Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Nov 21.
Published in final edited form as: Birth Defects Res A Clin Mol Teratol. 2010 Feb;88(2):84–93. doi: 10.1002/bdra.20639

Testing reported associations of genetic risk factors for oral clefts in a large Irish study population

Tonia C Carter 1, Anne M Molloy 2, Faith Pangilinan 3, James F Troendle 1, Peadar N Kirke 4, Mary R Conley 1, David J A Orr 5, Michael Earley 5, Eamon McKiernan 5, Ena C Lynn 4, Anne Doyle 4, John M Scott 2, Lawrence C Brody 3, James L Mills 1
PMCID: PMC3503531  NIHMSID: NIHMS418550  PMID: 19937600

Abstract

BACKGROUND

Suggestive, but not conclusive, studies implicate many genetic variants in oral cleft etiology. We used a large, ethnically homogenous study population to test whether reported associations between nonsyndromic oral clefts and 12 genes (CLPTM1, CRISPLD2, FGFR2, GABRB3, GLI2, IRF6, PTCH1, RARA, RYK, SATB2, SUMO1, TGFA) could be confirmed.

METHODS

Thirty-one single nucleotide polymorphisms (SNPs) in exons, splice sites, and conserved non-coding regions were studied in 509 patients with cleft lip with or without cleft palate (CLP), 383 with cleft palate only (CP), 838 mothers and 719 fathers of patients with oral clefts, and 902 controls from Ireland. Case-control and family-based statistical tests were performed using isolated oral clefts for the main analyses.

RESULTS

In case-control comparisons, the minor allele of PTCH1 A562A (rs2066836) was associated with reduced odds of CLP (OR: 0.29, 95% CI: 0.13–0.64 for homozygotes) whereas the minor allele of PTCH1 L1315P (rs357564) was associated with increased odds of CLP (OR: 1.36, 95% CI: 1.07–1.74 for heterozygotes and OR: 1.56, 95% CI: 1.09–2.24 for homozygotes). The minor allele of one SUMO1 SNP, rs3769817 located in intron 2, was associated with increased odds of CP (OR: 1.45, 95% CI: 1.06–1.99 for heterozygotes). Transmission disequilibrium was observed for the minor allele of TGFA V159V (rs2166975) which was over-transmitted to CP cases (P=0.041).

CONCLUSIONS

For 10 of the 12 genes, this is the largest candidate gene study of nonsyndromic oral clefts to date. The findings provide further evidence that PTCH1, SUMO1, and TGFA contribute to nonsyndromic oral clefts.

Keywords: cleft lip, cleft palate, congenital abnormalities

INTRODUCTION

Oral clefts are among the most common birth defects. Population-based estimates of the prevalence of nonsyndromic cleft lip with or without cleft palate (CLP) and cleft palate only (CP) are 4–13/10,000 births and 2–5/10,000 births, respectively, with worldwide estimates varying by ethnicity and geographic region (Hashmi et al., 2005). Research suggests that multiple genes are involved in oral clefts etiology (Schliekelman and Slatkin, 2002) and that various signaling pathways are important in lip and palate development (Marazita and Mooney, 2004), but the specific genes in these pathways that play a role in causing oral clefts in humans remain unknown. There have been consistent findings of an association with nonsyndromic oral clefts for only one gene, interferon regulatory factor 6 (IRF6), however this gene does not account for the majority of the genetic contribution to nonsyndromic oral clefts and more genes remain to be identified. Studies using a variety of approaches have produced inconclusive or conflicting results, possibly because of inadequate power and population diversity. To address these limitations we have assembled a large, ethnically homogenous study population in Ireland to examine genetic risk factors for oral clefts. The intent of the current study was to follow-up on previously published reports of genes examined for an association with oral clefts in humans and to confirm whether an association was present in an independent study population. Therefore, we performed case-control and trio-based analyses to examine whether selected SNPs in genes that were previously suggested to be risk factors for oral clefts were associated with nonsyndromic CLP and CP.

MATERIALS AND METHODS

Cases

Study subject enrollment and data collection have been described in detail previously (Mills et al., 2008). Subjects with CLP or CP were identified by one of two methods. First, patients at the Dublin Cleft Centre which treats most patients with oral clefts (80%) in Ireland, were recruited. The attending surgeons examined all patients, reported a diagnosis of CLP or CP, and confirmed the presence or absence of other birth defects. Second, patients were identified through membership in the Cleft Lip and Palate Association of Ireland, a patient support group. Medical record data on the diagnosis of oral clefts were available for 87% of all patients. Recruited case-parent trios gave either blood or buccal samples. Study participants, or in the case of minors, their parents, provided written informed consent. The Research Ethics Committees of the Health Research Board of Ireland, the participating hospitals, and the Institutional Review Board of the National Human Genome Research Institute gave ethical approval for the study.

Controls

Blood samples had previously been collected from 56,049 pregnant women who had attended any of the three main maternity hospitals in Dublin between 1986 and 1990. These hospitals deliver approximately 90% of Dublin births and about one-third of all Irish births occur in Dublin. Controls were a random sample (N=926) from this group. Women were not eligible to be controls if they had delivered a child with a known birth defect or if they had previously had a child with a neural tube defect. All the genes selected for this study are located on autosomes and, among the controls, allele frequencies for genes on the autosomes reflect those of the underlying Irish population (Parle-McDermott et al., 2003; Kirke et al., 2004).

Genes and SNPs

We selected 11 candidate genes that have been implicated in previous studies that produced suggestive, but not conclusive, evidence that they were associated with nonsyndromic oral clefts. We attempted to focus on genes that did not have a large body of published literature concerning their association with oral clefts as well as genes with previous controversial findings related to oral clefts. Genes were chosen if there were a) any reports of mutations or chromosomal breakage within the gene in patients with oral clefts; b) expression or gene-knockout data from animal studies linking the gene to oral clefts; or c) previous conflicting reports concerning their associations with oral clefts. Also, either the gene or the chromosomal region in which it is located had to have been examined in at least one human study. The selected genes were cleft lip and palate associated transmembrane protein 1 (CLPTM1), cysteine-rich secretory protein LCCL domain containing 2 (CRISPLD2), fibroblast growth factor receptor 2 (FGFR2), gamma-aminobutyric acid A receptor, beta 3 (GABRB3), GLI-Kruppel family member GLI2 (GLI2), patched homolog 1 (Drosophila) (PTCH1), retinoic acid receptor alpha (RARA), RYK receptor-like tyrosine kinase (RYK), SATB homeobox 2 (SATB2), SMT3 suppressor of mif two 3 homolog 1 (Saccharomyces cerevisiae) (SUMO1), and transforming growth factor alpha (TGFA). In addition to the 11 genes, IRF6 was also included in this study to investigate gene-gene interactions and because the gene has not been examined in the Irish population.

SNPs were selected for study based on published reports and/or their locations in the genes. Our aim was to target SNPs that might affect function rather than to SNP-tag or examine all the SNPs within these genes. Thirty-one SNPs were selected located in exons, splice sites, or conserved non-coding regions, locations in which nucleotide changes could have functional consequences. SNPs in conserved regions were identified from multi-species alignments of DNA sequences using the University of California at Santa Cruz Genome Browser (http://genome.ucsc.edu/).

Genotyping

Genomic DNA was extracted from blood or buccal samples using a QIAamp DNA Blood Mini Kit (Qiagen, Sussex, UK). KBiosciences (Herts, United Kingdom) performed SNP genotyping using a competitive allele-specific PCR genotyping system. Quality control measures included the use of blank wells, the repetition of one human DNA sample across multiple plates, and repeat genotyping for approximately 4% of study subjects. Fifteen of the 31 SNPs were called successfully >95% of the time, 14 were called successfully 93–95% of the time, and the remaining two SNPs, CRISPLD2 S105G (rs12051468) and SATB2 Intron 9 T>G (rs4675475), were called successfully 91% and 90% of the time, respectively. Tests of Hardy-Weinberg Equilibrium (HWE) were carried out for all SNPs (independently for cases, case-mothers, case-fathers, and controls). The level of agreement for repeat genotyping was >96% for all SNPs.

