X, et al. Modeling non-syndromic autism and the impact of TRPC6 disruption in human neurons. Mol Psychiatry. 2015;20:1350–65. doi:10.1038/mp.2014.141.PubMedCrossRef Griesi-Oliveira K, Acab A, Gupta AR, Sunaga DY, Chailangkarn T, Nicol X, et al. Modeling non-syndromic autism and the impact of TRPC6 disruption in human neurons. Mol Psychiatry. 2015;20:1350–65. doi:10.​1038/​mp.​2014.​141.PubMedCrossRef
150.
Zurück zum Zitat Mahe MM, Workman M, Trisno S, Poling H, Watson CL, Sundaram N, et al. Functional enteric nervous system in human small intestine derived from pluripotent stem cells. Neurogastroenterol Motil. 2016;28:6–6. Mahe MM, Workman M, Trisno S, Poling H, Watson CL, Sundaram N, et al. Functional enteric nervous system in human small intestine derived from pluripotent stem cells. Neurogastroenterol Motil. 2016;28:6–6.
151.
154.
160.
161.
Zurück zum Zitat Streckfuss-Bömeke K, Wolf F, Azizian A, Stauske M, Tiburcy M, Wagner S, et al. Comparative study of human-induced pluripotent stem cells derived from bone marrow cells, hair keratinocytes, and skin fibroblasts. Eur Heart J. 2013;34:2618–29. doi:10.1093/eurheartj/ehs203.PubMedCrossRef Streckfuss-Bömeke K, Wolf F, Azizian A, Stauske M, Tiburcy M, Wagner S, et al. Comparative study of human-induced pluripotent stem cells derived from bone marrow cells, hair keratinocytes, and skin fibroblasts. Eur Heart J. 2013;34:2618–29. doi:10.​1093/​eurheartj/​ehs203.PubMedCrossRef
173.
180.
Zurück zum Zitat Alarcón M, Cantor RM, Liu J, Gilliam TC, Geschwind DH. Autism Genetic Research Exchange Consortium. Evidence for a language quantitative trait locus on chromosome 7q in multiplex autism families. Am J Hum Genet. 2002;70:60–71. doi:10.1086/338241.PubMedCrossRef Alarcón M, Cantor RM, Liu J, Gilliam TC, Geschwind DH. Autism Genetic Research Exchange Consortium. Evidence for a language quantitative trait locus on chromosome 7q in multiplex autism families. Am J Hum Genet. 2002;70:60–71. doi:10.​1086/​338241.PubMedCrossRef
182.
183.
Zurück zum Zitat Allen-Brady K, Cannon D, Robison R, McMahon WM, Coon H. A unified theory of autism revisited: linkage evidence points to chromosome X using a high-risk subset of AGRE families. Autism Res. 2010;3:47–52. doi:10.1002/aur.119.PubMed Allen-Brady K, Cannon D, Robison R, McMahon WM, Coon H. A unified theory of autism revisited: linkage evidence points to chromosome X using a high-risk subset of AGRE families. Autism Res. 2010;3:47–52. doi:10.​1002/​aur.​119.PubMed
184.
Zurück zum Zitat Anitha A, Nakamura K, Yamada K, Suda S, Thanseem I, Tsujii M, et al. Genetic analyses of roundabout (ROBO) axon guidance receptors in autism. Am J Med Genet B Neuropsychiatr Genet. 2008;147B:1019–27. doi:10.1002/ajmg.b.30697.PubMedCrossRef Anitha A, Nakamura K, Yamada K, Suda S, Thanseem I, Tsujii M, et al. Genetic analyses of roundabout (ROBO) axon guidance receptors in autism. Am J Med Genet B Neuropsychiatr Genet. 2008;147B:1019–27. doi:10.​1002/​ajmg.​b.​30697.PubMedCrossRef
188.
190.
Zurück zum Zitat Baron CA, Tepper CG, Liu SY, Davis RR, Wang NJ, Schanen NC, et al. Genomic and functional profiling of duplicated chromosome 15 cell lines reveal regulatory alterations in UBE3A-associated ubiquitin-proteasome pathway processes. Hum Mol Genet. 2006;15:853–69. doi:10.1093/hmg/ddl004.PubMedCrossRef Baron CA, Tepper CG, Liu SY, Davis RR, Wang NJ, Schanen NC, et al. Genomic and functional profiling of duplicated chromosome 15 cell lines reveal regulatory alterations in UBE3A-associated ubiquitin-proteasome pathway processes. Hum Mol Genet. 2006;15:853–69. doi:10.​1093/​hmg/​ddl004.PubMedCrossRef
192.
Zurück zum Zitat Benayed R, Gharani N, Rossman I, Mancuso V, Lazar G, Kamdar S, et al. Support for the homeobox transcription factor gene ENGRAILED 2 as an autism spectrum disorder susceptibility locus. Am J Hum Genet. 2005;77:851–68. doi:10.1086/497705.PubMedPubMedCentralCrossRef Benayed R, Gharani N, Rossman I, Mancuso V, Lazar G, Kamdar S, et al. Support for the homeobox transcription factor gene ENGRAILED 2 as an autism spectrum disorder susceptibility locus. Am J Hum Genet. 2005;77:851–68. doi:10.​1086/​497705.PubMedPubMedCentralCrossRef
193.
