Skip to main content
Erschienen in: Molecular Autism 1/2011

Open Access 01.12.2011 | Research

A noise-reduction GWAS analysis implicates altered regulation of neurite outgrowth and guidance in autism

verfasst von: John P Hussman, Ren-Hua Chung, Anthony J Griswold, James M Jaworski, Daria Salyakina, Deqiong Ma, Ioanna Konidari, Patrice L Whitehead, Jeffery M Vance, Eden R Martin, Michael L Cuccaro, John R Gilbert, Jonathan L Haines, Margaret A Pericak-Vance

Erschienen in: Molecular Autism | Ausgabe 1/2011

Abstract

Background

Genome-wide Association Studies (GWAS) have proved invaluable for the identification of disease susceptibility genes. However, the prioritization of candidate genes and regions for follow-up studies often proves difficult due to false-positive associations caused by statistical noise and multiple-testing. In order to address this issue, we propose the novel GWAS noise reduction (GWAS-NR) method as a way to increase the power to detect true associations in GWAS, particularly in complex diseases such as autism.

Methods

GWAS-NR utilizes a linear filter to identify genomic regions demonstrating correlation among association signals in multiple datasets. We used computer simulations to assess the ability of GWAS-NR to detect association against the commonly used joint analysis and Fisher's methods. Furthermore, we applied GWAS-NR to a family-based autism GWAS of 597 families and a second existing autism GWAS of 696 families from the Autism Genetic Resource Exchange (AGRE) to arrive at a compendium of autism candidate genes. These genes were manually annotated and classified by a literature review and functional grouping in order to reveal biological pathways which might contribute to autism aetiology.

Results

Computer simulations indicate that GWAS-NR achieves a significantly higher classification rate for true positive association signals than either the joint analysis or Fisher's methods and that it can also achieve this when there is imperfect marker overlap across datasets or when the closest disease-related polymorphism is not directly typed. In two autism datasets, GWAS-NR analysis resulted in 1535 significant linkage disequilibrium (LD) blocks overlapping 431 unique reference sequencing (RefSeq) genes. Moreover, we identified the nearest RefSeq gene to the non-gene overlapping LD blocks, producing a final candidate set of 860 genes. Functional categorization of these implicated genes indicates that a significant proportion of them cooperate in a coherent pathway that regulates the directional protrusion of axons and dendrites to their appropriate synaptic targets.

Conclusions

As statistical noise is likely to particularly affect studies of complex disorders, where genetic heterogeneity or interaction between genes may confound the ability to detect association, GWAS-NR offers a powerful method for prioritizing regions for follow-up studies. Applying this method to autism datasets, GWAS-NR analysis indicates that a large subset of genes involved in the outgrowth and guidance of axons and dendrites is implicated in the aetiology of autism.
Begleitmaterial
Hinweise

Electronic supplementary material

The online version of this article (doi:10.​1186/​2040-2392-2-1) contains supplementary material, which is available to authorized users.

Competing interests

The authors declare that they have no competing interests.

Authors' contributions

All co-authors contributed to writing the manuscript. JPH was the primary author of the manuscript, developed the statistical methods and the design for their implementation and contributed pathway analysis of candidate genes. RHC contributed to the development of the statistical methods and study design and also conducted statistical analyses. AJG contributed to molecular analysis and interpretation. JMJ, DS and DM conducted statistical analyses. IK and PLW performed molecular analysis and interpretation. JMV performed molecular analysis and contributed to the study design. ERM provided input into study design, methods development, statistical analyses, and interpretation of findings. MLC analysed clinical data and contributed to the study design. JRG performed molecular analysis, interpreted data and contributed to the study design. JLH provided input to study design and statistical analyses. MPV contributed to the design of the study, development of methods, coordination of statistical and molecular analysis and interpretation of data. She is also the primary investigator on the parent study.
Abkürzungen
ADI-R
Autism Diagnostic Interview - Revised
AGRE
Autism Genetic Resource Exchange
APL
association in the presence of linkage
AUC
area under the curve
CNV
copy number variation
DAVID
Database for Annotation, Visualization and Discovery
GTP
guanosine triphosphate
LD
linkage disequilibrium
GWAS
Genome-wide association studies
NR
noise reduction
RefSeq
Reference Sequence
ROC
receiver operating characteristic
SNP
single nucleotide polymorphism
TPM
truncated product method.

Background

Genome-wide association studies (GWAS) have provided a powerful tool for identifying disease susceptibility genes. However, analysis of GWAS data has been focused on single-point tests, such as the traditional allele-based chi-squared test or the Cochran-Armitage Trend test [1], which proceed by testing each single nucleotide polymorphism (SNP) independently. As it is likely that the disease variants have not been directly genotyped in a GWAS, tests that account for multiple flanking SNPs in linkage disequilibrium (LD) with the disease variants may increase the power to detect association [2].
Several approaches have been proposed in order to test for association based on multiple markers, which include the haplotype-based approach [35] and the multivariate approach [6, 7]. Akey et al. [8] used analytical approaches to demonstrate that multilocus haplotype tests can be more powerful than single-marker tests. For the multivariate approach, tests such as Hotelling's T2 test are often used to account for multiple markers jointly [6, 9]. Although statistical power can be increased by such multi-marker approaches, it is not a straightforward operation to select markers for testing. Including all markers in a gene or region may not be feasible since it greatly increases the degrees of freedom in the test, which can reduce the power.
Follow-up studies, such as fine mapping and sequencing, are necessary in order to validate association signals and they are also challenging [2]. Prioritization of genes or regions for follow-up studies is often decided by a threshold of P-values or ranking for significant markers [10, 11]. However, many false positives can still exist in the markers classified as significant for follow-up as a result of statistical noise and genome-wide multiple testing. Joint and/or meta-analysis of GWAS data can achieve greater power if these data or P-values are available from different datasets. If P-values from individual and joint analyses are available, it is possible to further increase the power by assigning more weight to markers with replicated association signals in several datasets or to markers that have flanking markers with an association signal.
We propose the use of the GWAS noise reduction (GWAS-NR) approach which uses P-values from individual analyses, as well as joint analysis of multiple datasets, and which accounts for association signals from surrounding markers in LD. GWAS-NR is a novel approach to extending the power of GWAS studies to detect association. Noise reduction is achieved by applying a linear filter within a sliding window in order to identify genomic regions demonstrating correlated profiles of association across multiple datasets. As noise reduction (NR) techniques are widely used to boost signal identification in applications such as speech recognition, data transmission and image enhancement, we expect that GWAS-NR may complement other GWAS analysis methods in identifying candidate loci that may then be prioritized for follow-up analysis or analysed in the context of biological pathways.
Enhancing statistical power is particularly important in the study of complex diseases such as autism. There is overwhelming evidence from twin and family studies for a strong genetic component to autism, with estimates of heritability greater than 80% [1214]. Autism is generally diagnosed before the age of 4, based on marked qualitative differences in social and communication skills, often accompanied by unusual patterns of behaviour (for example, repetitive, restricted, stereotyped) [15]. Altered sensitivity to sensory stimuli and difficulties of motor initiation and coordination also are frequently present. Identifying the underlying genes and characterizing the molecular mechanisms of autism will provide immensely useful guidance in the development of effective clinical interventions.
Numerous autism candidate genes have been reported based on association evidence, expression analysis, copy number variation (CNV), and cytogenetic screening. These genes involve processes including cell adhesion (NLGN3, NLGN4 [16], NRXN1 [17], CDH9/CDH10 [18, 19]), axon guidance (SEMA5A [20]), synaptic scaffolding (SHANK2, DLGAP2 [21], SHANK3 [22]), phosphatidylinositol signalling (PTEN [23], PIK3CG [24]), cytoskeletal regulation (TSC1/TSC2 [24, 25], EPAC2/RAPGEF4 [26], SYNGAP1 [21]), transcriptional regulation (MECP2 [27], EN2 [28]) and excitatory/inhibitory balance (GRIN2A [29], GABRA4, GABRB1 [30]). However, aside from rare mutations and 'syndromic' autism secondary to known genetic disorders, the identification of specific genetic mechanisms in autism has remained elusive.
Over the past decade, the vast majority of genetic studies of autism (both linkage and focused candidate gene studies) have failed to broadly replicate suspected genetic variations. For this reason, the assumption that autism is governed by strong and pervasive genetic variations has given way to the view that autism may involve numerous genetic variants, each having a small effect size at the population level. This may arise from common variations having small individual effects in a large number of individuals (the common disease-common variant [CDCV] hypothesis) or rare variations having large individual effects in smaller subsets of individuals (the rare variant [RV] hypothesis).
Given the potential genetic heterogeneity among individuals with autism and the likely involvement of numerous genes of small effect at the population level, we expected that the GWAS-NR could improve the power to identify candidate genes for follow-up analysis. We applied GWAS-NR to autism GWAS data from multiple sources and conducted simulation studies in order to compare the performance of GWAS-NR with traditional joint and meta-analysis approaches. These data demonstrate that GWAS-NR is a useful tool for prioritizing regions for follow-up studies such as next-generation sequencing.

Methods

GWAS-NR

The GWAS-NR algorithm produces a set of weighted P-values for use in prioritizing genomic regions for follow-up study. Roeder and Wasserman [31] characterize the statistical properties of such weighting approaches in GWAS, observing that informative weights can improve power substantially, while the loss in power is usually small even if the weights are uninformative. The GWAS-NR algorithm computes a weight at each locus based on the strength and correlation of association signals at surrounding markers and in multiple datasets, without relying on prior information or scientific hypotheses. The weights are applied to the P-values derived from joint analysis of the complete data and the resulting weighted P-values are then used to prioritize regions for follow-up analysis.
Noise reduction methods are frequently applied when extracting a common signal from multiple sensors. The filter used by GWAS-NR is similar to the method proposed by de Cheveigné and Simon [32] for sensor noise suppression in magneto- and electro-encephalograph recordings. Each sensor is projected onto the other sensors and the fitted values from these regressions are used in place of the original values. The fitted values of such regressions retain sources of interest that are common to multiple sensors. As the regression residuals are orthogonal to the fitted values, uncorrelated components are suppressed.
In a genomic context, the 'sensors' take the form of probit-transformed P-values derived from independent datasets, as well as P-values derived from joint analysis of the full dataset. The filter inherently highlights cross-validating associations, by preserving signals that jointly occur in a given genomic region and attenuating spikes that are not correlated across subsets of the data. However, GWAS-NR can achieve no advantage over simple joint analysis when an association signal is restricted to a single marker and flanking markers provide no supplementary information.
We estimate ordinary least-squares regressions of the form
Z i j = α j k + β j k Z i k + v j k
and compute projections
Z i j ^ = α j k + β j k Z i k
where Z i and Z ik are the probits Φ-1(1 - p) of the P-values at locus i in two datasets j and k. Φ-1(⋅) denotes the inverse of the cumulative standard normal distribution. The estimates are computed within a centred sliding window of w markers and β jk are constrained to be nonnegative which sets Z i j ^ to the mean Z i j ¯ in regions having zero or negative correlation across sensors. As β jk is driven by the covariance between probits in datasets j and k, probits that demonstrate positive local correlation will tend to be preserved, while probits demonstrating weak local correlation will be attenuated. One local regression is computed for each locus and is used to compute a single fitted value Z i j ^ for that locus. The same method is used to compute projections Z i k ^ .
In order to capture association signals at adjacent loci in different datasets without estimating numerous parameters, the regressor at each locus is taken to be the probit of the lowest P-value among that locus and its two immediate neighbours. Quality control (QC) failure or different genotyping platforms can cause SNP genotypes to be missing in different datasets. Missing genotypes for a locus having no immediately flanking neighbours are assigned a probit of zero. The window width w is calculated as w = 2h + 1, where h is the lag at which the autocorrelation of the probits declines below a pre-defined threshold. In practice, we estimate the autocorrelation profile for each series of probits and use the average value of h with an autocorrelation threshold of 0.20.
After computing the projections of Z j and Z k , the resulting values are converted back to P-values and a set of filtered P-values is computed from these projections using Fisher's method. The same algorithm is executed again, this time using the probits of the filtered P-values and the P-values obtained from the joint association analysis of the complete data. The resulting Fisher P-values are then treated as weighting factors and are multiplied by the corresponding raw P-values from the joint analysis, producing a set of weighted P-values. To aid interpretation, we apply a monotonic transformation to these weighted P-values, placing them between 0 and 1 by fitting parameters of an extreme value distribution. The GWAS-NR algorithm was executed as a Matlab script.

