Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Dec 1.
Published in final edited form as: J Psychosom Res. 2009 Sep 30;67(6):533–545. doi: 10.1016/j.jpsychores.2009.06.006

The Genetics of Tourette Syndrome: A review

Julia A O’Rourke 1,2, Jeremiah M Scharf 1,2, Dongmei Yu 1,2, David L Pauls 1,2
PMCID: PMC2778609  NIHMSID: NIHMS149658  PMID: 19913658

Abstract

Objectives

This paper summarizes and evaluates recent advances in the genetics of Gilles de la Tourette Syndrome (GTS).

Methods

This is a review of recent literature focusing on: 1) the genetic etiology of GTS; 2) common genetic components of GTS, Attention Deficit Hyperactivity Disorder (ADHD), and Obsessive Compulsive Disorder (OCD); 3) recent linkage studies of GTS; 4) chromosomal translocations in GTS; and 5) candidate gene studies.

Results

Family, twin and segregation studies provide strong evidence for the genetic nature of GTS. GTS is a heterogeneous disorder with complex inheritance patterns and phenotypic manifestations. Family studies of GTS and OCD indicate that an early onset form of OCD is likely to share common genetic factors with GTS. While there apparently is an etiological relationship between GTS and ADHD, it appears that the common form of ADHD does not share genetic factors with GTS. The largest genome wide linkage study to date observed evidence for linkage on chromosome 2p23.2 (P=3.8 × 10−5). No causative candidate genes have been identified, and recent studies suggest that the newly identified candidate gene SLITRK1 is not a major risk gene for the majority of individuals with GTS.

Conclusion

The genetics of GTS are complex and not well understood. The Genome Wide Association Study (GWAS) design can hopefully overcome the limitations of linkage and candidate gene studies. However, large-scale collaborations are needed to provide enough power to utilize the GWAS design for discovery of causative mutations. Knowledge of susceptibility mutations and biological pathways involved should eventually lead to new treatment paradigms for GTS.

Keywords: Tourette’s Disorder, Genetics, Family study, Review


In the original description of the syndrome that bears his name, Georges Gilles de la Tourette observed that the disorder was familial [1]. Subsequently, there has been considerable research devoted to systematically examining whether that original observation could be replicated and whether the observed familiality is due in part to genetic factors. These studies have included family studies, twin studies, genetic linkage studies and genetic association (candidate gene) studies. This review is based on the comprehensive literature search of Pubmed database with keywords such as Gilles de la Tourette, Tourette Syndrome, Tourette Disorder, GTS, ADHD, and OCD.

Family Studies

Family studies have repeatedly demonstrated that Gilles de la Tourette Syndrome (GTS) is highly familial. Establishing that there is familial aggregation does not “prove” that the disorder is influenced only by genetic factors, since family members also share common environmental factors. Nevertheless, results from these studies provide an important first step for determining whether genetic factors are important in the manifestation of the condition. Since first degree relatives share on average 50% of their genetic material, it is expected that the first degree relatives of an individual affected with a genetic disorder will have a greater chance of also being affected with that disorder compared to the general population [2]. Results from GTS family studies consistently show a 10 to 100 fold increase in the rates of GTS in first-degree relatives when compared to those rates in the general population [312] making it one of the most heritable childhood onset neuropsychiatric conditions [13]. Furthermore, chronic tics (CT) also occur significantly more frequently (reports range from 7 to 22%) among first degree relatives of GTS probands compared to relatives of control probands, suggesting that CT is a manifestation of the same underlying genetic susceptibility as GTS [312].

The vast majority of clinically referred GTS individuals have comorbid psychiatric conditions; in fact only about 10–15% of individuals with GTS have no comorbidities [14, 15]. Obsessive Compulsive Disorder (OCD) and Attention Deficit Hyperactivity Disorder (ADHD) are the most common GTS-related comorbidities, each affecting over 50% of GTS individuals [11, 14, 16, 17]. These rates are significantly greater than those in the general population (e.g., 1.1 to 3.3% for OCD [18, 19], and 5% for ADHD in school age children [20]), suggesting that these disorders may also have some shared underlying etiological mechanisms with GTS. Understanding the etiology of comorbid GTS conditions is especially critical, as children with comorbid diagnoses exhibit greater psychopathological burden [21], have more severe tic symptoms [15], and are at higher risk for aggressive behaviors and frequent anger outbursts [14, 22, 23].

In addition to demonstrating that GTS is familial, the family study design can be utilized to examine the familial relationship between GTS and its comorbid disorders. That is, if the hypothesis of shared genes is correct, then first degree relatives of a proband with GTS will have an increased risk of the comorbid disorder when compared to the general population.

Examining the relationship between GTS and OCD

OCD is a common psychiatric disorder, with both childhood and adult onset forms, and is characterized by intrusive thoughts and images (i.e., obsessions) and ritualized repetitive behaviors (i.e., compulsions). Twin and family studies have demonstrated that there is a significant genetic component in the etiology of OCD, especially in the childhood onset form [24].

Increased rates of OCD have been reported in family studies of GTS-ascertained probands. Pauls and colleagues reported increased rates of OCD without tics among relatives of GTS probands even when the probands did not have OCD themselves [25]. In a larger follow-up study, Pauls and colleagues observed that the gender of the proband did not affect the rate of either OCD or tics in the relatives [6]. However, the gender of the relative was associated with the risk of GTS and OCD, such that female relatives were more likely to develop OCD without tics, while male relatives were more likely to develop tic disorders [6]. Subsequent family studies support this finding [79, 26, 27], providing evidence that OCD may be a sex-influenced phenotype of GTS or CT. Thus it appears that GTS, CT, and some forms of OCD are likely to have a common underlying susceptibility.

Results from family studies of OCD-ascertained families have also supported this conclusion, though the relationship is more complex. It has been observed that childhood onset OCD in probands is associated with greater rates of GTS and tics in their relatives [28, 29]. Furthermore, individuals with early onset of OCD appear more likely to have a tic disorder than individuals with adult-onset OCD [28, 30], and relatives of female OCD probands have a greater risk of developing tic disorders [28]. These investigators have also observed a much higher rate of GTS and tics among relatives of OCD probands who had family histories of OCD (7.4% vs. 1.4%) compared to relatives of probands without a family history of OCD [28], suggesting that non-familial OCD is not associated with tic disorders. However, in contrast to the family studies of GTS summarized above in which all GTS families had higher rates of OCD regardless of whether the proband had GTS and OCD, some OCD family studies have demonstrated that OCD probands with tics have higher rates of tics in first degree relatives (10.6%), compared with only 3.2% of the relatives of OCD probands without tics [28], though this finding was not fully replicated in a subsequent sample [31]. Thus results from OCD family studies suggest that there may be different types of OCD (adult onset and childhood-onset), with only the childhood-onset form likely to share common genetic factors with GTS. However it is still unclear whether there is an additional subtype of early onset OCD that is unrelated to tics.

Examining the relationship between GTS and ADHD

ADHD is one of the most common childhood-onset developmental disorders [20], characterized by hyperactivity, impulsivity, and inattentiveness. ADHD has a significant genetic component, with 70–80% of risk believed to be due to genetic causes [32]. The familial relationship between GTS and ADHD is not fully understood, but it is becoming clear that, unlike the relationship between GTS and OCD, the most common form of ADHD does not appear to have a shared underlying susceptibility with GTS. Comings and Comings [33] proposed that ADHD represented a variant expression of the same genetic factors responsible for GTS [33, 34]. However, several later studies did not find support for this hypothesis [2, 12, 35]. While an increased rate of ADHD has been observed among the first degree relatives of GTS probands, even when the probands themselves did not have ADHD, that increase appears to be due to an increased rate of comorbid GTS+ADHD among those relatives who also had a diagnosis of GTS. In other words, rates of ADHD alone were not elevated in the relatives of GTS probands who did not themselves also have ADHD, suggesting that the “pure” form of ADHD is not present at an increased frequency in these families [2, 12, 35], but instead exists only when these disorders co-occur in the same individual. Recent work [36] demonstrated that a comorbid diagnosis of GTS and ADHD in a relative of a GTS or ADHD proband was strongly associated with an OCD diagnosis of that proband, and that comorbid GTS and ADHD diagnoses in a relative was associated with some degree of OC symptoms in the same individual [36]. These results suggest that there may be a GTS/OCD/ADHD familial subtype, which might be associated with an increased genetic burden and could represent a more severe form of GTS [37].

Several observations support this hypothesis. First, it has been reported that comorbid conditions are more frequent in GTS than in other less severe tic disorders. In a recent Swedish study of school-aged children, 66% of the GTS cases had comorbid ADHD compared to 33% of children with chronic vocal tics, 12% of children with chronic motor tics, and just 4% of children with transient motor tics [17]. A similar gradient was observed in the rates of OCD, as well as in the total number of comorbid disorders [17]. Another study found 44% of children with GTS had comorbid ADHD, compared with 23% of children with chronic tics, and 54% of children with GTS had comorbid OCD, compared with 8% of children with chronic tics [11]. Second, an increased genetic burden appears to influence the risk of GTS and comorbid disorders. In a prospective study of children at risk for GTS, children of two GTS-affected parents had three times greater risk of developing ADHD compared to the children of one affected parent, and two times greater risk of developing either tics, ADHD, or OCD [38]. Third, there is some evidence that comorbid GTS/OCD/ADHD may be heritable. A recent latent class analysis of 952 individuals from 222 GTS families was performed to identify GTS subphenotypes based on diagnoses of GTS, OCD, OC symptoms and ADHD [39]. The investigators identified five classes of categorical GTS subphenotypes, of which only the comorbid GTS/OCD/ADHD class was highly heritable [39]. In addition, 34% of all GTS affected individuals had comorbid OCD and ADHD, while only 10% had comorbid ADHD without OCD [39]. Another study of almost 6,000 GTS affected individuals also found a significant increase in OCD and other psychiatric disorders in individuals with comorbid GTS and ADHD compared to individuals with GTS without ADHD [16]. Finally, individuals with comorbid conditions may exhibit a specific subset of symptoms. The evaluation of GTS symptoms in 410 GTS patients utilizing principal component analysis found that individuals who have GTS comorbid with ADHD or OCD are more likely to exhibit socially inappropriate behaviors and other complex vocal tics [40]. Further family studies of GTS–ascertained probands which would include GTS only, OCD only, ADHD only, comorbid GTS and ADHD, comorbid GTS and OCD, and comorbid GTS/OCD/ADHD probands and their relatives would help in understanding of the genetics of GTS and its comorbid disorders.

Segregation Analyses

Results from segregation analyses of family studies have been consistent with the hypothesis that GTS is genetically transmitted [7, 8, 26, 4149]. The majority of these studies support the hypothesis of at least one genetic locus with major effect, though in retrospect each indicate the likelihood of many additional genetic loci and/or the presence of genetic heterogeneity [7, 8, 26, 42, 44, 4649]. Several investigators have observed bilineal transmission [27, 5052] in a number of GTS and CT families, raising the possibility of nonrandom selection of partners for marriage (assortative mating). This fact further complicates the interpretation of segregation analyses, since the majority of these studies were performed under the assumption of random mating.