Zurück zum Zitat Buxbaum JD, Silverman JM, Smith CJ, Kilifarski M, Reichert J, Hollander E, et al. Evidence for a susceptibility gene for autism on chromosome 2 and for genetic heterogeneity. Am J Hum Genet. 2001;68:1514–20. doi:10.1086/320588.PubMedPubMedCentralCrossRef Buxbaum JD, Silverman JM, Smith CJ, Kilifarski M, Reichert J, Hollander E, et al. Evidence for a susceptibility gene for autism on chromosome 2 and for genetic heterogeneity. Am J Hum Genet. 2001;68:1514–20. doi:10.​1086/​320588.PubMedPubMedCentralCrossRef
194.
Zurück zum Zitat Buxbaum JD, Silverman J, Keddache M, Smith CJ, Hollander E, Ramoz N, et al. Linkage analysis for autism in a subset families with obsessive-compulsive behaviors: evidence for an autism susceptibility gene on chromosome 1 and further support for susceptibility genes on chromosome 6 and 19. Mol Psychiatry. 2004;9:144–50. doi:10.1038/sj.mp.4001465.PubMedCrossRef Buxbaum JD, Silverman J, Keddache M, Smith CJ, Hollander E, Ramoz N, et al. Linkage analysis for autism in a subset families with obsessive-compulsive behaviors: evidence for an autism susceptibility gene on chromosome 1 and further support for susceptibility genes on chromosome 6 and 19. Mol Psychiatry. 2004;9:144–50. doi:10.​1038/​sj.​mp.​4001465.PubMedCrossRef
199.
Zurück zum Zitat Chang S-C, Pauls DL, Lange C, Sasanfar R, Santangelo SL. Common genetic variation in the GAD1 gene and the entire family of DLX homeobox genes and autism spectrum disorders. Am J Med Genet B Neuropsychiatr Genet. 2011;156:233–9. doi:10.1002/ajmg.b.31148.PubMedCrossRef Chang S-C, Pauls DL, Lange C, Sasanfar R, Santangelo SL. Common genetic variation in the GAD1 gene and the entire family of DLX homeobox genes and autism spectrum disorders. Am J Med Genet B Neuropsychiatr Genet. 2011;156:233–9. doi:10.​1002/​ajmg.​b.​31148.PubMedCrossRef
200.
Zurück zum Zitat Chang S-C, Pauls DL, Lange C, Sasanfar R, Santangelo SL. Sex-specific association of a common variant of the XG gene with autism spectrum disorders. Am J Med Genet B Neuropsychiatr Genet. 2013;162B:742–50. doi:10.1002/ajmg.b.32165.PubMedCrossRef Chang S-C, Pauls DL, Lange C, Sasanfar R, Santangelo SL. Sex-specific association of a common variant of the XG gene with autism spectrum disorders. Am J Med Genet B Neuropsychiatr Genet. 2013;162B:742–50. doi:10.​1002/​ajmg.​b.​32165.PubMedCrossRef
202.
Zurück zum Zitat Cheslack-Postava K, Fallin MD, Avramopoulos D, Connors SL, Zimmerman AW, Eberhart CG, et al. beta2-Adrenergic receptor gene variants and risk for autism in the AGRE cohort. Mol Psychiatry. 2007;12:283–91. doi:10.1038/sj.mp.4001940.PubMed Cheslack-Postava K, Fallin MD, Avramopoulos D, Connors SL, Zimmerman AW, Eberhart CG, et al. beta2-Adrenergic receptor gene variants and risk for autism in the AGRE cohort. Mol Psychiatry. 2007;12:283–91. doi:10.​1038/​sj.​mp.​4001940.PubMed
203.
206.
Zurück zum Zitat Connors SL, Crowell DE, Eberhart CG, Copeland J, Newschaffer CJ, Spence SJ, et al. beta2-adrenergic receptor activation and genetic polymorphisms in autism: data from dizygotic twins. J Child Neurol. 2005;20:876–84. doi:10.1177/08830738050200110401.PubMedCrossRef Connors SL, Crowell DE, Eberhart CG, Copeland J, Newschaffer CJ, Spence SJ, et al. beta2-adrenergic receptor activation and genetic polymorphisms in autism: data from dizygotic twins. J Child Neurol. 2005;20:876–84. doi:10.​1177/​0883073805020011​0401.PubMedCrossRef
207.
Zurück zum Zitat D’Amelio M, Ricci I, Sacco R, Liu X, D’Agruma L, Muscarella LA, et al. Paraoxonase gene variants are associated with autism in North America, but not in Italy: possible regional specificity in gene-environment interactions. Mol Psychiatry. 2005;10:1006–16. doi:10.1038/sj.mp.4001714.PubMedCrossRef D’Amelio M, Ricci I, Sacco R, Liu X, D’Agruma L, Muscarella LA, et al. Paraoxonase gene variants are associated with autism in North America, but not in Italy: possible regional specificity in gene-environment interactions. Mol Psychiatry. 2005;10:1006–16. doi:10.​1038/​sj.​mp.​4001714.PubMedCrossRef
212.
Zurück zum Zitat Fatemi SH, Stary JM, Egan EA. Reduced blood levels of reelin as a vulnerability factor in pathophysiology of autistic disorder. Cell Mol Neurobiol. 2002;22:139–52.PubMedCrossRef Fatemi SH, Stary JM, Egan EA. Reduced blood levels of reelin as a vulnerability factor in pathophysiology of autistic disorder. Cell Mol Neurobiol. 2002;22:139–52.PubMedCrossRef
213.
215.
Zurück zum Zitat Fradin D, Cheslack-Postava K, Ladd-Acosta C, Newschaffer C, Chakravarti A, Arking DE, et al. Parent-of-origin effects in autism identified through genome-wide linkage analysis of 16,000 SNPs. PLoS ONE. 2010;5. doi:10.1371/journal.pone.0012513. Fradin D, Cheslack-Postava K, Ladd-Acosta C, Newschaffer C, Chakravarti A, Arking DE, et al. Parent-of-origin effects in autism identified through genome-wide linkage analysis of 16,000 SNPs. PLoS ONE. 2010;5. doi:10.​1371/​journal.​pone.​0012513.
