Genetics of dyslexia: the evolving landscape (2024)

As a library, NLM provides access to scientific literature. Inclusion in an NLM database does not imply endorsem*nt of, or agreement with, the contents by NLM or the National Institutes of Health.
Learn more: PMC Disclaimer | PMC Copyright Notice

Genetics of dyslexia: the evolving landscape (1)

Link to Publisher's site

J Med Genet. 2007 May; 44(5): 289–297.

Published online 2007 Feb 16. doi:10.1136/jmg.2006.046516

PMCID: PMC2597981

PMID: 17307837

Author information Article notes Copyright and License information PMC Disclaimer

Abstract

Dyslexia is among the most common neurodevelopmental disorders, with a prevalence of 5–12%. At the phenotypic level, various cognitive components that enable reading and spelling and that are disturbed in affected individuals can be distinguished. Depending on the phenotype dimension investigated, inherited factors are estimated to account for up to 80%. Linkage findings in dyslexia are relatively consistent across studies in comparison to findings for other neuropsychiatric disorders. This is particularly true for chromosome regions 1p34–p36, 6p21–p22, 15q21 and 18q11. Four candidate genes have recently been identified through systematic linkage disequilibrium studies in linkage region 6p21–p22, and through cloning approaches at chromosomal breakpoints. Results indicate that a disturbance in neuronal migration is a pathological correlate of dyslexia at the functional level. This review presents a summary of the latest insights into the genetics of dyslexia and an overview of anticipated future developments.

Dyslexia is among the most common neurodevelopmental disorders, with a prevalence of 5–12%.1,2 The prevalence varies with the use of different diagnostic criteria and, since reading and spelling are normally distributed in the population, is influenced by the cut‐off point applied to the psychometric tests. According to the International Classification of Diseases‐10, dyslexia is “a disorder manifested by difficulty learning to read despite conventional instruction, adequate intelligence and sociocultural opportunity”.3 Longitudinal studies have shown that the disorder involves an extremely stable developmental disturbance that does not, in contrast to popular opinion, disappear with adolescence.4 The psychosocial consequences are correspondingly grave. Affected individuals attain a much lower educational level and have substantially higher rates of unemployment and psychosocial stress than would be expected for their level of intelligence.5,6,7 In childhood, approximately 20% of those with dyslexia also present with attention‐deficit hyperactivity disorder (ADHD),8,9,10,11,12 whereas in adolescence depressive disorders and disorders of social behaviour are often associated with dyslexia.13,14,15 Whether dyslexia is more common among boys than girls has been part of a controversial discussion in the past, although recent epidemiological studies indicate a twofold increase in the risk for boys compared with that in girls.2,16 The sex ratio may be influenced by severity, IQ and assessed cognitive profiles.17

Familial clustering in dyslexia was recognised a few years after the first description of the disorder by Hinshelwood in 1895.18,19,20 A child with an affected parent has a risk of 40–60% of developing dyslexia. This risk is increased when other family members are also affected.19,21,22,23,24,25 There is an estimated 3–10‐fold increase in the relative risk for a sibling (λs), with an increase in λs observed when strict criteria are applied.25 Twin studies have confirmed that genetic factors are substantially responsible for the familial clustering of dyslexia.17,26 The proportion of inherited factors involved in the development of dyslexia is between 40% and 80%, the highest estimates being reported for the phenotype dimensions word reading (up to 58%) and spelling (70%).17,26,27 Twin studies have allowed for the estimation of heritabilities and also the impact of shared and non‐shared environmental factors. Although shared environmental effects are low for word reading, they are substantially higher (at about 14%) for reading and spelling correlated traits—for example, phonological awareness.27

Whether or not sex has an influence on heritability is controversial. Although the results of a US American twin study (Colorado Twin Study) showed similar heritability between the sexes,28,29 Harlaar et al30 found a higher heritability for boys in a UK sample (London Twins Early Development Study).

Through molecular genetic linkage studies in families with dyslexia, chromosome regions have been identified in which the presence of dyslexia susceptibility genes is suspected. As with all complex disorders, linkage findings are not completely overlapping between independent studies. However, greater consistency is reported for dyslexia than for most other neuropsychiatric disorders, and the identification of the first candidate genes therefore came as no surprise.

This review presents the current state of molecular genetic research on dyslexia, including discussion of the phenotypic aspects and neuropsychological concepts of dyslexia that have received increasing consideration in genetic research over recent years. Finally, the extent to which our understanding of dyslexia is likely to be increased through the results of current and future molecular genetic research is discussed.

Phenotypic aspects and neurophysiological theories

In general, the cognitive processes on which reading and spelling are based are complex, and differing cognitive dimensions ease the separate skills of reading and spelling. Such processes include those of short‐term memory, phonological awareness, rapid naming, and phonological and orthographic coding (table 11).). In recent years, several theories have been developed with the aim of characterising the basic processes underlying dyslexia. These have taken into consideration the increasing body of knowledge obtained from neurophysiological and imaging research (eg, event‐related potentials, functional MRI). The phonological deficit theory,32 which assumes a disturbance in phonological processing, is currently the most salient theory. According to this theory, affected individuals have difficulties in perceiving and segmenting phonemes, leading to difficulties in establishing a connection between phonemes and graphemes. The rapid auditory processing theory is another theory33 that proposes that phonological deficits are secondary to an auditory deficit in the perception of short or rapidly varying sounds. Many individuals with dyslexia perform poorly on auditory tasks including frequency discrimination34,35 and temporal order judgement.36 Abnormal neurophysiological responses to auditory stimuli have also been reported.36,37,38 However, individuals with dyslexia also have visual perceptual deficits which these theories cannot adequately explain. The magnocellular theory accounts for disturbances in visual processing.39,40,41,42 The theory proposes that in a proportion of individuals with dyslexia, the perception of visual, rapid moving stimuli and stimuli of low spatial frequency and low contrast is impaired. This deficit is associated at the central nervous system (CNS) level with impaired sensitivity of cells within the retinocortical magnocellular pathway and in the extrastriate areas in the dorsal stream to which they project. The cerebellar deficit theory suggests that the automatisation of cognitive processes and motor control in the cerebellum are disturbed in individuals with dyslexia.43 The double deficit hypothesis,44 which assumes disturbances in phonological processing and the speed of processing, should also be mentioned in this context.

Table 1 Cognitive components involved in reading and writing

ComponentRemark
Visual processingThe magnocellular system responds to moving stimuli and stimuli of low spatial frequency and low contrast. Impaired perception of moving stimuli and the neurophysiological correlates of this have been found repeatedly in individuals with dyslexia. The exact nature of this deficiency and its potential relationship to dyslexia is not yet clear
Phonological awarenessThe ability to perceive, segment and manipulate the sounds of spoken words.31 Phonemes are the smallest meaningfully distinct sounds from which an acoustic speech flow can be constructed. The word dog, for example, consists of three phonemes /d/, /o/, /g/. The capacity for phonological awareness is often tested through a phoneme deletion task
Verbal short‐term memoryVarious aspects of memory are required for reading. Many known words are no longer dissected into their phonemes, but are recalled directly from memory. Processing of unknown words into their phonemes occurs in short‐term memory. Short‐term memory is often examined by a digit span task
Phonological codingThe ability to put together the phonemes and then verbally express words which have never been previously read or heard. This ability is tested through reading of pseudowords
Orthographic codingOrthographic coding refers to the assumed process of recognising a word by its holistic form. Orthographic coding is measured by a pseudohom*ophone task where an orally presented word has to be compared with a visual presentation of two phonologically indistinguishable words, of which one may be orthographically correct
Rapid namingRapid naming is a measure of the speed of processing. The naming of objects, numbers, letters and colours is typically measured

