ArticlePDF Available

Sequence analysis of 21 genes located in the Kartagener syndrome linkage region on chromosome 15q

Authors:
  • Intitute of Human Genetics, Polish Academy of Sciences, Poznan, Poland

Abstract and Figures

Primary ciliary dyskinesia (PCD) is a rare genetic disorder, which shows extensive genetic heterogeneity and is mostly inherited in an autosomal recessive fashion. There are four genes with a proven pathogenetic role in PCD. DNAH5 and DNAI1 are involved in 28 and 10% of PCD cases, respectively, while two other genes, DNAH11 and TXNDC3, have been identified as causal in one PCD family each. We have previously identified a 3.5 cM (2.82 Mb) region on chromosome 15q linked to Kartagener syndrome (KS), a subtype of PCD characterized by the randomization of body organ positioning. We have now refined the KS candidate region to a 1.8 Mb segment containing 18 known genes. The coding regions of these genes and three neighboring genes were subjected to sequence analysis in seven KS probands, and we were able to identify 60 single nucleotide sequence variants, 35 of which resided in mRNA coding sequences. However, none of the variations alone could explain the occurrence of the disease in these patients.
Content may be subject to copyright.
ARTICLE
Sequence analysis of 21 genes located in the
Kartagener syndrome linkage region on chromosome
15q
Maciej Geremek
1,2
, Frederieke Schoenmaker
1
, Ewa Zietkiewicz
2
, Andrzej Pogorzelski
3
,
Scott Diehl
4
, Cisca Wijmenga*
,1,5,7
and Michal Witt
2,6,7
1
Complex Genetics Group, Department of Biomedical Genetics, University Medical Center Utrecht, Utrecht, the
Netherlands;
2
Department of Molecular and Clinical Genetics, Institute of Human Genetics, Poznan, Poland;
3
Institute
of Tuberculosis and Lung Diseases, Rabka, Poland;
4
New Jersey Dental School, University of Medicine and Dentistry
New Jersey, Newark, NJ, USA;
5
Department of Genetics, University Medical Center Groningen, University of Groningen,
Groningen, the Netherlands;
6
International Institute of Molecular and Cell Biology, Warsaw, Poland
Primary ciliary dyskinesia (PCD) is a rare genetic disorder, which shows extensive genetic heterogeneity
and is mostly inherited in an autosomal recessive fashion. There are four genes with a proven pathogenetic
role in PCD. DNAH5 and DNAI1 are involved in 28 and 10% of PCD cases, respectively, while two other
genes, DNAH11 and TXNDC3, have been identified as causal in one PCD family each. We have previously
identified a 3.5 cM (2.82 Mb) region on chromosome 15q linked to Kartagener syndrome (KS), a subtype of
PCD characterized by the randomization of body organ positioning. We have now refined the KS
candidate region to a 1.8 Mb segment containing 18 known genes. The coding regions of these genes and
three neighboring genes were subjected to sequence analysis in seven KS probands, and we were able to
identify 60 single nucleotide sequence variants, 35 of which resided in mRNA coding sequences. However,
none of the variations alone could explain the occurrence of the disease in these patients.
European Journal of Human Genetics (2008) 16, 688 695; doi:10.1038/ejhg.2008.5; published online 13 February 2008
Keywords: Kartagener syndrome; PCD; gene mapping; sequencing; mutation screening
Introduction
Primary ciliary dyskinesia (PCD, MIM no. 242650) is a rare
genetic disorder caused as far as is known by mutations
in genes encoding proteins important for ciliary beating.
1
The major clinical consequences fall into three categories:
(1) in the respiratory tract, immotility or dyskinetic beating
of cilia in epithelial cells impairs the mucociliary clearance,
leading to recurrent infections, sinusitis and bronchiectasis;
(2) in the urogenital tract, infertility occurs in some
patients because of the dysmotility of sperm tails; (3)
mirror reversal of body organ positioning (situs inversus) is
seen in approximately half of PCD patients. It is believed to
occur at random and to be caused by the immotility of
primary cilia of cells on the ventral surface of the
embryonic node. The presence of situs inversus in an
affected family defines a PCD subtype, known as
Kartagener syndrome (KS, MIM no. 244400).
For diagnostic purposes, cilia in bronchial scrapings can
be visualized under a light microscope, whereas the clinical
diagnosis of PCD is routinely verified by electron micro-
scopy analysis of respiratory cilia ultrastructure. About 20
different ultrastructural defects have been described
Received 8 June 2007; revised 11 December 2007; accepted 6 January
2008; published online 13 February 2008
*Correspondence: Professor C Wijmenga, Department of Genetics,
University Medical Center Groningen, University of Groningen, PO Box
30001, RB Groningen 9700, the Netherlands.
Tel: þ31 50 3617 100; Fax: þ31 50 3617 230;
E-mail: c.wijmenga@umcutrecht.nl
7
These authors have contributed equally to this work.
European Journal of Human Genetics (2008) 16, 688 695
&
2008 Nature Publishing Group All rights reserved 1018-4813/08
$30.00
www.nature.com/ejhg
in PCD patients, with lesions of outer and/or inner dynein
arms being the most frequent defects of the internal
anatomy of cilia.
2,3
Recently, it has been shown that
immunofluorescence staining of ciliated epithelium with
antibodies targeting DNAH5 can detect outer dynein arms
defects and therefore aid diagnosis of PCD.
4
Inheritance of PCD is autosomal recessive in most cases,
although pedigrees with an X-linked mode of inheritance
have also been described.
5,6
To date, mutations in four
genes have been found as causative for PCD, exclusively in
patients with outer dynein arm defects. DNAI1, coding
for intermediate dynein chain 1, was selected for mutation
screening based on homology with the Chlamydomonas
reinhardtii gene, which causes a slow-swimming phenotype
when mutated,
7,8
whereas DNAH5, coding for heavy
dynein chain 5, was identified by homozygosity mapping
in a large family and subsequent sequence analysis.
9
DNAH5 and DNAI1 are responsible for 28 and 10% of
PCD cases, respectively,
10,11
clearly indicating that other
genes are involved in PCD/KS etiology. DNAH11,
12
coding
for dynein heavy chain, and TXNDC3,
13
coding for a
thioredoxin family member, were also found to be
responsible for PCD, but so far each has been found in
only one family.
Analysis of the ciliary proteome in unicellular alga
C. reinhardtii led to the identification of a number of proteins
potentially involved in cilia formation and maintenance:
360 proteins with high and 293 with moderate confi-
dence.
14
Besides the known components of ciliary ultra-
structure, there were also 90 signal transduction proteins,
and many membrane proteins and metabolic enzymes.
Given that our knowledge of ciliary genes comes mostly
from studies of simple organisms and that the functional
spectrum of the possible candidate genes is very wide, it is
difficult to predict which of these should be considered
as candidate genes for PCD.