217.
Zurück zum Zitat Goin-Kochel RP, Porter AE, Peters SU, Shinawi M, Sahoo T, Beaudet AL. The MTHFR 677C→T polymorphism and behaviors in children with autism: exploratory genotype-phenotype correlations. Autism Res. 2009;2:98–108. doi:10.1002/aur.70.PubMedCrossRef Goin-Kochel RP, Porter AE, Peters SU, Shinawi M, Sahoo T, Beaudet AL. The MTHFR 677C→T polymorphism and behaviors in children with autism: exploratory genotype-phenotype correlations. Autism Res. 2009;2:98–108. doi:10.​1002/​aur.​70.PubMedCrossRef
220.
Zurück zum Zitat Hettinger JA, Liu X, Schwartz CE, Michaelis RC, Holden JJA. A DRD1 haplotype is associated with risk for autism spectrum disorders in male-only affected sib-pair families. Am J Med Genet B Neuropsychiatr Genet. 2008;147B:628–36. doi:10.1002/ajmg.b.30655.PubMedCrossRef Hettinger JA, Liu X, Schwartz CE, Michaelis RC, Holden JJA. A DRD1 haplotype is associated with risk for autism spectrum disorders in male-only affected sib-pair families. Am J Med Genet B Neuropsychiatr Genet. 2008;147B:628–36. doi:10.​1002/​ajmg.​b.​30655.PubMedCrossRef
221.
Zurück zum Zitat Higashida H, Yokoyama S, Huang J-J, Liu L, Ma W-J, Akther S, et al. Social memory, amnesia, and autism: brain oxytocin secretion is regulated by NAD+ metabolites and single nucleotide polymorphisms of CD38. Neurochem Int. 2012;61:828–38. doi:10.1016/j.neuint.2012.01.030.PubMedCrossRef Higashida H, Yokoyama S, Huang J-J, Liu L, Ma W-J, Akther S, et al. Social memory, amnesia, and autism: brain oxytocin secretion is regulated by NAD+ metabolites and single nucleotide polymorphisms of CD38. Neurochem Int. 2012;61:828–38. doi:10.​1016/​j.​neuint.​2012.​01.​030.PubMedCrossRef
223.
Zurück zum Zitat Hu VW, Frank BC, Heine S, Lee NH, Quackenbush J. Gene expression profiling of lymphoblastoid cell lines from monozygotic twins discordant in severity of autism reveals differential regulation of neurologically relevant genes. BMC Genomics. 2006;7:118. doi:10.1186/1471-2164-7-118.PubMedPubMedCentralCrossRef Hu VW, Frank BC, Heine S, Lee NH, Quackenbush J. Gene expression profiling of lymphoblastoid cell lines from monozygotic twins discordant in severity of autism reveals differential regulation of neurologically relevant genes. BMC Genomics. 2006;7:118. doi:10.​1186/​1471-2164-7-118.PubMedPubMedCentralCrossRef
224.
Zurück zum Zitat Hu VW, Sarachana T, Kim KS, Nguyen A, Kulkarni S, Steinberg ME, et al. Gene expression profiling differentiates autism case-controls and phenotypic variants of autism spectrum disorders: evidence for circadian rhythm dysfunction in severe autism. Autism Res. 2009;2:78–97. doi:10.1002/aur.73.PubMedPubMedCentralCrossRef Hu VW, Sarachana T, Kim KS, Nguyen A, Kulkarni S, Steinberg ME, et al. Gene expression profiling differentiates autism case-controls and phenotypic variants of autism spectrum disorders: evidence for circadian rhythm dysfunction in severe autism. Autism Res. 2009;2:78–97. doi:10.​1002/​aur.​73.PubMedPubMedCentralCrossRef
226.
Zurück zum Zitat Kistner-Griffin E, Brune CW, Davis LK, Sutcliffe JS, Cox NJ, Cook EH. Parent-of-origin effects of the serotonin transporter gene associated with autism. Am J Med Genet B Neuropsychiatr Genet. 2011;156:139–44. doi:10.1002/ajmg.b.31146.PubMedCrossRef Kistner-Griffin E, Brune CW, Davis LK, Sutcliffe JS, Cox NJ, Cook EH. Parent-of-origin effects of the serotonin transporter gene associated with autism. Am J Med Genet B Neuropsychiatr Genet. 2011;156:139–44. doi:10.​1002/​ajmg.​b.​31146.PubMedCrossRef
227.
229.
231.
Zurück zum Zitat Liu X, Solehdin F, Cohen IL, Gonzalez MG, Jenkins EC, Lewis MES, et al. Population- and family-based studies associate the MTHFR gene with idiopathic autism in simplex families. J Autism Dev Disord. 2011;41:938–44. doi:10.1007/s10803-010-1120-x.PubMedCrossRef Liu X, Solehdin F, Cohen IL, Gonzalez MG, Jenkins EC, Lewis MES, et al. Population- and family-based studies associate the MTHFR gene with idiopathic autism in simplex families. J Autism Dev Disord. 2011;41:938–44. doi:10.​1007/​s10803-010-1120-x.PubMedCrossRef
235.
