ArticlePDF Available

Genetic and environmental risk factors for submucous cleft palate

Authors:

Abstract and Figures

Reiter R, Brosch S, Lüdeke M, Fischbein E, Haase S, Pickhard A, Assum G, Schwandt A, Vogel W, Högel J, Maier C. Genetic and environmental risk factors for submucous cleft palate. Eur J Oral Sci 2012; 120: 97–103. © 2012 Eur J Oral Sci A multifactorial aetiology with genetic and environmental factors is assumed for orofacial clefts. Submucous cleft palate (SMCP), a subgroup of cleft palates with insufficient median fusion of the muscles of the soft palate hidden under the mucosa, has a prevalence of 1:1,250–1:5,000. We described the prevalence of risk factors among 103 German patients with the subtype SMCP and genotyped 24 single nucleotide polymorphisms (SNPs) from 12 candidate genes for orofacial clefts. Analysis of risk factors yielded a positive history for maternal cigarette smoking during pregnancy in 25.2% of the patients, and this was significantly more frequent than in the normal population. The group of patients differed in allele frequencies at SNP rs3917192 of the gene TGFB3 (nominal P = 0.053) and at SNP rs5752638 of the gene MN1 (nominal P = 0.075) compared with 279 control individuals. Our results indicate a potential role of maternal smoking during pregnancy for the formation of SMCP. The analysis of genetic variants hints at the contribution of TGFB3 and MN1 in the aetiology of SMCPs.
Content may be subject to copyright.
Genetic and environmental
risk factors for submucous
cleft palate
Rudolf Reiter
1
, Sibylle Brosch
1
,
Manuel Ldeke
2
, Elena Fischbein
2
,
Stephan Haase
3
, Anja Pickhard
4
,
Gnter Assum
2
, Anke Schwandt
2
,
Walther Vogel
2
, Josef Hçgel
2
*,
Christiane Maier
2
*
1
Section of Phoniatrics and Pedaudiology,
Department of Otolaryngology, Head and Neck
Surgery, University of Ulm, Ulm;
2
Institute of
Human Genetics, University of Ulm, Ulm;
3
Department of Cranio-Maxillo-Facial Surgery,
University of Ulm, Ulm;
4
Department of
Otolaryngology Head and Neck Surgery,
Technical University Munich, Munich, Germany
*Authors who contributed equally to this article.
Cleft lip with or without cleft palate (CL/P) and cleft
palate only (CP) are the most common congenital cra-
niofacial disorders, with a combined prevalence of 1:700
in newborns. The submucous cleft palate (SMCP) is a
subgroup of cleft palates with insufficient median fusion
of the muscles of the soft palate hidden under the mucosa
(Fig. 1) and has a prevalence of 1:1,250–1:5,000 (1).
About 70% of all orofacial celfts can occur as an isolated
anomaly and the rest are a part of a complex mal-
formation syndrome. A multifactorial aetiology with
genetic and environmental factors is assumed (2, 3).
Maternal exposure to potential teratogens during the
first trimester of pregnancy is a known risk factor for
complete clefts. Among the known teratogenic sub-
stances are alcohol (2, 4) and tobacco smoke (5, 6). In
fact, maternal smoking may be responsible for about 4%
of all clefts (7, 8). The use of anticonvulsants (9) and of
corticoids (10) during pregnancy is also associated with
an increased risk of clefts. Additionally, the risk for
OFCs increases with parental age (11, 12), and a positive
family history of OFCs in infants or parents is a known
risk factor for both CL/P and CP (13). A number of
genes have been reported to be associated with non-
syndromic CL/P; however, only a few reports on genes
associated with CP exist (3, 14). Cleft lip with or without
cleft palate and CP are assumed to have different aeti-
ologies (14), and SMCP occurs frequently in syndromes
but also as a non-syndromal form (15). However, no
association studies have yet been performed for the
SMCP phenotype.
We therefore performed such an association study
with 12 candidate genes. These were selected for one or
more of the following reasons: their expression is asso-
ciated with palate growth during embryogenesis, knock
out of the gene leads to OFCs (mainly CP or SMCP) in
mice, or they are known to be associated in humans with
syndromal or non-syndromal OFCs. In these 12 genes we
genotyped 24 single nucleotide polymorphisms (SNPs) in
a sample of 103 patients, of German origin, with non-
syndromic SMCPs and compared the allele frequencies
with 279 healthy (population) controls. Furthermore, the
possible role of environmental risk factors, such as
alcohol and nicotine exposure, of the mothers during
pregnancy was evaluated.
Material and methods
Population
Collection of clinical data for the patients with SMCP was
performed according to the International Consortium for
Oral Clefts Genetics (16). We studied 103 patients with
SMCP (58 male subjects and 45 female subjects, median age
15.0 yr, range 3.2–68.0 yr) of German origin for this case–
control study. German background was assumed if their
grandparents had been born in Germany. Clinical assess-
Reiter R, Brosch S, Lu
¨deke M, Fischbein E, Haase S, Pickhard A, Assum G, Schwandt
A, Vogel W, Ho
¨gel J, Maier C. Genetic and environmental risk factors for submucous
cleft palate.
Eur J Oral Sci 2012; 120: 97–103. 2012 Eur J Oral Sci
A multifactorial aetiology with genetic and environmental factors is assumed for
orofacial clefts. Submucous cleft palate (SMCP), a subgroup of cleft palates with
insufficient median fusion of the muscles of the soft palate hidden under the mucosa,
has a prevalence of 1:1,250–1:5,000. We described the prevalence of risk factors among
103 German patients with the subtype SMCP and genotyped 24 single nucleotide
polymorphisms (SNPs) from 12 candidate genes for orofacial clefts. Analysis of risk
factors yielded a positive history for maternal cigarette smoking during pregnancy in
25.2% of the patients, and this was significantly more frequent than in the normal
population. The group of patients differed in allele frequencies at SNP rs3917192 of
the gene TGFB3 (nominal P= 0.053) and at SNP rs5752638 of the gene MN1
(nominal P= 0.075) compared with 279 control individuals. Our results indicate a
potential role of maternal smoking during pregnancy for the formation of SMCP. The
analysis of genetic variants hints at the contribution of TGFB3 and MN1 in the
aetiology of SMCPs.
Rudolf Reiter, Section of Phoniatrics and
Pedaudiology, Department of Otolaryngology,
Head and Neck Surgery, University of Ulm,
Frauensteige 12, 89070 Ulm, Germany
Telefax: +49–731–50039702
E-mail: rudolf.reiter@uniklinik-ulm.de
Key words: genetics; MN1; smoking;
submucous cleft palate; TGFB3
Accepted for publication January 2012
Eur J Oral Sci 2012; 120: 97–103
DOI: 10.1111/j.1600-0722.2012.00948.x
Printed in Singapore. All rights reserved
2012 Eur J Oral Sci
European Journal of
Oral Sciences
ment of the patients was carried out by an ear, nose, and
throat (ENT) surgeon, a specialist for phoniatrics, a surgeon
for oral and maxillofacial surgery and, if necessary, by a
medical geneticist to exclude a possible syndrome. A de-
tailed questionnaire was filled out to identify possible
additional malformations and any possible risk factors for
clefts, including a positive family history for OFCs, mater-
nal smoking, and ingestion of known teratogenic medica-
tions or toxins such as alcohol.
For the association study, we used a sample of 279
unrelated healthy young persons (mostly between 20 and
30 yr of age) as population controls. They had a gender
distribution about equal to that of the study group and were
of German origin, as determined by first and last name, and
by maiden name in married women. DNA was originally
prepared for paternity testing between 1995 and 2000, as
described below, and these data were later (after the legal
period) made anonymous to allow this sample to be used as
population controls for association studies. A screening for
cleft status was not made in these controls, but an appre-
ciable reduction in power was not expected because the
prevalence of OFCs is very low in the general population
(17, 18). Information on maternal exposure to risk factors
during pregnancy was not available for these controls. In
order to ascertain the influence of environmental factors on
the occurrence of SMCP, we had to rely on the literature
and chose a population-based study of 3,103 pregnant wo-
men of German origin (19) that did not include the popu-
lation controls used for the genetic association study.
However, these pregnant women were not screened for
having a child with any malformation.
The study was approved by the local Ethics Committee.