On the basis of genetic linkage studies, extensive genetic
heterogeneity of PCD is postulated. A total-genome scan
performed in 31 multiplex families did not reveal any
predominant locus, rather it showed several peaks with
suggestive and indicative LOD scores.
15
Another genome
scan, performed in five families of Arabic origin and with
reported consanguinity, revealed linkage with chromo-
some 19q13.3 in only three of the families, thus
confirming locus heterogeneity.
16
Two additional loci, on
16p12 and 15q, were indicated by studies on genetically
isolated, but heterogeneous, populations from the Faroe
Islands and the Israeli Druze.
17
We earlier performed genome-wide linkage analysis in 52
families with KS and reported a region on chromosome
15q24 25, between D15S973 and D15S1037, linked to
KS.
18
After obtaining a significant LOD score of 4.34 with
D15S154 (using a strictly defined disease model), we
defined a region of 3.5 cM (2.82 Mb) as containing the
candidate gene.
18
In the present study, we have performed
further fine-mapping of the region of interest and screened
all 18 genes in the linked region and three neighboring
genes for mutations in seven KS patients from the families
that contributed most to the linkage results.
18
Materials and methods
Biological material
DNA samples from the existing PCD collection
18
came
from 31 Caucasian families (25 of Polish and 6 of Slovak
origin) classified as KS families; none was from a genetically
isolated population. Each family had at least one
member diagnosed with PCD and exhibiting situs inversus,
but with no other major anomalies or dysmorphologies
present. The group studied consisted of 38 affected
individuals and 99 unaffected family members. The
primary complaints in these patients were derived from
the respiratory tract and the clinical picture included
symptoms of sinusitis, nasal polyps, bronchiectasis, recur-
rent infections of the upper respiratory tract. All cases were
confirmed by a low concentration of nitric oxide (NO)
measured in exhaled air.
19
NO was measured from the
nasal cavity by a chemiluminescence analyzer with a
threshold value of 200 ppb for diagnosing PCD. In 60%
of the families, the clinical diagnosis was confirmed by
transmission electron microscopy analysis of bronchial
cilia ultrastructure; the analysis revealed defects of outer or
outer and inner dynein arms; a representative EM picture is
shown in Supplementary Figure 1. Direct microscopy of
bronchial scrapings was used to confirm the clinical
diagnosis of PCD in the patients that did not have electron
microscopy imaging of cilia. KS inheritance in these
pedigrees was in agreement with general assumptions for
an autosomal recessive disease model. KS in these families
has previously been shown to be linked with the region
between D15S973 and D15S1037, assuming full pene-
trance, 0.0001 phenocopies and the disease allele
frequency of 0.005. The genome-wide significant LOD
score of 4.34 was obtained using a cohort of 52 families.
18
Pair-wise LOD score analyses were performed using the
FASTLINK program, and multipoint LOD scores for markers
on chromosome 15 were calculated using GENEHUNTER.
Multipoint LOD scores allowing for locus heterogeneity
were calculated using Simwalk2. The 95% confidence
interval localized the KS locus to a 3.5 cM region between
D15S973 and D15S1037.
Recombination mapping using microsatellite markers
To refine the candidate gene region between D15S973
and D15S1037, seven markers (D15S1027–380 kb–
D15S524–33 kb–AFMA085AWB9–580 kb–D15S989–219 kb–
D15S1005–58 kb–D15S969–695 kb–D15S551), were geno-
typed in all 38 affected individuals and 99 unaffected
family members. Amplification products, obtained using
50end fluorescent-dye-labeled primers, were pooled and
Sequencing of 21 genes in 15q KS locus
M Geremek et al
689
European Journal of Human Genetics
analyzed on an automated ABI Prism 3730 DNA sequencer
(Applied Biosystems, Foster City, CA, USA). Genotypes
were scored using GENMAPPER (v3.5NT) software. Inheri-
tance was checked using PEDCHECK
20
and haplotypes in
the 31 pedigrees were constructed from microsatellite data
using SIMWALK2.
21
Analysis of candidate genes
Sequence analysis was performed in seven affected indivi-
duals from the families with the highest LOD score
for marker D15S1005.
18
Sequences of 135 exons and
surrounding intron sequences of the 20 genes located in
the region were obtained using Ensembl genome browser
version 42 (Figure 1; Table 1). AL109678, not present in
Ensembl, was also sequenced as it is an mRNA coding
sequence annotated in the UCSC genome browser. At least
200 bp upstream of the transcription start site were also
analyzed and the 30UTR was sequenced in all genes except
for TMC3,KIAA1199,ARNT2 and TMED3. The Primer3
program was used for designing PCR primers yielding
products of 400 500 bp;
22
primers were in the intronic
sequences flanking the exons to be analyzed. PCR reactions
were performed using standard protocols (data available
upon request). Amplification products were purified,
quantified on a 2% agarose gel, and diluted for direct
sequencing on an automated ABI Prism 3730 DNA
sequencer using BigDye Terminator Sequencing Standards
(Applied Biosystems). Sequences were assembled using
ContigExpress (Vector NTI suite v8.0, Informax Inc.).
Sequencing reactions were performed first with the
forward primer, but if the quality of the sequencing
electropherogram was unsatisfactory, or if single-nucleo-
tide polymorphisms (SNPs) were found, they were repeated
with the reverse primer. Each of the newly identified
variants was validated through an independent PCR and
sequencing reaction. The sequenced fragments were only
included in the results if data were obtained for at least six
of the seven patients. Seven reactions could not be
amplified despite at least three attempts with two different
sets of primers (nucleotides in mRNA sequence: BCL2A1,
602 710; FAM108C1, 1 343; C15ORF37, 421 –640;
TMED3, 260 299; ENSG00000180725, 457 618; FAH,
683 783, 991 1037). Reactions failed to use control as
well as patients’ DNA, indicating PCR problems related to
the local sequence composition of those genomic frag-
ments. All variations found in the protein-coding sequence
and not previously described in public databases were
sequenced in a control population group consisting of 48
unrelated, healthy individuals of Polish origin.
Results
Seven microsatellite markers located in the 3.5 cM region
reported to be linked to KS were genotyped in 31 families
that had positive LOD scores for chromosome 15q in our
genome linkage scan
18
(Figure 1). Haplotype analysis was
concordant with linkage to chromosome 15q24 25 (data
not shown). A recombination was mapped in family
numbers 114 and 126 (Figure 2), 1 Mb more telomerically
from the previous location defined by marker D15S937,
thus narrowing the minimal gene-containing region
down to 1.8 Mb, between markers AFMA085AWB9 and
D15S1037.
The analysis of haplotype sharing between nonrelated KS
patients was inconclusive. The most promising block,
encompassing three markers (D15S524, AFMA085WB9
and D15S989) and spanning 0.6 Mb, was present on five
nonrelated, affected chromosomes (families 103, 117, 136,
142 and 147) and on one untransmitted parental chromo-
some (data not shown).