Twin Studies

Twin studies provide strong evidence for the genetic nature of GTS. The largest study included 30 monozygotic and 13 dizygotic pairs of twins [53]. These investigators utilized phone-based assessment and found that 77% of monozygotic twins were concordant for tic disorders (CT or GTS), but only 23% of dizygotic twins were concordant for these disorders [53]. Furthermore, the concordance rate of monozygotic twins reached 100% for GTS or CT when direct observational interviews were conducted [47]. In a second smaller study of 16 pairs of monozygotic twins, 56% of MZ twins were concordant for GTS and 94% were concordant for tic disorders [54]. High concordance rates for tic disorders in monozygotic twins suggest that GTS is a genetic disorder and that CMT and GTS are genetically related.

Linkage Analyses

Five genome wide linkage analyses have been performed to date [5559]. The Tourette Syndrome Association International Consortium for Genetics (TSAICG) has conducted the largest of these genetic linkage studies. Their sample represents a joint analysis of most of the individuals contained in the four previous studies and consists of 238 nuclear families and 18 large multigenerational families totaling 2,040 individuals [59]. Both parametric and non-parametric analyses were performed using two diagnostic classifications: 1) GTS alone; 2) combination of either GTS or CT. Strong evidence for linkage was observed in the multigenerational families for markers on chromosome 2p23.2 using the combined phenotype of GTS or CT, with suggestive evidence for linkage on chromosomes 5p, and 6p. A combined analysis including both affected sib pair and multigenerational families showed a slightly increased linkage signal on 2p, and the fine mapping of the 15 centiMorgan (cM) critical region yielded a signal with an empirical p=3.8 × 10−5.

Subsequent linkage analyses of the large families provide additional evidence that GTS may be genetically heterogeneous. A heat map, shown in Figure 1, indicates the individual family Z scores on chromosome 2 for each of fifteen multigenerational families. Most of the families have positive linkage signals on chromosome 2p; however comparison across families indicates the absence of linkage in some families (Family 5, and likely Families 1 and 6). Furthermore, some families appear to have strong (Z score >3.719) linkage signals further along chromosome 2: Family 11 at D2S2216, Family 14 at D2S352 and D2S2368, and Family 15 at D2S2259. Thus, heterogeneity could partially explain the inconsistent results from previous linkage studies, as recently reviewed by Scharf and Pauls [60].

Figure 1.

Figure 1

Heat map of chromosome 2 linkage analysis from 15 large multigenerational pedigrees [59].

X = −log(P) if Z>0 and X = log(1−P) if Z<0

Positive Z scores indicate an increased likelihood of linkage, while negative Z scores indicate a decreased likelihood of linkage.

No single locus shows moderate or significant linkage across all pedigrees, suggesting that there is genetic heterogeneity of GTS, although low positive Z scores can also indicate the lack of power in a particular family to reach a significant threshold for linkage. The most significant overall linkage at D2S319 (Z score >3.719) resulted from one pedigree with a high Z score (3.09~3.719), and 8 pedigrees with moderate Z scores (2.326~3.09). The adjacent marker D2S2211 has only a moderately significant overall Z score because fewer pedigrees with moderate Z scores (2.326~3.09) contribute.

Furthermore the map shows three different pedigrees with strong evidence for linkage (Z score >3.719) at adjacent loci on chromosome 2: Family 11 at D2S2216, Family 14 at D2S352 and D2S2368, and Family 15 at D2S2259. However, the summed Z scores for these markers across all 15 pedigrees are only moderately positive, indicating that in the majority of tested pedigrees these loci either do not have enough power to detect linkage or do not contribute to the susceptibility to GTS.

Chromosomal Translocations

Another valuable approach for identifying disease genes is the identification of chromosomal aberrations in patients. Translocations, duplications or deletions of large chromosome segments can be visualized by karyotyping or fluorescence in situ hybridization (FISH). Newer methods for detection of duplications or deletions ranging from ~1kb to several Mb in size, referred to as copy number variations (CNVs), have identified de novo and inherited CNVs associated with risk of many neuropsychiatric disorders [6165].

The only study that systematically examined structural variation of chromosomes in GTS has karyotyped 68 consecutive patients and has identified one individual with XYY chromosome, and two individuals with normal heterochromatin variations on chromosomes 1 and 9 [66]. Since karyotyping can only detect duplications and deletions larger than 5 megabases, the rate of structural abnormalities could be significantly greater if newer techniques such as chromosomal microarrays (CMA) are utilized. In fact, a recent study of children with Autism Spectrum Disorders has detected abnormal karyotype in 2.2% of patients, and CMA identified deletions or duplications in 18.2% of patients [67]. Although, less than a half of the detected CMA deletions or duplications could be considered as abnormal CNV (variants associated with known genetic disorders), possibly reflecting normal structural variations in the genome [68].

Three chromosomal regions (7q22-q31, 8q13-q22, and 18q22) have been reported in the literature in at least two independent translocations and have been observed to co-segregate with GTS, CT, or OCD in several family members. In addition, two independent cases of translocation and deletion of 17p11 in GTS affected individuals were recently reported [69, 70].

7q22-q31: Three groups have reported rearrangements on 7q22-q31: one large family consisted of nine individuals with a balanced translocation t(7;18)(q22-31;q22), in which six individuals were affected with GTS or CT [71]; a single GTS patient had a duplication of 7q22-q31 [72]. Lastly, a third GTS patient was reported with a de novo duplication of the long arm of chromosome 7 [46, XY, dup(7)(q22.1-q31.1)] [73]. Mapping of the breakpoints implicated disruption of the inner mitochondrial membrane peptidase 2-like (IMMP2L); however recent screening of 39 GTS individuals did not find any coding mutations [74].

8q13-8q22: Three separate instances of a translocation breakpoint on 8q have been described in GTS: two isolated cases with breakpoints on 8q13 [75] and a the third which included a father and six children with a t(1;8)(q21;q22) translocation [76], with four children reported to have GTS or tics.

18q22: Four independent translocations or deletions involving 18q22 have been reported: an 18q22 deletion [77], an 18q21-q22 inversion [78], a t(2;18)(p12;q22) translocation [79]; and the t(7;18)(q22-31;q22) translocation [71].

Two children with GTS, OCD, and mental retardation had an insertion of 2p21-23 within 7q35-36, resulting in trisomy of 2p21-23 and disruption of 7q35-36 [80]. The trisomy of 2p21-23 is especially notable, since this region overlaps with the 2p23.2 linkage signal in the TSAICG linkage study as discussed above [59]. Another recent study described the translocation t(7;15)(q35;q26.1) in phenotypically normal individuals [81].

Finally, two earlier reports have indicated possible involvement of 9p locus: 9p terminal deletion in a 16 year old male with GTS, developmental delays, and dysmorphic features [82]; and triple X and 9p mosaicism in a woman with mild mental retardation, seizures and aggressive outbursts [83].

Candidate Gene Studies

Candidate gene studies have focused mainly on genes involved in the dopaminergic pathway due to the fact that dopamine antagonists are the most effective medications for tic suppression. As shown in Table 1, similar to findings in other complex disorders [84], candidate gene association studies in GTS have not yielded any clearly replicated results that unequivocally identify a causative GTS susceptibility gene. It is likely that most of these candidate genes tested are not involved in the etiology of GTS, since candidate gene selection is a subjective process, complicated by the lack of clear understanding of the biological pathways involved in GTS. In addition small study sizes are significantly underpowered to detect genes with small-to-medium effect sizes.

Table 1.

Summary of the candidate gene studies, adopted from Scharf and Pauls [60].

Gene Findings References
Dopamine Receptor D1 (DRD1) No linkage/association to GTS/CMT [112115]
Positive association with tic severity [116]
Dopamine Receptor D2 (DRD2) * No linkage/association to GTS/CMT [117121]
Positive association with GTS [122125]
Dopamine Receptor D3 (DRD3) No linkage/association to GTS/CMT [112, 120, 126128]
Positive association with GTS [129]
Dopamine Receptor D4 (DRD4) * No linkage/association to GTS/CMT [112, 116, 130132]
Positive association with GTS/CMT; positive association with comorbid OCD and tics [120, 133, 134]
Dopamine Receptor D5 (DRD5) No linkage/association to GTS/CMT [112, 135]
Tyrosine hydroxylase (TH) No linkage to GTS/CMT [112, 130]
Dopamine β–hydroxylase (DBH) No linkage to GTS/CMT [112, 136]
Positive association with GTS [123, 136]
Dopamine-associated transporter (DAT1, SLC6A3) No linkage/association to GTS/CMT [116, 120, 137, 138]
Positive association with GTS [123, 139, 140]
Monoamine oxidase A (MAOA) * Positive association with GTS [120, 141]
Cathechol-O-methyltransferase (COMT) No linkage/association to GTS/CMT [116, 142, 143]
α1c- and α2c- and α2a- Adrenergic receptors No linkage/association to GTS/CMT [144, 145]
Norepinephrine transporter (NET, SLC6A2) No association with GTS [146, 147]
Serotonin transporter (5HTTLPR) No association with GTS [143]
β2-Adrenergic receptor No linkage to GTS/CMT [148]
Serotonin receptor (5-HT1A) No linkage, association to GTS/CMT [149, 150]
Serotonin receptor (5-HT2A) * Positive association with GTS and OCD [151, 152]
Serotonin receptor 5-HT3A and 5-HT3B No association with GTS [153]
Serotonin receptor 5-HT7 No linkage to GTS/CMT [154]
Tryptophan oxidase No linkage to GTS/CMT [149]
Positive association with GTS [155]
γ-Aminobutyric acid-A receptor α1, α2, α6, β1, γ2 subunits No linkage to GTS/CMT [148]
Androgen Receptor (AR) Positive association with comorbid GTS and ADHD [156]
Huntingtin No association with GTS [157]
HLA-DR No association with GTS [158]
Myelin oligodendrocyte glycoprotein (MOG) Positive weak association with GTS [159]
SLITRK1 pathway genes (ROBO3 and ROBO4) No association with GTS [160]
TPH2 Positive association with GTS [161]
*

Indicates positive associations that have been reproduced in more than one study

Recently, Abelson and colleagues identified a patient affected with GTS and ADHD with a de novo chromosome 13 inversion, inv(13)(q31.1;q33.1) [85, 86]. Out of three genes that mapped within 500 kilobases of the chromosomal breakpoints, the Slit and Trk-like family member 1 (SLITRK1) was selected for further study based on its homology to the axon guidance molecule SLIT. SLITRK1 was screened in 174 GTS affected individuals, and three subjects were found to carry functional variations in this gene: One frameshift mutation and two unrelated occurrences of a sequence variation in the 3′ untranslated region of the gene (var321) that disrupted the binding site for a microRNA miR-189. These variants were absent from 3,600 and 4,296 control chromosomes respectively [85]. While another study [87] detected two independent, novel, non-synonymous sequence changes in SLITRK1 in a set of 44 families with trichotillomania, an OCD-spectrum disorder that has been previously hypothesized to be genetically related to GTS, additional studies in GTS and OCD samples have failed to replicate association with these variants (Table 2) [86, 8893]. The largest study to date screened over 1,000 patients with GTS and found only two var321-positive individuals, one in a GTS patient and the other in a mother of a GTS patient who herself had OCD, but no tics. Both patients failed to transmit this mutation to their GTS affected offspring. While this study does not support a role for var321 in GTS, it should also be noted that due to the low frequency of SLITRK1 var321 in the general population (0.1%), even this sample size of 1000 cases is markedly underpowered to detect an association of such a rare allele [88]. Another large study evaluated 307 Costa Rican and 515 Ashkenazi patients for association between GTS and SLITRK1 [86]. No var321 alleles were identified in the Costa Rican sample, but five instances of var321 were found in Ashkenazi GTS patients, two of whom transmitted var321 to affected children. This high number of SLITRK1 var321 polymorphisms in the Ashkenazi sample prompted analysis of this variant in 256 Ashkenazi control individuals. One unaffected Ashkenazi individual was identified with SLITRK1 var321, suggesting overrepresentation of the var321 polymorphism in Ashkenazi Jews and raising the possibility that population stratification of Ashkenazis in GTS cases compared to controls might account for the association between GTS and SLITRK1 var321 in the original study [86]. Finally, a recent study by Miranda et al. [94] screened 208 GTS affected children from 154 nuclear families for association between common single nucleotide polymorphisms (SNPs) in SLITRK1 and GTS. This study detected a significant association of a common single polymorphism and of a haplotype of three tagging SNPs located in SLITRK1, albeit at an experiment-wide threshold of p<0.05 following permutation. Future studies of larger samples will be needed to attempt to replicate this finding.

