ArticlePDF Available

AC and AG dinucleotide repeats in the PAX6 P1 promoter are associated with high myopia

Authors:

Abstract and Figures

The PAX6 gene, located at the reported myopia locus MYP7 on chromosome 11p13, was postulated to be associated with myopia development. This study investigated the association of PAX6 with high myopia in 379 high myopia patients and 349 controls. High myopia patients had refractive errors of -6.00 diopters or greater and axial length longer than 26 mm. Control subjects had refractive errors less than -1.00 diopter and axial length shorter than 24 mm. The P1 promoter, all coding sequences, and adjacent splice-site regions of the PAX6 gene were screened in all study subjects by polymerase chain reaction and direct sequencing. PAX6 P1 promoter-luciferase constructs with variable AC and AG repeat lengths were prepared and transfected into human ARPE-19 cells prior to assaying for their transcriptional activities. No sequence alterations in the coding or splicing regions showed an association with high myopia. Two dinucleotide repeats, (AC)(m) and (AG)(n), in the P1 promoter region were found to be highly polymorphic and significantly associated with high myopia. Higher repeat numbers were observed in high myopia patients for both (AC)(m) (empirical p = 0.013) and (AG)(n) (empirical p = 0.012) dinucleotide polymorphisms, with a 1.327-fold increased risk associated with the (AG)(n) repeat (empirical p = 0.016; 95% confidence interval: 1.059-1.663). Luciferase-reporter analysis showed elevated transcription activity with increasing individual (AC)(m) and (AG)(n) and combined (AC)(m)(AG)(n) repeat lengths. Our results revealed an association between high myopia and AC and AG dinucleotide repeat lengths in the PAX6 P1 promoter, indicating the involvement of PAX6 in the pathogenesis of high myopia.
Content may be subject to copyright.
AC and AG dinucleotide repeats in the PAX6 P1 promoter are
associated with high myopia
Tsz Kin Ng,1 Ching Yan Lam,1 Dennis Shun Chiu Lam,1 Sylvia Wai Yee Chiang,1 Pancy Oi Sin Tam,1
Dan Yi Wang,1,2 Bao Jian Fan,1,2 Gary Hin-Fai Yam,1 Dorothy Shu Ping Fan,1 Chi Pui Pang1
(The first two authors contributed equally to this work.)
1Department of Ophthalmology & Visual Sciences, The Chinese University of Hong Kong, Hong Kong S.A.R.; 2Present affiliation:
Department of Ophthalmology, Harvard Medical School, Massachusetts Eye and Ear Infirmary, Boston, MA
Purpose: The PAX6 gene, located at the reported myopia locus MYP7 on chromosome 11p13, was postulated to be
associated with myopia development. This study investigated the association of PAX6 with high myopia in 379 high
myopia patients and 349 controls.
Methods: High myopia patients had refractive errors of –6.00 diopters or greater and axial length longer than 26 mm.
Control subjects had refractive errors less than –1.00 diopter and axial length shorter than 24 mm. The P1 promoter, all
coding sequences, and adjacent splice-site regions of the PAX6 gene were screened in all study subjects by polymerase
chain reaction and direct sequencing. PAX6 P1 promoter-luciferase constructs with variable AC and AG repeat lengths
were prepared and transfected into human ARPE-19 cells prior to assaying for their transcriptional activities.
Results: No sequence alterations in the coding or splicing regions showed an association with high myopia. Two
dinucleotide repeats, (AC)m and (AG)n, in the P1 promoter region were found to be highly polymorphic and significantly
associated with high myopia. Higher repeat numbers were observed in high myopia patients for both (AC)m (empirical
p = 0.013) and (AG)n (empirical p = 0.012) dinucleotide polymorphisms, with a 1.327-fold increased risk associated with
the (AG)n repeat (empirical p = 0.016; 95% confidence interval: 1.059–1.663). Luciferase-reporter analysis showed
elevated transcription activity with increasing individual (AC)m and (AG)n and combined (AC)m(AG)n repeat lengths.
Conclusions: Our results revealed an association between high myopia and AC and AG dinucleotide repeat lengths in
the PAX6 P1 promoter, indicating the involvement of PAX6 in the pathogenesis of high myopia.
Myopia, one of the most common refractive errors of the
eye worldwide, is an important public health issue, especially
in Asia, because of its higher prevalence in Asians than in
other populations [1]. The progression of myopia in Chinese
children in Hong Kong and Singapore is also much higher than
in Caucasians [2,3]. In Hong Kong, the prevalence of myopia
in Chinese schoolchildren aged 11–16 was 36.7%, according
to a 2004 report, which is several times higher than among
Caucasian children of similar ages [4]. The prevalence of high
myopia, defined as a refractive error equal to or greater than
–6.00 diopters (D), is also higher in Chinese than in
Caucasians [5,6]. Individuals with high myopia are more
prone to develop serious ocular complications, such as retinal
detachment, glaucoma, premature cataracts, and macular
degeneration, which may lead to visual impairment or even
blindness [7-10].
Myopia is a complex disorder. Multiple interacting
environmental and genetic causes are implicated. Myopia
Correspondence to: Prof. C.P. Pang, Department of Ophthalmology
& Visual Sciences, The Chinese University of Hong Kong,
University Eye Center, Hong Kong Eye Hospital, 147K Argyle
Street, Kowloon, Hong Kong; Phone: +852 27623129; FAX: +852
27159490; email: cppang@cuhk.edu.hk
development in schoolchildren has been attributed to
environmental factors, such as near work, reading habits, and
school achievement [3,11,12]. In addition, high heritability of
refractive errors has been observed in dizygotic and
monozygotic twin studies [13-17]. Family and sibling studies
have shown that children of myopic parents have greater
chances of developing myopia than those with nonmyopic
parents [11,18]. Twenty-four chromosomal loci have been
identified for myopia: Xq28 (MYP1) [19], 18p11.31 (MYP2)
[20,21], 12q21-31 (MYP3) [22], 7q36 (MYP4) [23], 17q21-22
(MYP5) [24], 22q37.1 (MYP6) [25], 11p13 (MYP7) [26], 3q26
(MYP8) [26], 4q12 (MYP9) [26], 8p23 (MYP10) [26], 4q22-
q27 (MYP11) [27], 2q37.1 (MYP12) [28], Xq23 (MYP13)
[29], 1p36 (MYP14) [30], 10q21.2 (MYP15) [31], 5p15.33-
p15.2 (MYP16) [32], 7p15 (MYP17) [33,34], 14q22.1-q24.2
(MYP18) [35], 15q12-13 [36], 21q22.3 [37], 12q24 [38], 4q21
[38], 9q34.11 [39] and 2q37 [40]. Among them, MYP1–5,11–
13,16, and 18 are linked to high myopia, and MYP2,11,13,
16, and 18 are found in the Chinese population. Some
candidate genes have been postulated for myopia, such as
TGIF [41], HGF [42], MMP3 [43], MMP9 [43], COL1A1
[44], COL2A1 [45], TGFB1 [46], TGFB2 [47], LUM [48], and
CMET [49].
Molecular Vision 2009; 15:2239-2248 <http://www.molvis.org/molvis/v15/a241>
Received 10 August 2009 | Accepted 27 October 2009 | Published 5 November 2009
© 2009 Molecular Vision
2239
A genome-wide scan in dizygotic twins revealed a
susceptibility locus for myopia on chromosome 11p13 [26].
The PAX6 gene at this locus, a member of the paired-domain
PAX family, has been postulated as a candidate gene for
myopia. PAX6 is expressed in the human eye [50] and plays
an evolutionarily conserved role in ocular development
[51-53]. PAX6 mutations are associated with ocular disorders,
such as aniridia (OMIM 106210), cataracts (OMIM 604219),
Peters anomaly (OMIM 604229), and optic nerve hypoplasia
(OMIM 16550). PAX6 encodes a transcriptional regulator
containing the DNA-binding paired domain, paired-type
homeodomain, and COOH-terminal transactivation domain.
The Pax6 protein regulates cell adhesion molecules, cell-to-
cell signaling molecules, hormones, and structural proteins
[54] through interactions with transcription factors such as
Mitf [55] and Sox2 [56]. Transcription of PAX6 is regulated
by at least two promoters, P0 and P1 [57-60]. Within the P1
promoter (promoter B in Okladnova et al. [59]), two
dinucleotide repeats, (AC)m and (AG)n, are located about 1 kb
from the transcription start site [58] and are highly
polymorphic in Caucasians. The poly AC and poly AG repeats
are independently polymorphic [60]. Luciferase analysis in
Cos-7 cells has shown that the longer the combined length of
the AC and AG repeats, the higher the transcriptional activity,
implying that the length of this dinucleotide repeat might
influence the transcriptional activity of promoter B, or P1, and
subsequently the transcription of PAX6.
Pax6 levels are tightly controlled. Both overexpression
and haploinsufficiency lead to abnormal phenotypes [61-63].
Polymorphisms or mutations in the PAX6 promoter could
influence PAX6 expressions that ultimately lead to a disease
phenotype. However, although PAX6 has been postulated to
be a candidate gene for myopia, several studies in Caucasian
populations could not find an association between PAX6 and
myopia [26,45,64]. Still, an Australian study suggested
PAX6 mutations might be associated with high myopia [65].
Intronic sequence alterations (SNPs) in PAX6 have been
reported to associate with high myopia in Han Chinese nuclear
families [66] and with extreme myopia in a Taiwan Chinese
population [67], but not in Caucasians. To attest the
association between PAX6 and high myopia, we should look
for mutations that may affect PAX6 expressions. We therefore
screened for sequence alterations in the P1 promoter, coding
exons, and adjacent splice-site regions of PAX6 in unrelated
high myopia patients and control subjects. We also examined
transcriptional effects of dinucleotide repeats within the P1
promoter in cultured human APRE-19 cells by a luciferase-
reporter assay and predicted the presence of transcription
factor binding sites within the repeats.
METHODS
Study subjects: We recruited 379 unrelated Han Chinese
patients with high myopia at the Hong Kong Eye Hospital.
They were given complete ophthalmoscopic examinations.