Written,informedconsentwasobtainedfromallpatientsortheir
parents. Peripheral venous blood samples were taken. DNA was
extracted from leucocytes as described by Miller et al. (20).
Selection of genes for association
The following 12 genes were selected for our study because
they have been reported to be associated with palate growth
in gene-expression analyses, with OFCs in knockout mice,
or in humans with (non-)syndromal clefts.
Expression of the T-box transcription factor gene 22
(TBX22) is associated with palate growth during embryo-
genesis. In healthy C57BL/6N mice, expression of TBX22
mRNA was observed during embryogenesis in distinct
regions of the posterior palatal shelves before fusion, indi-
cating its function in palatogenesis (21). This functional rel-
evance was confirmed in TBX22
)/)
knockout mice, which
exhibit an SMCP phenotype caused by reduced palatal bone
formation during embryogenesis (22), and in patients with
X-linked cleft palate and ankyloglossia (CPX) who had
mutations in TBX22 (23). TBX22 mutations are also often
found in non-syndromic cleft palate in the Thai population
(24). Sequencing of the TBX22 promoter region in 137 pa-
tients with CP revealed seven SNPs, two of which (rs41307258
and rs7055763) are associated with CP (25).
Knockout of the following genes led to OFCs (mainly CP
or SMCP) in mice. The meningeoma 1 gene (MN1) has been
reported to be associated with OFCs in MN1
)/)
or MN1
+/)
knockout mice, which showed a CP phenotype as a result of
reduced palatal bone formation during embryogenesis (26).
Mice lacking the transforming growth factor b3 gene
(TGFB3
)/)
) exhibit a failure of the palate shelves to fuse,
causing CP with incomplete penetrance (27).
In a mouse model for induction of CP with 2,3,7,8,-ter-
achlorodibenzo-p-dioxin (TCDD) the fibroblast growth
factor receptor 1 gene (FGFR1) was not expressed in the
medial edge epithelium after the fusion phase of the palate
shelves (28), and 80% of mice with a homozygous knockout
of FGFR1 have craniofacial defects, such as CP with open
palatine shelves. It is assumed that palatal shelf elevation is
blocked (29). In humans, association studies/genotyping of
SNPs in FGFR1 found borderline significant associations
between non-syndromal CL/P and the SNP rs13317 in
FGFR1 in 294 multiplex Filipino families. Furthermore a
nonsense mutation (R609X) was found in FGFR1 in
patients with CL/P (30).
Association of genes with syndromal or non-syndromal
OFCs in humans were also used as a source for selection of
our genes. In syndromes with CP or SMCP, some genes were
found to be mutated, such as interferon regulatory factor 6
(IRF6), which was present as a nonsense mutation (Y67Z) in
a German family with Van der Woude syndrome (31).
Variants of IRF6 were associated with non-syndromal CL/P
in a case–control association study of 460 patients and 952
controls from central Europe (17). Genotyping of 1,536
SNPs from 357 candidate genes in samples from Norway
(562 case–parent and 592 control–parent triads) and Den-
mark (235 case–parent triads) showed significant associa-
tions of IRF6 and alcohol dehydrogenase 1C (ADH1C) with
CL/P and of aristaless-like homeobox 3 (ALX3), ETS var-
iant gene 5 (ETV5), and platelet-derived growth factor C
(PDGFC) with CP. (3). Associations with non-syndromic
CL/P were also found for methylenetetrahydrofolate reduc-
tase (MTHFR) and muscle segment homeobox 1 (MSX1)ina
study of 176 haplotype-tagging SNPs in 18 candidate genes in
a case–control study in an Estonian sample (32). In addition,
transforming growth factor a(TGFA) influences the risk for
CL/P with an apparent parent-of-origin effect, as shown in a
case–parent trio design (33).
Selection of SNPs
For genes known to be associated with OFCs we did not use
the SNPs already studied, but selected tagging SNPs con-
sidering linkage disequilibrium (LD) patterns within the
candidate gene loci by using the HapMap genome browser
(release #24, phase 1 and 2 data sets). For every gene
obviously covered by a single block of LD (D¢> 0.9) we
selected a first SNP, preferably with two equally common
Fig. 1. Submucous cleft palate with the three main morpho-
logical symptoms: (A) bifid uvula and (B) a transluscent zone
lacking (C) a posterior nasal spine, leading to a bony notch in
the posterior end of the hard palate.
98 Reiter et al.
alleles. If, in the single block gene, a second SNP was
present that was likely to split the allele of one of the first
SNPs into two equally frequent haplotypes, this additional
SNP was also genotyped. For genes with interrupted LD
patterns, at least one SNP was chosen per block, and one
further SNP from the same block was selected if considered
useful for defining equally common haplotypes. One
exception to this selection scheme was the MN1 gene, which
did not exhibit a defined block structure but rather diffuse
LD patterning. From this gene locus two SNPs with com-
mon alleles were arbitrarily selected for genotyping. Can-
didate genes and genotyped SNPs are shown in Table 1.
Genotyping was performed using predesigned allelic dis-
crimination probes on a TaqMan 7900HT instrument
(Applied Biosystems, Darmstadt, Germany) in a 384-well
plate format. Reactions were carried out, according to the
manufacturerÕs instructions, in a sample reaction volume of
5ll using approximately 10 ng of genomic DNA as
template.
All markers were checked for deviations from Hardy–
Weinberg equilibrium. For association tests, allele or hap-
lotype frequencies were compared between cases and con-
trols using the chi-square test of association, and genotype
frequencies were compared using the Cochran–Armitage
test for trend. Allelic ORs are given together with their 95%
CI. Deviations of proportions (e.g. the proportion of
smokers) from population values were tested using the (two-
sided) chi-square goodness-of-fit test. All analyses were
considered to be explorative without adjustments for mul-
tiple testing. Calculations were carried out using plink
(version 1.03, http://pngu.mgh.harvard.edu/purcell/plink )
and the reported significance (P-value) is nominal. None of
the associations reached nominal significance and the
threshold for Bonferroni corrections would be much lower
(P< 0.002). plink was also used to estimate haplotype
frequencies and to test for differences between cases and
controls.
Results
Description of risk factors
Description of maternal and paternal risk factors for
OFCs yielded a positive family history (up to second-
degree relatives) in 14 out of 103 patients with SMCP.
The distribution of cleft types among relatives was as
follows: nine out of 14 had SMCP, four out of 14 had
CP, and one out of 14 had CL/P. Twenty-nine mothers
of patients (28.2%) with SMCP smoked cigarettes during
pregnancy and 18 mothers (17.5%) took alcohol.
Smoking during pregnancy appeared to be significantly
more frequent in this cohort of mothers of patients with
SMCP (prevalence = 28.2%) compared with the prev-
alence in a previously reported and independent study of
mothers in Bavaria (9.8%, P< 0.0001) (19), whereas
alcohol consumption was less frequent (two-sided
P= 0.068, Table 2). Two of these mothers took folic
acid, five took iodine, three took analgesics, and 18 took
medicinal drugs without any known risk for OFC. The
median age of the mothers at the birth of their children
with SMCP was 30.2 (range 20–43) yr.
Table 1
Candidate genes and genotyped single nucleotide polymorphisms (SNPs), together with the international SNP identification (rs) number
Gene rs number
Alcohol dehydrogenase 1C (ADH1C) rs1614972, rs3133158
Aristales-like homeobox 3 (ALX3) rs2360635, rs3754443
Ets variant gene 5 (ETV5) rs1356292, rs7433760
Fibroblast growth factor receptor 1 (FGFR1) rs2304000, rs6987534, rs6996321
Interferon regulatory factor 6 (IRF6) rs6685182, rs861019
Meningioma 1 (MN1) rs2189132, rs5752638
Muscle segment homeobox 1 (MSX1) rs12532
Methylenetetrahydrofolate reductase (MTHFR) rs1801132
Platelet-derived growth factor C (PDGFC) rs342311, rs6822796, rs6819797
Transforming growth factor, alpha (TGFA) rs6546610, rs7561997, rs7606793
Transforming growth factor, beta-3 (TGFB3) rs3917192
T-box 22 (TBX22) rs1429591, rs195294
Table 2
Frequency of the risk factors smoking and alcohol consumption during pregnancy of mothers of children with a submucous cleft palate
(SMCP) compared with population-based data of a control group of pregnant women of German origin (19)
Risk factor
Mothers of patients with
SMCP (n= 103), %
Controls (n= 3103)
P-value
Exposure to risk factor at
any time during pregnancy (%)
Exposure to risk factor
during first trimester only (%)
Smoking 28.2 9.8 9.2 <0.0001
Alcohol consumption 17.5 25.2 15.3 0.068
The controls are not the same set of controls that were used for the analysis of genetic markers. Deviations of proportions (e.g. the
proportion of smokers) from population values were tested using the (two-sided) chi-square goodness-of-fit test.