Simulations

Although noise reduction has been shown to be useful in other biomedical applications [32], understanding its properties for identifying the true positives in disease association studies is also important. We used computer simulations to compare the performance of GWAS-NR with the joint association in the presence of linkage (APL) analysis and Fisher's method under a variety of disease models. We used genomeSIMLA [33] to simulate LD structures based on the Affymetrix 5.0 chip and performed the sliding-window haplotype APL [34] test to measure association. Detailed descriptions for the simulation settings are provided in Additional File 1 and detailed haplotype configurations can be found in Additional File 2.
An important goal for the proposed approach is to help prioritize candidate regions for follow-up studies such as next-generation sequencing. Top regions or genes ranked by their P-values are often considered priority regions for follow-up studies. In order to investigate the proportion of true positives that occur in the top regions, we treated the association tests as binary classifiers. The markers were ranked by their P-values and markers that occurred in the top k ranking were classified as significant, where k was pre-specified as a cut-off threshold. The markers that were not in the top k ranking were classified as non-significant. We then compared the sensitivity and specificity of GWAS-NR with the joint and Fisher's tests. The sensitivity was calculated based on the proportion of the three markers associated with the disease that were correctly classified as significant. The specificity was calculated based on the proportion of markers not associated with the disease that were correctly classified as non-significant. The sensitivity and specificity were averaged over 1000 replicates.

Ascertainment and sample description

We ascertained autism patients and their affected and unaffected family members through the Hussman Institute for Human Genomics (HIHG, University of Miami Miller School of Medicine, FL, USA), and the Vanderbilt Center for Human Genetics Research (CHGR, Vanderbilt University Medical Center, Tennessee, USA; UM/VU). Participating families were enrolled through a multi-site study of autism genetics and recruited via support groups, advertisements and clinical and educational settings. All participants and families were ascertained using a standard protocol. These protocols were approved by appropriate Institutional Review Boards. Written informed consent was obtained from parents, as well as from minors who were able to give informed consent; in individuals unable to give assent due to age or developmental problems, assent was obtained whenever possible.
The core inclusion criteria were as follows: (1) chronological age between 3 and 21 years of age; (2) presumptive clinical diagnosis of autism; and (3) expert clinical determination of autism diagnosis using Diagnostic and Statistical Manual of Mental Disorders (DSM)-IV criteria supported by the Autism Diagnostic Interview-Revised (ADI-R) in the majority of cases and all available clinical information. The ADI-R is a semi-structured diagnostic interview which provides diagnostic algorithms for classification of autism [35]. All ADI-R interviews were conducted by formally trained interviewers who have achieved reliability according to established methods. Thirty-eight individuals did not have an ADI-R and, for those cases, we implemented a best-estimate procedure to determine a final diagnosis using all available information from the research record and data from other assessment procedures. This information was reviewed by a clinical panel led by an experienced clinical psychologist and included two other psychologists and a paediatric medical geneticist - all of whom were experienced in autism. Following a review of case material, the panel discussed the case until a consensus diagnosis was obtained. Only those cases in which a consensus diagnosis of autism was reached were included. (4) The final criterion was a minimal developmental level of 18 months as determined by the Vineland Adaptive Behavior Scale (VABS) [36] or the VABS-II [37] or intelligence quotient equivalent >35. These minimal developmental levels assure that ADI-R results are valid and reduce the likelihood of including individuals with severe mental retardation only. We excluded participants with severe sensory problems (for example, visual impairment or hearing loss), significant motor impairments (for example, failure to sit by 12 months or walk by 24 months) or identified metabolic, genetic or progressive neurological disorders.
A total of 597 Caucasian families (707 individuals with autism) were genotyped at HIHG. This dataset consisted of 99 multiplex families (more than one affected individual) and 498 singleton (parent-child trio) families. A subset of these data had been previously reported [19]. In addition, GWAS data were obtained from the Autism Genetic Resource Exchange (AGRE) [35] as an additional dataset for analysis. The full AGRE dataset is publicly available and contains families with the full spectrum of autism spectrum disorders. From AGRE, we selected only families with one or more individuals diagnosed with autism (using DSM-IV and ADI-R); affected individuals with non-autism diagnosis within these families were excluded from the analysis. This resulted in a dataset of 696 multiplex families (1240 individuals with autism) from AGRE [35].

Genotyping and quality control and population stratification

We extracted DNA for individuals from whole blood by using Puregene chemistry (QIAGEN, MD, USA). We performed genotyping using the Illumina Beadstation and the Illumina Infinium Human 1 M beadchip following the recommended protocol, only with a more stringent GenCall score threshold of 0.25. Genotyping efficiency was greater than 99%, and quality assurance was achieved by the inclusion of one CEPH control per 96-well plate that was genotyped multiple times. Technicians were blinded to affection status and quality-control samples. The AGRE data were genotyped using the Illumina HumanHap550 BeadChip with over 550,000 SNP markers. All samples and SNPs underwent stringent GWAS quality control measures as previously described in detail in Ma et al. [19].
Although population substructure does not cause a type I error in family-based association tests, multiple founder effects could result in a reduced power to detect an association in a heterogeneous disease such as autism. Thus, we conducted EIGENSTRAT [38] analysis on all parents from analysed families for evidence of population substructure using the overlapping SNPs genotyped in both the UM/VU and AGRE datasets. In order to ensure the most homogeneous groups for association screening and replication, we excluded all families with outliers defined by EIGENSTRAT [38] out of four standard deviations of principal components 1 and 2.

Haplotype block definition

We used haplotype blocks to define regions of interest. Significant regions can be used for follow-up analysis such as next-generation sequencing. We applied the haplotype block definition method proposed by Gabriel et al. [39] to the UM/VU dataset. We performed GWAS-NR based on single-marker APL P-values from UM/VU, AGRE and joint tests. We also performed GWAS-NR on P-values obtained from sliding-window haplotype tests with a haplotype length of three markers for the UM/VU, AGRE and joint datasets. Since the true haplotype length is not known, we chose a fixed length of three markers across the genome and used GWAS-NR to sort out true signals from the P-values. Blocks containing the top 5000 markers, as ranked by the minimum values (MIN_NR) of the GWAS-NR P-values obtained from single-marker tests, and the GWAS-NR P-values obtained from tests of three-marker haplotypes, were selected for further analysis.

Combined P-values for haplotype block scoring

In order to test for the significance of the haplotype blocks, we calculated the combined P-value for each block using a modified version of the Truncated Product Method (TPM) [40]. TPM has been shown to have correct type I error rates and more power than other methods combining P-values [40] under different simulation models. Briefly, a combined score was calculated from the markers in each block, based on the product of MIN_NR that were below a threshold of 0.05. We used the Monte Carlo algorithm [40] with a slight modification to test the significance of the combined score. Specifically, a correlation matrix was applied to account for correlation among P-values for the markers in the same block. The null hypothesis is that none of the markers in the haplotype block are associated with the disease. In order to simulate the null distribution for the combined score, we generated two correlated sets of L uniform numbers based on the correlation of 0.67 for CAPL and HAPL P-values, where L denotes the number of tests in the block. The minimum values were selected from each pair in the two sets, which resulted in a vector of L minimum values. Then the correlation matrix was applied to the vector of L minimum values and a null combined GWAS-NR score was calculated for the haplotype block.

Functional analysis

In order to investigate functional relationships among genes in the candidate set, each candidate was manually annotated and cross-referenced, based on a review of current literature, with attention to common functions, directly interacting proteins and binding domains. Supplementary functional annotations were obtained using DAVID (The Database for Annotation, Visualization and Integrated Discovery) version 6.7 [4143].

Results

Simulations

We present the simulation results for the three-marker haplotype disease models in Figures 1 and 2. Figure 1 presents receiver operating characteristic (ROC) curves to show the sensitivity and specificity of GWAS-NR, the joint APL analysis and Fisher's tests, based on varying cut-off values of ranking for significance. The Fisher's test to combine P-values was used here as a standard meta-analysis approach. The performance of a classification model can be judged based on the area under the ROC curve (AUC). For scenario 1 (identical marker coverage in each dataset), GWAS-NR produced a greater AUC than the joint and Fisher's tests. It can also be observed from the figure that, given the same specificity, GWAS-NR achieved a higher sensitivity for classifying true positives as significant as the joint and Fisher's tests.
As independent datasets may have an imperfect overlap of markers, which is true of the UM/VU and AGRE autism data, and the omission of the closest disease-related polymorphism from the data can have substantial negative impact on the power of GWAS [44], we also compared the performance of GWAS-NR with the joint APL tests and Fisher's tests under a range of missing marker scenarios: 20% of the simulated markers in one dataset were randomly omitted for the recessive and multiplicative models and 50% of the simulated markers were randomly omitted in one dataset for the dominant and additive models. This performance is shown in Figure 2. Again, the GWAS-NR produced a greater AUC than the joint and Fisher's tests and achieved a higher sensitivity for classifying true positives at each level of specificity.
The results for the two-marker haplotype disease models are shown in Additional File 3. The same pattern is also observed in Additional File 3 that GWAS-NR produced greater AUC than the joint and Fisher's tests.
We also evaluated the type I error rates of the modified TPM for identifying significant LD blocks using a truncation threshold of 0.05. For the scenario assuming full marker coverage as described in Additional File 1, the modified TPM had type I error rates of 0.035 and 0.004 at the significance levels of 0.05 and 0.01, respectively. For the missing-marker scenario, the type I error rates for the modified TPM were 0.046 and 0.007 at the significance levels of 0.05 and 0.01, respectively.