Table 2.

Summary of the findings for SLITRK1 gene.

Analysis Results Reference
Sequencing, SNP genotyping Identified one frameshift mutation and two var321 alleles in 174 unrelated probands, these variants were absent from 3600 and 4296 control chromosomes respectively. [85]
SNP genotyping Out of 1048 GTS or CT affected individuals, only one had var321 present, but did not transmit it to the affected offspring. [88]
Sequencing Sequenced 82 Caucasian patients with TS from North America. No var321 alleles found. Novel Ile236Ile variant found in one GTS patient. [89]
Sequencing No var321 alleles found in 160 Taiwanese children with GTS. [90]
SNP Genotyping No var321 alleles found in 307 Costa Rican patients. Five var321 alleles found in 515 Ashkenazi patients, two of whom transmitted var321 to affected children. One in 256 Ashkenazi control individuals also had var321, suggesting overrepresentation of this variant in this population. [86]
SNP Genotyping No var321 or frameshift mutation found in 208 affected children. Haplotype analysis found significant association with GTS, making this the first study to support the original study that found SLITRK1 association with GTS. [94]
Sequencing No var321 alleles or frameshift mutations found in 92 Austrian patients with GTS. One female patient and two affected patients were found to carry a variant within 3′ untranslated region, which was absent from 192 controls. [93]

Analysis of three large multigenerational GTS families also has not detected any SLITRK1 mutations [91, 9597]. However, in large families only founder individuals are informative and thus it would be extremely unlikely to detect this rare mutation in large families.

Future directions

Recent advances in cataloging human genetic polymorphisms, in addition to the decreasing cost of high throughput SNP genotyping and the development of statistical methodology to analyze large sample sets in a rigorous manner have made genome-wide association studies (GWAS) a feasible method for genetic studies of complex disorders [84]. Based on the hypothesis that a proportion of the genetic susceptibility for common diseases may be caused by common genetic variants that arose early in human history and thus are shared across members of a population derived from a common set of ancestors (the common-disease, common variant or CDCV hypothesis), this approach has been remarkably successful, with over 150 common variants identified within the past 2 years [84, 98].

Genome-Wide Association Studies provide several advantages over linkage and candidate gene studies. With a sufficiently large sample, it is possible to overcome the lack of power in linkage analyses to detect common alleles with low penetrance. GWAS can also detect much smaller associated DNA regions compared to linkage studies, since linkage analyses are based on rare recombination events in only a few generations, thus resulting in large linked regions. Unlike linkage analyses, GWAS rely on historical recombination events in populations over the course of human history, thus resulting in much smaller regions of association. Furthermore, a great advantage of GWAS over candidate genes studies is the fact that GWAS assume no prior biological knowledge of the disease process, but instead test for association across the whole genome in an agnostic approach. Newer GWAS genotyping platforms now also have the additional benefit of containing copy-number probes to allow examination of both SNPs and CNVs in a single experiment [99, 100].

For conducting GWAS it is critical that the sample is sufficiently large to provide enough statistical power to reduce Type I error and detect an association. The power of GWAS to detect an association is dependent both on the allele frequency of the disease variant as well as its effect size, expressed as a genotype relative risk (GRR). GRR is defined as the probability of a person with a specific disease variant to have a disease compared to the probability of a person without that gene variant. Larger sample sizes are required to detect alleles of low frequency or of small effect.

As can be seen in Figure 2, for GRR=1.2 the number of cases and controls needed to detect a causative allele reduces from about 23,000 at minor allele frequency (MAF) =5% to ~ 8,000 at MAF=20% and to ~6,000 at MAF=40%. The effect of GRR on the study size is even greater: at MAF =10% and GRR 1.1, about 47,000 cases and controls are needed, but as GRR increases to 1.2, 1.3, 1.5, 1.7, and 2.0 the required number of individuals decreases to ~12,500, ~6,000, 2,300, 1,300, and 700 respectively. For instance, a GWAS for age-related macular degeneration (AMD) of only 96 cases and 50 controls was able to detect an association with a common variant in the complement factor H gene (CFH) at a nominal P value <10−7 [101]. The power to detect this association was due to the high GRR caused by CFH variant: the presence of two risk alleles in an individual increased risk of developing AMD by a factor of 7.4 [101]. On the other hand, in order to detect alleles with modest effect sizes, tens of thousands of cases and controls could be needed, as demonstrated by studies on human height variation with a combined sample size of ~63,000 individuals [102104]. These studies found 54 variants, each with an average size effect of 0.4 cm per allele, indicating that an even larger sample size maybe needed to detect common, small-effect alleles responsible for the residual population variance [105].

Figure 2.

Figure 2

The number of cases and controls required in an association study to achieve 80% power across a range of minor allele frequencies (MAF) and genotype relative risks (GRR) for a multiplicative model. GTS prevalence was set at 2% (1% GTS genes + 1% OCD genes) and linkage disequilibrium between marker and causative allele was set to 1. Power calculations were performed for Type I error rate (alpha) of p< 5 × 10−8 as required to reach a Bonferroni corrected p<0.05 significance level for ~ 1million independent tests performed in a GWAS [162]. Calculations were performed with the Genetic Power Calculator [163].

As described above, GWAS are designed specifically to detect association of common disease variants (typically with allele frequencies ≥5% in a population) [106]; GWAS have essentially no power to detect multiple rare variants in either the same gene or in different genes (the multiple rare variants hypothesis) [107]. However the ability to detect CNVs is a promising parallel approach, since these polymorphisms appear to exist both in common and rare forms and have been demonstrated to contribute to the manifestation of other psychiatric disorders such as autism [62] and schizophrenia [65]. In addition, rare, highly penetrant CNVs can be detected in a relatively small sample size, as demonstrated by studies of schizophrenia which consisted of about 150 individuals [63, 108].

What has become evident in GWAS of other diseases is that large-scale, multi-center collaborations are needed to obtain a sample sufficiently large enough to provide adequate power for a GWAS. A recent genome-wide association analysis of bipolar disorder, which combined three datasets [109111] totaling over 4300 cases and 6200 controls [109], detected two markers with genome wide significant association. Combining the datasets was instrumental in providing enough power to reach a genome-wide level of statistical significance, since none of these associations were detected in the individual samples.

In collaboration with other investigators, the TSAICG has undertaken a GWAS for GTS. The GTS GWAS should help identify short genomic segments (either SNPs or CNVs) harboring susceptibility genes for GTS. Once these genes are identified, research can be initiated to elucidate the biological pathways and processes influencing the development of the GTS phenotype. These pathways will hopefully reveal cellular and molecular mechanisms previously unsuspected in GTS pathogenesis, and thus could help to develop new treatment paradigms to significantly reduce the suffering experienced by individuals with GTS and related disorders.