Submucous cleft palate 99
Case–control association results on allele level for the
investigated SNPs
The distribution of genotypes at selected markers was in
accordance with Hardy–Weinberg equilibrium (HWE),
except for the SNP rs6996321 in FGFR1 (P= 0.0083 vs.
controls). Therefore, this marker was excluded from the
analysis. A borderline nominally significant deviation
(P= 0.053) in allele frequency was observed for the SNP
rs3917192 of TGFB3 between cases (allele C= 0.888,
and allele T= 0.112) and controls (allele C= 0.831, and
allele T= 0.169). The OR of being a case given the allele
C was 1.61 (95% CI: 0.99–2.63). A tendency towards a
significant deviation (P= 0.075) was also present for the
SNP rs5752638 of MN1, for which allele frequencies were
T= 0.820 and C= 0.180 in cases vs. T= 0.760 and
C= 0.240 in controls (Table 3). The OR of MN1 being a
case, given the presence of allele T, was 1.44 (95% CI:
0.96–2.17). The Cochrane–Armitage trend test confirmed
the test results based on allele frequency (TGFB3,
P= 0.046; and MN1,P= 0.077). These P-values are all
nominal and apply only to single tests. After adjustment
for multiple testing, the P-values were not significant and
a trend towards significance was not seen (all P-values
>0.05).
Association results at the haplotype level
Candidate genes with multiple SNPs were additionally
analysed at the haplotype level using two (ADH1C,
ALX3, ETV5, IRF-6, MN1, and TBX22) or three
(FGFR1, PDGFC, and TGFA) SNPs. No significant
differences in haplotype frequencies were observed
between cases and controls. TGFB3 was one gene in
which an SNP approached significance; however, this
was the only SNP investigated in this gene and we do not
have haplotype information. In MN1, the other gene
with a suggestive association of an SNP, two SNPs were
analyzed and all four possible haplotypes were observed.
In this gene, the SNP rs5752638 had revealed a tendency
towards (nominally) a significant association. The C
(ÔprotectiveÕ) allele of this SNP (rs5752638) is split by the
alleles of the second SNP (rs2189132, alleles A/G) into
unequal parts, and the haplotype (C-A) with the lowest
frequency (controls = 0.056 and cases = 0.026) was
associated with a P-value (P= 0.072) similar to
rs5752638 alone (P= 0.076). Thus, the haplotype data
are consistent with the observation for the single SNPs.
Discussion
Like all OFCs, SMCPs occur as an isolated anomaly or
in the context of malformation syndromes. The velo-
cardio-facial syndrome (VCFS; 22q11 deletion syn-
drome) represents one of the most common examples of
a syndrome in which SMCP is frequently seen, occurring
in about 1/2,000 births. In a recent study of over 300
patients with VCFS, 77% had a cleft palate and, of these,
68% were either SMCP or occult SMCP (15). In addi-
Table 3
Genes, single nucleotide polymorphisms (SNPs), alleles, and allele frequencies in patients with submucous cleft palate (SMCP) and
controls
Gene SNP
Alleles
(major/minor)
Controls
(n= 279)
SMCP
(n= 103)
P-value
MAF MAF
TGFA rs7606793 T/C 0.274 0.262 0.739
rs7561997 G/A 0.453 0.437 0.683
rs6546610 G/A 0.444 0.452 0.863
TGFB3 rs3917192 C/T 0.169 0.112 0.053
MSX1 rs12532 A/G 0.292 0.248 0.225
IRF6 rs861019 A/G 0.439 0.418 0.593
rs6685182 C/A 0.351 0.387 0.359
FGFR1 rs2304000 C/G 0.208 0.223 0.556
rs6987534 G/C 0.437 0.442 0.644
MTHFR rs1801133 G/A 0.328 0.364 0.912
ADH1C rs1614972 C/T 0.312 0.345 0.349
rs3133158 C/G 0.286 0.277 0.388
TBX22 rs1429591 G/T 0.126 0.118 0.801
rs195294 T/C 0.353 0.313 0.378
MN1 rs5752638 T/C 0.240 0.180 0.075
rs2189132 G/A 0.328 0.309 0.617
ALX3 rs2360635 G/A 0.350 0.325 0.531
rs3754443 G/T 0.394 0.427 0.410
PDGFC rs6819797 T/C 0.367 0.417 0.215
rs342311 A/G 0.181 0.209 0.384
rs6822796 C/T 0.335 0.369 0.382
ETV5 rs1356292 T/C 0.195 0.194 0.971
rs7433760 A/G 0.283 0.325 0.257
Nominal significance of the different P-values between the patients with SMCP and controls was determined using the chi-square test.
All but two P-values were >0.2. P-values smaller than 0.1 are shown in bold.
MAF, minor allele frequency.
100 Reiter et al.
tion, in ankyloblepharon-ectodermal defects-cleft lip/
palate (AEC) syndrome, 17% of the patients were noted
to have SMCPs (34).
In patients with SMCP as an isolated anomaly the
aetiological relationship of the SMCP with other clefts is
much less clear. In our 14 familial cases, nine had a
relative with SMCP. This observation may indicate a
specific genetic factor for SMCP, which may be inter-
esting to study specifically. We chose 12 candidate genes
and analyzed 24 SNPs, but none of these reached the
usual level of nominal significance (P< 0.05). Never-
theless, two SNPs – one in TGFB3 and one in MN1
yielded P-values between 0.1 and 0.05, an observation
that may indicate a true association. However, this
remains to be shown in a larger sample size.
The above-mentioned borderline significant associa-
tions with SMCPs seem to be in accordance with
observations for TGFB3 in humans with complete OFCs.
Variants in TGFB3 have previously been studied for an
association with complete OFCs. Whereas some of these
studies revealed unequivocal evidence for an association
between OFCs and TGFB3 in ethnically different popu-
lations (35–37), others did not (38, 39). One European
study of trios (40) did not observe an association with
TGFB3, but found a clearly increased risk for CL/P with
paternal transmission. This observation may explain the
borderline association seen in other studies, including the
present one. In contrast to observations in humans, the
role of TGFB3 in OFCs in a murine model is obvious.
Mice lacking TGFB3 (TGFB3
)/)
knockout mice) exhibit
a cleft palate phenotype (27). Transforming growth fac-
tor signalling may also play a role in the subphenotype of
SMCP, as heterozygous germline mutations in TGFBR1
and TGFBR2 have recently been found to cause Loeys–
Dietz syndrome, which includes familial aortic aneu-
rysms and shows SMCP in this phenotype (41). To our
knowledge, the present study is the first that associates
TGFB3 with non-syndromal SMCP.
MN1, which encodes a cofactor for activating the
nuclear receptor of vitamin D and retinoic acid, and by
this means influences cell proliferation and cell differen-
tiation (26), is the second gene with borderline signifi-
cance in our study. Variants in this gene associated with
palate growth and clefts have been described in animal
models, but not yet in humans (2, 42). Mice deficient for
the transcription factor MN1 (MN1
)/)
mutant mice)
exhibit a complete cleft of the secondary palate (cleft
palate), whereas no significant growth deficiency was
found in the anterior plate at any stage of palate devel-
opment. The palatal shelves failed to elevate, in partic-
ular in the middle and posterior regions of MN1
)/)
mutant mice. It was also shown that the MN1 tran-
scription factor regulates expression of the transcription
factor, TBX22, during posterior palate growth in mice
(42). TBX22 knockout mice have SMCP phenotypes.
These mice exhibit reduced bone formation of the pos-
terior hard palate with a typical notch associated with
SMCP (22). Investigations in humans show that TBX22
mutations represent the most common single cause of
cleft palate known to date (24, 25). Therefore, the
TBX22/MN1 axis is an excellent candidate pathway for
SMCP susceptibility, and the trend towards an associa-
tion of MN1 variants would be worth following up with
larger study sizes.