Even though evidence for one or the other of these theories is typically reported in affected individuals, there is no evidence so far for specific subgroups of dyslexia. A reason for this could be that although some of the deficits found in affected individuals are correlated with reading and spelling, they may not be causally associated with dyslexia. Findings from genetic research may have the potential to help delineate which cognitive and neurophysiological processes are causally related.45

Linkage findings in dyslexia

To date, linkage analyses in families with dyslexia have identified nine chromosome regions (dyslexia susceptibility 1(DYX1)–dyslexia susceptibility 9 (DYX9)) listed by the HUGO Gene Nomenclature Committee in which the presence of susceptibility genes is suspected (table 22).). There was initially great hope that it would be possible to correlate the respective cognitive components of dyslexia (table 11)) with specific linkage regions. Many studies accordingly investigated individual phenotype components as categorical or quantitative (quantitative trait loci (QTL)) subdimensions, and linkages with specific chromosomal regions have been claimed; unfortunately, with little support from independent studies so far. Nevertheless, the consistency of linkage findings is impressive in comparison to those for other neuropsychiatric disorders. This is particularly true of findings in chromosome regions 1p34–p36, 6p21–p22, 15q21 and 18q11, with support for each of these regions coming from the investigation of at least two large family samples.

Table 2 Summary of linkage findings in dyslexia

LocusNumber of families (individuals, sib pairs), countryLinkage evidenceLinkage evidence with components of the phenotypeStudy designReferences
DYX19 multiplex families (84 individuals), USALOD score 3.20ReadingCategoricalSmith et al46
6 multiplex families (94 individuals), USALOD score 3.15Single‐word readingCategoricalGrigorenko et al47
7 multiplex families (67 individuals), GermanyLOD score 1.78SpellingCategoricalSchulte‐Körne et al48
90 families (611 individuals), USALOD score 2.34Single‐word readingCategoricalChapman et al49
DYX219 multiplex families (358 individuals*, 50 dizygotic twin‐pair), USAp Values between 0.009–0.04DyslexiaQTLCardon et al50
82 families (181 sib pairs), UKp Values between 0.0004–0.038Orthographic and phonological processesQTLFisher et al51
79 families (126 sib pairs), USALOD scores between 2.42 and 3.10Orthographic and phonological processesQTLGayan et al52
89 families (195 sib pairs)†, UKp Values between 0.00001–0.042Phonological decodingQTLFisher et al53
119 families (180 sib pairs)‡, USAp Values between 0.002–0.006Phonological decoding
104 families (392 individuals)‡, USAp Values between 0.0005–0.05Orthographic and phonological processesQTLKaplan et al54
8 multiplex families (176 individuals)§LOD scores of 1.52 and 2.56Single‐word reading, phoneme awarenessCategoricalGrigorenko et al55
DYX31 multiplex family (36 individuals), NorwayLOD score 4.32DyslexiaCategoricalfa*gerheim et al56
89 families (195 sib pairs), UKp Value <0.001Orthographic choiceQTLFisher et al53
119 families (180 sib pairs), USAp Value of 0.001Phonological awarenessQTL
96 families (877 individuals), CanadaLOD scores of 1.13 and 3.82Phonological coding,spellingQTL and categoricalPetryshen et al57
11 multiplex families (97 individuals), FinlandLOD score 3.01DyslexiaCategaricalKaminen et al58
DYX496 families (877 individuals), CanadaLOD scores of 2.08 and 3.34Phonological coding,spellingQTL and categoricalPetryshen et al59
DYX51 multiplex family (74 individuals), FinlandLOD score 3.84DyslexiaCategoricalNopola‐Hemmi et al60
DYX689 families (195 sib pairs), UKp Value <0.001Single‐word readingQTLFisher et al53
119 families (180 sib pairs), USAp Value <0.001Single‐word readingQTL
84 families (143 sib pairs), UKp Value <0.001Phoneme awarenessQTL
DYX7100 families (914 individuals), Canadap Value <0.001DyslexiaCategoricalHsiung et al61
DYX89 families, USALOD score of 1.95 and 2.33DyslexiaQTLRabin et al62
8 multiplex families (165 individuals), USALOD scores of 3.00 and 2.30Single‐word reading, phonological decodingCategoricalGrigorenko et al63
100 families (914 individuals), CanadaLOD scores of 4.01 and 1.65Spelling, phonological codingQTL and categoricalTzenova et al64
DYX91 multiplex family (29 individuals), NetherlandsLOD score 3.68DyslexiaCategoricalde Kovel et al65
89 families (195 sib pairs), UKp Value of 0.001Single‐word readingQTLFisher et al53

LOD, logarithm of the odds; QTL, quantitative trait loci.

*Including 18 families with linkage evidence at DYX2 previously reported by Smith et al.46

†Including 82 families with linkage evidence at DYX2 previously reported by Fisher et al.51

‡Including 39 families with linkage evidence at DYX2 previously reported by Cardon et al50 and 70 families with linkage evidence at DYX2 previously reported by Gayan et al.52

§Including 8 families with linkage evidence at DYX2 previously reported by Grigorenko et al.47,55

The largest family samples reported in the literature are from the USA (Colorado, Seattle and Yale samples), the UK (Cardiff and Oxford samples), Canada (Toronto and Vancouver samples) and Germany (German sample). For the sake of clarity, these samples will be named according to their origin in the following sections. Results from genomewide linkage studies have been reported so far from the Seattle,49,66,67 Oxford and Colorado samples.53 In addition, genomewide linkage studies of large multiply affected families from Holland,65 Norway56 and Finland58,60 have been reported. The following section presents results for the individual regions, and discussion is limited to positive findings only.

DYX1—chromosome 15q21

DYX1 (MIM 127700) lies in chromosome region 15q21, and a total of four research groups have reported linkage in their family samples (table 22).46,47,48,49 Evidence for linkage was found for word reading and related phenotype dimensions in three samples (Colorado, Yale and Seattle samples),46,47,49 and one sample showed evidence of linkage for spelling (German sample).48 Two linkage disequilibrium (LD) studies have been carried out in region DYX1 using short tandem repeat markers,68,69 and positive evidence for association was obtained for one region of approximately 4 Mb. In both studies, a three‐marker haplotype was associated in a total of three independent trio‐samples, two samples of British origin (Cardiff sample) and one of Italian origin.68,69 Region 15q21 has also shown evidence of linkage to ADHD. A genome scan carried out in 164 Dutch sib pairs with ADHD showed the strongest evidence for linkage in this region.70 The risk‐conferring gene in DYX1 may contribute to the comorbidity reported between the two disorders.

DYX2— chromosome 6p21–p22

The chromosome region 6p21–p22 (DYX2, MIM 600202) is considered to be among the best‐replicated regions of linkage for dyslexia (table 22).). Evidence of linkage has been reported using a QTL approach in both a US‐American (Colorado) and a UK (Oxford) sample.53,50,51,52,54 Positive evidence for linkage was also reported from a US‐American subsample (Yale sample) in which categorical phenotype dimensions had been considered.55 A more precise containment of the phenotype subdimensions associated with DYX2 was not possible. Linkage was found with the phenotypes phonological processing 51,52,53,54 and orthographic processing.51,52,54,55 Meanwhile, LD mapping in DYX2 led to the identification of two strong candidate genes (DCDC2 (doublecortin domain containing protein 2) and K1AA0319).71,72,73,74,75 Interestingly, evidence for linkage has also been found in the chromosome region 6p21–p22 for ADHD.76

DYX3—chromosome 2p15–p16

The chromosome region 2p15–p16 (DYX3, MIM 604254) has been identified through linkage analyses in five family samples (including the Oxford, Colorado and Vancouver samples; table 22).53,56,57,58,77 The linkage peaks of the individual studies lie far apart from each other, however, and so it is not clear whether they indicate the same susceptibility locus. As with DYX2, no phenotype dimension has been found to be specifically linked with this locus, although not all studies have analysed subdimensions.