Seven families with the highest positive LOD scores were
selected for sequence analysis of all the genes located in
TMC3
LOC390616
TMED3
K1024
MTHFS
Q9HBF5
C15ORF3
BCL2A1
ZFAND6
FAH
ENSG00000196548
ENSG00000205296 Q8N817 MESDC2
AL109678
ARNT2
FAM108C1
KIAA1199
MESDC1
C15ORF26
ENSG00000180725
D15S524
AFMA085WB9 D15S989 D15S969
D15S1005
D15S1037
D15S551
100kb
Figure 1 The location of 21 genes in the linkage region on chromosome 15q24-25. The minimal KS gene-containing region is indicated by an
arrow.
Sequencing of 21 genes in 15q KS locus
M Geremek et al
690
European Journal of Human Genetics
the 1.8 Mb region (Figure 2). These families contributed
70% of the LOD score for marker D15S1005 on chromo-
some 15q,
18
and it was interesting that the DNAI1
mutation analysis was negative in all seven families. The
DNAH5 contribution was excluded in four families (114,
117, 136 and 147; Figure 2), based on the analysis of SNPs
located within the gene (discordant genotypes in the
affected sibs), while in the three remaining families, the
analyzed SNPs in DNAH5 were uninformative (unpub-
lished data). As the microsatellite data confirmed linkage of
the disease to 15q24 25 in all seven families, only one
affected individual, representative of each pedigree, was
subjected to DNA sequencing. Supplementary Table 2
summarizes the clinical findings of the seven families
included for sequencing.
We analyzed the 18 coding sequences located between
markers AFMA085WB9 and D15S1037 (Table 1). We also
sequenced three genes located in the vicinity of the
region, which could be functionally related to cilia:
MESDC2, which is important in embryonic development;
C15ORF26, which is expressed in lung and testis and
displays homology to a flagellar protein identified in
Chlamydomonas;
14
and TMC3, a membrane protein with
partial homology to nephrocystine, as indicated by Blast
search of the NCBI nr database (Figure 1). Except for
C15ORF26, none of the genes have been reported to be a
part of cilium.
In total, we analyzed 83 kb of genomic sequence in seven
individuals (35 kb of coding sequences and 48 kb of
intronic/flanking DNA). We found 60 SNPs, 45 of which
were already present in public SNP databases. Twelve SNPs,
including the four previously unknown ones, were located
in the protein coding regions. None of the identified SNPs
in the coding regions resulted in a stop codon change or
Table 1 Genes included in the Kartagener syndrome sequencing study
Gene symbol Gene name GO description No. of exons
Transcript
length (bp)
LOC390616 Unknown Unknown 1 1608
AL109678 Unknown Unknown 1 2540
TMED3 Transmembrane emp24 domain-containing
protein 3 precursor (membrane protein p24B)
Protein transport; membrane;
membrane; protein carrier activity
3 1420
K1024 UPF0258 protein KIAA1024 Membrane; integral to membrane 4 6732
MTHFS 5-formyltetrahydrofolate cyclo-ligase (EC 6.3.3.2)
(5,10-methenyl- tetrahydrofolate synthetase)
(methenyl-THF synthetase) (MTHFS)
Magnesium ion binding; folic acid
binding; ligase activity; metabolism
3 857
ENS-
G00000205296
Unknown Unknown 2 153
Q9HBF5 Cervical cancer suppressor-1 Unknown 2 240
C15ORF37 C15orf37 protein Unknown 1 2082
BCL2A1 Bcl-2-related protein A1 (protein BFL-1)
(hemopoietic-specific early response protein)
(protein GRS)
Regulation of apoptosis 3 624
ENS-
G00000180725
Unknown Unknown 2 698
ZFAND6 Zinc-finger A20 domain-containing protein 3
(associated with PRK1 protein)
DNA binding; zinc ion binding;
signal transduction; protein
binding
7 1680
ENS-
G00000196548
Unknown Unknown 3 123
FAH Fumarylacetoacetase (EC 3.7.1.2)
(fumarylacetoacetate hydrolase) (beta-diketonase)
(FAA)
Fumarylacetoacetase activity;
magnesium ion binding; regulation
of transcription, DNA-dependent;
metabolism
15 1471
ARNT2 Aryl hydrocarbon receptor nuclear translocator 2
(ARNT protein 2)
Transcription factor activity; signal
transducer activity; aryl
hydrocarbon receptor nuclear
translocator activity; response to
hypoxia; embryonic development
19 6550
Q8N817 CDNA FLJ40133 fis, clone TESTI2012231 Unknown 4 675
FAM108C1 Unknown Unknown 3 2326
KIAA1199 KIAA1199 Sensory perception of sound 29 6602
MESDC2 Mesoderm development candidate 2 (NY-REN-61
antigen)
Mesoderm development 5 2938
MESDC1 Mesoderm development candidate 1 Unknown 1 2373
C15ORF26 Uncharacterized protein C15orf26 Unknown 7 1574
TMC3 Transmembrane channel-like protein 3 Integral to membrane 19 2443
Sequencing of 21 genes in 15q KS locus
M Geremek et al
691
European Journal of Human Genetics
Figure 2 Pedigrees of the KS families included in the sequencing. The results of the pair-wise LOD score for marker D15S1005 are indicated below
each family. The recombination (marked by a red arrow) defining the minimal KS gene-containing region can be observed in pedigrees 114 and 126.
Individuals included in the sequencing are indicated by black arrows. Red dots indicate individuals exhibiting situs inversus. The numbers beside
marker names indicate the intermarker distances in kilobase pairs. Missing data are denoted by zeros. The genotypes of the parents in pedigrees 117
and 126 were deduced from the genotypes of the siblings.
