The American Psychiatric Association (APA) has updated its Privacy Policy and Terms of Use, including with new information specifically addressed to individuals in the European Economic Area. As described in the Privacy Policy and Terms of Use, this website utilizes cookies, including for the purpose of offering an optimal online experience and services tailored to your preferences.

Please read the entire Privacy Policy and Terms of Use. By closing this message, browsing this website, continuing the navigation, or otherwise continuing to use the APA's websites, you confirm that you understand and accept the terms of the Privacy Policy and Terms of Use, including the utilization of cookies.

×
PerspectivesFull Access

Integrated Genome-Wide Association Study Findings: Identification of a Neurodevelopmental Network for Attention Deficit Hyperactivity Disorder

Abstract

Objective:

Attention deficit hyperactivity disorder (ADHD) is a highly heritable neuropsychiatric disorder. In the present study, the authors investigated the presence of genomic convergence in the top findings of the five published genome-wide association studies (GWASs) of ADHD.

Method:

The authors carried out bioinformatics pathway analyses, using the Ingenuity and BiNGO tools, as well as a systematic literature analysis of 85 genes from the five published GWASs containing single nucleotide polymorphisms associated with ADHD at a p value <0.0001.

Results:

Findings revealed that 45 of the 85 top-ranked ADHD candidate genes encode proteins that fit into a neurodevelopmental network involved in directed neurite outgrowth. Data on copy number variations in patients with ADHD and data from animal studies provide further support for the involvement of this network in ADHD etiology. Several network proteins are also directly modulated by stimulants, the most commonly used psychopharmacological treatment for ADHD.

Conclusions:

The authors have identified a protein network for ADHD that contributes to our understanding of the molecular basis of the disorder. In addition, the data suggest new candidate genes for ADHD and provide clues to future research into psychopharmacological ADHD treatments.

Attention deficit hyperactivity disorder (ADHD) is a common neuropsychiatric disorder that is observed in children and adults. ADHD is characterized by age-inappropriate levels of hyperactive, impulsive, and inattentive behaviors and leads to significant clinical and psychosocial impairments (DSM-IV-TR). The disorder has a worldwide prevalence of 5%–6% in children (1) and 1%–4% in adults (2, 3). The most effective and commonly used medications for ADHD are the stimulants methylphenidate and amphetamine and the nonstimulant drug atomoxetine, which (predominantly) modulate dopaminergic and noradrenergic neurotransmission (4, 5).

Converging neurobiological evidence suggests that ADHD involves alterations of catecholaminergic brain circuits (6). Neuroimaging studies in children with ADHD also show neurodevelopmental brain anomalies, such as reduced cortical volume and folding and gray matter heterotopia (710), which indicate that, neuroanatomically, ADHD is a disorder of early brain development. In addition, specific neuroimaging modalities, such as diffusion tensor imaging (1113), (resting-state) functional magnetic resonance imaging (fMRI) (1416), and EEG (17), have shown that both structural and functional brain connectivity are impaired in ADHD patients.

Evidence from family and twin studies shows that ADHD is a highly heritable disorder, and approximately 76% of the phenotypic variance can be explained by genetic factors (18). ADHD behaves as a multifactorial (or complex) disorder, in which different combinations of genetic and environmental factors contribute to the overall risk of developing the disorder. The genetic model underlying most cases of ADHD is likely one in which multiple genetic factors of small-to-moderate effect size contribute to the disease etiology (19). Recent data indicate that rare copy number variations of genes may also be relevant to the etiology of ADHD (2022), but pending further studies, it is unclear which proportion of ADHD cases could be explained by these variants.

To date, eight independent genome-wide linkage scans have been conducted for ADHD (23, 24). A recent meta-analysis of seven of these studies identified genome-wide significant linkage for a chromosomal region on 16q (23). In addition, a large number of candidate gene association studies have been published, and these have primarily focused on genes involved in catecholaminergic neurotransmission. Recent meta-analyses of these studies have yielded significant meta-association for six ADHD candidate genes (i.e., the dopamine and serotonin transporter genes DAT1/SLC6A3 and 5-HTT/SLC6A4, the DRD4 and DRD5 dopamine receptor genes, the HTR1B serotonin receptor gene, and the SNAP25 gene involved in neurotransmission) (25, 26). In 2008, the results of three genome-wide association studies (GWASs) for ADHD were published (2729) in two independent data sets. Lasky-Su et al. (28) reported the following two findings that reached (trait-specific) genome-wide significance: 1) a single nucleotide polymorphism (SNP) in CDH13 and 2) a SNP in GFOD1.

Very recently, the results of two additional GWASs for ADHD in two independent samples were reported (30, 31). Apart from CDH13, which maps to the only chromosomal region that yielded genome-wide significant linkage for ADHD in a recent meta-analysis (23), overlap between GWAS findings and any of the reported linkage or candidate gene-based association findings has been limited (32), although this is probably due in part to the fact that the published GWASs for ADHD are strongly underpowered.

In the present study, we attempted to integrate the most significant findings from the five reported GWASs for ADHD into protein signaling networks that would not only increase our understanding of the genetic etiology of ADHD but also provide clues for future research into more effective (psychopharmacological) treatments for the disorder. Based on bioinformatics analyses and a systematic literature analysis of the (putative) function of the proteins encoded by the 85 top-ranked genes emerging from the five GWASs for ADHD, we describe a signaling network involved in neurite outgrowth that includes 45 of these proteins. The other 40 proteins could not be convincingly linked to the network. Corroborating evidence for the involvement of this gene network from chromosomal aberrations in humans, from animal models, and from psychopharmacological studies is also presented.

Method

Identification of ADHD Candidate Genes From GWASs

To date, the results of five GWASs of ADHD have been reported. Neale et al. (27) reported GWAS results for a categorical ADHD phenotype in 909 Caucasian case-parent trios that were collected as part of the International Multicenter ADHD Genetics study in children. Lasky-Su et al. (28) performed a GWAS on the same sample using quantitative phenotypes of ADHD.

Lesch et al. (29) conducted a GWAS for a categorically defined ADHD phenotype using pooled DNA from 343 ADHD-affected adults and 304 comparison subjects. Recently, Mick et al. (30) also reported the results of a GWAS of a categorical ADHD phenotype in a sample of 735 trios from three different U.S. clinical sites (30). In addition, the International Multicenter ADHD Genetics II consortium reported the results of a GWAS using a case-control design in 896 unrelated participants with ADHD and 2,455 comparison subjects (31). From these five GWASs, we compiled a list of top SNPs, taking into account SNPs located within exonic, intronic, or untranslated regions of genes and in association with the ADHD phenotype at p<1.00E-04 (Table 1).

TABLE 1. The Top Single Nucleotide Polymorphisms (SNPs) From the Five Reported Genome-Wide Association Studies of Attention Deficit Hyperactivity Disorder (ADHD)a