Acknowledgments

Acknowledgments for funding: Supported in part by grants from the National Institute of Neurological Disease and Stroke, NS16648 and NS40024.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Tourette Gdl. Etude sur une affection nerveuse caracterisee par de l’indoordination motrice accompagnee d’echolalie et al copralalie. Archives or Neurology. 1885;9:19–42. [Google Scholar]
  • 2.Pauls DL, Hurst CR, Kruger SD, Leckman JF, Kidd KK, Cohen DJ. Gilles de la Tourette’s syndrome and attention deficit disorder with hyperactivity. Evidence against a genetic relationship. Arch Gen Psychiatry. 1986;43(12):1177–9. doi: 10.1001/archpsyc.1986.01800120063012. [DOI] [PubMed] [Google Scholar]
  • 3.Kidd KK, Prusoff BA, Cohen DJ. Familial pattern of Gilles de la Tourette syndrome. Arch Gen Psychiatry. 1980;37(12):1336–9. doi: 10.1001/archpsyc.1980.01780250022001. [DOI] [PubMed] [Google Scholar]
  • 4.Pauls DL, Cohen DJ, Heimbuch R, Detlor J, Kidd KK. Familial pattern and transmission of Gilles de la Tourette syndrome and multiple tics. Arch Gen Psychiatry. 1981;38(10):1091–3. doi: 10.1001/archpsyc.1981.01780350025002. [DOI] [PubMed] [Google Scholar]
  • 5.Pauls DL, Kruger SD, Leckman JF, Cohen DJ, Kidd KK. The risk of Tourette’s syndrome and chronic multiple tics among relatives of Tourette’s syndrome patients obtained by direct interview. J Am Acad Child Psychiatry. 1984;23(2):134–7. doi: 10.1097/00004583-198403000-00003. [DOI] [PubMed] [Google Scholar]
  • 6.Pauls DL, Raymond CL, Stevenson JM, Leckman JF. A family study of Gilles de la Tourette syndrome. Am J Hum Genet. 1991;48(1):154–63. [PMC free article] [PubMed] [Google Scholar]
  • 7.Eapen V, Pauls DL, Robertson MM. Evidence for autosomal dominant transmission in Tourette’s syndrome. United Kingdom cohort study. Br J Psychiatry. 1993;162:593–6. doi: 10.1192/bjp.162.5.593. [DOI] [PubMed] [Google Scholar]
  • 8.Walkup JT, LaBuda MC, Singer HS, Brown J, Riddle MA, Hurko O. Family study and segregation analysis of Tourette syndrome: evidence for a mixed model of inheritance. Am J Hum Genet. 1996;59(3):684–93. [PMC free article] [PubMed] [Google Scholar]
  • 9.Hebebrand J, Klug B, Fimmers R, Seuchter SA, Wettke-Schafer R, Deget F, Camps A, Lisch S, Hebebrand K, von Gontard A, Lehmkuhl G, Poustka F, Schmidt M, Baur MP, Remschmidt H. Rates for tic disorders and obsessive compulsive symptomatology in families of children and adolescents with Gilles de la Tourette syndrome. J Psychiatr Res. 1997;31(5):519–30. doi: 10.1016/s0022-3956(97)00028-9. [DOI] [PubMed] [Google Scholar]
  • 10.Kano Y, Ohta M, Nagai Y, Pauls DL, Leckman JF. A family study of Tourette syndrome in Japan. Am J Med Genet. 2001;105(5):414–21. doi: 10.1002/ajmg.1436. [DOI] [PubMed] [Google Scholar]
  • 11.Saccomani L, Fabiana V, Manuela B, Giambattista R. Tourette syndrome and chronic tics in a sample of children and adolescents. Brain Dev. 2005;27(5):349–52. doi: 10.1016/j.braindev.2004.09.007. [DOI] [PubMed] [Google Scholar]
  • 12.Stewart SE, Illmann C, Geller DA, Leckman JF, King R, Pauls DL. A controlled family study of attention-deficit/hyperactivity disorder and Tourette’s disorder. J Am Acad Child Adolesc Psychiatry. 2006;45(11):1354–62. doi: 10.1097/01.chi.0000251211.36868.fe. [DOI] [PubMed] [Google Scholar]
  • 13.Smaller JW, Sheidley BR, Tsuang MT. Psychiatric Genetics Applications in Clinical Practice. Washington, DC: American Psychiatric Publishing, Inc; 2008. [Google Scholar]
  • 14.Freeman RD, Fast DK, Burd L, Kerbeshian J, Robertson MM, Sandor P. An international perspective on Tourette syndrome: selected findings from 3,500 individuals in 22 countries. Dev Med Child Neurol. 2000;42(7):436–47. doi: 10.1017/s0012162200000839. [DOI] [PubMed] [Google Scholar]
  • 15.Mol Debes NM, Hjalgrim H, Skov L. Validation of the presence of comorbidities in a Danish clinical cohort of children with Tourette syndrome. J Child Neurol. 2008;23(9):1017–27. doi: 10.1177/0883073808316370. [DOI] [PubMed] [Google Scholar]
  • 16.Freeman RD. Tic disorders and ADHD: answers from a world-wide clinical dataset on Tourette syndrome. Eur Child Adolesc Psychiatry. 2007;16 (Suppl 1):15–23. doi: 10.1007/s00787-007-1003-7. [DOI] [PubMed] [Google Scholar]
  • 17.Khalifa N, von Knorring AL. Psychopathology in a Swedish population of school children with tic disorders. J Am Acad Child Adolesc Psychiatry. 2006;45(11):1346–53. doi: 10.1097/01.chi.0000251210.98749.83. [DOI] [PubMed] [Google Scholar]
  • 18.Karno M, Golding JM, Sorenson SB, Burnam MA. The epidemiology of obsessive-compulsive disorder in five US communities. Arch Gen Psychiatry. 1988;45(12):1094–9. doi: 10.1001/archpsyc.1988.01800360042006. [DOI] [PubMed] [Google Scholar]
  • 19.Weissman MM, Bland RC, Canino GJ, Greenwald S, Hwu HG, Lee CK, Newman SC, Oakley-Browne MA, Rubio-Stipec M, Wickramaratne PJ, et al. The cross national epidemiology of obsessive compulsive disorder. The Cross National Collaborative Group. J Clin Psychiatry. 1994;55 (Suppl):5–10. [PubMed] [Google Scholar]
  • 20.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(6):942–8. doi: 10.1176/ajp.2007.164.6.942. [DOI] [PubMed] [Google Scholar]
  • 21.Pringsheim T, Lang A, Kurlan R, Pearce M, Sandor P. Understanding disability in Tourette syndrome. Dev Med Child Neurol. 2008:468–472. doi: 10.1111/j.1469-8749.2008.03168.x. [DOI] [PubMed] [Google Scholar]
  • 22.Budman CL, Bruun RD, Park KS, Lesser M, Olson M. Explosive outbursts in children with Tourette’s disorder. J Am Acad Child Adolesc Psychiatry. 2000;39(10):1270–6. doi: 10.1097/00004583-200010000-00014. [DOI] [PubMed] [Google Scholar]
  • 23.Budman CL, Bruun RD, Park KS, Olson ME. Rage attacks in children and adolescents with Tourette’s disorder: a pilot study. J Clin Psychiatry. 1998;59(11):576–80. doi: 10.4088/jcp.v59n1103. [DOI] [PubMed] [Google Scholar]
  • 24.Pauls DL. The genetics of obsessive compulsive disorder: a review of the evidence. Am J Med Genet C Semin Med Genet. 2008;148(2):133–9. doi: 10.1002/ajmg.c.30168. [DOI] [PubMed] [Google Scholar]
  • 25.Pauls DL, Towbin KE, Leckman JF, Zahner GE, Cohen DJ. Gilles de la Tourette’s syndrome and obsessive-compulsive disorder. Evidence supporting a genetic relationship. Arch Gen Psychiatry. 1986;43(12):1180–2. doi: 10.1001/archpsyc.1986.01800120066013. [DOI] [PubMed] [Google Scholar]
  • 26.Curtis D, Robertson MM, Gurling HM. Autosomal dominant gene transmission in a large kindred with Gilles de la Tourette syndrome. Br J Psychiatry. 1992;160:845–9. doi: 10.1192/bjp.160.6.845. [DOI] [PubMed] [Google Scholar]
  • 27.McMahon WM, van de Wetering BJ, Filloux F, Betit K, Coon H, Leppert M. Bilineal transmission and phenotypic variation of Tourette’s disorder in a large pedigree. J Am Acad Child Adolesc Psychiatry. 1996;35(5):672–80. [PubMed] [Google Scholar]
  • 28.Pauls DL, Alsobrook JP, 2nd, Goodman W, Rasmussen S, Leckman JF. A family study of obsessive-compulsive disorder. Am J Psychiatry. 1995;152(1):76–84. doi: 10.1176/ajp.152.1.76. [DOI] [PubMed] [Google Scholar]
  • 29.Grados MA, Riddle MA, Samuels JF, Liang KY, Hoehn-Saric R, Bienvenu OJ, Walkup JT, Song D, Nestadt G. The familial phenotype of obsessive-compulsive disorder in relation to tic disorders: the Hopkins OCD family study. Biol Psychiatry. 2001;50(8):559–65. doi: 10.1016/s0006-3223(01)01074-5. [DOI] [PubMed] [Google Scholar]
  • 30.Leonard HL, Lenane MC, Swedo SE, Rettew DC, Gershon ES, Rapoport JL. Tics and Tourette’s disorder: a 2- to 7-year follow-up of 54 obsessive-compulsive children. Am J Psychiatry. 1992;149(9):1244–51. doi: 10.1176/ajp.149.9.1244. [DOI] [PubMed] [Google Scholar]
  • 31.Nestadt G, Samuels J, Riddle M, Bienvenu OJ, 3rd, Liang KY, LaBuda M, Walkup J, Grados M, Hoehn-Saric R. A family study of obsessive-compulsive disorder. Arch Gen Psychiatry. 2000;57(4):358–63. doi: 10.1001/archpsyc.57.4.358. [DOI] [PubMed] [Google Scholar]
  • 32.Biederman J, Faraone SV. Attention-deficit hyperactivity disorder. Lancet. 2005;366(9481):237–48. doi: 10.1016/S0140-6736(05)66915-2. [DOI] [PubMed] [Google Scholar]
  • 33.Comings DE, Comings BG. A controlled study of Tourette syndrome. I. Attention-deficit disorder, learning disorders, and school problems. Am J Hum Genet. 1987;41(5):701–41. [PMC free article] [PubMed] [Google Scholar]
  • 34.Knell ER, Comings DE. Tourette’s syndrome and attention-deficit hyperactivity disorder: evidence for a genetic relationship. J Clin Psychiatry. 1993;54(9):331–7. [PubMed] [Google Scholar]
  • 35.Pauls DL, Leckman JF, Cohen DJ. Familial relationship between Gilles de la Tourette’s syndrome, attention deficit disorder, learning disabilities, speech disorders, and stuttering. J Am Acad Child Adolesc Psychiatry. 1993;32(5):1044–50. doi: 10.1097/00004583-199309000-00025. [DOI] [PubMed] [Google Scholar]
  • 36.ORourke JA, Scharf JM, Stewart E, Platko J, Illmann C, Geller D, King RA, Leckman JF, Pauls D. The Familial Association of Tourette’s Disorder and ADHD: the Impact of OCD Symptoms. doi: 10.1002/ajmg.b.31195. Submitted. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Spencer T, Biederman J, Harding M, O’Donnell D, Wilens T, Faraone S, Coffey B, Geller D. Disentangling the overlap between Tourette’s disorder and ADHD. J Child Psychol Psychiatry. 1998;39(7):1037–44. [PubMed] [Google Scholar]
  • 38.McMahon WM, Carter AS, Fredine N, Pauls DL. Children at familial risk for Tourette’s disorder: Child and parent diagnoses. Am J Med Genet B Neuropsychiatr Genet. 2003;121B(1):105–11. doi: 10.1002/ajmg.b.20065. [DOI] [PubMed] [Google Scholar]