None of them had known diseases predisposing them to
myopia, such as Stickler or Marfan syndromes. Their
refractive errors were equal to or greater than –6.00 D, and
their axial length was longer than 26 mm. We also recruited
349 unrelated Chinese control subjects who visited the
hospital for ophthalmic examinations. They had no eye
diseases except senile cataracts and slight floaters. All of them
had refractive errors of less than –1.00 D and axial length
shorter than 24 mm. The study protocol was approved by the
Ethics Committee for Human Research at the Chinese
University of Hong Kong and was in accordance with the
tenets of the Declaration of Helsinki. Informed consent was
obtained from the study subjects after explanation of the
nature and possible consequences of the study.
Construction of PAX6 P1 promoter-luciferase constructs: A
1,851 bp genomic fragment (from –1278 to +573) containing
the PAX6 P1 promoter was cloned into an empty pGL3-Basic
vector, pGL3 (Promega, Madison, WI) between the SacI and
BglII sites (OriGene Technologies, Rockville, MD).
Constructs with different repeat lengths were generated.
Genomic DNA from the study subjects was amplified by PCR
(forward primer 5'-ACA CAC AGA TGA CCG GTG G-3';
reverse primer 5'-AAG CCT AGG CCG AGA GGA-3'). AgeI
and AvrII digested products were ligated into a linearized
pGL3-Basic vector containing the P1 promoter (pGL3-
Pax6p). A positive control construct was made by cloning a
pCMV5 promoter [68] into the pGL3-Basic vector (pGL3-
pCMV). All constructs were verified by direct sequencing.
Cell culture and transfection: The human retinal pigment
epithelial cell line ARPE-19 (American Type Culture
Collection, Manassas, VA) [69] was cultured in Dulbecco’s
modified Eagle’s medium and F-12 nutrient mixture
Molecular Vision 2009; 15:2239-2248 <http://www.molvis.org/molvis/v15/a241> © 2009 Molecular Vision
2240
PAX6 genotyping: The whole blood specimens (5 ml) from
all the patients and controls were collected in EDTA tube and
stored at -80 °C for fewer than two months. Genomic DNA
was extracted (QIAamp DNA kit; Qiagen, Hiden, Germeny)
according to the supplier’s instructions. All samples were
screened for sequence alterations in the P1 promoter region
flanking –3,433 to –118, coding exons, and intron-exon
boundaries of PAX6 (ENSG00000007372 and
ENST00000241001; Ensembl genome browser) by
polymerase chain reaction (PCR) with primer sets [61]. PCR
was performed in a final volume of 25 μl containing 1X PCR
buffer (Invitrogen™ Life Technology, Carlsbad, CA), 1.5
mM MgCl2, 0.2 mM of dNTP (Roche, Indianapolis, IN), 0.2
mM of each primers, 0.5 U of Platinum® Taq DNA
polymerase (Invitrogen). After the initial denaturation at 95
°C for 2 min, 40 PCR cycles were conducted: 95 °C for 45 s,
57 °C for 45 s and 72 °C for 45 s. The final extension lasted
for 5 min at 72 °C. Direct sequencing was performed using a
BigDye Terminator Cycle Sequencing Reaction Kit (v3.1,
Applied Biosystems, Foster City, CA) on an ABI 3130XL
capillary DNA sequencer (Applied Biosystems).
supplemented with 10% fetal bovine serum (Gibco BRL,
Rockville, MD). Cells were plated in 60 mm tissue culture
dishes at a density of 2–3×105 cells/dish one day before
transfection. At 60–80% confluence, cells were transfected
with 2 μg luciferase constructs in 6 μl FuGene HD (Roche)
transfection reagent per dish. Empty pGL3 and pGL3-pCMV
were used as negative and positive controls, respectively. At
36 h after transfection, cell lysates were extracted using Cell
Culture Lysis Reagent (Promega, Madison, WI) for
immunoblotting.
Immunoblotting: The denatured cell lysates of the transfected
cells were resolved on 10% SDS-polyacrylamide gel and
electro-transferred to nitrocellulose membranes for probing
with a rabbit polyclonal primary antibody against firefly
luciferase (Sigma-Aldrich, St. Louis, MO) and a secondary
antibody against rabbit IgG conjugated with horseradish
peroxidase (Jackson Immuno Res., West Grove, PA). The
chemiluminescence was detected by an enhanced
chemiluminescence system (Amersham Pharmacia,
Cleveland, OH) and quantified by ChemiDoc (BioRad,
Hercules, CA). Normalized luciferase intensities were
calculated by dividing the quantified luciferase intensities by
the housekeeping β-actin intensities. Triplicates were
performed.
Statistical analysis: The χ2 test or Fisher exact test was used
to compare the allele and genotype frequencies of SNPs in
patients and control subjects. For the comparison of (AC)m
and (AG)n repeat alleles and genotypes between high myopia
patients and control subjects, the χ2 test was performed using
the CLUMP program (version 2.3) [70]. For multiple testing
corrections, 10,000 Monte Carlo permutations were chosen to
simulate the empirical significance levels of the statistics
produced by the program, resulting in an empirical p-value.
Due to low frequencies of some alleles, and in order to
determine whether the transcriptional activities were affected
by the thresholds, (AC)m and (AG)n repeats were collapsed
into groups for association study and immunoblotting analysis
[71,72]. The risk of high myopia was also determined by odds
ratio using the χ2 test. Activity of each allelic construct was
expressed relative to (AC)20(AG)6. One-way ANOVA and
independent T-testing were used to compare the means among
(AC)m groups and between (AG)n repeats, respectively. SNP-
trait association, odds ratio calculation, and immunoblotting
analysis were performed on SPSS version 16.0 (SPSS
Science, Chicago, IL). Significance was defined as p < 0.05.
Transcription factor binding site prediction: The DNA
sequence of the cloned PAX6 P1 promoter was used to predict
transcription factor binding sites. The Transcription Element
Search System (TESS: University of Pennsylvania,
Philadelphia, PA) [73,74] was used to predict the transcription
factors that would bind to the region of the dinucleotide
repeats in the PAX6 P1 promoter. Predictions for different
lengths of dinucleotide repeats were also performed. As in the
statistical analysis for immunoblotting, (AC)20(AG)6 was set
Within the PAX6 P1 promoter, two dinucleotide repeats,
(AC)m and (AG)n, were observed about 1 kb from the
transcription start site, both highly polymorphic (Table 1).
The AC repeats ranged from 16 to 26 in high myopia patients
and from 7 to 26 in control subjects, while 5 to 8 AG repeats
were observed in patients and 4 to 8 in controls. The median
numbers of AC and AG repeats were 20 and 6, respectively,
in both patients and controls. Distribution of the allele
frequencies was slightly skewed in patients for both AC and
AG repeats. Allele frequencies of the AC and AG repeats were
significantly different between patients and controls
(empirical p = 0.013 and 0.012, respectively; Table 1).
Because the frequencies of some of the alleles were low, the
AC and AG repeats were collapsed into groups. The grouped
repeat lengths were longer in patients than in controls
(empirical p = 0.016 for (AC)m and empirical p = 0.016 for
(AG)n; Table 2). In terms of risk analysis, individuals with
(AG)7-8 repeats had a 1.327-fold increased risk of developing
high myopia compared with the those with (AG)4-6 repeats
(empirical p = 0.016; 95% confidence interval = 1.059–
1.663). Both grouped AC and grouped AG genotypes were
significantly different between high myopia patients and
control (empirical p = 0.004 and 0.039, respectively; Table 3).
We found that the dinucleotide repeats affected the
transcriptional activity of the PAX6 P1 promoter (Figure 1).
For a given (AG)n repeat length, elevated transcriptional
Molecular Vision 2009; 15:2239-2248 <http://www.molvis.org/molvis/v15/a241> © 2009 Molecular Vision
2241
RESULTS
In our study cohort, high myopia patients had a mean age of
39.52±14.96 years and a male-to-female ratio of 1.2:1.
Refractive errors ranged from –6.00 to –30.00 D. For the
controls, the mean age was 64.85±14.85 years, with a male-
to-female ratio of 1.6:1. There was no significant difference
in the sex ratio between high myopia patients and controls.
Two sequence changes were identified in coding exons
with the intron-exon boundary of PAX6. One novel
heterozygous silent variant, 678A>G (R67R), was found in
one high myopia patient, and a noncoding sequence change,
rs667773, was found in both patients and controls. Allelic and
genotypic frequencies of both polymorphisms showed no
significant difference (p > 0.05) between patients and controls
(data not shown). Within the P1 promoter region, 20
polymorphisms were identified, with no significant difference
in frequencies between patients and controls: –186C>T,
215G>A, –242G>A, –263A>G, –292A>G, –331A>G,
337A>T, –354A>G, –382G>A, –407G>A, –409G>A,
692A>G, –758C>T, –782A>G, –933C>G, –3050C>A,
3070C>A, –3078A>G, –3090C>T, and –3282T>C (data not
shown). For -186C>T, -292A>G, -331A>G, -933C>G, and
-3282T>C, each SNP was only found in 1 high myopia patient.
Therefore, they were statistically not significant under
Pearson’s χ2 test (p > 0.05).
activity was observed with increasing length of (AC)m repeats
(p = 0.004, one-way ANOVA; post-hoc tests adjusted by
Tukey HSD: (AC)Below20–22 versus (AC)20–22, p = 0.033; and
(AC)Below20–22 versus (AC)Above20–22, p = 0.004; Figure 1A,B).
Similarly, at a given (AC)m repeat length, transcriptional
activity of (AG)8 was increased when compared with (AG)6,
although the increase was not significant, likely due to the
substantial standard deviation (p = 0.205, independent T-test;
Figure 1C,D). For combined repeats of the same length,
transcriptional activity of (AC)23(AG)6 was similar to that of
(AC)21(AG)8 (p = 0.627, independent T-test; Figure 1E,F).
Thus, both AC and AG repeats contributed to the
transcriptional activity of the PAX6 P1 promoter.
Our luciferase-reporter analysis showed that
transcription activity increased with AC and AG repeat length.