Exclusions

Of the 1,088 cases recruited, 196 were excluded from analyses for the following reasons (number of cases in parenthesis): maternal diabetes (5), maternal epilepsy (12), maternal teratogenic drug exposure (27), chromosomal abnormalities (9), known syndromes (49), type of cleft could not be classified (3), possible non-Mendelian inheritance (8), having two or more SNPs with discordant results upon repeat genotyping (1), and samples could not be genotyped (82). Among the 926 controls, two that had two or more SNPs with discordant results upon repeat genotyping and 22 that had samples that could not be genotyped were excluded from the analysis.

Statistical analysis

Patients with isolated oral clefts (absence of other non-cleft defects) were used for the main analysis and were categorized into isolated CLP and isolated CP case groups. We also divided the CLP case group into cleft lip only and cleft lip with cleft palate, and performed separate analyses for these subgroups, because epidemiological and molecular data indicate that these could be distinct conditions (Harville et al., 2005; Rahimov et al., 2008; Genisca et al., 2009). Analyses were repeated adding cases with multiple defects (presence of other non-cleft defects) to determine if they influenced the results. Tests of HWE were performed using a chi-squared test for genotype proportions. Allele frequencies were compared between cases and controls using a chi-squared test. Case-control comparisons of genotype frequency were performed using logistic regression to estimate odds ratios (OR) and 95% confidence intervals (CI) for each SNP. Effect estimates were reported separately for being heterozygous and homozygous for the minor allele, with homozygosity for the major allele as the reference category. We also used the multiplicative disease model in logistic regression analyses to generate P-values for the overall effect of minor allele copy number on the association. The transmission disequilibrium test (TDT) was used to assess allele transmission distortion among case-parent trios. Adjustment for multiple comparisons was done using permutation (N=99,999), separately by gene, because the associations between oral clefts and each gene were considered to be distinct a priori hypotheses derived from the literature. Multivariate permuting of the trios for the TDT was performed by treating the test as a one-sample test, and permuting the allele of interest. Permutations of cases and controls were independent of permutations of trios, and the results combined by Bonferroni adjustment so that the resulting adjusted P-values accounted for all comparisons on that gene while controlling the probability of any false-positives (family-wise error rate). Gene-gene interactions were examined between IRF6 SNPs and those in genes associated with oral clefts in individual analyses. A disease inheritance model for each SNP was estimated from the OR and 95% CI for heterozygotes and homozygotes in individual analyses and was then used in logistic regression analysis to test gene-gene interaction.

Linkage disequilibrium

Measures of linkage disequilibrium (LD) were estimated using Haploview (http://www.broad.mit.edu/mpg/haploview/) (Barrett et al., 2005). LD measures were based on genotypes of control samples (N=902) from this study or publicly available data (www.hapmap.org) as indicated in Results and discussion.

RESULTS

There were 892 cases (765 of them isolated), 838 case-mothers, 719 case-fathers, and 902 controls included in the analysis (Table 1). The birth year of the cases ranged from 1931–2004. The proportion of case-mothers who reported smoking and alcohol use during the affected pregnancy was 38% and 55%, respectively. There was a positive family history of oral clefts among first-degree relatives for 7.5% (67/892) of cases.

Table 1.

Irish case families.

Oral cleft case group Family member Isolated defect Multiple defects Total
Cleft lip with or without cleft palate Case 472 37 509
Mother 442 32 474
Father 370 28 398
Cleft lip only Case 147 5 152
Mother 140 5 145
Father 123 3 126
Cleft lip with cleft palate Case 325 32 357
Mother 302 27 329
Father 247 25 272
Cleft palate only Case 293 90 383
Mother 275 89 364
Father 243 78 321

The selected SNPs and their minor allele frequencies among controls and patients with isolated CLP or CP are presented in Table 2. Genotype distributions among study groups were in HWE with one exception: PTCH1 N555N (rs1805155) among controls (P=9×10−14). This could have been due to genotyping error; therefore, we did not report any results for this SNP. Quality control results indicated that genotyping was performed satisfactorily on the other two PTCH1 SNPs.

Table 2.

Allele frequencies of single nucleotide polymorphisms (SNPs) in controls and patients with isolated cleft lip with or with cleft palate (CLP) or isolated cleft palate only (CP).

Gene SNP Description Allelesa Minor allele frequency
Controls (N=902) CLP (N=472) CP (N=293)
CLPTM1 rs3786505 P309P A/G 0.316 0.288 0.293
rs204468 G331G C/T 0.366 0.345 0.350
rs7257610 Intron 7, splice site acceptor G/A 0.036 0.029 0.042
CRISPLD2 rs12051468 S105G A/G 0.460 0.463 0.423
rs8061351 P157P T/C 0.239 0.233 0.285b
rs721005 T322S C/G 0.393 0.402 0.382
rs1546124 5′ untranslated region C/G 0.329 0.334 0.358
rs16974880 3′ untranslated region T/G 0.338 0.357 0.314
rs4783099 3′ untranslated region C/T 0.394 0.428 0.377
FGFR2 rs1047100 V232V G/A 0.197 0.211 0.185
GABRB3 rs2017247 3′ untranslated region G/A 0.256 0.271 0.257
rs11637141 3′ untranslated region C/T 0.229 0.236 0.233
GLI2 rs12711538 D1306N A/G 0.280 0.287 0.319
IRF6 rs2235371 V274I C/T 0.014 0.004b 0.009
rs2013162 S153S C/A 0.356 0.318 0.376
rs7552506 Intron 3, splice site acceptor G/C 0.339 0.323 0.377
rs2235377 Intron 1, splice site acceptor T/C 0.013 0.006 0.007
PTCH1 rs2066836 A562A C/T 0.211 0.165c 0.219
rs357564 P1315L G/A 0.323 0.379c 0.350
RARA rs2229773 3′ untranslated region C/T 0.031 0.032 0.040
RYK rs1131262 Stop *98* G/A 0.115 0.111 0.117
rs4470517 Intron 13 T/C 0.356 0.383 0.349
SATB2 rs4673313 Intron 8 T/C 0.103 0.121 0.115
rs17199393 Intron 3 G/A 0.038 0.052 0.047
rs16831366 Intron 6 G/T 0.061 0.084b 0.086b
rs1446636 Intron 10 G/T 0.120 0.147 0.133
rs4675475 Intron 9 T/G 0.224 0.254 0.242
SUMO1 rs3769817 Intron 2 G/A 0.118 0.143 0.158b
rs12470401 Intron 1 G/A 0.116 0.138 0.153b
TGFA rs2166975 V159V G/A 0.254 0.245 0.291
a

Major allele is listed first

b

Significant difference in minor allele frequency compared with controls; P<0.05, uncorrected

c

Significant difference in minor allele frequency compared with controls; P<0.01, uncorrected