  • 39.Grados MA, Mathews CA. Latent Class Analysis of Gilles de la Tourette Syndrome Using Comorbidities: Clinical and Genetic Implications. Biol Psychiatry. 2008 doi: 10.1016/j.biopsych.2008.01.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Robertson MM, Althoff RR, Hafez A, Pauls DL. Principal components analysis of a large cohort with Tourette syndrome. Br J Psychiatry. 2008;193(1):31–6. doi: 10.1192/bjp.bp.107.039909. [DOI] [PubMed] [Google Scholar]
  • 41.Comings DE, Comings BG, Devor EJ, Cloninger CR. Detection of major gene for Gilles de la Tourette syndrome. Am J Hum Genet. 1984;36(3):586–600. [PMC free article] [PubMed] [Google Scholar]
  • 42.Baron M, Shapiro E, Shapiro A, Rainer JD. Genetic analysis of Tourette syndrome suggesting major gene effect. Am J Hum Genet. 1981;33(5):767–75. [PMC free article] [PubMed] [Google Scholar]
  • 43.Devor EJ. Complex segregation analysis of Gilles de la Tourette syndrome: further evidence for a major locus mode of transmission. Am J Hum Genet. 1984;36(3):704–9. [PMC free article] [PubMed] [Google Scholar]
  • 44.Hasstedt SJ, Leppert M, Filloux F, van de Wetering BJ, McMahon WM. Intermediate inheritance of Tourette syndrome, assuming assortative mating. Am J Hum Genet. 1995;57(3):682–9. [PMC free article] [PubMed] [Google Scholar]
  • 45.Kidd KK, Pauls DL. Genetic hypotheses for Tourette syndrome. Adv Neurol. 1982;35:243–9. [PubMed] [Google Scholar]
  • 46.Pauls DL, Leckman JF. The inheritance of Gilles de la Tourette’s syndrome and associated behaviors. Evidence for autosomal dominant transmission. N Engl J Med. 1986;315(16):993–7. doi: 10.1056/NEJM198610163151604. [DOI] [PubMed] [Google Scholar]
  • 47.Pauls DL, Pakstis AJ, Kurlan R, Kidd KK, Leckman JF, Cohen DJ, Kidd JR, Como P, Sparkes R. Segregation and linkage analyses of Tourette’s syndrome and related disorders. J Am Acad Child Adolesc Psychiatry. 1990;29(2):195–203. doi: 10.1097/00004583-199003000-00007. [DOI] [PubMed] [Google Scholar]
  • 48.Price RA, Pauls DL, Kruger SD, Caine ED. Family data support a dominant major gene for Tourette syndrome. Psychiatry Res. 1988;24(3):251–61. doi: 10.1016/0165-1781(88)90107-2. [DOI] [PubMed] [Google Scholar]
  • 49.Seuchter SA, Hebebrand J, Klug B, Knapp M, Lehmkuhl G, Poustka F, Schmidt M, Remschmidt H, Baur MP. Complex segregation analysis of families ascertained through Gilles de la Tourette syndrome. Genet Epidemiol. 2000;18(1):33–47. doi: 10.1002/(SICI)1098-2272(200001)18:1<33::AID-GEPI3>3.0.CO;2-2. [DOI] [PubMed] [Google Scholar]
  • 50.Comings DE, Comings BG, Knell E. Hypothesis: homozygosity in Tourette syndrome. Am J Med Genet. 1989;34(3):413–21. doi: 10.1002/ajmg.1320340318. [DOI] [PubMed] [Google Scholar]
  • 51.Hanna PA, Janjua FN, Contant CF, Jankovic J. Bilineal transmission in Tourette syndrome. Neurology. 1999;53(4):813–8. doi: 10.1212/wnl.53.4.813. [DOI] [PubMed] [Google Scholar]
  • 52.Kurlan R, Eapen V, Stern J, McDermott MP, Robertson MM. Bilineal transmission in Tourette’s syndrome families. Neurology. 1994;44(12):2336–42. doi: 10.1212/wnl.44.12.2336. [DOI] [PubMed] [Google Scholar]
  • 53.Price RA, Kidd KK, Cohen DJ, Pauls DL, Leckman JF. A twin study of Tourette syndrome. Arch Gen Psychiatry. 1985;42(8):815–20. doi: 10.1001/archpsyc.1985.01790310077011. [DOI] [PubMed] [Google Scholar]
  • 54.Hyde TM, Aaronson BA, Randolph C, Rickler KC, Weinberger DR. Relationship of birth weight to the phenotypic expression of Gilles de la Tourette’s syndrome in monozygotic twins. Neurology. 1992;42(3 Pt 1):652–8. doi: 10.1212/wnl.42.3.652. [DOI] [PubMed] [Google Scholar]
  • 55.Barr CL, Wigg KG, Pakstis AJ, Kurlan R, Pauls D, Kidd KK, Tsui LC, Sandor P. Genome scan for linkage to Gilles de la Tourette syndrome. Am J Med Genet. 1999;88(4):437–45. doi: 10.1002/(sici)1096-8628(19990820)88:4<437::aid-ajmg24>3.0.co;2-e. [DOI] [PubMed] [Google Scholar]
  • 56.Curtis D, Brett P, Dearlove AM, McQuillin A, Kalsi G, Robertson MM, Gurling HM. Genome scan of Tourette syndrome in a single large pedigree shows some support for linkage to regions of chromosomes 5, 10 and 13. Psychiatr Genet. 2004;14(2):83–7. doi: 10.1097/01.ypg.0000107927.32051.f5. [DOI] [PubMed] [Google Scholar]
  • 57.Leppert M, Peiffer A, Snyder B. Two loci of interest in a family with Tourette syndrome. Am J Hum Genet Suppl. 1996;59:A225. [Google Scholar]
  • 58.(TSAICG) TSAICfG. A complete genome screen in sib pairs affected by Gilles de la Tourette syndrome. The Tourette Syndrome Association International Consortium for Genetics. Am J Hum Genet. 1999;65(5):1428–36. doi: 10.1086/302613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.(TSAICG) TSAICfG. Genome scan for Tourette disorder in affected-sibling-pair and multigenerational families. Am J Hum Genet. 2007;80(2):265–72. doi: 10.1086/511052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Scharf JM, Pauls DL. Genetics of Tic Disorders. In: Rimoin D, Connor JM, Pyeritz RE, Korf BR, editors. Emery and Rimoin’s Principles and Practices of Medical Genetics. Philadelphia: Churchill Livingstone/Elsevier; 2007. pp. 2737–2754. [Google Scholar]
  • 61.Cook EH, Jr, Scherer SW. Copy-number variations associated with neuropsychiatric conditions. Nature. 2008;455(7215):919–23. doi: 10.1038/nature07458. [DOI] [PubMed] [Google Scholar]
  • 62.Weiss LA, Shen Y, Korn JM, Arking DE, Miller DT, Fossdal R, Saemundsen E, Stefansson H, Ferreira MA, Green T, Platt OS, Ruderfer DM, Walsh CA, Altshuler D, Chakravarti A, Tanzi RE, Stefansson K, Santangelo SL, Gusella JF, Sklar P, Wu BL, Daly MJ. Association between microdeletion and microduplication at 16p11.2 and autism. N Engl J Med. 2008;358(7):667–75. doi: 10.1056/NEJMoa075974. [DOI] [PubMed] [Google Scholar]
  • 63.Walsh T, McClellan JM, McCarthy SE, Addington AM, Pierce SB, Cooper GM, Nord AS, Kusenda M, Malhotra D, Bhandari A, Stray SM, Rippey CF, Roccanova P, Makarov V, Lakshmi B, Findling RL, Sikich L, Stromberg T, Merriman B, Gogtay N, Butler P, Eckstrand K, Noory L, Gochman P, Long R, Chen Z, Davis S, Baker C, Eichler EE, Meltzer PS, Nelson SF, Singleton AB, Lee MK, Rapoport JL, King MC, Sebat J. Rare structural variants disrupt multiple genes in neurodevelopmental pathways in schizophrenia. Science. 2008;320(5875):539–43. doi: 10.1126/science.1155174. [DOI] [PubMed] [Google Scholar]
  • 64.Stefansson H, Rujescu D, Cichon S, Pietilainen OP, Ingason A, Steinberg S, Fossdal R, Sigurdsson E, Sigmundsson T, Buizer-Voskamp JE, Hansen T, Jakobsen KD, Muglia P, Francks C, Matthews PM, Gylfason A, Halldorsson BV, Gudbjartsson D, Thorgeirsson TE, Sigurdsson A, Jonasdottir A, Jonasdottir A, Bjornsson A, Mattiasdottir S, Blondal T, Haraldsson M, Magnusdottir BB, Giegling I, Moller HJ, Hartmann A, Shianna KV, Ge D, Need AC, Crombie C, Fraser G, Walker N, Lonnqvist J, Suvisaari J, Tuulio-Henriksson A, Paunio T, Toulopoulou T, Bramon E, Di Forti M, Murray R, Ruggeri M, Vassos E, Tosato S, Walshe M, Li T, Vasilescu C, Muhleisen TW, Wang AG, Ullum H, Djurovic S, Melle I, Olesen J, Kiemeney LA, Franke B, Sabatti C, Freimer NB, Gulcher JR, Thorsteinsdottir U, Kong A, Andreassen OA, Ophoff RA, Georgi A, Rietschel M, Werge T, Petursson H, Goldstein DB, Nothen MM, Peltonen L, Collier DA, St Clair D, Stefansson K. Large recurrent microdeletions associated with schizophrenia. Nature. 2008;455(7210):232–6. doi: 10.1038/nature07229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.International_Schizophrenia_Consortium. Rare chromosomal deletions and duplications increase risk of schizophrenia. Nature. 2008;455(7210):237–41. doi: 10.1038/nature07239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Robertson MM, Trimble MR. Normal chromosomal findings in Gilles de la Tourette syndrome. Psychiatric Genetics. 1993;3:95–99. [Google Scholar]
  • 67.Shen Y, Dies KA, Holm IA, Bridgemohan C, Sobeih MM, Caronna EB, Miller KJ, Frazier JA, Silverstein I, Picker J, Weissman L, Raffalli P, Jeste S, Demmer LA, Peters HK, Brewster SJ, Kowalczyk SJ, Rosen-Sheidley B, McGowan C, DAW, Lincoln SA, Lowe KR, Schonwald A, Robbins M, Wolff R, Becker R, Nasir R, Milunsky JM, Rappaport L, Gusella JF, Walsh CA, Wu BL, Miller DT. Clinical Genetic Testing for Patients with Autism Spectrum Disorders. Pediatrics. doi: 10.1542/peds.2009-1684. Submitted. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Feuk L, Carson AR, Scherer SW. Structural variation in the human genome. Nat Rev Genet. 2006;7(2):85–97. doi: 10.1038/nrg1767. [DOI] [PubMed] [Google Scholar]
  • 69.Shelley BP, Robertson MM, Turk J. An individual with Gilles de la Tourette syndrome and Smith-Magenis microdeletion syndrome: is chromosome 17p11.2 a candidate region for Tourette syndrome putative susceptibility genes? J Intellect Disabil Res. 2007;51(Pt 8):620–4. doi: 10.1111/j.1365-2788.2006.00943.x. [DOI] [PubMed] [Google Scholar]
  • 70.Dehning S, Riedel M, Muller N. Father-to-son transmission of 6;17 translocation in Tourette’s syndrome. Am J Psychiatry. 2008;165(8):1051–2. doi: 10.1176/appi.ajp.2008.07111828. [DOI] [PubMed] [Google Scholar]
  • 71.Boghosian-Sell L, Comings DE, Overhauser J. Tourette syndrome in a pedigree with a 7;18 translocation: identification of a YAC spanning the translocation breakpoint at 18q22.3. Am J Hum Genet. 1996;59(5):999–1005. [PMC free article] [PubMed] [Google Scholar]