DYX4—chromosome 6q11–q12

The chromosome region 6q11–q12 (DYX4, MIM 127700) was identified in the context of a chromosome‐wide linkage study of a large Canadian family sample (Vancouver sample; table 22).59 The most strongly linked phenotype dimensions were phonological coding and spelling. There has so far been no independent replication of this finding for DYX4.

DYX5—chromosome 3p12–q13

The chromosome region 3p12–q13 (DYX5, MIM 606 896) showed linkage in a large Finnish family (table 22).60ROBO1 (roundabout Drosophila hom*olog of 1) has been identified as a possible candidate gene in this region. DYX5 also showed a positive evidence for linkage in 77 US‐American families with speech–sound disorder (SSD).78 SSD involves impairments in phonological processing, as with dyslexia.

DYX6—chromosome 18p11

DYX6 (MIM 606616), which lies in chromosome region 18p11, was identified in two independent family samples (Oxford and Colorado samples) through a genome scan applying a QTL approach (table 22).53 The strongest evidence for linkage was found for word reading. This finding was replicated in a third family sample (expanded Oxford sample), the strongest evidence for linkage being found for the phenotype subdimension phoneme awareness.53 The results of a subsequent multivariate analysis in the two Oxford samples indicate that a QTL in DYX6 influences multiple aspects of reading ability and is not correlated with specific phenotype subdimensions.79

DYX7—chromosome 11p15

Linkage with markers in the region of DYX7 (MIM 127700), which lies in chromosome region 11p15, has been described only in one family sample to date (Vancouver sample; table 22).61 The authors selected DYX7 as a candidate region on the basis that the gene for the dopamine D4 receptor (DRD4) is localised there. DRD4 is a possible risk gene for ADHD.80

DYX8—chromosome 1p34–1p36

Three research groups in total have reported linkage between DYX8 (MIM 608995) in chromosome region 1p34–p36 and dyslexia (including the Yale and Vancouver samples; table 22).62,63,64 Even though individual studies have shown linkage to differing phenotype subdimensions of dyslexia, linkage evidence from two studies was particularly strong when focus was placed on the phonological aspects of dyslexia.63,64

DYX9—chromosome Xq26–q27

Evidence for linkage was found in chromosome region Xq27 (DYX9, MIM 300509) in a Dutch multiplex family with dyslexia (table 22).65 The same research group failed to replicate their result in 67 affected sib pairs. However, positive evidence for linkage was found in region DYX9 in one of the UK samples (Oxford sample; table 22).53

Additional linkage regions in dyslexia

In addition to the HGNC‐listed DYX1–DYX9 regions, linkage with dyslexia has also been reported for other regions, although without replication in independent samples. This includes evidence for linkage on chromosome 13q12 for word reading,66 and on chromosome 2q22 for phonological decoding efficiency.67 Two further studies have been conducted which aimed to identify chromosomal loci with pleiotropic effects on dyslexia and ADHD. In the Colorado sample, families with dyslexia having ADHD problems showed evidence for linkage in chromosome regions 14q32, 13q32 and 20q11.81 In families with ADHD, evidence for linkage is shown for reading ability in regions 10q11, 16p12 and 17q22.82

Candidate gene findings in dyslexia

Of the newly identified candidate genes, DCDC2 and K1AA0319 seem to be of most significance for dyslexia. Both were identified through systematic investigation of LD (LD mapping) within DYX2 on chromosome 6p22. Initial findings for both genes have been replicated in independent samples, with the strongest findings being reported among severely affected individuals. By contrast, the genes DYX1C1 (dyslexia susceptibility 1 candidate 1) and ROBO1, which were identified through breakpoint mapping in Finnish patients, seem to be less involved in the development of dyslexia across different populations. Their contribution may be limited to a few families in the Finnish population.

DCDC2 (doublecortin domain containing protein 2)

Initial evidence for the involvement of DCDC2 (MIM 605755) and dyslexia was obtained through gene‐based LD mapping in a gene‐dense 680 kb section of linkage region 6p22 (DXY2; table 33).72 The sample was drawn from 114 US‐American nuclear families of predominantly European origin (Colorado sample). Positive evidence for association was found in two genome loci, in which a total of six genes were localised: VMP/DCDC2/KAAG1 and K1AA0319/TTRAP/THEM2. In a subsequently expanded Colorado sample (153 nuclear families), the strongest evidence for association was found in DCDC2 (table 33).74 Additionally, a deletion of 2.4 kb in intron 2 of DCDC2, which encodes tandem repeats of putative brain‐associated transcription factor binding sites, was identified, which had an allele frequency of 8.5% in the parents.74 The tandem repeats in the deleted region demonstrate several alleles. For the purposes of the association study, the authors combined the deletion and the rare repeat alleles into one allele, for which they reported a strong association with reading performance.

Table 3 Summary of association findings in dyslexia

GeneStudy designSample characteristics, countryGenotyped variantsResultsReferences
DYX1C1Case–control55 cases* vs 113 controls, Finland8 SNPsSignificant association at the single‐marker and haplotype levelTaipale et al83
Case–control54 cases* vs 82 controls, Finland8 SNPs†Significant association at the single‐marker level
Family based148 nuclear families, Canada6 SNPs†Significant association at the single‐marker and haplotype level, association in the opposite direction compared to Taipale et al83Wigg et al84
Family based264 nuclear families, UK8 SNPs†Significant association at the single‐marker level, association in the opposite direction compared to Taipale et al83Scerri et al85
Family based150 nuclear families, USA2 SNPs†No associationMeng et al86
Family based158 nuclear families, Italy3 SNPs†No associationMarino et al87
Case–control57 cases vs 96 controls, Italy3 SNPs†No associationBellini et al88
Family based247 nuclear families, UK3 SNPs†No associationCope et al89
DCDC2Family based114 nuclear families, USA31 SNPs within 680 kb (including VMP, DCDC2, KAAG1, KIAA0319, TTRAP and THEM2)Strongest association at the single‐marker and haplotype level within the VMP/DCDC2/KAAG1 locusDeffenbacher et al72
Family based153 nuclear families, USA147 SNPs within 1.5 Mb (including VMP, DCDC2, KAAG1, KIAA0319, TTRAP and THEM2)Strongest association at the single‐marker and haplotype level within DCDC2Meng et al74
Case–control240 cases vs 312 controls, UK‡137 SNPs within VMP, DCDC2, KAAG1, KIAA0319, TTRAP and THEM2No association within the VMP/DCDC2/KAAG1locusCope et al71
Family based137 triads, Germany18 SNPs and 4 STRs within the VMP/DCDC2/KAAG1‐locusStrongest association at the single‐marker and haplotype level within DCDC2, strongest results with severity selectionSchumacher et al75
Family based239 triads, Germany2 SNPs, 1 STR within DCDC2Significant association at the haplotype level, strongest results with severity selection
KIAA0319Family based114 nuclear families, USA31 SNPs within 680 kb (including VMP, DCDC2, KAAG1, KIAA0319, TTRAP and THEM2)Significant association at the single‐marker and haplotype level within the KIAA0319/TTRAP/THEM2 locusDeffenbacher et al72
Family based42 nuclear families, UK§31 SNPs within 680 kb (including the KIAA0319/TTRAP/THEM2 locus)Strongest association at the single‐marker and haplotype level within KIAA0319 and TTRAPFrancks et al73
Family based84 nuclear families, UK§20 SNPs within KIAA0319 and TTRAPSignificant association at the single‐marker and haplotype level
Family based124 nuclear families, USA§21 SNPs within KIAA0319 and TTRAPSignificant association at the single‐marker and haplotype levelCope et al71
Case– control240 cases vs 312 controls, UK‡¶137 SNPs within VMP, DCDC2, KAAG1, KIAA0319, TTRAP and THEM2Strongest association at the single‐marker and haplotype level within KIAA0319
Case–control223 cases vs 273 controls, UK‡¶10 SNPs within the KIAA0319/TTRAP/THEM2‐locusStrongest association at the single‐marker and haplotype level within KIAA0319Schumacher et al75
Family based376 triads, Germany10 SNPs within the KIAA0319/TTRAP/THEM2 locusNominal significant association at the single‐marker level for 1 variant within KIAA0319 in the most severely affected subsampleHarold et al90
Family based126 nuclear families, UK§16 SNPs within DCDC2, KIAA0319 and flanking regionStrongest association at the single‐marker level within 20 kB in intron1 of KIAA0319
Case–control350 cases vs 273 controls, UK28 SNPs and 1 STR within DCDC2, KIAA0319 and flanking regionsEvidence for gene–gene interaction betweenKIAA0319 and DCDC2
Case–control419 cases vs 273 controls, UK4 SNPs and 1 STR within DCDC2 and 5 SNPS within KIAA0319