Zurück zum Zitat Malenfant P, Liu X, Hudson ML, Qiao Y, Hrynchak M, Riendeau N, et al. Association of GTF2i in the Williams-Beuren syndrome critical region with autism spectrum disorders. J Autism Dev Disord. 2012;42:1459–69. doi:10.1007/s10803-011-1389-4.PubMedCrossRef Malenfant P, Liu X, Hudson ML, Qiao Y, Hrynchak M, Riendeau N, et al. Association of GTF2i in the Williams-Beuren syndrome critical region with autism spectrum disorders. J Autism Dev Disord. 2012;42:1459–69. doi:10.​1007/​s10803-011-1389-4.PubMedCrossRef
236.
Zurück zum Zitat Martin CL, Duvall JA, Ilkin Y, Simon JS, Arreaza MG, Wilkes K, et al. Cytogenetic and molecular characterization of A2BP1/FOX1 as a candidate gene for autism. Am J Med Genet B Neuropsychiatr Genet. 2007;144B:869–76. doi:10.1002/ajmg.b.30530.PubMedCrossRef Martin CL, Duvall JA, Ilkin Y, Simon JS, Arreaza MG, Wilkes K, et al. Cytogenetic and molecular characterization of A2BP1/FOX1 as a candidate gene for autism. Am J Med Genet B Neuropsychiatr Genet. 2007;144B:869–76. doi:10.​1002/​ajmg.​b.​30530.PubMedCrossRef
238.
Zurück zum Zitat Maussion G, Carayol J, Lepagnol-Bestel A-M, Tores F, Loe-Mie Y, Milbreta U, et al. Convergent evidence identifying MAP/microtubule affinity-regulating kinase 1 (MARK1) as a susceptibility gene for autism. Hum Mol Genet. 2008;17:2541–51. doi:10.1093/hmg/ddn154.PubMedCrossRef Maussion G, Carayol J, Lepagnol-Bestel A-M, Tores F, Loe-Mie Y, Milbreta U, et al. Convergent evidence identifying MAP/microtubule affinity-regulating kinase 1 (MARK1) as a susceptibility gene for autism. Hum Mol Genet. 2008;17:2541–51. doi:10.​1093/​hmg/​ddn154.PubMedCrossRef
239.
Zurück zum Zitat McCauley JL, Li C, Jiang L, Olson LM, Crockett G, Gainer K, et al. Genome-wide and Ordered-Subset linkage analyses provide support for autism loci on 17q and 19p with evidence of phenotypic and interlocus genetic correlates. BMC Med Genet. 2005;6:1. doi:10.1186/1471-2350-6-1.PubMedPubMedCentralCrossRef McCauley JL, Li C, Jiang L, Olson LM, Crockett G, Gainer K, et al. Genome-wide and Ordered-Subset linkage analyses provide support for autism loci on 17q and 19p with evidence of phenotypic and interlocus genetic correlates. BMC Med Genet. 2005;6:1. doi:10.​1186/​1471-2350-6-1.PubMedPubMedCentralCrossRef
240.
Zurück zum Zitat McInnes LA, Nakamine A, Pilorge M, Brandt T, Jiménez González P, Fallas M, et al. A large-scale survey of the novel 15q24 microdeletion syndrome in autism spectrum disorders identifies an atypical deletion that narrows the critical region. Mol Autism. 2010;1:5. doi:10.1186/2040-2392-1-5.PubMedPubMedCentralCrossRef McInnes LA, Nakamine A, Pilorge M, Brandt T, Jiménez González P, Fallas M, et al. A large-scale survey of the novel 15q24 microdeletion syndrome in autism spectrum disorders identifies an atypical deletion that narrows the critical region. Mol Autism. 2010;1:5. doi:10.​1186/​2040-2392-1-5.PubMedPubMedCentralCrossRef
243.
Zurück zum Zitat Muscarella LA, Guarnieri V, Sacco R, Militerni R, Bravaccio C, Trillo S, et al. HOXA1 gene variants influence head growth rates in humans. Am J Med Genet B Neuropsychiatr Genet. 2007;144B:388–90. doi:10.1002/ajmg.b.30469.PubMedCrossRef Muscarella LA, Guarnieri V, Sacco R, Militerni R, Bravaccio C, Trillo S, et al. HOXA1 gene variants influence head growth rates in humans. Am J Med Genet B Neuropsychiatr Genet. 2007;144B:388–90. doi:10.​1002/​ajmg.​b.​30469.PubMedCrossRef
244.
Zurück zum Zitat Nabi R, Zhong H, Serajee FJ, Huq AHMM. No association between single nucleotide polymorphisms in DLX6 and Piccolo genes at 7q21-q22 and autism. Am J Med Genet B Neuropsychiatr Genet. 2003;119B:98–101. doi:10.1002/ajmg.b.10012.PubMedCrossRef Nabi R, Zhong H, Serajee FJ, Huq AHMM. No association between single nucleotide polymorphisms in DLX6 and Piccolo genes at 7q21-q22 and autism. Am J Med Genet B Neuropsychiatr Genet. 2003;119B:98–101. doi:10.​1002/​ajmg.​b.​10012.PubMedCrossRef
245.
246.
Zurück zum Zitat Nakamura K, Anitha A, Yamada K, Tsujii M, Iwayama Y, Hattori E, et al. Genetic and expression analyses reveal elevated expression of syntaxin 1A (STX1A) in high functioning autism. Int J Neuropsychopharmacol. 2008;11:1073–84. doi:10.1017/S1461145708009036.PubMedCrossRef Nakamura K, Anitha A, Yamada K, Tsujii M, Iwayama Y, Hattori E, et al. Genetic and expression analyses reveal elevated expression of syntaxin 1A (STX1A) in high functioning autism. Int J Neuropsychopharmacol. 2008;11:1073–84. doi:10.​1017/​S146114570800903​6.PubMedCrossRef
247.