  • 72.Kroisel PM, Petek E, Emberger W, Windpassinger C, Wladika W, Wagner K. Candidate region for Gilles de la Tourette syndrome at 7q31. Am J Med Genet. 2001;101(3):259–61. doi: 10.1002/1096-8628(20010701)101:3<259::aid-ajmg1374>3.0.co;2-#. [DOI] [PubMed] [Google Scholar]
  • 73.Petek E, Windpassinger C, Vincent JB, Cheung J, Boright AP, Scherer SW, Kroisel PM, Wagner K. Disruption of a novel gene (IMMP2L) by a breakpoint in 7q31 associated with Tourette syndrome. Am J Hum Genet. 2001;68(4):848–58. doi: 10.1086/319523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Petek E, Schwarzbraun T, Noor A, Patel M, Nakabayashi K, Choufani S, Windpassinger C, Stamenkovic M, Robertson MM, Aschauer HN, Gurling HM, Kroisel PM, Wagner K, Scherer SW, Vincent JB. Molecular and genomic studies of IMMP2L and mutation screening in autism and Tourette syndrome. Mol Genet Genomics. 2007;277(1):71–81. doi: 10.1007/s00438-006-0173-1. [DOI] [PubMed] [Google Scholar]
  • 75.Crawford FC, Ait-Ghezala G, Morris M, Sutcliffe MJ, Hauser RA, Silver AA, Mullan MJ. Translocation breakpoint in two unrelated Tourette syndrome cases, within a region previously linked to the disorder. Hum Genet. 2003;113(2):154–61. doi: 10.1007/s00439-003-0942-4. [DOI] [PubMed] [Google Scholar]
  • 76.Matsumoto N, David DE, Johnson EW, Konecki D, Burmester JK, Ledbetter DH, Weber JL. Breakpoint sequences of an 1;8 translocation in a family with Gilles de la Tourette syndrome. Eur J Hum Genet. 2000;8(11):875–83. doi: 10.1038/sj.ejhg.5200549. [DOI] [PubMed] [Google Scholar]
  • 77.Donnai D. Gene location in Tourette syndrome. Lancet. 1987;1(8533):627. doi: 10.1016/s0140-6736(87)90263-7. [DOI] [PubMed] [Google Scholar]
  • 78.State MW, Greally JM, Cuker A, Bowers PN, Henegariu O, Morgan TM, Gunel M, DiLuna M, King RA, Nelson C, Donovan A, Anderson GM, Leckman JF, Hawkins T, Pauls DL, Lifton RP, Ward DC. Epigenetic abnormalities associated with a chromosome 18(q21-q22) inversion and a Gilles de la Tourette syndrome phenotype. Proc Natl Acad Sci U S A. 2003;100(8):4684–9. doi: 10.1073/pnas.0730775100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Cuker A, State MW, King RA, Davis N, Ward DC. Candidate locus for Gilles de la Tourette syndrome/obsessive compulsive disorder/chronic tic disorder at 18q22. Am J Med Genet A. 2004;130A(1):37–9. doi: 10.1002/ajmg.a.30066. [DOI] [PubMed] [Google Scholar]
  • 80.Verkerk AJ, Mathews CA, Joosse M, Eussen BH, Heutink P, Oostra BA. CNTNAP2 is disrupted in a family with Gilles de la Tourette syndrome and obsessive compulsive disorder. Genomics. 2003;82(1):1–9. doi: 10.1016/s0888-7543(03)00097-1. [DOI] [PubMed] [Google Scholar]
  • 81.Belloso JM, Bache I, Guitart M, Caballin MR, Halgren C, Kirchhoff M, Ropers HH, Tommerup N, Tumer Z. Disruption of the CNTNAP2 gene in a t(7;15) translocation family without symptoms of Gilles de la Tourette syndrome. Eur J Hum Genet. 2007;15(6):711–3. doi: 10.1038/sj.ejhg.5201824. [DOI] [PubMed] [Google Scholar]
  • 82.Taylor LD, Krizman DB, Jankovic J, Hayani A, Steuber PC, Greenberg F, Fenwick RG, Caskey CT. 9p monosomy in a patient with Gilles de la Tourette’s syndrome. Neurology. 1991;41(9):1513–5. doi: 10.1212/wnl.41.9.1513. [DOI] [PubMed] [Google Scholar]
  • 83.Singh DN, Howe GL, Jordan HW, Hara S. Tourette’s syndrome in a black woman with associated triple X and 9p mosaicism. J Natl Med Assoc. 1982;74(7):675–82. [PMC free article] [PubMed] [Google Scholar]
  • 84.Altshuler D, Daly MJ, Lander ES. Genetic mapping in human disease. Science. 2008;322(5903):881–8. doi: 10.1126/science.1156409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Abelson JF, Kwan KY, O’Roak BJ, Baek DY, Stillman AA, Morgan TM, Mathews CA, Pauls DL, Rasin MR, Gunel M, Davis NR, Ercan-Sencicek AG, Guez DH, Spertus JA, Leckman JF, Dure LSt, Kurlan R, Singer HS, Gilbert DL, Farhi A, Louvi A, Lifton RP, Sestan N, State MW. Sequence variants in SLITRK1 are associated with Tourette’s syndrome. Science. 2005;310(5746):317–20. doi: 10.1126/science.1116502. [DOI] [PubMed] [Google Scholar]
  • 86.Keen-Kim D, Mathews CA, Reus VI, Lowe TL, Herrera LD, Budman CL, Gross-Tsur V, Pulver AE, Bruun RD, Erenberg G, Naarden A, Sabatti C, Freimer NB. Overrepresentation of rare variants in a specific ethnic group may confuse interpretation of association analyses. Hum Mol Genet. 2006;15(22):3324–8. doi: 10.1093/hmg/ddl408. [DOI] [PubMed] [Google Scholar]
  • 87.Zuchner S, Cuccaro ML, Tran-Viet KN, Cope H, Krishnan RR, Pericak-Vance MA, Wright HH, Ashley-Koch A. SLITRK1 mutations in trichotillomania. Mol Psychiatry. 2006;11(10):887–9. doi: 10.1038/sj.mp.4001898. [DOI] [PubMed] [Google Scholar]
  • 88.Scharf JM, Moorjani P, Fagerness J, Platko JV, Illmann C, Galloway B, Jenike E, Stewart SE, Pauls DL. Lack of association between SLITRK1var321 and Tourette syndrome in a large family-based sample. Neurology. 2008;70(16 Pt 2):1495–6. doi: 10.1212/01.wnl.0000296833.25484.bb. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Deng H, Le WD, Xie WJ, Jankovic J. Examination of the SLITRK1 gene in Caucasian patients with Tourette syndrome. Acta Neurol Scand. 2006;114(6):400–2. doi: 10.1111/j.1600-0404.2006.00706.x. [DOI] [PubMed] [Google Scholar]
  • 90.Chou IC, Wan L, Liu SC, Tsai CH, Tsai FJ. Association of the Slit and Trk-like 1 gene in Taiwanese patients with Tourette syndrome. Pediatr Neurol. 2007;37(6):404–6. doi: 10.1016/j.pediatrneurol.2007.06.017. [DOI] [PubMed] [Google Scholar]
  • 91.Verkerk AJ, Cath DC, van der Linde HC, Both J, Heutink P, Breedveld G, Aulchenko YS, Oostra BA. Genetic and clinical analysis of a large Dutch Gilles de la Tourette family. Mol Psychiatry. 2006;11(10):954–64. doi: 10.1038/sj.mp.4001877. [DOI] [PubMed] [Google Scholar]
  • 92.Wendland JR, Kruse MR, Murphy DL. Functional SLITRK1 var321, varCDfs and SLC6A4 G56A variants and susceptibility to obsessive-compulsive disorder. Mol Psychiatry. 2006;11(9):802–4. doi: 10.1038/sj.mp.4001848. [DOI] [PubMed] [Google Scholar]
  • 93.Zimprich A, Hatala K, Riederer F, Stogmann E, Aschauer HN, Stamenkovic M. Sequence analysis of the complete SLITRK1 gene in Austrian patients with Tourette’s disorder. Psychiatr Genet. 2008;18(6):308–9. doi: 10.1097/YPG.0b013e3283060f6f. [DOI] [PubMed] [Google Scholar]
  • 94.Miranda DM, Wigg K, Kabia EM, Feng Y, Sandor P, Barr CL. Association of SLITRK1 to Gilles de la Tourette Syndrome. Am J Med Genet B Neuropsychiatr Genet. 2008 doi: 10.1002/ajmg.b.30840. [DOI] [PubMed] [Google Scholar]
  • 95.Fabbrini G, Pasquini M, Aurilia C, Berardelli I, Breedveld G, Oostra BA, Bonifati V, Berardelli A. A large Italian family with Gilles de la Tourette syndrome: clinical study and analysis of the SLITRK1 gene. Mov Disord. 2007;22(15):2229–34. doi: 10.1002/mds.21697. [DOI] [PubMed] [Google Scholar]
  • 96.Orth M, Djarmati A, Baumer T, Winkler S, Grunewald A, Lohmann-Hedrich K, Kabakci K, Hagenah J, Klein C, Munchau A. Autosomal dominant myoclonus-dystonia and Tourette syndrome in a family without linkage to the SGCE gene. Mov Disord. 2007;22(14):2090–6. doi: 10.1002/mds.21674. [DOI] [PubMed] [Google Scholar]
  • 97.Pasquini M, Fabbrini G, Berardelli I, Bonifati V, Biondi M, Berardelli A. Psychopathological features of obsessive-compulsive disorder in an Italian family with Gilles de la Tourette syndrome not linked to the SLITRK1 gene. Psychiatry Res. 2008;161(1):109–11. doi: 10.1016/j.psychres.2008.02.012. [DOI] [PubMed] [Google Scholar]
  • 98.Manolio TA, Brooks LD, Collins FS. A HapMap harvest of insights into the genetics of common disease. J Clin Invest. 2008;118(5):1590–605. doi: 10.1172/JCI34772. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.McCarroll SA, Kuruvilla FG, Korn JM, Cawley S, Nemesh J, Wysoker A, Shapero MH, de Bakker PI, Maller JB, Kirby A, Elliott AL, Parkin M, Hubbell E, Webster T, Mei R, Veitch J, Collins PJ, Handsaker R, Lincoln S, Nizzari M, Blume J, Jones KW, Rava R, Daly MJ, Gabriel SB, Altshuler D. Integrated detection and population-genetic analysis of SNPs and copy number variation. Nat Genet. 2008;40(10):1166–74. doi: 10.1038/ng.238. [DOI] [PubMed] [Google Scholar]
  • 100.Wang K, Li M, Hadley D, Liu R, Glessner J, Grant SF, Hakonarson H, Bucan M. PennCNV: an integrated hidden Markov model designed for high-resolution copy number variation detection in whole-genome SNP genotyping data. Genome Res. 2007;17(11):1665–74. doi: 10.1101/gr.6861907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Klein RJ, Zeiss C, Chew EY, Tsai JY, Sackler RS, Haynes C, Henning AK, SanGiovanni JP, Mane SM, Mayne ST, Bracken MB, Ferris FL, Ott J, Barnstable C, Hoh J. Complement factor H polymorphism in age-related macular degeneration. Science. 2005;308(5720):385–9. doi: 10.1126/science.1109557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Gudbjartsson DF, Walters GB, Thorleifsson G, Stefansson H, Halldorsson BV, Zusmanovich P, Sulem P, Thorlacius S, Gylfason A, Steinberg S, Helgadottir A, Ingason A, Steinthorsdottir V, Olafsdottir EJ, Olafsdottir GH, Jonsson T, Borch-Johnsen K, Hansen T, Andersen G, Jorgensen T, Pedersen O, Aben KK, Witjes JA, Swinkels DW, den Heijer M, Franke B, Verbeek AL, Becker DM, Yanek LR, Becker LC, Tryggvadottir L, Rafnar T, Gulcher J, Kiemeney LA, Kong A, Thorsteinsdottir U, Stefansson K. Many sequence variants affecting diversity of adult human height. Nat Genet. 2008;40(5):609–15. doi: 10.1038/ng.122. [DOI] [PubMed] [Google Scholar]