Autism GWAS-NR results

We applied the GWAS-NR in autism data using UM/VU, AGRE and the joint (UM/VU)/AGRE datasets. A flow diagram (Additional File 4) for the data analysis process is found in the supplemental data. The selection of haplotype blocks based on Gabriel's definition resulted in a total of 2680 blocks based on the top 5000 markers. Moreover, 141 markers out of the 5000 markers which are not in any blocks were also selected. Blocks of LD were scored based on the truncated product of P-values below a threshold of 0.05 and a P-value for each block was obtained through Monte Carlo simulation. The P-values for 141 markers not in any blocks were also calculated using the Monte Carlo algorithm to account for the minimum statistics. All of the 141 markers had P-values less than 0.05 and were selected. 725 LD blocks achieved a significance threshold of P < = 0.01, and an additional 810 blocks achieved a threshold of P < = 0.05. A complete list of these blocks is presented in Additional File 5.
In order to determine what genes reside within the 1535 significant LD blocks, we used the University of California Santa Cruz (UCSC) Genome Browser Table Browser. The 1535 regions were converted into start and end positions based on the SNP positions in the March 2006 (NCBI36/hg18) human genome assembly. These start and end positions were used to define regions in the UCSC Table Browser. We searched each region for overlap with the RefSeq annotation track in the UCSC Browser. This search resulted in 431 unique genes which mapped back to 646 significant LD blocks and 50 single markers. These genes are presented in Additional File 6. For the remaining 839 LD blocks that did not overlap a RefSeq gene, we identified the nearest RefSeq gene using Galaxy [45]. The distance to these nearest genes averaged 417,377 bp with a range from 5296 to 5,547,466 bp. These nearest genes include candidate genes for which strong proximal associations with autism have previously been reported, such as CDH9 [18, 19] and SEMA5A [20]. We considered these genes for follow-up because GWAS-NR, by construction, may capture association information from nearby regions that may not be in strict LD with a given SNP and because these proximal locations may also incorporate regulatory elements. These genes are presented in Additional File 7. Combining these sets resulted in a candidate set of 860 unique genes (presented in Additional File 8). For genes assigned to more than one significant LD block, the lowest P-value among these blocks is used for sorting and discussion purposes.
The most significant LD block we identified is located at 2p24.1 (ch2 204444539-20446116; P = 1.8E-06) proximal to PUM2. One LD block located within the PUM2 exon also had nominally significant association (P = 0.024). Additional top-ranking candidates, in order of significance, include CACNA1I (P = 1.8E-05), EDEM1 (P = 1.8E-05), DNER (P = 2.7E-05), A2BP1 (P = 3.6E-05), ZNF622 (P = 8.11E-05), SEMA4D (P = 9.09E-05) and CDH8 (P = 9.09E-05). Gene ontology classifications and InterPro binding domains reported by DAVID [4143] to be most enriched in the candidate gene set are presented in Tables 1 and 2, respectively, providing a broad functional characterization of the candidate genes identified by the GWAS-NR in autism.
Table 1
Common functions of autism candidate genes identified by genome-wide association studies-noise reduction (GWAS-NR)
Gene ontology (GO) term
No. of genes
GO term identification
P-value1
Examples
Cell adhesion
76
0007155
6.29E-13
CDH8, NCAM2
Biological adhesion
76
0022610
6.64E-13
CDH2, CTNNB1
Cell-cell adhesion
35
0016337
6.24E-08
CTNNA2, AMIGO2
Homophilic cell adhesion
21
0007156
1.21E-06
PTPRM, FAT1
Cell motion
44
0006928
6.65E-06
SEMA5A, FYN
Neuron differentiation
41
0030182
1.14E-05
EN2, NRXN1
Enzyme linked receptor protein signalling pathway
33
0007167
5.40E-05
NCK2, FGFR2
Neuron development
32
0048666
1.07E-04
ROBO2, RTN4R
Negative regulation of gene expression
42
0010629
1.27E-04
SIX3, CUX2
Axonogenesis
22
0007409
1.31E-04
SEMA6A, SLITRK5
Cell morphogenesis involved in differentiation
25
0000904
2.16E-04
PRKCA, PTK2
Cell motility
29
0048870
2.40E-04
DNER, PPAP2B
Localization of cell
29
0051674
2.40E-04
PTEN, NRP2
Negative regulation of transcription
38
0016481
3.19E-04
RBPJ, MEIS2
Cell morphogenesis involved in neuron differentiation
22
0048667
3.94E-04
PARD3, KALRN
Transmembrane receptor protein tyrosine kinase signalling
23
0007169
3.98E-04
SOCS2, DOK5
Neuron projection development
25
0031175
4.40E-04
RTN4R, NGF
Neuron projection morphogenesis
22
0048812
5.07E-04
PVRL1, CDH4
Regulation of cell projection organization
13
0031344
5.33E-04
SEMA4D, CDC42EP4
Negative regulation of nucleobase, nucleoside, nucleotide, and nucleic acid metabolic process
40
0045934
6.79E-04
BCL6, ZHX2
Table 2
Common binding domains of autism candidate genes identified by genome-wide association studies-noise reduction (GWAS-NR).
INTERPRO term
No. of genes
INTERPRO identification
P-value1
Immunoglobulin I-set
20
IPR013098
8.97E-06
Cadherin
16
IPR002126
6.98E-05
Cadherin cytoplasmic region
7
IPR000233
1.14E-04
Pleckstrin homology
26
IPR001849
5.03E-04
Immunoglobulin
21
IPR013151
5.61E-04
Immunoglobulin subtype 2
21
IPR003598
6.77E-04
Fibronectin, type III-like fold
19
IPR008957
1.19E-03
Fibronectin, type III
19
IPR003961
1.72E-03
Epidermal growth factor (EGF)
14
IPR006209
3.71E-03
Meprin/A5-protein/PTPmu (MAM)
5
IPR000998
6.78E-03
Protein-tyrosine phosphatase, receptor/non-receptor type
7
IPR000242
7.36E-03
Pleckstrin homology-type
24
IPR001993
7.41E-03
von Willebrand factor, type A
10
IPR002035
7.41E-03
Immunoglobulin-like
35
IPR007110
7.57E-03
Cell adhesion represented the most common functional annotation reported for the candidate gene set, with a second set of common functional annotations relating to neuronal morphogenesis and motility, including axonogenesis and neuron projection development. Given the enrichment scores reported by DAVID [4143] implicating neurite development and motility, and because numerous cell adhesion molecules are known to regulate axonal and dendritic projections [46, 47], we examined the known functional roles of the individual candidate genes responsible for these enrichment scores. A total of 183 candidate genes were represented among the top 20 functional classifications reported by DAVID [4143]. Based on annotations manually curated from a review of current literature, we observed that 76 (41.5%) of these genes have established roles in the regulation of neurite outgrowth and guidance. These include 39 (51.3%) of the candidate genes contained in the cell adhesion, biological adhesion, cell-cell adhesion and homophilic cell adhesion pathways.
Gene ontology [48] specifically associates two pathways with the narrow synonym 'neurite outgrowth': the neuron projection development (pathway 0031175); and the transmembrane receptor protein tyrosine kinase activity (pathway 0004714). To further test for functional enrichment of genes related to neurite outgrowth, we formed a restricted composite of these two pathways. Enrichment analysis using the EASE function of DAVID [4143] rejected the hypothesis that this composite pathway is randomly associated with the autism candidate set (P = 2.07E-05).
Although many of the candidate genes identified by the GWAS-NR remain uncharacterized or have no known neurological function, we identified 125 genes within the full candidate set having established and interconnected roles in the regulation of neurite outgrowth and guidance. These genes are involved in diverse sub-processes including cell adhesion, axon guidance, phosphatidylinositol signalling, establishment of cell polarity, Rho-GTPase signalling, cytoskeletal regulation and transcription. Table 3 presents a summary of these genes by functional category. Additional File 9 presents annotations for these 125 candidates. Additional File 10 presents 104 additional candidates which have suggestive roles in neurite regulation based on putative biological function or homology to known neurite regulators but where we did not find evidence specific to neurite outgrowth and guidance in the current literature.
Table 3
Autism candidate genes with known roles in neurite outgrowth and guidance.
Function
Candidate gene (by lowest P-value)
Cadherin-catenin function
CDH8, CDH2, CDH11, CTNNB1, CTNNA2, PKP4, CTNND2, CDH4, CTNND1, CTNNA3
Cell adhesion
NCAM2, CNTN3, OPCML, ODZ4, NID1, CNTN5, F3, PVRL1, PTPRG, PARVA, FLRT2, ODZ2, NRXN1, ITGA9, ELMO1, FUT9, AMIGO2, KIRREL3, CNTNAP2, NTM
Ion channel
CACNA1I, CACNA1G
Axon guidance
SEMA4D, RTN4R, ROBO2, SEMA5A, PLXDC2, SLITRK5, SEMA6A, RGMA, UNC5D, ALCAM, NTNG2, RTN4RL1, PLXNC1, NRP2
Vesicle transport
STX2, STX16, STXBP5, SYT6
Post-synaptic scaffold
DLGAP2, MAGI1, MAGI2
Signal transduction
DNER, SPRY4, FRK, PRKCA, DOK6, PDE3A, FER, IRS2, SOCS2, SPRY2, FRS3, DOK5, FYN, LZTS1, PTPRD, FGFR2, NRG3, PPP2R2B ALK, RYR2, PALM2-AKAP2, MAP3K7, NTRK3, NGF, PPM1H, GDNF, CXCR4, PTK2, NEDD9, PTPN1, LEPR
Phosphatidylinositol signalling
PLA2G6, PIK3C2B, PTEN, PLA2G4A
Cell polarity
FAT1, PARD3, PARD6G, DCHS2
Rho-GTPase signalling
NCK2, DOCK1, PREX1, CDC42EP4, RND3, RGNEF, DOCK8, CIT, SRGAP3, KALRN, IQGAP2
Cytoskeletal regulation
SGK1, MYLK, GPR56, APBB1IP, PTPRM, WIPF3, PTPRT, MAP3K8, MICAL2, DGKG, COBL, CALD1
Transcription
PUM2, A2BP1, NKX6-1, SOX14, EN2, EBF1, MAP3K1, FOXG1, NFIC, BCL11A
Outside of functions relating to neuritogenesis, the most significant functional annotation reported by DAVID for the candidate gene set relates to transmission of nerve impulses (p = 9.02E-04). We identified 40 genes in the candidate set related to neurotransmission (synaptogenesis, neuronal excitability, synaptic plasticity, and vesicle exocytosis) which did not have overlapping roles in neurite regulation. Table 4 presents a summary of these genes by functional category.
Table 4
Autism candidate genes with roles in synaptic function.
Function
Candidate gene (by lowest P-value)
Synaptogenesis
LRRTM4, SYN3
Excitatory/inhibitory balance
KCNIP1, KCNQ1, KCNQ5, KCNJ4, SLC6A13, IQCF1, GABBR2, GRIK4, OAT, KCNN3, GRM3, GCOM1, CACNA2D1, GRM7, ADRB2, KCNH7, KCNIP4, GRIK2, CACNG2, KCNMA1, KCNG1
Synaptic plasticity
RIMS1, PTGER2, SLC24A2, NETO1, PTGS2
Vesicle exocytosis
PTPRN2, AMPH, RAB11B, SYNPR
Other
TPH2, CHRNA9, RIMBP2, ATXN1, CHRNB4, NOVA1, SNCAIP, CHRM3
In order to investigate how the GWAS-NR results compared with the joint APL tests and Fisher's tests, we examined the lists of top 5000 markers selected based on GWAS-NR, joint APL test and Fisher's test P-values. A total of 3328 of the markers are overlapped between the lists for the GWAS-NR and joint APL tests, while 1951 of the markers are overlapped between the lists for the GWAS-NR and Fisher's tests. Thus, GWAS-NR had a higher concordance with the joint APL tests than the Fisher's tests. The results suggested that Fisher's test may have the lowest sensitivity to identify the true positives, which is consistent with our simulation results. Moreover, 120 markers that are not overlapped between Illumina Infinium Human 1M beadchip and Illumina HumanHap550 BeadChip were among the top 5000 markers selected based on GWAS-NR. Some of the 120 markers are in the significant genes identified by haplotype blocks such as PUM2, A2BP1, DNER and SEMA4D.
In order to similarly investigate the overlap of candidate genes indentified by GWAS-NR and joint APL tests, we repeated the haplotype block scoring method with the top 5000 markers as identified by joint APL: this analysis resulted in 1924 significant LD blocks. Of these, 1257 overlapped with the blocks selected by GWAS-NR analysis. Identification of the RefSeq genes within with these 1257 shared regions showed that 380 potential candidate genes were shared by the two methods. In addition, GWAS-NR analysis produced 53 non-overlapping genes while the joint APL analysis produced 349 non-overlapping genes.
As GWAS-NR amplifies association signals that are replicated in multiple flanking markers and across data sets, the method can be expected to produce a reduced list of higher confidence candidate regions for follow-up, compared with standard single-locus methods. At the same time, GWAS-NR does not generate a large number of significant candidates in regions that would otherwise be ranked as insignificant. While it is not possible to exclude a role in autism for the 349 additional candidate genes produced by the joint APL analysis, it is notable that among the top 20 gene ontology pathways reported by DAVID [4143] for this set of genes, not one is specific to neuronal function (data not shown). This analysis highlights the utility of GWAS-NR to narrow and prioritize follow-up gene lists.