GWASSNPpLocusGenePosition Within GeneGene Size
Neale et al. (27)b
rs96764471.00E-0519q13.33NUCB1intronic23 kb
rs8764772.70E-054p15.31KCNIP4intronic1,220 kb
rs76576083.00E-054q32.3SPOCK3intronic501 kb
rs37823095.50E-0512p11.23ITPR2intronic498 kb
rs9227815.50E-0515q22.2RORAintronic741 kb
rs125055025.60E-054q22.1FAM190Aintronic1,360 kb
rs10627937.00E-056q14.1SH3BGRL23′ untranslated region72 kb
rs7039659.20E-0510q22.3ZMIZ1intronic248 kb
rs41496019.30E-0518q21.31NEDD4Lintronic357 kb
Lasky-Su et al. (28)c
rs6565113d16q23.3CDH13intronic948 kb
rs552655d6p24.1GFOD1intronic130 kb
rs41287671.28E-0615q25.1IL16intronic117 kb
rs117196642.48E-063p24.3ZNF385Dintronic955 kb
rs75779252.55E-062q21.2NAP5intronic897 kb
rs74950522.83E-0615q26.1SLCO3A1intronic319 kb
rs172818133.46E-0616q12.1ZNF423intronic336 kb
rs3635123.89E-0621q21.3GRIK1intronic403 kb
rs80416753.98E-0615q14MEIS2intronic210 kb
rs176410784.73E-069p24.3DMRT2nonsynonymous, coding7 kb
rs9304215.64E-062p21MTA33′ untranslated region262 kb
rs176519786.05E-063p13FOXP1intronic629 kb
rs4785978.08E-0612q24.22NOS1intronic163 kb
rs67916448.32E-063p14.2FHITintronic1,502 kb
rs2604618.38E-0619q13.43ZNF544intronic35 kb
rs22904168.51E-068q24.3NAPRT1synonymous, coding4 kb
rs12021998.52E-066p22.3MBOAT1intronic110 kb
Lesch et al. (29)e
rs8646431.30E-083p22.2MOBP3′ untranslated region59 kb
rs79952151.35E-0813q31.3GPC6intronic1,180 kb
rs112438975.63E-089q34.13C9orf98intronic153 kb
rs71643351.30E-0715q23ITGA11intronic130 kb
rs2204701.34E-0717p13.2ITGAEintronic87 kb
rs109832381.37E-079q33.1ASTN2intronic990 kb
rs25876952.73E-072q14.2PCDP1intronic118 kb
rs22815975.41E-071p35.1CSMD2intronic652 kb
rs105146048.06E-0716q24.1ATP2C2intronic96 kb
rs26777449.69E-0715q26.1MAN2A2intronic18 kb
rs25027311.58E-069q34.11DNM1intronic52 kb
rs21991611.62E-065q13.2MAP1Bintronic100 kb
rs107862841.96E-0610q24.1TLL2intronic149 kb
rs15553223.60E-0620q11.22MMP24intronic50 kb
rs169285293.90E-0610q22.1UNC5Bintronic90 kb
rs5159104.36E-0610q25.1C10orf79intronic102 kb
rs22373494.63E-067p15.1CREB5intronic527 kb
rs49648054.74E-0612q23.3NT5DC3intronic71 kb
rs37999774.90E-066p21.1SUPT3Hintronic569 kb
rs116464117.40E-0616q23.3CDH13intronic948 kb
rs4697277.54E-065q22.2REEP5intronic46 kb
rs22416858.11E-062p25.3MYT1Lintronic542 kb
rs22420738.27E-062q33.3CRYGCintronic2 kb
rs133950229.68E-062p12CTNNA2intronic1,460 kb
rs115940821.00E-0510q22.1CDH23intronic419 kb
rs38932152.56E-0511p15.1KCNC1intronic47 kb
Mick et al. (30)f
rs28238196.70E-0721q21.1C21orf34intronic539 kb
rs110748896.71E-0716p13.13EMP2intronic52 kb
rs80747511.18E-0617q24.1CCDC46intronic556 kb
rs18591562.05E-064q22.3BMPR1Bintronic400 kb
rs4382593.71E-066p24.3AL024506.1intronic122 kb
rs26023813.85E-062q37.1UGT1A9intronic101 kb
rs100119268.07E-064q25ELOVL6intronic153 kb
rs119533461.59E-055q15MCTP1intronic578 kb
rs77021781.84E-055q35.1DUSP1intronic3 kb
rs170232181.88E-053p24.1RBMS3intronic1,166 kb
rs65616862.08E-0513q14.3LECT1intronic37 kb
International Multicenter ADHD Genetics II consortium (31)g
rs108239736.30E-0710q21.1PRKG1intronic1,307 kb
rs22915672.00E-067q32.1FLNCintronic29 kb
rs132388315.60E-067q32.1KCPintronic48 kb
rs115932415.70E-0610q26.3TCERG1Lintronic219 kb
rs123175529.10E-0612q14.1PPM1H3′ untranslated region291 kb
rs171518211.10E-057p21.3NXPH1intronic319 kb
rs170956901.60E-051p31.1C1orf173intronic106 kb
rs80450062.30E-0516q23.3CDH13intronic948 kb
rs10235562.60E-057p14.3AC005493.1intronic14 kb
rs23945382.80E-0510q22.1HK1intronic132 kb
rs12787762.90E-0513q34ATP11Aintronic197 kb
rs174953663.10E-052p16.3NRXN1intronic1,114 kb
rs9062193.70E-0510q22.1HKDC1nonsynonymous, coding47 kb
rs14721474.10E-057q32.1ATP6V1Fintronic3 kb
rs5899434.90E-0511q22.3DYNC2H1intronic370 kb
rs29623935.60E-055q34GABRB2intronic260 kb
rs8469715.60E-056q21SOBPintronic170 kb
rs104458616.50E-052q14.3CNTNAP5intronic890 kb
rs42616727.20E-052q22.2LRP1Bintronic1,900 kb
rs69865787.30E-058q13.2PREX2intronic280 kb
rs120900007.40E-051q25.2FAM5Bintronic111 kb
rs15656308.00E-057q32.1CCDC136intronic31 kb
rs45332678.30E-0515q26.3ADAMTS17intronic370 kb
rs29793079.00E-053q21.2OSBPL11intronic66 kb

a Data show SNPs from the GWASs (27–31) located within genes and with a p value <1.00E-04 for association with ADHD after correction for linkage disequilibrium. The 45 (different) genes encoding proteins that could be directly placed in the putative ADHD network are indicated in bold.

b GWAS of a categorical DSM-IV ADHD phenotype in 909 Caucasian case-parent trios collected as part of the International Multicentre ADHD Genetics study in children.

c GWAS of quantitative phenotypes of ADHD carried out in the same International Multicentre ADHD Genetics sample used in the GWAS conducted by Neale et al. (27).

d Findings revealed genome-wide significance for a single quantitative trait using the family-based association testing principal component algorithm.

e GWAS of a categorical DSM-IV ADHD phenotype using pooled DNA from 343 ADHD-affected adults and 304 comparison subjects. (Reported p values in this study were ranked according to the mean rank calculated across three statistics.)

f GWAS of DSM-IV-TR ADHD in a combined sample of 735 trios from three U.S. clinical sites.

g GWAS conducted in 896 unrelated case patients with DSM-IV ADHD and 2,455 comparison subjects using a case-control design.

TABLE 1. The Top Single Nucleotide Polymorphisms (SNPs) From the Five Reported Genome-Wide Association Studies of Attention Deficit Hyperactivity Disorder (ADHD)a

Enlarge table

Ingenuity and BiNGO Pathway Analyses

In order to detect significantly enriched gene categories in the selected ADHD candidate genes from the five GWASs, we performed analyses using the Ingenuity Pathway Analysis software package (http://www.ingenuity.com) and the BiNGO (Biological Network Gene Ontology) bioinformatics tool (33). We performed similar analyses on a set of top genes from four GWASs for diabetes type I and Crohn's disease and compared the results with those for ADHD in order to identify possible overrepresentation bias of large (brain-expressed) genes in the top findings of the ADHD GWASs.

The Ingenuity software package (http://www.ingenuity.com) uses the Ingenuity Knowledge Base, which is based on information from the published literature as well as many other sources, including gene expression and gene annotation databases, to assign genes to different groups and categories of functionally related genes. Each of these categories can be further divided into many subcategories. Ingenuity calculates single p values for the enrichment of each gene category using the right-tailed Fisher's exact test and taking into consideration both the total number of molecules from the analyzed data set and the total number of molecules that is linked to the same gene category according to the Ingenuity Knowledge Base. Furthermore, for each gene category, a multiple testing corrected p value of enrichment, calculated using the Benjamini-Hochberg correction, is provided.

The BiNGO tool also assigns genes to different functional categories, but these are specifically linked to gene ontology terms that group functionally related genes (33) and can be found in the Gene Ontology database ( http://www.geneontology.org/). The gene ontology terms can be further assigned to three main gene ontology subgroups or domains (i.e., cellular component, biological process, and molecular function). As with Ingenuity, the BiNGO tool provides single p values, calculated with the hypergeometric test and taking into consideration both the total number of molecules from the analyzed data set and the total number of molecules that is linked to the same gene ontology term, as well as multiple testing corrected p values, calculated using the Benjamini-Hochberg correction, for the enrichment of each gene ontology term (33).

In the presentation of the results of the Ingenuity and BiNGO analyses for ADHD, only categories/processes with significant enrichment (i.e., multiple testing corrected p<0.05) and containing more than one gene were taken into account. Only in the comparison between ADHD, diabetes type I, and Crohn's disease were the functional categories that only contain one gene also considered.

Literature Analysis

Subsequent to the bioinformatics analysis, we systematically searched the literature for the (proposed) function of all 85 proteins derived from the ADHD candidate genes using PubMed (http://www.ncbi.nlm.nih.gov/sites/entrez) and the UniProt Protein Knowledge Base (UniProtKB), a comprehensive catalog of protein sequences and functional annotations that is updated every 3 weeks and can be accessed online (http://www.uniprot.org/uniprot [34]). For each ADHD candidate, we first looked at the available information in UniProtKB, which in several cases already provided a general indication of the (putative) function of the gene/protein in question. Subsequently, we searched PubMed using the search terms “brain,” “neuron,” and “neurite” in combination with the name of each of the candidate genes/proteins from the GWASs. Lastly, guided by the literature we found, we searched PubMed for functional interactions between several of the candidate genes/proteins.

Results

Table 1 shows the 85 genes from the five GWASs for ADHD fulfilling the inclusion criteria of the present study. When more than one SNP in a single gene was found among the top findings, only the SNP yielding the lowest p value for association with (a quantitative phenotype of) ADHD is presented. As can be derived from Table 1, the only overlap between the top findings of the five published GWASs was for CDH13, in which (three different) SNPs were associated with ADHD at p<1.00E-04 in three GWASs (28, 29, 31).