This phenomenon may be due to influences of transcription
factor binding sites within this region. Thus, we used
(AC)20(AG)6 as a reference and predicted one binding site for
T-cell factor/Lymphoid enhancer factor family transcription
factors, one glucocorticoid receptor binding site, and four
transcription factor (TF) II-I binding sites (Figure 2B). With
decreasing AG repeat lengths, the T-cell factor/Lymphoid
enhancer factor and glucocorticoid receptor sites were
unchanged, but the TFII-I sites were reduced. Only two
predicted TFII-I sites were observed in (AC)15(AG)4 (Figure
2A). No alteration was observed with a decrease in AC repeat
TABLE 1. ALLELIC FREQUENCIES OF PAX6 P1 PROMOTER DINUCLEOTIDE REPEATS IN HIGH MYOPIA (HM) AND CONTROL SUBJECTS.
(AC)m repeat
Allelic count (%) Empirical
p-value (AG)n repeat
Allelic count (%) Empirical
p-value
HM n=750 Control n=678 HM n=758 Control n=698
(AC)70 (0.0) 1 (0.1) 0.013 (AG)40 (0.0) 1 (0.1) 0.012
(AC)15 0 (0.0) 2 (0.3) (AG)545 (5.9) 51 (7.3)
(AC)16 10 (1.3) 9 (1.3) (AG)6464 (61.2) 458 (65.6)
(AC)17 43 (5.7) 41 (6.0) (AG)7218 (28.8) 176 (25.2)
(AC)18 80 (10.7) 67 (9.9) (AG)831 (4.1) 12 (1.7)
(AC)19 100 (13.3) 138 (20.4)
(AC)20 155 (20.7) 134 (19.8)
(AC)21 149 (19.9) 99 (14.6)
(AC)22 161 (21.5) 138 (20.4)
(AC)23 29 (3.9) 33 (4.9)
(AC)24 13 (1.7) 13 (1.9)
(AC)25 6 (0.8) 2 (0.3)
(AC)26 4 (0.5) 1 (0.1)
TABLE 2. ALLELIC FREQUENCIES OF PAX6 P1 PROMOTER GROUPED DINUCLEOTIDE REPEATS, (AC)m AND (AG)n, IN HIGH MYOPIA (HM) AND CONTROL SUBJECTS.
Grouped
(AC) m repeat
Allelic count (%) Empirical
p-value
Grouped
(AG)n repeat
Allelic count (%) Empirical
p-value
HM n=750 Control n=678 HM n=758 Control n=698
(AC)Below 20-22 233 (31.1) 258 (38.1) 0.016 (AG)4-6 509 (67.2) 510 (73.1) 0.016
(AC)20-22 465 (62.0) 371 (54.7) (AG)7-8 249 (32.8) 188 (26.9)
(AC)Above 20-22 52 (6.9) 49 (7.2)
TABLE 3. GENOTYPIC FREQUENCIES OF PAX6 P1 PROMOTER GROUPED DINUCLEOTIDE REPEATS IN HIGH MYOPIA (HM) AND CONTROL SUBJECTS.
Grouped
(AC) m
genotype
Genotypic count (%) Empirical
p-value
Grouped
(AG)n
genotype
Genotypic count (%) Empirical
p-value
HM n=375 Control n=339 HM n=379 Control n=349
(AC)Below 20-22 /
(AC)Below 20-22
16 (4.3%) 40 (11.8%) 0.004 (AG)4-6 /
(AG)
173 (45.6%) 192 (55.0%) 0.039
(AC)Below 20-22 /
(AC)20-22
178 (47.5%) 149 (44.0%) (AG )4-6 /
(AG) 7-8
163 (43.0%) 126 (36.1%)
(AC)Below 20-22 /
(AC)Above 20-22
24 (6.4%) 29 (8.6%) (AG) 7-8 /
(AG) 7-8
43 (11.3%) 31 (8.9%)
(AC)20-22 / (AC)20-22 130 (34.7%) 103 (34.7%)
(AC)20-22 / (AC)Above
20-22
26 (6.9%) 16 (4.7%)
(AC)Above 20-22 /
(AC)Above 20-22
1 (0.3%) 2 (0.6%)
Molecular Vision 2009; 15:2239-2248 <http://www.molvis.org/molvis/v15/a241> © 2009 Molecular Vision
2242
4-6
lengths. Accordingly, more TFII-I sites were predicted with
increasing AG repeat lengths. Multiple sites for Wilms’ tumor
transcription factor without lysine-threonine-serine [Wt1(–
KTS)] were observed with an increase in AC repeat length,
and one GAGA factor binding site appeared with an increase
in AG repeat length. In (AC)26(AG)8, six TFII-I sites, six
Wt1(–KTS) sites, and one GAGA factor site were predicted
(Figure 2C).
DISCUSSION
We found no myopia mutations in the coding regions and
splice sites in PAX6 in our cohort of Chinese high myopia
patients. Some SNPs were detected in the P1 promoter, exon
7, and intron 10, but these were not statistically significant
(data not shown). In a recent report, two intronic SNPs
(rs3026390 and rs3026393, located in introns 12 and 13,
respectively) have been shown to be associated with high
myopia in Han Chinese nuclear families [66]. SNP
rs667773, located in intron 10, is in the same linkage
disequilibrium block with rs3026390 and rs3026393 [66].
However, in our study, no significant association was found
for rs667773 between high myopia patients and controls,
which was consistent with a previous case-control association
study in a Taiwan Chinese population [67]. The discrepancy
might be due to the much lower minor allele frequency of
rs667773 (0.137) than of rs3026390 and rs3026393 (0.472
and 0.493, respectively) [66]. Other studies have suggested
that rs667773, as a neural polymorphism, is an unlikely cause
of overt phenotypes such as aniridia [75,76].
The (AC)m(AG)n dinucleotide repeat sequence, located
about 1 kb from the transcription start site of the PAX6 P1
promoter, is highly polymorphic. The AC dinucleotide
polymorphism ranged from 18 to 31 repeats and AG ranged
from 5 to 7 repeats in a Caucasian population [60]. In our
Chinese cohort, the AC repeats ranged from 7 to 26 and the
AG repeats from 4 to 8 (Table 1). The allele size of the AG
repeats was similar in Caucasians and Chinese, but the AC
repeat length was longer in Caucasians. Notably, one (AC)7
allele was found in a control subject, far from the common
range of repeats between 15 and 26. In addition, many of the
dinucleotide repeats were heterozygous in both poly AC and
poly AG repeats (AC: 55.3% in controls and 75.9% in
patients; AG: 42.9% in controls and 53.4% in patients). The
observed heterozygosity rate was 65% in a Caucasian
population [60]. Although the allele number in that study was
defined as combined units of AC and AG repeats instead of
independent AC and AG alleles, the trend of heterozygosity
was similar to that in our work. These two dinucleotide repeats
Figure 1. Transcriptional activity of
dinucleotide repeats in the PAX6 P1
promoter. A 1,851 bp genomic fragment
(from –1278 to +573) containing the
PAX6 P1 promoter with different
dinucleotide repeats was cloned into an
empty pGL3-Basic vector (pGL3) and
transfected into ARPE-19 cells. The
activity of each allelic construct is
expressed relative to the construct
(AC)20(AG)6. Data are represented as
mean±SD for five independent
experiments. A and B: Immunoblotting
results and a bar chart show relative
luciferase activity for grouped (AC)m
repeats with a stable (AG)6. C and D:
Immunoblotting results and a bar chart
show relative luciferase activity for
(AG) repeats with (AC) . E and F:
Immunoblotting results and a bar chart
show relative luciferase activity for
combined (AC) (AG) repeats.
Molecular Vision 2009; 15:2239-2248 <http://www.molvis.org/molvis/v15/a241> © 2009 Molecular Vision
2243
21
n
mn
are, therefore, highly polymorphic both in Caucasians and in
Chinese.
The PAX6 P1, containing CCAAT boxes and a TATA-
like box, is likely a real promoter [58-60]. We evaluated the
influence of (AC)m(AG)n dinucleotide repeats on PAX6 P1
promoter activity by a luciferase-reporter assay and examined
the effects of repeat lengths as obtained from our high myopia
patients and controls. Since retinal pigment epithelium (RPE)
has been shown to have PAX6 P1 promoter activity [57], we
used an RPE cell line, ARPE-19, for transfection.
Immunoblotting showed that longer lengths of (AC)m have a
significant trend of increasing luciferase expression compared
with shorter lengths (Figure 1A,B), although this was not
observed for (AG)n, likely due to the substantial standard
deviation (Figure 1C,D).
We confirmed that transcriptional activity of
(AC)23(AG)6 was similar to that of (AC)21(AG)8 (Figure 1E,F),
suggesting that both (AC)m and (AG)n dinucleotide repeats
within the PAX6 P1 promoter contribute to transcriptional
activity and might work cooperatively as an unit. Previous
studies on luciferase-reporter assays assessed the promoter
activity invisibly by a luminometer [60,77]. In our study, we
monitored the luciferase-reporter assay by immunoblotting
using a commercially available antibody against firefly
luciferase and luciferase overexpression by pGL3-pCMV as
a positive control. There are technical advantages to this
method. The promoter activity could be visualized, and co-
transfection with another normalizing vector was not required,
as the luciferase intensity could be directly normalized with
the housekeeping protein, assuming the same transfection
efficiency among the constructs. The limitation of the
luciferase-reporter assay is that the effect of the dinucleotide
repeats on the transcriptional activity was performed using
RPE cells from normal controls, which might not truly reflect
the situation in high myopia unless the experiment were
performed using cells from a highly myopic individual.
Since levels of Pax6 are tightly controlled, small and
seemingly insignificant changes in the levels of Pax6 may lead
to significant phenotypic consequences [78]. Moreover, the
Pax6 protein could upregulate PAX6 P1 promoter activity
Figure 2. Transcription factor binding
site prediction for dinucleotide repeats
in the PAX6 P1 promoter. The cloned
PAX6 P1 promoter DNA sequence was
used to predict transcription factor
binding sites. Predicted transcription
factor binding sites around the region of
the dinucleotide repeats are shown, and
different lengths of AC and AG repeats
are assessed. As in the immunoblotting
analysis, [(AC)20(AG)6] was set as a
reference. A: Predicted transcription
factor binding sites for (AC) (AG)
are shown. B: Predicted transcription
factor binding sites for (AC) (AG)
are shown. C: Predicted transcription
factor binding sites for (AC) (AG)
are shown.