Discussion

We propose the use of GWAS-NR, a noise-reduction method for genome-wide association studies which aims to enhance the power to detect true positive associations for follow-up analysis. Our results demonstrate that GWAS-NR is a powerful method for the enhancement of the detection of genetic associations. Simulation evidence using a variety of disease models indicates that, when markers are ranked by P-values and candidates are selected based on a threshold rank, GWAS-NR achieves higher classification rates than the use of joint P-values or Fisher's method. In simulated data, the GWAS-NR also achieves strong performance when there is imperfect marker overlap across datasets and when the closest disease-related polymorphism is not typed. As Müller-Myhsok and Abel have observed, when less-than-maximum LD exists between a disease locus and the closest biallelic marker, the required sample size to achieve a given level of power may increase dramatically, particularly if there is a substantial difference in allele frequencies at the disease marker and the analysed marker [49].
In the context of allelic association, noise can be viewed as observed but random association evidence (for example, false positives) that is not the result of true LD with a susceptibility or causative variant. Such noise is likely to confound studies of complex disorders, where genetic heterogeneity among affected individuals or complex interactions among multiple genes may result in modest association signals that are difficult to detect. The influence of positive noise components is also likely to contribute to the so-called 'winner's curse' phenomenon, whereby the estimated effect of a putatively associated marker is often exaggerated in the initial findings, compared with estimated effects in follow-up studies [50]. GWAS-NR appears to be a promising approach to address these challenges.
By amplifying signals in regions where association evidence is locally correlated across datasets, the GWAS-NR captures information that may be omitted or underutilized in single-marker analysis. However, the GWAS-NR can achieve no advantage over simple joint analysis when flanking markers provide no supplementary information. This is likely to be true when a true risk locus is typed directly and a single-marker association method is used or when a true risk haplotype is typed directly and the number of markers examined in a haplotype-based analysis is of the same length.
Joint analysis generally has more power than individual tests due to the increase of sample size. Therefore, GWAS-NR, which uses P-values from individual analyses as well as joint analysis of multiple datasets, is expected to have more power than individual tests. However, if there are subpopulations in the sample and the association is specific to a subpopulation, joint analysis may not be as powerful as an individual test for the subpopulation with the association signal. If samples from multiple populations are analysed jointly, test results for individual datasets should also be carefully examined with the GWAS-NR results.
It is common for linear filters to include a large set of estimated parameters to capture cross-correlations in the data at multiple leads and lags. However, in a genomic context, the potentially uneven spacing of markers and varying strength of linkage disequilibrium between markers encouraged us to apply a parsimonious representation that would be robust to data structure. We expect that a larger, well-regularized parameterization may enhance the performance of the noise filter, particularly if the filter is constructed to adapt to varying linkage disequilibrium across the genome. This is a subject of further research.
Our simulation results indicate that applying the modified TPM to select LD blocks based on GWAS-NR can have conservative type I error rates. The original TPM reported by Zaykin et al. [40] produced the expected level of type I error, as a known correlation matrix was used in the simulations to account for correlation among P-values due to LD among markers. However, the true correlation is unknown in real datasets. Accordingly, we estimated correlations in our simulations and analysis by bootstrapping replicates of samples, as well as using the sample correlation between P-values obtained though single marker APL and sliding window haplotype analysis. It is possible that the use of estimated correlations may introduce extra variations in the Monte-Carlo simulations of TPM, which may contribute to conservative type I error rates. As we have demonstrated that GWAS-NR achieves higher sensitivity at each level of specificity, the resulting regions with top rankings can be expected to be enriched for true associations when such associations are actually present in the data, even if the LD block selection procedure is conservative. Overall, the simulation results suggest that GWAS-NR can be expected to produce a condensed set of higher confidence follow-up regions, and that this prioritization strategy can control the number of false positives at or below the expected number in analysis.

Autism

Our data identify potential candidate genes for autism that encode a large subset of proteins involved in the outgrowth and guidance of axons and dendrites to their appropriate synaptic targets. Our results also suggest secondary involvement of genes involved in synaptogenesis and neurotransmission which further contribute to the assembly and function of neural circuitry. Taken together, these findings augment existing genetic, epigenetic and neuropathological evidence suggestive of altered neurite morphology, cell migration, synaptogenesis and excitatory-inhibitory balance in autism [49].
Altered dendritic structure is among the most consistent neuroanatomical findings in autism [51, 52] and several other neurodevelopmental syndromes including Down, Rett and fragile-X [53, 54]. Recent neuroanatomical findings include evidence of subcortical, periventricular, hippocampal and cerebellar heterotopia [55] and altered microarchitecture of cortical minicolumns [56], suggestive of dysregulated neuronal migration and guidance. In recent years, evidence from neuroanatomical and neuroimaging studies has led a number of researchers to propose models of altered cortical networks in autism, emphasizing the possible disruption of long-range connectivity and a developmental bias toward the formation of short-range connections [57, 58].
Neurite regulation is a common function of numerous top-ranking candidates. PUM2 codes for pumilio homolog 2, which regulates dendritic outgrowth, arborization, spine formation and filopodial extension of developing and mature neurons [59]. DNER regulates the morphogenesis of cerebellar Purkinje cells [60] and acts as an inhibitor to retinoic-acid induced neurite outgrowth [61]. A2BP1 binds with ATXN2 (SCA2), a dosage-sensitive regulator of actin filament formation that is suggested to mediate the loss of cytoskeleton-dependent dendritic structure [62]. SEMA4D induces axonal growth cone collapse [63] and promotes dendritic branching and complexity in later stages of development [64, 65]. CDH8 regulates hippocampal mossy fibre axon fasciculation and targeting, complementing N-cadherin (CDH2) in the assembly of synaptic circuits [66].
Neurite outgrowth and guidance can be conceptualized as a process whereby extracellular signals are transduced to cytoplasmic signalling molecules which, in turn, regulate membrane protrusion and neuronal growth cone navigation by reorganizing the architecture of the neuronal cytoskeleton. In general, neurite extension is dependent on microtubule organization, while the extension and retraction of finger-like filopodia and web-like lamellipodia from the neuronal growth cone is dependent on actin dynamics. Gordon-Weeks [67] and Bagnard [68] provide excellent overviews relating to growth cone regulation and axon guidance. Figure 3 provides a simplified overview of some of these molecular interactions.
The autism gene candidates identified by GWAS-NR show functional enrichment in processes, including adhesion, cell motility, axonogenesis, cell morphogenesis and neuron projection development. Notably, a recent analysis of rare CNVs in autism by the Autism Genome Project Consortium indicates similar functional enrichment in the processes of neuronal projection, motility, proliferation, and Rho/Ras GTPase signalling [21].
We propose that, in autism, these processes are not distinct functional classifications but instead cooperate as interacting parts of a coherent molecular pathway regulating the outgrowth and guidance of axons and dendrites. Consistent with this view, the candidate set is enriched for numerous binding domains commonly found in proteins that govern neuritogenesis. These include immunoglobulin, cadherin, pleckstrin homology, MAM, fibronectin type-III and protein tyrosine phosphatase (PTP) domains [6971].
The cytoskeletal dynamics of extending neurites are largely governed by the activity of Rho-GTPases, which act as molecular switches to induce actin remodelling. Molecular evidence suggests that disassociation of catenin from cadherin promotes the activation of Rho-family GTPases Rac and Cdc42, resulting in cytoskeletal rearrangement [72]. Guanine nucleotide exchange factors (GEFs) such as DOCK1 [73] and KALRN [74] activate Rho-GTPases by exchanging bound guanosine diphosphate (GDP) for guanosine triphosphate (GTP), while GTPase activating proteins (GAPs) such as SRGAP3 [75] increase the rate of intrinsic GTP hydrolysis to inactivate GTPases. Pleckstrin homology domains, characteristic of several GEFs and GAPs, bind to phosphoinositides to establish membrane localization and also may play a signalling role in GTPase function [76]. Certain GTPases outside of the Rho family, particularly Rap and Ras, also exert an influence on cytoskeletal dynamics and neurite differentiation [77, 76].
Several genes in the candidate set with established roles in neurite formation and guidance have been previously implicated in autism. These include A2BP1 (P = 3.60E-05), ROBO2 (2.00E-03), SEMA5A (2.30E-03), EN2 (4.00E-03), CACNA1G (6.00E-03), PTEN (8.00E-03), NRXN1 (1.10E-02), FUT9 (1.80E-02), DOCK8 (2.10E-02), NRP2 (2.60E-02) and CNTNAP2 (2.70E-02). Other previously reported autism candidate genes with suggestive roles in neurite regulation include PCDH9 (1.76E-03), CDH9 (6.00E-03) and CSMD3 (2.10E-02).
The enriched presence of transcription factors in the candidate set is intriguing, as many of these candidates, including CUX2, SIX3, MEIS2 and ZFHX1B have established roles in the specification of GABAergic cortical interneurons [76]. Many guidance mechanisms in the neuritogenic pathway, such as Slit-Robo, semaphorin-neuropilin, and CXCR4 signalling also direct the migration and regional patterning of interneurons during development. Proper targeting of interneurons is vital to the organization of cortical circuitry, including minicolumnar architecture which is reported to be altered in autism [78]. Thus, the functional roles of the candidate genes we identify may embrace additional forms of neuronal motility and targeting.

Conclusions

We proposed a noise-reduction methodology, GWAS-NR, to enhance the ability to detect associations in GWAS data. By amplifying signals in regions where association evidence is locally correlated across datasets, the GWAS-NR captures information that may be omitted or underutilized in single-marker analysis. Simulation evidence demonstrates that under a variety of disease models, GWAS-NR achieves higher classification rates for true positive associations, compared with the use of joint p-values or Fisher's method.
The GWAS-NR method was applied to autism data, with the objective of prioritizing regions of association for follow-up analysis. Gene set analysis was conducted in order to examine if the identified autism candidate genes were over-represented in any biological pathway relative to the background genes. The significance of a given pathway suggests that the pathway may be associated with autism due to the enrichment of autism candidate genes in that pathway. We find that many of the implicated genes cooperate within a coherent molecular mechanism. This neuritogenic pathway regulates the transduction of membrane-associated signals to downstream cytoskeletal effectors that induce the directional protrusion of axons and dendrites. This mechanism provides a framework that embraces numerous genetic findings in autism to date, and is consistent with neuroanatomical evidence. While confirmation of this pathway will require additional evidence such as the identification of functional variants, our results suggest that autistic pathology may be mediated by the dynamic regulation of the neuronal cytoskeleton, with resulting alterations in dendritic and axonal connectivity.

Acknowledgements

We thank the patients with autism and their family members who participated in this study and personnel at the John P Hussman Institute for Human Genomics. We also wish to gratefully acknowledge the resources provided by the Autism Genetic Resource Exchange (AGRE) consortium and the participating AGRE families. The AGRE is supported by Autism Speaks. The authors gratefully acknowledge Drs Harry H Wright and Ruth K Abramson for their contribution of patient samples and clinical data to this study. The authors also thank Katherine Savage for graphic design and Dongmin Kim for programming support. This research was supported by grants from the National Institutes of Health (NIH) (NS26630 and MH080647), the Medical Research Council's (MRC) Autism Genome Project (AGP), and by a gift from the Hussman Foundation. A subset of the participants was ascertained while Dr Pericak-Vance was a faculty member at Duke University.
Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://​creativecommons.​org/​licenses/​by/​2.​0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Competing interests

The authors declare that they have no competing interests.