Analysis of these 85 top ADHD candidate genes using the Ingenuity pathway software revealed a significant enrichment (p=4.11E-08 after correction for multiple testing) of the functional gene category neurological disease, with 44 of the 85 genes falling into this category (Table 2). Furthermore, analysis with the BiNGO bioinformatics tool revealed that the gene ontology processes (calcium) ion binding and hexokinase activity were significantly enriched in the 85 ADHD genes (2.94E-03<p<5.43E-03 and p=2.69E-02, respectively) (Table 2).

TABLE 2. Enrichment Analyses of the Top 85 Attention Deficit Hyperactivity Disorder (ADHD) Candidate Genes in the Five Reported Genome-Wide Association Studies of ADHDa

CategoryGenesSignificanceAdjusted Significance
Ingenuity Pathway Toolb
Neurological disease (44/85 genes)ADAMTS17, ASTN2, ATP2C2, BMPR1B, C10ORF79, C21ORF34, C9ORF98, CCDC46,CDH13, CDH23, CNTNAP5,CREB5, CSMD2, CTNNA2, DYNC2H1, FAM190A, FHIT,FOXP1, GABRB2, GFOD1, GRIK1,ITGA11, ITPR2,KCNIP4, LRP1B, MAP1B, MCTP1, MEIS2, MOBP, MYT1L, NAPRT1,NOS1, NRXN1, NXPH1, PPM1H, PRKG1, RBMS3,RORA, SLCO3A1, SPOCK3, TLL2, UNC5B,ZNF385D,ZNF4234.67E-104.11E-08
Biological Network Gene Ontology toolc
Gene ontology termd
    Ion bindingADAMTS17, ATP11A,ATP2C2,BMPR1B,CDH13, CDH23,CREB5,DMRT2, DUSP1, FHIT, FOXP1, GABRB2,HK1,ITGA11,ITGAE, ITPR2, KCNC1,KCNIP4,LRP1B,MAN2A2, MCTP1,MMP24, MTA3,MYT1L,NOS1, NRXN1,NT5DC3, NUCB1,RORA,SOBP,SPOCK3, TLL2,ZMIZ1,ZNF385D,ZNF423,ZNF5444.10E-062.94E-03
    Metal ion bindingADAMTS17, ATP11A,ATP2C2,BMPR1B,CDH13, CDH23,CREB5,DMRT2, DUSP1, FHIT, FOXP1,HK1,ITGA11,ITGAE, ITPR2, KCNC1,KCNIP4, LRP1B,MAN2A2, MCTP1,MMP24, MTA3,MYT1L,NOS1, NRXN1,NT5DC3, NUCB1, RORA, SOBP,SPOCK3, TLL2,ZMIZ1,ZNF385D,ZNF423,ZNF5447.45E-062.94E-03
    Cation bindingADAMTS17,ATP2C2,BMPR1B,CDH13, CDH23,CREB5,DMRT2, DUSP1, FHIT, FOXP1,HK1,ITGA11,ITGAE, ITPR2, KCNC1,KCNIP4,LRP1B,MAN2A2, MCTP1,MMP24, MTA3,MYT1L,NOS1, NRXN1, NUCB1, RORA,SOBP,SPOCK3, TLL2,ZMIZ1,ZNF385D,ZNF423,ZNF5441.19E-053.14E-03
    Calcium ion bindingATP2C2,CDH13, CDH23,ITGA11,ITGAE, ITPR2,KCNIP4,LRP1B,MCTP1,MMP24,NRXN1, NUCB1, SPOCK3, TLL22.76E-055.43E-03
    Hexokinase activityHK1, HKDC11.71E-042.69E-02

a The genes (in the reported GWASs [27–31]) encoding proteins that could be directly placed in the putative ADHD network are indicated in bold.

b Analyses were conducted using Ingenuity pathway software (www.ingenuity.com). Significance was determined from a single test p value calculated using the right-tailed Fisher's exact test and taking into consideration both the total number of molecules from the analyzed data set and the total number of molecules linked to the same gene category according to the Ingenuity Knowledge Base. Adjusted significance was determined from multiple test-corrected p values using the Benjamini-Hochberg correction (only categories reaching a corrected statistical significance of <0.05 are shown).

c Analyses were conducted using the Biological Network Gene Ontology tool (33). Significance was determined from a single test p value calculated using the hypergeometric test and taking into consideration both the total number of molecules from the analyzed data set and the total number of molecules linked to the same gene ontology term. Adjusted significance was determined from multiple test-corrected p values using the Benjamini-Hochberg correction (only categories reaching a corrected statistical significance of <0.05 are shown).

d All significantly enriched gene ontology terms can be further assigned to the molecular function gene ontology subgroup/domain.

TABLE 2. Enrichment Analyses of the Top 85 Attention Deficit Hyperactivity Disorder (ADHD) Candidate Genes in the Five Reported Genome-Wide Association Studies of ADHDa

Enlarge table

Further literature analysis of the (putative) function of the proteins encoded by the genes in the enriched BiNGO terms revealed that 21 out of 36 genes in the (calcium) ion binding category and two out of two hexokinase activity genes play a role in neurite outgrowth. Subsequently, we investigated the function of the entire set of 85 genes further and found that a total of 45 of the 85 included genes (indicated in bold in tables, figure legend, and throughout this article) fit into a neurodevelopmental network that is involved in directed neurite outgrowth, which is illustrated in Figure 1. The proposed signaling network links axonal guidance and neuronal cell adhesion proteins at the neuronal cell membrane with downstream acting adaptor proteins and neuronal cytoskeleton/extracellular matrix-associated proteins.

FIGURE 1.

FIGURE 1. Schematic Representation of the ADHD Neurodevelopmental Signaling Network for Directed Neurite Outgrowtha

a The genes/proteins that emerged from the five reported genome-wide association studies for ADHD are indicated in yellow. The proteins encoded by genes found in copy number variations in patients with ADHD are indicated by a blue border. The proteins encoded by genes implicated in the etiology of ADHD through gene knockout studies in mice are indicated by a green border. The genes/proteins of which the expression and/or function is regulated by stimulants are indicated by a red border. Evidence placing the genes/proteins into this network is presented in the data supplement.

A detailed description of the evidence linking the genes in the network to neurite outgrowth is provided in the data supplement accompanying the online version of this article.

Involvement of the identified network in ADHD is also supported by the finding that several of the identified genes (i.e., CTNNA2, NRXN1, NAP5, SERPINI1, NOS1, ERK1, ZNF423, NEDD4L, and BMP2) are located within copy number variations-and are hence deleted or duplicated-in people with ADHD (2022, 3541) (Table 3, Figure 1).

Moreover, partial knockout models of the NOS1, ERK1, MAP1B, and RORA genes in mice provide further evidence of the involvement of these genes in ADHD etiology (Figure 1).

TABLE 3. Genes Encoding Proteins From the Putative Attention Deficit Hyperactivity Disorder (ADHD) Network and Reported to be Deleted and/or Duplicated in (Genome-Wide) Copy Number Variations Among Individuals With ADHD

LocusGeneCopy Number VariantPhenotypeAdditional CommentsStudy
2p12CTNNA2adeletedADHDGenome-wide copy number variant analysisElia et al. (20)
2p16.3NRXN1adeletedADHDGenome-wide copy number variant analysisBradley et al. (35)
2q21.2NAP5adeletedAttention deficit disorderBaker et al. (36)
3q26.1SERPINI1duplicatedADHDGenome-wide copy number variant analysisElia et al. (20)
3q26.1SERPINI1deletedADHDGenome-wide copy number variant analysisLesch et al. (21)
12q24.22NOS1aduplicatedADHDCappellacci et al. (37)
16p11.2ERK1duplicatedADHDSix/eight unrelated people with ADHDShinawi et al. (38)
16p11.2ERK1duplicatedADHDGenome-wide copy number variant analysisWilliams et al. (22)
16p11.2ERK1deletedADHDFour/15 unrelated people with ADHDShinawi et al. (38)
16p11.2ERK1deletedADHDGenome-wide copy number variant analysisWilliams et al. (22)
16q12.1ZNF423aduplicatedADHDvan den Berg et al. (39)
18q21.31NEDD4LaduplicatedAttention deficit disorderTwo siblings with attention deficit disorderCeccarini et al. (40)
20p12.3BMP2deletedAttention deficit disorderLalani et al. (41)

a One of the six genes from the reported GWAS (Table 1) that could be directly placed in the putative ADHD network.