Furthermore, not only genetic factors, but also envi-
ronmental factors, might influence the risk for SMCP. A
multifactorial mode of inheritance is often discussed (2,
43, 44). Known risk factors for OFCs are maternal
smoking and alcohol consumption during the first tri-
mester of pregnancy. We observed these risk factors in
25.2% (smoking) and 17.5% (alcohol consumption) of
our mothers. Smoking at any time during pregnancy was
significantly associated with SMCP (P< 0.0001).
Unfortunately, we did not have any timing information
for exposure to nicotine or alcohol consumption of
pregnant mothers in our patients (throughout pregnancy
vs. first trimester only). It is known that fetuses in the
first trimester (when fusion of the palate takes place) are
most susceptible to these teratogens (45).
Shi et al. (46) demonstrated, in a large collection of
1,244 cleft patients, that maternal nicotine consumption in
the 3-month period prior to, and during, pregnancy is a
significant risk factor for CL/P and CP, and risk occurs in
a dose-dependent manner. A meta-analysis confirmed this
observation (7). In particular, the combination of mater-
nal smoking during pregnancy and carrying a specific gene
variant of TGFB3 was a risk factor for CP (46, 47).
However, we did not see such an association in our pa-
tients with SMCP. The intake of drugs, such as anticon-
vulsants (9), and the maternal use of corticoids (10) are
also known risk factors for oral clefts. A positive history
for maternal drug intake was present in 17.5% of our
patients; however, drugs with a known specific risk were
not included. In our cohort, age was also considered to be
a risk factor (11, 12), but age was quite similar to that of
pregnant mothers in the general population (19).
A limitation of our investigation was that information
on possible risk factors for clefts was not available for
the control group in the genetic study. Therefore, a
comparison of samples (e.g. children of smoking and
non-smoking mothers) was not possible and the envi-
ronmental factors of our patients were compared with
population-based data of pregnant women of German
origin (19).
The genetic background may also influence our results
because allele frequencies are known to vary in different
populations and different ethnic backgrounds. Although
geographical origin was verified in our case–control
study, we cannot rule out hidden population stratifica-
tion as a cause for false-positive association.
It is known that the risk for a subsequent sibling to be
affected increases with the severity of cleft (13). Submu-
cous cleft palate is a mild form of cleft palates with a
hidden defect below the mucosa, compared with con-
tinuous/open clefts. This mild form might correlate with
a specific genetic susceptibility. On the other hand, pa-
tients with the VCFS had either an SMCP or a CP
phenotype (15). This indicates that there may be other
factors that influence the cleft type.
However, the milder (SMCP) form would probably
require an even larger group of patients in order to detect
differences of risk factors between cases and controls.
Submucous cleft palate 101
Thus, sample size is the most critical limitation of our
study, as a lack of power may cause no or insignificant
associations, even for actually causal genes.
Our results show that maternal smoking during preg-
nancy may contribute to the formation of SMCPs. The
analysis of genetic variants hints at the contribution of
TGFB3 and MN1 to SMCPs. These results need further
confirmation by larger case–control groups or functional
studies.
Acknowledgements – We thank all patients and their families for
participation in the study. Margot Brugger and Indira Wiest are
acknowledged for excellent technical assistance.
Conflicts of interest – The authors report no conflicts of interest.
References
1. Gosain AK, Conley SF, Marks S, Larson DL. Submucous
cleft palate: diagnostic methods and outcomes of surgical
treatment. Plast Reconstr Surg 1996; 97: 1497–1509.
2. Jugessur A, Murray JC. Orofacial clefting: recent insights
into a complex trait. Curr Opin Genet Dev 2005; 15: 270–278.
3. Jugessur A, Shi M, Gjessing HK, Lie RT, Wilcox AJ,
Weinberg CR, Christensen K, Boyles AL, Daak-Hirsch S,
Trung TN, Bille C, Lidral AC, Murray JC. Genetic
determinants of facial clefting: analysis of 357 candidate genes
using two national cleft studies from scandinavia. PLoS ONE
2009; 4: e5385.
4. Grewal J, Carmichael SL, Ma C, Lammer EJ, Shaw GM.
Maternal periconceptional smoking and alcohol consumption
and risk for select congenital anomalies. Birth Defects Res A
Clin Mol Teratol 2008; 82: 519–526.
5. Bille C, Olsen J, Vach W, Knudsen VK, Olsen SF,
Rasmussen K, Murray JC, Andersen AM, Christensen K.
Oral clefts and life style factors-a case-cohort study based on
prospective Danish data. Eur J Epidemiol 2007; 22: 173–181.
6. Shi M, Christensen K, Weinberg CR, Romitti P, Bathum
L, Lozada A, Morris RW, Lovett M, Murrray JC.
Orofacial cleft risk is increased with maternal smoking and
specific detoxification-gene variants. Am J Hum Genet 2007;
80: 76–90.
7. Honein MA, Rasmussen SA, Reefhuis J, Romitti PA, Lam-
mer EJ, Sun L, Correa A. Maternal smoking and environ-
mental tobacco smoke exposure and the risk of orofacial clefts.
Epidemiology 2007; 18: 226–233.
8. Vieira AR. Unraveling human cleft lip and palate research.
J Dent Res 2008; 87: 119–125.
9. Holmes LB, Baldwin EJ, Smith CR, Habecker E, Glassman
L, Wong SL, Wyszynski DF. Increased frequency of isolated
cleft palate in infants exposed to lamotrigine during pregnancy.
Neurology 2008; 70: 2152–2158.
10. Carmichael SL, Shaw GM. Maternal corticosteroid use and
risk of selected congenital anomalies. Am J Med Genet 1999; 86:
242–244.
11. Materna-Kiryluk A, Wisniewska K, Badura-Stronka M,
Mejnartowicz J, Wieckowska B, Balcar-Boron A, Czer-
wionka-Szaflarska M, Gajewska E, Godula-Stuglik U,
Krawczynski M, Limon J, Rusin J, Sawulicka-Oleszczuk H,
Szwalkiewicz-Warowicka E, Walczak M, Latos-Bielenska
A. Parental age as a risk factor for isolated congenital mal-
formations in a Polish population. Paediatr Perinat Epidemiol
2009; 23: 29–40.
12. Bille C, Skytthe A, Vach W, Knudsen LB, Andersen AM,
Murray JC, Christensen K. ParentÕs age and the risk of oral
clefts. Epidemiology 2005; 16: 311–316.
13. Bernheim N, Georges M, Malevez C, De Mey A, Mansbach
A. Embryology and epidemiology of cleft lip and palate.
B-ENT 2006; 2(Suppl): 11S–19S.
14. Carinci F, Scapoli L, Palmieri A, Zollino I, Pezzetti F.
Human genetic factors in nonsyndromic cleft lip and palate: an
update. Int J Pediatr Otorhinolaryngol 2007; 71: 1509–1519.
15. Friedman MA, Miletta N, Roe C, Wang D, Morrow BE,
Kates WR, Higgins AM, Shprintzen RJ. Cleft palate, retro-
gnathia and congenital heart disease in velo-cardio-facial syn-
drome: a phenotype correlation study. Int J Pediatr
Otorhinolaryngol 2011; 75: 1167–1172.
16. Mitchell LE, Beaty TH, Lidral AC, Munger RG, Murray
JC, Saal HM, Wyszynski DF. Guidelines for the design and
analysis of studies on nonsyndromic cleft lip and cleft palate in
humans: summary report from a workshop of the international
consortium for oral clefts genetics. Cleft Palate Craniofac J
2002; 39: 93–100.
17. Birnbaum S, Ludwig KU, Reutter H, Herms S, De Assis
NA, Diaz-Lacava A, Barth S, Lauster C, Schmidt G,
Scheer M, Saffar M, Martini M, Reich RH, Schiefke F,
Hemprich A, Poetzsch S, Wienker TF, Hoffmann P, Knapp
M, Kramer FJ, Noethen MM, Mangold E. IRF6 Gene
variants in central European patients with non-syndromic cleft
lip with or without cleft palate. Eur J Oral Sci 2009; 117:
766–769.
18. Moskvina V, Holmans P, Schmidt KM, Craddock N. Design
of case-controls studies with unscreened controls. Ann Hum
Genet 2005; 69: 566–576.