Sequencing of 21 genes in 15q KS locus
M Geremek et al
692
European Journal of Human Genetics
Table 2 List of all sequence variations found in 7 Kartagener syndrome patients
Gene Variation Amino acid Exon\intron
ID in SNP
database
Minor allele
frequency in NCBI
genome
database 35
Minor allele
frequency in 48
control individuals
of Polish origin
Number of patient
chromosomes with
the minor allele
LOC39016 c.735C4T P245P Exon 1 rs8038778 T ¼0.367 7/14
TMED c.659G4A A172A Exon 3 rs906439 A ¼0.0 0/14
KIAA1024 c.847A4C N258H
a
Exon 2 FC¼0.06 2/14
c.918C4A A281A Exon 2 F1/14
c.2569A4G I832V
b
Exon 3 rs2297773 G ¼0.208 9/14
c*1799A4CF30UTR rs17266017 C ¼0.375 2/14
c.*3771_3772insATG F30UTR FIns ¼0.06 2/14
BCL2A1 c.238G4A C19Y
c
Exon 1 rs1138357 A ¼0.146 3/14
c.299T4G N39K
d
Exon 1 rs1138358 G ¼0.146 3/14
MTHFS g.411A4GFUpstream rs2865825 F1/14
g.421C4GFUpstream rs2903105 F3/14
g.483delG FUpstream rs35401897 F2/14
C15ORF37 c.903C4A P166Q
f
Exon 1 rs3803540 A ¼0.217 1/14
c.*46G4TF30UTR rs17214656 T ¼0.167 2/14
c.*124A4CF30UTR rs36070199 F2/14
c.*206A4GF30UTR rs12442408 G ¼0.250 1/14
c.*413A4TF30UTR rs2733101 T ¼0.283 2/14
AWP1 c.29590T4CFIntron 2 rs1916048 C ¼0.217 5/12
C574+11C4GFIntron 5 rs1522636 G ¼0.261 9/14
c.46660A4TFIntron 4 rs12915043 F1/14
c.506A4G Q65Q Exon 4 FG¼0.0 1/14
FAH c.915173delC FIntron 10 rs3835063 F6/14
c.1133C4T S352S Exon 13 rs1801374 T ¼0.058 2/14
c.103835A4CFIntron 12 rs2043691 C ¼0.08 1/14
c.1038223C4TFIntron 12 rs2043692 T ¼0.15 1/14
c.*37A4CF30UTR rs1049181 F2/14
c.*38C4TF30UTR FF 2/14
c.*39C4TF30UTR rs3210172 F2/14
c.*40C4TF30UTR FF 2/14
c.*41C4GF30UTR FF 2/14
c.*50delC F30UTR FF 2/14
c. *93T4CF30UTR rs1049194 F2/14
ARNT2 c.312+15C4TFIntron 3 rs2278708 T ¼0.058 1/14
c.36128insC FIntron 3 rs5814026 F3/14
C574+36C4GFIntron 4 rs8033507 G ¼0.017 1/14
c.89274C4TFIntron 7 rs11072922 T ¼0.175 2/14
c.1255+21C4TFIntron 11 rs3924894 T ¼0.405 8/14
c.1255+22T4CFIntron 11 rs3924893 C ¼0.4 6/14
c.1279T4A W399R
e
Exon 12 - A ¼0.0 1/14
FAM108C1 c.*692T4CF30UTR rs1046417 C ¼0.130 2/14
c.*1115C4AF30UTR rs11072940 A ¼0.104 2/14
KIAA1199 c.255T4AF50UTR rs2273886 A ¼0.295 4/14
c.1029+26C4GFIntron 7 FF 7/14
c.1346+240C4TFIntron 9 FF 1/14
c.1479+83T4AFIntron 10 FF 1/14
c.2057+57A4GFIntron 13 rs2271164 G ¼0.417 5/12
c.2142T4C C28C Exon 14 rs2271160 C ¼0.261 3/14
c.2462A4CFIntron 16 rs2271162 C ¼0.167 3/14
MESDC2 c.*3G4TF30UTR rs8039607 T ¼0.121 1/14
c.*53G4TF30UTR rs8039384 T ¼0.142 1/14
c.136920_136918del-
CTT
FIntron 4 FDel ¼0.08 1/14
c.*1416C4TF30UTR FT¼0.069 1/14
c.*2153T4CF30UTR rs10519305 C ¼0.017 1/14
C15ORF26 G496C4GFUpstream rs2683255 F1/14
c.505105insA FIntron 3 FF 3/14
c.50587C4G Intron 3 FF 3/14
c.66450insA Intron 5 rs11395672 F7/14
TMC3 c.13T4CF50UTR FC¼0.08 1/14
c.580+55T4C Intron 6 rs12903128 C ¼0.345 1/14
c.82366A4G Intron 7 rs7167034 A ¼0.153 7/14
Nucleotide variations are presented according to the HUGO nomenclature. Translation start site was used for each gene as the +1 nucleotide. Putative
effect on protein:
a
Small, polar4polar, charged, positive, aromatic;
b
Aliphatic4aliphatic, small;
c
Tiny4aromatic;
d
Small4positive;
e
Aromatic, polar4polar, positive;
f
Small4polar.
Sequencing of 21 genes in 15q KS locus
M Geremek et al
693
European Journal of Human Genetics
a frame shift; six were amino-acid substitutions and seven
were synonymous codon changes. Twenty-one SNPs,
including seven novel ones, were located in the untrans-
lated regions. Four SNPs were located in regions upstream
of the transcription start site. Twenty-seven variations were
located in noncoding regions, six of which were previously
unknown. None of the SNPs found were closer than 10 bp
to a splicing site. Variations in the protein-coding
sequence, previously not described in SNP databases, were
sequenced in 48 individuals of Polish origin to estimate
population frequencies. Except for two SNPs (W399R in
ARNT2 and A506G in AWP1), all the variations were found
in the healthy controls. A complete list of the SNPs found is
given in Table 2 and the distribution of all SNPs in the
patients can be found in Supplementary Table 1.
Discussion
We have performed a follow-up study to our previous
finding
18
of linkage to chromosome 15q24 25 in families
with KS. Fine-mapping of the 3.5 cM (2.82 Mb) linked
region in 31 KS families allowed us to narrow down the
minimal candidate gene region to 1.8 Mb. Analysis of
haplotypes composed of microsatellite markers revealed
that there was no extensive haplotype sharing among the
affected individuals, arguing against the presence of only
one or a few frequent mutations occurring on a common
haplotype background. On the other hand, the absence of
haplotype sharing does not exclude the possibility of the
presence of a causative gene or genes in the linked region.
The linkage analysis is independent of a founder effect.
There are examples in literature of genes involved in rare
recessive diseases, but not showing evidence of haplotype
sharing. This happens because most of the patients are
compound heterozygous for unique mutations in the given
gene.
23
In searching for the KS gene in the 15q24-25 region, we
found 37 SNPs in the mRNA-coding sequences and
promoters of 14 genes, including 6 nonsynonymous codon
changes. No evidently pathogenic mutations, such as the
introduction of stop codons or frame-shift mutations, were
detected.
Given the rare occurrence of the disease (approximately
1:40 000), the expected cumulative frequency of all the
disease variant(s) in the general population should be
below 1%. We also sequenced fragments harboring variants
in mRNA-coding sequence in 96 control chromosomes
(Table 2). One nonsynonymous amino-acid change
(W399R in ARNT2), and one silent codon change (A506G
in AWP1) were not seen in the controls, suggesting that
their frequencies might be as low as 1%. However, each of
these two variants (in two different genes) was found in
heterozygous form in only a single patient, with no
other possibly pathogenic complementary allele; these
are therefore unlikely to represent the disease-causing
mutations. None of the remaining SNPs were particularly
rare and their occurrence in patients did not significantly
deviate from population frequencies (Table 2).