TABLE 3. Genes Encoding Proteins From the Putative Attention Deficit Hyperactivity Disorder (ADHD) Network and Reported to be Deleted and/or Duplicated in (Genome-Wide) Copy Number Variations Among Individuals With ADHD

Enlarge table

Mice in which the Nos1 and Erk1 genes have been knocked out display hyperactive behavior (42, 43), while, in contrast, Map1b- (44) and Rora-knockout mice (45) show hypoactivity.

Lastly, several proteins from the network appear to be under control of the stimulants methylphenidate and amphetamine that are used to treat ADHD symptoms (4, 5). Both stimulants have been shown to stimulate neurite outgrowth (46, 47) and directly or indirectly regulate the expression and/or function of several genes/proteins implicated in neurite outgrowth, including a number of genes/proteins from the identified network (Figure 1).

Methylphenidate exposure upregulates the expression of AVP (48) and CREB5 (49) (Figure 1). In addition, methylphenidate upregulates the expression of a large number of other genes not directly found in the identified network but functionally implicated in neurite outgrowth (50). Examples of these genes are MMP14, TIMP2, and TIMP3. The MMP14 gene encodes a neuronal extracellular matrix metalloproteinase, such as MMP24, and TIMP2 and TIMP3 encode two direct inhibitors of MMP14 (34, 50). Amphetamine upregulates the expression of UNC5B (51) and PAK1 (52) and downregulates the expression of CTNNB (53). Amphetamine also activates neuronal ERK1 and ERK2 (54) and increases the level of arachidonic acid in neocortical, extrapyramidal, and limbic brain regions (55). Deficiencies or imbalances of arachidonic acid have been shown to be associated with ADHD (56). Lastly, the adenylate cyclase/cAMP/protein kinase A/CREB-signaling cascade (including protein kinase A activating CREB5) is also activated by amphetamine (57).

An important potential bias of which to be aware in the analysis of the top findings from the reported GWASs of ADHD is the fact that large genes, which are often brain expressed, may be more likely found as associated with a phenotype in a GWAS as a result of chance, since more SNPs are present in these large genes (58). Indeed, the genes found among the top findings of the ADHD GWASs were considerably larger than the average gene size of the human genome ([Table 1] average size of 85 ADHD genes: 373 kb versus average gene size in human genome: 27 kb [59]).

If large genes are more likely found to be associated with a phenotype of a GWAS because of chance, this would be expected to be the case for GWASs of both nonpsychiatric disorders and ADHD, and therefore we compared the results of our analyses with the Ingenuity and BiNGO software tool results for the 20 top ADHD candidate genes with those for the top 20 candidate genes from four published GWASs for diabetes mellitus type I and Crohn's disease (Table 4, Table 5 [also see the data supplement]). This comparison showed that the neurological disease category was not enriched in Crohn's disease and only contained a single gene in diabetes type I (Table 4). In addition, the gene ontology processes that were enriched in the top GWAS findings were clearly different for the three diseases (Table 5).

TABLE 4. Nervous System-Related Gene Functional Categories Significantly Enriched in the Top 20 Candidate Genes From Genome-Wide Association Studies of Attention Deficit Hyperactivity Disorder (ADHD), Diabetes Type I, and Crohn's Diseasea

CategoryGenesSignificancebAdjusted Significancec
ADHD
    Neurological disease (13/20 genes)ASTN2, ATP2C2, C21ORF34, C9ORF98, CCDC46, CDH13, CSMD2, GFOD1, ITGA11, MAP1B, MOBP, PRKG1, TLL21.53E-047.55E-03
    Biogenesis of lamellipodiadCDH135.49E-032.10E-02
Diabetes type I
    Neurological disease (1/20 genes)TH1.42E-023.02E-02
    Long-term depressionINS1.30E-038.11E-03
Crohn's disease
    Neurological disease (0/20 genes)

a Enrichment analyses were conducted using Ingenuity pathway software (www.ingenuity.com). (Further details are presented in Table 1 of the data supplement accompanying the online version of this article.)

b Data were determined using a single test p value calculated with the right-tailed Fisher's exact test and taking into consideration both the total number of molecules from the analyzed data set and the total number of molecules linked to the same gene category according to the Ingenuity Knowledge Base.

c Data were determined using multiple test-corrected p values using the Benjamini-Hochberg correction (only categories reaching a corrected statistical significance of <0.05 are shown).

d Refers to protrusions of the neuronal growth cone that are formed during the process of neurite outgrowth (60).

TABLE 4. Nervous System-Related Gene Functional Categories Significantly Enriched in the Top 20 Candidate Genes From Genome-Wide Association Studies of Attention Deficit Hyperactivity Disorder (ADHD), Diabetes Type I, and Crohn's Diseasea

Enlarge table

TABLE 5. The Top Five Gene Functional Categories Significantly Enriched in the Top 20 Candidate Genes From Genome-Wide Association Studies of Attention Deficit Hyperactivity Disorder (ADHD), Diabetes Type I, and Crohn's Diseasea

CategoryGenesSignificancebAdjusted Significancec
ADHDd
Gene ontology term
    Regulation of chemotaxisCDH13, IL163.58E-054.40E-03
    Regulation of behaviorCDH13, IL165.25E-054.40E-03
    Regulation of response to external stimulusCDH13, IL162.77E-041.32E-02
    Integrin complexITGA11, ITGAE2.98E-041.32E-02
    Cell migrationCDH13, IL16, ITGA116.38E-042.34E-02
Diabetes type Ie
Gene ontology term
    Growth factor bindingERBB3, IL7R, INS4.46E-052.27E-02
    Regulation of secretionERBB3, INS3.35E-043.54E-02
    Anion transportC1QTNF6, CFTR, COL1A24.75E-043.54E-02
    Regulation of insulin receptor signalingINS9.76E-043.54E-02
    Tyrosine 3-monooxygenase activityTH9.76E-043.54E-02
Crohn's diseasef
Gene ontology term
    Protein kinase cascadeGCKR, LRRK2, MST1, STAT3, TNFSF155.10E-062.28E-03
    Intracellular signaling cascadeGCKR, LRRK2, MST1, NOD2, STAT3, TNFSF156.10E-042.42E-02
    Cell communicationATG16L1, BSN, GCKR, ITLN1, LRRK2, MST1, NOD2, PTPN22, STAT3, TNFSF157.95E-042.42E-02
    Protein amino acid autophosphorylationLRRK2, MST19.08E-042.42E-02
    Macrophage activation during immune responseSBNO29.76E-042.42E-02

a Enrichment analyses were conducted using the Biological Network Gene Ontology tool (33); genes from the reported GWAS (Table 1) encoding proteins that could be directly placed in the putative ADHD network are indicated in bold (further details are presented in Table 1 of the data supplement accompanying the online version of this article).

b Data were determined from a single test p value calculated using the hypergeometric test and taking into consideration both the total number of molecules from the analyzed data set and the total number of molecules linked to the same gene ontology term.

c Data were determined from multiple test-corrected p values using the Benjamini-Hochberg correction.

d The gene ontology terms regulation of chemotaxis, regulation of behavior, regulation of response to external stimulus, and cell migration can be further assigned to the biological process gene ontology subgroup/domain, while integrin complex falls under the cellular component domain.

e The gene ontology terms growth factor binding and tyrosine 3-monooxygenase activity can be further assigned to the molecular function gene ontology subgroup/domain, while regulation of secretion, anion transport, and regulation of insulin receptor signaling fall under the biological process domain.

f All gene ontology terms can be further assigned to the biological process gene ontology subgroup/domain.

TABLE 5. The Top Five Gene Functional Categories Significantly Enriched in the Top 20 Candidate Genes From Genome-Wide Association Studies of Attention Deficit Hyperactivity Disorder (ADHD), Diabetes Type I, and Crohn's Diseasea

Enlarge table

Discussion

In the present study, we used both bioinformatics tools and manual literature mining to attempt to integrate the top findings of the currently available GWAS data for ADHD. We found that 45 of the 85 top ADHD candidate genes (or approximately 53%) fit into a network that is involved in neurite outgrowth. It should be noted that our decision to only include SNPs within exonic or intronic regions of genes and those with association at p<1.00E-04 is essentially arbitrary. By only including genes with SNPs in exons and introns, we have ignored the possible role of upstream and downstream regulatory sequences (61), which could have increased the number of (potentially relevant) genes to be included in the study. However, we do not expect that this would have altered the general tendency in our results (i.e., the considerable overrepresentation of neurite outgrowth-related genes).

Surprisingly, the enrichment of neurite outgrowth-related genes was not reflected by the finding of gene functional categories and/or gene ontology terms directly related to neurite outgrowth with the bioinformatics tools used in this study (Table 2). This shows the limitations of the currently available bioinformatics software, which is a result of the current incompleteness of the annotation of the human genome.