Molecular Vision 2009; 15:2239-2248 <http://www.molvis.org/molvis/v15/a241> © 2009 Molecular Vision
2244
15 4
20 6
26 8
[77]. Results of our genotyping and promoter activity analyses
indicate that longer lengths of dinucleotide repeats increase
the expression of PAX6, which increases the risk of high
myopia. This postulation may be supported by several
assertions: (1) PAX6 gene expression has been shown to be
significantly higher in the retinas of optical defocused eyes
than in contralateral eyes in the rhesus monkey [79], and
expression of PAX6 was also increased in posthatch chicken
eyes with form-deprivation myopia [78]. (2) In another study,
the number of dividing retinal progenitor cells, of which
PAX6 is a marker, was highly correlated with axial elongation
of the eye, resulting in myopic refractive errors in primates
with form-deprivation myopia [80]. (3) Pax6 has been shown
to transactivate insulin promoters [81] and promote proinsulin
processing [82]. As insulin is a strong stimulator of axial
myopia in chicks [83], elevated PAX6 expression may
increase the risk of developing myopia through increased
expression of insulin. Chronic hyperinsulinemia has been
proposed as a key player in the pathogenesis of juvenile-onset
myopia [84]. Although Pax6 also transactivates the glucagon
promoter [81], which is a “stop” for myopia [85], insulin
might overcome the effects of glucagon in the development
of myopia [86].
The transcription factor binding site prediction (Figure 2)
showed that an increase in AC repeat length created additional
Wt1(–KTS) binding sites, while an increase in AG repeat
length created TFII-I and GAGA factor binding sites. If the
AG repeat length was reduced, TFII-I sites were also reduced.
Wt1(–KTS) is necessary for normal retina formation in mice
[87], while TFII-I is a signal-induced multifunctional
transcription factor that plays a key role in the regulation of
cell proliferation [88]. Moreover, the GAGA factor, a
transcription activator, is activated by epidermal growth
factors, platelet-derived growth factors, and insulin [89].
These growth factors could regulate PAX6 transcription
through the GAGA factor binding site.
In summary, we found no association between
polymorphisms in the PAX6 coding region and high myopia
in our Hong Kong Chinese cohort. Two dinucleotide repeats,
AC and AG, in the PAX6 P1 promoter were associated with
high myopia. These two repeats were also associated with the
elevation of PAX6 P1 promoter activity, and hence an increase
in the transcriptional activity of PAX6. Our results provide
evidence for the role of PAX6 in the pathogenesis of high
myopia.
ACKNOWLEDGMENTS
We express our greatest appreciation to all the participants in
the study. This study was supported by a block Grant, The
Chinese University of Hong Kong and the Lim Por Yen Eye
Foundation Endowment Fund.
REFERENCES
1. Wong TY, Foster PJ, Hee J, Ng TP, Tielsch JN, Chew SJ,
Johnson GJ. Seah SKL. Prevalence and risk factors for
refractive errors in adult Chinese in Singapore. Invest
Ophthalmol Vis Sci 2000; 41:2486-94. [PMID: 10937558]
2. Kleinstein RN, Jones LA, Hullett S, Kwon S, Lee RL, Friedman
NE, Manny RE, Mutti DO, Yu JA, Zadnik K, Collaborative
Longitudinal Evaluarion of Ethnicity and Refractive Error
Study Group. Refractive error and ethnicity in children. Arch
Ophthalmol 2003; 121:1141-7. [PMID: 12912692]
3. Saw SM, Tong L, Chua WH, Chia KS, Koh D, Tan DT, Katz J.
Incidence and progression of myopia in Singaporean school
children. Invest Ophthalmol Vis Sci 2005; 46:51-7. [PMID:
15623754]
4. Fan DS, Lam DS, Lam RF, Lau JT, Chong KS, Cheung EY, Lai
RY, Chew SJ. Prevalence, incidence, and progression of
myopia of school children in Hong Kong. Invest Ophthalmol
Vis Sci 2004; 45:1071-5. [PMID: 15037570]
5. Lam CS, Goldschmidt E, Edwards MH. Prevalence of myopia
in local and international schools in Hong Kong. Optom Vis
Sci 2004; 81:317-22. [PMID: 15181356]
6. Saw SM, Shankar A, Tan SB, Taylor H, Tan DT, Stone RA,
Wong TY. A cohort study of incident myopia in Singaporean
children. Invest Ophthalmol Vis Sci 2006; 47:1839-44.
[PMID: 16638989]
7. Mastropasqua L, Lobefalo L, Mancini A, Ciancaglini M, Palma
S. Prevalence of myopia in open angle glaucoma. Eur J
Ophthalmol 1992; 2:33-5. [PMID: 1638164]
8. Hochman MA, Seery CM, Zarbin MA. Pathophysiology and
management of subretinal hemorrhafe. Surv Ophthalmol
1997; 42:195-213. [PMID: 9406367]
9. Lim R, Mitchell P, Cumming RG. Refractive associations with
cataract: the Blue Mountains Eye Study. Invest Ophthalmol
Vis Sci 1999; 40:3021-6. [PMID: 10549667]
10. Banker AS, Freeman WR. Retinal detachment. Ophthalmol
Clin North Am 2001; 14:695-704. [PMID: 11787748]
11. Mutti DO, Mitchell GL, Oeschberge ML, Jones LA, Zadnik K.
Parental myopia, near work, school schievment, and
children’s refractive error. Invest Ophthalmol Vis Sci 2002;
43:3633-40. [PMID: 12454029]
12. Saw SM, Chua WH, Hong CY, Wu HM, Chan WY, Chia KS,
Stone RA, Tan D. Nearwork in early-onset myopia. Invest
Ophthalmol Vis Sci 2002; 43:332-9. [PMID: 11818374]
13. Hammond CJ, Snieder H, Gilbert CE, Spector TD. Genes and
environment in refractive error: the twin eye study. Invest
Ophthalmol Vis Sci 2001; 42:1232-6. [PMID: 11328732]
14. Lyhne N, Sjolie AK, Kyvik KO, Green A. The important of
genes and environment for ocular refractive and its
determiners: a population based study among 20-45 year old
twins. Br J Ophthalmol 2001; 85:1470-6. [PMID: 11734523]
15. Dirani M, Chamberlain M, Shekar SN, Islam AF, Garoufalis P,
Chen CY, Guymer RH, Baird PN. Heritability of refractive
error and ocular biometrics: the Genes in Myopia (GEM) twin
study. Invest Ophthalmol Vis Sci 2006; 47:4756-61. [PMID:
17065484]
16. Lopes MC, Andrew T, Carbonaro F, Spector TD, Hammond CJ.
Estimating heritability and shared environmental effects for
refractive error in twin and family studies. Invest Ophthalmol
Vis Sci 2009; 50:126-31. [PMID: 18757506]
17. Angi MR, Clementi M, Sardei C, Piattelli E, Bisantis C.
Heritability of myopic refractive errors in identical and
fraternal twins. Graefes Arch Clin Exp Ophthalmol 1993;
231:580-5. [PMID: 8224933]
Molecular Vision 2009; 15:2239-2248 <http://www.molvis.org/molvis/v15/a241> © 2009 Molecular Vision
2245
18. Liang CL, Yen E, Su JY, Liu C, Chang TY, Park N, Wu MJ,
Lee S, Flynn JT, Jou SH. Impact of family history of high
myopia on level and onset of myopia. Invest Ophthalmol Vis
Sci 2004; 45:3446-52. [PMID: 15452048]
19. Schwartz M, Haim M, Skarsholm D. X-linked myopia:
Bornholm eye disease. Linkage to DNA markers on the distal
part of Xq. Clin Genet 1990; 38:281-6. [PMID: 1980096]
20. Young TL, Ronana SM, Drahozal LA, Wildenberg SC, Alvear
AB, Oetting WS, Atwood LD, Wilkin DJ, King RA. Evidence
that a locus for familial high myopia maps to chromosome
18p. Am J Hum Genet 1998; 63:109-19. [PMID: 9634508]
21. Lam DS, Tam PO, Fan DS, Baum L, Leung YF, Pang CP.
Familial high myopia linkage to chromosome 18p.
Ophthalmologica 2003; 217:115-8. [PMID: 12592049]
22. Young TL, Ronan SM, Alvear AB, Wildenberg SC, Oetting
WS, Atwood LD, Wilkin DJ, King RA. A second locus for
familial high myopia maps to chromosome 12q. Am J Hum
Genet 1998; 63:1419-24. [PMID: 9792869]
23. Naiglin L, Gazagne C, Dallongeville F, Thalamas C, Idder A,
Rascol O, Malecaze F, Calvas P. Genome wide scan for
familial high myopia suggests a novel locus on chromosome
7q36. J Med Genet 2002; 39:118-24. [PMID: 11836361]
24. Paluru P, Ronan SM, Heon E, Devoto M, Wildenberg SC,
Scavello G, Holleschau A, Makitie O, Cole WG, King RA,
Young TL. New locus for autosomal dominant high myopia
maps to the long arm of chromosome 17. Invest Ophthalmol
Vis Sci 2003; 44:1830-6. [PMID: 12714612]
25. Stambolian D, Ibay G, Reider L, Dana D, Moy C, Schlifka M,
Holmes T, Ciner E, Bailey-Wilson JE. Genomewide linkage
scan for myopia susceptibility loci among Ashkenazi Hewish
families shows evidence of linkage on chromosome 22q12.
Am J Hum Genet 2004; 75:448-59. [PMID: 15273935]
26. Hammond CJ, Andrew T, Mak YT, Spector TD. A
susceptibility locus for myopia in the normal population is
linked to the PAX6 gene region on chromosome 11: a
genomewide scan of dizygotic twins. Am J Hum Genet 2004;
75:294-304. [PMID: 15307048]
27. Zhang Q, Guo X, Xiao X, Jia X, Li S, Hejtmancik JF. A new
locus for autosomal dominant high myopia maps to 4q22-q27
between D2S1578 and D4S1612. Mol Vis 2005; 11:554-60.