In conclusion, we did not find any variants that directly
fulfilled the criteria of a pathogenic mutation for KS in
any of the 21 genes located in the minimal gene-contain-
ing region, as indicated by linkage analysis (while
assuming the rare occurrence of the disease and the
recessive mode of inheritance). The analyzed region may,
however, contain pathogenic variants that we failed to
find. We did not sequence introns that could contain
regulatory elements or harbor gain-of-splicing site muta-
tions. It is also possible that some of the genes have
alternative exons not annotated in the current databases
and thus not analyzed in our study, and we cannot exclude
the possibility that some of the genes harbor heterozygous
deletion(s) of one or more exons. We did not detect any
extended homozygous regions, in neither the micro-
satellite data nor in the SNP data, but the relatively low
density of heterozygous SNPs could have left some regions
noninformative with respect to their purported hemi-
zygosity (see Supplementary Table 1).
As the pedigree structures in the cohort we studied were
simple, we cannot exclude the possibility that some of the
families are false positives, although none of the families
was crucial for the significant LOD score. Hence, despite
the statistical significance of the LOD score value, it cannot
entirely be excluded that the 15q locus is a false positive
locus. The current location of the linkage interval is based
on two families (114 and 126; Figure 2) that contribute to
the left recombination and on one family
18
that contri-
butes to the right recombination. This last family contains
only one affected individual and, therefore, contributed
very little to the overall LOD score. Under less stringent
disease models, the linkage peak would still be significant
with a 95% confidence interval matching the minimal
gene-containing region. However, the linkage curve would
be positive over a larger region with more than 200 genes.
Hence, we cannot exclude that the KS gene is located closer
to the 15q telomere. Given that this larger region does not
contain obvious candidate genes, more families will be
needed to establish the exact KS region.
Acknowledgements
We thank all the Polish and Slovak PCD families who participated in
this study for their invaluable cooperation, and J Pawlik (Rabka,
Poland) and A Kapellerova (Bratislava, Slovakia) for providing study
material and clinical data. We gratefully acknowledge the help of Ewa
Rutkiewicz in collecting biological material, of Jackie Senior in editing
the manuscript and of Behrooz Alizadeh in interpreting the statistical
data. This research was partly financed by Grant PBZ-KBN-122/
P05-1 from the Polish Ministry of Science and Higher Education. MG
is supported by the International PhD Program of the University of
Utrecht.
Sequencing of 21 genes in 15q KS locus
M Geremek et al
694
European Journal of Human Genetics
References
1 Geremek M, Witt M: Primary ciliary dyskinesia: genes,
candidate genes and chromosomal regions. J Appl Genet 2004;
3: 347 361.
2 Zariwala MA, Knowles MR, Omran H: Genetic defects in ciliary
structure and function. Annu Rev Physiol 2007; 69: 423– 450.
3 Van’s Gravesande KS, Omran H: Primary ciliary dyskinesia:
clinical presentation, diagnosis and genetics. Ann Med 2005; 37:
439 449.
4 Fliegauf M, Olbrich H, Horvath J et al: Mislocalization of DNAH5
and DNAH9 in respiratory cells from patients with primary ciliary
dyskinesia. Am J Respir Crit Care Med 2005; 171: 1343 1349.
5 Krawczynski MR, Witt M: PCD and RP: X-linked inheritance of
both disorders? Pediatr Pulmonol 2004; 38: 88– 89.
6 Moore A, Escudier E, Roger G et al: RPGR is mutated in patients
with a complex X linked phenotype combining primary ciliary
dyskinesia and retinitis pigmentosa. J Med Genet 2006; 43:
326 333.
7 Guichard C, Harricane M-C, Lafitte J-J et al: Axonemal dynein
intermediate-chain gene (DNAI1) mutations result in situs
inversus and primary ciliary dyskinesia (Kartagener syndrome).
Am J Hum Genet 2001; 68: 1030 1035.
8 Pennarun G, Escudier E, Chapelin C et al: Loss-of-function
mutations in a human gene related to Chlamydomonas reinhardtii
dynein IC78 result in primary ciliary dyskinesia. Am J Hum Genet
1999; 65: 1508 1519.
9 Olbrich H, Ha
¨ffner K, Kispert A et al: Mutations in DNAH5 cause
primary ciliary dyskinesia and randomization of left-right
asymmetry. Nat Genet 2002; 30: 143 144.
10 Hornef N, Olbrich H, Horvath J: DNAH5 mutations are a
common cause of primary ciliary dyskinesia with outer dynein
arm defects. Am J Respir Crit Care Med 2006; 174: 120 126.
11 Zariwala MA, Leigh MW, Ceppa F: Mutations of DNAI1 in
primary ciliary dyskinesia: evidence of founder effect in a
common mutation. Am J Respir Crit Care Med 2006; 174:
858 866.
12 Bartoloni L, Blouin JL, Pan Y et al: Mutations in the DNAH11
(axonemal heavy chain dynein type 11) gene cause one form
of situs inversus totalis and most likely primary ciliary dyskinesia.
Proc Natl Acad Sci 2002; 99: 10282 10286.
13 Duriez B, Duquesnoy P, Escudier E et al: A common variant in
combination with a nonsense mutation in a member of the
thioredoxin family causes primary ciliary dyskinesia. Proc Natl
Acad Sci 2007; 104: 3336 3341.
14 Pazour GJ, Agrin N, Leszyk J, Witman GB: Proteomic analysis of
a eukaryotic cilium. J Cell Biol 2005; 170: 103– 113.
15 Blouin JL, Meeks M, Radhakrishna U et al: Primary ciliary
dyskinesia: a genome-wide linkage analysis reveals extensive
locus heterogeneity. Eur J Hum Genet 2000; 8: 109 118.
16 Meeks M, Walne A, Spiden S et al: A locus for primary
ciliary dyskinesia maps to chromosome 19q. J Med Genet 2000;
37: 241 244.
17 Jeganathan D, Chodhari R, Meeks M et al: Loci for primary ciliary
dyskinesia map to chrom osome 16p12.1 12.2 and 15q13.1 15.1
in Faroe Islands and Israeli Druze genetic isolates. J Med Genet
2004; 41: 233 240.
18 Geremek M, Zietkiewicz E, Diehl SR, Alizadeh BZ, Wijmenga C,
Witt M: Linkage analysis localises a Kartagener syndrome gene to
a 3.5 cM region on chrom osome 15q24 25. J Med Genet 2006;
43:e1.
19 Karadag B, James AJ, Gultekin E et al: Nasal and lower airway level
of nitric oxide in children with primary ciliary dyskinesia. Eur
Respir J 1999; 13: 1402 1405.
20 O’Connell JR, Weeks DE: PedCheck: a program for identifying
genotype incompatibilities in linkage analysis. Am J Hum Genet
1998; 63: 259 266.
21 Sobel E, Lange K: Descent graphs in pedigree analysis: applica-
tions to haplotyping, location scores, and marker sharing
statistics. Am J Hum Genet 1996; 58: 1323– 1337.
22 Rozen S, Skaletsky HJ: Primer3 on the WWW for general users
and for biologist programmers; in Krawetz S, Misener S (eds):
Bioinformatics Methods and Protocols: Methods in Molecular Biology.
Totowa, NJ: Humana Press, 2000, pp 365 386.