19. Rebhan B, Kohlhuber M, Schwegler U, Koletzko B,
Fromme H. Smoking, alcohol and caffeine consumption of
mothers before, during and after pregnancy – results of the
study Ôbreast-feeding habits in Bavaria. Gesundheitswesen 2009;
71: 391–398.
20. Miller SA, Dykes DD, Polesky HF. A simple salting out
procedure for extracting DNA from human nucleated cells.
Nucleic Acids Res 1988; 16: 1215.
21. Kim SM, Lee JH, Jabaiti S, Lee SK, Choi JY. Tbx22 expres-
sions during palatal development in fetuses with glucocorti-
coid-/alcohol-induced C57BL/6N cleft palates. J Craniofac
Surg 2009; 20: 1316–1326.
22. Pauws E, Hoshino A, Bentley L, Prajapati S, Keller C,
Hammond P, Martinez-Barbera JP, Moore GE, Stanier P.
Tbx22null mice have a submucous cleft palate due to reduced
palatal bone formation and also display ankyloglossia and
choanal atresia phenotypes. Hum Mol Genet 2009; 18: 4171–
4179.
23. Braybrook C, Doudney K, Marcano AC, Arnason A,
Bjornsson A, Patton MA, Goodfellow PJ, Moore GE,
Stanier P. The T-box transcription factor gene TBX22 is
mutated in X-linked cleft palate and ankyloglossia. Nat Genet
2001; 29: 179–183.
24. Suphapeetiporn K, Tongkobpetch S, Siriwan P, Shoteler-
suk V. TBX22 Mutations are a frequent cause of non-syndro-
mic cleft palate in the Thai population. Clin Genet 2007; 72:
478–483.
25. Pauws E, Moore GE, Stanier P. A functional haplotype
variant in the TBX22 promoter is associated with cleft palate
and ankyloglossia. J Med Genet 2009; 46: 555–561.
26. Meester-Smoor MA, Vermeij M, Van Helmond MJ, Molijn
AC, Van Wely KH, Hekman AC, Vermey-Keers C, Riegman
PH, Zwarthoff EC. Targeted disruption of the Mn1 oncogene
results in severe defects in development of membranous bones
of the cranial skeleton. Mol Cell Biol 2005; 25: 4229–4236.
27. Proetzel G, Pawlowski SA, Wiles MV, Yin M, Boivin GP,
Howles PN, Ding J, Ferguson MW, Doetschman T. Trans-
forming growth factor-beta 3 is required for secondary palate
fusion. Nat Genet 1995; 11: 409–414.
28. Fujiwara K, Yamada T, Mishima K, Imura H, Sugahara T.
Morphological and immunohistochemical studies on cleft pal-
ates induced by 2,3,7,8-tetrachlorodibenzo-p-dioxin in mice.
Congenit Anom (Kyoto) 2008; 48: 68–73.
29. Trokovic N, Trokovic R, Mai P, Partanen J. Fgfr1 regulates
patterning of the pharyngeal region. Genes Dev 2003; 17:
141–153.
30. Riley BM, Schultz RE, Cooper ME, GOLDSTEIN-
Mchenry T, Daack-Hirsch S, Lee KT, Dragan E, Vieira
AR, Lidral AC, Marazita ML, Murray JC. A genome-wide
102 Reiter et al.
linkage scan for cleft lip and cleft palate identifies a novel locus
on 8p11-23. Am J Med Genet A 2007; 143: 846–852.
31. Brosch S, Baur M, Blin N, Reinert S, Pfister M. A novel
IRF6 nonsense mutation (Y67X) in a German family with Van
der Woude syndrome. Int J Mol Med 2007; 20: 85–89.
32. Jagoma
¨gi T, Nikopensius T, Krjutskov K, Tammekivi V,
Viltrop T, Saag M, Metspalu A. MTHFR and MSX1 con-
tribute to the risk of nonsyndromic cleft lip/palate. Eur J Oral
Sci 2010; 118: 213–220.
33. Sull JW, Liang KY, Hetmanski JB, Wu T, Fallin MD,
Ingersoll RG, Park JW, Wu-Chou YH, Chen PK, Chong
SS, Cheah F, Yeow V, Park BY, Jee SH, Jabs EW, Redett R,
Scott AF, Beaty TH. Evidence that TGFA influences risk to
cleft lip with/without cleft palate through unconventional
genetic mechanisms. Hum Genet 2009; 126: 385–394.
34. Cole P, Hatef DA, Kaufmann Y, Magruder A, Bree A,
Friedmann E, Sindwani R, Holler LH. Facial clefting and
oroauditory pathway manifestations in Ankyloblepharon-
Ectodermal Defects-Cleft Lip/Palate (AEC) syndrome. Am J
Med Genet A 2009; 149: 1910–1915.
35. Ichikawa E, Watanabe A, Nakano Y, Akita S, Hirano A,
Kinoshita A, Kondo S, Kishino T, Uchiyama T, Niikawa N,
Yoshiura K. PAX9 and TGFB3 are linked to susceptibility to
nonsyndromic cleft lip with or without cleft palate in the Jap-
anese: population-based and family-based candidate gene
analyses. J Hum Genet 2006; 51: 38–46.
36. Vieira AR, Orioli IM, Castilla EE, Cooper ME, Marazita
ML, Murray JC. MSX1 and TGFB3 contribute to clefting in
South America. J Dent Res 2003; 82: 289–292.
37. Lidral AC, Romitti PA, Basart AM, Doetschman T, Ley-
sens NJ, Daack-Hirsch S, Semina EV, Johnson LR, Mach-
ida J, Burds A, Parnell TJ, Rubenstein JL, Murray JC.
Association of MSX1 and TGFB3 with nonsyndromic clefting
in humans. Am J Hum Genet 1998; 63: 557–568.
38. Jugessur A, Lie RT, Wilcox AJ, Murray JC, Taylor JA,
Saugstad OD, Vindenes HA, Abyholm F. Variants of devel-
opmental genes (TGFA, TGFB3, and MSX1) and their asso-
ciations with orofacial clefts: a case-parent triad analysis. Genet
Epidemiol 2003; 24: 230–239.
39. Salahshourifar I, Halim AS, Wan Sulaiman WA, Zilfalil
BA. Contribution of MSX1 variants to the risk of non-syn-
dromic cleft lip and palate in a Malay population. J Hum Genet
2011; 56: 755–758.
40. Reutter H, Birnbaum S, Mende M, Lauster C, Schmidt
G, Henschke H, Saffar M, Martini M, Lauster R,
Schiefke F, Reich RH, Braumann B, Scheer M, Knapp M,
Noethen MM, Kramer FJ, Mangold E. TGFB3 displays
parent-of-origin effects among central Europeans with non-
syndromic cleft lip and palate. J Hum Genet 2008; 53:
656–661.
41. Kirmani S, Tebben PJ, Lteif AN, Gordon D, Clarke BL,
Hefferan TE, Yaszemski MJ, McGrann PS, Lindor NM,
Ellison JW. Germline TGF-Beta receptor mutations and
skeletal fragility: a report on two patients with Loeys-Dietz
Syndrome. Am J Med Genet A 2010; 152: 1016–1019.
42. Liu W, Lan Y, Pauwes E, Meester-Smoor MA, Stanier P,
Zwarthoff EC, Jiang R. The Mn1 transcription factor acts
upstream of Tbx22 and preferentially regulates posterior palate
growth in mice. Development 2008; 135: 3959–3968.
43. Jugessur A, Farlie PG, Kilpatrick N. The genetics of iso-
lated orofacial clefts: from genotypes to subphenotypes. Oral
Dis 2009; 15: 437–453.
44. Eppley BL, Van Aalst JA, Robey A, Havlik RJ, Sadove AM.
The spectrum of orofacial clefting. Plast Reconstr Surg 2005;
115: 101e–114e.
45. Meng L, Bian Z, Torensma R, Von Den Hoff JW. Biological
mechanisms in palatogenesis and cleft palate. J Dent Res 2009;
88: 22–33.
46. Shi M, Wehby GL, Murray JC. Review on genetic variants
and maternal smoking in the etiology of oral clefts and other
birth defects. Birth Defects Res C Embryo Today 2008; 84:
16–29.
47. Romitti PA, Lidral AC, Munger RG, Daack-Hirsch S,
Burns TL, Murray JC. Candidate genes for nonsyndromic
cleft lip and palate and maternal cigarette smoking and alcohol
consumption: evaluation of genotype-environment interactions
from a population-based case-control study of orofacial clefts.