[PMID: 16052171]
28. Paluru PC, Nallasamy S, Devoto M, Rappaport EF, Young TL.
Identification of a novel locus on 2q for autosomal dominant
high-grade myopia. Invest Ophthalmol Vis Sci 2005;
46:2300-7. [PMID: 15980214]
29. Zhang Q, Guo X, Xiao X, Jia X, Li S, Hejtmancik JF. Novel
locus for X linked recessive high myopia maps to Xq23-q25
but outside MYP1. J Med Genet 2006; 43:e20. [PMID:
16648373]
30. Wojciechowski R, Moy C, Ciner E, Ibay G, Reider L, Bailey-
Wilson JE, Stambolian D. Genomewide scan in Ashkenazi
Jewish families demonstrates evidence of linkage of ocular
refraction to a QTL on chromosome 1p36. Hum Genet 2006;
119:389-99. [PMID: 16501916]
31. Nallasamy S, Paluru PC, Devoto M, Wasserman NF, Zhou J,
Young TL. Genetic linkage study of high-grade myopia in a
Hutterite population from South Dakota. Mol Vis 2007;
13:229-36. [PMID: 17327828]
32. Lam CY, Tam PO, Fan DS, Fan BJ, Wang DY, Lee CW, Pang
CP, Lam DS. A genome-wide scan maps a novel high myopia
locus to 5p15. Invest Ophthalmol Vis Sci 2008; 49:3768-78.
[PMID: 18421076]
33. Ciner E, Wojciechowski R, Ibay G, Bailey-Wilson JE,
Stambolian D. Genomewide scan of ocular refraction in
African-American families shows significant linkage to
chromosome 7p15. Genet Epidemiol 2008; 32:454-63.
[PMID: 18293391]
34. Paget S, Julia S, Vitezica ZG, Soler V, Malecaze F, Calvas P.
Linkage analysis of high myopia susceptibility locus in 26
families. Mol Vis 2008; 14:2566-74. [PMID: 19122830]
35. Yang Z, Xiao X, Li S, Zhang Q. Clinical and linkage study on
a consanguineous Chinese family with autosomal recessive
high myopia. Mol Vis 2009; 15:312-8. [PMID: 19204786]
36. Yu ZQ, Li YB, Huang CX, Chu RY, Hu DN, Shen ZH, Huang
W. A genome-wide screening for pathological myopia
suggests a novel locus on chromosome 15q12-13. Zhonghua
Yan Ke Za Zhi 2007; 43:233-8. [PMID: 17605906]
37. Nishizaki R, Ota M, Inoko H, Meguro A, Shiota T, Okada E,
Mok J, Oka A, Ohno S, Mizuki N. New susceptibility locus
for high myopia is linked to the uromodulin-like 1
(UMODL1) gene region on chromosome 21q22.3. Eye 2009;
23:222-9. [PMID: 18535602]
38. Wojciechowski R, Stambolian D, Ciner E, Ibay G, Holmes TN,
Bailey-Wilson JE. Genomewide linkage scans for ocular
refraction and meta-analysis of four populations in the
Myopia Family Study. Invest Ophthalmol Vis Sci 2009;
50:2024-32. [PMID: 19151385]
39. Li YJ, Guggenheim JA, Bulusu A, Metlapally R, Abbott D,
Malecaze F, Calvas P, Rosenberg T, Paget S, Creer RC, Kirov
G, Owen MJ, Zhao B, White T, Mackey DA, Young TL. An
international collaborative family-based whole-genome
linkage scan for high-grade myopia. Invest Ophthalmol Vis
Sci 2009; 50:3116-27. [PMID: 19324860]
40. Schache M, Chen CY, Pertile KK, Richardson AJ, Dirani M,
Mitchell P, Baird PN. Fine mapping linkage analysis
identifies a novel susceptibility locus for myopia on
chromosome 2q37 adjacent to but not overlapping MYP12.
Mol Vis 2009; 15:722-30. [PMID: 19365569]
41. Lam DS, Lee WS, Leung YF, Tam PO, Fan DS, Fan BJ, Pang
CP. TGFbeta-induced factor: a candidate gene for high
myopia. Invest Ophthalmol Vis Sci 2003; 44:1012-5. [PMID:
12601022]
42. Yanovitch T, Li YJ, Metlapally R, Abbott D, Viet KN, Young
TL. Hepatocyte growth factor and myopia: genetic
association analyses in a Caucasian population. Mol Vis
2009; 15:1028-35. [PMID: 19471602]
43. Hall NF, Gale CR, Ye S, Martyn CN. Myopia and
polymorphisms in genes for matrix metalloproteinases. Invest
Ophthalmol Vis Sci 2009; 50:2632-6. [PMID: 19279308]
44. Inamori Y, Ota M, Inoko H, Okada E, Nishizaki R, Shiota T,
Mok J, Oka A, Ohno S, Mizuki N. The COL1A1 gene and
high myopia susceptibility in Japanese. Hum Genet 2007;
122:151-7. [PMID: 17557158]
45. Mutti DO, Cooper ME, O’Brien S, Jones LA, Marazita ML,
Murray JC, Zadnik K. Candidate gene and locus analysis of
myopia. Mol Vis 2007; 13:1012-9. [PMID: 17653045]
46. Lin HJ, Wan L, Tsai Y, Tsai YY, Fan SS, Tsai CH, Tsai FJ. The
TGFβ1 gene codon 10 polymorphism contributes to the
genetic predisposition to high myopia. Mol Vis 2006;
12:698-703. [PMID: 16807529]
Molecular Vision 2009; 15:2239-2248 <http://www.molvis.org/molvis/v15/a241> © 2009 Molecular Vision
2246
47. Lin HJ, Wan L, Tsai Y, Liu SC, Chen WC, Tsai SW, Tsai FJ.
Sclera-related gene polymorphisms in high myopia. Mol Vis
2009; 15:1655-63. [PMID: 19710942]
48. Lin HJ, Kung YJ, Lin YJ, Sheu JJ, Chen BH, Lan YC, Lai CH,
Shu YA, Wan L, Tsai FJ. Association of the Lumican gene
functional 3’ UTR polymorphism with high myopia. Invest
Ophthalmol Vis Sci. 2009 [PMID: 19643966]
49. Khor CC, Grignani R, Ng DP, Toh KY, Chia KS, Tan D, Goh
DL, Saw SM. cMET and refractive error progression in
children. Ophthalmology 2009; 116:1469-74. [PMID:
19500853]
50. Nishina S, Kohsaka S, Yamaguchi Y, Handa H, Kawakami A,
Fujisawa H, Azuma N. PAX6 expression in the developing
human eye. Br J Ophthalmol 1999; 83:723-7. [PMID:
10340984]
51. Xu PX, Zhang X, Heaney S, Yoon A, Michelson AM, Maas RL.
Regulation of Pax6 expression is conserved between mice and
flies. Development 1999; 126:383-95. [PMID: 9847251]
52. Morgan R. Conservation of sequence and function in the Pax6
regulatory elements. Trends Genet 2004; 20:283-7. [PMID:
15219391]
53. Lakowski J, Majumder A, Lauderdale JD. Mechanisms
controlling Pax6 isoform expression in the retina have been
conserved between teleosts and mammals. Dev Biol 2007;
307:498-520. [PMID: 17509554]
54. Simpson TI, Price DJ. Pax6; a pleiotropic player in
development. Bioessays 2002; 24:1041-51. [PMID:
12386935]
55. Hill RE, Favor J, Hogan BL, Ton CC, Saunder GF, Hanson IM,
Prosser J, Jordan T, Hastie ND, van Heyningen V. Mouse
small eye results from mutations in a paired-like homeobox-
containing gene. Nature 1991; 354:522-5. [PMID: 1684639]
56. Kamachi Y, Uchikawa M, Tanouchi A, Sekido R, Kondoh H.
Pax6 and SOX2 form a co-DNA-binding partner complex that
regulates initiation of lens development. Genes Dev 2001;
15:1272-86. [PMID: 11358870]
57. Plaza S, Dozier C, Turque N, Saule S. Quail Pax-6 (Pax-QNR)
mRNAs are expressed from two promoters used differentially
during retina development and neuronal differentiation. Mol
Cell Biol 1995; 15:3344-53. [PMID: 7760830]
58. Xu ZP, Saunders GF. Transcriptional regulation of the human
PAX6 gene promoter. J Biol Chem 1997; 272:3430-6.
[PMID: 9013587]
59. Okladnova O, Syagailo YV, Mossner R, Riederer P, Lesch KP.
Regulation of PAX-6 gene transcription: alternate promoter
usage in human brain. Brain Res Mol Brain Res 1998;
60:177-92. [PMID: 9757029]
60. Okladnova O, Syagailo YV, Tranitz M, Stober G, Riederer P,
Mossner R, Lesch KP. A promoter-associated polymorphic
repeat modulates PAX-6 expression in human brain. Biochem
Biophys Res Commun 1998; 248:402-5. [PMID: 9675149]
61. Glaser T, Jepeal L, Edwards JG, Young SR, Favor J, Maas RL.
PAX6 gene dosage effect in a family with congenital
cataracts, aniridia, anophthalmia and central nervous system
defects. Nat Genet 1994; 7:463-71. [PMID: 7951315]
62. Schedl A, Ross A, Lee M, Engelkamp D, Rashbass P, van
Heyningen V, Hastie ND. Influence of PAX6 gene dosage on
development: overexpression causes severe eye
abnormalities. Cell 1996; 86:71-82. [PMID: 8689689]
63. Favor J, Gloeckner CJ, Neuhauser-Klaus A, Pretsch W,
Sandulache R, Saule S, Zaus I. Relationship of Pax6 activity
levels to the extent of eye development in the mouse, Mus
musculus. Genetics 2008; 179:1345-55. [PMID: 18562673]
64. Simpson CL, Hysi P, Bhattacharya SS, Hammond CJ, Webster
A, Peckham CS, Sham PC, Rahi JS. The roles of PAX6 and
SOX2 in myopia: lessons from the 1958 British birth cohort.