Therefore, use of bioinformatics tools should always be accompanied by manual and systematic literature and database mining in order to fully understand the processes involved in the etiology of multifactorial disorders. Moreover, the enrichment of neurite outgrowth-related genes in the ADHD GWAS data (53%) should be viewed from the perspective that only 3% (N=576) of all currently annotated human genes (N=18,589) are involved in neurogenesis (gene ontology term: GO:0022008).

Several additional lines of evidence suggest that the identified network provides an important contribution to the etiology of ADHD. Apart from implicating the same genes as the GWASs, these other lines of evidence point to involvement of several other genes (and their corresponding proteins) that were not found (at the p<1.00E-04 level) in the GWASs but that functionally connect the genes from the GWAS findings.

First, nine proteins that act in the identified neurite outgrowth network are encoded by genes that were also found in copy number variations in people with ADHD (2022, 3541) (Table 3). Second, knockout mouse models of four network genes reveal ADHD-related behavior (4245). Third, and most importantly, the expression and/or function of a number of genes/proteins in the identified network is regulated by the stimulants methylphenidate (4850) and amphetamine (5155, 57), which are the most commonly used psychopharmacological treatments for ADHD symptoms (4, 5) and directly regulate neurite outgrowth (46, 47). The latter finding might not only increase our understanding of the working mechanism of these drugs but also provide directions for future research into new and more effective psychopharmacological ADHD treatments.

The fact that stimulants regulate neurite outgrowth potentially implicates a number of classic ADHD candidate genes in our network. Methylphenidate and amphetamine mediate effects on both dopaminergic and serotonergic neurotransmission in the brain (4, 5, 62). In this respect, the genes encoding the SLC6A3, DRD4, DRD5, SLC6A4, and HTR1B proteins could be putatively linked to the identified network. In addition, the link of these classic candidate genes with the neurite outgrowth network may help to localize the network to dopaminergic and serotonergic brain regions.

The findings from this study fit very well with literature about disturbed brain structure and function in ADHD patients (710) and, even more so, with recent reports describing aberrant structural and functional brain connectivity in these patients (1117). Because differences in brain connectivity have been shown to underlie interindividual variability in complex cognition-related processes, including ADHD (63, 64), relatively minor alterations in neurite outgrowth efficiency or direction may provide a major contribution to the cognitive deficits observed in ADHD.

A possibly important bias in our study may arise from the fact that the brain-expressed genes in the top findings of the ADHD GWASs are substantially larger than the average gene size in the human genome, and seven of the nine very large genes (i.e., genes with a size of >1 Mb) from the list of 85 ADHD genes indeed encode proteins that fit into the proposed ADHD network. It has been argued that the overrepresentation of large genes in the top findings of GWASs might result from bias because one is more likely to find an associated SNP in a larger, usually brain-expressed gene (58). If this is the case, it would be expected that similar results would be found in GWASs of polygenic disorders that are assumed not to primarily originate in the brain, such as diabetes mellitus and Crohn's disease. However, our Ingenuity and BiNGO analyses on the 20 top candidate genes from the GWASs for ADHD, diabetes mellitus type I, and Crohn's disease (Table 4, Table 5) show that this bias does not likely explain the findings for ADHD. In this respect, we would like to submit that the strong enrichment of neurite outgrowth-related genes in the top findings from the five ADHD GWASs, with 45 out of 85 genes fitting into our proposed network, also argues against the fact that these genes are spurious association findings. Nevertheless, we cannot completely exclude the possibility that some of the genes placed in the network still might have been chance findings.

Is the involvement of neurite outgrowth genes specific to the etiology of ADHD? This seems rather unlikely, since we recently also found neurite outgrowth to be implicated in dyslexia (65). Following an approach similar to the one used in the present study, we found that 10 of 14 dyslexia candidate genes reported to date (i.e., DCDC2, DIP2A, DOCK4, DYX1C1, FMR1, GTF2I, KIAA0319, KIAA0319L, ROBO1, S100B ) fit into a protein network that is involved in neurite outgrowth and neuronal migration. As shown in Figure 2 and explained in further detail in the data supplement, the proteins encoded by six of these genes (i.e., DCDC2, DYX1C1, FMR1, TFII-I [encoded by GTF2I], ROBO1, and S100B) fit directly into the identified network for ADHD. It is also interesting to note that decreased serum concentrations of S100B correlate with increasing hyperactivity symptoms in children with ADHD relative to normal comparison subjects (66) and that knockout mouse models of the FMR1 (67) and GTF2I (68) genes show ADHD-like behavior. One possible explanation for the observed genetic overlap between ADHD and dyslexia is that these neurodevelopmental disorders are highly comorbid and genetically correlated with one another (6971). The same also holds true for autism (72), and recent evidence suggests involvement of neurite outgrowth genes in this neurodevelopmental disorder as well (73). The different functional consequences of disturbed or abnormal neurite outgrowth in the different brain regions that are most affected in ADHD, dyslexia, and autism, respectively, could help to explain why a disturbance of the same neurodevelopmental process could lead to these partially overlapping but still distinct clinical phenotypes.

FIGURE 2.

FIGURE 2. Schematic Representation of How Six Neurite Outgrowth-Related Dyslexia Candidate Genes Can be Directly Placed in the Signaling Network for ADHDa

a The six neurite outgrowth-related dyslexia candidate genes (65) that can be placed in the signaling network for ADHD are indicated in orange. The genes/proteins that emerged from the five reported genome-wide association studies for ADHD are indicated in yellow. The proteins encoded by genes found in copy number variations in patients with ADHD are indicated by a blue border. The proteins encoded by genes implicated in the etiology of ADHD through gene knockout studies in mice are indicated by a green border. The genes/proteins of which the expression and/or function is regulated by stimulants are indicated by a red border. Evidence placing the six dyslexia candidate genes/proteins into the ADHD network is presented in the data supplement.

Several genes that are major players in the identified neurite outgrowth network have not been directly observed in the top findings of the GWASs. An example of such a gene is ERK1, which not only has an important function in the network but has also been implicated in ADHD through its location in ADHD-related copy number variations (22, 38) as well as the fact that Erk1-knockout mice display hyperactive behavior (43) and that amphetamine directly activates neuronal ERK1 (54). Other examples are CTNNB and CDC42. These and other genes from the network may be strong candidates for future association studies.

Lastly, our data make a compelling case for the use of polygenic types of association analyses to explain a higher percentage of the heritability of ADHD through the GWAS findings. In this respect, all the individually associated genes from the identified neurite outgrowth network could be fitted simultaneously into one polygenic risk test, as illustrated in the recent study by Yang et al. (74) of all SNPs that have been individually associated in GWASs of variation in human height as well as in two other recent studies applying similar approaches to GWAS data for schizophrenia (75) and functional gene groups for cognitive ability in ADHD (76).

In conclusion, by integrating the top findings of the five GWASs for ADHD with copy number variation data, (psychopharmacologically induced) expression data, and data from animal studies, we have identified a protein signaling network for ADHD that results in directed neurite outgrowth. Systems biology approaches like those used in this study are needed to yield genetic findings that are useful for clinical purposes, such as the prediction of disease prognosis and the identification of new treatment targets and strategies.

From the Department of Cognitive Neuroscience and Department of Psychiatry, Donders Institute for Brain, Cognition and Behaviour, and the Department of Human Genetics, Radboud University Nijmegen Medical Centre, Nijmegen, the Netherlands; and the Psychiatric and Neurodevelopmental Genetics Unit, Center for Human Genetic Research, Massachusetts General Hospital, Harvard Medical School, Boston.
Address correspondence and reprint requests to Dr. Franke,
Department of Human Genetics (855), Radboud University Nijmegen Medical Centre, P.O. Box 9101, 6500 HB Nijmegen, the Netherlands
; (e-mail).

Previously presented as a poster at the First International Attention Deficit Hyperactivity Disorder Congress of the European Network of Hyperkinetic Disorders (Eunethydis), Amsterdam, May 26–28, 2010 (where it won an award for best poster).

Received July 6, 2010; revisions received Aug. 20 and Oct. 19, 2010; accepted Oct. 29, 2010.

In the past 3 years, Dr. Buitelaar has been a consultant to, advisory board member of, and/or speaker for Bristol-Myers Squibb, Eli Lilly, Janssen-Cilag B.V., Medice, Organon/Shering Plough, Servier, Shire, and UCB. All other authors report no financial relationships with commercial interests.

The authors thank the families who made all the studies possible. The authors also thank the many investigators whose work drives the ADHD genetics field forward.