Zurück zum Zitat Nicholas B, Rudrasingham V, Nash S, Kirov G, Owen MJ, Wimpory DC. Association of Per1 and Npas2 with autistic disorder: support for the clock genes/social timing hypothesis. Mol Psychiatry. 2007;12:581–92. doi:10.1038/sj.mp.4001953.PubMedCrossRef Nicholas B, Rudrasingham V, Nash S, Kirov G, Owen MJ, Wimpory DC. Association of Per1 and Npas2 with autistic disorder: support for the clock genes/social timing hypothesis. Mol Psychiatry. 2007;12:581–92. doi:10.​1038/​sj.​mp.​4001953.PubMedCrossRef
250.
Zurück zum Zitat Petek E, Schwarzbraun T, Noor A, Patel M, Nakabayashi K, Choufani S, et al. Molecular and genomic studies of IMMP2L and mutation screening in autism and Tourette syndrome. Mol Genet Genomics. 2007;277:71–81. doi:10.1007/s00438-006-0173-1.PubMedCrossRef Petek E, Schwarzbraun T, Noor A, Patel M, Nakabayashi K, Choufani S, et al. Molecular and genomic studies of IMMP2L and mutation screening in autism and Tourette syndrome. Mol Genet Genomics. 2007;277:71–81. doi:10.​1007/​s00438-006-0173-1.PubMedCrossRef
251.
Zurück zum Zitat Philippi A, Roschmann E, Tores F, Lindenbaum P, Benajou A, Germain-Leclerc L, et al. Haplotypes in the gene encoding protein kinase c-beta (PRKCB1) on chromosome 16 are associated with autism. Mol Psychiatry. 2005;10:950–60. doi:10.1038/sj.mp.4001704.PubMedCrossRef Philippi A, Roschmann E, Tores F, Lindenbaum P, Benajou A, Germain-Leclerc L, et al. Haplotypes in the gene encoding protein kinase c-beta (PRKCB1) on chromosome 16 are associated with autism. Mol Psychiatry. 2005;10:950–60. doi:10.​1038/​sj.​mp.​4001704.PubMedCrossRef
252.
Zurück zum Zitat Philippi A, Tores F, Carayol J, Rousseau F, Letexier M, Roschmann E, et al. Association of autism with polymorphisms in the paired-like homeodomain transcription factor 1 (PITX1) on chromosome 5q31: a candidate gene analysis. BMC Med Genet. 2007;8:74. doi:10.1186/1471-2350-8-74.PubMedPubMedCentralCrossRef Philippi A, Tores F, Carayol J, Rousseau F, Letexier M, Roschmann E, et al. Association of autism with polymorphisms in the paired-like homeodomain transcription factor 1 (PITX1) on chromosome 5q31: a candidate gene analysis. BMC Med Genet. 2007;8:74. doi:10.​1186/​1471-2350-8-74.PubMedPubMedCentralCrossRef
254.
256.
Zurück zum Zitat Ramoz N, Cai G, Reichert JG, Silverman JM, Buxbaum JD. An analysis of candidate autism loci on chromosome 2q24-q33: evidence for association to the STK39 gene. Am J Med Genet B Neuropsychiatr Genet. 2008;147B:1152–8. doi:10.1002/ajmg.b.30739.PubMedCrossRef Ramoz N, Cai G, Reichert JG, Silverman JM, Buxbaum JD. An analysis of candidate autism loci on chromosome 2q24-q33: evidence for association to the STK39 gene. Am J Med Genet B Neuropsychiatr Genet. 2008;147B:1152–8. doi:10.​1002/​ajmg.​b.​30739.PubMedCrossRef
258.
Zurück zum Zitat Ramoz N, Cai G, Reichert JG, Corwin TE, Kryzak LA, Smith CJ, et al. Family-based association study of TPH1 and TPH2 polymorphisms in autism. Am J Med Genet B Neuropsychiatr Genet. 2006;141B:861–7. doi:10.1002/ajmg.b.30356.PubMedCrossRef Ramoz N, Cai G, Reichert JG, Corwin TE, Kryzak LA, Smith CJ, et al. Family-based association study of TPH1 and TPH2 polymorphisms in autism. Am J Med Genet B Neuropsychiatr Genet. 2006;141B:861–7. doi:10.​1002/​ajmg.​b.​30356.PubMedCrossRef
262.
Zurück zum Zitat Russo AJ, Neville L, Wroge C. Low Serum Alpha-1 Antitrypsin (AAT) in Family Members of Individuals with Autism Correlates with PiMZ Genotype. Biomark Insights. 2009;4:45–56.PubMedPubMedCentral Russo AJ, Neville L, Wroge C. Low Serum Alpha-1 Antitrypsin (AAT) in Family Members of Individuals with Autism Correlates with PiMZ Genotype. Biomark Insights. 2009;4:45–56.PubMedPubMedCentral
263.
272.
Zurück zum Zitat Silverman JM, Buxbaum JD, Ramoz N, Schmeidler J, Reichenberg A, Hollander E, et al. Autism-related routines and rituals associated with a mitochondrial aspartate/glutamate carrier SLC25A12 polymorphism. Am J Med Genet B Neuropsychiatr Genet. 2008;147:408–10. doi:10.1002/ajmg.b.30614.PubMedCrossRef Silverman JM, Buxbaum JD, Ramoz N, Schmeidler J, Reichenberg A, Hollander E, et al. Autism-related routines and rituals associated with a mitochondrial aspartate/glutamate carrier SLC25A12 polymorphism. Am J Med Genet B Neuropsychiatr Genet. 2008;147:408–10. doi:10.​1002/​ajmg.​b.​30614.PubMedCrossRef
274.