Invest Ophthalmol Vis Sci 2007; 48:4421-5. [PMID:
17898260]
65. Hewitt AW, Kearns LS, Jamieson RV, Williamson KA, van
Heyningen V, Mackey DA. PAX6 mutations may be
associated with high myopia. Ophthalmic Genet 2007;
28:179-82. [PMID: 17896318]
66. Han W, Leung KH, Fung WY, Mak JY, Li YM, Yap MK, Yip
SP. Association of PAX6 polymorphisms with high myopia
in Han Chinese nuclear families. Invest Ophthalmol Vis Sci
2009; 50:47-56. [PMID: 19124844]
67. Tsai YY, Chiang CC, Lin HJ, Lin JM, Wan J, Tsai FJA. PAX6
gene polymorphism is associated with genetic predisposition
to extreme myopia. Eye 2008; 22:576-81. [PMID: 17948041]
68. Andersson S, Davis DL, Dahlback H, Jornvall H, Russell DW.
Cloning, structure, and expression of the mitochondrial
cytochrome P-450 sterol 26-hydroxylase, a bile acid
biosynthetic enzyme. J Biol Chem 1989; 264:8222-9. [PMID:
2722778]
69. Dunn KC, Aotaki-Keen AE, Putkey FR, Hjelmeland LM.
ARPE-19, a human retinal pigment epithelial cell line with
differentiated properties. Exp Eye Res 1996; 62:155-69.
[PMID: 8698076]
70. Sham PC, Curtis D. Monte Carlo tests for association between
disease and alleles at highly polymorphic loci. Ann Hum
Genet 1995; 59:97-105. [PMID: 7762987]
71. Chen YH, Lin SJ, Lin MW, Tsai HL, Kuo SS, Chen JW, Charng
MJ, Wu TC, Chen LC, Ding PY, Pan WH, Jou YS, Chau LY.
Microsatellite polymorphism in promoter of heme
oxygenase-1 gene is associated with susceptibility to
coronary artery disease in type 2 diabetic patients. Hum Genet
2002; 111:1-8. [PMID: 12136229]
72. Rantner B, Kollerits B, Anderwald-Stadler M, Klein-Weigel P,
Gruber I, Gehringer A, Haak M, Schnapka-Kopf M, Fraedrich
G, Kronenberg F. Association between the UGT1A1 TA-
repeat polymorphism and bilirubin concentration in patients
with intermittent claudication: Results from the CAVASIC
study. Clin Chem 2008; 54:851-7. [PMID: 18375480]
73. Schug J. Using TESS to predict transcription factor binding sites
in DNA sequence. In: Baxevanis AD, editor. Current
Protocols in Bioinformatics. J. Wiley and Sons; 2003.
74. TESS. Transcription Element Search Software on the WWW.
Schug J, Overton GC. Technical Report CBIL-
TR-1997-1001-v0.0. Computational Biology and Informatics
Laboratory, School of Medicine, University of Pennsylvania,
1997. <http://www.cbil.upenn.edu/tess>.
75. Neethirajan G, Krishnadas SR, Vijayalakshmi P, Shashikant S,
Sundaresan P. PAX6 gene variations associated with aniridia
in south India. BMC Med Genet 2004; 5:9. [PMID:
15086958]
76. Brown A, McKie M, van Heyningen V, Prosser J. The human
PAX6 mutation database. Nucleic Acids Res 1998;
26:259-64. [PMID: 9399848]
Molecular Vision 2009; 15:2239-2248 <http://www.molvis.org/molvis/v15/a241> © 2009 Molecular Vision
2247
77. Grocott T, Frost V, Maillard M, Johansen T, Wheeler GN,
Dawes LJ, Wormstone IM, Chantry A. The MH1 domain of
Smad3 interacts with Pax6 and represses autoregulation of the
Pax6 P1 promoter. Nucleic Acids Res 2007; 35:890-901.
[PMID: 17251190]
78. Bhat SP, Rayner SA, Chau SC, Ariyasu RG. Pax-6 expression
in posthatch chick retina during and recovery from form-
deprivation myopia. Dev Neurosci 2004; 26:328-35. [PMID:
15855761]
79. Zhong XW, Ge J, Deng W, Chen X, Huang J. Expression of
pax-6 in rhesus monkey of optical defocus induced myopia
form deprivation myopia. Chin Med J 2004; 117:722-6.
[PMID: 15161541]
80. Tkatchenko AV, Walsh PA, Tkatchenko TV, Gustincich S,
Raviola E. Form deprivation modulates retinal neurogenesis
in primate experimental myopia. Proc Natl Acad Sci USA
2006; 103:4681-6. [PMID: 16537371]
81. Sander M, Neubuser A, Kalamaras J, Ee HC, Martin GR,
German MS. Genetic analysis reveals that PAX6 is required
for normal transcription of pancreatic hormone genes and islet
development. Genes Dev 1997; 11:1662-73. [PMID:
9224716]
82. Wen JH, Chen YY, Song SJ, Ding J, Gao Y, Hu QK, Feng RP,
Liu YZ, Ren GC, Zhange CY, Hong TP, Gao X, Li LS. Paired
box 6 (PAX6) regulates glucose metabolism via proinsulin
processing mediated by prohormone convertase 1/3 (PC1/3).
Diabetologia 2009; 52:504-13. [PMID: 19034419]
83. Feldkaemper MP, Neacsu I, Schaeffel F. Insulin acts as a
powerful stimulator of axial myopia in chicks. Invest
Ophthalmol Vis Sci 2009; 50:13-23. [PMID: 18599564]
84. Cordain L, Eaton SB, Brand Miller J, Lindeberg S, Jensen C.
An evolutionary analysis of the aetiology and pathogenesis of
juvenile-onset myopia. Acta Ophthalmol Scand 2002;
80:125-35. [PMID: 11952477]
85. Feldkaemper MP, Schaeffel F. Evidence for a potential role of
glucagon during eye growth regulation in chicks. Vis
Neurosci 2002; 19:755-66. [PMID: 12688670]
86. Zhu X, Wallman J. Opposite effects of glucagon and insulin on
compensation for spectacle lenses in chicks. Invest
Ophthalmol Vis Sci 2009; 50:24-36. [PMID: 18791176]
87. Wagner KD, Wagner N, Vidal VP, Schley G, Wilhelm D,
Schedl A, Englert C, Scholz H. The Wilms’ tumor gene Wt1
is required for normal development of the retina. EMBO J
2002; 21:1398-405. [PMID: 11889045]
88. Roy AL. Signal-induced functions of the transcription factor
TFII-I. Biochim Biophys Acta 2007; 1769:613-621. [PMID:
17976384]
89. Wyse BD, Linas SL, Thekkumkara TJ. Functional role of a
novel cis-acting element (GAGA box) in human type-1
angiotensin II receptor gene transcription. J Mol Endocrinol
2000; 25:97-108. [PMID: 10915222]
Molecular Vision 2009; 15:2239-2248 <http://www.molvis.org/molvis/v15/a241> © 2009 Molecular Vision
The print version of this article was created on 3 November 2009. This reflects all typographical corrections and errata to the
article through that date. Details of any changes may be found in the online version of the article.
2248
... Genetic studies, exploring the association of genes with myopia, provide insights on the mechanism of myopia and myopia progression. We have previously identified 4q25, 15q14 and MIPEP (mitochondrial intermediate peptidase) variants associated with different severity of myopia in Chinese population, jointly increasing the risk predisposing to high myopia by 10 folds (5). Moreover, we also identified the chromosome 13q12.12 ...
... There were several limitations in this study. First, the ratio of total myopia subjects to the control subjects was 3.65:1; yet, the ratio of myopia subjects with particular severity to the control subjects was close to 1:1, which is similar to our previous study (5). Second, the current study was just based on a cohort of Han Chinese population in southern China. ...
Article
Background: Myopia is the most prevalent ocular disorder in the world, and corneal parameters have been regarded as key ocular biometric parameters determining the refractive status. Here, we aimed to determine the association of genome-wide association study-identified corneal curvature (CC)-related gene variants with different severity of myopia and ocular biometric parameters in Chinese population. Methods: Total 2,101 unrelated Han Chinese subjects were recruited, including 1,649 myopia and 452 control subjects. Five previously reported CC-associated gene variants (PDGFRA, MTOR, WNT7B, CMPK1 and RBP3) were genotyped by TaqMan assay, and their association with different myopia severity and ocular biometric parameters were evaluated. Results: Joint additive effect analysis showed that MTOR rs74225573 paired with PDGFRA rs2114039 (P = .009, odds ratio (OR) = 4.91) or CMPK1 rs17103186 (P = .002, OR = 13.03) were significantly associated with higher risk in mild myopia. Critically, mild myopia subjects had significantly higher frequency in MTOR rs74225573 C allele than high myopia subjects (P = .003), especially in male subjects (P = .001, OR = 0.49). High myopia subjects carrying MTOR rs74225573 C allele have significant flatter CC (P = .035) and longer corneal radius (P = .044) than those carrying TT genotype. Conclusion: This study revealed that male high myopia subjects are more prone to carry CC-related MTOR rs74225573 T allele, whereas mild myopia subjects are prone to carry the C allele. MTOR rs7422573 variant could be a genetic marker to differentiate mild from high myopia in risk assessment. Abbreviations: ACD: anterior chamber depth; AL: axial length; AL/CR: axial length/corneal radius ratio; ANOVA: analysis of variance; CC: corneal curvature; CCT: central corneal thickness; C.I.: confidence interval; CMPK1: cytidine/uridine monophosphate kinase 1; CR: corneal radius; D: diopter; GWAS: genome-wide association studies; HWE: Hardy-Weinberg equilibrium; LT: lens thickness; MIPEP: mitochondrial intermediate peptidase; MTOR: mechanistic target of rapamycin kinase; OR: odds ratio; PDGFRA: platelet-derived growth factor receptor-α; RBP3: retinol-binding protein 3; SD: standard deviation; SE: spherical equivalence; SNTB1: syntrophin beta 1; VCD: vitreous chamber depth; VIPR2: vasoactive intestinal peptide receptor 2; WNT7B: wingless/integrated family member 7B.
... Taip pat nustatyta, kad AC ir AG dinukleotidų pasikartojimai PAX 6 P1 promotoriuje sukelia transkripcijos aktyvaciją, kurie ir yra siejami su didelio laipsnio trumparegyste (Ng, 2009). ...
... Moreover, the risk allele could reduce PAX6 protein levels, which may be involved in the underlying mechanism of myopia pathogenesis [81]. Increases in the length of the PAX6 P1 promoter AG dinucleotide repeatedly affect the transcription activity and are associated with high myopia [82]. ...