  • 103.Lettre G, Jackson AU, Gieger C, Schumacher FR, Berndt SI, Sanna S, Eyheramendy S, Voight BF, Butler JL, Guiducci C, Illig T, Hackett R, Heid IM, Jacobs KB, Lyssenko V, Uda M, Boehnke M, Chanock SJ, Groop LC, Hu FB, Isomaa B, Kraft P, Peltonen L, Salomaa V, Schlessinger D, Hunter DJ, Hayes RB, Abecasis GR, Wichmann HE, Mohlke KL, Hirschhorn JN. Identification of ten loci associated with height highlights new biological pathways in human growth. Nat Genet. 2008;40(5):584–91. doi: 10.1038/ng.125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Weedon MN, Lango H, Lindgren CM, Wallace C, Evans DM, Mangino M, Freathy RM, Perry JR, Stevens S, Hall AS, Samani NJ, Shields B, Prokopenko I, Farrall M, Dominiczak A, Johnson T, Bergmann S, Beckmann JS, Vollenweider P, Waterworth DM, Mooser V, Palmer CN, Morris AD, Ouwehand WH, Zhao JH, Li S, Loos RJ, Barroso I, Deloukas P, Sandhu MS, Wheeler E, Soranzo N, Inouye M, Wareham NJ, Caulfield M, Munroe PB, Hattersley AT, McCarthy MI, Frayling TM. Genome-wide association analysis identifies 20 loci that influence adult height. Nat Genet. 2008;40(5):575–83. doi: 10.1038/ng.121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Visscher PM. Sizing up human height variation. Nat Genet. 2008;40(5):489–90. doi: 10.1038/ng0508-489. [DOI] [PubMed] [Google Scholar]
  • 106.Reich DE, Lander ES. On the allelic spectrum of human disease. Trends Genet. 2001;17(9):502–10. doi: 10.1016/s0168-9525(01)02410-6. [DOI] [PubMed] [Google Scholar]
  • 107.Pritchard JK. Are rare variants responsible for susceptibility to complex diseases? Am J Hum Genet. 2001;69(1):124–37. doi: 10.1086/321272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Xu Z, He Z, Huang K, Tang W, Li Z, Tang R, Xu Y, Feng G, He L, Shi Y. No genetic association between NCAM1 gene polymorphisms and schizophrenia in the Chinese population. Prog Neuropsychopharmacol Biol Psychiatry. 2008;32(7):1633–6. doi: 10.1016/j.pnpbp.2008.06.007. [DOI] [PubMed] [Google Scholar]
  • 109.Ferreira MA, Sham P, Daly MJ, Purcell S. Ascertainment through family history of disease often decreases the power of family-based association studies. Behav Genet. 2007;37(4):631–6. doi: 10.1007/s10519-007-9149-0. [DOI] [PubMed] [Google Scholar]
  • 110.Sklar P, Smoller JW, Fan J, Ferreira MA, Perlis RH, Chambert K, Nimgaonkar VL, McQueen MB, Faraone SV, Kirby A, de Bakker PI, Ogdie MN, Thase ME, Sachs GS, Todd-Brown K, Gabriel SB, Sougnez C, Gates C, Blumenstiel B, Defelice M, Ardlie KG, Franklin J, Muir WJ, McGhee KA, MacIntyre DJ, McLean A, VanBeck M, McQuillin A, Bass NJ, Robinson M, Lawrence J, Anjorin A, Curtis D, Scolnick EM, Daly MJ, Blackwood DH, Gurling HM, Purcell SM. Whole-genome association study of bipolar disorder. Mol Psychiatry. 2008;13(6):558–69. doi: 10.1038/sj.mp.4002151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.WTCCC, Consortium WTCC. Genome-wide association study of 14,000 cases of seven common diseases and 3,000 shared controls. Nature. 2007;447(7145):661–78. doi: 10.1038/nature05911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Brett PM, Curtis D, Robertson MM, Gurling HM. The genetic susceptibility to Gilles de la Tourette syndrome in a large multiple affected British kindred: linkage analysis excludes a role for the genes coding for dopamine D1, D2, D3, D4, D5 receptors, dopamine beta hydroxylase, tyrosinase, and tyrosine hydroxylase. Biol Psychiatry. 1995;37(8):533–40. doi: 10.1016/0006-3223(94)00161-U. [DOI] [PubMed] [Google Scholar]
  • 113.Gelernter J, Kennedy JL, Grandy DK, Zhou QY, Civelli O, Pauls DL, Pakstis A, Kurlan R, Sunahara RK, Niznik HB, et al. Exclusion of close linkage of Tourette’s syndrome to D1 dopamine receptor. Am J Psychiatry. 1993;150(3):449–53. doi: 10.1176/ajp.150.3.449. [DOI] [PubMed] [Google Scholar]
  • 114.Thompson M, Comings DE, Feder L, George SR, O’Dowd BF. Mutation screening of the dopamine D1 receptor gene in Tourette’s syndrome and alcohol dependent patients. Am J Med Genet. 1998;81(3):241–4. [PubMed] [Google Scholar]
  • 115.Chou IC, Tsai CH, Lee CC, Kuo HT, Hsu YA, Li CI, Tsai FJ. Association analysis between Tourette’s syndrome and dopamine D1 receptor gene in Taiwanese children. Psychiatr Genet. 2004;14(4):219–21. doi: 10.1097/00041444-200412000-00010. [DOI] [PubMed] [Google Scholar]
  • 116.Tarnok Z, Ronai Z, Gervai J, Kereszturi E, Gadoros J, Sasvari-Szekely M, Nemoda Z. Dopaminergic candidate genes in Tourette syndrome: association between tic severity and 3′ UTR polymorphism of the dopamine transporter gene. Am J Med Genet B Neuropsychiatr Genet. 2007;144B(7):900–5. doi: 10.1002/ajmg.b.30517. [DOI] [PubMed] [Google Scholar]
  • 117.Devor EJ, Grandy DK, Civelli O, Litt M, Burgess AK, Isenberg KE, van de Wetering BJ, Oostra B. Genetic linkage is excluded for the D2-dopamine receptor lambda HD2G1 and flanking loci on chromosome 11q22-q23 in Tourette syndrome. Hum Hered. 1990;40(2):105–8. doi: 10.1159/000153914. [DOI] [PubMed] [Google Scholar]
  • 118.Gelernter J, Pakstis AJ, Pauls DL, Kurlan R, Gancher ST, Civelli O, Grandy D, Kidd KK. Gilles de la Tourette syndrome is not linked to D2-dopamine receptor. Arch Gen Psychiatry. 1990;47(11):1073–7. doi: 10.1001/archpsyc.1990.01810230089014. [DOI] [PubMed] [Google Scholar]
  • 119.Gelernter J, Pauls DL, Leckman J, Kidd KK, Kurlan R. D2 dopamine receptor alleles do not influence severity of Tourette’s syndrome. Results from four large kindreds. Arch Neurol. 1994;51(4):397–400. doi: 10.1001/archneur.1994.00540160099012. [DOI] [PubMed] [Google Scholar]
  • 120.Diaz-Anzaldua A, Joober R, Riviere JB, Dion Y, Lesperance P, Richer F, Chouinard S, Rouleau GA. Tourette syndrome and dopaminergic genes: a family-based association study in the French Canadian founder population. Mol Psychiatry. 2004;9(3):272–7. doi: 10.1038/sj.mp.4001411. [DOI] [PubMed] [Google Scholar]
  • 121.Nothen MM, Hebebrand J, Knapp M, Hebebrand K, Camps A, von Gontard A, Wettke-Schafer R, Lisch S, Cichon S, Poustka F, et al. Association analysis of the dopamine D2 receptor gene in Tourette’s syndrome using the haplotype relative risk method. Am J Med Genet. 1994;54(3):249–52. doi: 10.1002/ajmg.1320540311. [DOI] [PubMed] [Google Scholar]
  • 122.Comings DE, Comings BG, Muhleman D, Dietz G, Shahbahrami B, Tast D, Knell E, Kocsis P, Baumgarten R, Kovacs BW, et al. The dopamine D2 receptor locus as a modifying gene in neuropsychiatric disorders. Jama. 1991;266(13):1793–800. [PubMed] [Google Scholar]
  • 123.Comings DE, Wu S, Chiu C, Ring RH, Gade R, Ahn C, MacMurray JP, Dietz G, Muhleman D. Polygenic inheritance of Tourette syndrome, stuttering, attention deficit hyperactivity, conduct, and oppositional defiant disorder: the additive and subtractive effect of the three dopaminergic genes--DRD2, D beta H, and DAT1. Am J Med Genet. 1996;67(3):264–88. doi: 10.1002/(SICI)1096-8628(19960531)67:3<264::AID-AJMG4>3.0.CO;2-N. [DOI] [PubMed] [Google Scholar]
  • 124.Devor EJ. The D2 dopamine receptor and Tourette’s syndrome. Jama. 1992;267(5):651. doi: 10.1001/jama.267.5.651b. author reply 652. [DOI] [PubMed] [Google Scholar]
  • 125.Lee CC, Chou IC, Tsai CH, Wang TR, Li TC, Tsai FJ. Dopamine receptor D2 gene polymorphisms are associated in Taiwanese children with Tourette syndrome. Pediatr Neurol. 2005;33(4):272–6. doi: 10.1016/j.pediatrneurol.2005.05.005. [DOI] [PubMed] [Google Scholar]
  • 126.Brett P, Robertson M, Gurling H, Curtis D. Failure to find linkage and increased homozygosity for the dopamine D3 receptor gene in Tourette’s syndrome. Lancet. 1993;341(8854):1225. doi: 10.1016/0140-6736(93)91064-s. [DOI] [PubMed] [Google Scholar]
  • 127.Devor EJ, Dill-Devor RM, Magee HJ. The Bal I and Msp I polymorphisms in the dopamine D3 receptor gene display, linkage disequilibrium with each other but no association with Tourette syndrome. Psychiatr Genet. 1998;8(2):49–52. doi: 10.1097/00041444-199800820-00003. [DOI] [PubMed] [Google Scholar]
  • 128.Hebebrand J, Nothen MM, Lehmkuhl G, Poustka F, Schmidt M, Propping P, Remschmidt H. Tourette’s syndrome and homozygosity for the dopamine D3 receptor gene. German Tourette’s Syndrome Collaborative Research Group. Lancet. 1993;341(8858):1483–4. doi: 10.1016/0140-6736(93)90931-6. [DOI] [PubMed] [Google Scholar]
  • 129.Comings DE, Muhleman D, Dietz G, Dino M, LeGro R, Gade R. Association between Tourette’s syndrome and homozygosity at the dopamine D3 receptor gene. Lancet. 1993;341(8849):906. doi: 10.1016/0140-6736(93)93123-i. [DOI] [PubMed] [Google Scholar]
  • 130.Barr CL, Wigg KG, Zovko E, Sandor P, Tsui LC. No evidence for a major gene effect of the dopamine D4 receptor gene in the susceptibility to Gilles de la Tourette syndrome in five Canadian families. Am J Med Genet. 1996;67(3):301–5. doi: 10.1002/(SICI)1096-8628(19960531)67:3<301::AID-AJMG6>3.0.CO;2-P. [DOI] [PubMed] [Google Scholar]
  • 131.Comings DE, Gonzalez N, Wu S, Gade R, Muhleman D, Saucier G, Johnson P, Verde R, Rosenthal RJ, Lesieur HR, Rugle LJ, Miller WB, MacMurray JP. Studies of the 48 bp repeat polymorphism of the DRD4 gene in impulsive, compulsive, addictive behaviors: Tourette syndrome, ADHD, pathological gambling, and substance abuse. Am J Med Genet. 1999;88(4):358–68. doi: 10.1002/(sici)1096-8628(19990820)88:4<358::aid-ajmg13>3.0.co;2-g. [DOI] [PubMed] [Google Scholar]