276.
Zurück zum Zitat Sutcliffe JS, Delahanty RJ, Prasad HC, McCauley JL, Han Q, Jiang L, et al. Allelic heterogeneity at the serotonin transporter locus (SLC6A4) confers susceptibility to autism and rigid-compulsive behaviors. Am J Hum Genet. 2005;77:265–79. doi:10.1086/432648.PubMedPubMedCentralCrossRef Sutcliffe JS, Delahanty RJ, Prasad HC, McCauley JL, Han Q, Jiang L, et al. Allelic heterogeneity at the serotonin transporter locus (SLC6A4) confers susceptibility to autism and rigid-compulsive behaviors. Am J Hum Genet. 2005;77:265–79. doi:10.​1086/​432648.PubMedPubMedCentralCrossRef
278.
280.
Zurück zum Zitat Vincent JB, Petek E, Thevarkunnel S, Kolozsvari D, Cheung J, Patel M, et al. The RAY1/ST7 Tumor-Suppressor Locus on Chromosome 7q31 Represents a Complex Multi-transcript System. Genomics. 2002;80:283–94. doi:10.1006/geno.2002.6835.PubMedCrossRef Vincent JB, Petek E, Thevarkunnel S, Kolozsvari D, Cheung J, Patel M, et al. The RAY1/ST7 Tumor-Suppressor Locus on Chromosome 7q31 Represents a Complex Multi-transcript System. Genomics. 2002;80:283–94. doi:10.​1006/​geno.​2002.​6835.PubMedCrossRef
281.
Zurück zum Zitat Vincent JB, Kolozsvari D, Roberts WS, Bolton PF, Gurling HMD, Scherer SW. Mutation screening of X-chromosomal neuroligin genes: no mutations in 196 autism probands. Am J Med Genet B Neuropsychiatr Genet. 2004;129B:82–4. doi:10.1002/ajmg.b.30069.PubMedCrossRef Vincent JB, Kolozsvari D, Roberts WS, Bolton PF, Gurling HMD, Scherer SW. Mutation screening of X-chromosomal neuroligin genes: no mutations in 196 autism probands. Am J Med Genet B Neuropsychiatr Genet. 2004;129B:82–4. doi:10.​1002/​ajmg.​b.​30069.PubMedCrossRef
282.
Zurück zum Zitat Vincent JB, Thevarkunnel S, Kolozsvari D, Paterson AD, Roberts W, Scherer SW. Association and transmission analysis of the FMR1 IVS10 + 14C-T variant in autism. Am J Med Genet B Neuropsychiatr Genet. 2004;125B:54–6. doi:10.1002/ajmg.b.20088.PubMedCrossRef Vincent JB, Thevarkunnel S, Kolozsvari D, Paterson AD, Roberts W, Scherer SW. Association and transmission analysis of the FMR1 IVS10 + 14C-T variant in autism. Am J Med Genet B Neuropsychiatr Genet. 2004;125B:54–6. doi:10.​1002/​ajmg.​b.​20088.PubMedCrossRef
285.
288.
290.
Zurück zum Zitat Ylisaukko-oja T, Alarcón M, Cantor RM, Auranen M, Vanhala R, Kempas E, et al. Search for autism loci by combined analysis of Autism Genetic Resource Exchange and Finnish families. Ann Neurol. 2006;59:145–55. doi:10.1002/ana.20722.PubMedCrossRef Ylisaukko-oja T, Alarcón M, Cantor RM, Auranen M, Vanhala R, Kempas E, et al. Search for autism loci by combined analysis of Autism Genetic Resource Exchange and Finnish families. Ann Neurol. 2006;59:145–55. doi:10.​1002/​ana.​20722.PubMedCrossRef
292.
Zurück zum Zitat Zandi PP, Kalaydjian A, Avramopoulos D, Shao H, Fallin MD, Newschaffer CJ. Rh and ABO maternal-fetal incompatibility and risk of autism. Am J Med Genet B Neuropsychiatr Genet. 2006;141B:643–7. doi:10.1002/ajmg.b.30391.PubMedCrossRef Zandi PP, Kalaydjian A, Avramopoulos D, Shao H, Fallin MD, Newschaffer CJ. Rh and ABO maternal-fetal incompatibility and risk of autism. Am J Med Genet B Neuropsychiatr Genet. 2006;141B:643–7. doi:10.​1002/​ajmg.​b.​30391.PubMedCrossRef
298.
Zurück zum Zitat Ben-David E, Shifman S. Combined analysis of exome sequencing points toward a major role for transcription regulation during brain development in autism. Mol Psychiatry. 2013;18:1054–6. doi:10.1038/mp.2012.148.PubMedCrossRef Ben-David E, Shifman S. Combined analysis of exome sequencing points toward a major role for transcription regulation during brain development in autism. Mol Psychiatry. 2013;18:1054–6. doi:10.​1038/​mp.​2012.​148.PubMedCrossRef
301.
Zurück zum Zitat Celestino-Soper PBS, Shaw CA, Sanders SJ, Li J, Murtha MT, Ercan-Sencicek AG, et al. Use of array CGH to detect exonic copy number variants throughout the genome in autism families detects a novel deletion in TMLHE. Hum Mol Genet. 2011;20:4360–70. doi:10.1093/hmg/ddr363.PubMedPubMedCentralCrossRef Celestino-Soper PBS, Shaw CA, Sanders SJ, Li J, Murtha MT, Ercan-Sencicek AG, et al. Use of array CGH to detect exonic copy number variants throughout the genome in autism families detects a novel deletion in TMLHE. Hum Mol Genet. 2011;20:4360–70. doi:10.​1093/​hmg/​ddr363.PubMedPubMedCentralCrossRef
304.