Article
Full-text available
Myopia is the most common eye condition leading to visual impairment and is greatly influenced by genetics. Over the last two decades, more than 400 associated gene loci have been mapped for myopia and refractive errors via family linkage analyses, candidate gene studies, genome-wide association studies (GWAS), and next-generation sequencing (NGS). Lifestyle factors, such as excessive near work and short outdoor time, are the primary external factors affecting myopia onset and progression. Notably, besides becoming a global health issue, myopia is more prevalent and severe among East Asians than among Caucasians, especially individuals of Chinese, Japanese, and Korean ancestry. Myopia, especially high myopia, can be serious in consequences. The etiology of high myopia is complex. Prediction for progression of myopia to high myopia can help with prevention and early interventions. Prediction models are thus warranted for risk stratification. There have been vigorous investigations on molecular genetics and lifestyle factors to establish polygenic risk estimations for myopia. However, genes causing myopia have to be identified in order to shed light on pathogenesis and pathway mechanisms. This report aims to examine current evidence regarding (1) the genetic architecture of myopia; (2) currently associated myopia loci identified from the OMIM database, genetic association studies, and NGS studies; (3) gene-environment interactions; and (4) the prediction of myopia via polygenic risk scores (PRSs). The report also discusses various perspectives on myopia genetics and heredity.
... Our study proved that alleles with six repeats of GTTT are associated with Alzheimer's disease (AD) and multiple sclerosis (MS) (Rezazadeh et al. 2015b). Table 2 provides a summary of details including STR repeat motifs, their genetic locations, their normal range, expansion/ deletion mechanisms, and gene and organism names (Ng et al. 2009;Rezazadeh et al. 2015a;Bahlo et al. 2018;Sawaya et al. 2012;Chen et al. 2009;Deelen et al. 1994;Ouyang et al. 2012 ...
Article
Full-text available
Short tandem repeats (STRs) are commonly defined as short runs of repetitive nucleotides, consisting of tandemly repeating 2–6- bp motif units, which are ubiquitously distributed throughout genomes. Functional STRs are polymorphic in the population, and their variations influence gene expression, which subsequently may result in pathogenic phenotypes. To understand STR phenotypic effects and their functional roles, we describe four different mutational mechanisms including the unequal crossing-over model, gene conversion, retrotransposition mechanism and replication slippage. Due to the multi-allelic nature, small length, abundance, high variability, codominant inheritance, nearly neutral evolution, extensive genome coverage and simple assaying of STRs, these markers are widely used in various types of biological research, including population genetics studies, genome mapping, molecular epidemiology, paternity analysis and gene flow studies. In this review, we focus on the current knowledge regarding STR genomic distribution, function, mutation and applications.
Article
Full-text available
Mutations that provide environment-dependent selective advantages drive adaptive divergence among species. Many phenotypic differences among related species are more likely to result from gene expression divergence rather than from non-synonymous mutations. In this regard, cis-regulatory mutations play an important part in generating functionally significant variation. Some proposed mechanisms that explore the role of cis-regulatory mutations in gene expression divergence involve microsatellites. Microsatellites exhibit high mutation rates achieved through symmetric or asymmetric mutation processes and are abundant in both coding and non-coding regions in positions that could influence gene function and products. Here we tested the hypothesis that microsatellites contribute to gene expression divergence among species with 50 individuals from five closely related Helianthus species using an RNA-seq approach. Differential expression analyses of the transcriptomes revealed that genes containing microsatellites in non-coding regions (UTRs and introns) are more likely to be differentially expressed among species when compared to genes with microsatellites in the coding regions and transcripts lacking microsatellites. We detected a greater proportion of shared microsatellites in 5′UTRs and coding regions compared to 3′UTRs and non-coding transcripts among Helianthus spp. Furthermore, allele frequency differences measured by pairwise FST at single nucleotide polymorphisms (SNPs), indicate greater genetic divergence in transcripts containing microsatellites compared to those lacking microsatellites. A gene ontology (GO) analysis revealed that microsatellite-containing differentially expressed genes are significantly enriched for GO terms associated with regulation of transcription and transcription factor activity. Collectively, our study provides compelling evidence to support the role of microsatellites in gene expression divergence.
Chapter
Full-text available
The Consortium for Refractive Error and Myopia (CREAM) is an international collaboration founded to increase knowledge on the genetic background of refractive error and myopia. The consortium was established in 2011 and consists of >50 studies from all over the world with epidemiological and genetic data on myopia endophenotypes. Due to these efforts, almost 200 genetic loci for refractive error and myopia have been identified. These genetic risk variants mostly carry low risk but are highly prevalent in the general population. The genetic loci are expressed in all retinal cell layers and play a role in different processes, e.g., in phototransduction or extracellular matrix remodeling. The work of CREAM over the years has implicated the major pathways in conferring susceptibility to myopia and supports the notion that myopia is caused by a light-dependent retina-to-sclera signaling cascade. The current genetic findings offer a world of new molecules involved in myopiagenesis. However, as the currently identified genetic loci explain only a fraction of the high heritability, further genetic advances are needed. It is recommended to expand large-scale, in-depth genetic studies using complementary big data analytics, to consider gene-environment effects by thorough measurements of environmental exposures, and to focus on subgroups with extreme phenotypes and high familial occurrence. Functional characterization of associated variants is simultaneously needed to bridge the knowledge gap between sequence variance and consequence for eye growth. The CREAM consortium will endeavor to play a pivotal role in these future developments.
Chapter
Common eye diseases, including myopia, cataract, glaucoma, and age-related macular degeneration, are the leading cause of blindness and visual impairment, affecting billions of people worldwide. Unlike monogenic diseases, the inheritance of common eye diseases is complex, interplaying with genetics and environmental factors. Genome-wide association studies (GWAS) have identified hundreds of associated genes for common eye diseases; yet, the biological correlation of these disease-associated genes with the pathogenesis of the common eye diseases remains elusive. Apart from the involvement of multiple genes, the epigenetic regulation by environmental factors, including cigarette smoking and sunlight exposure, also determines the occurrence and etiology of the complex diseases. A gene promoter is composed of multiple transcription factor binding sites, which time-dependently regulates the spatial expression of a gene. Genetic variants in the promoter region, creating or disrupting the transcription factor binding sites, could impair the expression of the disease-associated genes and contribute to the pathogenesis of the common eye diseases. In this chapter, the association of the gene variants in the promoter region with the common eye diseases was summarized, with the focus on myopia, cataract, glaucoma, and age-related macular generation. In addition, the contribution of the promoter variants to the pathogenesis of these complex common eye diseases would also be discussed.
Article
Background: The aim of this study was to investigate the prevalence of myopia in key (university-oriented) and non-key elementary schools in China using a traditional and a new criterion for myopia diagnosis in an epidemiological study. Methods: This school-based, cross-sectional study examined students from four key schools and seven non-key schools. Non-cycloplegic autorefraction and visual acuity (VA) were performed on each student. Myopia was defined as a spherical equivalent (SE) refractive error not better than -1.00 D. A questionnaire was also administered. Results: Of the 13,220 students examined, 6,546 (49.5 per cent) had myopia using the criterion of SE not better than -1.00 D. However, 2,246 (34.3 per cent) of these myopes had VA ≥ 0 logMAR in both eyes, indicating they were not functioning as myopes. Thus, a second myopia criterion was adopted: SE refractive error not better than -1.00 D + uncorrected VA ≥ 0 logMAR in at least one eye. By this definition, only 32.5 per cent of the overall sample had myopia. Students in key schools had a higher prevalence of myopia than those in non-key schools (53.8 per cent versus 44.7 per cent) by the initial criterion. By the new criterion, the prevalence of myopia was 41.2 per cent versus 22.7 per cent. Myopia was equal in grade 1 of both school types, but accelerated faster in key schools, where there was a much higher prevalence of myopia by fourth grade, and continued up to 79.2 per cent prevalence by sixth grade based on SE refractive error not better than -1.00 D. Conclusion: Students in more competitive university-oriented elementary schools developed myopia much faster than those in regular schools, although they started with the same level of myopia. Since one-third of the 'myopes' had VA ≥ 0 logMAR in both eyes, they would not be prescribed a correction, or be clinically treated as myopes. A new criterion of SE refractive error not better than -1.00 D + uncorrected VA ≥ 0 logMAR in at least one eye was tested. This criterion is more clinically appropriate and could be used in future epidemiological studies.
Article
Full-text available
Transforming growth factor-beta2 (TGF-beta2), basic fibroblast growth factor (bFGF), and fibromodulin (FMOD) are important extracellular matrix components of the sclera and have been shown to be associated with the development of high myopia. Our aim was to examine the association between myopia and the polymorphisms within TGF-beta2, bFGF, and FMOD. The study group comprised of patients (n=195; age range: 17-24 years) with a spherical equivalent of -6.5 diopters (D) or a more negative refractive error. The control group comprised of individuals (n=94; age range: 17-25 years) with a spherical equivalent ranging from -0.5 D to +1.0 D. The subjects with astigmatism over -0.75 D were excluded from the study. High resolution melting (HRM) genotyping and restriction fragment length polymorphism (RFLP) genotyping were used to detect single nucleotide polymorphisms (SNPs). The polymorphisms detected were TGF-beta2 (rs7550232 and rs991967), bFGF (rs308395 and rs41348645), and FMOD (rs7543418). Moreover, a stepwise logistic regression procedure was used to detect which of the significant SNPs contributed to the main effects of myopia development. There were significant differences in the frequency of the A allele and A/A genotype in TGF-beta2 (rs7550232; p=0.0178 and 0.03, respectively). Moreover, the haplotype distribution of haplotype 2 (Ht2; A/A) of TGF-beta2 differed significantly between the two groups (p=0.014). The results of the stepwise logistic regression procedure revealed that TGF-beta2 (rs7550232) contributed significantly to the development of high myopia. TGF-beta2 is an important structure of sclera and might contribute to the formation of myopia. TGF-beta2 (rs7550232) polymorphisms, A allele and A/A genotype, had a protective role against the development of high myopia.