Authors' contributions

All co-authors contributed to writing the manuscript. JPH was the primary author of the manuscript, developed the statistical methods and the design for their implementation and contributed pathway analysis of candidate genes. RHC contributed to the development of the statistical methods and study design and also conducted statistical analyses. AJG contributed to molecular analysis and interpretation. JMJ, DS and DM conducted statistical analyses. IK and PLW performed molecular analysis and interpretation. JMV performed molecular analysis and contributed to the study design. ERM provided input into study design, methods development, statistical analyses, and interpretation of findings. MLC analysed clinical data and contributed to the study design. JRG performed molecular analysis, interpreted data and contributed to the study design. JLH provided input to study design and statistical analyses. MPV contributed to the design of the study, development of methods, coordination of statistical and molecular analysis and interpretation of data. She is also the primary investigator on the parent study.
Anhänge

Electronic supplementary material

Authors’ original submitted files for images

Literatur
1.
Zurück zum Zitat Armitage P: Test for linear trends in proportions and frequencies. Biometrics. 1955, 11: 375-386. 10.2307/3001775.CrossRef Armitage P: Test for linear trends in proportions and frequencies. Biometrics. 1955, 11: 375-386. 10.2307/3001775.CrossRef
2.
Zurück zum Zitat McCarthy MI, Abecasis GR, Cardon LR, Goldstein DB, Little J, Ioannidis JP, Hirschhorn JN: Genome-wide association studies for complex traits: consensus, uncertainty and challenges. Nat Rev Genet. 2008, 9: 356-369. 10.1038/nrg2344.CrossRefPubMed McCarthy MI, Abecasis GR, Cardon LR, Goldstein DB, Little J, Ioannidis JP, Hirschhorn JN: Genome-wide association studies for complex traits: consensus, uncertainty and challenges. Nat Rev Genet. 2008, 9: 356-369. 10.1038/nrg2344.CrossRefPubMed
3.
Zurück zum Zitat Zaykin DV, Westfall PH, Young SS, Karnoub MA, Wagner MJ, Ehm MG: Testing association of statistically inferred haplotypes with discrete and continuous traits in samples of unrelated individuals. Hum Hered. 2002, 53: 79-91. 10.1159/000057986.CrossRefPubMed Zaykin DV, Westfall PH, Young SS, Karnoub MA, Wagner MJ, Ehm MG: Testing association of statistically inferred haplotypes with discrete and continuous traits in samples of unrelated individuals. Hum Hered. 2002, 53: 79-91. 10.1159/000057986.CrossRefPubMed
4.
Zurück zum Zitat Tzeng JY, Wang CH, Kao JT, Hsiao CK: Regression-based association analysis with clustered haplotypes through use of genotypes. Am J Hum Genet. 2006, 78: 231-242. 10.1086/500025.PubMedCentralCrossRefPubMed Tzeng JY, Wang CH, Kao JT, Hsiao CK: Regression-based association analysis with clustered haplotypes through use of genotypes. Am J Hum Genet. 2006, 78: 231-242. 10.1086/500025.PubMedCentralCrossRefPubMed
5.
Zurück zum Zitat Sha Q, Chen HS, Zhang S: A new association test using haplotype similarity. Genet Epidemiol. 2007, 31: 577-593. 10.1002/gepi.20230.CrossRefPubMed Sha Q, Chen HS, Zhang S: A new association test using haplotype similarity. Genet Epidemiol. 2007, 31: 577-593. 10.1002/gepi.20230.CrossRefPubMed
7.
Zurück zum Zitat Rakovski CS, Xu X, Lazarus R, Blacker D, Laird NM: A new multimarker test for family-based association studies. Genet Epidemiol. 2007, 31: 9-17. 10.1002/gepi.20186.CrossRefPubMed Rakovski CS, Xu X, Lazarus R, Blacker D, Laird NM: A new multimarker test for family-based association studies. Genet Epidemiol. 2007, 31: 9-17. 10.1002/gepi.20186.CrossRefPubMed
8.
Zurück zum Zitat Akey J, Jin L, Xiong M: Haplotypes vs single marker linkage disequilibrium tests: what do we gain?. Eur J Hum Genet. 2001, 9: 291-300. 10.1038/sj.ejhg.5200619.CrossRefPubMed Akey J, Jin L, Xiong M: Haplotypes vs single marker linkage disequilibrium tests: what do we gain?. Eur J Hum Genet. 2001, 9: 291-300. 10.1038/sj.ejhg.5200619.CrossRefPubMed
9.
10.
Zurück zum Zitat Grant SF, Qu HQ, Bradfield JP, Marchand L, Kim CE, Glessner JT, Grabs R, Taback SP, Frackelton EC, Eckert AW, DCCT/EDIC Research Group: Follow-up analysis of genome-wide association data identifies novel loci for type 1 diabetes. Diabetes. 2009, 58: 290-295. 10.2337/db08-1022.PubMedCentralCrossRefPubMed Grant SF, Qu HQ, Bradfield JP, Marchand L, Kim CE, Glessner JT, Grabs R, Taback SP, Frackelton EC, Eckert AW, DCCT/EDIC Research Group: Follow-up analysis of genome-wide association data identifies novel loci for type 1 diabetes. Diabetes. 2009, 58: 290-295. 10.2337/db08-1022.PubMedCentralCrossRefPubMed
11.
Zurück zum Zitat International Multiple Sclerosis Genetics Consortium (IMSGC): Comprehensive follow-up of the first genome-wide association study of multiple sclerosis identifies KIF21B and TMEM39A as susceptibility loci. Hum Mol Genet. 2010, 19: 953-962. 10.1093/hmg/ddp542.CrossRef International Multiple Sclerosis Genetics Consortium (IMSGC): Comprehensive follow-up of the first genome-wide association study of multiple sclerosis identifies KIF21B and TMEM39A as susceptibility loci. Hum Mol Genet. 2010, 19: 953-962. 10.1093/hmg/ddp542.CrossRef
12.
Zurück zum Zitat Steffenburg S, Gillberg C, Hellgren L, Andersson L, Gillberg IC, Jakobsson G, Bohman M: A twin study of autism in Denmark, Finland, Iceland, Norway, and Sweden. J Child Psychol Psychiatry. 1989, 30: 405-416. 10.1111/j.1469-7610.1989.tb00254.x.CrossRefPubMed Steffenburg S, Gillberg C, Hellgren L, Andersson L, Gillberg IC, Jakobsson G, Bohman M: A twin study of autism in Denmark, Finland, Iceland, Norway, and Sweden. J Child Psychol Psychiatry. 1989, 30: 405-416. 10.1111/j.1469-7610.1989.tb00254.x.CrossRefPubMed
13.
Zurück zum Zitat Bailey A, Le Couteur A, Gottesman I, Bolton P, Simonoff E, Yuzda E, Rutter M: Autism as a strongly genetic disorder: evidence from a British twin study. Psychol Med. 1995, 25: 63-77. 10.1017/S0033291700028099.CrossRefPubMed Bailey A, Le Couteur A, Gottesman I, Bolton P, Simonoff E, Yuzda E, Rutter M: Autism as a strongly genetic disorder: evidence from a British twin study. Psychol Med. 1995, 25: 63-77. 10.1017/S0033291700028099.CrossRefPubMed
14.
Zurück zum Zitat Bolton P, Macdonald H, Pickles A, Rios P, Goode S, Crowson M, Bailey A, Rutter M: A case-control family history study of autism. J Child Psychol Psychiatry Allied Disciplines. 1994, 35: 877-900. 10.1111/j.1469-7610.1994.tb02300.x.CrossRef Bolton P, Macdonald H, Pickles A, Rios P, Goode S, Crowson M, Bailey A, Rutter M: A case-control family history study of autism. J Child Psychol Psychiatry Allied Disciplines. 1994, 35: 877-900. 10.1111/j.1469-7610.1994.tb02300.x.CrossRef
15.
Zurück zum Zitat American Psychiatric Association: Diagnostic and Statistical Manual of Mental Disorders (DSM-IV), Text Revision. 2000, Washington, DC: American Psychiatric Press American Psychiatric Association: Diagnostic and Statistical Manual of Mental Disorders (DSM-IV), Text Revision. 2000, Washington, DC: American Psychiatric Press
16.
Zurück zum Zitat Jamain S, Quach H, Betancur C, Rastam M, Colineaux C, Gillberg IC, Soderstrom H, Giros B, Leboyer M, Gillberg C, Autism Research International Sibpair Study: Mutations of the X-linked genes encoding neuroligins NLGN3 and NLGN4 are associated with autism. Nat Genet. 2003, 34: 27-29. 10.1038/ng1136.PubMedCentralCrossRefPubMed Jamain S, Quach H, Betancur C, Rastam M, Colineaux C, Gillberg IC, Soderstrom H, Giros B, Leboyer M, Gillberg C, Autism Research International Sibpair Study: Mutations of the X-linked genes encoding neuroligins NLGN3 and NLGN4 are associated with autism. Nat Genet. 2003, 34: 27-29. 10.1038/ng1136.PubMedCentralCrossRefPubMed
17.
Zurück zum Zitat Marshall CR, Noor A, Vincent JB, Lionel AC, Feuk L, Skaug J, Shago M, Moessner R, Pinto D, Ren Y: Structural variation of chromosomes in autism spectrum disorder. Am J Hum Genet. 2008, 82: 477-488. 10.1016/j.ajhg.2007.12.009.PubMedCentralCrossRefPubMed Marshall CR, Noor A, Vincent JB, Lionel AC, Feuk L, Skaug J, Shago M, Moessner R, Pinto D, Ren Y: Structural variation of chromosomes in autism spectrum disorder. Am J Hum Genet. 2008, 82: 477-488. 10.1016/j.ajhg.2007.12.009.PubMedCentralCrossRefPubMed
18.
Zurück zum Zitat Wang K, Zhang H, Ma D, Bucan M, Glessner JT, Abrahams BS, Salyakina D, Imielinski M, Bradfield JP, Sleiman PM: Common genetic variants on 5p14.1 associate with autism spectrum disorders. Nature. 2009, 459: 528-533. 10.1038/nature07999.PubMedCentralCrossRefPubMed Wang K, Zhang H, Ma D, Bucan M, Glessner JT, Abrahams BS, Salyakina D, Imielinski M, Bradfield JP, Sleiman PM: Common genetic variants on 5p14.1 associate with autism spectrum disorders. Nature. 2009, 459: 528-533. 10.1038/nature07999.PubMedCentralCrossRefPubMed
19.
Zurück zum Zitat Ma D, Salyakina D, Jaworski JM, Konidari I, Whitehead PL, Andersen AN, Hoffman JD, Slifer SH, Hedges DJ, Cukier HN: A genome-wide association study of autism reveals a common novel risk locus at 5p14.1. Ann Hum Genet. 2009, 73: 263-273. 10.1111/j.1469-1809.2009.00523.x.PubMedCentralCrossRefPubMed Ma D, Salyakina D, Jaworski JM, Konidari I, Whitehead PL, Andersen AN, Hoffman JD, Slifer SH, Hedges DJ, Cukier HN: A genome-wide association study of autism reveals a common novel risk locus at 5p14.1. Ann Hum Genet. 2009, 73: 263-273. 10.1111/j.1469-1809.2009.00523.x.PubMedCentralCrossRefPubMed
20.