23 Wijmenga C, Hansen RS, Gimelli G et al: Genetic variation in ICF
syndrome: evidence for genetic heterogeneity. Hum Mutat 2000;
16: 509 517.
Supplementary Information accompanies the paper on European Journal of Human Genetics website (http://www.nature.com/ejhg)
Sequencing of 21 genes in 15q KS locus
M Geremek et al
695
European Journal of Human Genetics
... Therefore, implementation of genetic and molecular diagnosis of PCD is rather a necessity, regardless of whether it is a confirmation of a clinical diagnosis or suspicion of PCD. The identification of PCD genes has been based on linkage studies, candidate gene approaches and proteomic analyses, combined with sequencing of potentially causative genes [16][17][18][19]. Recently, the availability of high-throughput sequencing techniques contributed to the swift identification of new PCD-causative genes, resulting in more than 40 genes to be identified so far, which allowed 65% of cases to be genetically described [15,[20][21][22][23]. ...
... This is the first study which includes the genetic and molecular profiling, as well as phenotypic characterisation of PCD patients of Serbian ethnicity. Although large cohorts of PCD patients have been reported in Europe, only a small number of studies included molecular analysis of PCD patients [17,20,21,27]. More often they included individual families or analysis of specific genes [22,25,28,29]. ...
Article
Full-text available
Primary ciliary dyskinesia (PCD) is a rare inherited autosomal recessive or X-linked disorder that mainly affects lungs. Dysfunction of respiratory cilia causes symptoms such as chronic rhinosinusitis, coughing, rhinitis, conductive hearing loss and recurrent lung infections with bronchiectasis. It is now well known that pathogenic genetic changes lead to ciliary dysfunction. Here we report usage of clinical-exome based NGS approach in order to reveal underlying genetic causes in cohort of 21 patient with diagnosis of PCD. By detecting 18 (12 novel) potentially pathogenic genetic variants, we established the genetic cause of 11 (9 unrelated) patients. Genetic variants were detected in six PCD disease-causing genes, as well as in SPAG16 and SPAG17 genes, that were not detected in PCD patients so far, but were related to some symptoms of PCD. The most frequently mutated gene in our cohort was DNAH5 (27.77%). Identified variants were in homozygous, compound heterozygous and trans-heterozygous state. For detailed characterization of one novel homozygous genetic variant in DNAI1 gene (c. 947_948insG, p. Thr318TyrfsTer11), RT-qPCR and Western Blot analysis were performed. Molecular diagnostic approach applied in this study enables analysis of 29 PCD disease-causing and related genes. It resulted in mutation detection rate of 50% and enabled discovery of twelve novel mutations and pointed two possible novel PCD candidate genes.
... This disorder is a result of multiple mutations in various genes, including DNAH5 and DNAI1, which are pivotal in ciliary motion [6]. Moreover, these genes are not the only ones responsible for this disorder, as evidence has identified that some other genetic factors might also contribute to the severity of this ciliary dysfunction [7,8]. Symptoms such as respiratory distress and repeated lung infections are prominent findings since birth in these patients; however, severity may vary across individuals. ...
Article
Full-text available
Kartagener's syndrome is an autosomal recessive disorder with symptoms varying from chronic sinusitis to bronchiectasis and situs inversus (a congenital condition in which the visceral organs are located in an opposite location). We describe a rare and complicated case of a 40-year-old female patient who presented to the emergency room with significant chest congestion and Kartagener's syndrome. This case demonstrates the value of individualized and proactive care as well as the challenge of managing this illness, particularly when it coexists with type II respiratory failure related to pneumothorax.
... A particular mutation of TPSD1 p.Ala92Thr was observed in colon cancer among those non-responders after 5-fluorouracil-based therapy [46]. CFAP161, located on chromosome 15q in the linkage region of Kartagener syndrome [47], was only studied in mice and Xenopus [48]. RNF149, demonstrated as a DEG between normal tissue and prostate cancer [49], was discovered to be associated with 2[18F]fluoro-2-deoxy-d-glucose (FDG) uptake during positron emission tomography (PET) and survival in non-small cell lung cancer patients [50]. ...
Article
Full-text available
Kidney renal clear cell carcinoma (KIRC) is among the major causes of cancer-caused mortality around the world. Transient receptor potential channels (TRPs), due to their role in various human diseases, might become potential drug targets in cancer. The mRNA expression, copy number variation, single-nucleotide variation, prognostic values, drug sensitivity, and pathway regulation of TRPs were studied across cancer types. The ArrayExpress and The Cancer Genome Atlas (TCGA) databases were used to retrieve KIRC samples. Simultaneously, training, internal, and external cohorts were grouped. In KIRC, a prognostic signature with superior survival prediction in contrast with other well-established signatures was created after a stepwise screening of optimized genes linked to TRPs using univariate Cox, weighted gene co-expression network analysis, multivariate Cox, and least absolute shrinkage and selection operator regression analyses. Subsequent to the determination of risk levels, the variations in the expression of immune checkpoint genes, tumor mutation burden, and immune subtypes and response between low-risk and high-risk subgroups were studied using a variety of bioinformatics algorithms, including ESTIMATE, XCELL, EPIC, CIBERSORT-ABS, CIBERSORT, MCPCOUNTER, TIMER, and QUANTISEQ. Gene set enrichment analysis helped in the identification of abnormal pathways across the low- and high-risk subgroups. Besides, high-risk KIRC patients might benefit from ABT888, AZD6244, AZD7762, Bosutinib, Camptothecin, CI1040, JNK inhibitor VIII, KU55933, Lenalidomide, Nilotinib, PLX4720, RO3306, Vinblastine, and ZM.447439; however, low-risk populations might benefit from Bicalutamide, FH535, and OSI906. Finally, calibration curves were used to validate the nomogram with a satisfactory predictive survival probability. In conclusion, this research provides useful insight that can aid and guide clinical practice and scientific research.
... Morpholino-mediated knock down of Cfap161 in zebrafish led to the loss of outer dynein arms, reduced beating frequency of pronephric cilia and strong ciliopathy phenotypes including left-right asymmetry defects in one study 37 , but only a curved body axis and hydrocephalus in a subsequent analysis 38 leaving some uncertainties concerning the ciliary function of CFAP161. Human CFAP161 (c15orf26) is located on chromosome 15q in the linkage region of Kartagener syndrome 39 , suggesting a potential involvement of Cfap161 mutations in this subtype of PCD. Here, we describe the analysis of CFAP161 function in Xenopus crispants and mutant mice generated by homologous recombination. ...