Teratology 1999; 59: 39–50.
Submucous cleft palate 103
... Submucous cleft palate (SMCP) is a rare congenital anomaly, with a prevalence of 1:1250 to 1:5000. 1 Calnan 2 presented the classic triad describing it, which includes a bifid uvula, a bony defect in the posterior end of the hard palate, and a translucent zone in the midline of the soft palate. Submucous cleft palate can be encountered as an isolated finding or as part of syndromic diseases. ...
... Submucous cleft palate is a special type of cleft palate, with insufficient median fusion of the muscles of the soft palate hidden under the mucosa. 1 Postnatally, radiological imaging and ultrasonography are both used to identify SMCP. The 3D CT reconstruction divides the bone defect of SMCP into 4 types: type I, involving a posterior nasal spine and no V-shaped hard-palate cleft; type II, involving part of the hard bone; type III, involving the complete palate and extending to the incisive foramen; and type IV, involving the complete palate and the alveolus. ...
Article
Full-text available
Objective We present a case with prenatal diagnosis of submucous cleft palate (SMCP) which was described using 2- and 3-dimensional (3D) ultrasonography in utero. Case Report A 25-year-old pregnant woman was referred to our department for fetal ultrasound screening. After the detection of cardiac and spinal malformations of fetal, further detailed examination detected SMCP, which showed a gap within the hard palate on axial transversal view with the soft palate visible on sagittal view. The imaging of a defective hard palate in prenatal 3D ultrasonography is similar to that in postmortem 3D computed tomography reconstruction. Conclusion A gap within the hard palate and verification of the visibility of the soft palate should be key points in the prenatal diagnosis of SMCP. Three-dimensional ultrasonic imaging is helpful for displaying the shape and extent of the bony defect in SMCP.
... 1 A submucosal cleft only involves defects in the bone and muscles of the palate, while the etiological risk factors appear to overlap partially with CLP and CP. 3 The formation of the secondary palate involves downward growth of the palatal shelves, elevation above the tongue, and fusion in the midline. 4 Subsequently, intramembranous ossification occurs. ...
Article
Full-text available
Objectives: Cleft lip and/or palate (CLP) is a common craniofacial birth defect caused by genetic as well as environmental factors. The phenotypic spectrum of CLP also includes submucous clefts with a defect in the palatal bone. To elucidate the contribution of vitamin A, we evaluated the effects of the vitamin A metabolite all-trans retinoic acid (ATRA) on the osteogenic differentiation and mineralization of mouse embryonic palatal mesenchymal cells (MEPM). Setting and sample population: MEPM cells were isolated from the prefusion palates of E13 mouse embryos from three different litters. Materials and methods: MEPM cells were cultured with and without 0.5 μM ATRA in osteogenic medium. Differentiation was analysed by the expression of osteogenic marker genes and alkaline phosphatase (ALP) activity after 1, 2, and 7 days. The expression of Wnt marker genes was also analysed. Mineralization was assessed by alizarin red staining after 7, 14, 21, and 28 days. Results: The bone marker genes Sp7, Runx2, Alpl, and Col1a1 were inhibited 10% ± 2%, 59% ± 7%, 79% ± 12% and 57% ± 20% (P < .05) at day 7. ALP activity was inhibited at days 1 and 7 by 35 ± 0% (P < .05) and 23 ± 6% (P < .001). ATRA also inhibited mineralization at 3 and 4 weeks. Finally, expression of the universal Wnt marker gene Axin2 was strongly reduced, by 31 ± 18% (P < .001), at day 7. Conclusion: Our data indicate that ATRA (vitamin A) inhibits bone formation by reducing Wnt signalling. This might contribute to the molecular aetiology of submucous clefting.
... Bagatin conducted similar research on 9720 Zagrebian school children and found 5 patients with SMCP, which resulted in an incidence of 1:1944 (Bagatin, 1985). Differences in reported incidences could partly be explained by variances in genes accountable for clefting among different study populations (Reiter et al., 2012). Moreover, Stewart et al. and Bagatin investigated asymptomatic children, where in the current study symptomatic children were examined (ie, solely cleft patients were registered). ...
Article
Full-text available
Objective To analyze the incidence of submucous cleft palate (SMCP) in a large national database and raise awareness among referring providers: pediatricians, speech pathologists, and dentists to minimize delay in diagnosis. Design Retrospective cohort study. Setting Tertiary setting. Patients Patients were extracted from the “Dutch Association for Cleft and Craniofacial Anomalies” database. A total of 6916 patients were included from 1997 until 2018 and divided into 2 groups (ie, SMCP versus cleft palate [CP]). Patients born before 1997 and adopted patients were excluded. Interventions Clefts were classified as either hard of soft palatal involvement based on anatomical landmarks at first consultation. Main Outcome Measures Primary outcomes were the patient characteristics in both groups (ie, gender, birth weight, gestational age, and additional anomalies). Secondary outcome was the time of diagnosis among subgroups. Results In total, 532 patients were diagnosed with SMCP (7.7%). Birth weight, gestational age, and additional anomalies did not differ between subgroups, but there were more males in the SMCP group ( P < .001). The median age of diagnosis of the SMCP group was significantly higher than of the CP group (987 vs 27 days; P < .001). Over the course of 22 years, the time of diagnosis for SMCP did not decrease. Conclusion Submucous cleft palate represents <10% of the Dutch cleft population and 19.4% of all CP. Time of diagnosis for SMCP is significantly longer when compared with time of diagnosis of CP, and this has not changed over the study period of 22 years.
... Expressed in the epithelium, TGFB3 regulates fusion at the MES and cell proliferation. Variants in TGFB3, in the context of smoking, have been associated with CL/P (Guo, Huang, Ding, Lin, & Gong, 2010;Lin et al., 2010;Romitti et al., 1999), CPO (Romitti et al., 1999), and submucous cleft palate (SMCP; Reiter et al., 2012). BMP4, a ligand for BMP signaling, has polymorphisms associated with CL/P and is involved in fusion of the medial and lateral nasal processes in this issue). ...
Article
Orofacial clefts (OFCs) are among the most common birth defects and impart a significant burden on afflicted individuals and their families. It is increasingly understood that many nonsyndromic OFCs are a consequence of extrinsic factors, genetic susceptibilities, and interactions of the two. Therefore, understanding the environmental mechanisms of OFCs is important in the prevention of future cases. This review examines the molecular mechanisms associated with environmental factors that either protect against or increase the risk of OFCs. We focus on essential metabolic pathways, environmental signaling mechanisms, detoxification pathways, behavioral risk factors, and biological hazards that may disrupt orofacial development.
... Many human cases involve submucous cleft palate, in which the cleft is covered by an intact mucosal membrane. Submucous clefts have a reported incidence of 1:1250-1:5000 (Reiter et al., 2012). Clinically, submucous cleft palate is characterized by a translucent zone (zona pellucida) along the midline of the posterior hard palate indicating a separation of the submucous palatal musculature, a palpable bony notch in the same area as the zona pellucida, and a bifid uvula (Kosowski et al., 2012). ...
Article
Cleft palate is one of the most common craniofacial congenital defects in humans. It is associated with multiple genetic and environmental risk factors, including mutations in the genes encoding signaling molecules in the sonic hedgehog (Shh) pathway, which are risk factors for cleft palate in both humans and mice. However, the function of Shh signaling in the palatal epithelium during palatal fusion remains largely unknown. Although components of the Shh pathway are localized in the palatal epithelium, specific inhibition of Shh signaling in palatal epithelium does not affect palatogenesis. We therefore utilized a hedgehog (Hh) signaling gain-of-function mouse model, K14-Cre;R26SmoM2, to uncover the role of Shh signaling in the palatal epithelium during palatal fusion. In this study, we discovered that constitutive activation of Hh signaling in the palatal epithelium results in submucous cleft palate and persistence of the medial edge epithelium (MEE). Further investigation revealed that precise downregulation of Shh signaling is required at a specific time point in the MEE during palatal fusion. Upregulation of Hh signaling in the palatal epithelium maintains the proliferation of MEE cells. This may be due to a dysfunctional p63/Irf6 regulatory loop. The resistance of MEE cells to apoptosis is likely conferred by enhancement of a cell adhesion network through the maintenance of p63 expression. Collectively, our data illustrate that persistent Hh signaling in the palatal epithelium contributes to the etiology and pathogenesis of submucous cleft palate through its interaction with a p63/Irf6-dependent biological regulatory loop and through a p63-induced cell adhesion network.