DCDC2, doublecortin domain containing protein 2gene; SNP, single‐nucleotide polymorphism.

*Some cases were extracted from the same families and are related.

†Including the two significantly associated SNPs reported by Taipale et al.83

‡Cases and controls were analysed using a DNA pooling approach.

§Samples represent subsamples selected for severity of the phenotype.

¶Individual genotyping of markers that were associated in the pooled samples (most of the cases and controls are identical).

Findings from two trio‐samples also indicate the involvement of DCDC2 in the development of dyslexia (German sample; table 33).75 Strong evidence for association was shown in both samples at the single‐marker and haplotype level. This effect seemed to be particularly substantial in severely affected individuals. In the pooled sample, severely affected individuals showed a genotypic relative risk of 4.88 on the basis of the hom*ozygous presence of the identified risk haplotype.

By contrast, investigation of the DCDC2 locus in the two UK samples (Oxford and Cardiff) had inconsistent results. In the Oxford sample, evidence of association between DCDC2 variants and various phenotype components of dyslexia were found, albeit with a weak level of significance. This association disappeared, however, when only severely affected cases were included in the analysis. Interestingly, the 2.4 kb deletion in intron 2 of DCDC2 was more common than by chance in severely affected patients. There was no association between dyslexia and DCDC2 in the Cardiff sample. Joint analysis of the two samples, however, produced evidence of a possible interaction between DCDC2 and K1AA0319.90

In summary, these results suggest that DCDC2 is involved in the development of dyslexia. It is unlikely that KAAG1 is the susceptibility gene at this locus. KAAG1 overlaps at the genomic level with exon 1 of DCDC2, although KAAG1 does not seem to be expressed in the CNS.74 By contrast, DCDC2 is widely expressed in the CNS, including areas of the brain in which lower activation patterns have been observed in individuals with dyslexia, such as the inferior temporal and medial temporal cortices.74,75,91,92,93

Functionally, DCDC2 is involved in processes of cortical neuronal migration during brain development and contains a double cortin hom*ology domain which is typical of this. RNA interference studies of in utero rats have shown that downregulation of DCDC2 leads to a significant reduction in neuronal migration.74 Determining whether the intron 2 deletion is one of the responsible variants will require further investigation in larger samples. There is no real rationale for combining the deletion with rare alleles of the STR polymorphism. Functional studies of the possible effect of the different alleles on expression or splicing are required to justify the combining of alleles.

KIAA0319

Besides evidence for association in the region of DCDC2, positive association with variants in the region of the K1AA0319/TTRAP/THEM2 gene cluster (MIM 609269) was reported in the Colorado sample.72 Association for the same gene cluster was reported by Francks et al in two independent samples (Oxford samples), which was particularly notable in severely affected individuals (table 33).73 Association in this region was replicated in a third UK sample (Cardiff sample; table 33).71 There was an association with SNPs in the region of KIAA0319 through the use of a DNA pooling screening step and subsequent replication through individual genotyping.

Meanwhile, further analyses of the two samples (Oxford and Cardiff) have shown that the responsible gene variant(s) is (are) probably localised near exon 1 of K1AA0319. Investigation of both UK samples has resulted in evidence of a gene–gene interaction between K1AA0319 and DCDC2.90

One sample (German sample), which had reported strong association with DCDC2, has so far produced no convincing evidence for association with the KIAA0319/TTRAP/THEM2 gene cluster.75 There was no further evidence of association at the K1AA0319 locus from the extended Colorado sample (153 nuclear families),74 although the genomic segment that had shown the strongest association findings in the two UK samples was insufficiently analysed.

The evidence of association for K1AA0319 obtained from independent samples is convincing. As with DCDC2, involvement of the KIAA0319 locus seems to be particularly marked in severely affected cases. Association findings, which were strongest around KIAA0319, and results from gene expression and functional studies suggest that KIAA0319 is the most likely susceptibility gene for dyslexia in this gene cluster. Allele‐specific expression analyses in lymphoblastoid cells have shown that carriers of the risk‐associated haplotype have a 40% reduction in the expression of KIAA0319, whereas the expression of other genes in this region remains unaffected.94 The expression of KIAA0319 is particularly strong in the cerebral neocortex of developing mouse and human brain tissue, and, similar to DCDC2, reduced expression of KIAA0319 through RNA interference leads to disturbed neuronal migration in rats in utero.94

DYX1C1

DYX1C1 (MIM 608706) was cloned in a two‐generation Finnish family with a translocation t(2;15)(q11;q21).83DYX1C1, which lies in chromosome region 15q21, is interrupted through the translocation breakpoint. All four family members in whom the translocation was detected showed reading‐associated problems.95 To determine whether DYX1C1 is of significance for affected cases outside of this family, a polymorphism discovery approach was used in 20 Finnish individuals with dyslexia. A total of eight SNPs were identified, which were then investigated in affected individuals and controls of Finnish origin. In an initial sample, two SNPs were found to be associated in the single‐marker and haplotype analysis. Replication was then achieved for one of the two variants in a second sample (table 33).). However, the sample sizes were limited, and a proportion of the affected individuals in the initial sample were related to each other.83

DYX1C1 is expressed in many tissues, including those of the CNS, where it is found in cortical neurones and white matter glial cells.83 Interestingly, it has recently been shown that DYX1C1, similar to KIAA0319 and DCDC2, functions in neuronal migration in rodent neocortex.96

Six other association analyses using independent samples of predominantly European origin have been carried out to date (including the Oxford, Cardiff, Colorado and Toronto samples).84,85,86,87,88,89 Overall, the results must be viewed as being negative, since the initial findings have not been replicated. Positive findings have been reported from two of these studies, although the association was with the opposite two‐marker haplotype (Oxford and Toronto samples; table 33).84,85 Given this failure to replicate, it is unlikely that DYX1C1 makes a significant contribution to the development of dyslexia in non‐Finnish European populations.