Article
Full-text available
Cilia are protrusions of the cell surface and composed of hundreds of proteins many of which are evolutionary and functionally well conserved. In cells assembling motile cilia the expression of numerous ciliary components is under the control of the transcription factor FOXJ1. Here, we analyse the evolutionary conserved FOXJ1 target CFAP161 in Xenopus and mouse. In both species Cfap161 expression correlates with the presence of motile cilia and depends on FOXJ1. Tagged CFAP161 localises to the basal bodies of multiciliated cells of the Xenopus larval epidermis, and in mice CFAP161 protein localises to the axoneme. Surprisingly, disruption of the Cfap161 gene in both species did not lead to motile cilia-related phenotypes, which contrasts with the conserved expression in cells carrying motile cilia and high sequence conservation. In mice mutation of Cfap161 stabilised the mutant mRNA making genetic compensation triggered by mRNA decay unlikely. However, genes related to microtubules and cilia, microtubule motor activity and inner dyneins were dysregulated, which might buffer the Cfap161 mutation.
... 7 El defecto es congénito y los síntomas se presentan desde los primeros años de vida, de ahí la importancia de que los pediatras conozcan esta enfermedad como diagnóstico diferencial sustancial, aunque poco frecuente en niños con síntomas recurrentes en las vías respiratorias superiores e inferiores. 8 Por otra parte, los Haemophilus y los Staphylococcus aureus son bacterias que afectan a estos pacientes. Según se ha visto, el tratamiento para las personas con esta alteración congénita rara incluye antibióticos por vías intravenosa u oral, a fin de aminorar las infecciones de las vías aéreas. ...
Article
Full-text available
The case report of a 12 year female adolescent who presented repeated breathing infections and bronchial obstruction, with a situs inversus, bronquiectasia and sinusitis is described; as well as recurrent pneumonia, bronchitis and bronchial hiperreactivity patterns. Taking into account these particularities and the complementary tests results, including the computed tomographic scan and the thorax x-ray, the Kartagener syndrome was diagnosed. Due to the uncommon of its presentation these findings are shared with the scientific community in general
Chapter
Pediatric pulmonary pathology is a challenging topic within pathology. There are several reasons, such as diseases of the lung are rare. Most pulmonary diseases within childhood are infections, which do not need a tissue analysis. Lung diseases in childhood do occur in two peaks: the first during the first 2 years of life, the second usually at the “Kindergarden” age. The first group comprises many developmental lung diseases, whereas the second group includes diseases based on gene mutations, immune system dysfunction, and infections. For the analysis of pediatric pulmonary diseases, the knowledge of lung development is essential, and the analysis of genetic abnormalities is becoming important.
Article
The axoneme of motile cilia is the largest macromolecular machine of eukaryotic cells. In humans, impaired axoneme function causes a range of ciliopathies. Axoneme assembly, structure, and motility require a radially arranged set of doublet microtubules, each decorated in repeating patterns with non-tubulin components. We use single-particle cryo-electron microscopy to visualize and build an atomic model of the repeating structure of a native axonemal doublet microtubule, which reveals the identities, positions, repeat lengths, and interactions of 38 associated proteins, including 33 microtubule inner proteins (MIPs). The structure demonstrates how these proteins establish the unique architecture of doublet microtubules, maintain coherent periodicities along the axoneme, and stabilize the microtubules against the repeated mechanical stress induced by ciliary motility. Our work elucidates the architectural principles that underpin the assembly of this large, repetitive eukaryotic structure and provides a molecular basis for understanding the etiology of human ciliopathies.
Article
This article is a case report of a 26 years old woman with history of high and low respiratory problems to repetition, which yielded with symptomatic treatment or the antibiotic use by oral route from the childhood. When she was 25 years old, she began with episodes of expulsion of red blood, brillant and bubble by the mouth of different quantity; these were made frequent more and more. By such reason she was hospitalized in the service of pneumology of the Benefico Juridico Hospital. In this occasion a simple x-ray of thorax was made to her, in which visualized dextrocardia accompanied of an accentuation of the bibasal broncho-vascular pattern. A history of infertility referred by the patient, the episodes of recurrent pansinusitis, respiratory pictures and the radiologic findings made diagnose of a primary ciliary dyskinesia, specifically Kartagener's syndrome. Our intention with this case is to reflect in that it is a possible diagnosis when pictures of hemoptisis and history of infertility are associated, although in our country and in the rest of the world the frequent of this disease is low.
Chapter
Electron microscopy is an important tool to know not only the presence but the nature of spermatozoa defects. The data it provides along with molecular studies, even if taken in isolation, are of higher value than simply mobility or coloring characteristics. An ultrastructural study is indicated in the following situations: (a) in all cases of moderate or severe persistent asthenozoospermia unexplained by sperm infection, varicocele, or the presence of anti-spermatozoa antibodies; (b) in teratozoospermia, provided that the number of spermatozoa showing the same anomaly is very high; and (c) in all patients with lengthy unexplained infertility, before including them in an IVF program. Anomalies considered to have a potential genetic basis are the following: globozoospermia, hypoplasia of the acrosome, macrocephalic spermatozoa syndrome, abnormal head-tail attachment, decapitated spermatozoa syndrome, fibrous sheath dysplasia, and primary ciliary dyskinesia. The different morphological abnormalities and the molecular basis of these anomalies are discussed in this chapter.
Chapter
Developmental and inherited lung diseases are rare events and therefore not many cases are seen in single institutions. This has resulted in a vast amount of single case reports, but only rarely have these been collected and studied with the focus on classification. A few studies came out from the Armed Forces Institute of Pathology (T. Stocker), which for a long time was one of the major pathology institutes with a vast amount of collected cases. Only recently initiatives were started in the USA and Europe to collect interstitial and developmental childhood cases and classify them accordingly [1–4]. As soon as cases have been analyzed, it was apparent that there are two peaks when diseases are recognized in children: The first peak occurs in the first 2 years of life comprising mainly developmental diseases and some acquired infections transmitted intrauterine or immediately postnatal. The second peak occurs in the period between 3 and 6 years of life and is composed predominantly by infections, genetic inherited diseases such as cystic fibrosis, and miscellaneous others.
Article
Full-text available
Rationale: Primary ciliary dyskinesia (PCD) is a rare, usually autosomal recessive, genetic disorder characterized by ciliary dysfunction, sino-pulmonary disease, and situs inversus. Disease-causing mutations have been reported in DNAI1 and DNAH5 encoding outer dynein arm (ODA) proteins of cilia. Objectives: We analyzed DNAI1 to identify disease-causing mutations in PCD and to determine if the previously reported IVS1+2_3insT (219+3insT) mutation represents a "founder" or "hot spot" mutation. Methods: Patients with PCD from 179 unrelated families were studied. Exclusion mapping showed no linkage to DNAI1 for 13 families; the entire coding region was sequenced in a patient from the remaining 166 families. Reverse transcriptase–polymerase chain reaction (RT-PCR) was performed on nasal epithelial RNA in 14 families. Results: Mutations in DNAI1 including 12 novel mutations were identified in 16 of 179 (9%) families; 14 harbored biallelic mutations. Deep intronic splice mutations were not identified by reverse transcriptase–polymerase chain reaction. The prevalence of mutations in families with defined ODA defect was 13%; no mutations were found in patients without a defined ODA defect. The previously reported IVS1+2_3insT mutation accounted for 57% (17/30) of mutant alleles, and marker analysis indicates a common founder for this mutation. Seven mutations occurred in three exons (13, 16, and 17); taken together with previous reports, these three exons are emerging as mutation clusters harboring 29% (12/42) of mutant alleles. Conclusions: A total of 10% of patients with PCD are estimated to harbor mutations in DNAI1; most occur as a common founder IVS1+2_3insT or in exons 13, 16, and 17. This information is useful for establishing a clinical molecular genetic test for PCD.