Article
Full-text available
Objective The aim was to provide epidemiological and clinical data on patients with orofacial clefts in Lower Saxony in Germany. Materials and methods The records of 404 patients with orofacial clefts treated surgically at the University Medical Center Goettingen from 2001 to 2019 were analyzed in this retrospective study. Prevalence of orofacial clefts in general, orofacial clefts as manifestation of a syndrome, sex distribution, and prevalence of different cleft types was evaluated and associated with the need for corrective surgery, family history, pregnancy complications, and comorbidities. Results The prevalence of orofacial clefts for Goettingen in Lower Saxony was 1:890. 231 patients were male and 173 were female. CLP was most common (39.1%) followed by CP (34.7%), CL (14.4%), CLA (9.9%), and facial clefts (2%). The left side was more frequently affected and unilateral cleft forms occurred more often than bilateral ones. Almost 10% of the population displayed syndromic CL/P. 10.9% of all patients had a positive family history regarding CL/P, predominantly from the maternal side. Pregnancy abnormalities were found in 11.4%, most often in the form of preterm birth. Comorbidities, especially of the cardiovascular system, were found in 30.2% of the sample. 2.2% of patients treated according to the University Medical Center Goettingen protocol corrective surgery was performed in form of a velopharyngoplasty or residual hole closure. Conclusions The epidemiological and clinical profile of the study population resembled the expected distributions in Western populations. The large number of syndromic CL/P and associated comorbidities supports the need for specialized cleft centers and interdisciplinary cleft care.
Chapter
Orofacial clefts (OFCs) are among the most common birth defects worldwide and are believed to result from both genetic and environmental risk factors. Orofacial development involves complex cellular mechanisms including migration, differentiation, apoptosis, and tissue fusion events largely involving cranial neural crest cells and orofacial epithelial cells. These activities culminate in the formation of the upper lip and palate and are tightly regulated by developmental signaling pathways. Disruption of these sensitive processes through genetic mutation, nutritional imbalance, and/or chemical exposure is the major factor driving OFC etiology. This chapter summarizes the field’s current understanding of OFC etiology with an examination of evidence from human population and animal studies. Genetic mutations that affect developmental signaling, transcriptional activity, extracellular matrix dynamics, and other processes are linked to both syndromic and nonsyndromic OFCs. Mutations that disrupt pathways such as transforming growth factor beta (TGF-β), bone morphogenic protein (BMP), fibroblast growth factor (FGF), retinoic acid (RA), hedgehog (Hh), and others are major contributors to OFC prevalence. Environmental factors including parental nutrition, pharmaceutical use, occupational and domestic chemical exposures, demographics, behavior, and health can influence the risk for OFCs. Epigenetic modification through processes such as DNA methylation and histone modifications, as well as microRNA expression, is an emerging mechanism of importance for orofacial development. We conclude with perspectives and future directions of OFC research.
Article
Meningioma-1 is a transcription activator that regulates mammalian palate development and is required for appropriate osteoblast proliferation, motility, differentiation, and function. Microdeletions involving the MN1 gene have been linked to syndromes including craniofacial anomalies, such as Toriello–Carey syndrome. Recently, truncating variants in the C-terminal portion of the MN1 transcriptional factor have been linked to a characteristic and distinct phenotype presenting with craniofacial anomalies and partial rhombencephalosynapsis, a rare brain malformation characterized by midline fusion of the cerebellar hemispheres with partial or complete loss of the cerebellar vermis. It has been called MN1 C-terminal truncation (MCTT) syndrome or CEBALID (Craniofacial defects, dysmorphic Ears, Brain Abnormalities, Language delay, and Intellectual Disability) and suggested to be caused by dominantly acting truncated protein MN1 instead of haploinsufficiency. As a proto-oncogene, MN1 is also involved in familial meningioma. In this study, we present a novel case of MCTT syndrome in a female patient presenting with craniofacial anomalies and rhombencephalosynapsis, harboring a de novo pathogenic variant in the MN1 gene: c.3686_3698del, p.(Met1229Argfs*87).
Article
Full-text available
This study examined the association between markers in transforming growth factor alpha (TGFA) and isolated, non-syndromic cleft lip with/without palate (CL/P) using a case–parent trio design, considering parent-of-origin effects. We also tested for gene–environmental interaction with common maternal exposures, and for gene–gene interaction using markers in TGFA and another recognized causal gene, IRF6. CL/P case–parent trios from four populations (76 from Maryland, 146 from Taiwan, 35 from Singapore, and 40 from Korea) were genotyped for 17 single nucleotide polymorphisms (SNPs) in TGFA. The transmission disequilibrium test was used to test individual SNPs, and the parent-of-origin likelihood ratio test (PO-LRT) was used to assess parent-of-origin effects. We also screened for possible gene–environment interaction using PBAT, and tested for gene–gene interaction using conditional logistic regression models. When all trios were combined, four SNPs showed significant excess maternal transmission, two of which gave significant PO-LRT values [rs3821261: P=0.004 and OR(imprinting)=4.17; and rs3771475: P=0.027 and OR(imprinting)=2.44]. Haplotype analysis of these two SNPS also supported excess maternal transmission. We saw intriguing but suggestive evidence of G×E interaction for several SNPs in TGFA when either individual SNPs or haplotypes of adjacent SNPs were considered. Thus, TGFA appears to influence risk of CL/P through unconventional means with an apparent parent-of-origin effect (excess maternal transmission) and possible interaction with maternal exposures.
Article
Full-text available
Oral clefts are clinically and genetically heterogeneous disorders that are influenced by both genetic and environmental factors. The present family-based association study investigated the role of the MSX1 and TGFB3 genes in the etiology of non-syndromic oral cleft in a Malay population. No transmission distortion was found in the transmission disequilibrium analysis for either MSX1-CA or TGFB3-CA intragenic markers, whereas TGFB3-CA exhibited a trend to excess maternal transmission. In sequencing the MSX1 coding regions in 124 patients with oral cleft, five variants were found, including three known variants (A34G, G110G and P147Q) and two novel variants (M37L and G267A). The P147Q and M37L variants were not observed in 200 control chromosomes, whereas G267A was found in one control sample, indicating a very rare polymorphic variant. Furthermore, the G110G variant displayed a significant association between patients with non-syndromic cleft lip, with or without cleft palate, and normal controls (P=0.001, odds ratio=2.241, 95% confidence interval, 1.357-3.700). Therefore, these genetic variants may contribute, along with other genetic and environmental factors, to this condition.
Article
Full-text available
Non-syndromic cleft lip, with or without cleft palate, is a heterogeneous, complex disease with a high incidence in the Asian population. Several association studies have been done on cleft candidate genes, but no reports have been published thus far on the Orofacial Cleft 1 (OFC1) genomic region in an Asian population. This study investigated the association between the OFC1 genomic region and non-syndromic cleft lip with or without cleft palate in 90 Malay father-mother-offspring trios. Results showed a preferential over-transmission of a 101-bp allele of marker D6S470 in the allele- and haplotype-based transmission disequilibrium test (TDT), as well as an excess of maternal transmission. However, no significant p-value was found for a maternal genotype effect in a log-linear model, although single and double doses of the 101-bp allele showed a slightly increased cleft risk (RR = 1.37, 95% CI, 0.527-3.4, p-value = 0.516). Carrying two copies of the 101-bp allele was significantly associated with an increased cleft risk (RR = 2.53, 95% CI, 1.06-6.12, p-value = 0.035). In conclusion, we report evidence of the contribution of the OFC1 genomic region to the etiology of clefts in a Malay population.
Article
We investigate the contribution of the Iberian bat fauna to the cryptic diversity in Europe using mitochondrial (cytb and ND1) and nuclear (RAG2) DNA sequences. For each of the 28 bat species known for Iberia, samples covering a wide geographic range within Spain were compared to samples from the rest of Europe. In this general screening, almost 20% of the Iberian species showed important mitochondrial discontinuities (K2P distance values > 5%) either within the Iberian or between Iberian and other European samples. Within Eptesicus serotinus and Myotis nattereri, levels of genetic divergence between lineages exceeded 16%, indicating that these taxa represent a complex of several biological species. Other well-differentiated lineages (K2P distances between 5–10%) appeared within Hypsugo savii, Pipistrellus kuhlii and Plecotus auritus, suggesting the existence of further cryptic diversity. Most unsuspected lineages seem restricted to Iberia, although two have crossed the Pyrenees to reach, at leas...