Case-control comparisons of genotype frequency indicated statistically significant associations between PTCH1 SNPs and CLP (Table 3). The PTCH1 A562A (rs2066836) minor allele was associated with reduced odds of CLP among homozygotes (OR: 0.29, 95% CI: 0.13–0.64; corrected P=0.038) which is equivalent to the major allele increasing the odds of CLP (OR: 3.51, 95% CI: 1.56–7.87). The PTCH1 P1315L (rs357564) minor allele was associated with increased odds of CLP (OR: 1.36, 95% CI: 1.07–1.74 for heterozygotes and OR: 1.56, 95% CI: 1.09–2.24 for homozygotes; corrected P=0.023). There was little linkage disequilibrium between these two PTCH1 SNPs (r2 = 0.10). The minor allele of the SUMO1 intron 2 G>A SNP (rs3769817) was associated with increased odds of CP among heterozygotes (OR: 1.45, 95% CI: 1.06–1.99; corrected P=0.038) but among homozygotes the confidence interval included the null (OR: 1.62, 95% CI: 0.65–4.03) (Table 4). The minor allele of the other SUMO1 SNP tested (intron 1 G>A, rs12470401) had a similar OR estimate for CP that was of borderline significance (OR = 1.44, 95% CI = 1.05–1.97 for heterozygotes; corrected P=0.056); it was in strong LD (r2= 0.98) with rs3769817. Based on the TDT, there was no transmission distortion of PTCH1 SNPs to CLP cases or of SUMO1 SNPs to CP cases (Table 4). TDT results indicated that the TGFA V159V (rs2166975) minor allele was over-transmitted to CP cases (corrected P=0.041) (Table 4). SNPs in the other nine genes were not associated with CLP or CP. When the CLP case group was divided into isolated cleft lip only and isolated cleft lip with cleft palate, the only statistically significant finding was for the association of the PTCH1 P1315L (rs357564) minor allele with cleft lip and palate in case-control comparisons (OR = 1.67, 95% CI = 1.13–2.49 for homozygotes; corrected P=0.040) (Table 5). Repeating the analyses after adding patients with multiple birth defects to the case groups produced no important changes in the results. The numbers of cases and controls in Table 1 are different to those in Table 3 due to the numbers of subjects with genotype data available. IRF6 S153S (rs2013162) was used in tests of gene-gene interaction because it was the only IRF6 SNP to show an association with CLP, based on its uncorrected P-value (P=0.046). There was no evidence of gene-gene interaction between this SNP and those in PTCH1, SUMO1, and TGFA.

Table 3.

Genotype distributions for single nucleotide polymorphisms (SNPs) in 12 candidate genes and tests of association in controls and patients with isolated cleft lip with or with cleft palate (CLP) or isolated cleft palate only (CP).

SNPs Control Genotypesa CLP Genotypesa CLP OR (95% CI)b Corrected P-Valued CP Genotypesa CP OR (95% CI)b Corrected P-valued
Heterozygotes Homozygotesc Heterozygotes Homozygotesc
CLPTM1
 rs3786505 414/375/91 235/172/44 0.81 (0.64, 1.03) 0.85 (0.57, 1.26) 0.523 146/108/29 0.82 (0.61, 1.09) 0.90 (0.57, 1.43) 1.000
 rs204468 360/401/123 196/203/55 0.93 (0.73, 1.19) 0.82 (0.57, 1.18) 0.981 118/128/34 0.97 (0.73, 1.30) 0.84 (0.55, 1.30) 1.000
 rs7257610 819/64/0 423/26/0 0.79 (0.49, 1.26) - 1.000 261/22/1 1.08 (0.65, 1.79) - 1.000
CRISPLD2
 rs12051468 234/406/170 125/209/93 0.96 (0.73, 1.27) 1.02 (0.73, 1.43) 1.000 87/142/45 0.94 (0.69, 1.29) 0.71 (0.47, 1.07) 0.850
 rs8061351 492/328/42 270/156/28 0.87 (0.68, 1.10) 1.22 (0.74, 2.00) 1.000 139/125/18 1.35 (1.02, 1.78) 1.52 (0.85, 2.72) 0.217
 rs721005 324/393/140 161/215/73 1.10 (0.86, 1.42) 1.05 (0.75, 1.48) 1.000 108/134/41 1.02 (0.76, 1.37) 0.88 (0.58, 1.33) 1.000
 rs1546124 385/369/95 193/201/47 1.09 (0.85, 1.39) 0.99 (0.67, 1.46) 1.000 110/137/31 1.30 (0.97, 1.73) 1.14 (0.72, 1.81) 1.000
 rs16974880 382/398/97 180/215/52 1.15 (0.90, 1.46) 1.14 (0.78, 1.67) 1.000 133/121/28 0.87 (0.66, 1.16) 0.83 (0.52, 1.32) 1.000
 rs4783099 321/425/134 143/224/79 1.18 (0.92, 1.53) 1.32 (0.94, 1.86) 0.840 110/134/40 0.92 (0.69, 1.23) 0.87 (0.58, 1.32) 1.000
FGFR2
 rs1047100 575/298/29 287/150/22 1.01 (0.79, 1.29) 1.52 (0.86, 2.69) 0.745 193/85/11 0.85 (0.64, 1.14) 1.13 (0.55, 2.31) 1.000
GABRB3
 rs2017247 496/329/63 244/168/38 1.04 (0.82, 1.32) 1.23 (0.80, 1.89) 1.000 156/110/18 1.06 (0.80, 1.41) 0.91 (0.52, 1.58) 1.000
 rs11637141 527/323/43 269/163/27 0.99 (0.78, 1.26) 1.23 (0.74, 2.04) 1.000 174/94/20 0.88 (0.66, 1.17) 1.41 (0.81, 2.46) 1.000
GLI2
 rs12711538 448/336/71 219/177/35 1.08 (0.85, 1.38) 1.01 (0.65, 1.56) 1.000 131/122/29 1.24 (0.93, 1.65) 1.40 (0.87, 2.24) 0.176
IRF6
 rs2235371 869/25/0 456/4/0 0.31 (0.11, 0.88) - 0.103 281/3/1 0.37 (0.11, 1.24) - 1.000
 rs2013162 350/409/102 204/202/41 0.85 (0.67, 1.08) 0.69 (0.46, 1.03) 0.275 109/129/40 1.01 (0.76, 1.36) 1.26 (0.82, 1.93) 1.000
 rs7552506 379/406/95 201/202/43 0.94 (0.74, 1.19) 0.85 (0.57, 1.27) 1.000 110/129/41 1.10 (0.82, 1.46) 1.49 (0.97, 2.27) 0.480
 rs2235377 847/23/0 440/5/0 0.42 (0.16, 1.11) - 0.338 271/4/0 0.54 (0.19, 1.59) - 1.000
PTCH1
 rs2066836 559/284/45 305/133/7 0.86 (0.67, 1.10) 0.29 (0.13, 0.64) 0.038 172/101/12 1.16 (0.87, 1.54) 0.87 (0.45, 1.68) 1.000
 rs357564 415/382/98 174/218/64 1.36 (1.07, 1.74) 1.56 (1.09, 2.24) 0.023 116/137/31 1.28 (0.97, 1.70) 1.13 (0.72, 1.78) 0.947
RARA
 rs2229773 825/53/1 429/29/0 1.05 (0.66, 1.68) - 1.000 262/23/0 1.37 (0.82, 2.27) - 0.637
RYK
 rs1131262 699/179/13 354/99/1 1.09 (0.83, 1.44) 0.15 (0.02, 1.17) 1.000 228/51/8 0.87 (0.62, 1.23) 1.89 (0.77, 4.61) 1.000
 rs4470517 354/376/111 178/198/73 1.05 (0.82, 1.34) 1.31 (0.93, 1.85) 0.597 119/120/36 0.95 (0.71, 1.27) 0.97 (0.63, 1.48) 1.000
SATB2
 rs4673313 694/147/15 338/94/6 1.31 (0.98, 1.75) 0.82 (0.32, 2.14) 1.000 216/60/2 1.31 (0.94, 1.84) 0.43 (0.10, 1.89) 1.000
 rs17199393 797/65/0 409/39/4 1.17 (0.77, 1.77) - 0.733 252/26/0 1.27 (0.79, 2.04) - 1.000
 rs16831366 762/96/5 367/68/3 1.47 (1.05, 2.06) 1.25 (0.30, 5.24) 0.309 232/46/1 1.57 (1.08, 2.30) 0.66 (0.08, 5.65) 0.408
 rs1446636 673/174/17 327/110/11 1.30 (0.99, 1.71) 1.33 (0.62, 2.88) 0.536 215/64/6 1.15 (0.83, 1.59) 1.11 (0.43, 2.84) 1.000
 rs4675475 530/291/50 223/158/24 1.29 (1.01, 1.66) 1.14 (0.68, 1.90) 0.877 152/87/19 1.04 (0.77, 1.41) 1.33 (0.76, 2.32) 1.000
SUMO1
 rs3769817 684/177/15 332/118/6 1.37 (1.05, 1.79) 0.82 (0.32, 2.14) 0.159 197/74/7 1.45 (1.06, 1.99) 1.62 (0.65, 4.03) 0.038
 rs12470401 694/117/14 337/114/6 1.33 (1.01, 1.74) 0.88 (0.34, 2.32) 0.227 199/73/6 1.44 (1.05, 1.97) 1.50 (0.57, 3.94) 0.056
TGFA
 rs2166975 472/338/49 253/161/28 0.89 (0.70, 1.13) 1.07 (0.65, 1.74) 1.000 138/107/25 1.08 (0.81, 1.45) 1.75 (1.04, 2.93) 0.165
a