Article
Designing PCR and sequencing primers are essential activities for molecular biologists around the world. This chapter assumes acquaintance with the principles and practice of PCR, as outlined in, for example, refs. 1, 2, 3, 4.
Article
ICF syndrome is a rare autosomal recessive immunoglobulin deficiency, sometimes combined with defective cellular immunity. Other features that are frequently observed in ICF syndrome patients include facial dysmorphism, developmental delay, and recurrent infections. The most diagnostic feature of ICF syndrome is the branching of chromosomes 1, 9, and 16 due to pericentromeric instability. Positional candidate cloning recently discovered the de novo DNA methyltransferase 3B (DNMT3B) as the responsible gene by identifying seven different mutations in nine ICF patients. DNMT3B specifically methylates repeat sequences adjacent to the centromeres of chromosome 1, 9, and 16. Our panel of 14 ICF patients was subjected to mutation analysis in the DNMT3B gene. Mutations in DNMT3B were discovered in only nine of our 14 ICF patients. Moreover, two ICF patients from consanguineous families who did not show autozygosity (i.e. homozygosity by descent) for the DNMT3B locus did not reveal DNMT3B mutations, suggesting genetic heterogeneity for this disease. Mutation analysis revealed 11 different mutations, including seven novel ones: eight different missense mutations, two different nonsense mutations, and a splice‐site mutation leading to the insertion of three aa’s. The missense mutations occurred in or near the catalytic domain of DNMT3B protein, indicating a possible interference with the normal functioning of the enzyme. However, none of the ICF patients was homozygous for a nonsense allele, suggesting that absence of this enzyme is not compatible with life. Compound heterozygosity for a missense and a nonsense mutation did not seem to correlate with a more severe phenotype. Hum Mutat 16:509–517, 2000. © 2000 Wiley‐Liss, Inc.
Article
ICF syndrome is a rare autosomal recessive immunoglobulin deficiency, sometimes combined with defective cellular immunity. Other features that are frequently observed in ICF syndrome patients include facial dysmorphism, developmental delay, and recurrent infections. The most diagnostic feature of ICF syndrome is the branching of chromosomes 1, 9, and 16 due to pericentromeric instability. Positional candidate cloning recently discovered the de novo DNA methyltransferase 3B (DNMT3B) as the responsible gene by identifying seven different mutations in nine ICF patients. DNMT3B specifically methylates repeat sequences adjacent to the centromeres of chromosome 1, 9, and 16. Our panel of 14 ICF patients was subjected to mutation analysis in the DNMT3B gene. Mutations in DNMT3B were discovered in only nine of our 14 ICF patients. Moreover, two ICF patients from consanguineous families who did not show autozygosity (i.e. homozygosity by descent) for the DNMT3B locus did not reveal DNMT3B mutations, suggesting genetic heterogeneity for this disease. Mutation analysis revealed 11 different mutations, including seven novel ones: eight different missense mutations, two different nonsense mutations, and a splice-site mutation leading to the insertion of three aa’s. The missense mutations occurred in or near the catalytic domain of DNMT3B protein, indicating a possible interference with the normal functioning of the enzyme. However, none of the ICF patients was homozygous for a nonsense allele, suggesting that absence of this enzyme is not compatible with life. Compound heterozygosity for a missense and a nonsense mutation did not seem to correlate with a more severe phenotype. Hum Mutat 16:509–517, 2000. © 2000 Wiley-Liss, Inc.
Chapter
Designing PCR and sequencing primers are essential activities for molecular biologists around the world. This chapter assumes acquaintance with the principles and practice of PCR, as outlined in, for example, refs. 1–4.
Article
Kartagener syndrome (KS) is a trilogy of symptoms (nasal polyps, bronchiectasis, and situs inversus totalis) that is associated with ultrastructural anomalies of cilia of epithelial cells covering the upper and lower respiratory tracts and spermatozoa flagellae. The axonemal dynein intermediate-chain gene 1 (DNAI1), which has been demonstrated to be responsible for a case of primary ciliary dyskinesia (PCD) without situs inversus, was screened for mutation in a series of 34 patients with KS. We identified compound heterozygous DNAI1 gene defects in three independent patients and in two of their siblings who presented with PCD and situs solitus (i.e., normal position of inner organs). Strikingly, these five patients share one mutant allele (splice defect), which is identical to one of the mutant DNAI1 alleles found in the patient with PCD, reported elsewhere. Finally, this study demonstrates a link between ciliary function and situs determination, since compound mutation heterozygosity in DNAI1 results in PCD with situs solitus or situs inversus (KS).
Article
The introduction of stochastic methods in pedigree analysis has enabled geneticists to tackle computations intractable by standard deterministic methods. Until now these stochastic techniques have worked by running a Markov chain on the set of genetic descent states of a pedigree. Each descent state specifies the paths of gene flow in the pedigree and the founder alleles dropped down each path. The current paper follows up on a suggestion by Elizabeth Thompson that genetic descent graphs offer a more appropriate space for executing a Markov chain. A descent graph specifies the paths of gene flow but not the particular founder alleles traveling down the paths. This paper explores algorithms for implementing Thompson's suggestion for codominant markers in the context of automatic haplotyping, estimating location scores, and computing gene-clustering statistics for robust linkage analysis. Realistic numerical examples demonstrate the feasibility of the algorithms.
Article
Prior to performance of linkage analysis, elimination of all Mendelian inconsistencies in the pedigree data is essential. Often, identification of erroneous genotypes by visual inspection can be very difficult and time consuming. In fact, sometimes the errors are not recognized until the stage of running linkage-analysis software. The effort then required to find the erroneous genotypes and to cross-reference pedigree and marker data that may have been recoded and renumbered can be not only tedious but also quite daunting, in the case of very large pedigrees. We have implemented four error-checking algorithms in a new computer program, PedCheck, which will assist researchers in identifying all Mendelian inconsistencies in pedigree data and will provide them with useful and detailed diagnostic information to help resolve the errors. Our program, which uses many of the algorithms implemented in VITESSE, handles large data sets quickly and efficiently, accepts a variety of input formats, and offers various error-checking algorithms that match the subtlety of the pedigree error. These algorithms range from simple parent-offspring-compatibility checks to a single-locus likelihood-based statistic that identifies and ranks the individuals most likely to be in error. We use various real data sets to illustrate the power and effectiveness of our program.