  • 132.Hebebrand J, Nothen MM, Ziegler A, Klug B, Neidt H, Eggermann K, Lehmkuhl G, Poustka F, Schmidt MH, Propping P, Remschmidt H. Nonreplication of linkage disequilibrium between the dopamine D4 receptor locus and Tourette syndrome. Am J Hum Genet. 1997;61(1):238–9. doi: 10.1016/S0002-9297(07)64298-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Cruz C, Camarena B, King N, Paez F, Sidenberg D, de la Fuente JR, Nicolini H. Increased prevalence of the seven-repeat variant of the dopamine D4 receptor gene in patients with obsessive-compulsive disorder with tics. Neurosci Lett. 1997;231(1):1–4. doi: 10.1016/s0304-3940(97)00523-5. [DOI] [PubMed] [Google Scholar]
  • 134.Grice DE, Leckman JF, Pauls DL, Kurlan R, Kidd KK, Pakstis AJ, Chang FM, Buxbaum JD, Cohen DJ, Gelernter J. Linkage disequilibrium between an allele at the dopamine D4 receptor locus and Tourette syndrome, by the transmission-disequilibrium test. Am J Hum Genet. 1996;59(3):644–52. [PMC free article] [PubMed] [Google Scholar]
  • 135.Barr CL, Wigg KG, Zovko E, Sandor P, Tsui LC. Linkage study of the dopamine D5 receptor gene and Gilles de la Tourette syndrome. Am J Med Genet. 1997;74(1):58–61. [PubMed] [Google Scholar]
  • 136.Ozbay F, Wigg KG, Turanli ET, Asherson P, Yazgan Y, Sandor P, Barr CL. Analysis of the dopamine beta hydroxylase gene in Gilles de la Tourette syndrome. Am J Med Genet B Neuropsychiatr Genet. 2006;141B(6):673–7. doi: 10.1002/ajmg.b.30393. [DOI] [PubMed] [Google Scholar]
  • 137.Gelernter J, Vandenbergh D, Kruger SD, Pauls DL, Kurlan R, Pakstis AJ, Kidd KK, Uhl G. The dopamine transporter protein gene (SLC6A3): primary linkage mapping and linkage studies in Tourette syndrome. Genomics. 1995;30(3):459–63. doi: 10.1006/geno.1995.1265. [DOI] [PubMed] [Google Scholar]
  • 138.Vandenbergh DJ, Thompson MD, Cook EH, Bendahhou E, Nguyen T, Krasowski MD, Zarrabian D, Comings D, Sellers EM, Tyndale RF, George SR, O’Dowd BF, Uhl GR. Human dopamine transporter gene: coding region conservation among normal, Tourette’s disorder, alcohol dependence and attention-deficit hyperactivity disorder populations. Mol Psychiatry. 2000;5(3):283–92. doi: 10.1038/sj.mp.4000701. [DOI] [PubMed] [Google Scholar]
  • 139.Yoon DY, Rippel CA, Kobets AJ, Morris CM, Lee JE, Williams PN, Bridges DD, Vandenbergh DJ, Shugart YY, Singer HS. Dopaminergic polymorphisms in Tourette syndrome: association with the DAT gene (SLC6A3) Am J Med Genet B Neuropsychiatr Genet. 2007;144B(5):605–10. doi: 10.1002/ajmg.b.30466. [DOI] [PubMed] [Google Scholar]
  • 140.Rowe DC, Stever C, Gard JM, Cleveland HH, Sanders ML, Abramowitz A, Kozol ST, Mohr JH, Sherman SL, Waldman ID. The relation of the dopamine transporter gene (DAT1) to symptoms of internalizing disorders in children. Behav Genet. 1998;28(3):215–25. doi: 10.1023/a:1021427314941. [DOI] [PubMed] [Google Scholar]
  • 141.Gade R, Muhleman D, Blake H, MacMurray J, Johnson P, Verde R, Saucier G, Comings DE. Correlation of length of VNTR alleles at the X-linked MAOA gene and phenotypic effect in Tourette syndrome and drug abuse. Mol Psychiatry. 1998;3(1):50–60. doi: 10.1038/sj.mp.4000326. [DOI] [PubMed] [Google Scholar]
  • 142.Barr CL, Wigg KG, Sandor P. Catechol-O-methyltransferase and Gilles de la Tourette syndrome. Mol Psychiatry. 1999;4(5):492–5. doi: 10.1038/sj.mp.4000549. [DOI] [PubMed] [Google Scholar]
  • 143.Cavallini MC, Di Bella D, Catalano M, Bellodi L. An association study between 5-HTTLPR polymorphism, COMT polymorphism, and Tourette’s syndrome. Psychiatry Res. 2000;97(2–3):93–100. doi: 10.1016/s0165-1781(00)00220-1. [DOI] [PubMed] [Google Scholar]
  • 144.Xu C, Ozbay F, Wigg K, Shulman R, Tahir E, Yazgan Y, Sandor P, Barr CL. Evaluation of the genes for the adrenergic receptors alpha 2A and alpha 1C and Gilles de la Tourette Syndrome. Am J Med Genet B Neuropsychiatr Genet. 2003;119B(1):54–9. doi: 10.1002/ajmg.b.20001. [DOI] [PubMed] [Google Scholar]
  • 145.Chou IC, Tsai CH, Wan L, Hsu YA, Tsai FJ. Association study between Tourette’s syndrome and polymorphisms of noradrenergic genes (ADRA2A, ADRA2C) Psychiatr Genet. 2007;17(6):359. doi: 10.1097/YPG.0b013e3281ac2358. [DOI] [PubMed] [Google Scholar]
  • 146.Stober G, Hebebrand J, Cichon S, Bruss M, Bonisch H, Lehmkuhl G, Poustka F, Schmidt M, Remschmidt H, Propping P, Nothen MM. Tourette syndrome and the norepinephrine transporter gene: results of a systematic mutation screening. Am J Med Genet. 1999;88(2):158–63. doi: 10.1002/(sici)1096-8628(19990416)88:2<158::aid-ajmg12>3.0.co;2-w. [DOI] [PubMed] [Google Scholar]
  • 147.Rippel CA, Kobets AJ, Yoon DY, Williams PN, Shugart YY, Bridges DD, Vandenbergh DJ, Singer HS. Norepinephrine transporter polymorphisms in Tourette syndrome with and without attention deficit hyperactivity disorder: no evidence for significant association. Psychiatr Genet. 2006;16(5):179–80. doi: 10.1097/01.ypg.0000218622.96127.71. [DOI] [PubMed] [Google Scholar]
  • 148.Brett PM, Curtis D, Robertson MM, Gurling HM. Neuroreceptor subunit genes and the genetic susceptibility to Gilles de la Tourette syndrome. Biol Psychiatry. 1997;42(10):941–7. doi: 10.1016/S0006-3223(97)00140-6. [DOI] [PubMed] [Google Scholar]
  • 149.Brett PM, Curtis D, Robertson MM, Gurling HM. Exclusion of the 5-HT1A serotonin neuroreceptor and tryptophan oxygenase genes in a large British kindred multiply affected with Tourette’s syndrome, chronic motor tics, and obsessive-compulsive behavior. Am J Psychiatry. 1995;152(3):437–40. doi: 10.1176/ajp.152.3.437. [DOI] [PubMed] [Google Scholar]
  • 150.Erdmann J, Shimron-Abarbanell D, Cichon S, Albus M, Maier W, Lichtermann D, Minges J, Reuner U, Franzek E, Ertl MA, et al. Systematic screening for mutations in the promoter and the coding region of the 5-HT1A gene. Am J Med Genet. 1995;60(5):393–9. doi: 10.1002/ajmg.1320600509. [DOI] [PubMed] [Google Scholar]
  • 151.Huang Y, Liu X, Li T, Guo L, Sun X, Xiao X, Ma X, Wang Y, Collier DA. Cases-Control association study and transmission disequilibrium test of T102C polymorphism in 5HT2A and Tourette syndrome. Zhonghua Yi Xue Yi Chuan Xue Za Zhi. 2001;18(1):11–3. [PubMed] [Google Scholar]
  • 152.Dickel DE, Veenstra-VanderWeele J, Bivens NC, Wu X, Fischer DJ, Van Etten-Lee M, Himle JA, Leventhal BL, Cook EH, Jr, Hanna GL. Association studies of serotonin system candidate genes in early-onset obsessive-compulsive disorder. Biol Psychiatry. 2007;61(3):322–9. doi: 10.1016/j.biopsych.2006.09.030. [DOI] [PubMed] [Google Scholar]
  • 153.Niesler B, Frank B, Hebebrand J, Rappold G. Serotonin receptor genes HTR3A and HTR3B are not involved in Gilles de la Tourette syndrome. Psychiatr Genet. 2005;15(4):303–4. doi: 10.1097/00041444-200512000-00015. [DOI] [PubMed] [Google Scholar]
  • 154.Gelernter J, Rao PA, Pauls DL, Hamblin MW, Sibley DR, Kidd KK. Assignment of the 5HT7 receptor gene (HTR7) to chromosome 10q and exclusion of genetic linkage with Tourette syndrome. Genomics. 1995;26(2):207–9. doi: 10.1016/0888-7543(95)80202-w. [DOI] [PubMed] [Google Scholar]
  • 155.Comings DE, Gade R, Muhleman D, Chiu C, Wu S, To M, Spence M, Dietz G, Winn-Deen E, Rosenthal RJ, Lesieur HR, Rugle L, Sverd J, Ferry L, Johnson JP, MacMurray JP. Exon and intron variants in the human tryptophan 2,3-dioxygenase gene: potential association with Tourette syndrome, substance abuse and other disorders. Pharmacogenetics. 1996;6(4):307–18. doi: 10.1097/00008571-199608000-00004. [DOI] [PubMed] [Google Scholar]
  • 156.Comings DE, Chen C, Wu S, Muhleman D. Association of the androgen receptor gene (AR) with ADHD and conduct disorder. Neuroreport. 1999;10(7):1589–92. doi: 10.1097/00001756-199905140-00036. [DOI] [PubMed] [Google Scholar]
  • 157.Hebebrand J, Nothen MM, Klug B, Wettke-Schafer R, Camps A, Lisch S, Hemmer S, von Gontard A, Poustka F, Lehmkuhl G, et al. No association between length of the (CAG)n repeat of the Huntington’s disease gene and Tourette’s syndrome. Biol Psychiatry. 1995;37(3):209–11. doi: 10.1016/0006-3223(94)00214-n. [DOI] [PubMed] [Google Scholar]
  • 158.Schoenian S, Konig I, Oertel W, Remschmidt H, Ziegler A, Hebebrand J, Bandmann O. HLA-DRB genotyping in Gilles de la Tourette patients and their parents. Am J Med Genet B Neuropsychiatr Genet. 2003;119B(1):60–4. doi: 10.1002/ajmg.b.20003. [DOI] [PubMed] [Google Scholar]
  • 159.Huang Y, Li T, Wang Y, Ansar J, Lanting G, Liu X, Zhao JH, Hu X, Sham PC, Collier D. Linkage disequilibrium analysis of polymorphisms in the gene for myelin oligodendrocyte glycoprotein in Tourette’s syndrome patients from a Chinese sample. Am J Med Genet B Neuropsychiatr Genet. 2004;124B(1):76–80. doi: 10.1002/ajmg.b.20079. [DOI] [PubMed] [Google Scholar]
  • 160.Miranda DM, Wigg K, Feng Y, Sandor P, Barr CL. Association study between Gilles de la Tourette Syndrome and two genes in the Robo-Slit pathway located in the chromosome 11q24 linked/associated region. Am J Med Genet B Neuropsychiatr Genet. 2008;147B(1):68–72. doi: 10.1002/ajmg.b.30580. [DOI] [PubMed] [Google Scholar]
  • 161.Mossner R, Muller-Vahl KR, Doring N, Stuhrmann M. Role of the novel tryptophan hydroxylase-2 gene in Tourette syndrome. Mol Psychiatry. 2007;12(7):617–9. doi: 10.1038/sj.mp.4002004. [DOI] [PubMed] [Google Scholar]
  • 162.Dudbridge F, Gusnanto A. Estimation of significance thresholds for genomewide association scans. Genet Epidemiol. 2008;32(3):227–34. doi: 10.1002/gepi.20297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Purcell S, Cherny SS, Sham PC. Genetic Power Calculator: design of linkage and association genetic mapping studies of complex traits. Bioinformatics. 2003;19(1):149–50. doi: 10.1093/bioinformatics/19.1.149. [DOI] [PubMed] [Google Scholar]

RESOURCES