Genotype values are numbers of individuals with homozygous major allele/heterozygous/homozygous minor allele

b

OR – odds ratio; CI – confidence interval; homozygous major allele was reference category

c

Homozygous for minor allele; homozygote odds ratios were not obtained when there were no cases and/or controls that were homozygous for the minor allele

d

Corrected P-value after adjustment for multiple comparisons; testing for the overall effect of minor allele copy number on the outcome based on the multiplicative disease model.

Table 4.

Test for distortion of allele transmission to patients that have isolated cleft lip with or with cleft palate or isolated cleft palate only.

SNPsa Cleft lip with or without cleft palate Cleft palate only
Nb T/NTc RR (95% CI)d Nb T/NTc RR (95% CI)d
CLPTM1
 rs3786505 216 128/148 0.87 (0.68, 1.10) 149 93/97 0.96 (0.72, 1.27)
 rs204468 229 137/160 0.86 (0.68, 1.08) 156 106/100 1.06 (0.81, 1.40)
 rs7257610 35 14/21 0.67 (0.34, 1.31) 30 17/15 1.13 (0.57, 2.27)
CRISPLD2
 rs12051468 224 158/133 1.19 (0.94, 1.50) 157 106/93 1.14 (0.86, 1.51)
 rs8061351 212 118/137 0.86 (0.67, 1.10) 130 86/72 1.19 (0.87, 1.63)
 rs721005 241 167/156 1.07 (0.86, 1.33) 154 103/96 1.07 (0.81, 1.42)
 rs1546124 244 149/169 0.88 (0.71, 1.10) 131 88/81 1.09 (0.80, 1.47)
 rs16974880 212 153/118 1.30 (1.02, 1.65) 136 94/90 1.04 (0.78, 1.39)
 rs4783099 221 159/131 1.21 (0.96, 1.53) 150 108/91 1.19 (0.90, 1.57)
FGFR2
 rs1047100 203 124/125 0.99 (0.77, 1.27) 121 66/74 0.89 (0.64, 1.24)
GABRB3
 rs2017247 193 124/123 1.01 (0.79, 1.29) 148 87/96 0.91 (0.68, 1.21)
 rs11637141 198 117/128 0.91 (0.71, 1.17) 144 88/83 1.06 (0.79, 1.43)
GLI2
 rs12711538 213 132/125 1.06 (0.83, 1.35) 142 93/90 1.03 (0.77, 1.38)
IRF6
 rs2235371 8 3/5 0.60 (0.14, 2.51) 3 1/2 0.50 (0.05, 5.51)
 rs2013162 236 128/171 0.75 (0.60, 0.94) 130 100/75 1.33 (0.99, 1.80)
 rs7552506 238 136/168 0.81 (0.65, 1.02) 135 104/76 1.37 (1.02, 1.84)
 rs2235377 8 2/6 0.33 (0.07, 1.65) 3 2/1 2.00 (0.18, 22.06)
PTCH1
 rs2066836 161 89/100 0.89 (0.67, 1.18) 120 78/66 1.18 (0.85, 1.64)
 rs357564 258 167/167 1.00 (0.81, 1.24) 165 111/106 1.05 (0.80, 1.37)
RARA
 rs2229773 45 26/20 1.30 (0.73, 2.33) 33 21/13 1.62 (0.81, 3.23)
RYK
 rs1131262 135 69/77 0.90 (0.65, 1.24) 78 45/47 0.96 (0.64, 1.44)
 rs4470517 231 157/156 1.01 (0.81, 1.26) 154 93/110 0.85 (0.64, 1.11)
SATB2
 rs4673313 119 68/63 1.08 (0.77, 1.52) 76 41/43 0.95 (0.62, 1.46)
 rs17199393 51 24/29 0.83 (0.48, 1.42) 38 19/21 0.91 (0.49, 1.68)
 rs16831366 81 42/43 0.98 (0.64, 1.49) 60 28/34 0.82 (0.50, 1.36)
 rs1446636 144 86/73 1.18 (0.86, 1.61) 98 57/58 0.98 (0.68, 1.42)
 rs4675475 171 105/99 1.06 (0.81, 1.40) 112 69/71 0.97 (0.70, 1.35)
SUMO1
 rs3769817 133 82/77 1.07 (0.78, 1.45) 102 65/53 1.23 (0.85, 1.76)
 rs12470401 141 84/80 1.05 (0.77, 1.43) 98 67/46 1.46 (1.00, 2.12)
TGFA
 rs2166975 177 111/109 1.01 (0.78, 1.33) 133 102/72 1.42 (1.05, 1.92)e
a

SNPs – single nucleotide polymorphisms

b

N – number of informative trios

c

T/NT – transmission/non-transmission of minor allele

d

RR – risk ratio; CI – confidence interval

e

Significant difference in allele transmission, corrected P<0.05, after adjustment for multiple comparisons

Table 5.

Tests of association between SNPs in 12 candidate genes and subgroups of isolated cleft lip with or without cleft palate: cleft lip only and cleft lip with cleft palate.