Zurück zum Zitat Weiss LA, Arking DE, Gene Discovery Project of Johns Hopkins & the Autism Consortium, Daly MJ, Chakravarti A: A genome-wide linkage and association scan reveals novel loci for autism. Nature. 2009, 461: 802-808. 10.1038/nature08490.PubMedCentralCrossRefPubMed Weiss LA, Arking DE, Gene Discovery Project of Johns Hopkins & the Autism Consortium, Daly MJ, Chakravarti A: A genome-wide linkage and association scan reveals novel loci for autism. Nature. 2009, 461: 802-808. 10.1038/nature08490.PubMedCentralCrossRefPubMed
21.
Zurück zum Zitat Pinto D, Pagnamenta AT, Klei L, Anney R, Merico D, Regan R, Conroy J, Magalhaes TR, Correia C, Abrahams BS: Functional impact of global rare copy number variation in autism spectrum disorders. Nature 2010. 2010, 466 (7304): 368-372. 10.1038/nature09146. Pinto D, Pagnamenta AT, Klei L, Anney R, Merico D, Regan R, Conroy J, Magalhaes TR, Correia C, Abrahams BS: Functional impact of global rare copy number variation in autism spectrum disorders. Nature 2010. 2010, 466 (7304): 368-372. 10.1038/nature09146.
22.
Zurück zum Zitat Durand CM, Betancur C, Boeckers TM, Bockmann J, Chaste P, Fauchereau F, Nygren G, Rastam M, Gillberg IC, Anckarsater H: Mutations in the gene encoding the synaptic scaffolding protein SHANK3 are associated with autism spectrum disorders. Nat Genet. 2007, 39: 25-27. 10.1038/ng1933.PubMedCentralCrossRefPubMed Durand CM, Betancur C, Boeckers TM, Bockmann J, Chaste P, Fauchereau F, Nygren G, Rastam M, Gillberg IC, Anckarsater H: Mutations in the gene encoding the synaptic scaffolding protein SHANK3 are associated with autism spectrum disorders. Nat Genet. 2007, 39: 25-27. 10.1038/ng1933.PubMedCentralCrossRefPubMed
23.
Zurück zum Zitat Goffin A, Hoefsloot LH, Bosgoed E, Swillen A, Fryns JP: PTEN mutation in a family with Cowden syndrome and autism. Am J Med Genet. 2001, 105: 521-524. 10.1002/ajmg.1477.CrossRefPubMed Goffin A, Hoefsloot LH, Bosgoed E, Swillen A, Fryns JP: PTEN mutation in a family with Cowden syndrome and autism. Am J Med Genet. 2001, 105: 521-524. 10.1002/ajmg.1477.CrossRefPubMed
24.
Zurück zum Zitat Serajee FJ, Nabi R, Zhong H, Mahbubul Huq AH: Association of INPP1, PIK3CG, and TSC2 gene variants with autistic disorder: implications for phosphatidylinositol signalling in autism. J Med Genet. 2003, 40: e119-10.1136/jmg.40.11.e119.PubMedCentralCrossRefPubMed Serajee FJ, Nabi R, Zhong H, Mahbubul Huq AH: Association of INPP1, PIK3CG, and TSC2 gene variants with autistic disorder: implications for phosphatidylinositol signalling in autism. J Med Genet. 2003, 40: e119-10.1136/jmg.40.11.e119.PubMedCentralCrossRefPubMed
25.
Zurück zum Zitat Folstein SE, Rosen-Sheidley B: Genetics of autism: complex aetiology for a heterogeneous disorder. Nat Rev Genet. 2001, 2: 943-955. 10.1038/35103559.CrossRefPubMed Folstein SE, Rosen-Sheidley B: Genetics of autism: complex aetiology for a heterogeneous disorder. Nat Rev Genet. 2001, 2: 943-955. 10.1038/35103559.CrossRefPubMed
26.
Zurück zum Zitat Bacchelli E, Blasi F, Biondolillo M, Lamb JA, Bonora E, Barnby G, Parr J, Beyer KS, Klauck SM, Poustka A, International Molecular Genetic Study of Autism Consortium (IMGSAC): Screening of nine candidate genes for autism on chromosome 2q reveals rare nonsynonymous variants in the cAMP-GEFII gene. Mol Psychiatry. 2003, 8: 916-924. 10.1038/sj.mp.4001340.CrossRefPubMed Bacchelli E, Blasi F, Biondolillo M, Lamb JA, Bonora E, Barnby G, Parr J, Beyer KS, Klauck SM, Poustka A, International Molecular Genetic Study of Autism Consortium (IMGSAC): Screening of nine candidate genes for autism on chromosome 2q reveals rare nonsynonymous variants in the cAMP-GEFII gene. Mol Psychiatry. 2003, 8: 916-924. 10.1038/sj.mp.4001340.CrossRefPubMed
27.
Zurück zum Zitat Abuhatzira L, Shemer R, Razin A: MeCP2 involvement in the regulation of neuronal alpha-tubulin production. Hum Mol Genet. 2009, 18: 1415-1423. 10.1093/hmg/ddp048.CrossRefPubMed Abuhatzira L, Shemer R, Razin A: MeCP2 involvement in the regulation of neuronal alpha-tubulin production. Hum Mol Genet. 2009, 18: 1415-1423. 10.1093/hmg/ddp048.CrossRefPubMed
28.
Zurück zum Zitat Gharani N, Benayed R, Mancuso V, Brzustowicz LM, Millonig JH: Association of the homeobox transcription factor, ENGRAILED 2, 3, with autism spectrum disorder. Mol Psychiatry. 2004, 9: 474-484. 10.1038/sj.mp.4001498.CrossRefPubMed Gharani N, Benayed R, Mancuso V, Brzustowicz LM, Millonig JH: Association of the homeobox transcription factor, ENGRAILED 2, 3, with autism spectrum disorder. Mol Psychiatry. 2004, 9: 474-484. 10.1038/sj.mp.4001498.CrossRefPubMed
29.
Zurück zum Zitat Barnby G, Abbott A, Sykes N, Morris A, Weeks DE, Mott R, Lamb J, Bailey AJ, Monaco AP: Candidate-gene screening and association analysis at the autism-susceptibility locus on chromosome 16p: evidence of association at GRIN2A and ABAT. Am J Hum Genet. 2005, 76: 950-966. 10.1086/430454.PubMedCentralCrossRefPubMed Barnby G, Abbott A, Sykes N, Morris A, Weeks DE, Mott R, Lamb J, Bailey AJ, Monaco AP: Candidate-gene screening and association analysis at the autism-susceptibility locus on chromosome 16p: evidence of association at GRIN2A and ABAT. Am J Hum Genet. 2005, 76: 950-966. 10.1086/430454.PubMedCentralCrossRefPubMed
30.
Zurück zum Zitat Collins AL, Ma D, Whitehead PL, Martin ER, Wright HH, Abramson RK, Hussman JP, Haines JL, Cuccaro ML, Gilbert JR, Pericak-Vance MA: Investigation of autism and GABA receptor subunit genes in multiple ethnic groups. Neurogenetics. 2006, 7: 167-174. 10.1007/s10048-006-0045-1.PubMedCentralCrossRefPubMed Collins AL, Ma D, Whitehead PL, Martin ER, Wright HH, Abramson RK, Hussman JP, Haines JL, Cuccaro ML, Gilbert JR, Pericak-Vance MA: Investigation of autism and GABA receptor subunit genes in multiple ethnic groups. Neurogenetics. 2006, 7: 167-174. 10.1007/s10048-006-0045-1.PubMedCentralCrossRefPubMed
31.
33.
Zurück zum Zitat Edwards TL, Bush WS, Turner SD, Dudek SM, Torstenson ES, Schmidt M, Martin E, Ritchie MD: Generating Linkage Disequilibrium Patterns in Data Simulations using genomeSIMLA. Lect Notes Comput Sci. 2008, 4973: 24-35. full_text.CrossRef Edwards TL, Bush WS, Turner SD, Dudek SM, Torstenson ES, Schmidt M, Martin E, Ritchie MD: Generating Linkage Disequilibrium Patterns in Data Simulations using genomeSIMLA. Lect Notes Comput Sci. 2008, 4973: 24-35. full_text.CrossRef
34.
Zurück zum Zitat Chung RH, Hauser ER, Martin ER: The APL test: extension to general nuclear families and haplotypes and the examination of its robustness. Hum Hered. 2006, 61: 189-199. 10.1159/000094774.CrossRefPubMed Chung RH, Hauser ER, Martin ER: The APL test: extension to general nuclear families and haplotypes and the examination of its robustness. Hum Hered. 2006, 61: 189-199. 10.1159/000094774.CrossRefPubMed
36.
Zurück zum Zitat Sparrow SS, Balla D, Cicchetti D: Vineland Adaptive Behavior Scales. 1984, MN: AGS Sparrow SS, Balla D, Cicchetti D: Vineland Adaptive Behavior Scales. 1984, MN: AGS
37.
Zurück zum Zitat Sparrow SS, Cicchetti DV, Balla D: Vineland Adaptive Behavior Scales. 2005, MN: AGS, 2 Sparrow SS, Cicchetti DV, Balla D: Vineland Adaptive Behavior Scales. 2005, MN: AGS, 2
39.
Zurück zum Zitat Gabriel SB, Schaffner SF, Nguyen H, Moore JM, Roy J, Blumenstiel B, Higgins J, DeFelice M, Lochner A, Faggart M: The structure of haplotype blocks in the human genome. Science. 2002, 296: 2225-2229. 10.1126/science.1069424.CrossRefPubMed Gabriel SB, Schaffner SF, Nguyen H, Moore JM, Roy J, Blumenstiel B, Higgins J, DeFelice M, Lochner A, Faggart M: The structure of haplotype blocks in the human genome. Science. 2002, 296: 2225-2229. 10.1126/science.1069424.CrossRefPubMed
40.
Zurück zum Zitat Zaykin DV, Zhivotovsky LA, Westfall PH, Weir BS: Truncated product method for combining P-values. Genet Epidemiol. 2002, 22: 170-185. 10.1002/gepi.0042.CrossRefPubMed Zaykin DV, Zhivotovsky LA, Westfall PH, Weir BS: Truncated product method for combining P-values. Genet Epidemiol. 2002, 22: 170-185. 10.1002/gepi.0042.CrossRefPubMed
41.
Zurück zum Zitat Huang da W, Sherman BT, Lempicki RA: Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat Protoc. 2009, 4: 44-57. 10.1038/nprot.2008.211.CrossRefPubMed Huang da W, Sherman BT, Lempicki RA: Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat Protoc. 2009, 4: 44-57. 10.1038/nprot.2008.211.CrossRefPubMed
42.
Zurück zum Zitat Dennis G, Sherman BT, Hosack DA, Yang J, Gao W, Lane HC, Lempicki RA: DAVID: Database for Annotation, Visualization, and Integrated Discovery. Genome Biol. 2003, 4: P3-10.1186/gb-2003-4-5-p3.CrossRefPubMed Dennis G, Sherman BT, Hosack DA, Yang J, Gao W, Lane HC, Lempicki RA: DAVID: Database for Annotation, Visualization, and Integrated Discovery. Genome Biol. 2003, 4: P3-10.1186/gb-2003-4-5-p3.CrossRefPubMed
44.
Zurück zum Zitat Tu IP, Whittemore AS: Power of association and linkage tests when the disease alleles are unobserved. Am J Human Genetics. 1999, 64: 641-649. 10.1086/302253.CrossRef Tu IP, Whittemore AS: Power of association and linkage tests when the disease alleles are unobserved. Am J Human Genetics. 1999, 64: 641-649. 10.1086/302253.CrossRef
46.
Zurück zum Zitat Andrews MR, Czvitkovich S, Dassie E, Vogelaar CF, Faissner A, Blits B, Gage FH, French-Constant C, Fawcett JW: Alpha9 integrin promotes neurite outgrowth on tenascin-C and enhances sensory axon regeneration. J Neurosci. 2009, 29: 5546-5557. 10.1523/JNEUROSCI.0759-09.2009.CrossRefPubMed Andrews MR, Czvitkovich S, Dassie E, Vogelaar CF, Faissner A, Blits B, Gage FH, French-Constant C, Fawcett JW: Alpha9 integrin promotes neurite outgrowth on tenascin-C and enhances sensory axon regeneration. J Neurosci. 2009, 29: 5546-5557. 10.1523/JNEUROSCI.0759-09.2009.CrossRefPubMed
47.