Article
Ziel der Studie: Die vorliegende Auswertung der Studie „Stillverhalten in Bayern” hat zum Ziel, das Rauchverhalten, den Alkohol- und den Koffeinkonsum von Müttern in Bayern vor, während und nach der Schwangerschaft zu beschreiben. Weiterhin wurde die Passivrauchbelastung der Schwangeren und der Säuglinge erfragt. Zudem soll der Einfluss des Genussmittelkonsums auf die Stilldauer quantifiziert und die Verbreitung des Risikoverhaltens in der Bevölkerung genauer charakterisiert werden. Methodik: Die Studie „Stillverhalten in Bayern” ist eine prospektive Kohortenstudie, die von April 2005 bis Januar 2006 durchgeführt wurde. An der Basiserhebung nahmen 3 822 Mütter aus ganz Bayern teil, die im April 2005 entbunden hatten. Methoden und erste Ergebnisse wurden bereits beschrieben. Die Studienteilnehmerinnen wurden in 4 Folgebefragungen zum Stillverhalten, zu Rauchgewohnheiten, zum Alkohol- und zum Koffeinkonsum befragt. Die Follow-Up Quote betrug 82%. In die vorliegende Auswertung wurden nur Studienteilnehmerinnen mit komplettem Follow-Up einbezogen (n=3 103). Ergebnisse: Vor der Schwangerschaft rauchten 23,7% der Mütter. Der Anteil der Frauen, die angaben, jemals in der Schwangerschaft geraucht zu haben, belief sich auf 9,8%. Ende des neunten Monats nach der Entbindung hatte bereits wieder mehr als die Hälfte der ehemaligen Raucherinnen (53%) mit dem Rauchen angefangen. Insgesamt 25,3% der Mütter gaben an, jemals in der Schwangerschaft Alkohol zu sich genommen zu haben. 69,0% der Schwangeren tranken jemals in der Schwangerschaft koffeinhaltige Getränke. Die Konsummengen wurden in der Schwangerschaft deutlich reduziert. Rauchen war dabei eher bei jüngeren Frauen mit niedrigerem Schulabschluss, Alkoholkonsum eher bei älteren Frauen mit höherem Schulabschluss verbreitet. In Deutschland geborene Mütter rauchten signifikant häufiger als Mütter mit Migrationshintergrund. Rauchen hatte einen dosisabhängigen, signifikant negativen Einfluss auf die Stilldauer von <4 vollen Monaten ausschließlichem Stillen (1–5 Zigaretten/Tag, Odds Ratio (OR) 2,04, 95%-Konfidenzintervall (KI) 1,31–3,18; >5 Zigaretten/Tag, OR 2,54, 95%-KI 1,42–4,54). Koffeinkonsum verkürzte die Stilldauer ebenfalls signifikant (OR 1,49, 95%-KI 1,25–1,79), wohingegen Alkoholkonsum die Stilldauer nicht beeinflusste. Schlussfolgerung: Alkoholkonsum, Rauchen und Koffeinkonsum in der Schwangerschaft sind weit verbreitetet. Neben den bereits etablierten Präventionsinitiativen können zusätzliche zielgruppenorientierte Maßnahmen zur Raucherentwöhnung für junge Schwangere mit niedriger Schulbildung die Raucherquoten in dieser Risikogruppe senken. Verstärkte Aufmerksamkeit sollte dem Thema Alkoholkonsum in Schwangerschaft und Stillzeit gewidmet werden.
Article
Evidence for the teratogenicity of corticosteroids in humans is limited and has resulted in inconsistent recommendations regarding their use during early pregnancy. We examined the association between women's corticosteroid use during the periconceptional period (1 month before to 3 months after conception) and delivering infants with selected congenital anomalies. Data were derived from a population-based case-control study that included cases of orafacial clefts (n = 662), conotruncal heart defects (n = 207), neural tube defects (n = 265), and limb reduction defects (n = 165). Information on medication use was collected via maternal telephone interviews. Corticosteroid use was associated with an increased risk for isolated cleft lip with or without cleft palate (odds ratio 4.3, 95% confidence interval 1.1–17.2) and isolated cleft palate (odds ratio 5.3, 95% confidence interval 1.1–26.5). Increased risks were not observed for the other anomaly groups studied. These data in conjunction with other epidemiologic data suggest a possible causal association between cleft lip and palate and corticosteroid use. Am. J. Med. Genet. 86:242–244, 1999. © 1999 Wiley-Liss, Inc.
Article
Objective: Velo-cardio-facial syndrome (VCFS) is caused by a microdeletion of approximately 40 genes from one copy of chromosome 22. Expression of the syndrome is a variable combination of over 190 phenotypic characteristics. As of yet, little is known about how these phenotypes correlate with one another or whether there are predictable patterns of expression. Two of the most common phenotypic categories, congenital heart disease and cleft palate, have been proposed to have a common genetic relationship to the deleted T-box 1 gene (TBX1). The purpose of this study is to determine if congenital heart disease and cleft palate are correlated in a large cohort of human subjects with VCFS. Methods: This study is a retrospective chart review including 316 Caucasian non-Hispanic subjects with FISH or CGH microarray confirmed chromosome 22q11.2 deletions. All subjects were evaluated by the interdisciplinary team at the Velo-Cardio-Facial Syndrome International Center at Upstate Medical University, Syracuse, NY. Each combination of congenital heart disease, cleft palates, and retrognathia was analyzed by Chi square or Fisher exact test. Results: For all categories of congenital heart disease and cleft palate or retrognathia no significant associations were found, with the exception of submucous cleft palate and retrognathia (nominal p=0.0325) and occult submucous cleft palate and retrognathia (nominal p=0.000013). Conclusions: Congenital heart disease and cleft palate do not appear to be correlated in human subjects with VCFS despite earlier suggestions from animal models. Possible explanations include modification of the effect of TBX1 by genes outside of the 22q11.2 region that may further influence the formation of the palate or heart, or the presence of epigenetic factors that may effect genes within the deleted region, modifying genes elsewhere, or polymorphisms on the normal copy of chromosome 22. Lastly, it is possible that TBX1 plays a role in palate formation in some species, but not in humans. In VCFS, retrognathia is caused by an obtuse angulation of the skull base. It is unknown if the correlation between retrognathia and cleft palate in VCFS indicates a developmental sequence related to skull morphology, or direct gene effects of both anomalies. Much work remains to be done to fully understand the complex relationships between phenotypic characteristics in VCFS.
Article
Jagomägi T, Nikopensius T, Krjutškov K, Tammekivi V, Viltrop T, Saag M, Metspalu A. MTHFR and MSX1 contribute to the risk of nonsyndromic cleft lip/palate. Eur J Oral Sci 2010; 118: 213–220. © 2010 The Authors. Journal compilation©2010 Eur J Oral Sci Recent studies suggest that multiple interacting loci, with possible additional environmental factors, influence the risk for nonsyndromic oral clefts, one of the most common birth defects in humans. Advances in high-throughput genotyping technology allow the testing of multiple markers, simultaneously, in many candidate genes. We tested for associations between 176 haplotype-tagging single nucleotide polymorphisms (SNPs) in 18 candidate genes/loci and nonsyndromic clefts in a case–control study in an Estonian sample (153 patients, 205 controls). The most significant associations with nonsyndromic cleft lip with or without cleft palate (CL/P) were found for SNPs in MSX1, MTHFR, and PVRL2, including several common haplotypes in the MTHFR and MSX1 genes. The strongest association was observed for rs6446693 in the MSX1 region, which remained statistically significant after Bonferroni correction. The strongest association with nonsyndromic cleft palate (CP) was found for the SNP rs11624283 in the JAG2 gene. Epistatic interactions were observed for SNPs within PVRL2, between BCL3 and EDN1, and between IRF6 and MSX1 genes. This study provides further evidence implicating MSX1 and MTHFR in the etiology of nonsyndromic CL/P across different populations.