SNPsa Cleft Lip Only Cleft Lip with Cleft Palate
Case-Control OR (95% CI)b TDT RR (95% CI)d Case-Control OR (95% CI)b TDT RR (95% CI)d
Heterozygotes Homozygotesc Heterozygotes Homozygotesc
CLPTM1
 rs3786505 0.81 (0.56, 1.19) 0.95 (0.52, 1.73) 0.86 (0.58, 1.29) 0.81 (0.61, 1.06) 0.81 (0.51, 1.28) 0.87 (0.65, 1.16)
 rs204468 0.83 (0.57, 1.21) 0.75 (0.43, 1.33) 0.70 (0.47, 1.04) 0.98 (0.74, 1.29) 0.86 (0.57, 1.30) 0.95 (0.72, 1.26)
 rs7257610 0.68 (0.30, 1.51) - 0.57 (0.17, 1.95) 0.84 (0.49, 1.42) - 0.71 (0.32, 1.61)
CRISPLD2
 rs12051468 0.94 (0.62, 1.43) 0.94 (0.56, 1.58) 0.94 (0.63, 1.40) 0.97 (0.71, 1.33) 1.07 (0.73, 1.56) 1.33 (1.00, 1.77)
 rs8061351 0.88 (0.60, 1.28) 1.53 (0.76, 3.10) 0.80 (0.53, 1.21) 0.86 (0.66, 1.14) 1.07 (0.60, 1.93) 0.90 (0.66, 1.22)
 rs721005 1.09 (0.74, 1.61) 0.98 (0.57, 1.67) 1.06 (0.73, 1.53) 1.10 (0.83, 1.47) 1.08(0.74, 1.60) 1.08 (0.82, 1.41)
 rs1546124 1.09 (0.75, 1.59) 0.98 (0.53, 1.80) 0.86 (0.60, 1.23) 1.08 (0.82, 1.43) 0.99 (0.63, 1.55) 0.89 (0.68, 1.18)
 rs16974880 0.87 (0.60, 1.27) 0.82(0.44, 1.53) 1.14 (0.76, 1.72) 1.31 (0. 99, 1.73) 1.32 (0.86, 2.04) 1.39 (1.03, 1.87)
 rs4783099 0.82 (0.56, 1.20) 0.80 (0.46, 1.38) 0.98 (0.66, 1.45) 1.45 (1.07, 1.96) 1.70 (1.15, 2.52) 1.36 (1.02, 1.81)
FGFR2
 rs1047100 0.99 (0.67, 1.45) 1.98 (0.91, 4.33) 1.05 (0.69, 1.59) 1.02 (0.77, 1.34) 1.31 (0.67, 2.57) 0.96 (0.71, 1.31)
GABRB3
 rs2017247 0.88 (0.60, 1.29) 1.26 (0.67, 2.40) 0.92 (0.59, 1.45) 1.12 (0.85, 1.47) 1.21 (0.74, 1.98) 1.05 (0.78, 1.41)
 rs11637141 0.73 (0.49, 1.09) 2.34 (1.28, 4.29) 0.98 (0.65, 1.47) 1.12 (0.85, 1.46) 0.68 (0.34, 1.38) 0.88 (0.64, 1.21)
GLI2
 rs12711538 1.04 (0.72, 1.51) 0.95 (0.48, 1.88) 1.22 (0.79, 1.90) 1.10 (0.83, 1.45) 1.04 (0.63, 1.71) 0.99 (0.74, 1.33)
IRF6
 rs2235371 0.48 (0.11, 2.06) - 0.33 (0.03, 3.20) 0.22 (0.05, 0.95) - 1.00 (0.14, 7.10)
 rs2013162 0.95 (0.66, 1.38) 0.78 (0.42, 1.44) 0.75 (0.50, 1.10) 0.80 (0.61, 1.06) 0.65 (0.41, 1.04) 0.75 (0.57, 1.00)
 rs7552506 0.98 (0.67, 1.42) 0.89 (0.48, 1.65) 0.79 (0.54, 1.15) 0.92 (0.70, 1.21) 0.84 (0.53, 1.33) 0.82 (0.62, 1.09)
 rs2235377 0.53 (0.12, 2.27) - 0.50 (0.05, 5.51) 0.37 (0.11, 1.23) - 0.25 (0.03, 2.24)
PTCH1
 rs2066836 0.83 (0.56, 1.23) 0.39 (0.12, 1.29) 0.97 (0.59, 1.60) 0.87 (0.66, 1.16) 0.24 (0.08, 0.67) 0.86 (0.60, 1.21)
 rs357564 1.67 (1.14, 2.45) 1.27 (0.69, 2.36) 0.88 (0.60, 1.28) 1.24 (0.94, 1.63) 1.67 (1.13, 2.49)e 1.06 (0.82, 1.38)
RARA
 rs2229773 1.68 (0.90, 3.11) - 1.86 (0.74, 4.65) 0.78 (0.43, 1.41) - 1.00 (0.46, 2.16)
RYK
 rs1131262 0.96 (0.62, 1.50) - 0.64 (0.36, 1.16) 1.16 (0.85, 1.58) 0.22 (0.03, 1.72) 1.04 (0.70, 1.54)
 rs4470517 1.10 (0.74, 1.62) 1.33 (0.78, 2.27) 1.11 (0.74, 1.66) 1.03 (0.77, 1.36) 1.30 (0.88, 1.92) 0.96 (0.74, 1.26)
SATB2
 rs4673313 1.31 (0.84, 2.04) 0.43(0.06, 3.28) 0.89 (0.51, 1.54) 1.31 (0.95, 1.83) 1.01 (0.36, 2.80) 1.22 (0.79, 1.90)
 rs17199393 1.25 (0.67, 2.32) - 0.64 (0.25, 1.64) 1.14 (0.71, 1.82) - 0.94 (0.49, 1.83)
 rs16831366 1.72 (1. 05, 2.80) 1.37 (0.16, 11.86) 1.00 (0.49, 2.05) 1.36 (0.93, 2.00) 1.19 (0.23, 6.17) 0.96 (0.57, 1.64)
 rs1446636 1.33 (0.88, 2.01) 0.76 (0.17, 3.31) 1.13 (0.65, 1.95) 1.29 (0.94, 1.76) 1.61 (0.71, 3.65) 1.20 (0.82, 1.76)
 rs4675475 1.41 (0.96, 2.06) 0.91 (0.38, 2.20) 0.97 (0.61, 1.56) 1.24 (0.93, 1.65) 1.25 (0.71, 2.20) 1.11 (0.79, 1.56)
SUMO1
 rs3769817 1.20 (0.79, 1.84) 1.29 (0.37, 4.53) 1.00 (0.57, 1.74) 1.45 (1.08, 1.96) 0.61 (0.17, 2.11) 1.10 (0.75, 1.60)
 rs12470401 1.21 (0.79, 1.85) 1.39 (0.39, 4.92) 1.04 (0.61, 1.78) 1.38 (1.02, 1.87) 0.65 (0.18, 2.27) 1.06 (0.73, 1.53)
TGFA
 rs2166975 0.68 (0.46, 0.99) 0.85 (0.39, 1.85) 0.88 (0.56, 1.38) 1.01 (0.77, 1.33) 1.19 (0.69, 2.06) 1.10 (0.80, 1.53)
a

SNPs – single nucleotide polymorphisms

b

OR – odds ratio; CI – confidence interval; homozygous major allele was reference category

c

Homozygous for minor allele; homozygote odds ratios were not obtained when there were no cases and/or controls that were homozygous for the minor allele

d

TDT – transmission disequilibrium test; RR – risk ratio; CI – confidence interval

e

Significant difference for test of association, corrected P<0.05, after adjustment for multiple comparisons

DISCUSSION

In a large, ethnically homogenous study population, we tested 12 genes that previously were suggested to be risk factors for oral clefts and found that SNPs in three genes had statistically significant associations with oral clefts, after adjusting for multiple comparisons. PTCH1 A562A (rs2066836) and P1315L (rs357564) were associated with CLP while a SUMO1 non-coding SNP (rs3769817) and TGFA V159V (rs2166975) were associated with CP. When the cleft lip only and cleft lip with cleft palate subgroups were examined separately, it was found that PTCH1 P1315L was associated with cleft lip with cleft palate.

PTCH1 is a receptor for sonic hedgehog (SHH), an important regulator of craniofacial morphogenesis and an etiologic factor in holoprosencephaly, a phenotype that often includes oral clefts (Cohen Jr., 2006). PTCH1 mutations have been identified in patients with holoprosencephaly, some with oral clefts (Ming et al., 2002). PTCH1 mutations also cause nevoid basal cell carcinoma syndrome, a condition in which approximately 3–5% of affected patients have oral clefts (Kimonis et al., 1997), a prevalence that is 30–50 times greater than in the general population. A study of CLP and CP cases combined observed that PTCH1 P1315L (rs357564) was associated with nonsyndromic oral clefts in Caucasians (71 cases, 66 controls) but not in Filipinos (84 cases, 414 controls) (Mansilla et al., 2006). PTCH1 A562A (rs2066836) was not tested in that study but based on analyses in 220 Filipino families it also found an association between oral clefts and a haplotype composed of two PTCH1 SNPs, C89T (rs2297088) and T86C (rs2236407), that was of borderline significance. The larger sample size in the current study permitted separate analyses to be conducted for CLP and CP and the results confirmed the association of CLP with PTCH1 P1315L observed previously among Caucasians.