Zurück zum Zitat Vessey JP, Karra D: More than just synaptic building blocks: scaffolding proteins of the post-synaptic density regulate dendritic patterning. J Neurochem. 2007, 102: 324-332. 10.1111/j.1471-4159.2007.04662.x.CrossRefPubMed Vessey JP, Karra D: More than just synaptic building blocks: scaffolding proteins of the post-synaptic density regulate dendritic patterning. J Neurochem. 2007, 102: 324-332. 10.1111/j.1471-4159.2007.04662.x.CrossRefPubMed
48.
Zurück zum Zitat Ashburner M, Ball CA, Blake JA, Botstein D, Butler H, Cherry JM, Davis AP, Dolinski K, Dwight SS, Eppig JT: Gene ontology: Tool for the unification of biology. the gene ontology consortium. Nat Genet. 2000, 5: 25-29.CrossRef Ashburner M, Ball CA, Blake JA, Botstein D, Butler H, Cherry JM, Davis AP, Dolinski K, Dwight SS, Eppig JT: Gene ontology: Tool for the unification of biology. the gene ontology consortium. Nat Genet. 2000, 5: 25-29.CrossRef
49.
Zurück zum Zitat Muller-Myhsok B, Abel L: Genetic analysis of complex diseases. Science. 1997, 275: 1328-9. author reply 1329-30PubMed Muller-Myhsok B, Abel L: Genetic analysis of complex diseases. Science. 1997, 275: 1328-9. author reply 1329-30PubMed
50.
Zurück zum Zitat Kraft P: Curses--winner's and otherwise--in genetic epidemiology. Epidemiology. 2008, 19: 649-51. 10.1097/EDE.0b013e318181b865. discussion 657-8CrossRefPubMed Kraft P: Curses--winner's and otherwise--in genetic epidemiology. Epidemiology. 2008, 19: 649-51. 10.1097/EDE.0b013e318181b865. discussion 657-8CrossRefPubMed
51.
Zurück zum Zitat Persico AM, Bourgeron T: Searching for ways out of the autism maze: genetic, epigenetic and environmental clues. Trends Neurosci. 2006, 29: 349-358. 10.1016/j.tins.2006.05.010.CrossRefPubMed Persico AM, Bourgeron T: Searching for ways out of the autism maze: genetic, epigenetic and environmental clues. Trends Neurosci. 2006, 29: 349-358. 10.1016/j.tins.2006.05.010.CrossRefPubMed
52.
Zurück zum Zitat Bauman ML, Kemper TL: Neuroanatomic observations of the brain in autism: a review and future directions. Int J Dev Neurosci. 2005, 23: 183-187. 10.1016/j.ijdevneu.2004.09.006.CrossRefPubMed Bauman ML, Kemper TL: Neuroanatomic observations of the brain in autism: a review and future directions. Int J Dev Neurosci. 2005, 23: 183-187. 10.1016/j.ijdevneu.2004.09.006.CrossRefPubMed
53.
Zurück zum Zitat Raymond GV, Bauman ML, Kemper TL: Hippocampus in autism: a Golgi analysis. Acta Neuropathol. 1996, 91: 117-119. 10.1007/s004010050401.CrossRefPubMed Raymond GV, Bauman ML, Kemper TL: Hippocampus in autism: a Golgi analysis. Acta Neuropathol. 1996, 91: 117-119. 10.1007/s004010050401.CrossRefPubMed
54.
Zurück zum Zitat Kaufmann WE, Moser HW: Dendritic anomalies in disorders associated with mental retardation. Cereb Cortex. 2000, 10: 981-991. 10.1093/cercor/10.10.981.CrossRefPubMed Kaufmann WE, Moser HW: Dendritic anomalies in disorders associated with mental retardation. Cereb Cortex. 2000, 10: 981-991. 10.1093/cercor/10.10.981.CrossRefPubMed
55.
Zurück zum Zitat Kaufmann WE, MacDonald SM, Altamura CR: Dendritic cytoskeletal protein expression in mental retardation: an immunohistochemical study of the neocortex in Rett syndrome. Cereb Cortex. 2000, 10: 992-1004. 10.1093/cercor/10.10.992.CrossRefPubMed Kaufmann WE, MacDonald SM, Altamura CR: Dendritic cytoskeletal protein expression in mental retardation: an immunohistochemical study of the neocortex in Rett syndrome. Cereb Cortex. 2000, 10: 992-1004. 10.1093/cercor/10.10.992.CrossRefPubMed
56.
Zurück zum Zitat Wegiel J, Kuchna I, Nowicki K, Imaki H, Wegiel J, Marchi E, Ma SY, Chauhan A, Chauhan V, Bobrowicz TW: The neuropathology of autism: defects of neurogenesis and neuronal migration, and dysplastic changes. Acta Neuropathol. 2010, 119: 755-770. 10.1007/s00401-010-0655-4.PubMedCentralCrossRefPubMed Wegiel J, Kuchna I, Nowicki K, Imaki H, Wegiel J, Marchi E, Ma SY, Chauhan A, Chauhan V, Bobrowicz TW: The neuropathology of autism: defects of neurogenesis and neuronal migration, and dysplastic changes. Acta Neuropathol. 2010, 119: 755-770. 10.1007/s00401-010-0655-4.PubMedCentralCrossRefPubMed
57.
Zurück zum Zitat Casanova MF, Buxhoeveden DP, Switala AE, Roy E: Minicolumnar pathology in autism. Neurology. 2002, 58: 428-432.CrossRefPubMed Casanova MF, Buxhoeveden DP, Switala AE, Roy E: Minicolumnar pathology in autism. Neurology. 2002, 58: 428-432.CrossRefPubMed
58.
Zurück zum Zitat Casanova M, Trippe J: Radial cytoarchitecture and patterns of cortical connectivity in autism. Philos Trans R Soc Lond B Biol Sci. 2009, 364: 1433-1436. 10.1098/rstb.2008.0331.PubMedCentralCrossRefPubMed Casanova M, Trippe J: Radial cytoarchitecture and patterns of cortical connectivity in autism. Philos Trans R Soc Lond B Biol Sci. 2009, 364: 1433-1436. 10.1098/rstb.2008.0331.PubMedCentralCrossRefPubMed
59.
Zurück zum Zitat Minshew NJ, Keller TA: The nature of brain dysfunction in autism: functional brain imaging studies. Curr Opin Neurol. 2010, 23: 124-130. 10.1097/WCO.0b013e32833782d4.PubMedCentralCrossRefPubMed Minshew NJ, Keller TA: The nature of brain dysfunction in autism: functional brain imaging studies. Curr Opin Neurol. 2010, 23: 124-130. 10.1097/WCO.0b013e32833782d4.PubMedCentralCrossRefPubMed
60.
Zurück zum Zitat Vessey JP, Schoderboeck L, Gingl E, Luzi E, Riefler J, Di Leva F, Karra D, Thomas S, Kiebler MA, Macchi P: Mammalian Pumilio 2 regulates dendrite morphogenesis and synaptic function. Proc Natl Acad Sci USA. 2010, 107: 3222-3227. 10.1073/pnas.0907128107.PubMedCentralCrossRefPubMed Vessey JP, Schoderboeck L, Gingl E, Luzi E, Riefler J, Di Leva F, Karra D, Thomas S, Kiebler MA, Macchi P: Mammalian Pumilio 2 regulates dendrite morphogenesis and synaptic function. Proc Natl Acad Sci USA. 2010, 107: 3222-3227. 10.1073/pnas.0907128107.PubMedCentralCrossRefPubMed
61.
Zurück zum Zitat Maeda N, Fukazawa N, Ishii M: Chondroitin sulfate proteoglycans in neural development and plasticity. Front Biosci. 2010, 15: 626-644. 10.2741/3637.CrossRef Maeda N, Fukazawa N, Ishii M: Chondroitin sulfate proteoglycans in neural development and plasticity. Front Biosci. 2010, 15: 626-644. 10.2741/3637.CrossRef
62.
Zurück zum Zitat Fukazawa N, Yokoyama S, Eiraku M, Kengaku M, Maeda N: Receptor type protein tyrosine phosphatase zeta-pleiotrophin signaling controls endocytic trafficking of DNER that regulates neuritogenesis. Mol Cell Biol. 2008, 28: 4494-4506. 10.1128/MCB.00074-08.PubMedCentralCrossRefPubMed Fukazawa N, Yokoyama S, Eiraku M, Kengaku M, Maeda N: Receptor type protein tyrosine phosphatase zeta-pleiotrophin signaling controls endocytic trafficking of DNER that regulates neuritogenesis. Mol Cell Biol. 2008, 28: 4494-4506. 10.1128/MCB.00074-08.PubMedCentralCrossRefPubMed
63.
Zurück zum Zitat Satterfield TF, Jackson SM, Pallanck LJ: A Drosophila homolog of the polyglutamine disease gene SCA2 is a dosage-sensitive regulator of actin filament formation. Genetics. 2002, 162: 1687-1702.PubMedCentralPubMed Satterfield TF, Jackson SM, Pallanck LJ: A Drosophila homolog of the polyglutamine disease gene SCA2 is a dosage-sensitive regulator of actin filament formation. Genetics. 2002, 162: 1687-1702.PubMedCentralPubMed
64.
Zurück zum Zitat Swiercz JM, Worzfeld T, Offermanns S: Semaphorin 4D signaling requires the recruitment of phospholipase C gamma into the plexin-B1 receptor complex. Mol Cell Biol. 2009, 29: 6321-6334. 10.1128/MCB.00103-09.PubMedCentralCrossRefPubMed Swiercz JM, Worzfeld T, Offermanns S: Semaphorin 4D signaling requires the recruitment of phospholipase C gamma into the plexin-B1 receptor complex. Mol Cell Biol. 2009, 29: 6321-6334. 10.1128/MCB.00103-09.PubMedCentralCrossRefPubMed
65.
Zurück zum Zitat Vodrazka P, Korostylev A, Hirschberg A, Swiercz JM, Worzfeld T, Deng S, Fazzari P, Tamagnone L, Offermanns S, Kuner R: The semaphorin 4D-plexin-B signalling complex regulates dendritic and axonal complexity in developing neurons via diverse pathways. Eur J Neurosci. 2009, 30: 1193-1208. 10.1111/j.1460-9568.2009.06934.x.CrossRefPubMed Vodrazka P, Korostylev A, Hirschberg A, Swiercz JM, Worzfeld T, Deng S, Fazzari P, Tamagnone L, Offermanns S, Kuner R: The semaphorin 4D-plexin-B signalling complex regulates dendritic and axonal complexity in developing neurons via diverse pathways. Eur J Neurosci. 2009, 30: 1193-1208. 10.1111/j.1460-9568.2009.06934.x.CrossRefPubMed
66.
Zurück zum Zitat Bekirov IH, Nagy V, Svoronos A, Huntley GW, Benson DL: Cadherin-8 and N-cadherin differentially regulate pre- and postsynaptic development of the hippocampal mossy fiber pathway. Hippocampus. 2008, 18: 349-363. 10.1002/hipo.20395.PubMedCentralCrossRefPubMed Bekirov IH, Nagy V, Svoronos A, Huntley GW, Benson DL: Cadherin-8 and N-cadherin differentially regulate pre- and postsynaptic development of the hippocampal mossy fiber pathway. Hippocampus. 2008, 18: 349-363. 10.1002/hipo.20395.PubMedCentralCrossRefPubMed
67.
Zurück zum Zitat Gordon-Weeks PR: Neuronal Growth Cones. 2000, Cambridge; New York: Cambridge University PressCrossRef Gordon-Weeks PR: Neuronal Growth Cones. 2000, Cambridge; New York: Cambridge University PressCrossRef
68.
Zurück zum Zitat Bagnard D: Axon Growth and Guidance. 2007, New York; Texas: Springer Science and Business Media; Landes BioscienceCrossRef Bagnard D: Axon Growth and Guidance. 2007, New York; Texas: Springer Science and Business Media; Landes BioscienceCrossRef
69.
Zurück zum Zitat Doherty P, Walsh FS: CAM-FGF receptor interactions: a model for axonal growth. Mol Cell Neurosci. 1996, 8: 99-111. 10.1006/mcne.1996.0049.CrossRef Doherty P, Walsh FS: CAM-FGF receptor interactions: a model for axonal growth. Mol Cell Neurosci. 1996, 8: 99-111. 10.1006/mcne.1996.0049.CrossRef
70.