Zurück zum Zitat Chaste P, Klei L, Sanders SJ, Hus V, Murtha MT, Lowe JK, et al. A genome-wide association study of autism using the Simons Simplex Collection: Does reducing phenotypic heterogeneity in autism increase genetic homogeneity? Biol Psychiatry. 2015;77:775–84. doi:10.1016/j.biopsych.2014.09.017.PubMedCrossRef Chaste P, Klei L, Sanders SJ, Hus V, Murtha MT, Lowe JK, et al. A genome-wide association study of autism using the Simons Simplex Collection: Does reducing phenotypic heterogeneity in autism increase genetic homogeneity? Biol Psychiatry. 2015;77:775–84. doi:10.​1016/​j.​biopsych.​2014.​09.​017.PubMedCrossRef
309.
Zurück zum Zitat Griswold AJ, Dueker ND, Van Booven D, Rantus JA, Jaworski JM, Slifer SH, et al. Targeted massively parallel sequencing of autism spectrum disorder-associated genes in a case control cohort reveals rare loss-of-function risk variants. Mol Autism. 2015;6:43. doi:10.1186/s13229-015-0034-z.PubMedPubMedCentralCrossRef Griswold AJ, Dueker ND, Van Booven D, Rantus JA, Jaworski JM, Slifer SH, et al. Targeted massively parallel sequencing of autism spectrum disorder-associated genes in a case control cohort reveals rare loss-of-function risk variants. Mol Autism. 2015;6:43. doi:10.​1186/​s13229-015-0034-z.PubMedPubMedCentralCrossRef
317.
Zurück zum Zitat Cross-Disorder Group of the Psychiatric Genomics Consortium, Lee SH, Ripke S, Neale BM, Faraone SV, Purcell SM, et al. Genetic relationship between five psychiatric disorders estimated from genome-wide SNPs. Nat Genet. 2013;45:984–94. doi:10.1038/ng.2711.PubMedCentralCrossRef Cross-Disorder Group of the Psychiatric Genomics Consortium, Lee SH, Ripke S, Neale BM, Faraone SV, Purcell SM, et al. Genetic relationship between five psychiatric disorders estimated from genome-wide SNPs. Nat Genet. 2013;45:984–94. doi:10.​1038/​ng.​2711.PubMedCentralCrossRef
322.
Zurück zum Zitat Moreno-De-Luca D, Sanders SJ, Willsey AJ, Mulle JG, Lowe JK, Geschwind DH, et al. Using large clinical data sets to infer pathogenicity for rare copy number variants in autism cohorts. Mol Psychiatry. 2013;18:1090–5. doi:10.1038/mp.2012.138.PubMedCrossRef Moreno-De-Luca D, Sanders SJ, Willsey AJ, Mulle JG, Lowe JK, Geschwind DH, et al. Using large clinical data sets to infer pathogenicity for rare copy number variants in autism cohorts. Mol Psychiatry. 2013;18:1090–5. doi:10.​1038/​mp.​2012.​138.PubMedCrossRef
323.
Zurück zum Zitat Murdoch JD, Gupta AR, Sanders SJ, Walker MF, Keaney J, Fernandez TV, et al. No evidence for association of autism with rare heterozygous point mutations in Contactin-Associated Protein-Like 2 (CNTNAP2), or in Other Contactin-Associated Proteins or Contactins. PLoS Genet. 2015;11:e1004852. doi:10.1371/journal.pgen.1004852.PubMedPubMedCentralCrossRef Murdoch JD, Gupta AR, Sanders SJ, Walker MF, Keaney J, Fernandez TV, et al. No evidence for association of autism with rare heterozygous point mutations in Contactin-Associated Protein-Like 2 (CNTNAP2), or in Other Contactin-Associated Proteins or Contactins. PLoS Genet. 2015;11:e1004852. doi:10.​1371/​journal.​pgen.​1004852.PubMedPubMedCentralCrossRef
326.
Zurück zum Zitat Peters SU, Hundley RJ, Wilson AK, Warren Z, Vehorn A, Carvalho CMB, et al. The behavioral phenotype in MECP2 duplication syndrome: a comparison with idiopathic autism. Autism Res. 2013;6:42–50. doi:10.1002/aur.1262.PubMedCrossRef Peters SU, Hundley RJ, Wilson AK, Warren Z, Vehorn A, Carvalho CMB, et al. The behavioral phenotype in MECP2 duplication syndrome: a comparison with idiopathic autism. Autism Res. 2013;6:42–50. doi:10.​1002/​aur.​1262.PubMedCrossRef
327.
Zurück zum Zitat Sagar A, Bishop JR, Tessman DC, Guter S, Martin CL, Cook EH. Co-occurrence of autism, childhood psychosis, and intellectual disability associated with a de novo 3q29 microdeletion. Am J Med Genet A. 2013;161A:845–9. doi:10.1002/ajmg.a.35754.PubMedCrossRef Sagar A, Bishop JR, Tessman DC, Guter S, Martin CL, Cook EH. Co-occurrence of autism, childhood psychosis, and intellectual disability associated with a de novo 3q29 microdeletion. Am J Med Genet A. 2013;161A:845–9. doi:10.​1002/​ajmg.​a.​35754.PubMedCrossRef
328.
329.
338.