Reports of SUMO1 haploinsufficiency (Alkuraya et al., 2006) and deletion (Shi et al., 2009) in patients with oral clefts implicate this gene in nonsyndromic oral clefts. SUMO1 participates in the post-translational modification of many target proteins including the products of genes with evidence of a role in oral clefts in humans: msh homeobox1 (MSX1) (Lee et al., 2006); SATB2 (Dobreva et al., 2003); T-box 22 (TBX22) (Andreou et al., 2007); and tumor protein 63 (TP63) (Ghioni et al., 2005). A study of 181 cases and 162 controls from China that examined four SUMO1 SNPs found an association with CLP for one haplotype but not for any individual SNPs (Song et al., 2008). HapMap data from a Caucasian population shows that one of the haplotype SNPs from that study (rs6761131) is in perfect LD (D′ = r2 = 1) with the two SUMO1 SNPs that we tested.

A number of studies have examined whether TGFA polymorphisms are related to oral clefts in humans but have produced conflicting results. However, these studies varied according to the type of study population, study design, sample size, and TGFA variants used (Vieira, 2006). The authors of a meta-analysis stated that the evidence concerning TGFA polymorphisms and oral clefts was inconclusive, and suggested that the different study outcomes could be due to variations in the clinical characteristics of cases and in the proportion of cases with a positive family history (Mitchell, 1997). TGFA V159V (rs2166975) was associated with CP in a study of 94 CLP trios and 18 CP trios from Lithuania (Morkuniene et al., 2007). The association with CP was confirmed by the current study in which the analysis was based on a larger sample size from another European population: 133 informative CP trios. Based on HapMap data for Caucasians, TGFA V159V was not in strong LD with two TGFA variants that have previously been examined for an association with oral clefts: the BamHI polymorphism (rs11466297; r2 = 0.253) and the T3851C polymorphism of Marker H2A (rs11466285; r2 = 0.021). Other TGFA variants that have been examined in studies of oral clefts, including the TaqI polymorphism, were not found in the HapMap database.

The V274I (rs2235371) variant and other SNPs in IRF6 have been shown to be associated with nonsyndromic CLP, and although the European populations studied were rarely polymorphic for V274I, there was a suggestive positive association between the V allele and CLP in these populations (Zucchero et al., 2004). Less than 3% of our study population had the I allele of the V274I variant, resulting in too few informative subjects. IRF6 S153S (rs2013162) has also been associated with nonsyndromic CLP in previous reports (Scapoli et al., 2005). In our study, case-control and TDT association tests for IRF6 S153S produced uncorrected P-values <0.05 for isolated CLP; however, the results were not statistically significant after adjustment for multiple comparisons. HapMap data for Caucasians indicated that the four IRF6 SNPs we tested were not in strong LD with IRF6 rs642961 (rs2235371, r2 = 0.014; rs2013162, r2 = 0.203; rs7552506, r2 = 0.171; rs2235377, r2 = 0.010), a common polymorphism (minor allele frequency in Europeans = 0.25) which is associated with cleft lip only and cleft lip with cleft palate in Europeans (Rahimov et al., 2008).

Previous reports have implicated a role in oral clefts for the other eight genes that we tested: CLPTM1 (Yoshiura et al., 1998); CRISPLD2 (Chiquet et al., 2007); FGFR2 (Riley et al., 2007); GABRB3 (Scapoli et al., 2002; Inoue et al., 2008; Vieira et al., 2008); GLI2 (Roessler et al., 2003); RARA (Chenevix-Trench et al., 1992; Maestri et al., 1997); RYK (Watanabe et al., 2006); and SATB2 (Fitzpatrick et al., 2003; Beaty et al., 2006). In contrast, several studies did not find strong evidence of an association between oral clefts and variants in some of these genes. A study of Japanese patients found no association between CLPTM1 and oral clefts (Ichikawa et al., 2006). Two FGFR2 SNPs showed no association with nonsyndromic CLP in 294 Filipino families (Riley et al., 2007). There was only a suggestive association between GABRB3 SNPs and CP for patients from Iowa (Vieira et al., 2008) and between SATB2 variants and oral clefts based on another Iowa study population (Vieira et al., 2005). Some studies observed no association between RARA polymorphisms and oral clefts (Kanno et al., 2002; Mitchell et al., 2003). Most of these findings were reported after adjustment for multiple comparisons and it is possible that some true positive associations became statistically non-significant in an attempt to reduce false-positive reports. We found no evidence of an association with oral clefts among the SNPs that we selected for testing in these eight genes however other variants in these genes should be examined for a possible role in oral clefts.

With the exception of CRISPLD2, the genes we selected were also included in a recent report that examined 357 candidate genes in two European study populations (Jugessur et al., 2009). There were only two SNPs in common between the two studies: IRF6 rs2013162 and RYK rs1131262. Their study found that IRF6 was one of seven genes associated with CLP. The other six genes found to be associated with CLP were not among the genes that we selected. Another recent study that examined SNPs in or near a number of candidate genes found that IRF6 rs2013162 was associated with cleft lip only and that forkhead box E1 (FOXE1) was associated with cleft lip with cleft palate (Marazita et al., 2009). These reports provide further evidence that IRF6 is involved in nonsyndromic oral clefts but, along with the results of our study, also implicate a role for other genes. Other recent reports include the first genome-wide association studies of oral clefts. Two such studies in people with European ancestry found that a SNP in the chromosome 8q24 region (rs987525) is associated with CLP, although the SNP is not located near any known genes (Birnbaum et al., 2009; Grant et al., 2009).

An advantage of our study was the large number of CLP and CP cases and parents available from an ethnically homogenous study population. A limitation of our study was that only one or a few SNPs were tested in each gene. Many haplotype blocks were not examined in the genes we selected because we did not use haplotype-tagging SNPs. Therefore, failure to find an association for SNPs in some of the genes studied does not provide conclusive evidence about whether the genes play a role in oral clefts. Within each gene, we considered whether SNPs in different LD blocks could represent separate a priori hypotheses regarding an association between possible functional effects of the SNP on the gene and oral clefts. Had we adjusted for multiple comparisons separately for SNPs in different LD blocks, without accounting for the other SNPs examined in the same gene, then the only altered finding would be that CRISPLD2 rs4783099 would be associated with cleft lip with cleft palate. We chose to be more conservative when adjusting for multiple comparisons and therefore, within each gene, adjusted for all comparisons performed on the SNPs selected for the gene. This was done to limit the possibility of a false-positive finding in our study. We used a permutation procedure to control the probability of a false-positive association at 5%, separately for each gene. Therefore it was expected that the chance of a Type I error would be small. It is difficult to determine whether our adjustment was over-conservative. Another shortcoming was that most of the SNPs (in PTCH1, SUMO1, TGFA) associated with clefts in this report alter codons but do not result in an amino acid change. However, synonymous changes can have a direct impact on gene function (that is, via splicing, altering translation rates). It is also possible that other variants in these genes are associated with oral clefts and were missed by our study.

Our control group was comprised of women only whereas the case group included both males and females. If the allele frequencies in the control group were not representative of those in the source population that gave rise to the cases then this would produce bias in the results. However, we had no reason to expect that genotype frequencies would differ by sex because the genes we selected were located on autosomes and not sex chromosomes. We compared the case-control and TDT results for SNPs in PTCH1, SUMO1, and TGFA to determine whether the findings based on use of our control group were in agreement with another test of association. The effect estimates for the association between the TGFA SNP (rs2166975) and CP were similar using the case-control and TDT methods, although the result from the case-control method was not statistically significant after adjustment for multiple comparisons. This similarity suggests that the use of our control group did not result in biased estimates of effect.