References

1. Polanczyk G , de Lima MS , Horta BL , Biederman J , Rohde LA: The worldwide prevalence of ADHD: a systematic review and metaregression analysis. Am J Psychiatry 2007; 164:942–948LinkGoogle Scholar

2. Kessler RC , Adler L , Barkley R , Biederman J , Conners CK , Demler O , Faraone SV , Greenhill LL , Howes MJ , Secnik K , Spencer T , Ustun TB , Walters EE , Zaslavsky AM: The prevalence and correlates of adult ADHD in the United States: results from the National Comorbidity Survey Replication. Am J Psychiatry 2006; 163:716–723LinkGoogle Scholar

3. Fayyad J , De Graaf R , Kessler R , Alonso J , Angermeyer M , Demyttenaere K , De Girolamo G , Haro JM , Karam EG , Lara C , Lepine JP , Ormel J , Posada-Villa J , Zaslavsky AM , Jin R: Cross-national prevalence and correlates of adult attention-deficit hyperactivity disorder. Br J Psychiatry 2007; 190:402–409Crossref, MedlineGoogle Scholar

4. Curatolo P , Paloscia C , D'Agati E , Moavero R , Pasini A: The neurobiology of attention deficit/hyperactivity disorder. Eur J Paediatr Neurol 2009; 13:299–304Crossref, MedlineGoogle Scholar

5. Wilens TE: Mechanism of action of agents used in attention-deficit/hyperactivity disorder. J Clin Psychiatry 2006; 67 (suppl 8):32–38Crossref, MedlineGoogle Scholar

6. Volkow ND , Wang GJ , Kollins SH , Wigal TL , Newcorn JH , Telang F , Fowler JS , Zhu W , Logan J , Ma Y , Pradhan K , Wong C , Swanson JM: Evaluating dopamine reward pathway in ADHD: clinical implications. JAMA 2009; 302:1084–1091Crossref, MedlineGoogle Scholar

7. Nopoulos P , Berg S , Castellenos FX , Delgado A , Andreasen NC , Rapoport JL: Developmental brain anomalies in children with attention-deficit hyperactivity disorder. J Child Neurol 2000; 15:102–108Crossref, MedlineGoogle Scholar

8. Valera EM , Faraone SV , Murray KE , Seidman LJ: Meta-analysis of structural imaging findings in attention-deficit/hyperactivity disorder. Biol Psychiatry 2007; 61:1361–1369Crossref, MedlineGoogle Scholar

9. Wolosin SM , Richardson ME , Hennessey JG , Denckla MB , Mostofsky SH: Abnormal cerebral cortex structure in children with ADHD. Hum Brain Mapp 2009; 30:175–184Crossref, MedlineGoogle Scholar

10. Shaw P , Lalonde F , Lepage C , Rabin C , Eckstrand K , Sharp W , Greenstein D , Evans A , Giedd JN , Rapoport J: Development of cortical asymmetry in typically developing children and its disruption in attention-deficit/hyperactivity disorder. Arch Gen Psychiatry 2009; 66:888–896Crossref, MedlineGoogle Scholar

11. Ashtari M , Kumra S , Bhaskar SL , Clarke T , Thaden E , Cervellione KL , Rhinewine J , Kane JM , Adesman A , Milanaik R , Maytal J , Diamond A , Szeszko P , Ardekani BA: Attention-deficit/hyperactivity disorder: a preliminary diffusion tensor imaging study. Biol Psychiatry 2005; 57:448–455Crossref, MedlineGoogle Scholar

12. Makris N , Buka SL , Biederman J , Papadimitriou GM , Hodge SM , Valera EM , Brown AB , Bush G , Monuteaux MC , Caviness VS , Kennedy DN , Seidman LJ: Attention and executive systems abnormalities in adults with childhood ADHD: A DT-MRI study of connections. Cereb Cortex 2008; 18:1210–1220Crossref, MedlineGoogle Scholar

13. Pavuluri MN , Yang S , Kamineni K , Passarotti AM , Srinivasan G , Harral EM , Sweeney JA , Zhou XJ: Diffusion tensor imaging study of white matter fiber tracts in pediatric bipolar disorder and attention-deficit/hyperactivity disorder. Biol Psychiatry 2009; 65:586–593Crossref, MedlineGoogle Scholar

14. Cao Q , Zang Y , Zhu C , Cao X , Sun L , Zhou X , Wang Y: Alerting deficits in children with attention deficit/hyperactivity disorder: event-related fMRI evidence. Brain Res 2008; 1219:159–168Crossref, MedlineGoogle Scholar

15. Wang L , Zhu C , He Y , Zang Y , Cao Q , Zhang H , Zhong Q , Wang Y: Altered small-world brain functional networks in children with attention-deficit/hyperactivity disorder. Hum Brain Mapp 2009; 30:638–649Crossref, MedlineGoogle Scholar

16. Castellanos FX , Kelly C , Milham MP: The restless brain: attention-deficit hyperactivity disorder, resting-state functional connectivity, and intrasubject variability. Can J Psychiatry 2009; 54:665–672Crossref, MedlineGoogle Scholar

17. Mazaheri A , Coffey-Corina S , Mangun GR , Bekker EM , Berry AS , Corbett BA: Functional disconnection of frontal cortex and visual cortex in attention-deficit/hyperactivity disorder. Biol Psychiatry 2010; 67:617–623Crossref, MedlineGoogle Scholar

18. Biederman J , Faraone SV: Attention-deficit hyperactivity disorder. Lancet 2005; 366:237–248Crossref, MedlineGoogle Scholar

19. Faraone SV , Perlis RH , Doyle AE , Smoller JW , Goralnick JJ , Holmgren MA , Sklar P: Molecular genetics of attention-deficit/hyperactivity disorder. Biol Psychiatry 2005; 57:1313–1323Crossref, MedlineGoogle Scholar

20. Elia J , Gai X , Xie HM , Perin JC , Geiger E , Glessner JT , D'arcy M , Deberardinis R , Frackelton E , Kim C , Lantieri F , Muganga BM , Wang L , Takeda T , Rappaport EF , Grant SF , Berrettini W , Devoto M , Shaikh TH , Hakonarson H , White PS: Rare structural variants found in attention-deficit hyperactivity disorder are preferentially associated with neurodevelopmental genes. Mol Psychiatry 2010; 15:637–646Crossref, MedlineGoogle Scholar

21. Lesch KP , Selch S , Renner TJ , Jacob C , Nguyen TT , Hahn T , Romanos M , Walitza S , Shoichet S , Dempfle A , Heine M , Boreatti-Hummer A , Romanos J , Gross-Lesch S , Zerlaut H , Wultsch T , Heinzel S , Fassnacht M , Fallgatter A , Allolio B , Schafer H , Warnke A , Reif A , Ropers HH , Ullmann R: Genome-wide copy number variation analysis in attention-deficit/hyperactivity disorder: association with neuropeptide Y gene dosage in an extended pedigree. Mol Psychiatry (Epub ahead of print, March 23, 2010)Google Scholar

22. Williams NM , Zaharieva I , Martin A , Langley K , Mantripragada K , Fossdal R , Stefansson H , Stefansson K , Magnusson P , Gudmundsson OO , Gustafsson O , Holmans P , Owen MJ , O'Donovan M , Thapar A: Rare chromosomal deletions and duplications in attention-deficit hyperactivity disorder: a genome-wide analysis. Lancet 2010; 376:1401–1408Crossref, MedlineGoogle Scholar

23. Zhou K , Dempfle A , Arcos-Burgos M , Bakker SC , Banaschewski T , Biederman J , Buitelaar J , Castellanos FX , Doyle A , Ebstein RP , Ekholm J , Forabosco P , Franke B , Freitag C , Friedel S , Gill M , Hebebrand J , Hinney A , Jacob C , Lesch KP , Loo SK , Lopera F , McCracken JT , McGough JJ , Meyer J , Mick E , Miranda A , Muenke M , Mulas F , Nelson SF , Nguyen TT , Oades RD , Ogdie MN , Palacio JD , Pineda D , Reif A , Renner TJ , Roeyers H , Romanos M , Rothenberger A , Schafer H , Sergeant J , Sinke RJ , Smalley SL , Sonuga-Barke E , Steinhausen HC , van der Meulen E , Walitza S , Warnke A , Lewis CM , Faraone SV , Asherson P: Meta-analysis of genome-wide linkage scans of attention deficit hyperactivity disorder. Am J Med Genet B Neuropsychiatr Genet 2008; 147B:1392–1398Crossref, MedlineGoogle Scholar

24. Amin N , Aulchenko YS , Dekker MC , Ferdinand RF , van Spreeken A , Temmink AH , Verhulst FC , Oostra BA , van Duijn CM: Suggestive linkage of ADHD to chromosome 18q22 in a young genetically isolated Dutch population. Eur J Hum Genet 2009; 17:958–966Crossref, MedlineGoogle Scholar

25. Gizer IR , Ficks C , Waldman ID: Candidate gene studies of ADHD: a meta-analytic review. Hum Genet 2009; 126:51–90Crossref, MedlineGoogle Scholar

26. Li D , Sham PC , Owen MJ , He L: Meta-analysis shows significant association between dopamine system genes and attention deficit hyperactivity disorder (ADHD). Hum Mol Genet 2006; 15:2276–2284Crossref, MedlineGoogle Scholar