It is highly probable that the linkage findings in chromosome region 15q21 (DYX1) cannot be traced back to DYX1C1, since DYX1C1 lies outside of the linkage peaks. Whether or not DYX1C1 contributes to dyslexia in the Finnish population requires clarification through larger association studies.

ROBO1

As with DYX1C1, the identification of ROBO1 (MIM 602430) was achieved through breakpoint mapping of a translocation. A translocation, which had probably occurred de novo, was diagnosed in an affected individual from Finland t(3;8)(p12:q11).97ROBO1 was interrupted through the translocation breakpoint, localised in linkage region 3p12 (DYX5). A rare ROBO1 haplotype was identified in the Finnish family, in which the original linkage finding for DYX5 had been found, and cosegregation of this haplotype with dyslexia was reported. Lymphocyte investigation of four affected family members showed that expression of the risk haploytpe was reduced.97 Investigation of the orthologous gene in Drosophila (robo) and mice (Robo1) suggests that ROBO1 functions as a neuronal axon guidance gene involved in brain development.98,99,100

Whether or not ROBO1 actually contributes to the development of dyslexia is currently not clear. A critical point is that the connection between the translocation and dyslexia in the original translocation patient was not imperative: A sibling of the translocation carrier also had dyslexia without carrying the translocation. Should the dyslexia of the Finnish multiplex family be based on a rare and highly penetrant mutation, the causal variant will not be easy to identify, given its size (990 kb of genomic DNA) and the difficulties involved in separating the effects of individual variants from the background variation characterising the haplotype.

Conclusions

Of the candidate genes discussed to date, the evidence for DCDC2 and KIAA0319 is the most convincing. Their identification represents an important step in our understanding of the molecular processes that lead to dyslexia. However, many outstanding questions will need to be addressed by future studies. It is necessary to clarify whether population‐specific genetic heterogeneity and/or phenotypic differences between samples have led to differing findings for the respective loci. Identifying which of the genetic changes in these candidate genes are causal is also important. The lack of associated variants in the coding regions suggests that it is variants influencing generegulation and expression which are responsible.

The nature of the genes identified to date suggests that a disturbance in cortical neurone migration and reduced activity in left‐hemispheric brain regions are pathophysiological correlates of dyslexia. With DCDC2, as with KIAA0319, inhibition leads to poorer neuronal migration in the neocortex of fetal rats through specific small interfering RNAs.74,94 This concept of disturbed neuronal migration is also supported by the few results available from postmortem brain studies of affected individuals, which report cortical malformations in the region of the perisylvian cortex.101,102,103 The concept of disturbed neuronal migration in dyslexia is intriguing and will stimulate further research in this area. In view of the fact that DCDC2 and K1AA0319 only contribute a limited part to the development of dyslexia and that most susceptibility genes are still unknown, it may be possible in the future to identify completely new pathophysiological mechanisms.

To date, no specific cognitive processes are known to be influenced by the proposed susceptibility genes. Some studies have already started to include neurophysiological (eg, event‐related potential) and imaging (eg, functional MRI) procedures in their phenotype characterisation of patients. Such samples are an important prerequisite for the identification of those processes that are most proximal to the effects of particular genes and their associated biological pathways.

Through the availability of detailed clinical data, it should be possible to associate special phenotype dimensions of dyslexia with specific risk genes (genotype–phenotype association). Phenotype subdimensions are, of course, correlated with each other, and the effects will not affect isolated subdimensions. Nor is it to be expected that specific genes will affect the whole spectrum of phenotype dimensions equally. Studies have not yet managed to establish genotype–phenotype relations convincingly, although samples may have been too small to demonstrate these effects. However, proof of genotype–phenotype associations could be facilitated through the joint analysis of larger samples and the identification of causative variants.

The molecular genetic studies conducted so far have not considered sex‐specific genetic effects. Differing prevalence rates between males and females could be suggestive of a sex‐specific geneeffect. A satisfactory power to detect such effects can be provided only when sex is taken into account during the analysis of results, 66 and this should be a feature of future studies.

Identification of susceptibility genes will allow research into the molecular background of clinically observed comorbidity. Eight loci have already been proposed as having pleiotropic effects on dyslexia and ADHD at a linkage level.46,47,48,49,50,51,52,53,54,70,76,81,82 The identification of susceptibility genes also allows examination of the extent to which dyslexia‐associated disorders, such as SSD and language impairment, are influenced by the same susceptibility genes. For SSD, overlapping linkage evidence in DYX5 already provides the first concrete evidence of such common gene effects.60,78

The identification of susceptibility genes will enable the analysis of gene–gene interactions, through which epistatic effects can be discovered. A first example of this might be the proposed interaction between DCDC2 and KIAA0319.90 A further aim of future research will be to establish a better understanding of gene–environment interactions in order to identify relevant exogenous risk factors. It has long been recognised that environmental factors are of great relevance to the development of dyslexia, but only some of these factors have been identified so far.104 If such factors can be modulated, future dyslexia prevention and individual genetic risk profiling could be envisaged.

The genes that accompany the development of dyslexia are naturally of great interest from an evolutionary perspective.105 Through the identification of the gene at the DNA level, comparison with species that are closely related to us but that do not have the same speech capacity could be carried out, as well as examination of sequence variability between humans. Speech‐associated genes may have been under a selection pressure, which proved advantageous for the development of modern man.

As is generally the case with research on complex genetic disorders, it can be assumed that the speed by which susceptibility genes are identified will be increased through increasing knowledge and huge technological advances (eg, genomewide association studies). Future research efforts will be of a collaborative nature, drawing on complementary expertise from various scientific disciplines and involving the combining of large samples, an approach exemplified by the large multidisciplinary European research consortium (www.neurodys.com) which integrates the work of research groups from nine countries.

Abbreviations

ADHD - attention‐deficit hyperactivity disorder

CNS - central nervous system

DCDC2 - doublecortin domain containing protein 2

DYX1 - dyslexia susceptibility 1

DYX9 - dyslexia susceptibility 9

DYX1C1 - dyslexia susceptibility 1 candidate 1

LD - linkage disequilibrium

QTL - quantitative trait loci

ROBO1 - roundabout Drosophila hom*olog of 1

SSD - speech–sound disorder

Footnotes

Funding: This work was supported by the Deutsche Forschungsgemeinschaft.

Competing interests: None.

References

1. Katusic S K, Colligan R C, Barbaresi W J.et al Incidence of reading disability in a population‐based birth cohort, 1976–1982, Rochester, Minn. Mayo Clin Proc 2001761081–1092. [PubMed] [Google Scholar]

2. Shaywitz S E, Shaywitz B A, Fletcher J M.et al Prevalence of reading‐disability in boys and girls—results of the Connecticut longitudinal‐study. JAMA 1990264998–1002. [PubMed] [Google Scholar]

3. World Health Organization The ICD‐10 classification of mental and behavioural disorders: diagnostic criteria for research. Geneva: World Health Organization, 1993

4. Shaywitz S E, Fletcher J M, Holahan J M.et al Persistence of dyslexia: the Connecticut Longitudinal Study at adolescence. Pediatrics 19991041351–1359. [PubMed] [Google Scholar]

5. Bruck M. Outcomes of adults with childhood histories of dyslexia. In: Hulme C, Joshi RM, eds. Reading and spelling: development and disorders Mahwah, NJ: L Erlbaum, 1998179–200.