For 10 of the 12 candidate genes, ours was the largest study to have examined associations between SNPs and nonsyndromic oral clefts. We observed associations with SNPs in PTCH1, SUMO1, and TGFA. PTCH1 A562A (rs2066836) is located in the protein’s sterol-sensing domain and P1315L (rs357564) is in the C-terminal domain; both domains are required for transduction of the SHH signal (Johnson et al., 2000; Martin et al., 2001). P1315L alters the third non-conserved amino acid in a motif, PPxY, that is a potential target for Nedd4 ubiquitin ligases, proteins that regulate the trafficking of intracellular cargo (Ingham et al., 2004). Mutation studies of the PPxY motif suggest that it is important for the regulation of PTCH1 turnover (Lu et al., 2006). TGFA V159V (rs2166975) encodes a terminal valine in the TGFA precursor protein and this residue is required for glycosylation leading to mature protein and for TGFA localization to the cell surface (Briley et al., 1997). The work reported here suggests that most of the 31 SNPs we examined do not play a role in oral clefts but that three genes, PTCH1, SUMO1, and TGFA, warrant further investigation as risk factors for oral clefts. We recommend that future research investigates whether the SNPs that showed positive associations with oral clefts in this study actually alter gene function.

Acknowledgments

This work was supported by the Intramural Research Programs of the National Institutes of Health, Eunice Kennedy Shriver National Institute of Child Health and Human Development and the National Human Genome Research Institute.

The authors gratefully acknowledge the patients and families who participated in the study, the Cleft Lip and Palate Association of Ireland, and the Dublin Cleft Center team.

Footnotes

Presented in poster form at the 22nd annual meeting of the Society for Pediatric and Perinatal Epidemiologic Research, June 22–23, 2009, Anaheim, California.