27. Neale BM , Lasky-Su J , Anney R , Franke B , Zhou K , Maller JB , Vasquez AA , Asherson P , Chen W , Banaschewski T , Buitelaar J , Ebstein R , Gill M , Miranda A , Oades RD , Roeyers H , Rothenberger A , Sergeant J , Steinhausen HC , Sonuga-Barke E , Mulas F , Taylor E , Laird N , Lange C , Daly M , Faraone SV: Genome-wide association scan of attention deficit hyperactivity disorder. Am J Med Genet B Neuropsychiatr Genet 2008; 147B:1337–1344Crossref, MedlineGoogle Scholar

28. Lasky-Su J , Neale BM , Franke B , Anney RJ , Zhou K , Maller JB , Vasquez AA , Chen W , Asherson P , Buitelaar J , Banaschewski T , Ebstein R , Gill M , Miranda A , Mulas F , Oades RD , Roeyers H , Rothenberger A , Sergeant J , Sonuga-Barke E , Steinhausen HC , Taylor E , Daly M , Laird N , Lange C , Faraone SV: Genome-wide association scan of quantitative traits for attention deficit hyperactivity disorder identifies novel associations and confirms candidate gene associations. Am J Med Genet B Neuropsychiatr Genet 2008; 147B:1345–1354Crossref, MedlineGoogle Scholar

29. Lesch KP , Timmesfeld N , Renner TJ , Halperin R , Roser C , Nguyen TT , Craig DW , Romanos J , Heine M , Meyer J , Freitag C , Warnke A , Romanos M , Schafer H , Walitza S , Reif A , Stephan DA , Jacob C: Molecular genetics of adult ADHD: converging evidence from genome-wide association and extended pedigree linkage studies. J Neural Transm 2008; 115:1573–1585Crossref, MedlineGoogle Scholar

30. Mick E , Todorov A , Smalley S , Hu X , Loo S , Todd RD , Biederman J , Byrne D , Dechairo B , Guiney A , McCracken J , McGough J , Nelson SF , Reiersen AM , Wilens TE , Wozniak J , Neale BM , Faraone SV: Family-based genome-wide association scan of attention-deficit/hyperactivity disorder. J Am Acad Child Adolesc Psychiatry 2010; 49:898–905Crossref, MedlineGoogle Scholar

31. Neale BM , Medland S , Ripke S , Anney RJ , Asherson P , Buitelaar J , Franke B , Gill M , Kent L , Holmans P , Middleton F , Thapar A , Lesch KP , Faraone SV , Daly M , Nguyen TT , Schäfer H , Steinhausen HC , Reif A , Renner TJ , Romanos M , Romanos J , Warnke A , Walitza S , Freitag C , Meyer J , Palmason H , Rothenberger A , Hawi Z , Sergeant J , Roeyers H , Mick E , Biederman J; IMAGE II Consortium Group: Case control genome-wide association of attention deficit/hyperactivity disorder. J Am Acad Child Adolesc Psychiatry 2010; 49:906–920Crossref, MedlineGoogle Scholar

32. Franke B , Neale BM , Faraone SV: Genome-wide association studies in ADHD. Hum Genet 2009; 126:13–50Crossref, MedlineGoogle Scholar

33. Maere S , Heymans K , Kuiper M: BiNGO: a Cytoscape plugin to assess overrepresentation of gene ontology categories in biological networks. Bioinformatics 2005; 21:3448–3449Crossref, MedlineGoogle Scholar

34. UniProt Consortium: The Universal Protein Resource (UniProt) in 2010. Nucleic Acids Res 2010; 38:D142–D148Crossref, MedlineGoogle Scholar

35. Bradley WE , Raelson JV , Dubois DY , Godin E , Fournier H , Prive C , Allard R , Pinchuk V , Lapalme M , Paulussen RJ , Belouchi A: Hotspots of large rare deletions in the human genome. PLoS One 2010; 5:e9401Crossref, MedlineGoogle Scholar

36. Baker KL , Rees MI , Thompson PW , Howell RT , Cole TR , Houghes HE , Upadhyaya M , Ravine D: Chromosome 2 interstitial deletion (del[2][q14.1q21]) associated with connective tissue laxity and an attention deficit disorder. J Med Genet 2001; 38:493–496Crossref, MedlineGoogle Scholar

37. Cappellacci S , Martinelli S , Rinaldi R , Martinelli E , Parisi P , Mancini B , Pescosolido R , Grammatico P: De novo pure 12q22q24.33 duplication: first report of a case with mental retardation, ADHD, and Dandy-Walker malformation. Am J Med Genet A 2006; 140:1203–1207Crossref, MedlineGoogle Scholar

38. Shinawi M , Liu P , Kang SH , Shen J , Belmont JW , Scott DA , Probst FJ , Craigen WJ , Graham B , Pursley A , Clark G , Lee J , Proud M , Stocco A , Rodriguez D , Kozel B , Sparagana S , Roeder E , McGrew S , Kurczynski T , Allison L , Amato S , Savage S , Patel A , Stankiewicz P , Beaudet A , Cheung SW , Lupski JR: Recurrent reciprocal 16p11.2 rearrangements associated with global developmental delay, behavioral problems, dysmorphism, epilepsy, and abnormal head size. J Med Genet 2010; 47:332–341Crossref, MedlineGoogle Scholar

39. van den Berg L , de Waal HD , Han JC , Ylstra B , Eijk P , Nesterova M , Heutink P , Stratakis CA: Investigation of a patient with a partial trisomy 16q including the fat mass and obesity associated gene (FTO): fine mapping and FTO gene expression study. Am J Med Genet A 2010; 152A:630–637Crossref, MedlineGoogle Scholar

40. Ceccarini C , Sinibaldi L , Bernardini L , De Simone R , Mingarelli R , Novelli A , Dallapiccola B: Duplication 18q21.31-q22.2. Am J Med Genet A 2007; 143:343–348Crossref, MedlineGoogle Scholar

41. Lalani SR , Thakuria JV , Cox GF , Wang X , Bi W , Bray MS , Shaw C , Cheung SW , Chinault AC , Boggs BA , Ou Z , Brundage EK , Lupski JR , Gentile J , Waisbren S , Pursley A , Ma L , Khajavi M , Zapata G , Friedman R , Kim JJ , Towbin JA , Stankiewicz P , Schnittger S , Hansmann I , Ai T , Sood S , Wehrens XH , Martin JF , Belmont JW , Potocki L: 20p12.3 microdeletion predisposes to Wolff-Parkinson-White syndrome with variable neurocognitive deficits. J Med Genet 2009; 46:168–175Crossref, MedlineGoogle Scholar

42. Tanda K , Nishi A , Matsuo N , Nakanishi K , Yamasaki N , Sugimoto T , Toyama K , Takao K , Miyakawa T: Abnormal social behavior, hyperactivity, impaired remote spatial memory, and increased D1-mediated dopaminergic signaling in neuronal nitric oxide synthase knockout mice. Mol Brain 2009; 2:19Crossref, MedlineGoogle Scholar

43. Engel SR , Creson TK , Hao Y , Shen Y , Maeng S , Nekrasova T , Landreth GE , Manji HK , Chen G: The extracellular signal-regulated kinase pathway contributes to the control of behavioral excitement. Mol Psychiatry 2009; 14:448–461Crossref, MedlineGoogle Scholar

44. Pangratz-Fuehrer S , Bubna-Littitz H , Propst F , Reitsamer H: Mice deficient in microtubule-associated protein MAP1B show a distinct behavioral phenotype and altered retina function. Behav Brain Res 2005; 164:188–196Crossref, MedlineGoogle Scholar

45. Deiss V , Strazielle C , Lalonde R: Regional brain variations of cytochrome oxidase activity and motor co-ordination in staggerer mutant mice. Neuroscience 2000; 95:903–911Crossref, MedlineGoogle Scholar

46. Lipton JW , Tolod EG , Thompson VB , Pei L , Paumier KL , Terpstra BT , Lynch KA , Collier TJ , Sortwell CE: 3,4-Methylenedioxy-N-methamphetamine (ecstasy) promotes the survival of fetal dopamine neurons in culture. Neuropharmacology 2008; 55:851–859Crossref, MedlineGoogle Scholar

47. Park YH , Kantor L , Guptaroy B , Zhang M , Wang KK , Gnegy ME: Repeated amphetamine treatment induces neurite outgrowth and enhanced amphetamine-stimulated dopamine release in rat pheochromocytoma cells (PC12 cells) via a protein kinase C- and mitogen activated protein kinase-dependent mechanism. J Neurochem 2003; 87:1546–1557Crossref, MedlineGoogle Scholar