Zurück zum Zitat Sallee JL, Wittchen ES, Burridge K: Regulation of cell adhesion by protein-tyrosine phosphatases: II. Cell-cell adhesion. J Biol Chem. 2006, 281: 16189-16192. 10.1074/jbc.R600003200.CrossRefPubMed Sallee JL, Wittchen ES, Burridge K: Regulation of cell adhesion by protein-tyrosine phosphatases: II. Cell-cell adhesion. J Biol Chem. 2006, 281: 16189-16192. 10.1074/jbc.R600003200.CrossRefPubMed
71.
Zurück zum Zitat Nakamura F, Tanaka M, Takahashi T, Kalb RG, Strittmatter SM: Neuropilin-1 extracellular domains mediate semaphorin D/III-induced growth cone collapse. Neuron. 1998, 21: 1093-1100. 10.1016/S0896-6273(00)80626-1.CrossRefPubMed Nakamura F, Tanaka M, Takahashi T, Kalb RG, Strittmatter SM: Neuropilin-1 extracellular domains mediate semaphorin D/III-induced growth cone collapse. Neuron. 1998, 21: 1093-1100. 10.1016/S0896-6273(00)80626-1.CrossRefPubMed
72.
Zurück zum Zitat Noren NK, Liu BP, Burridge K, Kreft B: p120 catenin regulates the actin cytoskeleton via Rho family GTPases. J Cell Biol. 2000, 150: 567-580. 10.1083/jcb.150.3.567.PubMedCentralCrossRefPubMed Noren NK, Liu BP, Burridge K, Kreft B: p120 catenin regulates the actin cytoskeleton via Rho family GTPases. J Cell Biol. 2000, 150: 567-580. 10.1083/jcb.150.3.567.PubMedCentralCrossRefPubMed
73.
Zurück zum Zitat Cote JF, Motoyama AB, Bush JA, Vuori K: A novel and evolutionarily conserved PtdIns(3,4,5)P3-binding domain is necessary for DOCK180 signalling. Nat Cell Biol. 2005, 7: 797-807. 10.1038/ncb1280.PubMedCentralCrossRefPubMed Cote JF, Motoyama AB, Bush JA, Vuori K: A novel and evolutionarily conserved PtdIns(3,4,5)P3-binding domain is necessary for DOCK180 signalling. Nat Cell Biol. 2005, 7: 797-807. 10.1038/ncb1280.PubMedCentralCrossRefPubMed
74.
Zurück zum Zitat Xie Z, Cahill ME, Penzes P: Kalirin loss results in cortical morphological alterations. Mol Cell Neurosci. 2010, 43: 81-89. 10.1016/j.mcn.2009.09.006.PubMedCentralCrossRefPubMed Xie Z, Cahill ME, Penzes P: Kalirin loss results in cortical morphological alterations. Mol Cell Neurosci. 2010, 43: 81-89. 10.1016/j.mcn.2009.09.006.PubMedCentralCrossRefPubMed
75.
Zurück zum Zitat Yang Y, Marcello M, Endris V, Saffrich R, Fischer R, Trendelenburg MF, Sprengel R, Rappold G: MEGAP impedes cell migration via regulating actin and microtubule dynamics and focal complex formation. Exp Cell Res. 2006, 312: 2379-2393. 10.1016/j.yexcr.2006.04.001.CrossRefPubMed Yang Y, Marcello M, Endris V, Saffrich R, Fischer R, Trendelenburg MF, Sprengel R, Rappold G: MEGAP impedes cell migration via regulating actin and microtubule dynamics and focal complex formation. Exp Cell Res. 2006, 312: 2379-2393. 10.1016/j.yexcr.2006.04.001.CrossRefPubMed
76.
Zurück zum Zitat Schwamborn JC, Puschel AW: The sequential activity of the GTPases Rap1B and Cdc42 determines neuronal polarity. Nat Neurosci. 2004, 7: 923-929. 10.1038/nn1295.CrossRefPubMed Schwamborn JC, Puschel AW: The sequential activity of the GTPases Rap1B and Cdc42 determines neuronal polarity. Nat Neurosci. 2004, 7: 923-929. 10.1038/nn1295.CrossRefPubMed
77.
Zurück zum Zitat Liu C, Takahashi M, Li Y, Song S, Dillon TJ, Shinde U, Stork PJ: Ras is required for the cyclic AMP-dependent activation of Rap1 via Epac2. Mol Cell Biol. 2008, 28: 7109-7125. 10.1128/MCB.01060-08.PubMedCentralCrossRefPubMed Liu C, Takahashi M, Li Y, Song S, Dillon TJ, Shinde U, Stork PJ: Ras is required for the cyclic AMP-dependent activation of Rap1 via Epac2. Mol Cell Biol. 2008, 28: 7109-7125. 10.1128/MCB.01060-08.PubMedCentralCrossRefPubMed
78.
Zurück zum Zitat Hernandez-Miranda LR, Parnavelas JG, Chiara F: Molecules and mechanisms involved in the generation and migration of cortical interneurons. ASN Neuro. 2010, 2: e00031-10.1042/AN20090053.PubMedCentralCrossRefPubMed Hernandez-Miranda LR, Parnavelas JG, Chiara F: Molecules and mechanisms involved in the generation and migration of cortical interneurons. ASN Neuro. 2010, 2: e00031-10.1042/AN20090053.PubMedCentralCrossRefPubMed
79.
Zurück zum Zitat Boscher C, Mege RM: Cadherin-11 interacts with the FGF receptor and induces neurite outgrowth through associated downstream signalling. Cell Signal. 2008, 20: 1061-1072. 10.1016/j.cellsig.2008.01.008.CrossRefPubMed Boscher C, Mege RM: Cadherin-11 interacts with the FGF receptor and induces neurite outgrowth through associated downstream signalling. Cell Signal. 2008, 20: 1061-1072. 10.1016/j.cellsig.2008.01.008.CrossRefPubMed
80.
Zurück zum Zitat Georgiev D, Taniura H, Kambe Y, Takarada T, Yoneda Y: A critical importance of polyamine site in NMDA receptors for neurite outgrowth and fasciculation at early stages of P19 neuronal differentiation. Exp Cell Res. 2008, 314: 2603-2617. 10.1016/j.yexcr.2008.06.009.CrossRefPubMed Georgiev D, Taniura H, Kambe Y, Takarada T, Yoneda Y: A critical importance of polyamine site in NMDA receptors for neurite outgrowth and fasciculation at early stages of P19 neuronal differentiation. Exp Cell Res. 2008, 314: 2603-2617. 10.1016/j.yexcr.2008.06.009.CrossRefPubMed
81.
Zurück zum Zitat Williams EJ, Furness J, Walsh FS, Doherty P: Activation of the FGF receptor underlies neurite outgrowth stimulated by L1, N-CAM, and N-cadherin. Neuron. 1994, 13: 583-594. 10.1016/0896-6273(94)90027-2.CrossRefPubMed Williams EJ, Furness J, Walsh FS, Doherty P: Activation of the FGF receptor underlies neurite outgrowth stimulated by L1, N-CAM, and N-cadherin. Neuron. 1994, 13: 583-594. 10.1016/0896-6273(94)90027-2.CrossRefPubMed
82.
Zurück zum Zitat Falk J, Bonnon C, Girault JA, Faivre-Sarrailh C: F3/contactin, a neuronal cell adhesion molecule implicated in axogenesis and myelination. Biol Cell. 2002, 94: 327-334. 10.1016/S0248-4900(02)00006-0.CrossRefPubMed Falk J, Bonnon C, Girault JA, Faivre-Sarrailh C: F3/contactin, a neuronal cell adhesion molecule implicated in axogenesis and myelination. Biol Cell. 2002, 94: 327-334. 10.1016/S0248-4900(02)00006-0.CrossRefPubMed
83.
Zurück zum Zitat Lin X, Ogiya M, Takahara M, Yamaguchi W, Furuyama T, Tanaka H, Tohyama M, Inagaki S: Sema4D-plexin-B1 implicated in regulation of dendritic spine density through RhoA/ROCK pathway. Neurosci Lett. 2007, 428: 1-6. 10.1016/j.neulet.2007.09.045.CrossRefPubMed Lin X, Ogiya M, Takahara M, Yamaguchi W, Furuyama T, Tanaka H, Tohyama M, Inagaki S: Sema4D-plexin-B1 implicated in regulation of dendritic spine density through RhoA/ROCK pathway. Neurosci Lett. 2007, 428: 1-6. 10.1016/j.neulet.2007.09.045.CrossRefPubMed
84.
Zurück zum Zitat Akiyama H, Matsu-ura T, Mikoshiba K, Kamiguchi H: Control of neuronal growth cone navigation by asymmetric inositol 1,4,5-trisphosphate signals. Sci Signal. 2009, 2: ra34-10.1126/scisignal.2000196.CrossRefPubMed Akiyama H, Matsu-ura T, Mikoshiba K, Kamiguchi H: Control of neuronal growth cone navigation by asymmetric inositol 1,4,5-trisphosphate signals. Sci Signal. 2009, 2: ra34-10.1126/scisignal.2000196.CrossRefPubMed
85.
Zurück zum Zitat Robles E, Woo S, Gomez TM: Src-dependent tyrosine phosphorylation at the tips of growth cone filopodia promotes extension. J Neurosci. 2005, 25: 7669-7681. 10.1523/JNEUROSCI.2680-05.2005.CrossRefPubMed Robles E, Woo S, Gomez TM: Src-dependent tyrosine phosphorylation at the tips of growth cone filopodia promotes extension. J Neurosci. 2005, 25: 7669-7681. 10.1523/JNEUROSCI.2680-05.2005.CrossRefPubMed
86.
Zurück zum Zitat Nishiyama M, Hoshino A, Tsai L, Henley JR, Goshima Y, Tessier-Lavigne M, Poo MM, Hong K: Cyclic AMP/GMP-dependent modulation of Ca2+ channels sets the polarity of nerve growth-cone turning. Nature. 2003, 423: 990-995. 10.1038/nature01751.CrossRefPubMed Nishiyama M, Hoshino A, Tsai L, Henley JR, Goshima Y, Tessier-Lavigne M, Poo MM, Hong K: Cyclic AMP/GMP-dependent modulation of Ca2+ channels sets the polarity of nerve growth-cone turning. Nature. 2003, 423: 990-995. 10.1038/nature01751.CrossRefPubMed
87.
Zurück zum Zitat Dawes AT, Edelstein-Keshet L: Phosphoinositides and Rho proteins spatially regulate actin polymerization to initiate and maintain directed movement in a one-dimensional model of a motile cell. Biophys J. 2007, 92: 744-768. 10.1529/biophysj.106.090514.PubMedCentralCrossRefPubMed Dawes AT, Edelstein-Keshet L: Phosphoinositides and Rho proteins spatially regulate actin polymerization to initiate and maintain directed movement in a one-dimensional model of a motile cell. Biophys J. 2007, 92: 744-768. 10.1529/biophysj.106.090514.PubMedCentralCrossRefPubMed
88.
Zurück zum Zitat Govek EE, Newey SE, Van Aelst L: The role of the Rho GTPases in neuronal development. Genes Dev. 2005, 19: 1-49. 10.1101/gad.1256405.CrossRefPubMed Govek EE, Newey SE, Van Aelst L: The role of the Rho GTPases in neuronal development. Genes Dev. 2005, 19: 1-49. 10.1101/gad.1256405.CrossRefPubMed
89.
Zurück zum Zitat Calabrese B, Wilson MS, Halpain S: Development and regulation of dendritic spine synapses. Physiology (Bethesda). 2006, 21: 38-47.CrossRef Calabrese B, Wilson MS, Halpain S: Development and regulation of dendritic spine synapses. Physiology (Bethesda). 2006, 21: 38-47.CrossRef
Metadaten
Titel
A noise-reduction GWAS analysis implicates altered regulation of neurite outgrowth and guidance in autism
verfasst von
John P Hussman
Ren-Hua Chung
Anthony J Griswold
James M Jaworski
Daria Salyakina
Deqiong Ma
Ioanna Konidari
Patrice L Whitehead
Jeffery M Vance
Eden R Martin
Michael L Cuccaro
John R Gilbert
Jonathan L Haines
Margaret A Pericak-Vance
Publikationsdatum
01.12.2011
Verlag
BioMed Central
Erschienen in
Molecular Autism / Ausgabe 1/2011
Elektronische ISSN: 2040-2392
DOI
https://doi.org/10.1186/2040-2392-2-1

Weitere Artikel der Ausgabe 1/2011

Molecular Autism 1/2011 Zur Ausgabe