Zurück zum Zitat Anitha A, Nakamura K, Thanseem I, Matsuzaki H, Miyachi T, Tsujii M, et al. Downregulation of the expression of mitochondrial electron transport complex genes in autism brains. Brain Pathol. 2013;23:294–302. doi:10.1111/bpa.12002.PubMedCrossRef Anitha A, Nakamura K, Thanseem I, Matsuzaki H, Miyachi T, Tsujii M, et al. Downregulation of the expression of mitochondrial electron transport complex genes in autism brains. Brain Pathol. 2013;23:294–302. doi:10.​1111/​bpa.​12002.PubMedCrossRef
341.
Zurück zum Zitat Fatemi SH, Folsom TD, Kneeland RE, Yousefi MK, Liesch SB, Thuras PD. Impairment of fragile X mental retardation protein-metabotropic glutamate receptor 5 signaling and its downstream cognates ras-related C3 botulinum toxin substrate 1, amyloid beta A4 precursor protein, striatal-enriched protein tyrosine phosphatase, and homer 1, in autism: a postmortem study in cerebellar vermis and superior frontal cortex. Mol Autism. 2013;4:21. doi:10.1186/2040-2392-4-21.PubMedPubMedCentralCrossRef Fatemi SH, Folsom TD, Kneeland RE, Yousefi MK, Liesch SB, Thuras PD. Impairment of fragile X mental retardation protein-metabotropic glutamate receptor 5 signaling and its downstream cognates ras-related C3 botulinum toxin substrate 1, amyloid beta A4 precursor protein, striatal-enriched protein tyrosine phosphatase, and homer 1, in autism: a postmortem study in cerebellar vermis and superior frontal cortex. Mol Autism. 2013;4:21. doi:10.​1186/​2040-2392-4-21.PubMedPubMedCentralCrossRef
342.
Zurück zum Zitat Fatemi SH, Reutiman TJ, Folsom TD, Rustan OG, Rooney RJ, Thuras PD. Downregulation of GABAA receptor protein subunits α6, β2, δ, ε, γ2, θ, and ρ2 in superior frontal cortex of subjects with autism. J Autism Dev Disord. 2014;44:1833–45. doi:10.1007/s10803-014-2078-x.PubMedCrossRef Fatemi SH, Reutiman TJ, Folsom TD, Rustan OG, Rooney RJ, Thuras PD. Downregulation of GABAA receptor protein subunits α6, β2, δ, ε, γ2, θ, and ρ2 in superior frontal cortex of subjects with autism. J Autism Dev Disord. 2014;44:1833–45. doi:10.​1007/​s10803-014-2078-x.PubMedCrossRef
344.
345.
Zurück zum Zitat Huang H-S, Cheung I, Akbarian S. RPP25 is developmentally regulated in prefrontal cortex and expressed at decreased levels in autism spectrum disorder. Autism Res. 2010;3:153–61. doi:10.1002/aur.141.PubMedCrossRef Huang H-S, Cheung I, Akbarian S. RPP25 is developmentally regulated in prefrontal cortex and expressed at decreased levels in autism spectrum disorder. Autism Res. 2010;3:153–61. doi:10.​1002/​aur.​141.PubMedCrossRef
347.
Zurück zum Zitat Nagarajan RP, Hogart AR, Gwye Y, Martin MR, LaSalle JM. Reduced MeCP2 expression is frequent in autism frontal cortex and correlates with aberrant MECP2 promoter methylation. Epigenetics. 2006;1:e1–11.PubMedPubMedCentralCrossRef Nagarajan RP, Hogart AR, Gwye Y, Martin MR, LaSalle JM. Reduced MeCP2 expression is frequent in autism frontal cortex and correlates with aberrant MECP2 promoter methylation. Epigenetics. 2006;1:e1–11.PubMedPubMedCentralCrossRef
350.
351.
Zurück zum Zitat Stamova B, Ander BP, Barger N, Sharp FR, Schumann CM. Specific Regional and Age-Related Small Noncoding RNA Expression Patterns Within Superior Temporal Gyrus of Typical Human Brains Are Less Distinct in Autism Brains. J Child Neurol. 2015;30:1930–46. doi:10.1177/0883073815602067.PubMedPubMedCentralCrossRef Stamova B, Ander BP, Barger N, Sharp FR, Schumann CM. Specific Regional and Age-Related Small Noncoding RNA Expression Patterns Within Superior Temporal Gyrus of Typical Human Brains Are Less Distinct in Autism Brains. J Child Neurol. 2015;30:1930–46. doi:10.​1177/​0883073815602067​.PubMedPubMedCentralCrossRef
354.
362.
Zurück zum Zitat Stover AE, Herculian S, Banuelos MG, Navarro SL, Jenkins MP, Schwartz PH. Culturing human pluripotent and neural stem cells in an enclosed cell culture system for basic and preclinical research. J Vis Exp. 2016. doi:10.3791/53685.PubMedPubMedCentral Stover AE, Herculian S, Banuelos MG, Navarro SL, Jenkins MP, Schwartz PH. Culturing human pluripotent and neural stem cells in an enclosed cell culture system for basic and preclinical research. J Vis Exp. 2016. doi:10.​3791/​53685.PubMedPubMedCentral
Metadaten
Titel
Bio-collections in autism research
verfasst von
Jamie Reilly
Louise Gallagher
June L. Chen
Geraldine Leader
Sanbing Shen
Publikationsdatum
01.12.2017
Verlag
BioMed Central
Erschienen in
Molecular Autism / Ausgabe 1/2017
Elektronische ISSN: 2040-2392
DOI
https://doi.org/10.1186/s13229-017-0154-8

Weitere Artikel der Ausgabe 1/2017

Molecular Autism 1/2017 Zur Ausgabe

Neu in den Fachgebieten Neurologie und Psychiatrie