References

  1. Alkuraya FS, Saadi I, Lund JJ, et al. SUMO1 haploinsufficiency leads to cleft lip and palate. Science. 2006;313:1751. doi: 10.1126/science.1128406. [DOI] [PubMed] [Google Scholar]
  2. Andreou AM, Pauws E, Jones MC, et al. TBX22 missense mutations found in patients with X-linked cleft palate affect DNA binding, sumoylation, and transcriptional repression. Am J Hum Genet. 2007;81:700–712. doi: 10.1086/521033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Barrett JC, Fry B, Maller J, Daly MJ. Haploview: analysis and visualization of LD and haplotype maps. Bioinformatics. 2005;21:263–265. doi: 10.1093/bioinformatics/bth457. [DOI] [PubMed] [Google Scholar]
  4. Beaty TH, Hetmanski JB, Fallin MD, et al. Analysis of candidate genes on chromosome 2 in oral cleft case-parent trios from three populations. Hum Genet. 2006;120:501–518. doi: 10.1007/s00439-006-0235-9. [DOI] [PubMed] [Google Scholar]
  5. Birnbaum S, Ludwig KU, Reutter H, et al. Key susceptibility locus for nonsyndromic cleft lip with or without cleft palate on chromosome 8q24. Nat Genet. 2009;41:473–477. doi: 10.1038/ng.333. [DOI] [PubMed] [Google Scholar]
  6. Briley GP, Hissong MA, Chiu ML, et al. The carboxyl-terminal valine residues of proTGF alpha are required for its efficient maturation and intracellular routing. Mol Biol Cell. 1997;8:1619–1631. doi: 10.1091/mbc.8.8.1619. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Chenevix-Trench G, Jones K, Green AC, et al. Cleft lip with or without cleft palate: associations with transforming growth factor alpha and retinoic acid receptor loci. Am J Hum Genet. 1992;51:1377–1385. [PMC free article] [PubMed] [Google Scholar]
  8. Chiquet BT, Lidral AC, Stal S, et al. CRISPLD2: a novel NSCLP candidate gene. Hum Mol Genet. 2007;16:2241–2248. doi: 10.1093/hmg/ddm176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Cohen MM., Jr Holoprosencephaly: clinical, anatomic, and molecular dimensions. Birth Defects Res A Clin Mol Teratol. 2006;76:658–673. doi: 10.1002/bdra.20295. [DOI] [PubMed] [Google Scholar]
  10. Dobreva G, Dambacher J, Grosschedl R. SUMO modification of a novel MAR-binding protein, SATB2, modulates immunoglobulin mu gene expression. Genes Dev. 2003;17:3048–3061. doi: 10.1101/gad.1153003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Fitzpatrick DR, Carr IM, McLaren L, et al. Identification of SATB2 as the cleft palate gene on 2q32-q33. Hum Mol Genet. 2003;12:2491–2501. doi: 10.1093/hmg/ddg248. [DOI] [PubMed] [Google Scholar]
  12. Genisca AE, Frias JL, Broussard CS, et al. Orofacial clefts in the National Birth Defects Prevention Study, 1997–2004. Am J Med Genet A. 2009;149A:1149–1158. doi: 10.1002/ajmg.a.32854. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Ghioni P, D’Alessandra Y, Mansueto G, et al. The protein stability and transcriptional activity of p63alpha are regulated by SUMO-1 conjugation. Cell Cycle. 2005;4:183–190. doi: 10.4161/cc.4.1.1359. [DOI] [PubMed] [Google Scholar]
  14. Grant SF, Wang K, Zhang H, et al. A genome-wide association study identifies a locus for nonsyndromic cleft lip with or without cleft palate on 8q24. J Pediatr. 2009 doi: 10.1016/j.jpeds.2009.06.020. (In Press) [DOI] [PubMed] [Google Scholar]
  15. Harville EW, Wilcox AJ, Lie RT, et al. Cleft lip and palate versus cleft lip only: are they distinct defets? Am J Epidemiol. 2005;162:448–453. doi: 10.1093/aje/kwi214. [DOI] [PubMed] [Google Scholar]
  16. Hashmi SS, Waller DK, Langlois P, et al. Prevalence of nonsyndromic oral clefts in Texas: 1995–1999. Am J Med Genet A. 2005;134A:368–372. doi: 10.1002/ajmg.a.30618. [DOI] [PubMed] [Google Scholar]
  17. Ichikawa E, Watanabe A, Nakano Y, et al. PAX9 and TGFB3 are linked to susceptibility to nonsyndromic cleft lip with or without cleft palate in the Japanese: population-based and family-based candidate gene analyses. J Hum Genet. 2006;51:38–46. doi: 10.1007/s10038-005-0319-8. [DOI] [PubMed] [Google Scholar]
  18. Ingham RJ, Gish G, Pawson T. The Nedd4 family of E3 ubiquitin ligases: functional diversity within a common modular architecture. Oncogene. 2004;23:1972–1984. doi: 10.1038/sj.onc.1207436. [DOI] [PubMed] [Google Scholar]
  19. Inoue H, Kayano S, Aoki Y, et al. Association of the GABRB3 gene with nonsyndromic oral clefts. Cleft Palate Craniofac J. 2008;45:261–266. doi: 10.1597/06-142. [DOI] [PubMed] [Google Scholar]
  20. Johnson RL, Milenkovic L, Scott MP. In vivo functions of the patched protein: requirement of the C terminus for target gene inactivation but not Hedgehog sequestration. Mol Cell. 2000;6:467–478. doi: 10.1016/s1097-2765(00)00045-9. [DOI] [PubMed] [Google Scholar]
  21. Jugessur A, Shi M, Gjessing HK, et al. Genetic determinants of facial clefting: analysis of 357 candidate genes using two national cleft studies from Scandinavia [electronic article] PLoS One. 2009;4:e5385. doi: 10.1371/journal.pone.0005385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Kanno K, Suzuki Y, Yang X, et al. Lack of evidence for a significant association between nonsyndromic cleft lip with or without cleft palate and the retinoic acid receptor alpha gene in the Japanese population. J Hum Genet. 2002;47:269–274. doi: 10.1007/s100380200038. [DOI] [PubMed] [Google Scholar]
  23. Kimonis VE, Goldstein AM, Pastakia B, et al. Clinical manifestations in 105 persons with nevoid basal cell carcinoma syndrome. Am J Med Genet. 1997;69:299–308. [PubMed] [Google Scholar]
  24. Kirke PN, Mills JL, Molloy AM, et al. Impact of the MTHFR C677T polymorphism on risk of neural tube defects: a case-control study. BMJ. 2004;328:1535–1536. doi: 10.1136/bmj.38036.646030.EE. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Lee H, Quinn JC, Prasanth KV, et al. PIAS1 confers DNA-binding specificity on the Msx1 homeoprotein. Genes Dev. 2006;20:784–794. doi: 10.1101/gad.1392006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Lu X, Liu S, Kornberg TB. The C-terminal tail of the Hedgehog receptor Patched regulates both localization and turnover. Genes Dev. 2006;20:2539–2551. doi: 10.1101/gad.1461306. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Maestri NE, Beaty TH, Hetmanski J, et al. Application of transmission disequilibrium tests to nonsyndromic oral clefts: including candidate genes and environmental exposures in the models. Am J Med Genet. 1997;73:337–344. doi: 10.1002/(sici)1096-8628(19971219)73:3<337::aid-ajmg21>3.0.co;2-j. [DOI] [PubMed] [Google Scholar]
  28. Mansilla MA, Cooper ME, Goldstein T, et al. Contributions of PTCH gene variants to isolated cleft lip and palate. Cleft Palate Craniofac J. 2006;43:21–29. doi: 10.1597/04-169R.1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Marazita ML, Mooney MP. Current concepts in the embryology and genetics of cleft lip and cleft palate. Clin Plast Surg. 2004;31:125–140. doi: 10.1016/S0094-1298(03)00138-X. [DOI] [PubMed] [Google Scholar]
  30. Marazita ML, Lidral AC, Murray JC, et al. Genome scan, fine-mapping, and candidate gene analysis of non-syndromic cleft lip with or without cleft palate reveals phenotype-specific differences in linkage and association results. Hum Hered. 2009;68:151–170. doi: 10.1159/000224636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Martin V, Carrillo G, Torroja C, et al. The sterol-sensing domain of Patched protein seems to control Smoothened activity through Patched vesicular trafficking. Curr Biol. 2001;11:601–607. doi: 10.1016/s0960-9822(01)00178-6. [DOI] [PubMed] [Google Scholar]
  32. Mills JL, Molloy AM, Parle-McDermott A, et al. Folate-related gene polymorphisms as risk factors for cleft lip and cleft palate. Birth Defects Res A Clin Mol Teratol. 2008;82:636–643. doi: 10.1002/bdra.20491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Ming JE, Kaupas ME, Roessler E, et al. Mutations in PATCHED-1, the receptor for SONIC HEDGEHOG, are associated with holoprosencephaly. Hum Genet. 2002;110:297–301. doi: 10.1007/s00439-002-0695-5. [DOI] [PubMed] [Google Scholar]
  34. Mitchell LE. Transforming growth factor alpha locus and nonsyndromic cleft lip with or without cleft palate: a reappraisal. Genet Epidemiol. 1997;14:231–240. doi: 10.1002/(SICI)1098-2272(1997)14:3<231::AID-GEPI2>3.0.CO;2-8. [DOI] [PubMed] [Google Scholar]
  35. Mitchell LE, Murray JC, O’Brien S, et al. Retinoic acid receptor alpha gene variants, multivitamin use, and liver intake as risk factors for oral clefts: a population-based case-control study in Denmark, 1991–1994. Am J Epidemiol. 2003;158:69–76. doi: 10.1093/aje/kwg102. [DOI] [PubMed] [Google Scholar]
  36. Morkuniene A, Steponaviciut D, Utkus A, et al. Few associations of candidate genes with nonsyndromic orofacial clefts in the population of Lithuania. J Appl Genet. 2007;48:89–91. doi: 10.1007/BF03194663. [DOI] [PubMed] [Google Scholar]
  37. Parle-McDermott A, Mills JL, Kirke PN, et al. Analysis of the MTHFR 1298A-->C and 677C-->T polymorphisms as risk factors for neural tube defects. J Hum Genet. 2003;48:190–193. doi: 10.1007/s10038-003-0008-4. [DOI] [PubMed] [Google Scholar]
  38. Rahimov R, Marazita ML, Visel A, et al. Disruption of an AP-2α binding site in an IRF6 enhancer is associated with cleft lip. Nat Genet. 2008;40:1341–1347. doi: 10.1038/ng.242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Riley BM, Mansilla MA, Ma J, et al. Impaired FGF signaling contributes to cleft lip and palate. Proc Natl Acad Sci USA. 2007;104:4512–4517. doi: 10.1073/pnas.0607956104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Roessler E, Du YZ, Mullor JL, et al. Loss-of-function mutations in the human GLI2 gene are associated with pituitary anomalies and holoprosencephaly-like features. Proc Natl Acad Sci USA. 2003;100:13424–13429. doi: 10.1073/pnas.2235734100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Scapoli L, Martinelli M, Pezzetti F, et al. Linkage disequilibrium between GABRB3 gene and nonsyndromic familial cleft lip with or without cleft palate. Hum Genet. 2002;110:15–20. doi: 10.1007/s00439-001-0639-5. [DOI] [PubMed] [Google Scholar]
  42. Scapoli L, Palmieri A, Martinelli M, et al. Strong evidence of linkage disequilibrium between polymorphisms at the IRF6 locus and nonsyndromic cleft lip with or without cleft palate, in an Italian population. Am J Hum Genet. 2005;76:180–183. doi: 10.1086/427344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Schliekelman P, Slatkin M. Multiplex relative risk and estimation of the number of loci underlying an inherited disease. Am J Hum Genet. 2002;71:1369–1385. doi: 10.1086/344779. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Shi M, Mostowska A, Jugessur A, et al. Identification of microdeletions in candidate genes for cleft lip and/or palate. Birth Defects Res A Clin Mol Teratol. 2009;85:42–51. doi: 10.1002/bdra.20571. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Song T, Li G, Jing G, et al. SUMO1 polymorphisms are associated with non-syndromic cleft lip with or without cleft palate. Biochem Biophys Res Commun. 2008;377:1265–1268. doi: 10.1016/j.bbrc.2008.10.138. [DOI] [PubMed] [Google Scholar]
  46. Vieira AR, Avila JR, Daack-Hirsch S, et al. Medical sequencing of candidate genes for nonsyndromic cleft lip and palate [electronic article] PLoS Genet. 2005;1:e64. doi: 10.1371/journal.pgen.0010064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Vieira AR. Association between the transforming growth factor alpha gene and nonsyndromic oral clefts: a HuGE review. Am J Epidemiol. 2006;163:790–810. doi: 10.1093/aje/kwj103. [DOI] [PubMed] [Google Scholar]
  48. Vieira AR, Howe A, Murray JC. Studies of gamma-aminobutyric acid type A receptor beta3 (GABRB3) and glutamic acid decarboxylase 67 (GAD67) with oral clefts. Am J Med Genet A. 2008;146A:2828–2830. doi: 10.1002/ajmg.a.32260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Watanabe A, Akita S, Tin NT, et al. A mutation in RYK is a genetic factor for nonsyndromic cleft lip and palate. Cleft Palate Craniofac J. 2006;43:310–316. doi: 10.1597/04-145.1. [DOI] [PubMed] [Google Scholar]
  50. Yoshiura K, Machida J, Daack-Hirsch S, et al. Characterization of a novel gene disrupted by a balanced chromosomal translocation t(2;19)(q11.2;q13.3) in a family with cleft lip and palate. Genomics. 1998;54:231–240. doi: 10.1006/geno.1998.5577. [DOI] [PubMed] [Google Scholar]
  51. Zucchero TM, Cooper ME, Maher BS, et al. Interferon regulatory factor 6 (IRF6) gene variants and the risk of isolated cleft lip or palate. N Engl J Med. 2004;351:769–780. doi: 10.1056/NEJMoa032909. [DOI] [PubMed] [Google Scholar]

RESOURCES