6. Maughan B. Annotation: long‐term outcomes of developmental reading problems. J Child Psychol Psychiatry 199536357–371. [PubMed] [Google Scholar]

7. Strehlow U, Kluge R, Möller H.et al Long‐term course of dyslexia beyond the school years: catamnesis from pediatric psychiatric ambulatory care. Z Kinder Jugendpsychiatr 199220254–265. [PubMed] [Google Scholar]

8. August G J, Garfinkel B D. Comorbidity of ADHD and reading disability among clinic‐referred children. J Abnorm Child Psychol 19901829–45. [PubMed] [Google Scholar]

9. Kaplan B J, Dewey D M, Crawford S G.et al The term comorbidity is of questionable value in reference to developmental disorders: data and theory. J Learn Disabil 200134555–565. [PubMed] [Google Scholar]

10. Purvis K L, Tannock R. Language abilities in children with attention deficit hyperactivity disorder, reading disabilities, and normal controls. J Abnorm Child Psychol 199725133–144. [PubMed] [Google Scholar]

11. Shaywitz S E. Dyslexia. N Engl J Med 1998338307–312. [PubMed] [Google Scholar]

12. Willcutt E G, Pennington B F, DeFries J C. Twin study of the etiology of comorbidity between reading disability and attention‐deficit/hyperactivity disorder. Am J Med Genet 200096293–301. [PubMed] [Google Scholar]

13. Frauenheim J G, Heckerl J R. A longitudinal study of psychological and achievement test performance in severe dyslexic adults. J Learn Disabil 198316339–347. [PubMed] [Google Scholar]

14. Naylor C F, Felton R H, Wood F B. Adult outcome in developmental dyslexia. In: Pavlidis G, ed. Perspectives on dyslexia: cognition, language and treatment Vol 2. Chichester, England: John Wiley & Sons, 199029

15. Schulte‐Körne G, Deimel W, Remschmidt H. Diagnosis of reading and spelling disorder. Z Kinder Jugendpsychiatr Psychother 200129113–116. [PubMed] [Google Scholar]

16. Rutter M, Caspi A, Fergusson D.et al Sex differences in developmental reading disability—new findings from 4 epidemiological studies. JAMA 20042912007–2012. [PubMed] [Google Scholar]

17. Olson R K. Dyslexia: nature and nurture. Dyslexia 20028143–159. [PubMed] [Google Scholar]

18. Hinshelwood J. Word—blindness and visual memory. Lancet 18951461564–1570. [Google Scholar]

19. Stephenson S. Six cases of congenital word‐blindness affecting three generations of one family. Ophthalmoscope 19075482–484. [Google Scholar]

20. Thomas C J. Congenital word blindness and its treatment. Ophthalmoscope 19053380–385. [Google Scholar]

21. Hallgren B. Specific dyslexia (congenital word‐blindness); a clinical and genetic study. Acta Neurol Scand (Suppl) 1950651–287. [PubMed] [Google Scholar]

22. Olson R K, Forsberg H, Wise B. Genes, environment, and development of orthographic skills. In: Berninger VW, ed. The varieties of orthographic knowledge I: theoretical and developmental issues Dordrecht, Netherlands: Kluwer Academic Publishers, 199427–71.

23. Schulte‐Körne G, Deimel W, Müller K.et al Familial aggregation of spelling disability. J Child Psychol Psychiatry 199637817–822. [PubMed] [Google Scholar]

24. Stevenson J. Which aspects of processing text mediate genetic‐effects. Read Writ 19913249–269. [Google Scholar]

25. Ziegler A, König I R, Deimel W.et al Developmental dyslexia—recurrence risk estimates from a German bi‐center study using the single proband sib pair design. Hum Hered 200559136–143. [PubMed] [Google Scholar]

26. Plomin R, Kovas Y. Generalist genes and learning disabilities. Psychol Bull 2005131592–617. [PubMed] [Google Scholar]

27. Gayán J, Olson R K. Genetic and environmental influences on orthographic and phonological skills in children with reading disabilities. Dev Neuropsychol 200120483–507. [PubMed] [Google Scholar]

28. Hawke J L, Wadsworth S J, DeFries J C. Genetic influences on reading difficulties in boys and girls: the Colorado twin study. Dyslexia 20061221–29. [PubMed] [Google Scholar]

29. Wadsworth S J, Knopik V S, DeFries J C. Reading disability in boys and girls: no evidence for a differential genetic etiology. Read Writ 200013133–145. [Google Scholar]

30. Harlaar N, Spinath F M, Dale P S.et al Genetic influences on early word recognition abilities and disabilities: a study of 7‐year‐old twins. J Child Psychol Psychiatry 200546373–384. [PubMed] [Google Scholar]

31. Goswami U, Bryant P.Phonological skills and learning to read. Hillsdale, NJ: L Erlbaum, 1990

32. Ramus F, Rosen S, Dakin S C.et al Theories of developmental dyslexia: insights from a multiple case study of dyslexic adults. Brain 2003126841–865. [PubMed] [Google Scholar]

33. Tallal P. Auditory temporal perception, phonics, and reading disabilities in children. Brain Lang 19809182–198. [PubMed] [Google Scholar]

34. Ahissar M, Protopapas A, Reid M.et al Auditory processing parallels reading abilities in adults. Proc Natl Acad Sci USA 2000976832–6837. [PMC free article] [PubMed] [Google Scholar]

35. McAnally K I, Stein J F. Auditory temporal coding in dyslexia. Proc Biol Sci 1996263961–965. [PubMed] [Google Scholar]

36. Nagarajan S, Mahncke H, Salz T.et al Cortical auditory signal processing in poor readers. Proc Natl Acad Sci USA 1999966483–6488. [PMC free article] [PubMed] [Google Scholar]

37. Kujala T, Myllyviita K, Tervaniemi M.et al Basic auditory dysfunction in dyslexia as demonstrated by brain activity measurements. Psychophysiology 200037262–266. [PubMed] [Google Scholar]

38. Schulte‐Körne G, Deimel W, Bartling J.et al Pre‐attentive processing of auditory patterns in dyslexic human subjects. Neurosci Lett 199927641–44. [PubMed] [Google Scholar]

39. Lovegrove W J, Bowling A, Badco*ck D.et al Specific reading disability: differences in contrast sensitivity as a function of spatial frequency. Science 1980210439–440. [PubMed] [Google Scholar]

40. Schulte‐Körne G, Bartling J, Deimel W.et al Motion‐onset VEPs in dyslexia. Evidence for visual perceptual deficit. Neuroreport 2004151075–1078. [PubMed] [Google Scholar]

41. Stein J, Walsh V. To see but not to read; the magnocellular theory of dyslexia. Trends Neurosci 199720147–152. [PubMed] [Google Scholar]

42. Talcott J B, Witton C, McLean M F.et al Dynamic sensory sensitivity and children's word decoding skills. Proc Natl Acad Sci USA 2000972952–2957. [PMC free article] [PubMed] [Google Scholar]

43. Nicolson R I, Fawcett A J, Dean P. Developmental dyslexia: the cerebellar deficit hypothesis. Trends Neurosci 200124508–511. [PubMed] [Google Scholar]

44. Wolf M, Bowers P G. The double‐deficit hypothesis for the developmental dyslexias. J Educ Psychol 199991415–438. [Google Scholar]