48. Appenrodt E , Schwarzberg H: Methylphenidate-induced motor activity in rats: modulation by melatonin and vasopressin. Pharmacol Biochem Behav 2003; 75:67–73Crossref, MedlineGoogle Scholar

49. Andersen SL , Arvanitogiannis A , Pliakas AM , LeBlanc C , Carlezon WA: Altered responsiveness to cocaine in rats exposed to methylphenidate during development. Nat Neurosci 2002; 5:13–14Crossref, MedlineGoogle Scholar

50. Adriani W , Leo D , Guarino M , Natoli A , Di Consiglio E , De Angelis G , Traina E , Testai E , Perrone-Capano C , Laviola G: Short-term effects of adolescent methylphenidate exposure on brain striatal gene expression and sexual/endocrine parameters in male rats. Ann N Y Acad Sci 2006; 1074:52–73Crossref, MedlineGoogle Scholar

51. Yetnikoff L , Labelle-Dumais C , Flores C: Regulation of netrin-1 receptors by amphetamine in the adult brain. Neuroscience 2007; 150:764–773Crossref, MedlineGoogle Scholar

52. Mlewski EC , Krapacher FA , Ferreras S , Paglini G: Transient enhanced expression of CDK5 activator p25 after acute and chronic d-amphetamine administration. Ann N Y Acad Sci 2008; 1139:89–102Crossref, MedlineGoogle Scholar

53. Alimohamad H , Sutton L , Mouyal J , Rajakumar N , Rushlow WJ: The effects of antipsychotics on beta-catenin, glycogen synthase kinase-3 and dishevelled in the ventral midbrain of rats. J Neurochem 2005; 95:513–525Crossref, MedlineGoogle Scholar

54. Choe ES , Wang JQ: CaMKII regulates amphetamine-induced ERK1/2 phosphorylation in striatal neurons. Neuroreport 2002; 13:1013–1016Crossref, MedlineGoogle Scholar

55. Bhattacharjee AK , Chang L , White L , Bazinet RP , Rapoport SI: D-amphetamine stimulates D2 dopamine receptor-mediated brain signaling involving arachidonic acid in unanesthetized rats. J Cereb Blood Flow Metab 2006; 26:1378–1388Crossref, MedlineGoogle Scholar

56. Schuchardt JP , Huss M , Stauss-Grabo M , Hahn A: Significance of long-chain polyunsaturated fatty acids (PUFAs) for the development and behaviour of children. Eur J Pediatr 2010; 169:149–164Crossref, MedlineGoogle Scholar

57. Jones DC , Kuhar MJ: Cocaine-amphetamine-regulated transcript expression in the rat nucleus accumbens is regulated by adenylyl cyclase and the cyclic adenosine 5′-monophosphate/protein kinase: a second messenger system. J Pharmacol Exp Ther 2006; 317:454–461Crossref, MedlineGoogle Scholar

58. Holmans P , Green EK , Pahwa JS , Ferreira MA , Purcell SM , Sklar P , Owen MJ , O'Donovan MC , Craddock N: Gene ontology analysis of GWA study data sets provides insights into the biology of bipolar disorder. Am J Hum Genet 2009; 85:13–24Crossref, MedlineGoogle Scholar

59. Strachan T , Read AP: Human Molecular Genetics. New York, Garland Publishing, 2004, p 253Google Scholar

60. Brown MD , Cornejo BJ , Kuhn TB , Bamburg JR: CDC42 stimulates neurite outgrowth and formation of growth cone filopodia and lamellipodia. J Neurobiol 2000; 43:352–364Crossref, MedlineGoogle Scholar

61. Veyrieras JB , Kudaravalli S , Kim SY , Dermitzakis ET , Gilad Y , Stephens M , Pritchard JK: High-resolution mapping of expression-QTLs yields insight into human gene regulation. PLoS Genet 2008; 4:e1000214Crossref, MedlineGoogle Scholar

62. Han DD , Gu HH: Comparison of the monoamine transporters from human and mouse in their sensitivities to psychostimulant drugs. BMC Pharmacol 2006; 6:6Crossref, MedlineGoogle Scholar

63. Chiang MC , Barysheva M , Shattuck DW , Lee AD , Madsen SK , Avedissian C , Klunder AD , Toga AW , McMahon KL , de Zubicaray GI , Wright MJ , Srivastava A , Balov N , Thompson PM: Genetics of brain fiber architecture and intellectual performance. J Neurosci 2009; 29:2212–2224Crossref, MedlineGoogle Scholar

64. Valera EM , Brown A , Biederman J , Faraone SV , Makris N , Monuteaux MC , Whitfield-Gabrieli S , Vitulano M , Schiller M , Seidman LJ: Sex differences in the functional neuroanatomy of working memory in adults with ADHD. Am J Psychiatry 2010; 167:86–94LinkGoogle Scholar

65. Poelmans G , Buitelaar JK , Pauls DL , Franke B: A theoretical molecular network for dyslexia: integrating available genetic findings. Mol Psychiatry 2010 (Epub ahead of print October 19, 2010)MedlineGoogle Scholar

66. Oades RD , Myint AM , Dauvermann MR , Schimmelmann BG , Schwarz MJ: Attention-deficit hyperactivity disorder (ADHD) and glial integrity: an exploration of associations of cytokines and kynurenine metabolites with symptoms and attention. Behav Brain Funct 2010; 6:32Crossref, MedlineGoogle Scholar

67. Moon J , Beaudin AE , Verosky S , Driscoll LL , Weiskopf M , Levitsky DA , Crnic LS , Strupp BJ: Attentional dysfunction, impulsivity, and resistance to change in a mouse model of fragile X syndrome. Behav Neurosci 2006; 120:1367–1379Crossref, MedlineGoogle Scholar

68. Lucena J , Pezzi S , Aso E , Valero MC , Carreiro C , Dubus P , Sampaio A , Segura M , Barthelemy I , Zindel MY , Sousa N , Barbero JL , Maldonado R , Perez-Jurado LA , Campuzano V: Essential role of the N-terminal region of TFII-I in viability and behavior. BMC Med Genet 2010; 1161Google Scholar

69. Stevenson J , Pennington BF , Gilger JW , DeFries JC , Gillis JJ: Hyperactivity and spelling disability: testing for shared genetic aetiology. J Child Psychol Psychiatry 1993; 34:1137–1152Crossref, MedlineGoogle Scholar

70. Light JG , Pennington BF , Gilger JW , DeFries JC: Reading disability and hyperactivity: evidence for a common genetic etiology. Dev Neuropsychol 1995; 11:323–335CrossrefGoogle Scholar

71. Chadwick O , Taylor E , Taylor A , Heptinstall E , Danckaerts M: Hyperactivity and reading disability: a longitudinal study of the nature of the association. J Child Psychol Psychiatry 1999; 40:1039–1050Crossref, MedlineGoogle Scholar

72. Rommelse NN , Franke B , Geurts HM , Hartman CA , Buitelaar JK: Shared heritability of attention-deficit/hyperactivity disorder and autism spectrum disorder. Eur Child Adolesc Psychiatry 2010; 19:281–295Crossref, MedlineGoogle Scholar

73. Piton A , Gauthier J , Hamdan FF , Lafreniere RG , Yang Y , Henrion E , Laurent S , Noreau A , Thibodeau P , Karemera L , Spiegelman D , Kuku F , Duguay J , Destroismaisons L , Jolivet P , Cote M , Lachapelle K , Diallo O , Raymond A , Marineau C , Champagne N , Xiong L , Gaspar C , Riviere JB , Tarabeux J , Cossette P , Krebs MO , Rapoport JL , Addington A , Delisi LE , Mottron L , Joober R , Fombonne E , Drapeau P , Rouleau GA: Systematic resequencing of X-chromosome synaptic genes in autism spectrum disorder and schizophrenia. Mol Psychiatry 2010 (Epub ahead of print, May 18, 2010)Google Scholar

74. Yang J , Benyamin B , McEvoy BP , Gordon S , Henders AK , Nyholt DR , Madden PA , Heath AC , Martin NG , Montgomery GW , Goddard ME , Visscher PM: Common SNPs explain a large proportion of the heritability for human height. Nat Genet 2010; 42:565–569Crossref, MedlineGoogle Scholar

75. Purcell SM , Wray NR , Stone JL , Visscher PM , O'Donovan MC , Sullivan PF , Sklar P: Common polygenic variation contributes to risk of schizophrenia and bipolar disorder. Nature 2009; 460:748–752Crossref, MedlineGoogle Scholar

76. Ruano D , Abecasis GR , Glaser B , Lips ES , Cornelisse LN , de Jong AP , Evans DM , Davey SG , Timpson NJ , Smit AB , Heutink P , Verhage M , Posthuma D: Functional gene group analysis reveals a role of synaptic heterotrimeric G proteins in cognitive ability. Am J Hum Genet 2010; 86:113–125Crossref, MedlineGoogle Scholar