Article
Full-text available
The lumican gene (LUM) encodes a major extracellular component of the fibrous mammalian sclera. Alteration in the expression levels of extracellular matrix components may influence scleral shape, which in turn could affect visual acuity. Single-nucleotide polymorphisms (SNPs) in the LUM gene were determined in an investigation of whether LUM gene polymorphisms correlate with high myopia. Sequences spanning all three exons, intron-exon boundaries, and promoter regions were determined in 50 normal individuals. Five SNPs were identified, one of which was found to be a newly identified polymorphism. Genomic DNA was prepared from peripheral blood obtained from 201 patients with high myopia and 86 control subjects. Genotypes of the SNPs -1554 T/C (rs3759223), -628 A/-(rs17018757), -59 CC/-(rs3832846), c.601 T/C (rs17853500), and the novel SNP c.1567 C>T were determined by polymerase chain reaction. Of the five SNPs, one showed a significant difference between patients and control subjects (c.1567 C>T, P = 0.0016). Haplotype analysis revealed a significantly higher presence of polymorphisms in patients with myopia (P < 0.0001). Moreover, the c.1567 T polymorphism was determined to have lower reporter gene activity than that of c.1567 C. These observations suggest that LUM gene polymorphisms contribute to the development of high myopia.
Article
Full-text available
Hepatocyte growth factor (HGF) and hepatocyte growth factor receptor (C-MET) genes have previously been reported to be associated with myopia in Asian family-based and case-control association studies, respectively. We examined whether these genes were associated with myopia in a Caucasian family dataset biased towards high myopia. Participating families had at least one offspring with high myopia (< or = -5.00 diopters [D]). Genotyping was performed with tagging single nucleotide polymorphisms (SNPs) for each candidate gene using Taqman allelic discrimination assays. The data were analyzed with two family-based association methods, the pedigree disequilibrium test (PDT) and the association in the presence of linkage (APL) test. Analyses compared 1) high myopia (<-5.00 D), 2) mild to moderate myopia (-0.50 to -5.00 D), 3) any myopia (<-0.50 D) and 4) extreme high myopia (< or =-10.00 D) versus emmetropia using refractive error as either sphere (SPH) or spherical equivalent (SE=sphere + [cylinder/2]). Bonferroni correction was applied to adjust for multiple testing leading to significance levels of 0.0125 for HGF and 0.008 for C-MET. Two and three-marker sliding window haplotype association tests using APL were also performed for HGF markers. Significance levels for haplotype association testing were set at 0.01 for the global tests, and 0.007 for the three marker haplotype specific tests and 0.0125 for the two marker haplotype specific tests. A total of 146 multiplex families consisting of 649 Caucasian subjects were included. The HGF SNP, rs3735520 (APL p=0.002768 for SPH and 0.005609 for SE), and the haplotypes, rs2286194-rs3735520-rs17501108 (APL p=0.007403 for SPH and 0.062685 for SE) and rs12536657-rs2286194 (APL p=0.004219 for SPH and 0.00518 for SE), showed significant association with mild to moderate myopia versus emmetropia. A promising association between extreme high myopia and the HGF SNP, rs2286194, was also found (APL p=0.005763 for SPH and 0.004103 for SE). No evidence of association was found in the SNPs tested for C-MET. This study supports a strong association between the mild to moderate myopia group and the HGF SNP rs3735520 and the HGF haplotypes rs2286194-rs3735520-rs17501108 and rs12536657-rs2286194, and a moderate association of the extreme high myopia with rs2286194. C-MET polymorphism statistical associations with myopia in an Asian study were not replicated in our Caucasian cohort. HGF may be a potential myopia candidate gene for further investigation.
Article
Full-text available
Myopia (shortsightedness) is one of the most common ocular conditions worldwide and results in blurred distance vision. It is a complex trait influenced by both genetic and environmental factors. We have previously reported linkage of myopia to a 13.01 cM region of chromosome 2q37 in three large multigenerational Australian families that initially overlapped with the known myopia locus, MYP12. The purpose of this study was to perform fine mapping of this region and identify single nucleotide polymorphisms (SNPs) associated with myopia. Fine mapping linkage analysis was performed on three multigenerational families with common myopia to refine the previously mapped critical interval. SNPs in the region were also genotyped to assess for association with myopia using an independent case-control cohort. The disease interval was refined to a 1.83 cM region that is adjacent to rather than overlapping with the MYP12 locus. Subsequent sequencing of all known and hypothetical genes as well as an association study using an independent myopia case-control cohort showed suggestive but not statistically significant association to two intronic SNPs. We have identified a novel locus for common myopia on chromosome 2q37.
Article
Full-text available
Several nonsyndromic high-grade myopia loci have been mapped primarily by microsatellite markers and a limited number of pedigrees. In this study, whole-genome linkage scans were performed for high-grade myopia, using single nucleotide polymorphisms (SNPs) in 254 families from five independent sites. Genomic DNA samples from 1411 subjects were genotyped (Linkage Panel IVb; Illumina, San Diego, CA). Linkage analyses were performed on 1201 samples from 10 Asian, 12 African-American, and 221 Caucasian families, screening for 5744 SNPs after quality-control exclusions. Two disease states defined by sphere (SPH) and spherical equivalence (SE; sphere+cylinder/2) were analyzed. Parametric and nonparametric two-point and multipoint linkage analyses were performed using the FASTLINK, HOMOG, and MERLIN programs. Multiple stratified datasets were examined, including overall, center-specific, and race-specific. Linkage regions were declared suggestive if they had a peak LOD score >or= 1.5. The MYP1, MYP3, MYP6, MYP11, MYP12, and MYP14 loci were replicated. The novel region q34.11 on chromosome 9 (max NPL= 2.07 at rs913275) was identified. Chromosome 12, region q21.2-24.12 (36.59 cM, MYP3 locus) showed significant linkage (peak HLOD = 3.48) at rs337663 in the overall dataset by SPH and was detected by the Duke, Asian, and Caucasian subsets as well. Potential shared interval was race dependent-a 9.4-cM region (rs163016-rs1520724) driven by the Asian subset and a 13.43-cM region (rs163016-rs1520724) driven by the Caucasian subset. The present study is the largest linkage scan to date for familial high-grade myopia. The outcomes will facilitate the identification of genes implicated in myopic refractive error development and ocular growth.
Article
purpose. To investigate the coding exons of transforming growth factor (TGF)-β–induced factor (TGIF) for mutations in Chinese patients with high myopia. methods. Seventy-one individuals with high myopia of −6.00 D or less and 105 control subjects were screened by DNA sequencing for sequence alterations. Univariate analysis and logistic regression were performed to identify single-nucleotide polymorphisms (SNPs) and their interactions in TGIF that may be associated with myopia. results. Six SNPs showed a significant difference (P < 0.05) between patient and control subject in univariate analysis. Four of them cause codon changes: G223R, G231S, P241T, and A262G. Among all the SNPs that entered multivariate analysis, only 657(T→G) showed statistical significance in the logistic regression model (odds ratio 0.133; 95% confidence interval 0.037–0.488; P = 0.002). conclusions. TGIF is a probable candidate gene for high myopia. Further studies are needed to identify the underlying mechanism.
Article
TESS, Transcription Element Search Software, is a web-based software tool for locating possible transcription factor binding sites in DNA sequence and for browsing the TRANSFAC database. It provides functionality beyond that of the TRANSFAC at les and web site. TESS allows the user to search sequence for possible bindings sites using either cis-element strings or weight matrix models. 1
Article
The original publication is available at www.springerlink.com.
Article
To assess whether genetic variation in cMET is associated with refractive error or change in refractive error over time. Cohort study. Discovery set (Set 1: N = 579 children; 403 cases, 176 controls). Confirmatory set (Set 2: N = 547 children; 338 cases, 209 controls). Children in the discovery set were genotyped for a panel of genetic markers within cMET. Markers that were found to be significantly associated with the presence of refractive error or more rapid change in refractive error were then genotyped in the confirmatory set. Presence or absence of myopia and the rate of change in refractive error over a 3-year follow-up period. Carriage of the variant cMET +110703 A allele was found to associate with increased susceptibility to myopia. The variant was also found to associate with a faster rate of change in refractive error in both the discovery set and the confirmatory cohort regardless of the initial refractory ability (School 1; chi(2) for trend P = 0.014) (Schools 2 and 3; chi(2) for trend = 5.42, P = 0.020) (combined N = 1126, overall chi(2) for trend = 10.90, P = 9.6 x 10(-4)). Carriage of the variant allele was also found to be significantly overrepresented in children within the fastest changing quartile (Q4: mean change of -3.01 D over 3 years) compared with the slowest (Q1: mean change of -0.28 D over 3 years) (P(Set1) = 0.004, P(Set2) = 0.02, Combined N = 559, P = 3.0 x 10(-4)). Our data implicate the involvement of cMET in the pathogenesis of myopia in general, as well as more rapid progression in refractive error regardless of the initial refractory ability. These results underline the importance of eye growth genes in the development of common myopia.
Article
To investigate the relation between myopia and variations in three genes coding for matrix metalloproteinases, enzymes that degrade matrix proteins and modulate scleral extensibility. Three hundred sixty-six men and women, from Sheffield, United Kingdom, were genotyped for the 1G/2G polymorphism in the MMP-1 gene, the 5A/6A polymorphism in the MMP-3 gene and the Arg-->Gln polymorphism in exon 6 of the MMP-9 gene and assessed for refractive error. Risk of myopia was increased in people homozygous for the 5A allele of the MMP-3 gene (odds ratio [OR], 3.1; 95% confidence interval [CI], 1.1-9.0) compared to those who were homozygous for the 6A allele, and in people homozygous for the Gln allele in exon 6 of the MMP-9 gene (OR, 2.8; 95% CI, 1.1-7.0) compared to those who were homozygous for the Arg allele. People who were homozygous for the 2G allele of the MMP-1 gene had an odds ratio for myopia of 2.3 (95% CI, 0.9-6.1), compared with those who were homozygous for the 1G allele, although this relation did not reach statistical significance. Risk of myopia increased progressively with the dose of these three alleles, showing a greater than 10-fold difference across the range. The results suggest that common variations in three of the genes that control breakdown of matrix proteins in the sclera may contribute to the development of simple myopia.