45. Schulte‐Körne G, Zucchelli M, Deimel W.et al Interrelationship and familiality of dyslexia related quantitative measures. Ann Hum Genet 2006701–16. [PubMed] [Google Scholar]

46. Smith S D, Kimberling W J, Pennington B F.et al Specific reading disability: identification of an inherited form through linkage analysis. Science 19832191345–1347. [PubMed] [Google Scholar]

47. Grigorenko E L, Chang J T. An extension of affected‐pedigree‐member analyses to triads of relatives. Genet Epidemiol 1997141005–1010. [PubMed] [Google Scholar]

48. Schulte‐Körne G, Grimm T, Nöthen M M.et al Evidence for linkage of spelling disability to chromosome 15. Am J Hum Genet 199863279–282. [PMC free article] [PubMed] [Google Scholar]

49. Chapman N H, Igo R P, Thomson J B.et al Linkage analyses of four regions previously implicated in dyslexia: confirmation of a locus on chromosome 15q. Am J Med Genet B Neuropsychiatr Genet 200413167–75. [PubMed] [Google Scholar]

50. Cardon L R, Smith S D, Fulker D W.et al Quantitative trait locus for reading disability on chromosome 6. Science 1994266276–279. [PubMed] [Google Scholar]

51. Fisher S E, Marlow A J, Lamb J.et al A quantitative‐trait locus on chromosome 6p influences different aspects of developmental dyslexia. Am J Hum Genet 199964146–156. [PMC free article] [PubMed] [Google Scholar]

52. Gayán J, Smith S D, Cherny S S.et al Quantitative‐trait locus for specific language and reading deficits on chromosome 6p. Am J Hum Genet 199964157–164. [PMC free article] [PubMed] [Google Scholar]

53. Fisher S E, Francks C, Marlow A J.et al Independent genome‐wide scans identify a chromosome 18 quantitative‐trait locus influencing dyslexia. Nat Genet 20023086–91. [PubMed] [Google Scholar]

54. Kaplan D E, Gayan J, Ahn J.et al Evidence for linkage and association with reading disability on 6p21.3‐22. Am J Hum Genet 2002701287–1298. [PMC free article] [PubMed] [Google Scholar]

55. Grigorenko E L, Wood F B, Golovyan L.et al Continuing the search for dyslexia genes on 6p. Am J Med Genet B Neuropsychiatr Genet 200311889–98. [PubMed] [Google Scholar]

56. fa*gerheim T, Raeymaekers P, Tonnessen F E.et al A new gene (DYX3) for dyslexia is located on chromosome 2. J Med Genet 199936664–669. [PMC free article] [PubMed] [Google Scholar]

57. Petryshen T L, Kaplan B J, Hughes M L.et al Supportive evidence for the DYX3 dyslexia susceptibility gene in Canadian families. J Med Genet 200239125–126. [PMC free article] [PubMed] [Google Scholar]

58. Kaminen N, Hannula‐Jouppi K, Kestilä M.et al A genome scan for developmental dyslexia confirms linkage to chromosome 2p11 and suggests a new locus on 7q32. J Med Genet 200340340–345. [PMC free article] [PubMed] [Google Scholar]

59. Petryshen T L, Kaplan B J, Fu Liu M.et al Evidence for a susceptibility locus on chromosome 6q influencing phonological coding dyslexia. Am J Med Genet 2001105507–517. [PubMed] [Google Scholar]

60. Nopola‐Hemmi J, Myllyluoma B, Haltia T.et al A dominant gene for developmental dyslexia on chromosome 3. J Med Genet 200138658–664. [PMC free article] [PubMed] [Google Scholar]

61. Hsiung G Y, Kaplan B J, Petryshen T L.et al A dyslexia susceptibility locus (DYX7) linked to dopamine D4 receptor (DRD4) region on chromosome 11p15.5. Am J Med Genet B Neuropsychiatr Genet 2004125112–119. [PubMed] [Google Scholar]

62. Rabin M, Wen X L, Hepburn M.et al Suggestive linkage of developmental dyslexia to chromosome 1p34‐p36. Lancet 1993342178 [PubMed] [Google Scholar]

63. Grigorenko E L, Wood F B, Meyer M S.et al Linkage studies suggest a possible locus for developmental dyslexia on chromosome 1p. Am J Med Genet 2001105120–129. [PubMed] [Google Scholar]

64. Tzenova J, Kaplan B J, Petryshen T L.et al Confirmation of a dyslexia susceptibility locus on chromosome 1p34‐p36 in a set of 100 Canadian families. Am J Med Genet B Neuropsychiatr Genet 2004127117–124. [PubMed] [Google Scholar]

65. de Kovel C G, Hol F A, Heister J G.et al Genomewide scan identifies susceptibility locus for dyslexia on Xq27 in an extended Dutch family. J Med Genet 200441652–657. [PMC free article] [PubMed] [Google Scholar]

66. Igo R P, Chapman N H, Berninger V W.et al Genomewide scan for real‐word reading subphenotypes of dyslexia: novel chromosome 13 locus and genetic complexity. Am J Med Genet B Neuropsychiatr Genet 200614115–27. [PMC free article] [PubMed] [Google Scholar]

67. Raskind W H, Igo R P, Chapman N H.et al A genome scan in multigenerational families with dyslexia: identification of a novel locus on chromosome 2q that contributes to phonological decoding efficiency. Mol Psychiatry 200510699–711. [PubMed] [Google Scholar]

68. Marino C, Giorda R, Vanzin L.et al A locus on 15q15‐15qter influences dyslexia: further support from a transmission/disequilibrium study in an Italian speaking population. J Med Genet 20044142–46. [PMC free article] [PubMed] [Google Scholar]

69. Morris D W, Robinson L, Turic D.et al Family‐based association mapping provides evidence for a gene for reading disability on chromosome 15q. Hum Mol Genet 20009843–848. [PubMed] [Google Scholar]

70. Bakker S C, van der Meulen E M, Buitelaar J K.et al A whole‐genome scan in 164 Dutch sib pairs with attention‐deficit/hyperactivity disorder: suggestive evidence for linkage on chromosomes 7p and 15q. Am J Hum Genet 2003721251–1260. [PMC free article] [PubMed] [Google Scholar]

71. Cope N, Harold D, Hill G.et al Strong evidence that KIAA0319 on chromosome 6p is a susceptibility gene for developmental dyslexia. Am J Hum Genet 200576581–591. [PMC free article] [PubMed] [Google Scholar]

72. Deffenbacher K E, Kenyon J B, Hoover D M.et al Refinement of the 6p21. 3 quantitative trait locus influencing dyslexia: linkage and association analyses, Hum Genet 2004115128–138. [PubMed] [Google Scholar]

73. Francks C, Paracchini S, Smith S D.et al A 77‐kilobase region of chromosome 6p22.2 is associated with dyslexia in families from the United Kingdom and from the United States. Am J Hum Genet 2004751046–1058. [PMC free article] [PubMed] [Google Scholar]

74. Meng H, Smith S D, Hager K.et alDCDC2 is associated with reading disability and modulates neuronal development in the brain. Proc Natl Acad Sci USA 200510217053–17058. [PMC free article] [PubMed] [Google Scholar]

75. Schumacher J, Anthoni H, Dahdouh F.et al Strong genetic evidence of DCDC2 as a susceptibility gene for dyslexia. Am J Hum Genet 20067852–62. [PMC free article] [PubMed] [Google Scholar]

76. Willcutt E G, Pennington B F, Smith S D.et al Quantitative trait locus for reading disability on chromosome 6p is pleiotropic for attention‐deficit/hyperactivity disorder. Am J Med Genet 2002114260–268. [PubMed] [Google Scholar]

77. Fisher S E, DeFries J C. Developmental dyslexia: genetic dissection of a complex cognitive trait. Nat Rev Neurosci 20023767–780. [PubMed] [Google Scholar]

78. Stein C M, Schick J H, Gerry Taylor H.et al Pleiotropic effects of a chromosome 3 locus on speech‐sound disorder and reading. Am J Hum Genet 200474283–297. [PMC free article] [PubMed] [Google Scholar]

79. Marlow A J, Fisher S E, Francks C.et al Use of multivariate linkage analysis for dissection of a complex cognitive trait. Am J Hum Genet 200372561–570. [PMC free article] [PubMed] [Google Scholar]

80. Faraone S V, Doyle A E, Mick E.et al Meta‐analysis of the association between the 7‐repeat allele of the dopamine D(4) receptor gene and attention deficit hyperactivity disorder. Am J Psychiatry 20011581052–1057. [PubMed] [Google Scholar]

81. Gayán J, Willcutt E G, Fisher S E.et al Bivariate linkage scan for reading disability and attention‐deficit/hyperactivity disorder localizes pleiotropic loci. J Child Psychol Psychiatry 2005461045–1056. [PubMed] [Google Scholar]

82. Loo S K, Fisher S E, Francks C.et al Genome‐wide scan of reading ability in affected sibling pairs with attention‐deficit/hyperactivity disorder: unique and shared genetic effects. Mol Psychiatry 20049485–493. [PubMed] [Google Scholar]

83. Taipale M, Kaminen N, Nopola‐Hemmi J.et al A candidate gene for developmental dyslexia encodes a nuclear tetratricopeptide repeat domain protein dynamically regulated in brain. Proc Natl Acad Sci USA 200310011553–11558. [PMC free article] [PubMed] [Google Scholar]

84. Wigg K G, Couto J M, Feng Y.et al Support for EKN1 as the susceptibility locus for dyslexia on 15q21. Mol Psychiatry 200491111–1121. [PubMed] [Google Scholar]

85. Scerri T S, Fisher S E, Francks C.et al Putative functional alleles of DYX1C1 are not associated with dyslexia susceptibility in a large sample of sibling pairs from the UK. J Med Genet 200441853–857. [PMC free article] [PubMed] [Google Scholar]

86. Meng H, Hager K, Held M.et al TDT‐association analysis of EKN1 and dyslexia in a Colorado twin cohort. Hum Genet 200511887–90. [PubMed] [Google Scholar]

87. Marino C, Giorda R, Lorusso M L.et al A family‐based association study does not support DYX1C1 on 15q21.3 as a candidate gene in developmental dyslexia. Eur J Hum Genet 200513491–499. [PubMed] [Google Scholar]

88. Bellini G, Bravaccio C, Calamoneri F.et al No evidence for association between dyslexia and DYX1C1 functional variants in a group of children and adolescents from Southern Italy. J Mol Neurosci 200527311–314. [PubMed] [Google Scholar]

89. Cope N A, Hill G, van den Bree M.et al No support for association between dyslexia susceptibility 1 candidate 1 and developmental dyslexia. Mol Psychiatry 200510237–238. [PubMed] [Google Scholar]

90. Harold D, Paracchini S, Scerri T.et al Further evidence that the KIAA0319 gene confers susceptibility to developmental dyslexia. Mol Psychiatry 2006111085–1091. [PubMed] [Google Scholar]

91. Horwitz B, Rumsey J M, Donohue B C. Functional connectivity of the angular gyrus in normal reading and dyslexia. Proc Natl Acad Sci USA 1998958939–8944. [PMC free article] [PubMed] [Google Scholar]

92. Shaywitz S E, Shaywitz B A, Pugh K R.et al Functional disruption in the organization of the brain for reading in dyslexia. Proc Natl Acad Sci USA 1998952636–2641. [PMC free article] [PubMed] [Google Scholar]

93. Silani G, Frith U, Demonet J F.et al Brain abnormalities underlying altered activation in dyslexia: a voxel based morphometry study. Brain 20051282453–2461. [PubMed] [Google Scholar]

94. Paracchini S, Thomas A, Castro S.et al The chromosome 6p22 haplotype associated with dyslexia reduces the expression of KIAA0319, a novel gene involved in neuronal migration. Hum Mol Genet 2006151659–1666. [PubMed] [Google Scholar]

95. Nopola‐Hemmi J, Taipale M, Haltia T.et al Two translocations of chromosome 15q associated with dyslexia. J Med Genet 200037771–775. [PMC free article] [PubMed] [Google Scholar]

96. Wang Y, Paramasivam M, Thomas A.et al DYX1C1 functions in neuronal migration in developing cortex. Neuroscience 2006143515–522. [PubMed] [Google Scholar]

97. Hannula‐Jouppi K, Kaminen‐Ahola N, Taipale M.et al The axon guidance receptor gene ROBO1 is a candidate gene for developmental dyslexia. PLoS Genet 20051e50 [PMC free article] [PubMed] [Google Scholar]

98. Kidd T, Bland K S, Goodman C S. Slit is the midline repellent for the robo receptor in Drosophila. Cell 199996785–794. [PubMed] [Google Scholar]

99. Seeger M, Tear G, Ferres‐Marco D.et al Mutations affecting growth cone guidance in Drosophila: genes necessary for guidance toward or away from the midline. Neuron 199310409–426. [PubMed] [Google Scholar]

100. Andrews W, Liapi A, Plachez C.et al Robo1 regulates the development of major axon tracts and interneuron migration in the forebrain. Development 20061332243–2252. [PubMed] [Google Scholar]

101. Galaburda A M, Kemper T L. Cytoarchitectonic abnormalities in developmental dyslexia: a case study. Ann Neurol 1979694–100. [PubMed] [Google Scholar]

102. Galaburda A M, Sherman G F, Rosen G D.et al Developmental dyslexia: four consecutive patients with cortical anomalies. Ann Neurol 198518222–233. [PubMed] [Google Scholar]

103. Galaburda A M. Developmental dyslexia and animal studies: at the interface between cognition and neurology. Cognition 199450133–149. [PubMed] [Google Scholar]

104. Kremen W S, Jacobson K C, Xian H.et al Heritability of word recognition in middle‐aged men varies as a function of parental education. Behav Genet 200535417–433. [PubMed] [Google Scholar]

105. Fisher S E, Marcus G F. The eloquent ape: genes, brains and the evolution of language. Nat Rev Genet 200679–20. [PubMed] [Google Scholar]

Articles from Journal of Medical Genetics are provided here courtesy of BMJ Publishing Group

Genetics of dyslexia: the evolving landscape (2024)
Top Articles
Latest Posts
Article information

Author: Kareem Mueller DO

Last Updated:

Views: 6054

Rating: 4.6 / 5 (46 voted)

Reviews: 85% of readers found this page helpful

Author information

Name: Kareem Mueller DO

Birthday: 1997-01-04

Address: Apt. 156 12935 Runolfsdottir Mission, Greenfort, MN 74384-6749

Phone: +16704982844747

Job: Corporate Administration Planner

Hobby: Mountain biking, Jewelry making, Stone skipping, Lacemaking, Knife making, Scrapbooking, Letterboxing

Introduction: My name is Kareem Mueller DO, I am a vivacious, super, thoughtful, excited, handsome, beautiful, combative person who loves writing and wants to share my knowledge and understanding with you.